05.01.2013 Views

RePoSS #11: The Mathematics of Niels Henrik Abel: Continuation ...

RePoSS #11: The Mathematics of Niels Henrik Abel: Continuation ...

RePoSS #11: The Mathematics of Niels Henrik Abel: Continuation ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>RePoSS</strong>: Research Publications on Science Studies<br />

<strong>RePoSS</strong> <strong>#11</strong>:<br />

<strong>The</strong> <strong>Mathematics</strong> <strong>of</strong> <strong>Niels</strong> <strong>Henrik</strong><br />

<strong>Abel</strong>: <strong>Continuation</strong> and New<br />

Approaches in <strong>Mathematics</strong><br />

During the 1820s<br />

<strong>Henrik</strong> Kragh Sørensen<br />

October 2010<br />

Centre for Science Studies, University <strong>of</strong> Aarhus, Denmark<br />

Research group: History and philosophy <strong>of</strong> science


Please cite this work as:<br />

<strong>Henrik</strong> Kragh Sørensen (Oct. 2010). <strong>The</strong> <strong>Mathematics</strong> <strong>of</strong> <strong>Niels</strong><br />

<strong>Henrik</strong> <strong>Abel</strong>: <strong>Continuation</strong> and New Approaches in <strong>Mathematics</strong><br />

During the 1820s. <strong>RePoSS</strong>: Research Publications on Science<br />

Studies 11. Aarhus: Centre for Science Studies, University <strong>of</strong><br />

Aarhus. url: ���������������������������.<br />

Copyright c○ <strong>Henrik</strong> Kragh Sørensen, 2010


<strong>The</strong> <strong>Mathematics</strong> <strong>of</strong><br />

NIELS HENRIK ABEL<br />

<strong>Continuation</strong> and New Approaches<br />

in <strong>Mathematics</strong> During the 1820s<br />

HENRIK KRAGH SØRENSEN


For Mom and Dad<br />

who were always there for me<br />

when I abandoned<br />

all good manners,<br />

good friends,<br />

and common sense<br />

to pursue my dreams.


<strong>The</strong> <strong>Mathematics</strong> <strong>of</strong><br />

NIELS HENRIK ABEL<br />

<strong>Continuation</strong> and New Approaches<br />

in <strong>Mathematics</strong> During the 1820s<br />

HENRIK KRAGH SØRENSEN<br />

PhD dissertation<br />

March 2002<br />

Electronic edition, October 2010<br />

History <strong>of</strong> Science Department<br />

<strong>The</strong> Faculty <strong>of</strong> Science<br />

University <strong>of</strong> Aarhus, Denmark


This dissertation was submitted to the Faculty <strong>of</strong> Science,<br />

University <strong>of</strong> Aarhus in March 2002 for the purpose <strong>of</strong> ob-<br />

taining the scientific PhD degree. It was defended in a public<br />

PhD defense on May 3, 2002. A second, only slightly revised<br />

edition was printed October, 2004.<br />

<strong>The</strong> PhD program was supervised by associate pr<strong>of</strong>essor<br />

KIRSTI ANDERSEN, History <strong>of</strong> Science Department, Univer-<br />

sity <strong>of</strong> Aarhus.<br />

Pr<strong>of</strong>essors UMBERTO BOTTAZZINI (University <strong>of</strong> Palermo,<br />

Italy), JEREMY J. GRAY (Open University, UK), and OLE<br />

KNUDSEN (History <strong>of</strong> Science Department, Aarhus) served<br />

on the committee for the defense.<br />

c○ <strong>Henrik</strong> Kragh Sørensen and the History <strong>of</strong> Science De-<br />

partment (Department <strong>of</strong> Science Studies), University <strong>of</strong><br />

Aarhus, 2002–2010.<br />

<strong>The</strong> dissertation was typeset in Palatino using pdfLATEX.<br />

First edition (March 2002, 7 copies) was copied and bound<br />

by the Printshop at the Faculty <strong>of</strong> Science, Aarhus.<br />

Second edition (October 2004, 5 copies) was printed and<br />

bound by the Printshop at Agder University College, Kris-<br />

tiansand.<br />

This electronic edition was compiled on October 28, 2010.<br />

For further information, additions, corrections, and<br />

contact to the author, please refer to the website<br />

������������������������������.<br />

<strong>The</strong> picture on the front page is a painting <strong>of</strong> NIELS HENRIK<br />

ABEL performed by the Norwegian painter JOHAN GØRB-<br />

ITZ during ABEL’s time in Paris 1826. It is the only authentic<br />

depiction <strong>of</strong> ABEL and is reproduced from (Ore, 1957).<br />

<strong>The</strong> picture on the reverse shows a curlicue frequently used<br />

by ABEL in his notebooks to mark the end <strong>of</strong> manuscripts. It<br />

is reproduced from (Stubhaug, 1996).


Contents<br />

Contents i<br />

List <strong>of</strong> Tables vii<br />

List <strong>of</strong> Figures ix<br />

List <strong>of</strong> Boxes xi<br />

List <strong>of</strong> <strong>The</strong>orems etc. xiii<br />

Summary xv<br />

Preface to the 2004 edition xvii<br />

<strong>Abel</strong>’s mathematics in the context <strong>of</strong> traditions and changes . . . . . . . . . . xvii<br />

Recent literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xviii<br />

Preface to the 2002 edition xxi<br />

Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi<br />

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxii<br />

I Introduction 1<br />

1 Introduction 3<br />

1.1 <strong>The</strong> historical and geographical setting <strong>of</strong> ABEL’s life . . . . . . . . . . . 4<br />

1.2 <strong>The</strong> mathematical topics involved . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.3 <strong>The</strong>mes from early nineteenth-century mathematics . . . . . . . . . . . . 7<br />

1.4 Reflections on methodology . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2 Biography <strong>of</strong> NIELS HENRIK ABEL 17<br />

2.1 Childhood and education . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.2 “Study the masters” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.3 <strong>The</strong> European tour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.4 Back in Norway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

i


3 Historical background 39<br />

3.1 Mathematical institutions and networks . . . . . . . . . . . . . . . . . . . 39<br />

3.2 ABEL’s position in mathematical traditions . . . . . . . . . . . . . . . . . 41<br />

3.3 <strong>The</strong> state <strong>of</strong> mathematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.4 ABEL’s legacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

II “My favorite subject is algebra” 47<br />

4 <strong>The</strong> position and role <strong>of</strong> ABEL’s works within the discipline <strong>of</strong> algebra 49<br />

4.1 Outline <strong>of</strong> ABEL’s results and their structural position . . . . . . . . . . . 50<br />

4.2 Mathematical change as a history <strong>of</strong> new questions . . . . . . . . . . . . . 53<br />

5 Towards unsolvable equations 57<br />

5.1 Algebraic solubility before LAGRANGE . . . . . . . . . . . . . . . . . . . . 59<br />

5.2 LAGRANGE’s theory <strong>of</strong> equations . . . . . . . . . . . . . . . . . . . . . . . 65<br />

5.3 Solubility <strong>of</strong> cyclotomic equations . . . . . . . . . . . . . . . . . . . . . . . 72<br />

5.4 Belief in algebraic solubility shaken . . . . . . . . . . . . . . . . . . . . . . 80<br />

5.5 RUFFINI’s pro<strong>of</strong>s <strong>of</strong> the insolubility <strong>of</strong> the quintic . . . . . . . . . . . . . . 84<br />

5.6 CAUCHY’ theory <strong>of</strong> permutations and a new pro<strong>of</strong> <strong>of</strong> RUFFINI’s theorem 90<br />

5.7 Some algebraic tools used by GAUSS . . . . . . . . . . . . . . . . . . . . . 95<br />

6 Algebraic insolubility <strong>of</strong> the quintic 97<br />

6.1 <strong>The</strong> first break with tradition . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.2 Outline <strong>of</strong> ABEL’s pro<strong>of</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

6.3 Classification <strong>of</strong> algebraic expressions . . . . . . . . . . . . . . . . . . . . 101<br />

6.4 ABEL and the theory <strong>of</strong> permutations . . . . . . . . . . . . . . . . . . . . . 108<br />

6.5 Permutations linked to root extractions . . . . . . . . . . . . . . . . . . . . 110<br />

6.6 Combination into an impossibility pro<strong>of</strong> . . . . . . . . . . . . . . . . . . . 112<br />

6.7 ABEL and RUFFINI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

6.8 Limiting the class <strong>of</strong> solvable equations . . . . . . . . . . . . . . . . . . . 124<br />

6.9 Reception <strong>of</strong> ABEL’s work on the quintic . . . . . . . . . . . . . . . . . . . 125<br />

6.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

7 Particular classes <strong>of</strong> solvable equations 141<br />

7.1 Solubility <strong>of</strong> <strong>Abel</strong>ian equations . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

7.2 Elliptic functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

7.3 <strong>The</strong> concept <strong>of</strong> irreducibility at work . . . . . . . . . . . . . . . . . . . . . 157<br />

7.4 Enlarging the class <strong>of</strong> solvable equations . . . . . . . . . . . . . . . . . . . 160<br />

8 A grand theory in spe 163<br />

8.1 Inverting the approach once again . . . . . . . . . . . . . . . . . . . . . . 163<br />

8.2 Construction <strong>of</strong> the irreducible equation . . . . . . . . . . . . . . . . . . . 165<br />

ii


8.3 Refocusing on the equation . . . . . . . . . . . . . . . . . . . . . . . . . . 171<br />

8.4 Further ideas on the theory <strong>of</strong> equations . . . . . . . . . . . . . . . . . . . 176<br />

8.5 General resolution <strong>of</strong> the problem by E. GALOIS . . . . . . . . . . . . . . 181<br />

IIIInterlude: ABEL and the ‘new rigor’ 189<br />

9 <strong>The</strong> nineteenth-century change in epistemic techniques 191<br />

10 Toward rigorization <strong>of</strong> analysis 193<br />

10.1 EULER’s vision <strong>of</strong> analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 193<br />

10.2 LAGRANGE’s new focus on rigor . . . . . . . . . . . . . . . . . . . . . . . 197<br />

10.3 Early rigorization <strong>of</strong> theory <strong>of</strong> series . . . . . . . . . . . . . . . . . . . . . 198<br />

10.4 New types <strong>of</strong> series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204<br />

11 CAUCHY’s new foundation for analysis 207<br />

11.1 Programmatic focus on arithmetical equality . . . . . . . . . . . . . . . . 207<br />

11.2 CAUCHY’s concepts <strong>of</strong> limits and infinitesimals . . . . . . . . . . . . . . . 209<br />

11.3 Divergent series have no sum . . . . . . . . . . . . . . . . . . . . . . . . . 210<br />

11.4 Means <strong>of</strong> testing for convergence <strong>of</strong> series . . . . . . . . . . . . . . . . . . 212<br />

11.5 CAUCHY’s pro<strong>of</strong> <strong>of</strong> the binomial theorem . . . . . . . . . . . . . . . . . . 214<br />

11.6 Early reception <strong>of</strong> CAUCHY’s new rigor . . . . . . . . . . . . . . . . . . . 218<br />

12 ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem 221<br />

12.1 ABEL’s critical attitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223<br />

12.2 Infinitesimals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228<br />

12.3 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229<br />

12.4 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233<br />

12.5 ABEL’s “exception” . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238<br />

12.6 A curious reaction: Lehrsatz V . . . . . . . . . . . . . . . . . . . . . . . . . 241<br />

12.7 From power series to absolute convergence . . . . . . . . . . . . . . . . . 246<br />

12.8 Product theorems <strong>of</strong> infinite series . . . . . . . . . . . . . . . . . . . . . . 251<br />

12.9 ABEL’s pro<strong>of</strong> <strong>of</strong> the binomial theorem . . . . . . . . . . . . . . . . . . . . 254<br />

12.10Aspects <strong>of</strong> ABEL’s binomial paper . . . . . . . . . . . . . . . . . . . . . . . 260<br />

13 ABEL and OLIVIER on convergence tests 265<br />

13.1 OLIVIER’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265<br />

13.2 ABEL’s counter example . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269<br />

13.3 ABEL’s general refutation . . . . . . . . . . . . . . . . . . . . . . . . . . . 271<br />

13.4 More characterizations and tests <strong>of</strong> convergence . . . . . . . . . . . . . . 272<br />

14 Reception <strong>of</strong> ABEL’s contribution to rigorization 277<br />

14.1 Reception <strong>of</strong> ABEL’s rigorization . . . . . . . . . . . . . . . . . . . . . . . 277<br />

iii


14.2 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281<br />

IVElliptic functions and the Paris mémoire 283<br />

15 Elliptic integrals and functions: Chronology and topics 285<br />

15.1 Elliptic transcendentals before the nineteenth century . . . . . . . . . . . 286<br />

15.2 <strong>The</strong> lemniscate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289<br />

15.3 LEGENDRE’s theory <strong>of</strong> elliptic integrals . . . . . . . . . . . . . . . . . . . . 292<br />

15.4 Left in the drawer: GAUSS on elliptic functions . . . . . . . . . . . . . . . 296<br />

15.5 Chronology <strong>of</strong> ABEL’s work on elliptic transcendentals . . . . . . . . . . 297<br />

16 <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals 299<br />

16.1 <strong>The</strong> importance <strong>of</strong> the lemniscate . . . . . . . . . . . . . . . . . . . . . . . 299<br />

16.2 Inversion in the Recherches . . . . . . . . . . . . . . . . . . . . . . . . . . . 300<br />

16.3 <strong>The</strong> division problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313<br />

16.4 Perspectives on inversion . . . . . . . . . . . . . . . . . . . . . . . . . . . 319<br />

17 Steps in the process <strong>of</strong> coming to “know” elliptic functions 321<br />

17.1 Infinite representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321<br />

17.2 Elliptic functions as ratios <strong>of</strong> power series . . . . . . . . . . . . . . . . . . 325<br />

17.3 Characterization <strong>of</strong> ABEL’s representations . . . . . . . . . . . . . . . . . 328<br />

17.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330<br />

18 Tools in ABEL’s research on elliptic transcendentals 331<br />

18.1 Transformation theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331<br />

18.2 Integration in logarithmic terms . . . . . . . . . . . . . . . . . . . . . . . . 339<br />

18.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345<br />

19 <strong>The</strong> Paris memoir 347<br />

19.1 ABEL’s approach to the Paris memoir . . . . . . . . . . . . . . . . . . . . . 347<br />

19.2 <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> . . . . . . . . . . . . . . 351<br />

19.3 Additional, tentative remarks on ABEL’s tools . . . . . . . . . . . . . . . . 371<br />

19.4 <strong>The</strong> fate <strong>of</strong> the Paris memoir . . . . . . . . . . . . . . . . . . . . . . . . . . . 375<br />

19.5 Reception <strong>of</strong> the Paris memoir . . . . . . . . . . . . . . . . . . . . . . . . . 376<br />

19.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377<br />

20 General approaches to elliptic functions 379<br />

20.1 ABEL’s version <strong>of</strong> a general theory <strong>of</strong> elliptic functions . . . . . . . . . . . 379<br />

20.2 Other ways <strong>of</strong> introducing elliptic functions in the nineteenth century . . 381<br />

20.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383<br />

iv


V ABEL’s mathematics and the rise <strong>of</strong> concepts 385<br />

21 ABEL’s mathematics and the rise <strong>of</strong> concepts 387<br />

21.1 From formulae to concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . 387<br />

21.2 Concepts and classes enter mathematics . . . . . . . . . . . . . . . . . . . 393<br />

21.3 <strong>The</strong> role <strong>of</strong> counter examples . . . . . . . . . . . . . . . . . . . . . . . . . 395<br />

21.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400<br />

A ABEL’s correspondence 403<br />

B ABEL’s manuscripts 407<br />

Bibliography 411<br />

Index <strong>of</strong> names 431<br />

Index 435<br />

v


List <strong>of</strong> Tables<br />

2.1 Time <strong>of</strong> publications <strong>of</strong> CRELLE’s Journal 1826–29 . . . . . . . . . . . . . . . . 32<br />

2.2 List <strong>of</strong> most productive authors in CRELLE’s Journal 1826–29 . . . . . . . . . 33<br />

2.3 Summary <strong>of</strong> NIELS HENRIK ABEL’s biography . . . . . . . . . . . . . . . . . 37<br />

5.1 RUFFINI’s classification <strong>of</strong> permutations . . . . . . . . . . . . . . . . . . . . . 86<br />

5.2 <strong>The</strong> N m<br />

circles formed by applying (As<br />

At ) to A1, . . . , AN. . . . . . . . . . . . . . 94<br />

6.1 <strong>The</strong> order and degree <strong>of</strong> some expressions in CARDANO’s solution to the<br />

general cubic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

10.1 LAGRANGE’s monographs on his algebraic analysis . . . . . . . . . . . . . . 197<br />

11.1 CAUCHY’s textbooks on the calculus . . . . . . . . . . . . . . . . . . . . . . . 208<br />

15.1 LEGENDRE’s publications on elliptic transcendentals . . . . . . . . . . . . . . 293<br />

15.2 LEGENDRE’s classification <strong>of</strong> elliptic integrals . . . . . . . . . . . . . . . . . . 294<br />

15.3 ABEL’s early unpublished works on elliptic integrals and related topics . . . 298<br />

18.1 Important dates in the ABEL-JACOBI-rivalry . . . . . . . . . . . . . . . . . . . 346<br />

A.1 Correspondence sorted by sender . . . . . . . . . . . . . . . . . . . . . . . . . 403<br />

A.1 Correspondence sorted by sender (cont.) . . . . . . . . . . . . . . . . . . . . . 404<br />

A.1 Correspondence sorted by sender (cont.) . . . . . . . . . . . . . . . . . . . . . 405<br />

B.1 <strong>Abel</strong> manuscript collections in Oslo . . . . . . . . . . . . . . . . . . . . . . . . 408<br />

B.2 <strong>Abel</strong> manuscripts in Oslo, MS:592 . . . . . . . . . . . . . . . . . . . . . . . . . 408<br />

B.3 <strong>Abel</strong> manuscripts in Oslo, MS:434 . . . . . . . . . . . . . . . . . . . . . . . . . 409<br />

B.4 <strong>Abel</strong> notebooks in Oslo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409<br />

B.5 <strong>Abel</strong> manuscripts in the Mittag-Leffler Institute, Djursholm . . . . . . . . . . 410<br />

vii


List <strong>of</strong> Figures<br />

2.1 NIELS HENRIK ABEL (1802–1829) . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.2 <strong>The</strong> southern part <strong>of</strong> Norway with Christiania and Finnøy marked . . . . . 19<br />

2.3 BERNT MICHAEL HOLMBOE (1795–1850) . . . . . . . . . . . . . . . . . . . . 20<br />

2.4 CARL FERDINAND DEGEN (1766–1825) . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.5 AUGUST LEOPOLD CRELLE (1780–1855) . . . . . . . . . . . . . . . . . . . . . 28<br />

5.1 JOSEPH LOUIS LAGRANGE (1736–1813) . . . . . . . . . . . . . . . . . . . . . . 65<br />

5.2 EDWARD WARING (1734–1798) . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

5.3 CARL FRIEDRICH GAUSS (1777–1855) . . . . . . . . . . . . . . . . . . . . . . 73<br />

5.4 PAOLO RUFFINI (1765–1822) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.5 AUGUSTIN-LOUIS CAUCHY (1789–1857) . . . . . . . . . . . . . . . . . . . . . 91<br />

6.1 Limiting the class <strong>of</strong> solvable equations . . . . . . . . . . . . . . . . . . . . . 125<br />

6.2 WILLIAM ROWAN HAMILTON (1805–1865) . . . . . . . . . . . . . . . . . . . 129<br />

7.1 N. H. ABEL’S (1802–1829) drawing <strong>of</strong> the lemniscate in one <strong>of</strong> his notebooks 153<br />

7.2 Last page <strong>of</strong> ABEL’S manuscript for Mémoire sur une classe particulière . . . . 156<br />

7.3 Extending the class <strong>of</strong> solvable equations: <strong>Abel</strong>ian equations . . . . . . . . . 160<br />

8.1 Page from ABEL’S notebook manuscript on algebraic solubility . . . . . . . 177<br />

8.2 EVARISTE GALOIS (1811–1832) . . . . . . . . . . . . . . . . . . . . . . . . . . . 182<br />

10.1 BERNARD BOLZANO (1781–1848) . . . . . . . . . . . . . . . . . . . . . . . . . 202<br />

10.2 JEAN BAPTISTE JOSEPH FOURIER (1768–1830) . . . . . . . . . . . . . . . . . . 205<br />

12.1 Comparison <strong>of</strong> CAUCHY’s and ABEL’s structures <strong>of</strong> the basic theory <strong>of</strong> in-<br />

finite series. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232<br />

12.2 Graphical representation <strong>of</strong> ABEL’s “Exception” . . . . . . . . . . . . . . . . 240<br />

13.1 L. OLIVIER’S geometrical argument . . . . . . . . . . . . . . . . . . . . . . . 268<br />

15.1 EULER’s rectification <strong>of</strong> an ellipse by infinite series . . . . . . . . . . . . . . . 288<br />

15.2 LEONHARD EULER (1707–1783) . . . . . . . . . . . . . . . . . . . . . . . . . . 291<br />

15.3 ADRIEN-MARIE LEGENDRE (1752–1833) . . . . . . . . . . . . . . . . . . . . . 293<br />

16.1 Stamp depicting Gauss and the construction <strong>of</strong> the regular 17-gon. . . . . . 300<br />

ix


16.2 ABEL’S extension to the complex rectangle . . . . . . . . . . . . . . . . . . . 305<br />

18.1 CARL GUSTAV JACOB JACOBI (1804–1851) . . . . . . . . . . . . . . . . . . . . 332<br />

19.1 GEORG FRIEDRICH BERNHARD RIEMANN (1826–1866) . . . . . . . . . . . . 374<br />

21.1 <strong>The</strong> equality <strong>of</strong> the concepts <strong>of</strong> explicit algebraic expressions and ABEL’s<br />

normal form. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389<br />

21.2 Delineating the border between a concept and its super-concept. . . . . . . . 396<br />

x


List <strong>of</strong> Boxes<br />

1 <strong>The</strong> algebraic reduction <strong>of</strong> the cubic equation . . . . . . . . . . . . . . . . . . 60<br />

2 Pro<strong>of</strong> <strong>of</strong> Lehrsatz I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230<br />

3 A modern reconstruction <strong>of</strong> DIRICHLET’s pro<strong>of</strong>. . . . . . . . . . . . . . . . . 239<br />

4 ABEL’s Lehrsatz V derived from DU BOIS-REYMOND’s theorem . . . . . . . . 249<br />

5 Pro<strong>of</strong> that the trigonometric coefficients cannot approach zero . . . . . . . . 256<br />

6 Rectification <strong>of</strong> the ellipse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288<br />

7 Rectification <strong>of</strong> the lemniscate . . . . . . . . . . . . . . . . . . . . . . . . . . . 290<br />

8 A.-M. LEGENDRE’S (1752–1833) reduction <strong>of</strong> elliptic integrals . . . . . . . . 295<br />

9 An important lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350<br />

10 Linear interdependence <strong>of</strong> q0, . . . , qn−1 . . . . . . . . . . . . . . . . . . . . . . 359<br />

xi


List <strong>of</strong> <strong>The</strong>orems etc.<br />

Problem 1 Division Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

Problem 2 Division Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

<strong>The</strong>orem 12 DU BOIS-REYMOND . . . . . . . . . . . . . . . . . . . . . . . . . . . 248<br />

<strong>The</strong>orem 13 Generalized Cauchy product theorem . . . . . . . . . . . . . . . . . 253<br />

Pro<strong>of</strong> 2 Pro<strong>of</strong> <strong>of</strong> lemma 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350<br />

<strong>The</strong>orem 16 Main <strong>The</strong>orem I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353<br />

<strong>The</strong>orem 17 Main <strong>The</strong>orem II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371<br />

xiii


Summary<br />

<strong>The</strong> present PhD dissertation uses the mathematics <strong>of</strong> the Norwegian N. H. ABEL<br />

(1802–1829) as a framework for describing and analyzing trends in the development<br />

<strong>of</strong> mathematics during the first half <strong>of</strong> the nineteenth century. ABEL’S mathematics<br />

is read and interpreted in its context and used to describe a fundamental change in<br />

mathematics in the early nineteenth century in which concepts replaced formulae as<br />

the basic objects <strong>of</strong> mathematics.<br />

<strong>The</strong> dissertation is structured into five parts: 1) an introductory part consisting <strong>of</strong><br />

biographical and other historical framework, 2) three descriptive parts each devoted<br />

to a particular theme analyzed from a particular discipline in ABEL’S mathematical<br />

production, and 3) a part comprising the syntheses <strong>of</strong> a general transition in mathe-<br />

matics in the early nineteenth century as seen from the perspective <strong>of</strong> ABEL’S works.<br />

Introduction. In the introductory part, ABEL’S biography is described to point out<br />

some <strong>of</strong> the formative instances in the creation <strong>of</strong> one <strong>of</strong> the important mathematicians<br />

<strong>of</strong> the first half <strong>of</strong> the nineteenth century. Because ABEL’S biography has been written<br />

repeatedly — and recently in an excellent cultural biography — the biography is only<br />

intended to locate ABEL’S production in its contexts <strong>of</strong> his life and the mathematics <strong>of</strong><br />

his time.<br />

New questions: Algebraic solubility. <strong>The</strong> first <strong>of</strong> the three studies <strong>of</strong> ABEL’S math-<br />

ematics deals with his contributions to the theory <strong>of</strong> equations. It is illustrated how<br />

ABEL was led to ask a new kind <strong>of</strong> question <strong>of</strong> the solubility <strong>of</strong> equations which<br />

would have seemed both counter intuitive and futile to mathematicians a few gen-<br />

erations before. With the foundation in works <strong>of</strong> J. L. LAGRANGE (1736–1813) and<br />

A.-L. CAUCHY (1789–1857), ABEL was able to prove that the algebraic solution <strong>of</strong><br />

general quintic equations was impossible. This result restricted the class <strong>of</strong> solvable<br />

equations and separated it from the class <strong>of</strong> all polynomial equations. In another line<br />

<strong>of</strong> research, ABEL proved that an extensive class <strong>of</strong> equations — later called <strong>Abel</strong>ian<br />

equations — were algebraically solvable. Compared with the previous result, the sol-<br />

ubility <strong>of</strong> the <strong>Abel</strong>ian equations showed that the extension <strong>of</strong> the concept <strong>of</strong> solvable<br />

equations did not collapse. Subsequently, this branch <strong>of</strong> the theory <strong>of</strong> equations be-<br />

came a question <strong>of</strong> delineating the extension <strong>of</strong> solvable equations, i.e. <strong>of</strong> drawing<br />

xv


the border between solvable and unsolvable equations by some other criterion. ABEL<br />

commenced research on this issue but had to leave it incomplete. In part II, ABEL’S<br />

research on all these issues is carefully analyzed based on the works <strong>of</strong> his main pre-<br />

decessors and contemporaries. <strong>The</strong> reception <strong>of</strong> ABEL’S research and the subsequent<br />

development <strong>of</strong> the theory is also addressed.<br />

New epistemic standards: Rigorization. ABEL’S devotion to and adaption <strong>of</strong> the<br />

rigorization movement spearheaded by CAUCHY is the topic <strong>of</strong> part III. By describing<br />

ABEL’S critical attitude towards the existing practices <strong>of</strong> rigor and his publications on<br />

the binomial theorem and a certain type <strong>of</strong> criteria <strong>of</strong> convergence, it is illustrated how<br />

the new epistemic standards were manifesting themselves in a period <strong>of</strong> rapid transi-<br />

tion in analysis. Starting with a description <strong>of</strong> the Eulerian focus on algebraic equality,<br />

it is described how CAUCHY’S new emphasis on arithmetical equality effected the cen-<br />

tral concepts <strong>of</strong> continuous functions and converget series. Furthermore, the so-called<br />

exception which ABEL presented to a theorem by CAUCHY is treated in some detail<br />

because it will later prove to be an important example in the description <strong>of</strong> the change<br />

from formula based to concept based mathematics.<br />

New objects: Elliptic and higher transcendentals. <strong>The</strong> third and final pillar <strong>of</strong> ABEL’S<br />

mathematics which is treated in the present dissertation concerns his works on ellip-<br />

tic and higher transcendentals. Selected aspects <strong>of</strong> his work are again presented and<br />

discussed from a diachronical viewpoint. In this case, special emphasis is given to the<br />

way ABEL was led to his formal inversion <strong>of</strong> elliptic integrals into elliptic functions.<br />

Furthermore, ABEL’S means <strong>of</strong> obtaining workable representations <strong>of</strong> the formally de-<br />

fined object is described. <strong>The</strong>reafter, special attention is paid to the techniques which<br />

he employed in studying transcendentals and it is illustrated how algebraic methods<br />

figured prominently in his toolbox. In the process, it is also described how his style <strong>of</strong><br />

argument <strong>of</strong>ten relied essentially on manipulations <strong>of</strong> formulae in ways which could<br />

sometimes lead to results which were true “in general”. Finally, the changing inter-<br />

nal relationships between definitions and results are illustrated by describing various<br />

ways towards a general theory <strong>of</strong> elliptic functions.<br />

Syntheses. In the ultimate part, the preceding descriptions and discussions <strong>of</strong> ABEL’S<br />

mathematics are thrown into perspective by arguing that the development <strong>of</strong> mathe-<br />

matics in the early nineteenth century can be understood as a change <strong>of</strong> paradigms:<br />

An Eulerian, formula based paradigm is contrasted with a concept based paradigm.<br />

Various aspects <strong>of</strong> ABEL’S works — including delineation problems, ABEL’S exception,<br />

and the nature <strong>of</strong> arguments which are only true “in general” — are then all inter-<br />

preted based on this transition in paradigms.<br />

xvi


Preface to the 2004 edition<br />

For this second edition, some minor changes have been made to the initial version,<br />

which was handed in on March 27, 2002 and defended on May 3 the same year. <strong>The</strong><br />

changes include a number <strong>of</strong> corrections <strong>of</strong> misprints, a revision <strong>of</strong> the layout and the<br />

figures, and the addition <strong>of</strong> a name index, a subject index, and a list <strong>of</strong> boxes. 1 Fur-<br />

thermore, I have included a new preface below, which elaborates on the general theme<br />

<strong>of</strong> dissertation, namely the analysis <strong>of</strong> NIELS HENRIK ABEL’S (1802–1829) mathemat-<br />

ics within a transition from formula based to concept based mathematics. This new<br />

“introduction” is an updated and distilled version <strong>of</strong> the lecture which I gave at the<br />

defence in May 2002.<br />

Funded by grants from the Faculty <strong>of</strong> Science (Aarhus) and from the Netherlands’<br />

Organization for Scientific Research (NWO), I have continued to elaborate on this<br />

transitional framework, producing two papers focusing on particular aspects such<br />

as critical revision, exceptions, and habituation — processes which are discussed in<br />

the present work. One <strong>of</strong> these papers is currently accepted for publication in His-<br />

toria Mathematica, (Sørensen, 2005); the other remains in the pipeline and should be<br />

submitted soon. In the future, I will continue to elaborate on the general analytical<br />

framework and its impact on analysing ABEL’S mathematics.<br />

<strong>Abel</strong>’s mathematics in the context <strong>of</strong> traditions and<br />

changes<br />

<strong>The</strong> aim <strong>of</strong> the dissertation was tw<strong>of</strong>old: to describe and analyse ABEL’S mathematics<br />

within its historical context and to draw perspectives on the general development <strong>of</strong><br />

mathematics in the early nineteenth century from the <strong>Abel</strong>ian corpus <strong>of</strong> mathematics.<br />

<strong>The</strong> analyses which are drawn from ABEL’S mathematics are — obviously — re-<br />

lated to it, and it has been my main ambition to point to general developments which<br />

shed light on ABEL’S mathematics. That is to say, I do not claim that these trends are<br />

independent <strong>of</strong> ABEL’S mathematics and could (or should) apply to other periods <strong>of</strong><br />

time, other topics <strong>of</strong> mathematics, or other mathematicians. However, it is my convic-<br />

1 I am grateful for the corrections which were pointed out to me by OLE HALD and by many other<br />

friendly people. However, if some misprints have endured, I would not be too surprised and therefore<br />

beg the forgiveness <strong>of</strong> the reader.<br />

xvii


tion that they are genuine themes, and I have sought to develop and document this<br />

position in this dissertation and in my subsequent research.<br />

<strong>The</strong> main theme, which I have idenfied and used to structure and analyse ABEL’S<br />

mathematics is one <strong>of</strong> a transition from a predominantly formula based approach to<br />

mathematics towards a more concept based one.<br />

<strong>The</strong> formula based approach to mathematics can be found most clearly in the<br />

works <strong>of</strong> such mathematicians <strong>of</strong> the eighteenth century as L. EULER (1707–1783) or<br />

A.-M. LEGENDRE (1752–1833). <strong>The</strong>se mathematicians approached analysis (including<br />

the theory <strong>of</strong> equations) in a formula based way, in which formulae were the central<br />

objects <strong>of</strong> mathematics. <strong>The</strong>ir mathematics — to a large extend — consisted <strong>of</strong> manip-<br />

ulations <strong>of</strong> formulae which produced new formulae as results.<br />

As a counterpart to the formula based approach, I introduce concept based ap-<br />

proach <strong>of</strong> the nineteenth century. This way <strong>of</strong> thinking about and doing mathemat-<br />

ics was championed by such mathematicians as G. P. L. DIRICHLET (1805–1859) and<br />

G. F. B. RIEMANN (1826–1866). <strong>The</strong>y thought mainly in terms <strong>of</strong> concepts, and their<br />

mathematics consisted <strong>of</strong> researching the relations between concepts, including repre-<br />

sentations <strong>of</strong> concepts, extensions <strong>of</strong> concepts, and the precise delineation <strong>of</strong> concepts.<br />

This framework <strong>of</strong> formula based and concept based mathematics has helped me<br />

to organise my study <strong>of</strong> ABEL’S mathematics because it explains several particular as-<br />

pects and provides an organising scheme. Thus, in the second part — on the theory<br />

<strong>of</strong> algebraic solubility — the framework provides an explanation and structuring <strong>of</strong><br />

what I term the delineation problem: finding the precise extension <strong>of</strong> the quality <strong>of</strong> solv-<br />

able equations. In the third part (rigorisation <strong>of</strong> analysis), the framework suggests that<br />

mathematicians were struggling with changing conceptions about their objects, not just<br />

the ways <strong>of</strong> manipulating them. This led to the necessity for critical revision which is<br />

also better understood within a framework <strong>of</strong> transitions. And in the fourth part (el-<br />

liptic and higher transcendentals), the framework finally suggests that habituation —<br />

coming to know new objects — was a prominent problem which played a part in the<br />

shaping <strong>of</strong> concept based mathematics.<br />

This extremely brief outline <strong>of</strong> the suggestive and explanatory powers <strong>of</strong> the frame-<br />

work is meant to justify and explain the organising principle <strong>of</strong> the dissertation. In my<br />

subsequent research, I have worked to expand more on the framework and its analyt-<br />

ical powers.<br />

Recent literature<br />

Since the dissertation was submitted in 2002, new literature on ABEL and his mathe-<br />

matics has emerged. In particular, ABEL’S Paris memoir was located — first partially,<br />

then completely; see (Del Centina, 2002; Del Centina, 2003). I am grateful to NILS<br />

VOJE JOHANSEN for providing me with a xerox copy <strong>of</strong> the famous treatise. Later,<br />

xviii


in connection with the <strong>Abel</strong> centennial in Oslo 2002, the proceedings were published<br />

(Laudal and Piene, 2004). <strong>The</strong>se contain a mathematical introduction by CHRISTIAN<br />

HOUZEL which elaborates on the previous publications <strong>of</strong> this scholar. Besides these,<br />

a number <strong>of</strong> other publications deal with ABEL’S mathematics and his topics; these<br />

include (Radl<strong>of</strong>f, 2002) which sheds interesting light on ABEL’S theory <strong>of</strong> solubility<br />

and (Pesic, 2003). In my future research and publishing on ABEL’S mathematics, these<br />

will provide good possibilities for discussions. Although this list is not complete and<br />

no use <strong>of</strong> these publications has been made in the present work, they are cited here for<br />

the convenience <strong>of</strong> the reader.<br />

xix


Preface to the 2002 edition<br />

<strong>The</strong> present PhD dissertation grew out <strong>of</strong> long held curiosity towards the multifaceted<br />

transformation which mathematics underwent during the nineteenth century. In this<br />

respect, the project is about how mathematics came to have a form recognizable to us<br />

as modern mathematics. As a pragmatic and useful tool, KIRSTI ANDERSEN — who<br />

supervised the project — suggested studying the mathematics <strong>of</strong> ABEL in order to get<br />

a firmer hold <strong>of</strong> the transition which was thereby also restricted to a shorter period<br />

in the first half <strong>of</strong> the nineteenth century. Based on detailed and <strong>of</strong>ten cumbersome<br />

studies <strong>of</strong> ABEL’S mathematics, some aspects <strong>of</strong> the transition stand out which are<br />

here described and analyzed from the perspective <strong>of</strong> concept based mathematics.<br />

Layout<br />

Quotations. Quotations are used extensively to convey the authentic arguments and<br />

thoughts. All quotations are presented in English in the text with the original included<br />

in a footnote. This method has been chosen because it increases the flow <strong>of</strong> the main<br />

text. <strong>The</strong> translations are sometimes based on existing translations; in such cases refer-<br />

ences are given. Otherwise, the translations have been made by the author. Through-<br />

out, small-caps have been reserved for names mentioned in the text. For this reason,<br />

small-caps found in quotations have been replaced by bold-face.<br />

References and footnotes. References to the bibliography are given in the footnotes<br />

and consist <strong>of</strong> the name(s) <strong>of</strong> the author(s) and the year <strong>of</strong> (original) publication. Some<br />

items are referred to through collected works or other compilations; in such cases, the<br />

bibliography contains both the publication in the collection (primary) and the original<br />

means <strong>of</strong> publication (secondary). For papers published in A. L. CRELLE’S Journal für<br />

die reine und angewandte Mathematik, the original publication is always the primary one.<br />

In all cases, page references relate to the primary method <strong>of</strong> publication mentioned in<br />

the bibliography. Thus, for instance, (N. H. <strong>Abel</strong>, 1826f, 311) refers to the first page <strong>of</strong><br />

ABEL’S binomial paper as published in CRELLE’S Journal für die reine und angewandte<br />

Mathematik whereas (L. Euler, 1760, 585) refers to the first page <strong>of</strong> EULER’S paper as<br />

printed in the Opera. In the bibliography, authors are listed alphabetically ordered<br />

by their last name. Items by the same author are listed chronologically and potential<br />

xxi


items published in the same year are separated by letters. <strong>The</strong>re are a few exceptions<br />

to these rules — mainly concerning ABEL’S publications. All collected works and a<br />

few other items have been given more illustrative names, e.g. (N. H. <strong>Abel</strong>, 1839; N.<br />

H. <strong>Abel</strong>, 1902e) which denote the first edition <strong>of</strong> ABEL’S Œuvres and the Norwegian<br />

Festschrift <strong>of</strong> 1902, respectively. Some <strong>of</strong> ABEL’S manuscripts have been dated and<br />

published posthumously; for these, the year they were written is included in brackets<br />

as in (N. H. <strong>Abel</strong>, [1828] 1839). References to ABEL’S manuscripts and notebooks are<br />

<strong>of</strong> the form (<strong>Abel</strong>, MS:351:A). Letters are refered to as (<strong>Abel</strong>→Holmboe, Kjøbenhavn,<br />

1823/06/15. In N. H. <strong>Abel</strong>, 1902a, 3–4), which denotes the letter from ABEL to B. M.<br />

HOLMBOE (1795–1850) sent from Copenhagen on June 15, 1823. I have only used<br />

published letters, and the references are given.<br />

<strong>Mathematics</strong> and notation. It has been my general ambition to unwrap and disen-<br />

tangle the mathematics presented in the dissertation to such a degree that the reader<br />

who holds no particular knowledge <strong>of</strong> the topics discussed but is familiar with math-<br />

ematical reasoning and mathematical notation should be able to benefit from the ar-<br />

guments and analyses. At the same time, it has been a high priority <strong>of</strong> mine to present<br />

the mathematics produced in the early nineteenth century in a way which respects and<br />

represents the way its creators thought about it. However, I have introduced a mini-<br />

mum <strong>of</strong> notational advances, in particular combining sums into the modern notation<br />

using the summation sign. I have also occasionally renumbered indices or replaced<br />

symbols to ease the notation. Throughout, I write Σn for the symmetric group on n<br />

symbols, which is elsewhere frequently referred to as Sn. Slightly <strong>of</strong>f-topic mathemat-<br />

ical themes have been placed in boxes shaded gray.<br />

Names and portraits. Upon first mention, historical actors are listed with their full<br />

Christian names and years <strong>of</strong> birth and death according to the Dictionary <strong>of</strong> Scientific<br />

Biography. 2 In situations where the person is not included in the Dictionary <strong>of</strong> Scien-<br />

tific Biography, other sources are employed and explicitly referred to. 3 Full names and<br />

dates <strong>of</strong> important persons are sometimes repeated in various parts. Unless other-<br />

wise noticed, all pictures stem from the history <strong>of</strong> mathematics internet archives at St.<br />

Andrews, Scotland. 4<br />

Acknowledgments<br />

My utmost gratitude extends towards my supervisor, KIRSTI ANDERSEN. For more<br />

than four years, KIRSTI has always kindly guided me and put up with my chang-<br />

ing moods. KIRSTI has read most <strong>of</strong> the chapters <strong>of</strong> the present dissertation while<br />

2 (Gillispie, 1970–80).<br />

3 Mostly (Biermann, 1988; Poggendorff, 1965; Stubhaug, 1996).<br />

4 ����������������������������������������������������������.<br />

xxii


they were under production. <strong>The</strong> endless effort which she put into her constructive<br />

criticism has made my presentation more easily understandable and sharpened my<br />

arguments considerably. For this, I am infinitely grateful.<br />

Benefitting from my supervisor’s elaborate network, I have had the opportunity to<br />

discuss aspects <strong>of</strong> my project with experts who have all been extremely forthcoming.<br />

During my visit with pr<strong>of</strong>essor H. J. M. BOS in Utrecht in November and Decem-<br />

ber 1998, the foundation for the overall structure <strong>of</strong> the present work was laid and I<br />

am greatly indebted to pr<strong>of</strong>essor BOS for his always timely suggestions and contin-<br />

ued kind interest. In May and June <strong>of</strong> 2000, I had the opportunity to visit pr<strong>of</strong>essor<br />

UMBERTO BOTTAZZINI in Palermo. Discussing my progress with one <strong>of</strong> the great-<br />

est experts in the field was an unforgettable experience and I owe a lot <strong>of</strong> inspiration<br />

to the kind counselling <strong>of</strong> pr<strong>of</strong>essor BOTTAZZINI. Closer to home, pr<strong>of</strong>essor JESPER<br />

LÜTZEN has critically and constructively sharpened my research through a number<br />

<strong>of</strong> discussions. In particular, LÜTZEN’S insightful and inspiring examination <strong>of</strong> my<br />

progress report (1999) helped me improve its contents and made me aware <strong>of</strong> inter-<br />

esting parallels also in need <strong>of</strong> consideration.<br />

In Norway, historians <strong>of</strong> mathematics have taken a kind interest in my work. In<br />

particular, I wish to thank pr<strong>of</strong>essor REINHARD SIEGMUND-SCHULTZE for his com-<br />

ments on a paper <strong>of</strong> mine; 5 the present dissertation has also benefitted from these<br />

comments. <strong>The</strong> unsurpassable biography <strong>of</strong> ABEL had been written by ARILD STUB-<br />

HAUG less than two years before I embarked on my project. 6 I deeply admire STUB-<br />

HAUG’S cultural biography and enjoyed meeting with him to discuss our common<br />

obsession. However, the greatest <strong>of</strong> my debts to him was that his work made it easier<br />

for me to focus on my project: the (almost) exclusive study <strong>of</strong> ABEL’S mathematics.<br />

For many years, I have enjoyed being part <strong>of</strong> the History <strong>of</strong> Science Department.<br />

I have had the chance to discuss my project, mathematics, science, and life in general<br />

with some very inspiring and kind people. I wish to thank the entire staff and various<br />

generations <strong>of</strong> students <strong>of</strong> the Department for their openness and support. I particu-<br />

lar, I wish to thank ANITA KILDEBÆK NIELSEN, LOUIS KLOSTERGAARD, TERESE M.<br />

O. NIELSEN, and BJARNE AAGAARD for interesting discussions from which this dis-<br />

sertation has benefitted, directly or indirectly. I am also indebted to THOMAS BRITZ<br />

for comments and corrections on part II and to my mother, KIRSTEN STENTOFT, for<br />

pro<strong>of</strong> reading and help in translating quotations into English.<br />

My thanks also go to SIGBJØRN GRINDHEIM at the manuscript collection <strong>of</strong> the<br />

university library in Oslo for providing me with copies <strong>of</strong> ABEL’S manuscripts, and<br />

to HANS ERIK JENSEN, Statsbiblioteket Aarhus University for tracing the Vienna re-<br />

view <strong>of</strong> ABEL’S impossibility pro<strong>of</strong>. 7 I also wish to thank KLAUS FROVIN JØRGENSEN<br />

for kindly taking the time to provide me with a list <strong>of</strong> the ABEL-manuscripts held in<br />

5 (Sørensen, 2002).<br />

6 (Stubhaug, 1996), translated into English (Stubhaug, 2000).<br />

7 Used in section 6.7.<br />

xxiii


the Mittag-Leffler archives in Djursholm, Sweden, 8 to UDAI VENEDEM for providing<br />

me with a list <strong>of</strong> items reviewed in BARON DE FERRUSAC’S Bulletin, and to OTTO B.<br />

BEKKEN for his interest in my work and for providing me references to the announcements<br />

by BRIOT and BOUQUET. 9<br />

It goes almost without saying that despite the generous help which I have received,<br />

any remaining mistakes, misprints, or misunderstandings are solely my responsibility.<br />

<strong>The</strong> material presented in part II constitutes the reworking <strong>of</strong> a 140-page chapter<br />

appended to my progress report (Sørensen, 1999). Aspects <strong>of</strong> the research leading up<br />

to the dissertation have been presented in various colloquia, courses and meetings.<br />

8 (I. Grattan-Guinness, 1971, 372–373).<br />

9 Touched upon in chapter 14.<br />

xxiv


Part I<br />

Introduction<br />

1


Chapter 1<br />

Introduction<br />

In the aftermath <strong>of</strong> the French Revolution <strong>of</strong> 1789, the political and scientific scenes in<br />

Paris and throughout Europe underwent radical changes. Social and educational re-<br />

forms introduced the first massive instruction in mathematics at the newly established<br />

École Polytechnique in Paris; and mathematics, itself, changed and developed into a<br />

form recognizable to modern mathematicians. In the first decades <strong>of</strong> the nineteenth<br />

century, the neo-humanist movement greatly influenced Prussian academia and as an<br />

effect, mathematics was promoted into a very prominent position in the curriculum<br />

<strong>of</strong> secondary schools. At the university level, mathematics gained a certain autonomy<br />

and started to evolve along a distinctly theoretical line with less focus on applications<br />

and mathematical physics.<br />

<strong>The</strong> present work centers on one <strong>of</strong> the main innovative figures in mathematics in<br />

the 1820s, the Norwegian NIELS HENRIK ABEL (1802–1829), and describes his con-<br />

tribution to and influence on the fermentation <strong>of</strong> the mathematical discipline in the<br />

early nineteenth century. Born at the periphery <strong>of</strong> the mathematical world and with<br />

a life-span <strong>of</strong> less than 27 years, ABEL nevertheless contributed importantly to the<br />

disciplines which he studied. <strong>The</strong> overall outline <strong>of</strong> this presentation is recapitulated<br />

in the following three sections which introduce ABEL’S pr<strong>of</strong>essional background and<br />

training, the mathematics <strong>of</strong> his works, and the treated themes <strong>of</strong> development in<br />

mathematics in the first half <strong>of</strong> the nineteenth century. Throughout, ABEL’S mathe-<br />

matics is seen in its mathematical context, and the influences <strong>of</strong> mathematicians such<br />

as A.-L. CAUCHY (1789–1857), C. F. GAUSS (1777–1855), J. L. LAGRANGE (1736–1813),<br />

and A.-M. LEGENDRE (1752–1833) is traced and described. This approach provides<br />

a background for discussing aspects <strong>of</strong> continuity and transformation in mathematics<br />

as can be envisioned from ABEL’S works.<br />

3


4 Chapter 1. Introduction<br />

1.1 <strong>The</strong> historical and geographical setting <strong>of</strong> ABEL’s<br />

life<br />

ABEL lived in a politically turbulent time during which his birthplace, Finnøy, be-<br />

longed to three different monarchies. When ABEL was born in 1802, Finnøy belonged<br />

to the Danish-Norwegian twin monarchy but in the wake <strong>of</strong> the Napoleonic Wars, the<br />

province <strong>of</strong> Norway was ceded to Sweden after a short spell <strong>of</strong> independence. Edu-<br />

cation in the twin monarchy was centered in Copenhagen, and only in 1813 was the<br />

university in Christiania (now Oslo) opened. <strong>The</strong> scientific climate was beginning to<br />

ripe, but mathematics was not studied at a high level.<br />

As was common practice for the sons <strong>of</strong> a minister, ABEL attended cathedral school<br />

in Christiania and soon got the young B. M. HOLMBOE (1795–1850) as a mathematics<br />

teacher. HOLMBOE was the first to notice ABEL’S affinity for and skills in mathe-<br />

matics and they began to study the works <strong>of</strong> the masters in special private lessons.<br />

In 1821, after graduating from the cathedral school, ABEL enrolled at the university<br />

but continued his private studies <strong>of</strong> the masters <strong>of</strong> mathematics. In 1824, he applied<br />

for a travel grant to go to the Continent and he embarked on his European tour in<br />

1825. It brought him to Berlin and Paris where he had the opportunities to meet some<br />

<strong>of</strong> the most prominent mathematicians <strong>of</strong> the time and frequent the well equipped<br />

continental libraries. More importantly, ABEL came into contact with A. L. CRELLE<br />

(1780–1855) in Berlin. CRELLE became ABEL’S friend and published most <strong>of</strong> ABEL’S<br />

works in the Journal für die reine und angewandte Mathematik, which he founded in 1826.<br />

When ABEL returned to Norway in 1827 he found himself without a permanent job<br />

and with no family fortune to cover his expenses, he took up tutoring in mathematics.<br />

He had suffered from a lung infection during his tour, and in 1829 he succumbed to<br />

tuberculosis.<br />

ABEL’S geographical background thus dictated his approach to mathematics; it<br />

forced him to study the masters and advance in isolation to do original work. In his<br />

short life span he carefully studied works <strong>of</strong> the previous generation and went be-<br />

yond those. During the months abroad, he came into contact with the newest trends<br />

in mathematics, and was immediately engaged in new research. Almost all his pub-<br />

lications were written during or after the tour. <strong>The</strong> presentation <strong>of</strong> ABEL’S historical<br />

and biographical background serves to provide a framework for tracing ideas, influ-<br />

ences, and connections in his work.<br />

1.2 <strong>The</strong> mathematical topics involved<br />

ABEL’S mathematical production span a wide range <strong>of</strong> topics and theories which were<br />

important in the early nineteenth century. His primary contributions are universally<br />

considered to be in the theory <strong>of</strong> algebraic solubility <strong>of</strong> equations, in the rigorization <strong>of</strong>


1.2. <strong>The</strong> mathematical topics involved 5<br />

the theory <strong>of</strong> series, and in the study <strong>of</strong> elliptic functions and higher transcendentals.<br />

However, some <strong>of</strong> ABEL’S other works (published or unpublished) also have their<br />

place in the contexts <strong>of</strong> other disciplines, e.g. in the solution <strong>of</strong> particular types <strong>of</strong><br />

differential equations, in the prehistory <strong>of</strong> fractional calculus, in the theory <strong>of</strong> integral<br />

equations, or in the study <strong>of</strong> generating functions. However, to keep the focus <strong>of</strong> the<br />

present dissertation, these “minor” topics have not been included and emphasis is put<br />

on equations, series, and elliptic and higher transcendentals.<br />

<strong>The</strong>ory <strong>of</strong> equations. <strong>The</strong> essentials <strong>of</strong> mathematics in the eighteenth century come<br />

down to the work <strong>of</strong> a single brilliant mind, L. EULER (1707–1783). Through a lifelong<br />

devotion to mathematics which spanned most <strong>of</strong> the century preceding the French<br />

Revolution, EULER reformulated the core <strong>of</strong> mathematics in pr<strong>of</strong>ound ways. Inspired<br />

by his attempts to demonstrate that any polynomial <strong>of</strong> degree n had n roots (the so-<br />

called Fundamental <strong>The</strong>orem <strong>of</strong> Algebra), EULER introduced another important math-<br />

ematical question: Can any root <strong>of</strong> a polynomial be expressed in the coefficients by<br />

radicals, i.e. by using only basic arithmetic and the extraction <strong>of</strong> roots? This ques-<br />

tion concerned the algebraic solubility <strong>of</strong> equations and to EULER it was almost self-<br />

evident. However, mathematicians strove to supply even the evident with pro<strong>of</strong>, and<br />

LAGRANGE developed an elaborated theory <strong>of</strong> equations based on permutations to<br />

answer the question. Though a believer in generality in mathematics, LAGRANGE<br />

came to recognize that the effort required to solve just the general fifth degree equa-<br />

tion might exceed the humanly possible. In LAGRANGE’S native country, Italy, an<br />

even more radical perception <strong>of</strong> the problem had emerged; around the turn <strong>of</strong> the cen-<br />

tury, P. RUFFINI (1765–1822) had made public his conviction that the general quintic<br />

equation could not be solved by radicals and provided his claim with lengthy pro<strong>of</strong>s.<br />

ABEL’S first and lasting romance with mathematics was with this topic, the theory<br />

<strong>of</strong> equations; his first independent steps out <strong>of</strong> the shadows <strong>of</strong> the masters were un-<br />

successful ones when in 1821 he believed to have obtained a general solution formula<br />

for the quintic equation. Provoked by a request to elaborate his argument, he realized<br />

that it was in err, and by 1824 he gave a pro<strong>of</strong> that no such solution formula could<br />

exist. <strong>The</strong> pro<strong>of</strong>, which was based on a detailed theory <strong>of</strong> permutations and a clas-<br />

sification <strong>of</strong> possible solutions, reached world (i.e. European) publicity in 1826 when<br />

it appeared in the first volume <strong>of</strong> CRELLE’S Journal für die reine und angewandte Math-<br />

ematik. But as so <strong>of</strong>ten happens, solving one question only leads to posing another.<br />

Realizing that the general fifth degree equation could not be solved by radicals, ABEL<br />

set out on a mission to investigate which equations could and which equations could<br />

not be solved algebraically. Despite his efforts — which were soon distracted to an-<br />

other subject — ABEL had to leave it to the younger French mathematician E. GALOIS<br />

(1811–1832) to describe the criteria for algebraic solubility.


6 Chapter 1. Introduction<br />

Elliptic functions. Since the emergence <strong>of</strong> the calculus toward the end <strong>of</strong> the sev-<br />

enteenth century, the mathematical discipline <strong>of</strong> analysis had been able to treat an<br />

increasing number <strong>of</strong> curves. In his textbook Introductio in analysin infinitorum <strong>of</strong> 1748,<br />

EULER elevated the concept <strong>of</strong> function to the central object <strong>of</strong> analysis. Concrete func-<br />

tions were studied through their power series expansions and the brilliant calculator<br />

EULER obtained series expansions for all known functions including the trigonomet-<br />

ric and exponential ones. However, EULER did not stop there but ventured into the<br />

territory <strong>of</strong> unknown functions <strong>of</strong> which he tried to get hold. One important type <strong>of</strong><br />

function which analysis had struggled to treat on a par with the rest was the so-called<br />

elliptic integrals which can measure the length <strong>of</strong> an arc <strong>of</strong> an ellipse.<br />

Mathematicians such as EULER and LEGENDRE felt and spoke <strong>of</strong> an unsatisfactory<br />

restriction <strong>of</strong> analysis because it was only able to treat a limited set <strong>of</strong> elementary tran-<br />

scendental functions. Admitting new functions into analysis meant obtaining the kind<br />

<strong>of</strong> knowledge about these functions that would allow them to be given as answers. If a<br />

function today is nothing more than a mapping <strong>of</strong> one set into another, the knowledge<br />

<strong>of</strong> a function then included tabulation <strong>of</strong> values, series expansions and other represen-<br />

tations, differential and integral relations, functional relations, and much more.<br />

When ABEL made elliptic integrals his main research topic, much knowledge con-<br />

cerning these objects had already been established. An algebraic approach which had<br />

pr<strong>of</strong>ound influence on ABEL was GAUSS’ study <strong>of</strong> the division problem for the circle<br />

(construction <strong>of</strong> regular n-gons) in the Disquisitiones arithmeticae. 1 GAUSS had hinted<br />

that his approach could be applied to the lemniscate integral, a particularly simple<br />

case <strong>of</strong> elliptic integrals, and ABEL took it upon himself to provide the claim with<br />

a pro<strong>of</strong>. By a new idea, soon to be praised as one <strong>of</strong> the greatest in analysis, ABEL<br />

inverted the study <strong>of</strong> elliptic integrals into the study <strong>of</strong> elliptic functions: Instead <strong>of</strong><br />

considering the value <strong>of</strong> an integral to be a function <strong>of</strong> its upper limit, he considered<br />

the upper limit to be a function <strong>of</strong> the value <strong>of</strong> the integral (compare arcsin and sin).<br />

Through formal substitutions and certain addition formulae, ABEL obtained elliptic<br />

functions <strong>of</strong> a complex variable. By this inversion <strong>of</strong> focus, ABEL managed to place<br />

the entire theory <strong>of</strong> elliptic integrals on a new and much more fertile footing. Fueled<br />

by a fierce competition between ABEL and the German mathematician C. G. J. JA-<br />

COBI (1804–1851), the new theory gained almost immediate momentum and became<br />

one <strong>of</strong> the central pillars <strong>of</strong> and main motivations for nineteenth century advances in<br />

mathematics.<br />

Although ABEL had presented the crucial idea <strong>of</strong> inverting elliptic integrals into<br />

elliptic functions, his impact on the further development <strong>of</strong> the theory stemmed as<br />

much from a vast generalization <strong>of</strong> the addition formulae presented in a paper which<br />

he handed in to the Parisian Académie des Sciences in 1826 (not published until 1841).<br />

In this paper, ABEL treated an even broader class <strong>of</strong> integrals generalizing the elliptic<br />

1 (C. F. Gauss, 1801).


1.3. <strong>The</strong>mes from early nineteenth-century mathematics 7<br />

ones and — again using primarily algebraic methods — proved more general versions<br />

<strong>of</strong> the addition theorems. <strong>The</strong> quest <strong>of</strong> later mathematicians to reapply ABEL’S dar-<br />

ing inversion <strong>of</strong> elliptic integrals to this broader class <strong>of</strong> integrals led to much <strong>of</strong> the<br />

important development in complex analysis and topology in the nineteenth century.<br />

Rigor. Although the theory <strong>of</strong> equations was closest to ABEL’S heart, and the the-<br />

ory <strong>of</strong> elliptic functions brought him fame in the nineteenth century, his mathemat-<br />

ical legacy remembered in the twentieth century is just as much about his intense<br />

reception <strong>of</strong> CAUCHY’S new rigor. Picking up from LAGRANGE’S theory <strong>of</strong> functions,<br />

CAUCHY had placed concepts such as continuity and convergence in the foreground<br />

and founded these concepts on a new interpretation <strong>of</strong> limits. Equally importantly,<br />

CAUCHY had shown a way <strong>of</strong> working with these concepts to deduce properties <strong>of</strong><br />

classes <strong>of</strong> objects (e.g.. continuous functions or convergent series) rather than explicit,<br />

<strong>of</strong>ten lengthy, studies <strong>of</strong> specific objects.<br />

In a memorable and <strong>of</strong>ten quoted letter dated 1826 (first published 1839), ABEL<br />

expressed his conversion to Cauchy-ism and gave the new rigor its dogmatic mani-<br />

festo. Apparently more radical than CAUCHY himself, ABEL helped determine the<br />

formulation <strong>of</strong> the new rigor through his interpretative readings <strong>of</strong> CAUCHY. In the<br />

process <strong>of</strong> re-founding analysis on rigorous grounds, central concepts were specified<br />

and changed (stretched) to an extent where they included elements whose behavior<br />

was deemed abnormal. <strong>The</strong> encounter and resolution <strong>of</strong> these abnormalities, excep-<br />

tions as they were <strong>of</strong>ten called, was an integrated part <strong>of</strong> the rigorization process; such<br />

exceptions — which a modern reader would consider counter examples — shed inter-<br />

esting light on the role and use <strong>of</strong> concepts in mathematics in the early nineteenth<br />

century.<br />

1.3 <strong>The</strong>mes from early nineteenth-century mathematics<br />

<strong>The</strong> early nineteenth century marks a period <strong>of</strong> transition and fermentation in mathe-<br />

matics which involves most layers <strong>of</strong> the discipline, external as well as internal. With<br />

the boundaries fixed, say, between 1790 and 1840, a definite change in the way mathe-<br />

matics was performed and presented is evident; research mathematicians began work-<br />

ing in institutions set up for instruction in mathematics and started presenting their<br />

results in pr<strong>of</strong>essional periodicals with substantial circulation. However, the change<br />

even effected the internal core <strong>of</strong> the discipline: how mathematics was done, what<br />

mathematics was, and which mathematical questions were interesting. Gradually,<br />

concepts and relations between concepts took an increasingly central position in math-<br />

ematics research; although the concern for concrete objects never ceased completely.


8 Chapter 1. Introduction<br />

Concept based mathematics. Concepts such as function, continuity <strong>of</strong> functions, irre-<br />

ducibility <strong>of</strong> equations, and convergence <strong>of</strong> series attained central importance in math-<br />

ematical research in the transitional period. CAUCHY’S contribution to the rigoriza-<br />

tion <strong>of</strong> the calculus laid as much in applying technical definitions <strong>of</strong> concepts to prove<br />

theorems as with providing the definitions, themselves. Generalization in the 1820s<br />

turned the attention from specific objects to classes <strong>of</strong> objects, which were then in-<br />

vestigated. This shift <strong>of</strong> attention toward collections <strong>of</strong> individual objects had a very<br />

direct influence on the style <strong>of</strong> presenting mathematical research. In the ‘old’ tradition,<br />

mathematical papers could easily be concerned with explicit derivations (calculations)<br />

pertaining to single mathematical objects. Although this presentational style far from<br />

ceased to fill periodicals, a less explicit style gained impetus in the first half <strong>of</strong> the<br />

nineteenth century. By deriving properties <strong>of</strong> classes instead <strong>of</strong> individual objects,<br />

the arguments became more abstract and <strong>of</strong>ten more comprehensible by lowering the<br />

load <strong>of</strong> calculations and simplifying the mathematical notation. <strong>The</strong> transition is ev-<br />

ident in ABEL’S works which show deep traces <strong>of</strong> the calculation based approach to<br />

doing mathematics as well as being markedly conceptual at times; his 1826 paper on<br />

the binomial theorem is a fascinating mixture <strong>of</strong> both approaches.<br />

Abstract definitions and coming to know mathematical objects. In many develop-<br />

ing fields <strong>of</strong> mathematics in the early nineteenth century, new concepts were specified<br />

by the use <strong>of</strong> abstract definitions based on previous pro<strong>of</strong>s, intentions, and intuitions.<br />

In the approach which I term concept based mathematics, the concepts were defined in<br />

the modern sense that there is nothing more to a concept than its definition. However,<br />

when abstract definitions determine the extent <strong>of</strong> a concept, representations and de-<br />

marcation criteria are required in order to get hold <strong>of</strong> properties <strong>of</strong> objects, and this<br />

quest for understanding, coming to know, the objects is an important aspect <strong>of</strong> early<br />

nineteenth century mathematics. In many ways, analogies may be drawn to the ef-<br />

fort <strong>of</strong> coming to know geometrical objects, e.g. curves, in the seventeenth century. To<br />

mathematicians <strong>of</strong> the seventeenth century, a curve meant more than any single given<br />

piece <strong>of</strong> information. In particular, an equation (or a method <strong>of</strong> constructing any num-<br />

ber <strong>of</strong> points on the curve) was not considered sufficient to accept the curve as known.<br />

Similarly, in the nineteenth century, knowledge <strong>of</strong> an elliptic function meant more<br />

than just a formal definition and included various representations, basic properties,<br />

and even tabulation <strong>of</strong> values.<br />

<strong>The</strong> question <strong>of</strong> coming to know a mathematical object relates to the problem <strong>of</strong><br />

accepting the object as solutions to problems. <strong>The</strong> reduction <strong>of</strong> properties <strong>of</strong> curves to<br />

questions pertaining those basic curves which were considered well known was im-<br />

portant in the seventeenth century. However, certain properties were not expressible<br />

in basic curves (or functions) but required higher transcendentals such as elliptic inte-<br />

grals. Thus, much <strong>of</strong> EULER’S research on elliptic integrals in the eighteenth century<br />

can be seen as an effort to make these integrals basic in the sense <strong>of</strong> acceptable so-


1.4. Reflections on methodology 9<br />

lutions to problems. This research program was continued and reformulated in the<br />

nineteenth century during which the foundations, definitions, and framework <strong>of</strong> el-<br />

liptic functions underwent repeated revolutions.<br />

Critical revision. <strong>The</strong> critical mode <strong>of</strong> thought, rooted in the Enlightenment, had a<br />

pr<strong>of</strong>ound impact on mathematics. Together with the demand for wider instruction in<br />

mathematics, the critical attitude brought about a deeply sceptical reading <strong>of</strong> the mas-<br />

ters which focused on the foundations. In geometry, some mathematicians began to<br />

believe in the possibility <strong>of</strong> a non-Euclidean version, and in analysis, the long-standing<br />

problem <strong>of</strong> the foundation <strong>of</strong> the calculus was made an important mathematical research<br />

topic. 2<br />

CAUCHY’S definition <strong>of</strong> the central concept <strong>of</strong> limits was itself a novelty, but <strong>of</strong><br />

equal importance was the outlook for a concept based version <strong>of</strong> the calculus. CAUCHY’S<br />

new foundation for the calculus was arithmetical and introduced the arithmetical con-<br />

cept <strong>of</strong> equality. In the wake <strong>of</strong> the change <strong>of</strong> foundations <strong>of</strong> the calculus, certain<br />

objects and methods could no longer be allowed into analysis, and it became a quest<br />

to prop up parts <strong>of</strong> the mathematical complex recently made insecure. In particular,<br />

CAUCHY had to abolish from analysis all divergent series which had formerly been<br />

interpreted by a formal concept <strong>of</strong> equality. However, divergent series had provided<br />

new insights to mathematicians which they were reluctant to abandon and it became<br />

a legitimate, albeit difficult, mathematical problem to investigate how problematic<br />

or outright unjust procedures had led to correct results. Resolution <strong>of</strong> this problem<br />

laid in further specification <strong>of</strong> concepts involved and a heightened awareness <strong>of</strong> the<br />

procedures employed in arguments. For instance, by the mid-nineteenth century, the<br />

unreflective interchange <strong>of</strong> orders <strong>of</strong> limit processes had been identified as problem-<br />

atic and concepts such as absolute and uniform convergence had been introduced and<br />

put to use in theorems and pro<strong>of</strong>s.<br />

1.4 Reflections on methodology<br />

<strong>The</strong> present study aims at illustrating important conceptual developments in math-<br />

ematics which took place in the first decades <strong>of</strong> the nineteenth century. In order to<br />

introduce a focus on the many-faceted aspects <strong>of</strong> these developments, the mathemat-<br />

ical production <strong>of</strong> ABEL has been taken as a starting point. In the present section,<br />

some <strong>of</strong> the methodological choices and considerations involved in the project are<br />

briefly discussed. It is not my ambition to present a coherent theoretical framework<br />

for historical enquiries but rather to make some considerations explicit and open for<br />

discussion.<br />

2 See e.g. (Grabiner, 1981b).


10 Chapter 1. Introduction<br />

1.4.1 Diachronic descriptions<br />

It is within the framework <strong>of</strong> mathematics developed by EULER and cultivated in the<br />

eighteenth century that ABEL’S production is rightfully seen. Although his mathe-<br />

matics has been perceived as a very important step forward in a linear development,<br />

ABEL’S mathematical ideas were rooted in the previous attitude and style; and many<br />

<strong>of</strong> the famed new trends are only barely recognizable in his work. <strong>The</strong>refore, anachro-<br />

nisms and teleological conceptions have to be dismissed in favor <strong>of</strong> a diachronic, more<br />

hermeneutic approach. In the process <strong>of</strong> tracing and describing this historical evolu-<br />

tion <strong>of</strong> mathematical content, it is <strong>of</strong> the utmost importance that ABEL’S works be<br />

studied within their contemporary framework — their mathematical context.<br />

Each <strong>of</strong> the main theories outlined above with which ABEL was involved is re-<br />

viewed with the purpose <strong>of</strong> illustrating how they were effected by currents <strong>of</strong> mathe-<br />

matical change in the early nineteenth century. To do so, ABEL’S mathematics is pre-<br />

sented and discussed based on a contextualized reading which emphasizes ABEL’S<br />

own methods and tools. To place these in their proper historical contexts, the theories<br />

and results will be traced back into the eighteenth century in search <strong>of</strong> the inspirations<br />

and their progressions in the nineteenth century will be followed. In the theory <strong>of</strong><br />

equations, for instance, the works <strong>of</strong> EULER on the fundamental theorem <strong>of</strong> algebra<br />

will be briefly introduced; more emphasis will be given to the works <strong>of</strong> LAGRANGE<br />

and GAUSS which served as the direct inspirations for ABEL and most <strong>of</strong> his contem-<br />

poraries in the field. <strong>The</strong>n, in the nineteenth century, special emphasis is given to<br />

those works which share their inspirations with the works <strong>of</strong> ABEL, in this case the<br />

works <strong>of</strong> RUFFINI and GALOIS. For each disciplinary theme, a sketch <strong>of</strong> the further<br />

development after the initial decades <strong>of</strong> the nineteenth century is then given in or-<br />

der to illustrate how the ideas and currents which were barely discernible in the first<br />

decades came to play very important roles in the conceptions <strong>of</strong> mathematicians. <strong>The</strong><br />

descriptions <strong>of</strong> the ensuing histories also serve to illustrate how ABEL’S works were<br />

valued and received by the following generations.<br />

Besides, it has been a secondary aim <strong>of</strong> the present study to make ABEL’S authentic<br />

mathematical thought available to the mathematically trained reader who is not famil-<br />

iar with early nineteenth century technicalities. In order to understand and evaluate<br />

ABEL’S role in the formation <strong>of</strong> modern mathematics, this presentation will always<br />

favor the original source over any modern approach.<br />

<strong>The</strong> setting <strong>of</strong> ABEL’S mathematics within the general view <strong>of</strong> mathematics ex-<br />

pressed by EULER will also be manifest in another way as a chronological mark. It has<br />

been necessary to trace many <strong>of</strong> the ideas and methods <strong>of</strong> ABEL’S mathematics back<br />

to the middle <strong>of</strong> the eighteenth century, but they might be even older. However, as<br />

this is a work on ABEL’S mathematics, such hypothesis will rarely be made explicit<br />

and EULER will be attributed things, which he did do — perhaps not as the first.


1.4. Reflections on methodology 11<br />

1.4.2 Philosophical theories and their applicability<br />

Philosophical theories enter the framework <strong>of</strong> the present study only rather implicitly.<br />

Written with the utopian goal <strong>of</strong> being an “account <strong>of</strong> things which happened”, the<br />

outlines <strong>of</strong> a certain perception <strong>of</strong> concepts such as change and transformation is never-<br />

theless discernible. <strong>The</strong> internal (and external) structure <strong>of</strong> scientific change has been<br />

subjected to many philosophical investigations over the past decades. In the present<br />

context, however, two <strong>of</strong> the founding theories have served as inspirations; those <strong>of</strong><br />

Kuhnian paradigms and revolutions developed for the sciences in general and those <strong>of</strong><br />

Lakatosian dialectics which were developed explicitly for mathematics and illustrated<br />

by examples from the nineteenth century. <strong>The</strong>se two philosophical positions are so<br />

well established within the history <strong>of</strong> science community that only a brief presenta-<br />

tion is given with an emphasis on their applicability to the present study.<br />

T. S. KUHN (1922–1996), paradigms, crises, and revolutions. In his very influential<br />

monograph <strong>The</strong> structure <strong>of</strong> scientific revolutions <strong>of</strong> 1962, 3 KUHN advocated an explana-<br />

tion <strong>of</strong> the dynamics <strong>of</strong> scientific change. <strong>The</strong> total mental and physical entourage <strong>of</strong><br />

a science at a given time was encompassed in the notion <strong>of</strong> paradigms. Paradigms are<br />

abruptly replaced through revolutions which are the responses to crises brought about<br />

by a compilation <strong>of</strong> anomalies inexplicable within the ruling paradigm. Once a revolu-<br />

tion has taken place, a new paradigm is introduced and communication between two<br />

distinct paradigms (e.g. over which paradigm to prefer) becomes irrational or extra-<br />

scientific. KUHN’S original model — although deeper than the present description —<br />

was a simplistic one which was amended and extended by numerous studies follow-<br />

ing its publication. Here, on the other hand, it will not be taken to serve as a complete<br />

model but rather as inspiration and terminological framework used to capture impor-<br />

tant aspects.<br />

As a model <strong>of</strong> mathematical change in the early nineteenth century, the Kuhnian<br />

system <strong>of</strong>fers some obvious advantages; the important position given to anomalies in<br />

bringing about crises and revolutions was further extended in the works <strong>of</strong> I. LAKA-<br />

TOS (1922–1974) (see below). However, as has been emphasized by many philosophers<br />

and historians <strong>of</strong> mathematics, no overthrow <strong>of</strong> knowledge seems to occur in mathematics;<br />

thus no truly Kuhnian revolutions seem possible in the mathematical realm. 4<br />

<strong>The</strong> remaining notions <strong>of</strong> the Kuhnian conceptual framework such as paradigms,<br />

anomalies and crises are, however, applicable and useful in the description and anal-<br />

ysis <strong>of</strong> mathematical change, even if KUHN’S dynamics are not always appropriate.<br />

In the first sections <strong>of</strong> part II, for instance, it is illustrated how a mathematical theory<br />

came into being by a change <strong>of</strong> focus (a paradigmatic change) which shifted emphasis<br />

to questions <strong>of</strong> solubility. <strong>The</strong> theme will recur even more distinctly in part III which<br />

3 (Kuhn, 1962).<br />

4 See (Gillies, 1992).


12 Chapter 1. Introduction<br />

documents ABEL’S role and position in the most Kuhnian <strong>of</strong> changes in mathematics<br />

during the early nineteenth century: the complete reformulation <strong>of</strong> analysis according<br />

to CAUCHY’S new program <strong>of</strong> arithmetical rigor.<br />

LAKATOS and the extension <strong>of</strong> concepts. Further philosophical inspiration is taken<br />

from LAKATOS’ Pro<strong>of</strong>s and Refutations published as a series <strong>of</strong> articles in 1963–64 and as<br />

a book in 1976. 5 LAKATOS described the dynamics <strong>of</strong> mathematical change in terms <strong>of</strong><br />

a dialectic between pro<strong>of</strong>s and counter examples by means <strong>of</strong> pro<strong>of</strong> revisions. In the main<br />

part <strong>of</strong> the Pro<strong>of</strong>s and Refutations, LAKATOS explained his theory by exhibiting a ratio-<br />

nally reconstructed development <strong>of</strong> the Eulerian polyhedral formula; in appendices,<br />

he further illustrated the theory by exhibiting applications to other concepts including<br />

the development <strong>of</strong> the concept <strong>of</strong> uniform convergence (see part III).<br />

LAKATOS saw the process <strong>of</strong> pro<strong>of</strong> as central to the mathematical endeavor. Incor-<br />

porating into mathematics a version <strong>of</strong> K. R. POPPER’S (1902–1994) falsificationism,<br />

LAKATOS described mathematical change as a continued revision <strong>of</strong> pro<strong>of</strong>s to reflect<br />

objections raised by counter examples. LAKATOS classified counter examples as either<br />

local (refuting only part <strong>of</strong> a pro<strong>of</strong>, but not the overall statement) or global (refuting<br />

the overall statement, but not necessarily any identifiable part <strong>of</strong> the pro<strong>of</strong>).<br />

Counter examples could, in LAKATOS’ description, be constructed from existing<br />

pro<strong>of</strong>s by a process <strong>of</strong> concept stretching by which a partially defined concept was rede-<br />

fined in an extended version which — although possibly more precise — encompassed<br />

instances (objects) not covered by the previous — <strong>of</strong>ten more intuitive — version <strong>of</strong> the<br />

concept.<br />

In response to such falsifications (refutations) by counter examples, LAKATOS sug-<br />

gested various strategies for refining the pro<strong>of</strong>s. A naive approach would try to ex-<br />

plain the counter examples away, either by arguing that they were too pathological<br />

to be taken seriously or (more interestingly) by restricting the theorem to a narrower<br />

domain for which it was believed to surely valid; the latter approach was named ex-<br />

ception barring by LAKATOS. A more fruitful response to the refutation by counter<br />

examples — and the one which LAKATOS’ philosophy dogmatized — was the method<br />

<strong>of</strong> pro<strong>of</strong> analysis which took the counter examples more seriously. By carefully ana-<br />

lyzing the counter example and the pro<strong>of</strong> which it refuted, pro<strong>of</strong> analysis produced a<br />

new pro<strong>of</strong> in which a refuted lemma was replaced by an unrefuted one which might<br />

cause an alteration <strong>of</strong> the overall statement. Thus, LAKATOS suggested, theorems were<br />

produced which had very explicit assumptions and were very hard to refute.<br />

Just as was the case with the Kuhnian model, LAKATOS’ model — in all its general-<br />

ity — is <strong>of</strong>ten found inadequate to describe the actual historical development <strong>of</strong> math-<br />

ematics. On the other hand, LAKATOS’ model <strong>of</strong>fers some further concepts which<br />

<strong>of</strong>ten ease the description and analysis <strong>of</strong> past events. Most importantly, LAKATOS’<br />

5 (Lakatos, 1976). A good description <strong>of</strong> LAKATOS’ life and philosophy can be found in (Larvor, 1998).


1.4. Reflections on methodology 13<br />

description <strong>of</strong> counter examples and the role that they play in mathematical change<br />

elaborates the role played by anomalies in the Kuhnian model and suggests a more<br />

refined view on the status <strong>of</strong> a mathematical theory in crisis.<br />

<strong>The</strong> Lakatosian theory <strong>of</strong> mathematical evolution is present as background through-<br />

out; it will surface sporadically in parts II–IV and become important again in the final,<br />

more analytical part V.<br />

M. EPPLE’S epistemic configurations. Quite recently, EPPLE has suggested the no-<br />

tion <strong>of</strong> epistemic configurations in order to be able to discuss change in mathematics<br />

in another context. 6 In EPPLE’S analysis, epistemic configurations consist <strong>of</strong> epistemic<br />

objects and epistemic techniques and are manipulated in mathematical workshops. <strong>The</strong><br />

concept <strong>of</strong> epistemic objects encompasses the immaterial objects with which mathe-<br />

matics deals. <strong>The</strong>se are manipulated and investigated by a number <strong>of</strong> methods <strong>of</strong><br />

producing (or obtaining) mathematical knowledge; these methods are the epistemic<br />

techniques. <strong>The</strong> precise applicability and range <strong>of</strong> EPPLE’S concepts and their use-<br />

fulness in historical analysis is not the primary objective here. Instead, as with the<br />

inspirations <strong>of</strong> KUHN and LAKATOS, I have taken the liberty <strong>of</strong> using EPPLE’S terms<br />

to ease the analysis and discussion <strong>of</strong> what I believe to be a fundamental change in<br />

mathematics in the early nineteenth century: the change from formula based to concept<br />

based mathematics which is addressed in chapter 21.<br />

1.4.3 Existing literature<br />

Being one <strong>of</strong> the important mathematicians <strong>of</strong> the nineteenth century, ABEL’S person<br />

and his mathematics have been subjected to study for a multitude <strong>of</strong> different rea-<br />

sons. A few general trends <strong>of</strong> the literature on ABEL can pr<strong>of</strong>itably be identified at<br />

this point. 7 At the relevant places in the subsequent parts, references are given to the<br />

secondary literature which is listed in the bibliography.<br />

ABEL in the history <strong>of</strong> mathematics literature. In the pr<strong>of</strong>essional literature in the<br />

history <strong>of</strong> mathematics, ABEL is <strong>of</strong>ten mentioned in order to illustrate one or more <strong>of</strong><br />

the following aspects:<br />

1. ABEL’S life story is invoked to illustrate the conditions <strong>of</strong> young mathematicians<br />

two centuries ago. This aspect is closely related to the biographies treated below.<br />

2. ABEL’S letters from Paris are used to illuminate how the confrontation with<br />

CAUCHY’S new rigor brought about a radical change. For instance, U. BOTTAZZ-<br />

INI (⋆1947) quotes in extenso from these letters in his comprehensive account <strong>of</strong><br />

the evolution <strong>of</strong> analysis in the nineteenth century. 8<br />

6 (Epple, 2000).<br />

7 For a thematic listing <strong>of</strong> the ABEL literature, see also (Sørensen, 2002).<br />

8 (Bottazzini, 1986).


14 Chapter 1. Introduction<br />

3. <strong>The</strong> modern highlights <strong>of</strong> ABEL’S production, e.g. the binomial theorem or the<br />

insolubility <strong>of</strong> the quintic, are described to shed some light on the evolution <strong>of</strong><br />

the theories and the involved concepts.<br />

4. ABEL’S mathematics is described per se in order to give a presentation <strong>of</strong> his<br />

production. Very good examples include articles by P. L. M. SYLOW (1832–1918)<br />

in the ABEL centennial memorial volume and the second edition <strong>of</strong> the collected<br />

works. 9<br />

<strong>The</strong> present study incorporates all these approaches to give a comprehensive overview<br />

<strong>of</strong> ABEL’S mathematical production as well as positioning it within a broader frame<br />

describing themes <strong>of</strong> mathematical change in the period.<br />

Two other types <strong>of</strong> studies treating the life and works <strong>of</strong> ABEL delineate them-<br />

selves: biographies and interpretations.<br />

Biographies — scientific or not. As should become clear in the next chapter, ABEL’S<br />

biography includes all the components <strong>of</strong> a truly romantic biography <strong>of</strong> a misun-<br />

derstood genius who rose from the dust to become a nobility <strong>of</strong> mathematics. Such<br />

biographies have been written; 10 but more interestingly, biographies have also been<br />

written which serve a purpose <strong>of</strong> their own. <strong>The</strong> first biographies appeared as obituar-<br />

ies written by ABEL’S friends soon after his death. Of primary importance in describ-<br />

ing ABEL’S mathematics are the obituaries written by HOLMBOE and CRELLE which<br />

include first hand descriptions <strong>of</strong> ABEL’S mathematical work. 11 Although a larger<br />

number <strong>of</strong> biographies could be listed, the most widely circulated and very well re-<br />

searched twentieth century biography was written by Ø. ORE (1899–1968); 12 it has<br />

been used mainly to help set the chronology straight. <strong>The</strong> human and cultural as-<br />

pects <strong>of</strong> ABEL’S life has most recently been very carefully researched and described<br />

by A. STUBHAUG (⋆1948) who meticulously sets the cultural scene <strong>of</strong> early nineteenth<br />

century Norway and Europe. 13 STUBHAUG’S biography has relieved me <strong>of</strong> any obli-<br />

gation to produce biographical news concerning ABEL’S person; the biography which<br />

is provided in chapter 2 serves merely to set the framework <strong>of</strong> the subsequent chap-<br />

ters. It is my hope that the present study <strong>of</strong> ABEL’S mathematics will complement<br />

STUBHAUG’S book on his life and environment to produce a picture <strong>of</strong> ABEL’S person<br />

and his mathematics.<br />

Renderings <strong>of</strong> ABEL’S work in modern theories. By the very nature <strong>of</strong> mathemat-<br />

ics, mathematical knowledge seems to accumulate and only change its presentational<br />

9 (N. H. <strong>Abel</strong>, 1881; L. Sylow, 1902).<br />

10 E.g. (Bell, 1953).<br />

11 (A. L. Crelle, 1829b; Holmboe, 1829).<br />

12 (Ore, 1954; Ore, 1957).<br />

13 (Stubhaug, 1996; Stubhaug, 2000).


1.4. Reflections on methodology 15<br />

form or its internal relations within mathematical structures. For this reason, mathe-<br />

maticians <strong>of</strong>ten hope to find inspiration in the works <strong>of</strong> their predecessors. Frequently,<br />

this leads to the publication <strong>of</strong> modernized versions <strong>of</strong> historical pro<strong>of</strong>s. By itself, this<br />

practice is very good as long as the author and the community recognize that it is pre-<br />

cisely a revisited pro<strong>of</strong> or theorem and precisely not a diachronic description <strong>of</strong> that<br />

pro<strong>of</strong> or theorem within its contemporary structure.<br />

Such revisits to ABEL’S production are most frequently made to his theory <strong>of</strong> alge-<br />

braic solubility <strong>of</strong> equations, more precisely to his pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the quin-<br />

tic equation. 14 ABEL’S other main contributions attract less attention; the binomial<br />

theorem because it has become an integral part <strong>of</strong> basic mathematical knowledge, and<br />

the <strong>Abel</strong>ian <strong>The</strong>orem (see part IV) because its original form has been surpassed and the<br />

result has been recast in a different theory.<br />

As should now be clear, the methodology <strong>of</strong> the present approach can be summa-<br />

rized thus: A diachronic reading <strong>of</strong> the original sources <strong>of</strong> ABEL’S mathematics with<br />

the purpose <strong>of</strong> analyzing themes <strong>of</strong> mathematical change in the early nineteenth cen-<br />

tury, in particular the rise <strong>of</strong> concept based mathematics.<br />

14 See e.g. (R. Ayoub, 1982; Radl<strong>of</strong>f, 1998).


Chapter 2<br />

Biography <strong>of</strong> NIELS HENRIK ABEL<br />

<strong>The</strong> life <strong>of</strong> NIELS HENRIK ABEL (1802–1829) was not always a happy one. Born in<br />

a time <strong>of</strong> national upheaval and into a family with few provisions against the hard<br />

times, his actions were always restricted by pecuniary concerns. A melancholic, he<br />

preferred to be surrounded by people, but due to his shy and modest nature he felt se-<br />

cure only with a score <strong>of</strong> friends including his elder brother, his sister, his mathematics<br />

teacher B. M. HOLMBOE (1795–1850), Mrs. C. A. B. HANSTEEN (1787–1840) — ABEL’S<br />

benefactor and the wife <strong>of</strong> his university pr<strong>of</strong>essor, and his mentor in Berlin A. L.<br />

CRELLE (1780–1855). ABEL fell in love with C. KEMP (1804–1862) in 1823 and they<br />

were engaged the following year. Unfortunately, ABEL’S position never became se-<br />

cure enough for them to marry. <strong>The</strong> last years <strong>of</strong> ABEL’S life were spent in uncertainty<br />

with hopes <strong>of</strong> a more stable future either at home or abroad. When he died, ABEL’S<br />

mathematical star was still rising, and years would pass before the world knew exactly<br />

how bright it had been.<br />

<strong>The</strong> short and yet very creative life <strong>of</strong> ABEL has caught the interest <strong>of</strong> many bi-<br />

ographers. Confined within the romantic period and exhibiting distinctly romantic<br />

features itself, the biographers have <strong>of</strong>ten focused on ABEL’S poverty and contempo-<br />

rary lack <strong>of</strong> acknowledgment; 1 both features frequently found in the romanticization<br />

<strong>of</strong> mathematicians and scientists. Another genre <strong>of</strong> biography has constructed and<br />

researched a controversy with C. G. J. JACOBI (1804–1851) over the priority <strong>of</strong> the<br />

inversion <strong>of</strong> elliptic integrals. 2 <strong>The</strong> most recent and excellent biography written by<br />

A. STUBHAUG (⋆1948) has taken a different angle, describing and bringing to life the<br />

cultural and political context in which ABEL lived and which is so important for Norwegian<br />

self-image. 3<br />

As STUBHAUG’S book is such a convincing description <strong>of</strong> the cultural and bio-<br />

graphical background <strong>of</strong> ABEL’S life, the present biography serves only to provide a<br />

self-contained presentation <strong>of</strong> the temporal framework in which ABEL’S mathemati-<br />

1 See e.g. (Bell, 1953) or, more soberly, (Ore, 1950; Ore, 1954; Ore, 1957).<br />

2 E.g. (Bjerknes, 1880; Bjerknes, 1885; Bjerknes, 1930). This debate was also the subject <strong>of</strong> (Koenigsberger,<br />

1879).<br />

3 (Stubhaug, 1996), translated into English in (Stubhaug, 2000).<br />

17


18 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

Figure 2.1: NIELS HENRIK ABEL (1802–1829)<br />

cal production was localized. All facts presented here have been taken from existing<br />

literature, primarily HOLMBOE’S obituary, C. A. BJERKNES’ (1825–1903) biographies,<br />

and STUBHAUG’S contextual biographical study. 4 References to these works will not<br />

always be made explicit. I will deliberately desist from giving detailed analyses or<br />

speculations concerning the personality and private life <strong>of</strong> ABEL except where sup-<br />

ported by the biographies and ABEL’S correspondence.<br />

2.1 Childhood and education<br />

ABEL was born as the second son into an incumbent’s family on 5 August 1802.<br />

ABEL’S father, S. G. ABEL (1772–1820), himself the son <strong>of</strong> a minister, had been ed-<br />

ucated in Copenhagen and received a call to the rural parish <strong>of</strong> Finnøy in 1800 (see<br />

figure 2.2).<br />

Also in 1800, SØREN GEORG married A. M. SIMONSEN (1781–1846), the daughter<br />

<strong>of</strong> a wealthy merchant, and together they had six children; five boys and a girl. In<br />

1804, SØREN GEORG took over the more lucrative parish <strong>of</strong> Gjerstad from his father<br />

who had died the year before. Nevertheless, due to the family increase, the costs <strong>of</strong><br />

educating the children, a nationalist sentiment to contribute to the founding <strong>of</strong> the<br />

university, and the troubled times for the nation, the ABEL-family remained without<br />

4 (Holmboe, 1829), (Bjerknes, 1880; Bjerknes, 1885; Bjerknes, 1930), (Stubhaug, 1996).


2.1. Childhood and education 19<br />

Figure 2.2: <strong>The</strong> southern part <strong>of</strong> Norway with Christiania and Finnøy marked. See<br />

also (Stubhaug, 2000, 136).<br />

fortune — a situation which only deteriorated after SØREN GEORG died in 1820 leav-<br />

ing his widow a thirty year commitment <strong>of</strong> financial donations to the university in<br />

Christiania.<br />

During his years in Copenhagen, SØREN GEORG had been influenced by the re-<br />

formist education and he took care <strong>of</strong> the primary education <strong>of</strong> his sons himself. Ac-<br />

cordingly, the focus was on the catechism and skills in reading, writing, and basic<br />

arithmetic. In 1815, when NIELS HENRIK was 13, he and his elder brother H. M. ABEL<br />

(1800–1842) were sent to the Cathedral School in Christiania. <strong>The</strong>re, they were taught<br />

classical and modern languages, as well as arithmetic and geometry. ABEL passed his<br />

examen artium in 1821 with first grades in the mathematical disciplines, second grade<br />

in French, and only third grades in German, Latin, and Greek. 5 In the lower classes,<br />

ABEL demonstrated no particular affinity for the mathematical disciplines but was a<br />

fair student in all subjects. However, this situation dramatically changed to the better<br />

due to an unfortunate event which took place in 1817–18.<br />

In November <strong>of</strong> 1817, H. P. BADER (1790–1819), ABEL’S mathematics teacher, phys-<br />

ically molested one <strong>of</strong> his pupils who later died from this hands-on approach to math-<br />

ematics education. BADER, who had a record <strong>of</strong> hot temper and violent teaching<br />

methods, was excused from his teaching obligations, and HOLMBOE stepped in as<br />

5 (N. H. <strong>Abel</strong>, 1902d, 3)


20 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

Figure 2.3: BERNT MICHAEL HOLMBOE (1795–1850)<br />

a replacement in 1818. 6 HOLMBOE, who was ABEL’S senior by only 7 years, had been<br />

educated at the cathedral school himself and had attended mathematics courses un-<br />

der S. RASMUSSEN (1768–1850) at the university. HOLMBOE soon noticed a special<br />

talent for mathematics in ABEL and they began reading extra-curricular mathematics<br />

together.<br />

2.2 “Study the masters”<br />

Soon, HOLMBOE began to realize what mathematical talent he had at hand and by<br />

the autumn <strong>of</strong> 1818, HOLMBOE urged ABEL to study the important works in mathe-<br />

matics on his own. In one <strong>of</strong> his notebooks, ABEL pursued calculations inspired by<br />

P.-S., MARQUIS DE LAPLACE’S (1749–1827) use <strong>of</strong> generating functions. Between all<br />

the calculations, he noted in the margin:<br />

“If one wants to know what one should do to obtain a result in more conformity<br />

with Nature one should consult the works <strong>of</strong> the famous Laplace where this<br />

theory is exposed with the most clarity and to an extent in accordance with the<br />

importance <strong>of</strong> the subject. It is also easy to see that a theory written by M. Laplace<br />

must be much superior to any other written by less bright mathematicians. By the<br />

way it seems to me that if one wants to progress in mathematics one should study<br />

the masters and not the pupils.” 7<br />

6 For details, see (Stubhaug, 1996, 172–174).


2.2. “Study the masters” 21<br />

HOLMBOE’S list <strong>of</strong> the masters whom they studied together sheds interesting light<br />

on the mathematical literature <strong>of</strong> the early nineteenth century as seen from the pe-<br />

riphery. 8 In the eyes <strong>of</strong> the two Norwegians, the two most influential writers were<br />

L. EULER (1707–1783) and J. L. LAGRANGE (1736–1813); but the list also included<br />

S. F. LACROIX (1765–1843), L. B. FRANCOEUR (1773–1849), S.-D. POISSON (1781–<br />

1840), C. F. GAUSS (1777–1855), and J. G. GARNIER (1766–1840). In the following<br />

paragraphs, the influences from these authors are outlined.<br />

EULER. EULER and HOLMBOE studied algebra and calculus from EULER’S works;<br />

ABEL’S first independent adventures into creative mathematics were greatly inspired<br />

by the great Swiss mathematician. Although the sources are not very explicit, it is be-<br />

yond doubts that ABEL studied EULER’S Introductio in analysin infinitorum [Introduc-<br />

tion to the infinite analysis] <strong>of</strong> 1748. 9 To what extent ABEL also knew <strong>of</strong> EULER’S other<br />

publications including his papers in the transactions <strong>of</strong> the St. Petersburg Academy<br />

is left for speculation; we shall return to the question when we see examples <strong>of</strong> EU-<br />

LER’S — possibly indirect — influence on ABEL in parts II and IV.<br />

<strong>The</strong> big four. LAGRANGE, LACROIX, POISSON, and GAUSS all belong to the heavy-<br />

weight division <strong>of</strong> mathematics in the late eighteenth century with massive and im-<br />

portant works on the calculus and algebra. Although a writer <strong>of</strong> very influential text-<br />

books on the calculus, 10 LAGRANGE mainly inspired ABEL through his work on the<br />

theory <strong>of</strong> equations which redefined the viewpoint from which this theory was to be<br />

attacked. 11 LACROIX’ effort laid more in organization and presentation than in cre-<br />

ative research; his three volume textbook on the calculus, Traité de calcul différentiel et<br />

intégral [Treatise on differential and integral calculus], ran multiple editions beginning<br />

in 1797–1800. 12 In the Traité, LACROIX presented an survey <strong>of</strong> the calculus based on<br />

the research <strong>of</strong> his contemporaries and picking up a variety <strong>of</strong> approaches and foun-<br />

dations from different authors.<br />

<strong>The</strong> lesser souls. <strong>The</strong> two authors in HOLMBOE’S list who today are lesser known,<br />

FRANCOEUR and GARNIER, both wrote textbooks on mathematics which found wide<br />

circulation toward the end <strong>of</strong> the eighteenth century. ABEL almost certainly studied<br />

7 “Si l’on veut savoir comment on doit faire pour parvenir à un resultat plus conforme à la nature il<br />

faut consulter l’ouvrage du celebre Laplace où cette theorie est exposée avec la plus grande clarté<br />

et dans une extension convenable à l’importance de la matière. Il est en outre aisé de voir que une<br />

theorie ecrite par M. Laplace doit être bien superieure à toute autre donnée des geometres d’une<br />

claire inferieure. Au reste il me parait que si l’on veut faire des progres dans les mathématiques il<br />

faut étudier les maitres et non pas les écoliers.” (<strong>Abel</strong>, MS:351:A, 79, marginal note).<br />

8 (Holmboe, 1829, 335).<br />

9 (L. Euler, 1748).<br />

10 E.g. (Lagrange, 1813).<br />

11 (Lagrange, 1770–1771).<br />

12 (Lacroix, 1797; Lacroix, 1798; Lacroix, 1800).


22 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

FRANCOEUR’S Cours complet de mathematiques pures [Complete course on pure mathe-<br />

matics], 13 which in two volumes dedicated to the emperor <strong>of</strong> Russia, introduced arith-<br />

metic, geometry, algebra, and differential and integral calculus. <strong>The</strong> textbook had been<br />

translated into German by one <strong>of</strong> ABEL’S mentors, C. F. DEGEN (1766–1825), 14 but<br />

since ABEL only came to master German during his tour 1825–27, he probably studied<br />

the French original.<br />

Besides writing textbooks on algebraic analysis (LAGRANGE’S approach to the cal-<br />

culus), in 1807, GARNIER translated EULER’S Vollständige Einleitung zur Algebra [Com-<br />

plete introduction to algebra] <strong>of</strong> 1770 into French. 15 With his limited knowledge<br />

<strong>of</strong> German, it is doubtful whether ABEL read EULER in the original language, but<br />

through the translations by LAGRANGE or GARNIER or even through FRANCOEUR’S<br />

complete course on pure mathematics, ABEL became acquainted with the elementary<br />

parts <strong>of</strong> contemporary mathematical knowledge, in particular the solution <strong>of</strong> cubic<br />

and bi-quadratic equations.<br />

2.2.1 An alleged solution formula<br />

While still in grammar school, ABEL approached one <strong>of</strong> the most prestigious prob-<br />

lems <strong>of</strong> contemporary mathematics: the search for an algebraic formula expressing<br />

the solution <strong>of</strong> the quintic equation. Since the Western Renaissance, similar formulae<br />

for equations <strong>of</strong> the first four degrees had been known. In 1821, ABEL believed to<br />

have found a closed algebraic expression solving the next case: the general quintic.<br />

He wrote down his result and showed it to his teacher HOLMBOE, who took it to C.<br />

HANSTEEN (1784–1873), one <strong>of</strong> the two pr<strong>of</strong>essors in science at the Christiania Univer-<br />

sity. 16 HANSTEEN, who together with HOLMBOE were among the few people in Nor-<br />

way competent enough to have a chance <strong>of</strong> understanding ABEL’S tedious argument,<br />

took it to the University’s collegium academicum. <strong>The</strong> collegium took note <strong>of</strong> ABEL’S ar-<br />

gument and wanted to make it public to a broader mathematical audience. However,<br />

as the young Norwegian state was itself without means <strong>of</strong> such a publication with a<br />

wide circulation, ABEL’S paper was sent to pr<strong>of</strong>essor DEGEN in Copenhagen with the<br />

hope that it be published in the transactions <strong>of</strong> Royal Danish Academy <strong>of</strong> Sciences and<br />

Letters. DEGEN’S assessment proved to have a pr<strong>of</strong>ound influence on ABEL’S career.<br />

Upon reception <strong>of</strong> the paper, DEGEN scrutinized ABEL’S solution, and DEGEN’S re-<br />

sponse to HANSTEEN is the only existing written source <strong>of</strong> ABEL’S adventure. <strong>The</strong>re,<br />

DEGEN requested an elaboration, a numerical example, and a rewriting <strong>of</strong> the manu-<br />

script for the other members <strong>of</strong> Videnskabernes Selskab to be able to read it. To ABEL,<br />

the refusal to immediately publish his result must have been disappointing. However,<br />

13 (L.-B. Francœur, 1809).<br />

14 (L.-B. Francœur, 1815). In 1839, it was again translated into German by KÜLP.<br />

15 It has been translated into English as (L. Euler, 1972).<br />

16 Very unfortunately, ABEL’S supposed solution has not survived and only speculative reconstructions<br />

can be suggested.


2.2. “Study the masters” 23<br />

as he sat down to provide the details, he must have realized that DEGEN had spared<br />

him a humiliating entry onto the mathematical scene. Before 1824, ABEL realized that<br />

no algebraic solution formula could be found for the general quintic, and thus that his<br />

solution had been flawed and his search in vain. ABEL never sent an elaboration to<br />

DEGEN, never published in the Transactions <strong>of</strong> the Royal Danish Academy <strong>of</strong> Science, and<br />

when ABEL and DEGEN eventually met in person two years after their initial corre-<br />

spondence, ABEL had other things on his mind.<br />

2.2.2 A student at the young university<br />

When ABEL enrolled at the university in 1821, the university was still in its constitu-<br />

tional phase. Founded in 1811 and opened in 1813 as only the third university in the<br />

twin monarchy (after Copenhagen and Kiel), the Christiania university initially only<br />

<strong>of</strong>fered degrees in theology, law, medicine, and philosophy. <strong>The</strong> study <strong>of</strong> science and<br />

mathematics was subsumed under the philosophical faculty and no course <strong>of</strong> studies<br />

led to any degree in the sciences. Thus, when ABEL enrolled, his determination to<br />

study mathematics defied the existing structure <strong>of</strong> academic qualification. He must<br />

have hoped that his extraordinary talents alone would be enough to secure him a fu-<br />

ture in academia.<br />

During his years at the university, ABEL attended lectures by the two pr<strong>of</strong>essors in<br />

mathematics and astronomy, RASMUSSEN and HANSTEEN. <strong>The</strong> mathematical lectures<br />

were primarily on elementary mathematics, spherical geometry, and applications to<br />

astronomy, and ABEL had soon learned all he could from these courses. As a comple-<br />

ment, he continued studying the works <strong>of</strong> the masters <strong>of</strong> mathematics. In 1823, ABEL<br />

came across the Disquisitiones arithmeticae <strong>of</strong> GAUSS, 17 which provided him with a rich<br />

source <strong>of</strong> inspiration and problems for his own research. Itself an immensely impor-<br />

tant work in the theory <strong>of</strong> numbers, the Disquisitiones arithmeticae influenced ABEL in<br />

two other fields: the theory <strong>of</strong> equations and the rectification <strong>of</strong> the lemniscate.<br />

During his years as a student, ABEL held a free room and board at the Regentsen,<br />

a student residence for the most needy students. Until he was given a stipend from<br />

the State in 1824, he was financially supported by some <strong>of</strong> the University pr<strong>of</strong>essors,<br />

including RASMUSSEN and HANSTEEN. 18<br />

First publications in Magazin for Naturvidenskaberne. In 1823, pr<strong>of</strong>essor HAN-<br />

STEEN, together with two fellow pr<strong>of</strong>essors at the university, tried to amend the lack <strong>of</strong><br />

Norwegian periodicals in natural science with the creation <strong>of</strong> the Magazin for Naturv-<br />

idenskaberne [Magazine for the natural sciences]. Its aim was to convey Norwegian<br />

research in the sciences to the educated lay audience and provide an emerging group<br />

<strong>of</strong> young scientists with a forum for publication.<br />

17 (C. F. Gauss, 1801).<br />

18 (Stubhaug, 1996, 244–245).


24 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

In the first issue <strong>of</strong> the Magazin, the 21 year old ABEL had his first publication.<br />

Inspired by EULER’S Institutiones calculi differentialis and A.-M. LEGENDRE’S (1752–<br />

1833) Exercises de calcul integral, ABEL solved two problems in the integral calculus. 19<br />

In the same year, a second publication by ABEL dealing with the theory <strong>of</strong> elimina-<br />

tion was published in the Magazin. 20 On this occasion, HANSTEEN found it necessary<br />

to add an introduction in which he — summoning G. GALILEI (1564–1642), I. NEW-<br />

TON (1642–1727), and C. HUYGENS (1629–1695) — argued that mathematics even in<br />

its purest form was within the scope <strong>of</strong> the magazine devoted to the natural sciences.<br />

Taking into account the limited circulation <strong>of</strong> the Magazin and HANSTEEN’S efforts to<br />

make ABEL’S work acceptable to the audience, it is doubtful how much ABEL gained<br />

from these publications. But to a young man — still nothing but a studiosus — getting<br />

his name on printed paper must have been a great satisfaction.<br />

When ABEL’S publications in the Magazin were first noticed, it was for all the<br />

wrong reasons. In 1824, ABEL published some computations pertaining to the influ-<br />

ence <strong>of</strong> the Moon on the movement <strong>of</strong> a pendulum. 21 This problem fitted nicely into a<br />

research project concerning the magnetic field <strong>of</strong> the Earth with which HANSTEEN was<br />

immensely involved. Upon HANSTEEN’S request, the paper was sent to H. C. SCHU-<br />

MACHER (1784–1873) in Altona, who edited the journal Astronomische Nachrichten [As-<br />

tronomical intelligencer], for possible republication therein. However, SCHUMACHER<br />

realized that ABEL had made a computational error which had led him to estimate the<br />

influence <strong>of</strong> the Moon to be ten times stronger than was rightfully supported. SCHU-<br />

MACHER refused to publish the paper, and a correction was subsequently inserted in<br />

the Magazin. 22<br />

2.2.3 Visiting Copenhagen<br />

In 1823, ABEL for the first time left Norway to go on his first educational tour. Sup-<br />

ported privately by pr<strong>of</strong>essor RASMUSSEN, ABEL traveled to Copenhagen to visit and<br />

discuss with the mathematicians there.<br />

<strong>Mathematics</strong> in the capital. <strong>The</strong> mathematical milieu in Copenhagen was not com-<br />

pletely different from the one in Christiania. 23 In 1823, the mathematical pr<strong>of</strong>ession<br />

was centered around the university and the academy <strong>of</strong> science; six years later, Den<br />

Polytekniske Læreanstalt [<strong>The</strong> polytechnic college] 24 was opened. Despite the university<br />

19 (N. H. <strong>Abel</strong>, 1823)<br />

20 (ibid.)<br />

21 (N. H. <strong>Abel</strong>, 1824c).<br />

22 (N. H. <strong>Abel</strong>, 1824a).<br />

23 <strong>The</strong> mathematical milieu in Copenhagen in the first half <strong>of</strong> the nineteenth century has been the subject<br />

<strong>of</strong> a subsequent study by the author in the Danish History <strong>of</strong> Science project undertaken at the<br />

History <strong>of</strong> Science Department at the University <strong>of</strong> Aarhus. <strong>The</strong> interested reader may also wish to<br />

consult the dissertation (in Danish) by KURT RAMSKOV (Ramskov, 1995, chapter 1) or STUBHAUG’S<br />

book (Stubhaug, 2000, chapter 31) for information.<br />

24 Today Danmarks Tekniske Universitet [Technical University <strong>of</strong> Denmark].


2.2. “Study the masters” 25<br />

Figure 2.4: CARL FERDINAND DEGEN (1766–1825)<br />

in Copenhagen being older and larger than the one in Christiania, the faculty <strong>of</strong> phi-<br />

losophy had only two pr<strong>of</strong>essorships in mathematics (and one in astronomy) which<br />

in 1823 were held by DEGEN and E. G. F. THUNE (1785–1829).<br />

Together, DEGEN and THUNE had meant a change <strong>of</strong> generations in the mathe-<br />

matical milieu at the university when they were appointed 1813 and 1815. 25 DEGEN,<br />

being the more creative <strong>of</strong> the two, had gained international reputation with a publi-<br />

cation <strong>of</strong> tables for the solution <strong>of</strong> Pellian equations in 1817. 26 To the mathematicians<br />

in Christiania — as to the general Norwegian public — Copenhagen was still to some<br />

extent considered the intellectual and cultural capital <strong>of</strong> the country even after the<br />

separation in 1814. Thus, by virtue <strong>of</strong> its former colonial power, the mathematical<br />

talents <strong>of</strong> DEGEN, and the circulation and status <strong>of</strong> the publications <strong>of</strong> the Royal Dan-<br />

ish Academy <strong>of</strong> Sciences and Letters, Copenhagen was the obvious choice for a first<br />

foreign trip for a young and promising Norwegian mathematician.<br />

<strong>The</strong> maturing mathematician. ABEL’S first two works in the sciences estimated im-<br />

portant enough to receive wider circulation by Norwegian scholars, the alleged solu-<br />

tion <strong>of</strong> the quintic and his computations concerning the magnetic field <strong>of</strong> the Earth,<br />

were both caught in the review system <strong>of</strong> the time. In the incipient scientific milieu in<br />

25 THUNE was originally appointed as pr<strong>of</strong>essor <strong>of</strong> astronomy 1815 before he transferred to the pr<strong>of</strong>essorship<br />

<strong>of</strong> mathematics 1819.<br />

26 (C. F. Degen, 1817)


26 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

Norway, the means <strong>of</strong> publication were limited and advanced knowledge <strong>of</strong> the sci-<br />

ences was confined to a few men. <strong>The</strong>refore, foreign experts were invited to judge —<br />

and possibly publish — ABEL’S first papers. DEGEN made reservations concerning the<br />

completeness <strong>of</strong> ABEL’S research on the solution <strong>of</strong> the quintic and refused to present<br />

ABEL’S paper to the Danish Academy until he had seen the method applied to a nu-<br />

merical example. <strong>The</strong> numerical example was explicitly requested as a lapis lydius —<br />

a test <strong>of</strong> correctness — and it is not improbable that the confrontation with an explicit,<br />

difficult problem was what later led ABEL to realize his being in error.<br />

<strong>The</strong> HANSTEEN-DEGEN period. In his historical introduction to the centennial memo-<br />

rial volume, 27 E. B. HOLST (1849–1915) has emphasized the influence <strong>of</strong> HANSTEEN<br />

and DEGEN on ABEL’S research 1821–24. P. L. M. SYLOW’S (1832–1918) analysis seems<br />

applicable on at least two levels: topics and methods. On the topical level, the two pro-<br />

tagonists exerted contrary influences. Responding to HANSTEEN’S suggestion, ABEL<br />

had briefly worked on a physical problem; either because <strong>of</strong> this failed encounter or<br />

because <strong>of</strong> a personal inclination, he subsequently focused exclusively on working<br />

within pure mathematics. Following an advice <strong>of</strong> DEGEN, ABEL ventured into the<br />

theory <strong>of</strong> integration in the tradition <strong>of</strong> EULER and LEGENDRE. 28 Being an impor-<br />

tant theory <strong>of</strong> the eighteenth century left open for further developments, DEGEN had,<br />

himself, spent some time studying elliptic integrals. But there is no real evidence to<br />

suggest that DEGEN could have foreseen what his new disciple would do for the disci-<br />

pline. Although their interests differed, both HANSTEEN and DEGEN were trained in<br />

the typical eighteenth century mathematical literature including the men whom ABEL<br />

considered his masters at the time. Formal manipulations and physical applicability<br />

were considered positive aspects <strong>of</strong> the approaches <strong>of</strong> EULER and his mid-eighteenth<br />

century contemporaries. <strong>The</strong> DEGEN-HANSTEEN period marks the end <strong>of</strong> ABEL’S<br />

youthful encounters with the formal approach to analysis and at the same time marks<br />

the beginning <strong>of</strong> a period <strong>of</strong> intense study <strong>of</strong> the theory <strong>of</strong> higher transcendentals<br />

which would be ABEL’S masterpiece when judged by his contemporaries.<br />

2.3 <strong>The</strong> European tour<br />

After ABEL returned to Christiania from his first trip to Copenhagen, he soon real-<br />

ized that there was little more for him to gain while isolated in the limited Norwegian<br />

mathematical community. In 1824, ABEL applied with the support <strong>of</strong> the pr<strong>of</strong>essors<br />

HANSTEEN and RASMUSSEN for a travel grant from the university. ABEL’S primary<br />

aim was to visit to the mathematical capital <strong>of</strong> his time, Paris. <strong>The</strong>re, in Paris, math-<br />

ematics had been institutionalized and cultivated to the highest level in the wake <strong>of</strong><br />

27 (Holst, 1902, 22).<br />

28 In part IV, the influence on ABEL <strong>of</strong> these mathematicians will be traced, documented, and analyzed.


2.3. <strong>The</strong> European tour 27<br />

the French Revolution. Later in the application procedure, Göttingen, the seat <strong>of</strong> the<br />

German champion <strong>of</strong> mathematics GAUSS was added to the travel plan.<br />

In his first period <strong>of</strong> creative mathematical production, 1823–24, while inspired<br />

by HANSTEEN and DEGEN (see above), ABEL devoted his attention to the theory <strong>of</strong><br />

integration in the tradition <strong>of</strong> EULER’S Institutiones calculi integralis and LEGENDRE’S<br />

Traité de calcul integral. By 1824, ABEL had documented his “exceptional abilities in the<br />

mathematical sciences” 29 an example <strong>of</strong> which HANSTEEN presented to the collegium<br />

in the form <strong>of</strong> a manuscript. ABEL had hoped that — besides the travel grant — the<br />

collegium would support publication <strong>of</strong> the result which he believed had international<br />

importance and could serve as a door opener on his tour. However, the collegium<br />

decided to only support a travel grant for ABEL to go to the Continent. <strong>The</strong>re was only<br />

one condition; the grant was only to begin in 1825; until then ABEL had to prepare by<br />

studying the “learned languages”, which in particular meant French. 30<br />

2.3.1 Objectives and plans<br />

In the application to the collegium, sources to the contemporary Norwegian rank-<br />

ing <strong>of</strong> the mathematical centres may be found. <strong>The</strong> choices were mainly made by<br />

ABEL’S benefactors, the mathematics pr<strong>of</strong>essors HANSTEEN and RASMUSSEN. When<br />

they first proposed sending ABEL abroad they suggested that he should go to “the<br />

places abroad where the most distinguished mathematicians <strong>of</strong> our time are located,<br />

perhaps primarily to Paris”. 31 In ABEL’S <strong>of</strong>ficial application to the King, he wrote:<br />

“After I have thus in this country by the use <strong>of</strong> the available tools sought to<br />

approach the erected goal, it would be very beneficial to me to acquaint myself,<br />

during a stay abroad at different universities in particular in Paris where so many<br />

distinguished mathematicians are located, with the newest creations in the science<br />

[mathematics] and enjoy the guidance <strong>of</strong> those men who in our time have brought<br />

it [mathematics] to such a remarkable height.” 32<br />

Only <strong>of</strong> July 4th 1825 did the idea <strong>of</strong> sending ABEL to Göttingen enter into the ap-<br />

plication when the collegium applied for the grant to be made effective. ABEL submit-<br />

ted a more detailed travel plan which RASMUSSEN was supposed to comment upon<br />

to the collegium; this plan is no longer extant. 33 Two months later, on September 7,<br />

ABEL embarked on his European tour.<br />

29 HANSTEEN’S opinion as expressed in the accommodating letter from the collegium academicum to<br />

the department <strong>of</strong> the church (N. H. <strong>Abel</strong>, 1902d, 7).<br />

30 (ibid., 12).<br />

31 (ibid., 7).<br />

32 “Efter at jeg saaledes her i Landet ved de her forhaanden værende Hjelpemidler har stræbet at<br />

nærme mig det foresatte Maal, vilde det være mig særdeles gavnligt ved et Ophold i Udlandet ved<br />

forskjellige Universiteter, især i Paris hvor saa mange i udmærkede Mathematikere findes, at blive<br />

bekjendt med de nyeste Frembringelser i Videnskaben og nyde de Mænds Veiledning som i vor<br />

Tidsalder have bragt den til en saa betydelig Høide.” (ibid., 13).<br />

33 (ibid., 20, footnote).


28 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

2.3.2 ABEL in Berlin<br />

Figure 2.5: AUGUST LEOPOLD CRELLE (1780–1855)<br />

After a brief stop to visit DEGEN in Copenhagen and another to visit SCHUMACHER in<br />

Hamburg, 34 ABEL’S first extended stay was in Berlin. He had not had any real plans<br />

<strong>of</strong> going to Berlin, but went there together with a group <strong>of</strong> friends, all <strong>of</strong> whom belong<br />

to the founding generation <strong>of</strong> Norwegian scientists, in particular in geology, and who<br />

were all on educational tours <strong>of</strong> Europe. ABEL wrote to HANSTEEN in his first report<br />

from the European tour where he obviously had to defend spending time in Berlin:<br />

“You may have wondered why I first traveled to Germany; I did so partly<br />

because I could thereby stay with friends and partly because I would be less likely<br />

not to make the most <strong>of</strong> my time since I can leave Germany at any time to go to<br />

Paris which should be the most important place for me.” 35<br />

In the same letter, ABEL described the acquaintance which he had made with one <strong>of</strong><br />

the local mathematicians, the pr<strong>of</strong>essional administrator Geheimrat [Privy Councilor]<br />

CRELLE whom he had been told about by the Copenhagen mathematician H. G. V.<br />

34 Part <strong>of</strong> ABEL’S obligation was to carry out experiments for HANSTEEN measuring the Earth’s magnetic<br />

field at different locations.<br />

35 “De har maaskee forundret Dem over hvorfor jeg først reiste til Tyskland; men dette gjorde jeg deels<br />

fordi jeg da kom til at leve sammen med Bekjendtere deels fordi jeg da var mindre udsat for ikke at<br />

anvende Tiden paa den bedste Maade, da jeg kan forlade Tyskland hvert Øjeblik det skal være for<br />

at reise til Paris, som bør være det vigtigste Sted for mig.” (<strong>Abel</strong>→Hansteen, Berlin, 1825/12/05.<br />

N. H. <strong>Abel</strong>, 1902a, 9–10).


2.3. <strong>The</strong> European tour 29<br />

SCHMIDTEN (1799–1831). Introducing himself in his stuttering German, ABEL called<br />

upon the busy Geheimrath soon after his arrival in Berlin. Initially, their conversation<br />

was staggering; only when CRELLE enquired what ABEL had already read, did he re-<br />

alize that he was facing a young man quite versed in the modern mathematics. ABEL<br />

presented CRELLE with a copy <strong>of</strong> his pamphlet on the insolubility <strong>of</strong> the quintic equa-<br />

tion, and CRELLE expressed his difficulty understanding the argument. In due time,<br />

ABEL would present CRELLE with an elaborated argument which would gain world<br />

wide circulation.<br />

“I am extremely pleased that I happened to go to Germany and in particular<br />

Berlin before I came to Paris; since — as you may have learned from my letter to<br />

Hansteen — I have made the splendid acquaintance with Geheimrath Crelle.” 36<br />

<strong>The</strong> founding <strong>of</strong> the Journal für die reine und angewandte Mathematik. Communi-<br />

cation <strong>of</strong> mathematics in the nineteenth century underwent rapid change. In the days<br />

<strong>of</strong> EULER, mathematics had been confined to pr<strong>of</strong>essional amateurs and academicians<br />

who communicated their results either privately in correspondence, in monographs,<br />

or in the periodicals <strong>of</strong> the academies. Only in the beginning <strong>of</strong> the nineteenth century<br />

did pr<strong>of</strong>essional, independent periodicals devoted to mathematics come into being;<br />

ABEL was instrumental in the creation <strong>of</strong> the first major German journal <strong>of</strong> mathemat-<br />

ics, which CRELLE founded in 1826.<br />

When ABEL first called upon CRELLE in Berlin, they discussed the relatively low<br />

status <strong>of</strong> mathematics in Germany (Prussia). When ABEL happened to mention his<br />

astonishment at the lack <strong>of</strong> German periodicals devoted to mathematics, he struck a<br />

nerve with CRELLE. For years, CRELLE had been engaged in an effort to promote<br />

mathematics in Prussia. In France, the first journal (Annales de mathématiques pures<br />

et appliquées) devoted entirely to mathematics had been initiated by J. D. GERGONNE<br />

(1771–1859) in 1810. 37 In 1822, CRELLE was forced to abandon plans for a German lan-<br />

guage journal <strong>of</strong> mathematics due to lack <strong>of</strong> contributors. 38 However, with the advent<br />

ABEL and other promising young mathematicians — all looking for a way <strong>of</strong> pub-<br />

lishing their results — the time was ripe for another attempt. Following an intensive<br />

and continuing campaign to secure funding and with substantial personal investment,<br />

CRELLE had the first volume <strong>of</strong> his Journal für die reine und angewandte Mathematik pub-<br />

lished in the spring <strong>of</strong> 1826.<br />

CRELLE’S initial idea for the Journal für die reine und angewandte Mathematik was to<br />

provide a broad German speaking audience with an instrument for presenting and<br />

keeping up to date with recent research in pure and applied mathematics — possibly<br />

36 “Overmaade vel fornøiet er jeg fordi jeg kom til at reise til Tyskland og navnligen til Berlin førend jeg<br />

kom til Paris; thi som Du maaskee har erfaret af mit Brev til Hansteen har jeg her gjort et fortræffeligt<br />

Bekjendtskab med Geiheimrath Crelle.” (<strong>Abel</strong>→Holmboe, 1826/01/16. ibid., 13).<br />

37 (Otero, 1997).<br />

38 (W. Eccarius, 1976, 233).


30 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

through translations. Soon, however, the attention devoted to applied mathematics<br />

declined and the Journal für die reine und angewandte Mathematik became the mouth-<br />

piece for a limited group <strong>of</strong> pure mathematicians. <strong>The</strong> change in CRELLE’S conception<br />

<strong>of</strong> his Journal für die reine und angewandte Mathematik is evident by comparing the in-<br />

troductions with which he prefaced the two first volumes. In the very first volume,<br />

CRELLE described the domain <strong>of</strong> the journal to include both pure mathematics (analy-<br />

sis, geometry, mechanics), and applied mathematics including optics, theories <strong>of</strong> heat,<br />

sound, and probability, and geography and geodesy. 39 This changed quickly, though,<br />

and in the introduction to the second volume, CRELLE stood down — both on nation-<br />

alistic and disciplinary ambitions.<br />

For the first volume, CRELLE took it upon himself to translate the French manu-<br />

scripts <strong>of</strong> ABEL and others into German before publication. 40 Despite having been<br />

taught German for four years at the Cathedral School (written German for two years), 41<br />

ABEL’S marks in written German were quite inconsistent: 1 and 4 on a scale from 1<br />

to 5 (1 best); for the Artium, he scored a 3. 42 ABEL was reluctant to write in that language.<br />

When he eventually prepared a paper in German, ABEL was very proud. 43<br />

However, CRELLE soon succumbed to the pressure <strong>of</strong> internationalizing his journal<br />

and accepted publishing papers in foreign languages. In response, after just a single<br />

paper prepared in German, ABEL returned to writing exclusively in French.<br />

CRELLE’S library. In Christiania, ABEL had access to a large section <strong>of</strong> French lit-<br />

erature on pure mathematics written by the masters and some <strong>of</strong> the servants <strong>of</strong> the<br />

subject in the eighteenth and early nineteenth century. However, circulation <strong>of</strong> re-<br />

sults to the geographical periphery was far from instant, and many <strong>of</strong> the products<br />

<strong>of</strong> the French reorganization <strong>of</strong> mathematics had not yet been brought to Norway.<br />

<strong>The</strong>refore, it was an explicit motivation for ABEL’S European tour to go to the largest<br />

libraries and bookstores on the Continent which he expected to find together with the<br />

rest <strong>of</strong> the mathematical milieu in Göttingen and Paris. However, one <strong>of</strong> his most in-<br />

fluential encounters with the libraries <strong>of</strong> the Continent took place in Berlin, probably<br />

in the private library <strong>of</strong> Geheimrath CRELLE.<br />

“<strong>The</strong> afore-mentioned Crelle also has a perfectly splendid mathematical library<br />

which I use as if it had been my own and from which I benefit particularly<br />

as it contains all the latest material which he gets as soon as possible.” 44<br />

39 (A. L. Crelle, 1826).<br />

40 (W. Eccarius, 1976, 236). CRELLE also occasionally edited the manuscripts.<br />

41 (Stubhaug, 1996, 520).<br />

42 (N. H. <strong>Abel</strong>, 1902d, 3).<br />

43 (<strong>Abel</strong>→Holmboe, Wien, 1826/04/16. N. H. <strong>Abel</strong>, 1902a, 27).<br />

44 “Den samme Crelle har ogsaa et aldeles fortræffeligt mathematisk Bibliothek, som jeg benytter som<br />

mit eget og som jeg har særdeles Nytte af da det indeholder alt det nyeste, som han faaer saa snart<br />

det er mueligt.” (<strong>Abel</strong>→Hansteen, Berlin, 1825/12/05. ibid., 11).


2.3. <strong>The</strong> European tour 31<br />

A.-L. CAUCHY’S (1789–1857) new program <strong>of</strong> founding analysis on the notion <strong>of</strong><br />

limits as expressed by inequalities had not reached Norway before ABEL left. CAUCHY<br />

had expressed his thoughts most influentially in the textbook Cours d’analyse intended<br />

(but never used) for instruction at the École Polytechnique which was printed in 1821. In<br />

a review <strong>of</strong> CAUCHY’S Exercises de mathématiques, CRELLE spoke highly <strong>of</strong> CAUCHY’S<br />

insights and his other works, and there can be little doubt that during ABEL’S time in<br />

Berlin, CAUCHY’S new analysis was discussed by a circle <strong>of</strong> mathematicians around<br />

CRELLE. One member <strong>of</strong> the circle, the mathematician M. OHM (1792–1872) 45 reacted<br />

to CAUCHY’S new rigorization by devising his own approach to algebraic analysis<br />

which was kept the formal aspect which CAUCHY had rejected. 46 In a letter, ABEL<br />

records how the circle discontinued its meetings because <strong>of</strong> G. S. OHM’S (1789–1854)<br />

arrogant mentality. 47 It is possible, that mathematical topics may also have played a<br />

role.<br />

“<strong>The</strong> work [Exercises de mathématiques] is full <strong>of</strong> deep analytical investigations<br />

as it would be expected from the acute and inventive author <strong>of</strong> Cours d’analyse,<br />

Leçons sur le calcul infinitésimal, Leçons sur l’application du calcul infinitésimal à la<br />

géométrie, etc.” 48<br />

ABEL discovered the works <strong>of</strong> CAUCHY in 1826. CAUCHY’S new approach to the<br />

theory <strong>of</strong> infinite series took ABEL by storm, and soon ABEL became one <strong>of</strong> CAUCHY’S<br />

most devoted missionaries (see part III). In print, ABEL first disclosed his sympathies<br />

in his work on the binomial theorem, which was printed in the early spring <strong>of</strong> 1827<br />

(see table 2.1), but ABEL’S letters allow us to date his encounter with the new Cauchian<br />

rigor in analysis more precisely. In a famous letter dated 16 January 1826, i.e. while in<br />

Berlin, ABEL wrote to HOLMBOE:<br />

“Taylor’s theorem, the foundation for the entire higher mathematics, is equally<br />

ill founded. I have found only one rigorous pro<strong>of</strong> which is by Cauchy in his Resumé<br />

des leçons sur le calcul infinitesimal.” 49<br />

It is likely, that it was also in CRELLE’S library that ABEL came across CAUCHY’S<br />

famous textbook Cours d’analyse, a work which had tremendous consequences for<br />

ABEL’S attitude toward rigor. 50<br />

45 For the years <strong>of</strong> birth and death, see (Jahnke, 1987, 103). OHM was the younger brother <strong>of</strong> the famous<br />

physicist OHM.<br />

46 (Jahnke, 1987; Jahnke, 1993).<br />

47 (<strong>Abel</strong>→Hansteen, Berlin, 1825/12/05. N. H. <strong>Abel</strong>, 1902a, 11).<br />

48 “Das Werk ist voller tiefer analytischer Untersuchungen, wie sie sich von dem scharfsinningen und<br />

an neuen Ideen reichen Verfasser des “Cours d’analyse”, “Leçons sur le calcul infinitésimal,” der “Leçons<br />

sur l’application du calcul infinitésimal à la géométrie, etc.” erwarten lassen.” (A. L. Crelle, 1827, 400).<br />

49 “Det Taylorske <strong>The</strong>orem, Grundlaget for hele den høiere Mathematik er ligesaa slet begrundet. Kun<br />

eet eneste strængt Beviis har jeg fundet og det er af Cauchy i hans Resumé des leçons sur le calcul<br />

infinitesimal.” (<strong>Abel</strong>→Holmboe, 1826/01/16. N. H. <strong>Abel</strong>, 1902a, 16).<br />

50 (I. Grattan-Guinness, 1970b, 79). Again, this influence will be documented in part III.


32 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

Volume Number Time <strong>of</strong> publication<br />

1 1 February–March 1826<br />

1 2 June 1826<br />

1 3<br />

1 4 February–March 1827<br />

2 1 5 June 1827<br />

2 2 20 September 1827<br />

2 3<br />

2 4 12 January 1828<br />

3 1 25 March 1828<br />

3 2 26 May 1828<br />

3 3<br />

3 4 3 December 1828<br />

4 1 25 January 1829<br />

4 2 28 March 1829<br />

4 3<br />

4 4 10 June 1829<br />

Table 2.1: Time <strong>of</strong> publications for CRELLE’S Journal für die reine und angewandte Mathematik<br />

1826–1829. Compiled from SYLOW’S notes in (N. H. <strong>Abel</strong>, 1881, vol. 2).<br />

<strong>The</strong> Berlin mathematical scene. In Berlin, mathematics was cultivated in three dis-<br />

tinct — and largely disjoint — circles: the academy, the university, and the circle around<br />

CRELLE. <strong>The</strong> Academy had played its role in the era <strong>of</strong> academies when it had housed<br />

such eminent mathematicians as EULER and LAGRANGE. However, after LAGRANGE<br />

had moved to Paris in 1784 without being suitably replaced, mathematics at the Academy<br />

inevitably and irreversibly declined. 51<br />

In the first ordinary mathematics chair at the university — which had only opened<br />

in 1810 — J. G. TRALLES (1763–1822) had resided. <strong>The</strong> brothers A. VON HUMBOLDT<br />

(1769–1859) and W. VON HUMBOLDT (1767–1835), who had been instrumental in bring-<br />

ing the university into being, had made efforts to call GAUSS to the chair, but he had<br />

to settle for less; TRALLES’ academic record shows a marked bias for applied mathe-<br />

matics, and during his reign pure mathematics was not well <strong>of</strong>f in Berlin not was it<br />

elsewhere in Germany except for Göttingen. 52 Besides the ordinary pr<strong>of</strong>essor, a num-<br />

ber <strong>of</strong> extraordinary pr<strong>of</strong>essors and Privatdozenten 53 <strong>of</strong>fered mathematics courses. Af-<br />

ter TRALLES’ death in 1822, E. H. DIRKSEN (1788–1850), who together with OHM had<br />

previously served as Privatdozenten, was appointed to the ordinary pr<strong>of</strong>essorship.<br />

<strong>The</strong> real forum for pure mathematics in Berlin in the 1820s centered around CRELLE<br />

and condensed around the Journal für die reine und angewandte Mathematik once it was<br />

initiated in 1826. CRELLE organized weekly meetings <strong>of</strong> a group <strong>of</strong> young mathemati-<br />

51 (Knobloch, 1998).<br />

52 (Biermann, 1988, 20–21), (Rowe, 1998)<br />

53 <strong>The</strong>se Privatdozenten include EYTELWEIN, GRUSON, LEHMUS, and LUBBE.


2.3. <strong>The</strong> European tour 33<br />

Author Vol. 1 Vol. 2 Vol. 3 Vol. 4 1826–29<br />

ABEL 97 (7) 100 (4) 56 (6) 125 (6) 378 (23)<br />

STEINER 93 (5) 45 (7) 11 (3) 149 (15)<br />

PONCELET 11 (2) 60 (1) 71 (1) 142 (4)<br />

JACOBI 7 (1) 60 (9) 32 (7) 29 (2) 128 (19)<br />

OLIVIER 66 (7) 54 (4) 1 (1) 121 (12)<br />

DIRICHLET 62 (4) 18 (2) 80 (6)<br />

MÖBIUS 36 (2) 36 (2) 72 (4)<br />

HILL 3 (1) 59 (1) 62 (2)<br />

CRELLE a 43 (3) 4 (1) 47 (4)<br />

RAABE 44 (5) 44 (5)<br />

PLÜCKER 13 (1) 22 (1) 35 (2)<br />

CLAUSEN 2 (1) 13 (5) 15 (5) 30 (11)<br />

HUMBOLDT 27 (1) 27 (1)<br />

OLTMANNS 10 (1) 13 (1) 23 (2)<br />

DIRKSEN 10 (2) 9 (1) 19 (3)<br />

HORN 18 (2) 1 (1) 19 (3)<br />

GRUNERT 17 (3) 1 (1) 18 (4)<br />

a Only papers explicitly authored by “Crelle” or “<strong>The</strong> editor” are included. Besides<br />

these, many anonymous papers must be attributed to CRELLE.<br />

Table 2.2: List <strong>of</strong> most productive (in pages, number <strong>of</strong> papers in parentheses) authors<br />

in CRELLE’S Journal 1826–29.<br />

cians, mainly contributors to his Journal für die reine und angewandte Mathematik; but in<br />

the year <strong>of</strong> the founding <strong>of</strong> the journal, the meetings had to be discontinued due to<br />

personal conflicts within the group. However, the group reorganized under CRELLE’S<br />

initiative and met each Monday for a social event combining music and mathematics,<br />

and in smaller groups they discussed while strolling the city.<br />

<strong>The</strong> members <strong>of</strong> CRELLE’S circle can only be identified indirectly from the list <strong>of</strong><br />

authors <strong>of</strong> the early volumes <strong>of</strong> the Journal and from the correspondence <strong>of</strong> the iden-<br />

tifiable central actors. 54 In ABEL’S letters to his friends in Norway, explicit mention is<br />

only made <strong>of</strong> OHM, the extraordinary pr<strong>of</strong>essor at the university who by an arrogant<br />

nature caused the weekly seances to be discontinued. However, ABEL wrote about a<br />

few unnamed young mathematicians with whom he entertained himself.<br />

2.3.3 Why never Göttingen?<br />

Already in ABEL’S first letter to HANSTEEN from Berlin, a certain ambiguity concern-<br />

ing his obligation to go to Göttingen can be found:<br />

“My winter quarters will be here in Berlin and I have not yet decided when I<br />

am going to leave. For the sake <strong>of</strong> Crelle and the Journal, I would like to stay here<br />

54 CRELLE’S Nachlass appears to have gone on auction shortly after CRELLE’S death in 1855 and must<br />

be considered lost. (W. Eccarius, 1975, 49, footnote)


34 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

as long as possible, and as I gather there is no other place in Germany where I<br />

would be better <strong>of</strong>f. <strong>The</strong>re is certainly a good library in Göttingen, but that is about<br />

all because Gauss who is the only capable one there is completely inapproachable.<br />

However, I must go to Göttingen.” 55<br />

ABEL’S devotion toward CRELLE and the Journal für die reine und angewandte Ma-<br />

thematik seems to overpower his ambition and obligation to go to Göttingen in accord<br />

with the stipend. A month later, a further reservation was expressed in a letter to<br />

HOLMBOE, where ABEL referred to the personality <strong>of</strong> the great GAUSS:<br />

“I am probably going to remain here in Berlin until the end <strong>of</strong> February or<br />

March; then I will travel via Leipzig and Halle to Göttingen, not to see Gauss because<br />

he is said to be intolerably reserved, but for the library which is apparently<br />

excellent.” 56<br />

<strong>The</strong> reservation concerning GAUSS was communicated to HANSTEEN a fortnight<br />

later, in a letter in which ABEL outlined his plans to accompany CRELLE on a trip to<br />

Göttingen. Furthermore, ABEL had already his eyes firmly set on Paris; he wrote the<br />

following to HANSTEEN after describing how he planed to go to Leipzig and Freiburg<br />

with one <strong>of</strong> his Norwegian friends:<br />

“<strong>The</strong>n I [will] travel back to Berlin in order to join Crelle on the tour to Göttingen<br />

and the Rhine area. In Göttingen, I shall only stay for a short period <strong>of</strong> time<br />

as there is nothing to be gained. Gauss is unapproachable and the library cannot<br />

possibly be better than those in Paris.” 57<br />

ABEL’S impression that GAUSS was not easily accessible seems to have been in-<br />

spired by rumors nourished in Berlin at the time. With his few but ripe policy concern-<br />

ing publications, GAUSS’ image was largely built from his network <strong>of</strong> corresponding<br />

mathematicians. Modern biographers <strong>of</strong> GAUSS use the published correspondence<br />

with e.g. the F. BOLYAI (1775–1856) s, ABEL, and others to support the picture <strong>of</strong><br />

the great mathematician as remote and even hostile. However, these events were<br />

only taking place, and the letters were not publicly known in the 1820s. <strong>The</strong> Berlin<br />

mathematicians certainly held GAUSS’ mathematics in high respect but thought less<br />

<strong>of</strong> the master’s personal qualities and openness. HUMBOLDT had — with the help <strong>of</strong><br />

CRELLE — tried to call GAUSS to the Berlin polytechnic and later the university, but<br />

GAUSS had declined all such <strong>of</strong>fers and remained in Göttingen.<br />

55 “Mit Vinterquarteer kommer jeg til at holde her i Berlin og jeg er endnu ikke ganske enig med mig<br />

selv naar jeg skal reise herfra. For Crelles og Journalens Skyld vilde jeg gjerne være her saalænge<br />

som muelig og eftersom jeg hører er der vel intet andet Sted i Tyskland som vil være mig gavnligere.<br />

Göttingen har rigtignok et godt Bibliothek, men det er ogsaa det eneste; thi Gauss som er<br />

den eneste der der kan noget, er aldeles ikke tilgjængelig. Dog til Göttingen maae jeg det forstaaer<br />

sig.” (<strong>Abel</strong>→Hansteen, Berlin, 1825/12/05. N. H. <strong>Abel</strong>, 1902a, 12).<br />

56 “Jeg kommer formodentlig til at blive her i Berlin til Enden af Februar eller Marts, og reiser da<br />

over Leipzig og Halle til Göttingen (ikke for Gauss Skyld, thi han skal være utaalelig stolt men for<br />

Bibliothekets Skyld som skal være fortræffeligt).” (<strong>Abel</strong>→Holmboe, 1826/01/16. ibid., 18).<br />

57 “Siden reiser jeg tilbage til Berlin for i Følge med Crelle at tage Touren til Göttingen og Rhin-Egnene.<br />

I Göttingen bliver jeg kun kort da der ikke er noget at hole. Gauss er utilgjængelig og Bibliotheket<br />

kan ikke være bedre end i Paris.” (<strong>Abel</strong>→Hansteen, Berlin, [1826]/01/30. ibid., 20).


2.3. <strong>The</strong> European tour 35<br />

2.3.4 <strong>The</strong> European detour<br />

During February 1826, CRELLE’S possibilities to go to Göttingen crumbled and ABEL’S<br />

plans for the rest <strong>of</strong> the tour changed. Complaining repeatedly <strong>of</strong> his melancholic<br />

nature, ABEL hesitated and stalled at the prospect <strong>of</strong> travelling to Paris alone. Instead,<br />

his Norwegian friends presented an inviting alternative. With their interests largely<br />

in geology, the Norwegian travelling entourage made it for the Alps, and ABEL chose<br />

to join them. His defence <strong>of</strong> this less-than-obvious decision can be read in a very<br />

charming letter to HANSTEEN:<br />

“Can I then be blamed for wanting to see some <strong>of</strong> the southern life. During my<br />

journey I can work quite hard. Once I am in Vienna and on the road to Paris, the<br />

straight route almost passes through Switzerland. Why should I not also see a bit<br />

<strong>of</strong> that country. My lord! I am not completely without feelings for the beauty <strong>of</strong><br />

Nature. <strong>The</strong> entire trip will postpone my arrival in Paris by two months and that<br />

does no harm. I will catch up. Do you not think that I would benefit from such a<br />

journey?” 58<br />

<strong>The</strong> reaction <strong>of</strong> the sponsor HANSTEEN can only be imagined.<br />

On his detour through Europe, ABEL did continue working on his mathematics,<br />

and he called upon the local mathematicians where he could. In Vienna, ABEL brought<br />

a letter <strong>of</strong> introduction from CRELLE to the mathematician K. L. VON LITTROW (1811–<br />

1877) at the observatory. His encounter with LITTROW was perhaps the only strictly<br />

mathematical benefit gained from the detour itself; mediated by LITTROW, ABEL man-<br />

aged to circulate his result on the insolubility <strong>of</strong> the quintic equation in an even wider<br />

(albeit still German speaking) audience. 59<br />

2.3.5 Isolation in Paris<br />

When ABEL eventually arrived in Paris in the summer <strong>of</strong> 1826, he found the mathe-<br />

matical scene abandoned: Most <strong>of</strong> the Paris mathematicians had left the city for the<br />

countryside. However, ABEL made a brief call to LEGENDRE, whom he described as “a<br />

really excellent old man”. 60 ABEL later met with CAUCHY without having more than a<br />

brief and non-technical conversation with him. Besides the opportunity to meet with<br />

the Parisian mathematicians, ABEL saw in the famous Académie des Sciences a possibil-<br />

ity for presenting his research. Before he left Norway, he had prepared a paper on the<br />

58 “Kan man da fortænke mig i at jeg ønsker ogsaa at see lidt af Sydens Liv og Færden. Paa min Reise<br />

kan jeg jo arbeide temmelig brav. Er jeg nu engang i Wien og jeg skal derfra til Paris saa gaaer jo<br />

den lige Vei næsten igjennem Schweitz. Hvorfor skulde jeg da ikke ogsaa see lidt deraf? Herre Gud!<br />

Jeg er dog ikke uden al Sands for Naturens Skjønheder. Den hele Reise vil gjøre at jeg kommer to<br />

Maaneder senere til Paris end ellers og det gjør da ikke noget. Jeg skal nok hente det ind igjen. Troer<br />

de ikke at jeg vil have godt af en saadan Reise?” (<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. ibid., 24).<br />

59 ABEL had his article from CRELLE’S Journal reviewed anonymously in the newly founded Vienna<br />

based journal Zeitschrift für Physik und Mathematik (Anonymous, 1826). <strong>The</strong> review will be described<br />

in part II.<br />

60 (ABEL to HOLMBOE, Paris 1826.8.12, (N. H. <strong>Abel</strong>, 1902a, 40))


36 Chapter 2. Biography <strong>of</strong> NIELS HENRIK ABEL<br />

insolubility <strong>of</strong> the quintic equation which had, however, already served its purpose as<br />

a door-opener with CRELLE in Berlin where it had been published. <strong>The</strong>refore, that re-<br />

sult was not eligible for the academy and neither did ABEL submit the manuscript on<br />

the integration <strong>of</strong> differentials which he had also prepared for the collegium in Chris-<br />

tiania. Instead, on October 10, ABEL presented a paper, the so-called Paris memoir,<br />

which he had worked out during the tour and his stay in Paris and which extended<br />

the collegium manuscript. As is described in much more detail in part IV, the Paris<br />

mémoire was an algebraic approach to the general theory <strong>of</strong> integration <strong>of</strong> algebraic<br />

differentials and extended the approach to elliptic transcendentals which ABEL had<br />

already embarked upon while in Germany.<br />

ABEL’S entire production in Paris was directly influenced by the works <strong>of</strong> LEGEN-<br />

DRE on elliptic integrals which had recently appeared. 61 ABEL took a new approach to<br />

that theory and extended and advanced the program which LEGENDRE had begun to<br />

include higher transcendentals into analysis. Where the influence <strong>of</strong> LEGENDRE was<br />

thus clearly reflected in ABEL’S interests, the influence from CAUCHY was less direct.<br />

We know from his letters that ABEL bought and read the first issues <strong>of</strong> CAUCHY’S<br />

Exercises de mathématiques which include the first presentation <strong>of</strong> CAUCHY’S theory <strong>of</strong><br />

complex integration. This particular theory was, however, completely without influ-<br />

ence on ABEL’S inversion <strong>of</strong> elliptic integrals as is described in part II.<br />

2.4 Back in Norway<br />

When ABEL ultimately left Paris around the New Year 1826/27, he headed for Berlin<br />

where he worked with CRELLE on editing the Journal for some months. <strong>The</strong>re was<br />

no permanent position for ABEL to return to in Norway; the only possibility, a teach-<br />

ing position at the University to replace pr<strong>of</strong>essor RASMUSSEN, had been given to<br />

HOLMBOE while ABEL was in Paris. During ABEL’S second visit to Berlin, CRELLE<br />

intensified his efforts to convince ABEL to come work for the Journal für die reine und<br />

angewandte Mathematik on a permanent level and CRELLE began lobbying for a posi-<br />

tion for ABEL at one <strong>of</strong> the new institutions <strong>of</strong> higher education in Berlin. In the spring<br />

<strong>of</strong> 1827, ABEL finally headed back north, back to uncertainty in Christiania.<br />

Back in his native country, ABEL made a living tutoring in mathematics and sub-<br />

stituting for HANSTEEN while he went on an expedition to Siberia. Throughout,<br />

ABEL continued his mathematical research and his production was intensified when<br />

he learned that the young German mathematician JACOBI was producing astonishing<br />

results on an important problem in the theory <strong>of</strong> elliptic functions. A hectic compe-<br />

tition ensued during which the two mathematicians advanced the theory far beyond<br />

its previous horizon. Initially, ABEL had been able to provide pro<strong>of</strong>s <strong>of</strong> some <strong>of</strong> the<br />

claims which JACOBI had advanced without giving the pro<strong>of</strong>s. Later, JACOBI’S results<br />

61 (A. M. Legendre, 1811–1817).


2.4. Back in Norway 37<br />

1802, Aug. 5 NIELS HENRIK ABEL was born at Finnøy<br />

1815 ABEL moved to Christiania to attend grammar<br />

school<br />

1818 BERNT MICHAEL HOLMBOE took over<br />

ABEL’s mathematics classes<br />

1821 ABEL graduated from grammar school and<br />

was immatriculated at the university in<br />

Christiania<br />

1823 ABEL visited DEGEN in Copenhagen<br />

1825, Sept. 7 ABEL left for the European tour making the<br />

first stop in Berlin<br />

1826, Jul. 10 ABEL arrived in Paris<br />

1826, Dec. ABEL left Paris and returned to Berlin<br />

1827, May 20 ABEL returned to Christiania<br />

1829, Apr. 6 NIELS HENRIK ABEL died at Froland<br />

Table 2.3: Summary <strong>of</strong> NIELS HENRIK ABEL’s biography<br />

provided inspiration for problems which ABEL tackled. <strong>The</strong> competition was fierce<br />

and only ended when ABEL fell ill around Christmas 1828. Even from his sick-bed,<br />

ABEL signed the last papers which he sent for publication in CRELLE’S Journal. ABEL<br />

died at Froland on April 6 1829 at the age <strong>of</strong> 26. A few dates later good news arrived<br />

from Berlin: CRELLE had finally secured a position for ABEL in Berlin.


Chapter 3<br />

Historical background<br />

<strong>The</strong> social and institutional conditions <strong>of</strong> mathematics resemble the general social con-<br />

text in undergoing abrupt transitions in the late eighteenth and early nineteenth cen-<br />

tury. <strong>The</strong> period is one <strong>of</strong> only a handful <strong>of</strong> instances where the influences <strong>of</strong> politi-<br />

cal and social changes can be witnessed directly in mathematical institutions and re-<br />

search. <strong>The</strong> main argument <strong>of</strong> the present chapter will be to establish that not only the<br />

external, outer shell <strong>of</strong> mathematics was transformed; the very core <strong>of</strong> mathematics,<br />

the ways mathematicians thought about their subject, was also influenced by social<br />

and political upheaval. 1<br />

3.1 Mathematical institutions and networks<br />

Mathematical circles. <strong>The</strong> metaphor <strong>of</strong> center and periphery has aptly been applied<br />

to describe science conducted in geographically “remote” regions such as Scandinavia<br />

during periods when the most prominent contributions were made at larger centers in<br />

Central Europe. 2 However, as described below, in the case <strong>of</strong> mathematics in the early<br />

nineteenth century a curious picture emerges with a very strong mathematical center<br />

in Paris and two lesser centers in Germany: the Gaussian ivory tower in Göttingen<br />

and the emerging mathematical scene in Berlin centered around the university, A. L.<br />

CRELLE (1780–1855), and his Journal. In the following section, N. H. ABEL’S (1802–<br />

1829) position within the two mathematical traditions unfolding at these centers —<br />

for short denoted the French and German traditions — is described.<br />

Norway. Knowledge <strong>of</strong> the Norwegian national history during the early nineteenth<br />

century is <strong>of</strong> importance for understanding the conditions under which ABEL grew<br />

1 Some <strong>of</strong> the historical facts and circumstances have already been touched upon in the preceding<br />

chapters but are taken up here again from a slightly different perspective.<br />

2 <strong>The</strong> terms were first employed in (Shils, 1961) and have since found their way into the history <strong>of</strong><br />

science.<br />

39


40 Chapter 3. Historical background<br />

up and evolved into a prominent mathematician. 3 In the middle <strong>of</strong> the nineteenth<br />

century, national states appeared as the natural units <strong>of</strong> political and military power.<br />

Before that, in the early nineteenth century, unitary states had been the most obvi-<br />

ous ways <strong>of</strong> organizing power. In Scandinavia, two unitary states had coexisted as<br />

rivals for centuries, the Danish-Norwegian monarchy and the Swedish monarchy. In<br />

1814, in the aftermath <strong>of</strong> the Napoleonic Wars, Norway was separated from the twin-<br />

monarchy and ceded to Sweden. During the war, a nationalist sentiment had grown<br />

strong in Norway; this national pride led to a brief period <strong>of</strong> independence before the<br />

transition to Swedish rule.<br />

Receiving all <strong>of</strong> his formal education within Norwegian institutions, ABEL belongs<br />

to the first generation <strong>of</strong> truly Norwegian scientists. Before the creation <strong>of</strong> the university<br />

in 1813, Norwegians who wanted any kind <strong>of</strong> higher education had to go to Copen-<br />

hagen; quite a number <strong>of</strong> them went and later returned to fill administrative or clerical<br />

positions in Norway.<br />

At the Christiania Cathedral School, ABEL received qualified and personal tutor-<br />

ing from B. M. HOLMBOE (1795–1850); and at the University he became the prodigy <strong>of</strong><br />

the pr<strong>of</strong>essors S. RASMUSSEN (1768–1850) and C. HANSTEEN (1784–1873). HOLMBOE<br />

was ABEL’S senior by only seven years and had been among the very first students<br />

to attend the Christiania university where he sat in the mathematics classes <strong>of</strong> RAS-<br />

MUSSEN. Such relations between prodigies and benefactors may well originate in the<br />

fact that the scientific community in Christiania was rather small and yet led by a few<br />

men with international relations.<br />

France. In France (in the early nineteenth century that almost exclusively meant<br />

Paris), by contemporary standards, the scientific community was anything but small.<br />

After the Revolution, educational reforms were introduced to develop the military<br />

and civil engineering in France. In order to achieve this goal, large-scale and very<br />

advanced instruction in mathematics was set up at two newly founded educational<br />

institutions, the École Normale and the École Polytechnique. A mathematical milieu <strong>of</strong><br />

substantial size and quality established itself in the French capital in the decades fol-<br />

lowing the Revolution. Teaching at either the École Polytechnique or the École Normale<br />

provided a means <strong>of</strong> living for mathematicians; and the Académie des Sciences pro-<br />

vided a possibility to communicate mathematical research. <strong>The</strong> focus on the teaching<br />

<strong>of</strong> engineers and the sheer volume <strong>of</strong> classes exerted influences on the contents <strong>of</strong><br />

the mathematics taught and researched; the eighteenth century prevalence <strong>of</strong> appli-<br />

cable mathematics continued to dominate, but the communication <strong>of</strong> the calculus to<br />

the previously un-initiated helped provoke research on the foundations <strong>of</strong> the topic.<br />

<strong>The</strong> liberal ideas <strong>of</strong> the Revolution meant a dramatic increase in the numbers <strong>of</strong> publi-<br />

cations. This general trend also influenced mathematics; mathematicians could more<br />

3 <strong>The</strong> cultural framework <strong>of</strong> Norwegian society in the first decades <strong>of</strong> the nineteenth century has been<br />

aptly described in (Stubhaug, 1996; Stubhaug, 2000); here only a few aspects need repetition.


3.2. ABEL’s position in mathematical traditions 41<br />

easily have their works either printed or published in one <strong>of</strong> the journals which were<br />

set up. 4 However, perhaps as a consequence <strong>of</strong> its size, the milieu in Paris was a<br />

very competitive one in the first decades <strong>of</strong> the nineteenth century; 5 teachers proba-<br />

bly seemed to a Norwegian more reserved there than in Christiania, and mathematical<br />

cooperation was actually rarely seen.<br />

German states. In the early nineteenth century, the German speaking region was<br />

organized in a multitude <strong>of</strong> sovereign states. One <strong>of</strong> the most influential and ambitious<br />

ones, Prussia, was dominated from its capital Berlin where a national awareness was<br />

spreading to the sciences and mathematics in the 1820s.<br />

<strong>The</strong> German reaction to the events in France around the turn <strong>of</strong> the century found<br />

a philosophical grounding in the neo-humanistic movement which sought to reintro-<br />

duce classical ideals <strong>of</strong> humanism and learning through educational reforms. <strong>The</strong>se<br />

reforms — promoted in Prussia by W. VON HUMBOLDT (1767–1835) and others — im-<br />

proved the position <strong>of</strong> mathematics within the curriculum and promoted a particular<br />

view <strong>of</strong> mathematics. 6 In the opinion <strong>of</strong> the neo-humanists, mathematics was not to be<br />

cultivated for its applicability in the sciences; instead, a “pure” form <strong>of</strong> mathematics<br />

was promoted with its own set <strong>of</strong> motivations and ideologies. <strong>The</strong> important mathe-<br />

matical branch <strong>of</strong> algebraic analysis found a unique form with these philosophically<br />

inspired mathematicians in the form <strong>of</strong> the German combinatorial school. 7<br />

In order to implement the educational reforms, new institutional constructions<br />

were devised. In Berlin, a university was opened in 1810 which included mathe-<br />

matics; plans for a polytechnic school — where mathematics should also be taught —<br />

had to be postponed but were carried out in the 1820s. Thus, in Prussia, instruction<br />

in mathematics <strong>of</strong> the future teachers — with its focus on pure thought and individ-<br />

ual contemplation and research — became institutionalized in the first decades <strong>of</strong> the<br />

nineteenth century. 8 Outside the realm <strong>of</strong> the university, a group <strong>of</strong> mathematicians<br />

gathered around CRELLE’S Journal für die reine und angewandte Mathematik; in the 1820s,<br />

this group constituted an alternative mathematical forum in Berlin and a place for<br />

younger mathematicians to meet.<br />

3.2 ABEL’s position in mathematical traditions<br />

Within the Continental tradition in mathematics. ABEL’S interactions with Euro-<br />

pean mathematics were almost exclusively confined to the centers in Berlin and Paris.<br />

In the 1820s, a few other peripheral communities <strong>of</strong> mathematics existed; in particu-<br />

4 (J. Dhombres, 1985; J. Dhombres, 1986).<br />

5 (I. Grattan-Guinness, 1982) and (I. Grattan-Guinness, 1990, e.g. 1227).<br />

6 (Pyenson, 1983).<br />

7 (Jahnke, 1990; Jahnke, 1996). <strong>The</strong> combinatorial school is briefly treated upon in Part III. For a fuller<br />

account, consult (Jahnke, 1987; Jahnke, 1993).<br />

8 (Begehr et al., 1998; Biermann, 1988; Mehrtens, 1981).


42 Chapter 3. Historical background<br />

lar on the British Isles, in the Austrian-Hungarian Empire, and in Russia. <strong>The</strong>re are,<br />

however, nearly no traces <strong>of</strong> any kinds <strong>of</strong> interactions with these mathematicians to<br />

be found in ABEL’S works or letters.<br />

In the first decades <strong>of</strong> the nineteenth century, British mathematicians (both outside<br />

and within the Analytical Society) were consciously trying to adapt the Continental cal-<br />

culus. 9 Only once — in an 1823-letter to HOLMBOE — did ABEL mention the two con-<br />

temporary British mathematicians T. YOUNG (1773–1829) and J. IVORY (1765–1842),<br />

neither <strong>of</strong> whom were associated with the Analytical Society. 10 From a single sentence<br />

in one <strong>of</strong> his letters, we know that ABEL was aware <strong>of</strong> the existence <strong>of</strong> the Czech<br />

theologian, philosopher, and mathematician B. BOLZANO (1781–1848). In one <strong>of</strong> his<br />

notebooks, ABEL makes the following rather sudden remark, which initially confused<br />

the P. L. M. SYLOW (1832–1918), who by 1902 still only knew <strong>of</strong> Bolzano as a town in<br />

the Alps: 11<br />

“Bolzano is a clever fellow from what I have studied” 12<br />

This remark made SYLOW speculate that ABEL had probably read BOLZANO’S Rein<br />

analytischer Beweis during the European tour. 13 <strong>The</strong>re is nothing to suggest that ABEL<br />

met personally with BOLZANO during his stop in Prague during the European de-<br />

tour. 14 In chapter 12.1, ABEL’S contribution to the rigorization <strong>of</strong> analysis is described<br />

in further detail, and a few more details concerning his relation to BOLZANO are dis-<br />

cussed.<br />

Although interested in topics central to the endeavors <strong>of</strong> the Analytical Society<br />

and BOLZANO (certainly independently), ABEL’S inspirations thus seem to come from<br />

somewhere else: the centers <strong>of</strong> Berlin and Paris.<br />

Between the German and the French traditions. ABEL’S mathematical production<br />

was confined to the discipline <strong>of</strong> algebra, the foundations <strong>of</strong> analysis, and the the-<br />

ory <strong>of</strong> higher transcendentals. Thus, his interests did not include geometry and —<br />

except for a youthful work — also excluded applied mathematics. Although these<br />

topics constitute important parts <strong>of</strong> the French and German traditions in mathemat-<br />

ics in the period, 15 ABEL’S work is nevertheless rightfully interpreted within these<br />

9 (Craik, 1999).<br />

10 (N. H. <strong>Abel</strong>, 1902a, 4–8). IVORY had studied mathematics in Scotland before taking up the subject as<br />

his pr<strong>of</strong>ession. In 1819, he retired to become a private mathematician living in London. YOUNG was<br />

an autodidact natural philosopher with an interest in mathematics. He was elected into the French<br />

Académie des Sciences in 1827; ABEL and JACOBI had also been nominated for the election.<br />

11 (L. Sylow, 1902, 12).<br />

12 “Bolzano er en dygtig Karl i hvad jeg [. . . ]” (<strong>Abel</strong>, MS:351:A, 61). <strong>The</strong> rest is crossed out and difficult<br />

to decipher. See (Stubhaug, 2000, 505, fig. 44).<br />

13 (Bolzano, 1817), see (L. Sylow, 1902, 12).<br />

14 Historical speculations based on the similarities <strong>of</strong> results and the possible personal rendezvous <strong>of</strong><br />

the authors have been found inadequate, as the responses to GRATTAN-GUINNESS’ provocative<br />

suggestion that CAUCHY plagiarized BOLZANO (I. Grattan-Guinness, 1970a); for a sober review <strong>of</strong><br />

the ensuing controversy, see (Bottazzini, 1986, 123–124 (note 10)).<br />

15 See e.g. (Dauben, 1981) or (Jahnke, 1994).


3.3. <strong>The</strong> state <strong>of</strong> mathematics 43<br />

traditions. <strong>The</strong> new French approach to rigor and the German focus on pure math-<br />

ematics (i.e. mathematics without the justification <strong>of</strong> applicability) were important<br />

backgrounds for ABEL’S mathematics; no expressed concerns for applicability can be<br />

found in ABEL’S writings; he was — in every respect — a pure mathematician. On<br />

the other hand, as will become evident in chapters 16 and 19, ABEL was no dogmatic<br />

rigorist when he aimed at producing new mathematical results.<br />

3.3 <strong>The</strong> state <strong>of</strong> mathematics<br />

Some tendencies in the mathematicians’ thoughts about mathematics just prior to the<br />

period <strong>of</strong> main interest merit attention. In particular, prominent mathematicians to-<br />

ward the end <strong>of</strong> the eighteenth century expressed the belief that mathematics was just<br />

about to reach its pinnacle <strong>of</strong> cultivation. From the eighteenth century perspective,<br />

where mathematical praxis was to a large part made up <strong>of</strong> formal and explicit ma-<br />

nipulations <strong>of</strong> known algebraic or analytic dependencies, these methods seemed <strong>of</strong><br />

limited scope. For instance, J. L. LAGRANGE (1736–1813) — a few years after his in-<br />

novative paper on polynomial equations from which our story will commence in the<br />

next chapter — wrote to J. LE R. D’ALEMBERT (1717–1783) in 1781,<br />

“It appears to me also that the mine [<strong>of</strong> mathematics] is already very deep and<br />

that unless one discovers new veins it will be necessary sooner or later to abandon<br />

it.” 16<br />

Similar dark visions seem to be recurring at intervals — <strong>of</strong>ten in the form <strong>of</strong> fin-<br />

de-siècle pessimism. Even into the nineteenth century, a similar view was expressed<br />

by J.-B. J. DELAMBRE (1749–1822). In 1810, DELAMBRE delivered a commissioned<br />

review <strong>of</strong> the progress made in the mathematical sciences after the French Revolution.<br />

He expressed his concern over the future <strong>of</strong> mathematics in the following way:<br />

“It would be difficult and rash to analyze the chances which the future <strong>of</strong>fers<br />

to the advancement <strong>of</strong> mathematics; in almost all its branches one is blocked by<br />

insurmountable difficulties; perfection <strong>of</strong> detail seems to be the only thing which<br />

remains to be done. [. . . ] All these difficulties appear to announce that the power<br />

<strong>of</strong> our analysis is practically exhausted in the same way as the power <strong>of</strong> the ordinary<br />

algebra was with respect to the geometry <strong>of</strong> transcendentals at the time <strong>of</strong><br />

Leibniz and Newton, and it is required that combinations are made which open a<br />

new field in the calculus <strong>of</strong> transcendentals and in the solution <strong>of</strong> equations which<br />

these [transcendentals] contain.” 17<br />

16 “Il me semble aussi que la mine est presque déjà trop pr<strong>of</strong>onde, et qu’à moins qu’on ne découvre<br />

de nouveaux filons il faudra tôt ou tard l’abandonner.” (Lagrange→d’Alembert, Berlin, 1781. Lagrange,<br />

1867–1892, vol. 13, 368); English translation from (Kline, 1990, 623).<br />

17 “Il seroit difficile et peut-être téméraire d’analyser les chances que l’avenir <strong>of</strong>fre à l’avancement des<br />

mathématiques: dans presque toutes les parties, on est arrêté par des difficultés insurmontables;<br />

des perfectionnemens de détail semblent la seule chose qui reste à faire; [. . . ] Toutes ces difficultés<br />

semblent annoncer que la puissance de notre analyse est à-peu-près épuisée, comme celle de


44 Chapter 3. Historical background<br />

Taken together, the two quotations hint at two possible ways out <strong>of</strong> the apparently<br />

imminent stagnation <strong>of</strong> mathematics: discovery <strong>of</strong> new questions (veins) and fusions<br />

<strong>of</strong> existing theories. After the evidence has been presented in the following three parts,<br />

it will be illustrated how the mathematical community <strong>of</strong> the early nineteenth century<br />

invoked precisely these approaches in a period <strong>of</strong> such mathematical creativity that<br />

the remarks <strong>of</strong> LAGRANGE and DELAMBRE afterwards seem well <strong>of</strong>f the mark.<br />

3.4 ABEL’s legacy<br />

It is well known that after ABEL’S death, his name was tied to a romantic picture <strong>of</strong> a<br />

neglected mathematical genius. His case was used as fuel for arguments ranging from<br />

nationalistic awareness to revolutionary issues. His influence on the ensuing century<br />

was vast — both as mathematical and even personal inspiration. Here, only ABEL’S<br />

mathematical legacy will be discussed, although certain aspects <strong>of</strong> his personality and<br />

destiny are present in the quotations given. More importantly than contributing new<br />

results to mathematics, ABEL’S programmatic approach caught the attention <strong>of</strong> the<br />

leading figures in mathematics in the nineteenth century, in particular C. G. J. JACOBI<br />

(1804–1851) and K. T. W. WEIERSTRASS (1815–1897).<br />

<strong>The</strong> judgement on ABEL’S legacy passed by A.-M. LEGENDRE (1752–1833) and<br />

JACOBI is legendary, itself. In a letter to CRELLE, LEGENDRE is reported to have said:<br />

“After having worked by myself, I have felt a great satisfaction paying homage<br />

to Mr. <strong>Abel</strong>’s talents, feeling all the merits <strong>of</strong> the beautiful theorem the discovery<br />

<strong>of</strong> which is due to him and to which the qualification monumentum aere perennius<br />

can be applied.” 18<br />

JACOBI, who reported LEGENDRE’S homage to ABEL in his review <strong>of</strong> the third sup-<br />

plement <strong>of</strong> LEGENDRE’S Théorie des fonctions elliptiques, at another place qualified and<br />

generalized the praise which should be given to ABEL’S contribution to mathematics:<br />

“<strong>The</strong> vast problems which he [ABEL] had proposed to himself — i.e. to establish<br />

sufficient and necessary criteria for any algebraic equation to be solvable, for<br />

any integral to be expressible in finite terms, his admirable discovery <strong>of</strong> the theorem<br />

encompassing all the functions which are the integrals <strong>of</strong> algebraic functions,<br />

etc. — marks a very special type <strong>of</strong> questions which nobody before him had dared<br />

to imagine. He has gone but he has left a grand example.” 19<br />

l’algèbre ordinaire l’étoit par rapport à la géométrie transcendante au temps de Leibnitz et de Newton,<br />

et qu’il faut des combinaisons qui ouvrent un nouveau champ au calcul des transcendantes et à<br />

la résolution des équations qui les contiennent.” (Delambre, 1810, 131); translation based on Kline,<br />

1990, 623.<br />

18 “En travaillant pour mon propre compte, j’ai éprouvé une grande satisfaction, de rendre un éclatant<br />

hommage au génie de Mr. <strong>Abel</strong>, en faisant sentir tout le mérite du beau théorème dont l’invention<br />

lui est due, et auquel on peut appliquer la qualification de monumentum aere perennius.” Legendre<br />

quoted in C. G. J. Jacobi, 1832a, 413.


3.4. ABEL’s legacy 45<br />

As can be seen from the quote, to JACOBI, ABEL’S legacy laid more in the way he<br />

asked questions than in the solutions and answers which he provided. ABEL’S ques-<br />

tions were, in the mind <strong>of</strong> JACOBI, questions <strong>of</strong> necessary and sufficient conditions<br />

for certain properties to hold. This interpretation highlights two aspects which find<br />

instances in the present work: rigorization as the process <strong>of</strong> making clear, useful, and<br />

precise the conditions <strong>of</strong> theorems, and the new, concept based questions which are<br />

guaranteed to be answerable, although the answers might defy contemporary intu-<br />

itions.<br />

In his youth, WEIERSTRASS was deeply influenced in his career by works written<br />

by ABEL. Throughout his life, WEIERSTRASS thought highly <strong>of</strong> ABEL; the following<br />

statements testify to WEIERSTRASS’ devotion which at times almost resembles envy.<br />

“<strong>The</strong> fortunate <strong>Abel</strong>; he has contributed something <strong>of</strong> lasting value! — He [ABEL]<br />

was used to always taking the most elevated point <strong>of</strong> view. — <strong>Abel</strong> was distinguished<br />

by the all-embracing vision directed at the highest position, the ideal.” 20<br />

<strong>The</strong>se quotations are, <strong>of</strong> course, equally good sources to WEIERSTRASS’ views on<br />

mathematics in the second half <strong>of</strong> the nineteenth century as to ABEL’S mathemati-<br />

cal production in the 1820s. However, the quotations touch upon the same themes as<br />

the quotation from JACOBI above; in due time it will be clearer, on which basis WEIER-<br />

STRASS could claim that ABEL had produced lasting results by taking the most general<br />

approach toward the idealistic goal <strong>of</strong> mathematics.<br />

19 “Les vastes problèmes qui’il s’était proposés, d’établir des critères suffisants et nécessaires pour<br />

qu’une équation algébrique quelconque soit résoluble, pour qu’une intégrale quelconque puisse être<br />

exprimée en quantités finies, son invention admirable de la propriété générale qui embrasse toutes<br />

les fonctions qui sont des intégrales de fonctions algébriques quelconques, etc., etc., marquent un<br />

genre de questions tout à fait particulier, et que personne avant lui n’a osé imaginer. Il s’en est allé,<br />

mais il a laissé un grand exemple.” (Jacobi→Legendre, Potsdam, 1829. Legendre and Jacobi, 1875,<br />

70–71); for a German translation, see (Pieper, 1998, 153).<br />

20 “<strong>Abel</strong> der Glückliche; er hat etwas Bleibendes geleistet! — Er [ ABEL] war gewohnt, überall den<br />

höchsten Standpunkt einzunehmen. — <strong>Abel</strong> zeichnete der allumfassende, auf das höchste, das Ideale<br />

gerichtete Blick aus.” <strong>The</strong> quotes are all taken from (Biermann, 1966, 218).


Part II<br />

“My favorite subject is algebra”<br />

47


Chapter 4<br />

<strong>The</strong> position and role <strong>of</strong> ABEL’s works<br />

within the discipline <strong>of</strong> algebra<br />

In the nineteenth century, the theory <strong>of</strong> equations acquired its status as a mathematical<br />

discipline with its own set <strong>of</strong> problems, methods, and legitimizations. In the process,<br />

N. H. ABEL (1802–1829) played an important role. His works on the algebraic in-<br />

solubility <strong>of</strong> the general quintic equation and his penetrating studies <strong>of</strong> the so-called<br />

<strong>Abel</strong>ian equations belong to the first results established within this incipient discipline.<br />

Although ABEL’S investigations raised new questions and answered some <strong>of</strong> them,<br />

his methods and his approach was deeply rooted in the works <strong>of</strong> mathematicians be-<br />

longing to the previous generations. In particular, ABEL drew upon the algebraic re-<br />

searches <strong>of</strong> L. EULER (1707–1783). <strong>The</strong>refore, in the following chapter 5, these works,<br />

similar approaches taken by A.-T. VANDERMONDE (1735–1796), and the even more<br />

influential works by J. L. LAGRANGE (1736–1813) and C. F. GAUSS (1777–1855) are<br />

described and analyzed. In the ensuing chapters 6–8, ABEL’S algebraic researches are<br />

described and their role and impact are analyzed. Focus in this part II will be on de-<br />

scribing the change in asking and answering questions pertaining to mathematical ob-<br />

jects; more precisely questions concerning the algebraic solubility <strong>of</strong> equations. Such<br />

questions have been central to mathematical development since the Renaissance, but<br />

starting in the second half <strong>of</strong> the eighteenth century, they gave rise to a new mathe-<br />

matical theory. Once this theory-building has been described, the attention is directed<br />

toward ABEL’S approach to algebraic questions. ABEL’S studies <strong>of</strong> the quintic equa-<br />

tion provide an example <strong>of</strong> how a change in the process <strong>of</strong> asking questions led to<br />

unexpected answers. <strong>The</strong>n, because <strong>of</strong> the similarity in methods and inspirations,<br />

ABEL’S questions concerning the geometric division <strong>of</strong> the lemniscate are treated to<br />

illustrate how an algebraic topic emerged within an apparently non-algebraic realm.<br />

Finally, the quest — taken up by ABEL and slightly later by E. GALOIS (1811–1832) —<br />

to completely describe solvable equations is outlined to provide a first illustration <strong>of</strong><br />

the new and more abstract kind <strong>of</strong> questions which C. G. J. JACOBI (1804–1851) saw<br />

49


50 Chapter 4. <strong>The</strong> position and role <strong>of</strong> ABEL’s works within the discipline <strong>of</strong> algebra<br />

as ABEL’S greatest legacy. 1<br />

4.1 Outline <strong>of</strong> ABEL’s results and their structural<br />

position<br />

In the penultimate year <strong>of</strong> the eighteenth century, the Italian P. RUFFINI (1765–1822)<br />

had published the first pro<strong>of</strong> <strong>of</strong> the impossibility <strong>of</strong> solving the general quintic alge-<br />

braically. Working within the same tradition as ABEL, RUFFINI published his investi-<br />

gations on numerous occasions; however, his presentations were generally criticized<br />

for lacking clarity and rigour. Not until 1826 — after ABEL had published his pro<strong>of</strong> <strong>of</strong><br />

this result 2 — did ABEL mention RUFFINI’S pro<strong>of</strong>s, and there is reason to believe that<br />

ABEL obtained his pro<strong>of</strong> independently <strong>of</strong> RUFFINI, yet from the same inspirations.<br />

From LAGRANGE’S comprehensive study <strong>of</strong> the solution <strong>of</strong> equations 3 originated<br />

the idea <strong>of</strong> studying the numbers <strong>of</strong> formally distinct values which a rational func-<br />

tion <strong>of</strong> multiple quantities could take when these quantities were permuted. <strong>The</strong> idea<br />

was cultivated and emancipated into an emerging theory <strong>of</strong> permutations by A.-L.<br />

CAUCHY (1789–1857) who in 1815 provided the theory <strong>of</strong> permutations with its ba-<br />

sic notation and terminology. 4 CAUCHY also established the first important theorem<br />

within this theory when he proved a generalization <strong>of</strong> one <strong>of</strong> RUFFINI’S results to the<br />

effect that no function <strong>of</strong> five quantities could have three or four different values under<br />

permutations <strong>of</strong> these quantities.<br />

Insolubility <strong>of</strong> the quintic. ABEL combined the results and terminology <strong>of</strong> CAUCHY’S<br />

theory <strong>of</strong> permutations with his own innovative investigations <strong>of</strong> algebraic expressions<br />

(radicals). ABEL’S pro<strong>of</strong> is a representation <strong>of</strong> his approach to mathematics. Once he<br />

had realized that the quintic might be unsolvable, he was led to study the “extent”<br />

<strong>of</strong> the class <strong>of</strong> algebraic expressions which could serve as solutions: the “expressive<br />

power” <strong>of</strong> algebraic solutions. Following a minimalistic definition <strong>of</strong> algebraic expres-<br />

sions, ABEL classified these newly introduced objects in a way imposing a hierarchic<br />

structure in the class <strong>of</strong> radicals. <strong>The</strong> classification enabled ABEL to link algebraic ex-<br />

pressions — formed from the coefficients — which occur in any supposed solution for-<br />

mula to rational functions <strong>of</strong> the roots <strong>of</strong> the equation. By the theory <strong>of</strong> permutations,<br />

which ABEL had taken over from CAUCHY, he reduced such rational functions to only<br />

a few standard forms. Considering these forms individually, ABEL demonstrated —<br />

by reductio ad absurdum — that no algebraic solution formula for the general quintic<br />

could exist.<br />

1 See p. 44, above.<br />

2 (N. H. <strong>Abel</strong>, 1824b; N. H. <strong>Abel</strong>, 1826a).<br />

3 (Lagrange, 1770–1771).<br />

4 (A.-L. Cauchy, 1815a).


4.1. Outline <strong>of</strong> ABEL’s results and their structural position 51<br />

In the first part <strong>of</strong> the nineteenth century, the century-long search for algebraic<br />

solution formulae was brought to a negative conclusion: no such formula could be<br />

found. To many mathematicians <strong>of</strong> the late eighteenth century such a conclusion had<br />

been counter-intuitive, but owing to the work and utterings <strong>of</strong> men like E. WARING<br />

(∼1736–1798), 5 LAGRANGE, and GAUSS the situation was different in the 1820s.<br />

ABEL’S pro<strong>of</strong> was also met with criticism and scrutiny. By and large, though, the<br />

criticism was confined to local parts <strong>of</strong> the pro<strong>of</strong>. <strong>The</strong> global statement — that the gen-<br />

eral quintic was unsolvable by radicals — was soon widely accepted.<br />

<strong>Abel</strong>ian equations. In his only other publication on the theory <strong>of</strong> equations, Mémoire<br />

sur une classe particulière d’équations résolubles algébriquement 1829, ABEL took a different<br />

approach. <strong>The</strong> paper was inspired by ABEL’S own research on the division problem<br />

for elliptic functions and GAUSS’ Disquisitiones arithmeticae. In it, ABEL demonstrated<br />

a positive result that an entire class <strong>of</strong> equations — characterized by relations between<br />

their roots — were algebraically solvable.<br />

For his 1829 approach, ABEL seamlessly abandoned the permutation theoretic pil-<br />

lar <strong>of</strong> the insolubility-pro<strong>of</strong>. Instead, he introduced the new concept <strong>of</strong> irreducibility<br />

and — with the aid <strong>of</strong> the Euclidean division algorithm — proved a fundamental the-<br />

orem concerning irreducible equations.<br />

<strong>The</strong> equations which ABEL studied in 1829 were characterized by having rational<br />

relations between their roots. 6 Using the concept <strong>of</strong> irreducibility, ABEL demonstrated<br />

that such irreducible equations <strong>of</strong> composite degree, m × n, could be reduced to equa-<br />

tions <strong>of</strong> degrees m and n in such a way that only one <strong>of</strong> these might not be solvable<br />

by radicals. Furthermore, he proved that if all the roots <strong>of</strong> an equation could be writ-<br />

ten as iterated applications <strong>of</strong> a rational function to one root, 7 the equation would be<br />

algebraically solvable.<br />

<strong>The</strong> most celebrated result contained in ABEL’S Mémoire sur une classe particulière<br />

was the algebraic solubility <strong>of</strong> a class <strong>of</strong> equations later named <strong>Abel</strong>ian by L. KRO-<br />

NECKER (1823–1891). <strong>The</strong>se equations were characterized by the following two prop-<br />

erties: (1) all their roots could be expressed rationally in one root, and (2) these ratio-<br />

nal expressions were “commuting” in the sense that if θ i (x) and θ j (x) were two roots<br />

given by rational expressions in the root x, then<br />

θ iθ j (x) = θ jθ i (x) .<br />

By reducing the solution <strong>of</strong> such an equation to the theory he had just developed,<br />

ABEL demonstrated that a chain <strong>of</strong> similar equations <strong>of</strong> decreasing degrees could be<br />

constructed. <strong>The</strong>reby, he proved the algebraic solubility <strong>of</strong> <strong>Abel</strong>ian equations.<br />

5 1734 is a more qualified guess for Waring’s year <strong>of</strong> birth than (Scott, 1976) giving “around 1736”.<br />

See (Waring, 1991, xvi).<br />

6 (N. H. <strong>Abel</strong>, 1829c).<br />

7 I.e. an equation in which the roots are x, θ (x) , θ (θ (x)) , . . . , θ n (x) for some rational function θ.


52 Chapter 4. <strong>The</strong> position and role <strong>of</strong> ABEL’s works within the discipline <strong>of</strong> algebra<br />

ABEL planned to apply this theory to the division problems for circular and elliptic<br />

functions. However, only his reworking <strong>of</strong> GAUSS’ study <strong>of</strong> cyclotomic equations was<br />

published in the paper.<br />

Together, the insolubility pro<strong>of</strong> and the study <strong>of</strong> <strong>Abel</strong>ian equations can be inter-<br />

preted as an investigation <strong>of</strong> the extension <strong>of</strong> the concept <strong>of</strong> algebraic solubility. On<br />

one hand, the insolubility pro<strong>of</strong> provided a negative result which limited this exten-<br />

sion by establishing the existence <strong>of</strong> certain equations in its complement. On the other<br />

hand, the <strong>Abel</strong>ian equations fell within the extension <strong>of</strong> the concept <strong>of</strong> algebraic solu-<br />

bility and thus ensured a certain power (or volume) <strong>of</strong> the concept.<br />

Grand <strong>The</strong>ory <strong>of</strong> Solubility. In a notebook manuscript — first published 1839 in the<br />

first edition <strong>of</strong> ABEL’S Œuvres — ABEL pursued further investigations <strong>of</strong> the exten-<br />

sion <strong>of</strong> the concept <strong>of</strong> algebraic solubility. In the introduction to the manuscript, he<br />

proposed to search for methods <strong>of</strong> deciding whether or not a given equation was solv-<br />

able by radicals. <strong>The</strong> realization <strong>of</strong> this program would, thus, have amounted to a<br />

complete characterization <strong>of</strong> the concept <strong>of</strong> algebraic solubility.<br />

ABEL’S own approach to this program was based upon his concept <strong>of</strong> irreducible<br />

equations. In the first part <strong>of</strong> the manuscript — which appears virtually ready for the<br />

press — ABEL gave his definition <strong>of</strong> this concept. Arguing from the definition, he<br />

proved some basic and important theorems concerning irreducible equations.<br />

In the latter part <strong>of</strong> the manuscript — which is less lucid and toward the end con-<br />

sists <strong>of</strong> nothing but equations — ABEL reduced the study <strong>of</strong> algebraic expressions sat-<br />

isfying a given equation <strong>of</strong> degree µ to the study <strong>of</strong> algebraic expressions which could<br />

satisfy an irreducible <strong>Abel</strong>ian equation whose degree divided µ − 1. However, ABEL’S<br />

researches were inconclusive. When ABEL’S attempt at a general theory <strong>of</strong> algebraic<br />

solubility eventually was published in 1839, another major player in the field, GA-<br />

LOIS, had also worked on the subject. Inspired by the same tradition and exemplary<br />

problems as ABEL had been, GALOIS put forth a very general theory with the help <strong>of</strong><br />

which the solubility <strong>of</strong> any equation could — at least in principle — be decided.<br />

GALOIS’ writings were inaccessible to the mathematical community until the mid-<br />

dle <strong>of</strong> the nineteenth century. His style was brief and — at times — obscure and unrig-<br />

orous. Many mathematicians <strong>of</strong> the second half <strong>of</strong> the nineteenth century — starting<br />

with J. LIOUVILLE (1809–1882) who first published GALOIS’ manuscripts in 1846 —<br />

invested large efforts in clarifying, elaborating, and extending GALOIS’ ideas. In the<br />

process, the theory <strong>of</strong> equations finally emerged in its modern form as a fertile subfield<br />

<strong>of</strong> modern algebra. Part <strong>of</strong> this evolution concerned mathematical styles. <strong>The</strong> highly<br />

computational mathematical style <strong>of</strong> the eighteenth century, to which ABEL had also<br />

adhered, was superseded. <strong>The</strong> old style had been marked by lengthy, rather concrete,<br />

and painstaking algebraic manipulations. In the nineteenth century, this was replaced<br />

by a more conceptual reasoning, early glimpses <strong>of</strong> which can be seen in ABEL’S works<br />

on the algebraic solubility <strong>of</strong> equations.


4.2. Mathematical change as a history <strong>of</strong> new questions 53<br />

4.2 Mathematical change as a history <strong>of</strong> new questions<br />

A permeating theme <strong>of</strong> the present work is the emergence <strong>of</strong> new questions in the<br />

early nineteenth century. <strong>The</strong> description and analyses <strong>of</strong> ABEL’S algebraic works<br />

serve to illustrate three aspects <strong>of</strong> this process:<br />

1. New questions may have unexpected answers which push mathematics for-<br />

ward.<br />

2. New and fertile questions may arise from importing methods or inspirations<br />

from one theoretical complex into another; entirely new theories may develop.<br />

3. A deliberate reformulation <strong>of</strong> hard but improperly formulated questions may<br />

transform them into forms more open to mathematical treatment. <strong>The</strong> process <strong>of</strong><br />

reformulating the question may involve a process <strong>of</strong> scrutinizing mathematical<br />

intuitions.<br />

New questions with unexpected answers. Ever since procedures to algebraically<br />

compute the roots <strong>of</strong> cubic and bi-quadratic equations were discovered in the mid-<br />

dle <strong>of</strong> the sixteenth century, the search had been on for a generalization to quin-<br />

tic equations. Once R. DU P. DESCARTES’ (1596–1650) new notational system trans-<br />

lated the problem into purely algebraic manipulations <strong>of</strong> symbols, the belief became<br />

widespread that such a generalization had to be obtainable. Although the goal defied<br />

even the greatest mathematicians for centuries, the belief remained intact as late as<br />

the second half <strong>of</strong> the eighteenth century. EULER, for instance, felt assured enough<br />

about the general algebraic solubility <strong>of</strong> equations to utilize it as the basis for pro<strong>of</strong>s<br />

<strong>of</strong> another almost self-evident result: the fundamental theorem <strong>of</strong> algebra.<br />

In 1770, LAGRANGE decided to study carefully the reasons behind the solubility<br />

<strong>of</strong> equations <strong>of</strong> degrees 1,2,3, and 4 with the hope <strong>of</strong> obtaining some kind <strong>of</strong> general<br />

procedure which could subsequently be applied to the fifth degree equation. LA-<br />

GRANGE’S investigations were important in two respects: firstly, they provided a the-<br />

orization <strong>of</strong> the problem into problems <strong>of</strong> permutations <strong>of</strong> the roots — a mathemat-<br />

ical tool which would become immensely important for the problem, and secondly,<br />

LAGRANGE envisioned that the powers <strong>of</strong> his analysis were not powerful enough to<br />

deduce the desired result. This second observation can be taken as the first hint that<br />

such solutions were beyond the reach <strong>of</strong> humans.<br />

In the last years <strong>of</strong> the eighteenth century, the full consequences <strong>of</strong> the failure to<br />

obtain algebraic solutions to the general quintic were realized and published indepen-<br />

dently by two mathematicians located at opposite ends <strong>of</strong> the pr<strong>of</strong>essional spectrum:<br />

the German “prince <strong>of</strong> mathematics” GAUSS and the much lesser known Italian RUF-<br />

FINI. In 1799, GAUSS remarked as a criticism <strong>of</strong> EULER that the algebraic solubility<br />

<strong>of</strong> equations should not be taken for granted. <strong>The</strong> same year, RUFFINI published the


54 Chapter 4. <strong>The</strong> position and role <strong>of</strong> ABEL’s works within the discipline <strong>of</strong> algebra<br />

first <strong>of</strong> a series <strong>of</strong> pro<strong>of</strong>s that the general fifth degree equation could not be solved al-<br />

gebraically. RUFFINI’S pro<strong>of</strong>s were, as noted, difficult and had little impact, although<br />

RUFFINI communicated with some <strong>of</strong> the Parisian mathematicians. Instead, math-<br />

ematicians took some notice <strong>of</strong> GAUSS’ 1801 claim in the prestigious Disquisitiones<br />

arithmeticae to possess a rigorous pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the quintic equation.<br />

During the third decade <strong>of</strong> the 19 th century, the question was finally resolved by<br />

ABEL and — more generally — by GALOIS. In some respect, RUFFINI had already ob-<br />

tained the answer in 1799, and comparing the pro<strong>of</strong>s <strong>of</strong> RUFFINI, ABEL, and GALOIS<br />

sheds interesting light on the intra- and extra-mathematical mechanisms behind the<br />

establishment <strong>of</strong> mathematical knowledge.<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the quintic is a fascinating combination <strong>of</strong> previ-<br />

ously established results and an approach designed to make the question addressable.<br />

Despite lacking in certain respects, ABEL’S pro<strong>of</strong> and its conclusion soon gained wide<br />

acceptance among the experts. However, for many years to come, some mathemati-<br />

cians found the conclusion so counter-intuitive that they had to doubt the result. It is<br />

in this respect that the question led to an unexpected answer.<br />

Asking algebraic questions <strong>of</strong> transcendental objects By 1823, ABEL had carefully<br />

studied GAUSS’ Disquisitiones arithmeticae, although no explicit reference to it was<br />

made in the insolubility-pro<strong>of</strong>s. When ABEL reacted upon a suggestion by C. F. DE-<br />

GEN (1766–1825) to turn his attention toward the study <strong>of</strong> higher transcendentals, he<br />

found ample inspiration from GAUSS’ work. In the Disquisitiones, GAUSS applied al-<br />

gebraic studies to the problem <strong>of</strong> constructing regular polygons with the help <strong>of</strong> ruler<br />

and compass. GAUSS suggested that the method could be carried over to the divi-<br />

sion problem for curves whose rectification depended on a simple elliptic integral, the<br />

lemniscate integral. In a large paper, which included the foundation <strong>of</strong> the new objects<br />

elliptic functions, ABEL provided the details supporting GAUSS’ claim and was led to a<br />

new class <strong>of</strong> polynomial equations which were always solvable.<br />

Thus, in the midst <strong>of</strong> a realm apparently inherited by highly transcendental objects,<br />

ABEL focuses upon algebraic relations pertaining to and existing among these objects.<br />

<strong>The</strong> theory <strong>of</strong> higher transcendentals was in a phase <strong>of</strong> transition in the period, and<br />

ABEL’S algebraic focus influenced the future developments during the second quarter<br />

<strong>of</strong> the nineteenth century.<br />

<strong>The</strong> art <strong>of</strong> asking answerable questions An important ingredient in bringing about<br />

the change in attitude toward the solubility <strong>of</strong> the quintic had been ABEL’S way <strong>of</strong><br />

asking questions. In a passage in one <strong>of</strong> his notebooks, ABEL emphasized that any<br />

mathematical problem, when formulated properly, is decidable — be it affirmatively<br />

or not. 8 Thus, for goals which had remained unattainable for years, ABEL suggested<br />

8 <strong>The</strong> belief is also present in HILBERT’S “Wir müssen wissen, wir werden wissen.” However, as the<br />

development in the twentieth century showed, such a belief has to take into account the accepted


4.2. Mathematical change as a history <strong>of</strong> new questions 55<br />

a reformulation <strong>of</strong> the problem to a question <strong>of</strong> the form “is this goal achievable?” In<br />

the case <strong>of</strong> the quintic equation, the search for an algebraic solution was reformulated<br />

to a question whether such a solution existed at all.<br />

Such change <strong>of</strong> attitude toward mathematical goals signal — as JACOBI soon real-<br />

ized — a change toward more general and abstract mathematics. In order to answer<br />

questions concerning possibility <strong>of</strong> existence, ABEL used implicit quantification over<br />

all possible solutions to the question. His approach was based upon the classification<br />

and normalization <strong>of</strong> these objects which were therefore studied — not individually —<br />

but as items belonging to a collection defined by a concept. Thus, a concept based ap-<br />

proach to doing mathematics was intimately connected to the kinds <strong>of</strong> questions asked<br />

and addressed.<br />

In the theory <strong>of</strong> equations, having established the existence <strong>of</strong> both algebraically<br />

solvable and unsolvable exemplars, ABEL raised the question <strong>of</strong> determining directly<br />

whether a given equation would be solvable or not. In a notebook manuscript, ABEL<br />

set out to address this question. For certain types <strong>of</strong> equations, he made some progress;<br />

however, it was left to GALOIS to outline a theory, based on the same inspirations as<br />

ABEL, which — when elaborated — was powerful enough to answer the question.<br />

rules <strong>of</strong> mathematical reasoning and the system <strong>of</strong> primitive truth from which deductions are made.


Chapter 5<br />

Towards unsolvable equations<br />

By the dawn <strong>of</strong> the nineteenth century, the theory <strong>of</strong> equations addressed a wide range<br />

<strong>of</strong> questions. For the present purpose, the main question is the one <strong>of</strong> algebraic solubil-<br />

ity, but in the eighteenth century, a multitude <strong>of</strong> other questions concerning existence<br />

and characterization <strong>of</strong> roots were intertwined with it. <strong>The</strong>refore, in order to broaden<br />

the perspective, aspects <strong>of</strong> the history <strong>of</strong> these approaches are briefly outlined.<br />

<strong>The</strong> existence <strong>of</strong> roots. When R. DU P. DESCARTES (1596–1650) in 1637 claimed that<br />

any equation <strong>of</strong> degree n possessed n roots an important theorem <strong>of</strong> algebra was for-<br />

mulated whose pro<strong>of</strong> became central to subsequent development. 1 His way out was<br />

a rather evasive one which consisted <strong>of</strong> distinguishing the real ones (real meaning “in<br />

existence”) from the imaginary ones which were products <strong>of</strong> human imagination. To<br />

DESCARTES the assertion that any equation <strong>of</strong> degree n had n roots took the form <strong>of</strong> a<br />

general property possessed by all equations and the trick <strong>of</strong> introducing the imagined 2<br />

roots saved him from further argument. 3<br />

“Neither the true nor the false roots are always real; sometimes they are imaginary;<br />

that is, while we can always conceive <strong>of</strong> as many roots for each equation as<br />

I have already assigned; yet there is not always a definite quantity corresponding<br />

to each root so conceived <strong>of</strong>.” 4<br />

To the next generations <strong>of</strong> mathematicians the character <strong>of</strong> the core <strong>of</strong> the theorem<br />

changed slightly. Where DESCARTES had not dealt with the nature <strong>of</strong> the imagined<br />

roots, they did. Soon the problem <strong>of</strong> demonstrating that all (imagined) roots <strong>of</strong> a<br />

1 In fact it had been formulated by GIRARD in 1629 (Gericke, 1970, 65).<br />

2 I shall use the term “imagined” to distinguish it from the current technical term “imaginary”. <strong>The</strong><br />

word “complex” will be used to denote “imaginary” in the historical sense, i.e. numbers <strong>of</strong> the form<br />

a + b √ −1 where a, b are real and b �= 0.<br />

3 Since the time <strong>of</strong> CARDANO, negative roots had been called false or fictuous roots. <strong>The</strong> true roots <strong>of</strong><br />

which DESCARTES spoke were the positive ones.<br />

4 “Au reste tant les vrayes racines que les fausses ne sont pas tousiours reelles; mais quelquefois<br />

seulement imaginaires; c’est a dire qu’on peut bien tousiours en imaginer autant que iay dit en<br />

chasque Equation; mais qu’il n’y a quelquefois aucune quantité, qui corresponde a celles qu’on<br />

imagine.” (Descartes, 1637, 380); English translation from (Smith and Latham, 1954, 175).<br />

57


58 Chapter 5. Towards unsolvable equations<br />

polynomial equations were complex, i.e. <strong>of</strong> the form a + b √ −1 for real a, b, was raised;<br />

and around the time <strong>of</strong> C. F. GAUSS (1777–1855), the theorem acquired the name <strong>of</strong><br />

the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra.<br />

When G. W. LEIBNIZ (1646–1716) doubted that the polynomial x 4 + c 4 could be<br />

split into two real factors <strong>of</strong> the second degree, 5 the validity <strong>of</strong> the result seemed for a<br />

moment in doubt. L. EULER (1707–1783) demonstrated in 1749 (published 1751) that<br />

the set <strong>of</strong> complex numbers was closed under all algebraic and numerous transcendental<br />

operations. 6 Thus, at least by 1751 it would implicitly be known that √ i = 1+i<br />

√ 2 .<br />

This made LEIBNIZ’S supposed counter-example evaporate, since he factorized his<br />

polynomial as<br />

x 4 + c 4 �<br />

= x 2 �<br />

− ic<br />

2�<br />

x 2 + ic 2�<br />

�<br />

= x − √ � �<br />

ic x + √ � �<br />

ic x + √ � �<br />

−ic x − √ �<br />

−ic .<br />

Numerous prominent mathematicians <strong>of</strong> the eighteenth century — among them no-<br />

tably J. LE R. D’ALEMBERT (1717–1783), EULER, and J. L. LAGRANGE (1736–1813) —<br />

sought to provide pro<strong>of</strong>s that any real polynomial could be split into linear and quadratic<br />

factors which would prove that any imagined roots were indeed complex. In the half-<br />

century 1799–1849 GAUSS gave a total <strong>of</strong> four pro<strong>of</strong>s 7 which, although belonging to an<br />

emerging trend <strong>of</strong> indirect existence pro<strong>of</strong>s, were considered to be superior in rigour<br />

when compared to those <strong>of</strong> his predecessors.<br />

Characterizing roots. <strong>The</strong> pro<strong>of</strong>s <strong>of</strong> the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra were mostly<br />

existence pro<strong>of</strong>s which did not provide any information on the computational aspect.<br />

Other similar, nonconstructive results were also pursued. An important subfield <strong>of</strong> the<br />

theory <strong>of</strong> equations was developed in order to characterize and describe properties <strong>of</strong><br />

the roots <strong>of</strong> a given equation from a priori inspections <strong>of</strong> the equation and without<br />

explicitly knowing the roots.<br />

LAGRANGE’S study <strong>of</strong> the properties <strong>of</strong> the roots <strong>of</strong> particular equations was an<br />

<strong>of</strong>fspring from his attempts to solve higher degree equations through algebraic ex-<br />

pressions (see below). 8 LAGRANGE’S interest in numerical equations, i.e. concrete<br />

equations in which some dependencies among the coefficients can exist, can be di-<br />

vided into three topics: the nature and number <strong>of</strong> the roots, limits for the values <strong>of</strong><br />

these roots, and methods for approximating these. LAGRANGE made use <strong>of</strong> analytic<br />

geometry, function theory, and the Lagrangian calculus in order to investigate these<br />

topics. 9<br />

5 (K. Andersen, 1999, 69).<br />

6 (ibid., 70).<br />

7 GAUSS’ pro<strong>of</strong>s can be found in (C. F. Gauss, 1863–1933, vol. 3) and have been collected in German<br />

translation in (C. F. Gauss, 1890).<br />

8 (Hamburg, 1976, 28).<br />

9 (ibid., 29–30).


5.1. Algebraic solubility before LAGRANGE 59<br />

Elementary symmetric relations. A different example <strong>of</strong> a priori properties <strong>of</strong> the<br />

roots <strong>of</strong> an equation was conceived <strong>of</strong> by men as G. CARDANO (1501–1576), F. VIÈTE<br />

(1540–1603), A. GIRARD (1595–1632), and I. NEWTON (1642–1727) in the sixteenth and<br />

seventeenth centuries. From inspection <strong>of</strong> equations <strong>of</strong> low degree they obtained (gen-<br />

erally by analogy and without general pro<strong>of</strong>s) by a tacit theorem on the factorization<br />

<strong>of</strong> polynomials the dependency <strong>of</strong> the coefficients <strong>of</strong> the equation<br />

on the roots x1, . . . , xn given by<br />

x n + an−1x n−1 + an−2x n−2 + · · · + a1x + a0 = 0<br />

an−1 = − (x1 + · · · + xn)<br />

an−2 = x1x2 + · · · + xn−1xn<br />

.<br />

a1 = ± (x1x2 . . . xn−1 + · · · + x2x3 . . . xn)<br />

a0 = ∓x1x2 . . . xn.<br />

(5.1)<br />

<strong>The</strong>se equations established the elementary symmetric relations between the roots and<br />

the coefficients <strong>of</strong> an equation. When pro<strong>of</strong>s <strong>of</strong> these relations first emerged, they were<br />

obtained through formal manipulations <strong>of</strong> the tacitly introduced factors and were,<br />

thus, firmly within the established algebraic style.<br />

<strong>The</strong> relations (5.1) were to become a central tool in the theory <strong>of</strong> equations once<br />

NEWTON and E. WARING (∼1736–1798) realized that they were the basic, or ele-<br />

mentary, ones upon which all other symmetric functions <strong>of</strong> the roots depended rationally.<br />

10<br />

5.1 Algebraic solubility before LAGRANGE<br />

Among the multitude <strong>of</strong> possible questions concerning the unknown roots, one is<br />

particularly linked to the question <strong>of</strong> solving equations algebraically. It arose when<br />

mathematicians began investigating the form in which the roots can be written and is<br />

thus a first step in the direction <strong>of</strong> asking general solubility questions. 11<br />

<strong>The</strong> general approach taken in solving equations <strong>of</strong> degrees 2, 3 or 4 had since the<br />

first attempts been to reduce their solution to the solution <strong>of</strong> equations <strong>of</strong> lower degree.<br />

<strong>The</strong> example <strong>of</strong> the third degree equation solved by S. FERRO (1465–1526) around<br />

1515, by N. TARTAGLIA (1499/1500–1557) in 1539, and by CARDANO, who published<br />

the solution in 1545, might be illustrative 12 . Although CARDANO’S arguments and<br />

style were geometric, its algebraic content is presented in algebraic notation in box 1.<br />

10 See section 5.2.4.<br />

11 This aspect shall be dealt with below (see page 62ff) and section 8.4.<br />

12 In the present form, revised to expose central concepts, CARDANO’S solution closely resembles the<br />

young school-boy’s notes found in the section Ligninger af tredje Grads Opløsning (af Cardan) in<br />

ABEL’S notebook (<strong>Abel</strong>, MS:829, 139–141).


60 Chapter 5. Towards unsolvable equations<br />

<strong>The</strong> algebraic reduction <strong>of</strong> the cubic equation When the general third degree equa-<br />

tion<br />

was subjected to the transformation<br />

x 3 + ax 2 + bx + c = 0<br />

x ↦−→ y − a<br />

3<br />

it took the canonical form (in which the term <strong>of</strong> the second highest degree did not<br />

appear)<br />

Letting y = u + v, CARDANO obtained<br />

and the equation could be satisfied if<br />

y 3 + ny + p = 0. (5.2)<br />

0 = (u + v) 3 + n (u + v) + p<br />

= u 3 + v 3 + (3uv + n) (u + v) + p,<br />

� u 3 + v 3 + p = 0, and<br />

3uv + n = 0.<br />

This system <strong>of</strong> equations could easily be reduced to the quadratic system (by letting<br />

U = u 3 , V = v 3 )<br />

or<br />

� U + V = −p<br />

27UV = −n 3<br />

U 2 + Up − n3<br />

27<br />

= 0, (5.3)<br />

the solution <strong>of</strong> which was well known. Thus, U and V could be found, and finding<br />

u, v was only a matter <strong>of</strong> extracting 3 rd roots<br />

giving one <strong>of</strong> the roots y <strong>of</strong> (5.2) as<br />

u = 3√ U and v = 3√ V,<br />

y = u + v = 3√ U + 3√ V.<br />

Box 1: <strong>The</strong> algebraic reduction <strong>of</strong> the cubic equation<br />


5.1. Algebraic solubility before LAGRANGE 61<br />

Purely formal methods were used in reducing the problem <strong>of</strong> the third degree<br />

equation to one <strong>of</strong> solving an equation <strong>of</strong> lower degree, here (5.3). A similar approach<br />

was adopted by L. FERRARI (1522–1565) in 1545 and by R. BOMBELLI (1526–1572) be-<br />

tween 1557 and 1560 to solve the general fourth degree equation. By the seventeenth<br />

century, the search for reductions <strong>of</strong> the general fifth degree equation into equations <strong>of</strong><br />

lower degrees was establishing itself as a prestigious mathematical problem. LEIBNIZ<br />

and E. W. TSCHIRNHAUS (1651–1708) worked on the problem. In 1683 TSCHIRNHAUS<br />

published a procedure which, if applied to the general fifth degree equation, would<br />

reduce it to a binomial one with the help <strong>of</strong> a polynomial equation <strong>of</strong> degree 4. How-<br />

ever, as LEIBNIZ soon demonstrated, determining the coefficients <strong>of</strong> that polynomial<br />

unavoidably involved solving an equation <strong>of</strong> degree 24 which rendered TSCHIRN-<br />

HAUS’S reduction useless for solving the fifth degree equation algebraically. 13 Another<br />

independent and unsuccessful attempt at reducing the fifth degree equation was made<br />

by J. GREGORY (1638–1675), whose proposed reduction was based on a sixth degree<br />

auxiliary (resolvent) equation. 14<br />

<strong>The</strong> procedure <strong>of</strong> reduction to lower degree equations — so naturally suggested<br />

by incomplete induction from low degree equations — thus failed to give results for<br />

higher degree equations. <strong>The</strong> search had largely been conducted in an empirical way<br />

by proposing different reducing functions. It was wanting <strong>of</strong> a general and theoretical<br />

investigation; this was initiated around 1770.<br />

<strong>The</strong> search for resolvent equations conducted throughout the sixteenth, seven-<br />

teenth, and eighteenth centuries is properly seen as the quest to find algebraic solu-<br />

tions for all polynomial equations, thereby explicitly and constructively demonstrat-<br />

ing their algebraic solubility. A polynomial equation <strong>of</strong> degree n such as<br />

x n + an−1x n−1 + an−2x n−2 + · · · + a1x + a0 = 0<br />

is said to be algebraically solvable if its roots x1, . . . , xn can all be expressed by algebraic<br />

expressions in the coefficients a0, . . . , an−1 — the roots must be expressible as finite com-<br />

binations <strong>of</strong> the coefficients and constants using the five algebraic operations addition,<br />

subtraction, multiplication, division, and root extraction.<br />

From the second half <strong>of</strong> the eighteenth century, the diverse and largely empirical<br />

attempts to provide concrete reductions was superseded by theoretical and general<br />

investigations, mainly by LAGRANGE 1770–1771. In the work <strong>of</strong> LAGRANGE, the incli-<br />

nation towards general investigations was accompanied by the idea <strong>of</strong> studying per-<br />

mutations. 15 Both parts were essential in finally establishing that the long sought-for<br />

algebraic solution <strong>of</strong> the quintic equation was impossible.<br />

13 (Kracht and Kreyszig, 1990, 27–28) and (Kline, 1990, 599–600).<br />

14 (Whiteside, 1972, 528).<br />

15 For LAGRANGE’S focus on the general, see (Grabiner, 1981a, 317) and (Grabiner, 1981b, 39).


62 Chapter 5. Towards unsolvable equations<br />

LEONHARD EULER. In his paper (L. Euler, 1732b), read to the St. Petersburg Academy<br />

and published in 1738, EULER gave his solutions to the equations <strong>of</strong> degree 2, 3, and 4<br />

and demonstrated that they could all be written in the form 16<br />

√ A for the second degree equation,<br />

3√ A + 3 √ B for the third degree equation, and<br />

4√ A + 4 √ B + 4√ C for the fourth degree equation,<br />

(5.4)<br />

where the quantities A, B, C were roots in certain resolvent equations <strong>of</strong> lower degree<br />

which could be obtained from the original equation. 17 EULER appears to have been<br />

the first to introduce the term “resolvent” and to attribute to it the central position it<br />

was to take in the future research on the solubility <strong>of</strong> equations. 18<br />

Extending these results, EULER conjectured that the resolvents also existed for the<br />

general equation <strong>of</strong> the fifth degree — and more generally for any higher degree equation<br />

— and that the roots could be expressed in analogy with (5.4). 19<br />

“Although this emphasizes the three particular cases [<strong>of</strong> equations <strong>of</strong> degrees<br />

2, 3, and 4], I, nevertheless, think that one could possibly, not without reason, conclude<br />

that also higher equations would possess similar solving equations. From<br />

the proposed equation<br />

x 5 = ax 3 + bx 2 + cx + d,<br />

I expect to obtain an equation <strong>of</strong> the fourth degree<br />

z 4 = αz 3 − βz 2 + γz − δ<br />

the roots <strong>of</strong> which will be A, B, C, and D,<br />

In the general equation<br />

x = 5√ A + 5√ B + 5√ C + 5√ D.<br />

x n = ax n−2 + bx n−3 + cx n−4 + etc.<br />

the resolvent equation will, I suspect, be <strong>of</strong> the form<br />

z n−1 = αz n−2 − βz n−3 + γz n−4 − etc.,<br />

whose n − 1 known roots will be A, B, C, D, etc.,<br />

z = n√ A + n√ B + n√ C + n√ D + etc.<br />

If this conjecture is valid and if the resolvent equations, which can obviously be<br />

said to have assignable roots, can be determined, I can obtain equations <strong>of</strong> lower<br />

degrees, and in continuing this process produce the true root <strong>of</strong> the equation.” 20<br />

16 (ibid., 7).<br />

17 <strong>The</strong> resolvent equation in the example <strong>of</strong> the third degree equation is (5.3).<br />

18 (F. Rudio, 1921, ix, footnote 2).<br />

19 According to (Eneström, 1912–1913, 346) already LEIBNIZ seemed conviced that the root <strong>of</strong> the general<br />

equation <strong>of</strong> the 5 th degree could be written in the form<br />

x = 5√ A + 5√ B + 5√ C + 5√ D.


5.1. Algebraic solubility before LAGRANGE 63<br />

<strong>The</strong> quotation illustrates how EULER’S conjecture amounted to the algebraic solu-<br />

bility <strong>of</strong> all polynomial equations. Returning to the problem, EULER sought to provide<br />

further evidence for his conjecture. 21<br />

EULER was led to a related problem concerning the multiplicity <strong>of</strong> values <strong>of</strong> rad-<br />

icals. By calculating the number <strong>of</strong> values <strong>of</strong> the multi-valued function consisting <strong>of</strong><br />

n − 1 radicals<br />

n√ A + n √ B + n√ C + n√ D + . . . ,<br />

EULER found that the function had n n−1 essentially different values, which apparently<br />

contradicted the fact that the equation <strong>of</strong> degree n should only have n roots. In a paper<br />

written in 1759, EULER refined his hypothesis <strong>of</strong> 1732 and conjectured that the roots<br />

<strong>of</strong> the resolvent A, B, C, D were dependent. EULER’S new conjecture was that the root<br />

would be expressible in the form<br />

x = ω + A n√ v + B n√ v 2 + C n√ v 3 + · · · + D n√ v n−1 ,<br />

where the coefficients ω, A, B, C, . . . , D were rational functions <strong>of</strong> the coefficients, and<br />

the n − 1 other roots would be obtained by attributing to n√ v the n − 1 other values<br />

a n√ v, b n√ v, c n√ v . . . where a, b, c were the different n th roots <strong>of</strong> unity. 22 As will be illus-<br />

trated in chapter 7.1.2, N. H. ABEL (1802–1829) used a similar kind <strong>of</strong> argument.<br />

20 “8. Ex his etiamsi tribus tantum casibus tamen non sine sufficienti ratione mihi concludere videor<br />

superiorum quoque aequationum dari huiusmodi aequationes resolventes. Sic proposita aequatione<br />

coniicio dari aequationem ordinis quarti<br />

cuius radices si sint A, B, C et D, fore<br />

Et generatim aequationis<br />

x 5 = ax 3 + bx 2 + cx + d<br />

z 4 = αz 3 − βz 2 + γz − δ,<br />

x = 5√ A + 5√ B + 5√ C + 5√ D.<br />

x n = ax n−2 + bx n−3 + cx n−4 + etc.<br />

aequatio resolvens, prout suspicor, erit huius formae<br />

z n−1 = αz n−2 − βz n−3 + γz n−4 − etc.<br />

cuius cognitis radicibus omnibus numero n − 1, quae sint A, B, C, D etc., erit<br />

x = n√ A + n√ B + n√ C + n√ D + etc.<br />

Haec igitur coniectura si esset veritati consentanea atque si aequationes resolventes possent determinari,<br />

cuiusque aequationis in promtu foret radices assignare; perpetuo enim pervenitur ad<br />

aequationem ordine inferiorem hocque modo progrediendo tandem vera aequationis propositae<br />

radix innotescet.” (L. Euler, 1732b, 7–8); for a German translation, see (L. Euler, 1788–1791, vol. 3,<br />

9–10).<br />

21 (F. Rudio, 1921, ix–x).<br />

22 (ibid., x–xi).


64 Chapter 5. Towards unsolvable equations<br />

ALEXANDRE-THÉOPHILE VANDERMONDE. Another very important component <strong>of</strong><br />

the theory <strong>of</strong> equations in the early nineteenth century was the turn towards focusing<br />

on the expressive powers <strong>of</strong> algebraic expressions. This approach can be traced back<br />

to VANDERMONDE who in 1770 presented the Académie des Sciences in Paris with a trea-<br />

tise entitled Mémoire sur la résolution des équations. 23 <strong>The</strong>re, he described the purpose<br />

<strong>of</strong> his investigations:<br />

“One seeks the most simple general values which can conjointly satisfy an<br />

equation <strong>of</strong> a certain degree.” 24<br />

As H. WUSSING has remarked, this weakly formulated program only gained importance<br />

through VANDERMONDE’S use <strong>of</strong> examples from low degree equations. 25<br />

VANDERMONDE’S aim was to build algebraic functions from the elementary symmet-<br />

ric ones which could assume the value <strong>of</strong> any root <strong>of</strong> the given equation. His approach<br />

was very direct, constructive, and computationally based. For example the elementary<br />

symmetric functions in the case <strong>of</strong> the general second degree equation<br />

(x − x1) (x − x2) = x 2 − (x1 + x2) x + x1x2 = 0<br />

are x1x2 and x1 + x2. <strong>The</strong> well known solution <strong>of</strong> the quadratic is<br />

� �<br />

1<br />

x1 + x2 + (x1 + x2)<br />

2<br />

2 �<br />

− 4x1x2 ,<br />

which gives the two roots x1 and x2 when the square root is considered to be a two-<br />

valued function. Similarly, although with greater computational difficulties, VANDER-<br />

MONDE treated equations <strong>of</strong> degree 3 or 4. In those cases, he also constructed algebraic<br />

expressions having the desired properties. When he attacked equations <strong>of</strong> degree 5,<br />

however, he ended up with having to solve a resolvent equation <strong>of</strong> degree 6. Simi-<br />

larly, his approach led from a sixth degree equation to resolvent equations <strong>of</strong> degrees<br />

10 and 15. Having seen the apparent unfruitfulness <strong>of</strong> the approach, VANDERMONDE<br />

abandoned it. Later, the idea <strong>of</strong> studying the algebraic expressions formed from the<br />

elementary symmetric functions became central to ABEL’S research.<br />

Both EULER’S and VANDERMONDE’S approaches are, in spite <strong>of</strong> their apparently<br />

unsuccessful outcome, interesting in interpreting ABEL’S work on the theory <strong>of</strong> equa-<br />

tions. Firstly, ABEL’S pro<strong>of</strong> <strong>of</strong> the impossibility <strong>of</strong> solving the general quintic by<br />

radicals (see chapter 6) is a fusion <strong>of</strong> ideas advanced by LAGRANGE and VANDER-<br />

MONDE, although there is no evidence that ABEL was familiar with VANDERMONDE’S<br />

work. Secondly, ABEL’S attempted general theory <strong>of</strong> algebraic solubility (see chapter<br />

8) bears resemblances to paths followed by EULER, VANDERMONDE and LAGRANGE.<br />

23 (Vandermonde, 1771). This paragraph on VANDERMONDE is largely based on (Wussing, 1969, 52–<br />

53).<br />

24 “On demande les valeurs générales les plus simples qui puissent satisfaire concurremment à une<br />

Équation [sic] d’un degré déterminé.” (Vandermonde, 1771, 366).<br />

25 (Wussing, 1969, 53)


5.2. LAGRANGE’s theory <strong>of</strong> equations 65<br />

Figure 5.1: JOSEPH LOUIS LAGRANGE (1736–1813)<br />

In section 8.4, I demonstrate how ABEL rigorized the assumptions <strong>of</strong> EULER’S conjec-<br />

ture and provided the conjecture with a pro<strong>of</strong>. Before going into ABEL’S impossibility<br />

pro<strong>of</strong>, it is necessary to present important results obtained by ABEL’S predecessors (in-<br />

cluding LAGRANGE) <strong>of</strong> which he made use, and demonstrate the change in approach<br />

and belief that made ABEL’S demonstration possible.<br />

5.2 LAGRANGE’s theory <strong>of</strong> equations<br />

Nobody influenced ABEL’S work on the theory <strong>of</strong> equations more than LAGRANGE.<br />

<strong>The</strong> present section briefly outlines the parts <strong>of</strong> LAGRANGE’S large and very influen-<br />

tial treatise Réflexions sur la résolution algébrique des équations which were <strong>of</strong> particular<br />

importance to ABEL’S work. 26 LAGRANGE’S work is well studied and has <strong>of</strong>ten —<br />

and rightfully so, I think — been seen as one <strong>of</strong> the first major steps towards linking<br />

the theory <strong>of</strong> equations to group theory. 27 However, with the focus mainly on ABEL’S<br />

approach, emphasis is given only to points <strong>of</strong> direct relevance for this.<br />

When LAGRANGE in 1770–1771 had his Réflexions sur la résolution algébrique des<br />

équations published in the Mémoires <strong>of</strong> the Berlin Academy, he was a well established<br />

mathematician held in high esteem. <strong>The</strong> Réflexions was a thorough summary <strong>of</strong> the<br />

26 (Lagrange, 1770–1771).<br />

27 See for instance (Wussing, 1969, 49–52, 54–56), (Kiernan, 1971), (Hamburg, 1976), or (Scholz, 1990,<br />

365–372).


66 Chapter 5. Towards unsolvable equations<br />

nature <strong>of</strong> solutions to algebraic equations which had been uncovered until then. Like<br />

EULER and VANDERMONDE had done, LAGRANGE investigated the known solutions<br />

<strong>of</strong> equations <strong>of</strong> low degrees hoping to discover a pattern feasible to generalizations to<br />

higher degree equations. Where EULER had sought to extend a particular algebraic<br />

form <strong>of</strong> the roots, and VANDERMONDE had tried to generalize the algebraic func-<br />

tions <strong>of</strong> the elementary symmetric functions, LAGRANGE’S innovation was to study<br />

the number <strong>of</strong> values which functions <strong>of</strong> the coefficients could obtain under permu-<br />

tations <strong>of</strong> the roots <strong>of</strong> the equation. Although he exclusively studied the values <strong>of</strong> the<br />

functions under permutations, his results marked a first step in the emerging indepen-<br />

dent theory <strong>of</strong> permutations. In turn, this permutation theory was soon, through its<br />

central role in E. GALOIS’S (1811–1832) theory <strong>of</strong> algebraic solubility, incorporated in<br />

an abstract theory <strong>of</strong> groups which grew out <strong>of</strong> progress made in the nineteenth and<br />

twentieth centuries. 28<br />

<strong>The</strong> work Réflexions sur la résolution algébrique des équations was divided into four<br />

parts reflecting the structure <strong>of</strong> LAGRANGE’S investigation.<br />

1. “On the solution <strong>of</strong> equations <strong>of</strong> the third degree” (Lagrange, 1770–1771, 207–<br />

254)<br />

2. “On the solution <strong>of</strong> equations <strong>of</strong> the fourth degree” (ibid. 254–304)<br />

3. “On the solution <strong>of</strong> equations <strong>of</strong> the fifth and higher degrees” (ibid. 305–355)<br />

4. “Conclusion <strong>of</strong> the preceding reflections with some general remarks concerning<br />

the transformation <strong>of</strong> equations and their reduction to a lower degree” (ibid. 355–<br />

421)<br />

Of these the latter part is <strong>of</strong> particular interest to the following discussion. Its aim<br />

was to provide a link between the number <strong>of</strong> values a function could obtain under<br />

permutations and the degree <strong>of</strong> the associated resolvent equation. Most accounts <strong>of</strong><br />

LAGRANGE’S contribution in the theory <strong>of</strong> equations emphasize the 100 th section deal-<br />

ing with the rational dependence <strong>of</strong> semblables fonctions, a topic which became central<br />

after the introduction <strong>of</strong> GALOIS theory. 29<br />

5.2.1 Formal values <strong>of</strong> functions<br />

Central to LAGRANGE’S treatment <strong>of</strong> the general equations <strong>of</strong> all degrees were his<br />

concepts <strong>of</strong> formal functional equality and formal appearance <strong>of</strong> expressions. 30 LA-<br />

GRANGE considered two rational functions (formally) equal only when they were<br />

given by the same algebraic formulae, in which xy and yx were considered equal<br />

28 (Wussing, 1969).<br />

29 For instance (J. Pierpont, 1898, 333–335) and (Scholz, 1990, 370).<br />

30 (Kiernan, 1971, 46).


5.2. LAGRANGE’s theory <strong>of</strong> equations 67<br />

because both multiplication (and addition) were implicitly assumed to be commuta-<br />

tive, associative, and distributive. This concept <strong>of</strong> formal equality was intertwined<br />

with LAGRANGE’S focus on the formal appearance <strong>of</strong> expressions which made the<br />

form and not the value the important aspect <strong>of</strong> expressions. Denoting the roots <strong>of</strong> the<br />

general µ th degree equation by x1, . . . , xµ, LAGRANGE considered the variables (roots)<br />

to be independent symbols. For example, in LAGRANGE’S view, the two expressions<br />

x1 − x2 and x2 − x1 were always (formally) different, although particular values could<br />

be given to x1 and x2 such that the values <strong>of</strong> the two expressions were equal. <strong>The</strong><br />

independence <strong>of</strong> the symbols x1, . . . , xµ reflected the fact that in a general equation, the<br />

coefficients were considered independent; to treat special, e.g. numerical equations, a<br />

modified approach had to be adapted.<br />

LAGRANGE’S formal approach reflects a general eighteenth century conception <strong>of</strong><br />

polynomials not as functional mappings but as expressions combined <strong>of</strong> various sym-<br />

bols: variables and constants, either known or unknown. LAGRANGE was not par-<br />

ticularly explicit about this notion <strong>of</strong> formal equality which occurs throughout his<br />

investigations; however, he emphasized that<br />

“it is only a matter <strong>of</strong> the form <strong>of</strong> these values and not their absolute [numerical]<br />

quantities.” 31<br />

<strong>The</strong> focus on formal values was lifted when GALOIS saw that in order to address<br />

special equations in which the coefficients were not completely general — some or all<br />

<strong>of</strong> them might be restricted to certain numerical values — he had to consider the nu-<br />

merical equality <strong>of</strong> the symbols in place <strong>of</strong> LAGRANGE’S formal equality.<br />

5.2.2 <strong>The</strong> emergence <strong>of</strong> permutation theory<br />

An important part <strong>of</strong> LAGRANGE’S approach was the introduction <strong>of</strong> symbols denot-<br />

ing the roots which enabled him to treat them as if they had been known. 32 This al-<br />

lowed him to focus his attention on the action <strong>of</strong> permutations on formal expressions<br />

in the roots. LAGRANGE set up a system <strong>of</strong> notation in which<br />

f �� x ′� � x ′′� � x ′′′��<br />

meant that the function f was (formally) altered by any (non-identity) permutation<br />

<strong>of</strong> x ′ , x ′′ , x ′′′ . 33 For instance, the expression x ′ + αx ′′ + α 2 x ′′′ would be altered by any<br />

non-identity permutation if α was an independent symbol (or a number, say, α = 2).<br />

If the function remained unaltered when x ′ and x ′′ were interchanged, LAGRANGE<br />

wrote it as<br />

f �� x ′ , x ′′� � x ′′′�� .<br />

31 “il s’agit ici uniquement de la forme de ces valeurs et non de leur quantité absolute.” (Lagrange,<br />

1770–1771, 385).<br />

32 (Kiernan, 1971, 45). Reminiscences <strong>of</strong> this can also be found with EULER.<br />

33 (Lagrange, 1770–1771, 358).


68 Chapter 5. Towards unsolvable equations<br />

For example, x ′ x ′′ + x ′′′ would remain unaltered by interchanging x ′ and x ′′ , but any<br />

permutation involving x ′′′ would alter it. Finally, if the function was symmetric (i.e.<br />

formally invariant under all permutations <strong>of</strong> x ′ , x ′′ , x ′′′ ), he wrote<br />

f �� x ′ , x ′′ , x ′′′�� .<br />

<strong>The</strong> most important examples <strong>of</strong> such functions were the elementary symmetric func-<br />

tions,<br />

x ′ + x ′′ + x ′′′ ,<br />

x ′ x ′′ + x ′ x ′′′ + x ′′ x ′′′ , and<br />

x ′ x ′′ x ′′′ .<br />

With this notation and his concept <strong>of</strong> formal equality, LAGRANGE derived far-<br />

reaching results on the number <strong>of</strong> (formally) different values which rational func-<br />

tions could assume under all permutations <strong>of</strong> the roots. With the hindsight that the<br />

set <strong>of</strong> permutations form an example <strong>of</strong> an abstract group, a permutation group, one<br />

may see that LAGRANGE was certainly involved in the early evolution <strong>of</strong> permutation<br />

group theory. As we shall see in the following section, he was led by this approach<br />

to Lagrange’s <strong>The</strong>orem, which in modern terminology expresses that the order <strong>of</strong> a sub-<br />

group divides the order <strong>of</strong> the group. However, since LAGRANGE dealt with the ac-<br />

tions <strong>of</strong> permutations on rational functions, he was conceptually still quite far from the<br />

concept <strong>of</strong> groups. LAGRANGE’S contribution to the later field <strong>of</strong> group theory laid in<br />

providing the link between the theory <strong>of</strong> equations and permutations which in turn<br />

led to the study <strong>of</strong> permutation groups from which (in conjunction with other sources)<br />

the abstract group concept was distilled. 34 More importantly, LAGRANGE’S idea <strong>of</strong> in-<br />

troducing permutations into the theory <strong>of</strong> equations provided subsequent generations<br />

with a powerful tool.<br />

5.2.3 LAGRANGE’s resolvents<br />

Another result found by LAGRANGE, <strong>of</strong> which ABEL later made eminent and frequent<br />

use in his investigations, concerned the polynomial having as its roots all the different<br />

values which a given function took when its arguments were permuted. Starting with<br />

the case <strong>of</strong> the quadratic equation having as roots x and y<br />

z 2 + mz + n = 0, (5.5)<br />

LAGRANGE studied the values f [(x) (y)] and f [(y) (x)] which were all the values a<br />

rational function f could obtain under permutations <strong>of</strong> x and y. He then demonstrated<br />

that the equation in t<br />

34 (Wussing, 1969).<br />

Θ = [t − f [(x) (y)]] × [t − f [(y) (x)]] = 0


5.2. LAGRANGE’s theory <strong>of</strong> equations 69<br />

had coefficients which depended rationally on the coefficients m and n <strong>of</strong> the original<br />

quadratic (5.5). 35 This may not be so surprising because today this is easily realized by<br />

observing that the coefficients are symmetric in f [(x) (y)] and f [(y) (x)]. However,<br />

this was precisely the result, which LAGRANGE was about to prove.<br />

Subsequently, LAGRANGE carried out the rather lengthy argument for the general<br />

cubic. <strong>The</strong>reby, he proved that the equation which had the six values <strong>of</strong> f under all<br />

permutations <strong>of</strong> the three roots <strong>of</strong> the cubic as its roots would be rationally expressible<br />

in the coefficients <strong>of</strong> the cubic.<br />

Based on these illustrative cases <strong>of</strong> equations <strong>of</strong> low (second and third) degrees,<br />

LAGRANGE could state the following two results as a general theorem generalizing<br />

the argument sketched for the quadratic above. 36 In the general case, the degree <strong>of</strong> the<br />

equation was denoted µ, and the polynomial having all the values which the given<br />

function f assumes under permutations <strong>of</strong> the µ roots was denoted Θ and its degree<br />

ϖ. LAGRANGE then stated:<br />

1. <strong>The</strong> degree ϖ <strong>of</strong> Θ divides µ! where µ is the degree <strong>of</strong> the proposed equation,<br />

and<br />

2. <strong>The</strong> coefficients <strong>of</strong> the equation Θ = 0 depend rationally on the coefficients <strong>of</strong><br />

the original equation.<br />

In his pro<strong>of</strong> <strong>of</strong> this general theorem, LAGRANGE’S notation and machinery re-<br />

stricted his argument slightly. Because he worked with permutations acting on func-<br />

tions and had no way <strong>of</strong> clarifying the underlying sets <strong>of</strong> permutations, his argu-<br />

ments — which contain all the necessary ideas — may seem to rely on analogies with<br />

the cases <strong>of</strong> low degrees. 37 Be that as it may, by any contemporary standards, LA-<br />

GRANGE’S argument must have been a convincing pro<strong>of</strong> and LAGRANGE’S general<br />

theorem became an immensely important tool in the investigations <strong>of</strong> future alge-<br />

braists.<br />

“From this it is clear that the number <strong>of</strong> different functions [i.e. different values<br />

obtained by permuting the arguments] must increase following the products <strong>of</strong><br />

natural numbers<br />

1, 1.2, 1.2.3, 1.2.3.4, . . . , 1.2.3.4.5 . . . µ.<br />

Having all these functions one will have the roots <strong>of</strong> the equation Θ = 0; thus, if<br />

it is represented as<br />

Θ = t ϖ − Mt ϖ−1 + Nt ϖ−2 − Pt ϖ−3 + · · · = 0,<br />

35 (Lagrange, 1770–1771, 361).<br />

36 (ibid., 369–370).<br />

37 Since LAGRANGE’S pro<strong>of</strong> can easily be adapted to newer frameworks <strong>of</strong> pro<strong>of</strong>, this interpretation<br />

may be a matter <strong>of</strong> personal taste. However, I do see a major difference between LAGRANGE’S pro<strong>of</strong><br />

by analogy and pattern and the pro<strong>of</strong> later given by CAUCHY (see below).


70 Chapter 5. Towards unsolvable equations<br />

one will have ϖ = 1.2.3.4 . . . µ and the coefficient M will equal the sum <strong>of</strong> all the<br />

obtained functions, the coefficient N will equal the sum <strong>of</strong> all products <strong>of</strong> these<br />

functions multiplied two by two, the coefficient P will equal the sum <strong>of</strong> all products<br />

<strong>of</strong> the functions multiplied three by three, and so on. [. . . ]<br />

And since we have demonstrated above that the expression Θ must necessarily<br />

be a rational function <strong>of</strong> t and the coefficients m, n, p, . . . <strong>of</strong> the proposed equation,<br />

it follows that the quantities M, N, P, . . . are necessarily rational functions<br />

<strong>of</strong> m, n, p, . . . which one can find directly as we have seen done in the preceding<br />

sections.” 38<br />

Expressed in modern mathematical language, the first part <strong>of</strong> the above result is<br />

the equivalent <strong>of</strong> the Lagrange’s <strong>The</strong>orem <strong>of</strong> group theory, which states that the order<br />

<strong>of</strong> any subgroup divides the order <strong>of</strong> the group. As we shall see in section 5.6, the<br />

first general pro<strong>of</strong> was given by A.-L. CAUCHY (1789–1857) based on his approach to<br />

working with permutations.<br />

<strong>The</strong> second part <strong>of</strong> the result was used extensively by ABEL, although he never<br />

gave references when applying it. ABEL used the result in a form equivalent to the<br />

following theorem, formulated in a compact notation.<br />

<strong>The</strong>orem 1 If φ � �<br />

x1, . . . , xµ is a rational function which takes on the values φ1, . . . , φϖ under<br />

all permutations <strong>of</strong> its arguments x1, . . . , xµ and the equation<br />

Θ =<br />

ϖ<br />

ϖ<br />

∏ (v − φk) = ∑ Akv k=1<br />

k=0<br />

k<br />

is formed, then all the coefficients A0, . . . , Aϖ are symmetric functions <strong>of</strong> x1, . . . , xµ. ✷<br />

(5.6)<br />

<strong>The</strong> link between the above theorem as used by ABEL and LAGRANGE’S second<br />

result can be obtained through a result which I denote Waring’s formulae. <strong>The</strong>se for-<br />

mulae, obtained by NEWTON and WARING by different routes and described in the<br />

next section, were incorporated by LAGRANGE in his work and must have been ac-<br />

cepted as common knowledge in LAGRANGE’S era. As quoted above, LAGRANGE’S<br />

38 “D’où l’on voit clairement que le nombre des fonctions différentes doit croître suivant les produits<br />

des nombres naturels<br />

1, 1.2, 1.2.3, 1.2.3.4, . . . , 1.2.3.4.5 . . . µ.<br />

Ayant toutes ces fonctions on aura donc les racines de l’équation Θ = 0; de sorte que, si on la<br />

représente par<br />

Θ = t ϖ − Mt ϖ−1 + Nt ϖ−2 − Pt ϖ−3 + · · · = 0,<br />

on aura ϖ = 1.2.3.4 . . . µ; et le coefficient M sera égal à la somme de toutes les fonctions trouvées,<br />

le coefficient N égal à la somme de tous les produits de ces fonctions multipliées deux à deux, le<br />

coefficient P égal à la somme de tous les produits des mêmes fonctions multipliées trois à trois, et<br />

ainsi de suite. [. . . ]<br />

Et comme nous avons démontré ci-dessus que l’expression de Θ doit être nécessairement une fonction<br />

rationnelle de t et des coefficients m, n, p, . . . de l’équation proposée, il s’ensuit que les quantités<br />

M, N, P, . . . seront nécessairement des fonctions ratinnelles de m, n, p, . . . qu’on pourra trouver<br />

directement, comme nous l’avons pratiqué dans les Sections précédentes.” (Lagrange, 1770–1771,<br />

369).


5.2. LAGRANGE’s theory <strong>of</strong> equations 71<br />

Figure 5.2: EDWARD WARING (1734–1798)<br />

theorem stated that the coefficients, here A0, . . . , A ¯ω, were rational functions <strong>of</strong> the co-<br />

efficients <strong>of</strong> the given equation. By Waring’s formulae, any such rational function <strong>of</strong> the<br />

coefficients was a symmetric function <strong>of</strong> the roots.<br />

5.2.4 Waring’s formulae<br />

<strong>The</strong> elementary symmetric functions <strong>of</strong> the roots <strong>of</strong> an equation, which since the times<br />

<strong>of</strong> VIÈTE and NEWTON had been known to agree with the coefficients (see section 5),<br />

was seen by the little known British mathematician WARING to provide a basis for the<br />

study <strong>of</strong> all symmetric functions <strong>of</strong> the equation’s roots. In his Miscellanea analytica <strong>of</strong><br />

1762, WARING demonstrated that all rational symmetric functions <strong>of</strong> the roots could<br />

be expressed rationally in the elementary symmetric functions. 39 In his other more<br />

influential work Meditationes algebraicae, 40 to which LAGRANGE referred, 41 the result<br />

was contained in the first chapter. <strong>The</strong>re, WARING dealt with the determination <strong>of</strong> the<br />

power sums <strong>of</strong> the roots x1, . . . , xµ (modern notation)<br />

39 (Waerden, 1985, 76–77).<br />

40 (Waring, 1770).<br />

41 (Lagrange, 1770–1771, 369–370).<br />

µ<br />

∑ x<br />

k=1<br />

m k<br />

for integer m


72 Chapter 5. Towards unsolvable equations<br />

from the coefficients <strong>of</strong> the equation. 42 <strong>The</strong> solution was the so-called Waring’s Formu-<br />

lae giving a procedure alternative to one given earlier by NEWTON. From this, WAR-<br />

ING proceeded to show how any function <strong>of</strong> the roots <strong>of</strong> the form (modern notation<br />

writing Σµ for the symmetric group)<br />

∑ x<br />

σ∈Σµ<br />

a1 σ(1) xa2 . . . xaµ<br />

σ(2) σ(µ) with a1, . . . , aµ non-negative integers (5.7)<br />

could be expressed as an integral function <strong>of</strong> the power sums <strong>of</strong> the roots. 43 Thus,<br />

WARING had demonstrated that all rational and symmetric functions <strong>of</strong> x1, . . . , xµ de-<br />

pended rationally on the power sums and thus on the coefficients <strong>of</strong> the equation by<br />

the preceding result. 44<br />

Although this important theorem was stated and proved by WARING, it entered<br />

the mathematical toolbox <strong>of</strong> the early nineteenth century mainly through LAGRANGE’S<br />

adaption <strong>of</strong> it in his Réflexions (which is the reason for treating it at this place). While<br />

WARING’S notation and letter-manipulating approach had hampered his presentation,<br />

LAGRANGE dealt with it in a clear and integrated fashion in the Réflexions. 45 <strong>The</strong>re, he<br />

observed that if the function f had the form<br />

f<br />

��x ′ ′′<br />

, x � � �<br />

x<br />

′′′�<br />

x iv�<br />

. . .<br />

indicating that x ′ and x ′′ appeared symmetrically, the roots <strong>of</strong> the equation Θ = 0<br />

(5.6) would be equal in pairs, whereby the degree could be reduced to µ!<br />

2 . After briefly<br />

studying a few other types <strong>of</strong> functions f , LAGRANGE concluded that if f had the form<br />

f<br />

�<br />

,<br />

��<br />

x ′ , x ′′ , x ′′′ , . . . , x (µ)��<br />

,<br />

i.e. was a symmetric function <strong>of</strong> the roots, the degree <strong>of</strong> the equation Θ = 0 (5.6) could<br />

be reduced to one and f would be given rationally in the coefficients <strong>of</strong> the original<br />

equation.<br />

5.3 Solubility <strong>of</strong> cyclotomic equations<br />

Thirty years after LAGRANGE’S creative studies on known solutions to low degree<br />

equations, and in particular properties <strong>of</strong> rational functions under permutations <strong>of</strong><br />

their arguments, another great master published a work <strong>of</strong> pr<strong>of</strong>ound influence on<br />

early nineteenth century mathematics. In Göttingen, GAUSS was located at a physical<br />

distance from the emerging centers <strong>of</strong> mathematical research in Paris and Berlin. By<br />

1801, the Parisian mathematicians had for some time been publishing their results in<br />

42 (Waring, 1770, 1–5).<br />

43 (ibid., 9–18).<br />

44 By formal equality, all terms <strong>of</strong> the same degree would have to have identical coefficients, and thus<br />

any rational symmetric function could be decomposed into functions <strong>of</strong> the form (5.7).<br />

45 (Lagrange, 1770–1771, 371–372).


5.3. Solubility <strong>of</strong> cyclotomic equations 73<br />

Figure 5.3: CARL FRIEDRICH GAUSS (1777–1855)<br />

French — and, within a generation, the German mathematicians would also be writ-<br />

ing in their native language, at least for publications intended for A. L. CRELLE’S<br />

(1780–1855) Journal für die reine und angewandte Mathematik. But GAUSS published his<br />

fundamental work Disquisitiones arithmeticae as a Latin monograph as was still cus-<br />

tomary for his generation <strong>of</strong> German scholars.<br />

<strong>The</strong> book cosisted <strong>of</strong> seven sections, although allusions and references were made<br />

to an eighth section which GAUSS never completed for publication. 46 <strong>The</strong> main part<br />

was concerned with the theory <strong>of</strong> congruences, the theory <strong>of</strong> forms, and related num-<br />

ber theoretic investigations. Together, these topics provided a new foundation, em-<br />

phasis, and disciplinary independence — as well as a wealth <strong>of</strong> results — for nine-<br />

teenth century number theorists — in particular G. P. L. DIRICHLET (1805–1859) —<br />

to elaborate. In dealing with the classification <strong>of</strong> forms and describing primitive roots,<br />

GAUSS made use <strong>of</strong> “implicit group theory”. 47 Despite the fact that both LAGRANGE<br />

and GAUSS worked with particular instances <strong>of</strong> groups, neither <strong>of</strong> them introduced<br />

an abstract concept <strong>of</strong> groups.<br />

One <strong>of</strong> the new tools applied by GAUSS in the theory <strong>of</strong> congruences was that <strong>of</strong><br />

primitive roots. In the articles 52–57, GAUSS gave his exposition <strong>of</strong> EULER’S treatment<br />

<strong>of</strong> primitive roots. A primitive root k <strong>of</strong> modulus µ is an integer 1 < k < µ such that<br />

46 (C. F. Gauss, 1863–1933, vol. 1, 477). It is, however, included among the Nachlass in the second<br />

volume <strong>of</strong> the Werke (ibid.).<br />

47 (Wussing, 1969, 40–44).


74 Chapter 5. Towards unsolvable equations<br />

the set <strong>of</strong> remainders <strong>of</strong> its powers k 1 , k 2 , . . . , k µ−1 modulo µ coincides with the set<br />

{1, 2, . . . , µ − 1}, possibly in a different order. A central result obtained was the existence<br />

<strong>of</strong> the p − 1 different primitive roots 1, 2, . . . ,<br />

p − 1 <strong>of</strong> modulus p if p were assumed to be prime.<br />

5.3.1 <strong>The</strong> division problem for the circle<br />

In the last section <strong>of</strong> his Disquisitiones arithmeticae, 48 GAUSS turned his investigations<br />

toward the equations defining the division <strong>of</strong> the periphery <strong>of</strong> the circle into equal<br />

parts. He was interested in the ruler-and-compass constructibility 49 <strong>of</strong> regular poly-<br />

gons and was therefore led to study in detail how, i.e. by the extraction <strong>of</strong> which roots,<br />

the binomial equations <strong>of</strong> the form<br />

x n − 1 = 0 (5.8)<br />

could be solved algebraically. If the roots <strong>of</strong> this equation could be constructed by ruler<br />

and compass, then so could the regular p-gon. It is evident from GAUSS’ mathematical<br />

diary that this problem had occupied him from a very early stage in his mathemati-<br />

cal career and had been a deciding factor in his choice <strong>of</strong> mathematics over classical<br />

philology. 50 <strong>The</strong> very first entry in his mathematical progress diary from 1796 read:<br />

“[1] <strong>The</strong> principles upon which the division <strong>of</strong> the circle depend, and geometrical<br />

divisibility <strong>of</strong> the same into seventeen parts, etc. [1796] March 30 Brunswick.” 51<br />

In his introductory remarks, GAUSS noticed that the approach which had led him<br />

to the division <strong>of</strong> the circle could equally well be applied to the division <strong>of</strong> other tran-<br />

scendental curves <strong>of</strong> which he gave the lemniscate as an example.<br />

“<strong>The</strong> principles <strong>of</strong> the theory which we are going to explain actually extend<br />

much farther than we will indicate. For they can be applied not only to circular<br />

functions but just as well to other transcendental functions, e.g. to those<br />

which depend on the integral � � 1/ √ � 1 − x 4�� dx and also to various types <strong>of</strong><br />

congruences.” 52<br />

48 (C. F. Gauss, 1801). For historical studies, see for instance (Wussing, 1969, 37–44), (Schneider, 1981,<br />

37–50), or (Scholz, 1990, 372–376).<br />

49 Throughout, I refer to Euclidean construction, i.e. by ruler and compass when I speak <strong>of</strong> constructions<br />

or constructibility.<br />

50 (Biermann, 1981, 16).<br />

51 “[1.] Principia quibus innititur sectio circuli, ac divisibilitas eiusdem geometrica in septemdecim<br />

partes etc. [1796] Mart. 30. Brunsv[igae]” (C. F. Gauss, 1981, 21, 41); English translation from (J. J.<br />

Gray, 1984, 106).<br />

52 “Ceterum principia theoriae, quam exponere aggredimur, multo latius patent, quam hic extenduntur.<br />

Namque non solum ad functiones circulares, sed pari successu ad multas functiones transscendentes<br />

applicari possunt, e.g. ad eas, quae ab integrali �<br />

√<br />

dx pendent, praetereaque etiam ad<br />

(1−x4 )<br />

varia congruentiarum genera.” (C. F. Gauss, 1801, 412–413); English translation from (C. F. Gauss,<br />

1986, 407).


5.3. Solubility <strong>of</strong> cyclotomic equations 75<br />

However, as he was preparing to write a treatise on these topics, GAUSS chose to<br />

leave this extension out <strong>of</strong> the Disquisitiones. GAUSS never wrote the promised treatise,<br />

and after ABEL had published his first work on elliptic functions culminating in the<br />

division <strong>of</strong> the lemniscate, 53 GAUSS gave him credit for bringing these results into<br />

print. 54<br />

A first simplification <strong>of</strong> the study <strong>of</strong> the constructibility <strong>of</strong> a regular n-gon was<br />

made when GAUSS observed that he needed only to consider cases in which n was a<br />

prime since any polygon with a composite number <strong>of</strong> edges could be constructed from<br />

the polygons with the associated prime numbers <strong>of</strong> edges. Equations expressing the<br />

sine, the cosine, and the tangent were well known, but none <strong>of</strong> those were as suitable<br />

for GAUSS’ purpose as the equation x n − 1 = 0 <strong>of</strong> which he knew that the roots were 55<br />

cos 2kπ 2kπ<br />

+ i sin = 1 when 0 ≤ k ≤ n − 1.<br />

n n<br />

Inspecting these roots, GAUSS observed that the equation xn − 1 = 0 for odd n<br />

had a single real root, x = 1, and the remaining imaginary roots were all given by the<br />

equation<br />

X = xn − 1<br />

x − 1 = xn−1 + x n−2 + · · · + x + 1 = 0, (5.9)<br />

the roots <strong>of</strong> which GAUSS thought <strong>of</strong> as forming the complex Ω. When GAUSS used<br />

the term “complex” (Latin: complexum) he thought <strong>of</strong> it as a collection <strong>of</strong> objects<br />

(here roots) without any structure imposed. 56 Initially, GALOIS used the French term<br />

groupe in a similar (naive) way before it later gradually acquired its status as a mathe-<br />

matical term. 57 This evolution <strong>of</strong> everyday words into mathematical concepts appears<br />

to be a recurring feature <strong>of</strong> mathematics in the early nineteenth century when so many<br />

terms became precisely defined and re-defined. 58 GAUSS demonstrated that if r desig-<br />

nated any root in Ω, all roots <strong>of</strong> (5.8) could be expressed as powers <strong>of</strong> r, thereby saying<br />

that any root in Ω was a primitive n th root <strong>of</strong> unity.<br />

5.3.2 Irreducibility <strong>of</strong> the equation xn −1<br />

x−1<br />

= 0<br />

An interesting feature <strong>of</strong> GAUSS’ approach was his focusing on the complex or system <strong>of</strong><br />

roots instead <strong>of</strong> the individual roots. This slight shift in the conception <strong>of</strong> roots enabled<br />

GAUSS (as it had enabled LAGRANGE) 59 to study properties <strong>of</strong> the equations which<br />

53 (N. H. <strong>Abel</strong>, 1827b).<br />

54 (Crelle→<strong>Abel</strong>, 1828/05/18. N. H. <strong>Abel</strong>, 1902a, 62).<br />

55 GAUSS wrote P (periphery) for 2π; however the use <strong>of</strong> i for √ −1 is his.<br />

56 Later, in 1831, GAUSS introduced the term complex numbers to denote numbers which had hitherto<br />

been designated imaginary; (Gericke, 1970, 57). I fail to see any connection between the term complexum<br />

as used here and the later, technical term.<br />

57 (Wussing, 1969, 78).<br />

58 See also section 21.2.<br />

59 In the preface, GAUSS briefly described his debt to the number theoretic investigations <strong>of</strong> the “modern<br />

authors” FERMAT, EULER, LAGRANGE, and LEGENDRE. If GAUSS had read LAGRANGE’S Réflexions,<br />

he did not refer explicitly to it.


76 Chapter 5. Towards unsolvable equations<br />

could only be captured in studies <strong>of</strong> the entire system <strong>of</strong> roots. To GAUSS, the most<br />

important properties were those <strong>of</strong> decomposability and irreducibility. GAUSS demon-<br />

strated through an ad hoc argument that the function X (5.9) could not be decomposed<br />

into polynomials <strong>of</strong> lower degree with rational coefficients. In modern terminology,<br />

he proved that the polynomial X was irreducible over Q.<br />

GAUSS’ pro<strong>of</strong> assumed that the function<br />

X = x n−1 + x n−2 + · · · + x + 1<br />

was divisible by a function <strong>of</strong> lower degree<br />

P = x λ + Ax λ−1 + Bx λ−2 + · · · + Kx + L, (5.10)<br />

in which the coefficients A, B, . . . , K, L were rational numbers. Assuming X = PQ,<br />

GAUSS introduced the two systems <strong>of</strong> roots P and Q <strong>of</strong> P and Q respectively. From<br />

these two systems GAUSS defined another two consisting <strong>of</strong> the reciprocal roots60 �<br />

ˆP = r −1 � �<br />

: r ∈ P and ˆQ = r −1 �<br />

: r ∈ Q .<br />

Although GAUSS consistently termed the roots <strong>of</strong> ˆP and ˆQ reciprocal roots, it is easy for<br />

us to see that they are what we would term conjugate roots since any root in P has unit<br />

length.<br />

GAUSS split the subsequent argument into four different cases. <strong>The</strong> opening one<br />

is the most interesting one, namely the case in which P = ˆP, i.e. when all roots <strong>of</strong><br />

P = 0 occur together with their conjugates. It may be surprising that GAUSS consid-<br />

ered other cases as we would expect him to know that in any polynomial with real<br />

coefficients the imaginary roots occur in conjugate pairs. K. JOHNSEN has argued that<br />

this apparently unnecessary complication in GAUSS’ argument can be traced back to<br />

a more general concept <strong>of</strong> irreducibility over fields different from Q, for instance the<br />

field Q (i). 61 If so, there are no explicit hints at such a concept in the Disquisitiones,<br />

and the result which GAUSS proved only served a very specific purpose in his larger<br />

argument, and did not give a general concept, general criteria, or a body <strong>of</strong> theo-<br />

rems concerning irreducibility over Q or any other field. <strong>The</strong> pro<strong>of</strong> relies more on<br />

number theory (higher arithmetic) than on general theorems and criteria concerning<br />

irreducible equations, let alone any general concept <strong>of</strong> fields distinct from the rational<br />

numbers Q.<br />

After observing that P was the product <strong>of</strong> λ 2<br />

(x − cos ω) 2 + sin 2 ω,<br />

paired factors <strong>of</strong> the form<br />

GAUSS concluded that these factors would assume real and positive values for all real<br />

values <strong>of</strong> x, which would then also apply to the function P (x). He then formed n − 1<br />

60 <strong>The</strong> notation ˆP and ˆQ for these is mine.<br />

61 (Johnsen, 1984).


5.3. Solubility <strong>of</strong> cyclotomic equations 77<br />

auxiliary equations 62<br />

P (k) = 0 where 1 ≤ k ≤ n − 1<br />

defined by their root systems P (k) consisting <strong>of</strong> k th powers <strong>of</strong> the roots <strong>of</strong> P = 0,<br />

P (k) =<br />

�<br />

r k �<br />

: r ∈ P ,<br />

P (k) (x) = ∏ (x − s) = ∏<br />

s∈P (k)<br />

r∈P<br />

�<br />

x − r k�<br />

.<br />

Following the introduction <strong>of</strong> the numbers pk defined by<br />

pk = P (k) �<br />

(1) = ∏ (1 − r) = ∏ 1 − r<br />

s∈P (k)<br />

r∈P<br />

k�<br />

,<br />

GAUSS used properties derived in a previous article to establish<br />

Furthermore,<br />

n−1<br />

∏ P<br />

k=1<br />

(k) (x) =<br />

n−1<br />

∏ pk =<br />

k=1<br />

n−1<br />

∏<br />

∏<br />

k=1 r∈P<br />

n−1<br />

∑ pk =<br />

k=1<br />

�<br />

x − r k�<br />

= ∏ r∈P<br />

n−1<br />

∑ P<br />

k=1<br />

(k) (1) = nA. (5.11)<br />

n−1 �<br />

∏ x − r<br />

k=1<br />

k� = ∏ X = X<br />

r∈P<br />

λ , and<br />

n−1<br />

∏ P<br />

k=1<br />

(k) (1) = X λ (1) = n λ since X (1) = n.<br />

From the article describing the construction <strong>of</strong> an equation with the k th powers<br />

<strong>of</strong> the roots <strong>of</strong> a given equation as its roots, GAUSS knew that the coefficients <strong>of</strong><br />

P (1) , . . . , P (n−1) would be rational numbers if the coefficients <strong>of</strong> P were rationals. Much<br />

earlier, in article 42, he had furthermore demonstrated that the product <strong>of</strong> two poly-<br />

nomials with rational but not integral coefficients could not be a polynomial with<br />

integral coefficients. Since X had integral coefficients and P had rational coefficients<br />

by assumption, it followed that the coefficients <strong>of</strong> P (1) , . . . , P (n−1) would indeed be in-<br />

tegers, since any P (k) was a factor <strong>of</strong> X λ with rational coefficients. Consequently, the<br />

quantities p k would have to be integral, and since their product was n λ and there were<br />

n − 1 > λ <strong>of</strong> them, at least n − 1 − λ <strong>of</strong> the quantities p k would have to be equal to 1<br />

and the others would have to equal n or some power <strong>of</strong> n since n was assumed to be<br />

prime. But if the number <strong>of</strong> quantities equal to 1 was g it would follow that<br />

n−1<br />

∑ pk ≡ g<br />

k=1<br />

( mod n) ,<br />

which GAUSS saw would contradict (5.11) since 0 < g < n.<br />

62 <strong>The</strong> notation P (k) and P (k) is mine.


78 Chapter 5. Towards unsolvable equations<br />

<strong>The</strong> other cases, which in the presently adopted notation can be described as<br />

2.P �= ˆP and P ∩ ˆP �= ∅,<br />

3.Q ∩ ˆQ �= ∅, and<br />

4.P ∩ ˆP = ∅ and Q ∩ ˆQ = ∅,<br />

could all be brought to a contradiction, either directly or by referring to the first case<br />

described above.<br />

<strong>The</strong> fruitfulness <strong>of</strong> the pro<strong>of</strong> <strong>of</strong> the irreducibility <strong>of</strong> X was that it demonstrated<br />

that if X was decomposed into factors <strong>of</strong> lower degrees (such as 5.10) some <strong>of</strong> these<br />

had to have irrational coefficients. Thus any attempt at determining the roots would<br />

have to involve equations <strong>of</strong> degree higher than one. <strong>The</strong> purpose <strong>of</strong> the following<br />

investigation was to gradually reduce the degree <strong>of</strong> these equations to minimal values<br />

by refining the system <strong>of</strong> roots.<br />

5.3.3 Outline <strong>of</strong> GAUSS’s pro<strong>of</strong><br />

Continuing from the result above that any root r in Ω was a primitive n th root <strong>of</strong> unity,<br />

GAUSS wrote [1] , [2] , . . . , [n − 1] for the associated powers <strong>of</strong> r. He introduced the con-<br />

cept <strong>of</strong> periods by defining the period ( f , λ) to be the set <strong>of</strong> the roots [λ] , [λg] , . . . , � λg f −1� ,<br />

where f was an integer, λ an integer not divisible by n, and g a primitive root <strong>of</strong> the<br />

modulus n. Connected to the period, he introduced the sum <strong>of</strong> the period, which he also<br />

designated ( f , λ),<br />

( f , λ) =<br />

f −1 �<br />

∑ λg<br />

k=0<br />

k�<br />

,<br />

and the first result, which he stated concerning these periods, was their independence<br />

<strong>of</strong> the choice <strong>of</strong> g.<br />

Throughout the following argument, GAUSS let g designate a primitive root <strong>of</strong><br />

modulus n and constructed a sequence <strong>of</strong> equations through which the periods (1, g),<br />

i.e. the roots in X = 0 (5.9), could be determined. Assuming that the number n − 1<br />

had been decomposed into primes as<br />

n − 1 =<br />

u<br />

∏ pk, k=1<br />

GAUSS partitioned the roots <strong>of</strong> Ω into n−1<br />

p 1 periods, each <strong>of</strong> p1 terms. From these, he<br />

formed p1 equations X ′ = 0 having the n−1<br />

p 1 sums <strong>of</strong> the form (p1, λ) as its roots. By<br />

a central theorem proved using symmetric functions, 63 he proved that the coefficients<br />

<strong>of</strong> these latter equations depended upon the solution <strong>of</strong> yet another equation <strong>of</strong> de-<br />

gree p1. Thus the solution <strong>of</strong> the original equation <strong>of</strong> degree n had been reduced to<br />

solving p1 equations X ′ = 0 each <strong>of</strong> degree n−1<br />

p 1 and a single equation <strong>of</strong> degree p1.<br />

63 (C. F. Gauss, 1801, §350).


5.3. Solubility <strong>of</strong> cyclotomic equations 79<br />

By repeating the procedure, GAUSS could solve the equation X ′ = 0 by solving p2<br />

equations <strong>of</strong> degree n−1<br />

p 1p2 and a single equation <strong>of</strong> degree p2. Similarly, the procedure<br />

could be iterated further until the solution <strong>of</strong> the equation X = 0 <strong>of</strong> degree n − 1 had<br />

been reduced to solving u equations <strong>of</strong> degrees p1, p2, . . . , pu since the other equations<br />

would ultimately have degree 1.<br />

A special case emerged if n − 1 was a power <strong>of</strong> 2. It was well known that square<br />

roots could always be constructed by ruler and compass. <strong>The</strong>refore, if n had the form<br />

n = 1 + 2 k ,<br />

the construction <strong>of</strong> the roots <strong>of</strong> (5.8) could be carried out by ruler and compass. By ap-<br />

plying this to k = 4, GAUSS demonstrated that the regular 17-gon could be constructed<br />

by ruler and compass giving the first new constructible regular polygon since the time<br />

<strong>of</strong> EUCLID (∼295 B.C.) (∼295BC). 64<br />

By the same argument he had devised to consider only prime values <strong>of</strong> n, GAUSS<br />

could also conclude that the construction <strong>of</strong> the regular n-gon was possible by ruler<br />

and compass when n had the form<br />

h<br />

m<br />

n = 2 ∏ (1 + 2<br />

k=1<br />

uk) when {u k} was a set <strong>of</strong> distinct integers such that {1 + 2 u k} were primes, the so-called<br />

Fermat primes. <strong>The</strong> converse implication, that only such n-gons were constructible,<br />

was claimed without detailed pro<strong>of</strong> by GAUSS:<br />

“Whenever n − 1 involves prime factors other than 2, we are always led to<br />

equations <strong>of</strong> higher degree, namely to one or more cubic equations when 3 appears<br />

once or several times among the prime factors <strong>of</strong> n − 1, to equations <strong>of</strong> the<br />

fifth degree when n − 1 is divisible by 5, etc. We can show with all rigor that<br />

these higher-degree equations cannot be avoided in any way nor can they be<br />

reduced to lower-degree equations. <strong>The</strong> limits <strong>of</strong> the present work exclude this<br />

demonstration here, but we issue this warning lest anyone attempt to achieve geometric<br />

constructions for sections other than the ones suggested by our theory (e.g.<br />

sections into 7, 11, 13, 19, etc. parts) and so spend his time uselessly.” 65<br />

64 GAUSS, himself, was very aware <strong>of</strong> the progress he had made, see (C. F. Gauss, 1986, 458) and<br />

(Schneider, 1981, 38–39). As always, the date given for EUCLID is taken from the Dictionary <strong>of</strong> Scientific<br />

Biography.<br />

65 “Quoties autem n − 1 alios factores primos praeter 2 implicat, semper ad aequationes altiores deferimur;<br />

puta ad unam pluresve cubicas, quando 3 semel aut pluries inter factores primos ipsius<br />

n − 1 reperitur, ad aequationes quinti gradus, quando n − 1 divisibilis est per 5 etc., omnique rigore<br />

demonstrare possumus, has aequationes elevatas nullo modo nec evitari nec ad inferiores reduci<br />

posse, etsi limites huius operis hanc demonstrationem hic tradere non patiantur, quod tamen monendum<br />

esse duximus, ne quis adhuc alias sectiones praeter eas, quas theoria nostra suggerit, e.g.<br />

sectiones in 7, 11, 13, 19 etc. partes, ad constructiones geometricas perducere speret, tempusque<br />

inutiliter terat.” (C. F. Gauss, 1801, 462); English translation from (C. F. Gauss, 1986, 459). Bold-face<br />

has been substituted for the original small-caps.


80 Chapter 5. Towards unsolvable equations<br />

<strong>The</strong> class <strong>of</strong> equations (the cyclotomic ones), which GAUSS had demonstrated had<br />

constructible roots, was also interesting from the point <strong>of</strong> algebraic solubility <strong>of</strong> equa-<br />

tions. In his pro<strong>of</strong>, GAUSS had demonstrated that they were indeed solvable by radi-<br />

cals including only square roots, whereby the first new non-elementary class <strong>of</strong> solv-<br />

able equations <strong>of</strong> high degrees had been established. By the time GAUSS wrote his<br />

Disquisitiones, he had come to suspect that not all equations were solvable by radicals.<br />

A few years later, ABEL could consider this newly found class to be a special example<br />

<strong>of</strong> equations having the nice property <strong>of</strong> being algebraically solvable.<br />

5.4 Belief in algebraic solubility shaken<br />

In the seventeenth century, the belief in the algebraic solubility (in radicals) <strong>of</strong> all poly-<br />

nomial equations seems to have been in little dispute. <strong>The</strong> question <strong>of</strong> solubility was<br />

not an issue when the prominent mathematicians such as TSCHIRNHAUS searched for<br />

a general solution. Half-way through the eighteenth century the problem had taken<br />

a slight turn when EULER in 1732 proposed to investigate the hypothesis that the roots<br />

<strong>of</strong> the general n th degree equation could be written as a sum <strong>of</strong> n − 1 root extractions<br />

<strong>of</strong> degree n − 1. Although he advanced this as a hypothesis and his search for definite<br />

pro<strong>of</strong> was in vain, he based his 1749 “pro<strong>of</strong>” <strong>of</strong> the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra on<br />

the belief that any polynomial equation could be reduced to pure equations. 66 Towards<br />

the end <strong>of</strong> the eighteenth century, the outspoken beliefs <strong>of</strong> the most prominent math-<br />

ematicians had changed, though. Mathematicians with a keen interest in the subject<br />

started to suspect that the reduction to pure equations was beyond — not only their<br />

grasp — but the power <strong>of</strong> their existing tools. At the turn <strong>of</strong> the century one <strong>of</strong> the most<br />

influential mathematicians, GAUSS, declared the reduction to be outright impossible.<br />

<strong>The</strong> belief in the algebraic solubility <strong>of</strong> general equations did not vanish completely<br />

with GAUSS’ proclamation <strong>of</strong> its impossibility. In chapter 6.9, where I discuss the<br />

reception <strong>of</strong> ABEL’S work on the theory <strong>of</strong> equations, I shall also discuss the “inertia”<br />

<strong>of</strong> the mathematical community in this respect.<br />

5.4.1 “Infinite labor”<br />

On the British Isles, WARING had recognized patterns which led him to the known<br />

solutions <strong>of</strong> low degree equations. Based on analogies, he thought that solutions to<br />

all equations could be formed but that the amount <strong>of</strong> involved computations would<br />

explode beyond anything practical.<br />

66 By pure equations EULER (and with him GAUSS) meant equations describing explicit functions, i.e. a<br />

pure equation for x is <strong>of</strong> the form<br />

x = some expression.


5.4. Belief in algebraic solubility shaken 81<br />

“From the preceding examples and earlier observations, we may compose resolutions<br />

appropriate to any given equation; but in equations <strong>of</strong> the fifth and higher<br />

degree the calculations require practically infinite labor.” 67<br />

<strong>The</strong> inner tension in WARING’S statement — that the solution was possible in prin-<br />

ciple but perhaps not in practice — is confusing. It remains unclear exactly what it<br />

meant to him that he could construct solutions but that the effort required would be<br />

infinite.<br />

In France, LAGRANGE felt strong confidence in his approach to the study <strong>of</strong> poly-<br />

nomial equations. His detailed studies <strong>of</strong> low degree equations led him to the conclu-<br />

sion that each root x1, . . . , xn <strong>of</strong> the general equation <strong>of</strong> degree n could be expressed<br />

through a resolvent equation <strong>of</strong> degree n − 1 which had the roots<br />

n<br />

∑ ω<br />

k=1<br />

k j xk (for j = 1, . . . , n − 1)<br />

where ω1, . . . , ωn−1 were the imaginary n th roots <strong>of</strong> unity. When LAGRANGE sought<br />

to prove this result for the fifth degree equation, however, he had to accept that his<br />

effort was inconclusive. If such a reduction was to be possible at all, other resolvents<br />

were required.<br />

“It thus appears that from this one can conclude by induction that every equation,<br />

<strong>of</strong> whatever degree, will also be solvable with the help <strong>of</strong> a resolvent [equation]<br />

whose roots are represented by the same formulae<br />

x ′ + y ′′ + y 2 x ′′′ + y 3 x iv + . . . .<br />

But, as we have demonstrated in the previous section in connection with the methods<br />

<strong>of</strong> MM. Euler and Bezout, these lead directly to the same resolvent equations,<br />

there seems to be reason to convince oneself in advance that this conclusion is defective<br />

for the fifth degree. From this it follows, that if the algebraic solution <strong>of</strong><br />

equations <strong>of</strong> degrees higher than four is not impossible, it must depend on certain<br />

functions <strong>of</strong> the roots, which are different from the preceding ones.” 68<br />

Although his investigations had not led to the goal <strong>of</strong> generalizing known solu-<br />

tions <strong>of</strong> low degree equations to a solution to the general fifth degree equation, LA-<br />

GRANGE was confident that he had presented and founded a true theory — based<br />

67 (Waring, 1770, 162). <strong>The</strong> Latin original has not been available. <strong>The</strong>refore, reference is given to the<br />

English translation (ibid.).<br />

68 “Il semble donc qu’on pourrait conclure de là par induction que toute équation, de quelque degré<br />

qu’elle soit, sera aussi résoluble à l’aide d’une réduite dont les racines soient représentées par la<br />

même formule<br />

x ′ + yx ′′ + y 2 x ′′′ + y 3 x iv + . . . .<br />

Mais, d’après ce que nous avons démontré dans la Section précédente à l’occasion des méthodes de<br />

MM. Euler et Bezout, lesquelles conduisent directement à de pareilles réduites, on a, ce semble, lieu<br />

de se convaincre d’avance que cette conclusion se trouvera en défaut dès le cinquième degré; d’où il<br />

s’ensuit que, si la résolution algébrique des équations des degrés supérieurs au quatrième n’est pas<br />

impossible, elle doit dépendre de quelques fonctions des racines, différentes de la précédente.” (Lagrange,<br />

1770–1771, 356–357).


82 Chapter 5. Towards unsolvable equations<br />

upon combinations, i.e. permutations — inside which the solution could be investi-<br />

gated. However, for equations <strong>of</strong> the fifth and higher degrees the required number <strong>of</strong><br />

calculations and combinations would be exceeding practical possibilities.<br />

“<strong>The</strong>se are, if I am not mistaken, the true principles <strong>of</strong> the solution <strong>of</strong> equations,<br />

and the most appropriate analysis leading to it. As one can see, it all comes<br />

down to a sort <strong>of</strong> calculus <strong>of</strong> combinations, by which one finds à priori the results<br />

for which one should be prepared. It should, by the way, be applicable to<br />

equations <strong>of</strong> the fifth degree and higher degrees, <strong>of</strong> which the solution is until<br />

now unknown. But this application demands a too great number <strong>of</strong> researches<br />

and combinations, <strong>of</strong> which the success is still in serious doubt, for us to follow<br />

this path in the present work. We hope, though, to be able to follow it at another<br />

time, and we content ourselves by having laid the foundations <strong>of</strong> a theory which<br />

appears to us to be new and general.” 69<br />

LAGRANGE never wrote the definitive work which he had reserved the right to<br />

do. By the time GALOIS had substantiated LAGRANGE’S claim for generality and ap-<br />

plicability <strong>of</strong> his theory <strong>of</strong> combinations (see chapter 8.5), LAGRANGE was no longer<br />

around to celebrate the ultimate vindication <strong>of</strong> his research in the field <strong>of</strong> algebraic<br />

solubility.<br />

Both WARING and LAGRANGE believed by 1770 that their theories were the nec-<br />

essary stepping stones towards the study <strong>of</strong> solutions to general equations. However,<br />

they both acknowledged that the amount <strong>of</strong> work required to apply these theories to<br />

the quintic equation was beyond their own limitations. Before the end <strong>of</strong> the century,<br />

even more radical opinions were voiced in print.<br />

5.4.2 Outright impossibility<br />

In the introduction to his first pro<strong>of</strong> (published 1799 but constructed two years earlier)<br />

<strong>of</strong> the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra, GAUSS gave detailed discussions and criticisms<br />

<strong>of</strong> previously attempted pro<strong>of</strong>s. In EULER’S attempt dating back to 1749, GAUSS found<br />

the implicit assumption that any polynomial equation could be solved by radicals.<br />

“In a few words: It is without sufficient reason assumed that the solution <strong>of</strong><br />

any equation can be reduced to the resolution <strong>of</strong> pure equations. Perhaps it would<br />

not be too difficult to prove the impossibility for the fifth degree with all rigor; I<br />

will communicate my investigations on this subject on another occasion. At this<br />

place, it suffices to emphasize that the general solution <strong>of</strong> equations, in this sense,<br />

69 “Voilà, si je me ne trompe, les vrais principes de la résolution des équations et l’analyse la plus<br />

propre à y conduire; tout se réduit, comme on voit, à une espèce de calcul des combinaisons, par<br />

lequel on trouve à priori les résultats auxquels on doit s’attendre. Il serait à propos d’en faire<br />

l’application aux équations du cinquième degré et des degrés supérieurs, dont la résolution est<br />

jusqu’à présent inconnue; mais cette application demande un trop grand nombre de recherches et<br />

de combinaisons, dont le succés est encore d’ailleurs fort douteux, pour que nous puissions quant<br />

à présent nous livrer à ce travail; nous espérons cependant pouvoir y revenir dans un autre temps,<br />

et nous nous contenterons ici d’avoir posé les fondements d’une théorie qui nous paraît nouvelle et<br />

générale.” (Lagrange, 1770–1771, 403).


5.4. Belief in algebraic solubility shaken 83<br />

remains very doubtful, and consequently that any pro<strong>of</strong> whose entire strength<br />

depends on this assumption in the current state <strong>of</strong> affairs has no weight.” 70<br />

In 1799, GAUSS’S aim was to scrutinize EULER’S pro<strong>of</strong> <strong>of</strong> the Fundamental <strong>The</strong>o-<br />

rem <strong>of</strong> Algebra. For this purpose, it was sufficient for GAUSS to express his suspicion<br />

that the algebraic solution <strong>of</strong> general equations was not established with the necessary<br />

rigor. Thus — at least as it stands — GAUSS’S criticism seems to confront the founda-<br />

tions and not the validity <strong>of</strong> this hidden assumption in EULER’S pro<strong>of</strong>. Although it<br />

is doubtful whether or not GAUSS possessed a demonstration that the validity could<br />

also be questioned, he certainly suggested the possibility. Two years later in his influ-<br />

ential Disquisitiones 1801, GAUSS addressed the problem again in connection with the<br />

cyclotomic equations (see quotation below). Possibly alluding to LAGRANGE’S “very<br />

great computational work” GAUSS described the solution <strong>of</strong> higher degree equations<br />

not merely beyond the existing tools <strong>of</strong> analysis but outright impossible.<br />

“<strong>The</strong> preceding discussion had to do with the discovery <strong>of</strong> auxiliary equations.<br />

Now we will explain a very remarkable property concerning their solution. Everyone<br />

knows that the most eminent geometers have been unsuccessful in the search<br />

for a general solution <strong>of</strong> equations higher than the fourth degree, or (to define the<br />

search more accurately) for the reduction <strong>of</strong> mixed equations to pure equations.<br />

And there is little doubt that this problem is not merely beyond the powers <strong>of</strong> contemporary<br />

analysis but proposes the impossible (cf. what we said on this subject<br />

in Demonstrationes nova, art. 9 [above]). Nevertheless it is certain that there are<br />

innumerable mixed equations <strong>of</strong> every degree which admit a reduction to pure<br />

equations, and we trust that geometers will find it gratifying if we show that our<br />

equations are always <strong>of</strong> this kind.” 71<br />

While GAUSS was voicing his opinion on the insolubility <strong>of</strong> higher degree equa-<br />

tions in Latin from his position in Göttingen, the support for the solubility <strong>of</strong> the<br />

quintic was shaken even more radically by an Italian. P. RUFFINI (1765–1822) had<br />

published his first pro<strong>of</strong> <strong>of</strong> the impossibility <strong>of</strong> solving the quintic in 1799, the same<br />

year GAUSS had first uttered his doubts about its possibility. But while GAUSS had<br />

70 “Seu, missis verbis, sine ratione sufficienti supponitur, cuiusvis aequationis solutionem ad solutionem<br />

aequationum purarum reduci posse. Forsan non ita difficile foret, impossibilitatem iam pro<br />

quinto gradu omni rigore demonstrare, de qua re alio loco disquisitiones meas fusius proponam.<br />

Hic sufficit, resolubilitatem generalem aequationum, in illo sensu acceptam, adhuc valde dubiam<br />

esse, adeoque demonstrationem, cuius tota vis ab illa suppositione pendet, in praesenti rei statu<br />

nihil ponderis habere.” (C. F. Gauss, 1799, 17–18); for a translation into German, see (C. F. Gauss,<br />

1890, 20–21).<br />

71 “Disquisitiones praecc. circa inventionem aequationum auxiliarium versabantur: iam de earum solutione<br />

proprietatem magnopere insignem explicabimus. Constat, omnes summorum geometrarum<br />

labores, aequationum ordinem quartum superantium resolutionem generalem, sive (ut accuratius<br />

quid desideretur definiam) affectarum reductionem ad puras, inveniendi semper hactenus irritos<br />

fuisse, et vix dubium manet, quin hocce problema non tam analyseos hodiernae vires superet, quam<br />

potius aliquid impossibile proponat (Cf. quae de hoc argumento annotavimus in Demonstr. nova<br />

etc. arg. 9). Nihilominus certum est, innumeras aequationes affectas cuiusque gradus dari, quae<br />

talem reductionem ad puras admittant, geometrisque gratum fore speramus, si nostras aequationes<br />

auxiliares semper huc referendas esse ostenderimus.” (C. F. Gauss, 1801, 449); English translation<br />

from (C. F. Gauss, 1986, 445). Bold-face has been substituted for the original small-caps.


84 Chapter 5. Towards unsolvable equations<br />

only alluded to a pro<strong>of</strong> without communicating it, RUFFINI had taken the step <strong>of</strong> pub-<br />

lishing his arguments.<br />

5.5 RUFFINI’s pro<strong>of</strong>s <strong>of</strong> the insolubility <strong>of</strong> the quintic<br />

In the 1820s, the search for algebraic solutions to equations <strong>of</strong> higher degree was<br />

proved to be in vain when ABEL demonstrated the algebraic insolubility <strong>of</strong> the quintic.<br />

However, ABEL was not the first to claim the insolubility; more than 25 years before<br />

him, the Italian RUFFINI had published his investigations which led him to the same<br />

conclusion and a pro<strong>of</strong> there<strong>of</strong>. RUFFINI’S works were not widely known, and during<br />

his investigations ABEL was unaware <strong>of</strong> their existence (see section 6.7). ABEL based<br />

his investigations on the analysis by LAGRANGE and works <strong>of</strong> CAUCHY on the theory<br />

<strong>of</strong> permutations. Although not directly inspired by RUFFINI’S works, RUFFINI played<br />

an indirect role in fertilizing the ground for ABEL’S work. <strong>The</strong> indirect influence <strong>of</strong><br />

RUFFINI through the very direct influence <strong>of</strong> CAUCHY on the development leading<br />

to ABEL’S work is two-fold. Firstly, these men smoothed the transition from the be-<br />

liefs described in the previous section to the rigorous knowledge <strong>of</strong> the insolubility <strong>of</strong><br />

the quintic. Secondly, their investigations took the still young theory <strong>of</strong> permutations<br />

to a more advanced level; and in doing so, they provided an important characteriza-<br />

tion <strong>of</strong> the number <strong>of</strong> values a rational function can obtain under permutations <strong>of</strong> its<br />

arguments.<br />

5.5.1 Insolubility proved<br />

Although GAUSS had proclaimed his belief that the insolubility <strong>of</strong> the quintic might<br />

not be difficult to prove with all rigor, the Italian RUFFINI, in 1799, was the first math-<br />

ematician to state the insolubility as a result and provide the claim with a pro<strong>of</strong>. RUF-<br />

FINI’S style <strong>of</strong> presentation was long, cumbersome, and at times not free <strong>of</strong> errors; and<br />

his initial pro<strong>of</strong> was met with immediate criticism for these reasons. But convinced <strong>of</strong><br />

the result and his pro<strong>of</strong>, RUFFINI kept elaborating and clarifying his theory in print for<br />

the next 20 years, producing a total <strong>of</strong> five different versions <strong>of</strong> the pro<strong>of</strong>. <strong>The</strong> pro<strong>of</strong>s<br />

were published in Italian as monographs in Bologna and in the mathematical mem-<br />

oirs <strong>of</strong> the Società Italiana delle Scienze, Modena. Although published and distributed,<br />

the impact <strong>of</strong> RUFFINI’S work was limited; among the few non-Italians to take a view-<br />

point on RUFFINI’S work was CAUCHY (see section 5.5.3). In 1810, J.-B. J. DELAMBRE<br />

(1749–1822) was aware <strong>of</strong> RUFFINI’S 1802-pro<strong>of</strong> which he described as “difficult” and<br />

“not suited for inclusion in works meant as a first introduction” to the subject. 72 I shall<br />

mainly deal with RUFFINI’S initial pro<strong>of</strong> given in his textbook, 73 which he elaborated<br />

72 (Delambre, 1810, 86–87).<br />

73 (Ruffini, 1799).


5.5. RUFFINI’s pro<strong>of</strong>s <strong>of</strong> the insolubility <strong>of</strong> the quintic 85<br />

Figure 5.4: PAOLO RUFFINI (1765–1822)<br />

on numerous occasions, and his final pro<strong>of</strong> published 1813. 74<br />

5.5.2 RUFFINI’s first pro<strong>of</strong><br />

<strong>The</strong> writings <strong>of</strong> RUFFINI were deeply inspired by LAGRANGE’S analysis <strong>of</strong> the solu-<br />

bility <strong>of</strong> equations described in section 5.2. 75 LAGRANGE’S ideas, concepts, and nota-<br />

tion permeate RUFFINI’S works; and on numerous occasions RUFFINI openly acknowl-<br />

edged his debt to LAGRANGE. 76 As LAGRANGE had done, RUFFINI studied equations<br />

<strong>of</strong> low degrees in order to establish patterns subjectable <strong>of</strong> generalization. Prior to<br />

applying his analysis to the fifth degree equation, RUFFINI propounded the corner<br />

stone <strong>of</strong> his investigation. Central to his line <strong>of</strong> argument was his classification <strong>of</strong> per-<br />

mutations. Founded in LAGRANGE’S studies <strong>of</strong> the behavior <strong>of</strong> functions when their<br />

arguments were permuted, RUFFINI set out to classify all such permutations <strong>of</strong> argu-<br />

ments which left the function (formally) unaltered. RUFFINI’S concept <strong>of</strong> permutation<br />

(Italian: “permutazione”) differed from the modern one, and can most easily be un-<br />

derstood if translated into the modern concept introduced by CAUCHY in the 1840s <strong>of</strong><br />

74 (Ruffini, 1813). <strong>The</strong> presentation <strong>of</strong> RUFFINI’S pro<strong>of</strong>s will largely rely on secondary sources, primarily<br />

(Burkhardt, 1892), (Wussing, 1969, 56–59), and (Kiernan, 1971, 56–60). In (R. G. Ayoub, 1980), his<br />

pro<strong>of</strong>s are interpreted using concepts from GALOIS theory.<br />

75 (Lagrange, 1770–1771).<br />

76 See for instance his preliminary discourse in (Ruffini, 1799, 3–4).


86 Chapter 5. Towards unsolvable equations<br />

simple permutations<br />

composite permutations<br />

�<br />

⎧<br />

⎨<br />

⎩<br />

powers <strong>of</strong> a cycle<br />

powers <strong>of</strong> a non-cycle<br />

intransitive ones<br />

transitive, imprimitive ones<br />

transitive, primitive ones<br />

Table 5.1: RUFFINI’s classification <strong>of</strong> permutations<br />

systems <strong>of</strong> conjugate substitutions. 77 Thus, a permutation for RUFFINI corresponded to a<br />

collection <strong>of</strong> interchangements (i.e. transitions from one arrangement <strong>of</strong> symbols e.g.<br />

123 . . . n to another e.g. 213 . . . n) which left the given function formally unaltered. 78<br />

Classification <strong>of</strong> permutations. RUFFINI divided his permutations into simple ones<br />

which were generated by iterations (i.e. powers) <strong>of</strong> a single interchangement 79 and<br />

composite ones generated by more than one interchangement. His simple permutations<br />

consisting <strong>of</strong> powers <strong>of</strong> a single interchangement were subdivided into two types dis-<br />

tinguishing the case in which the single interchangement consisted <strong>of</strong> a single cycle<br />

from the case in which it was the product <strong>of</strong> more than one cycle.<br />

RUFFINI’S composite permutations were subsequently subdivided into three types. 80<br />

A permutation (i.e. set <strong>of</strong> interchangements) was said to be <strong>of</strong> the first type if two<br />

arrangements existed which were not related by an interchangement from the permu-<br />

tation. 81 Translated into the modern terminology <strong>of</strong> permutation groups, this type<br />

corresponds to intransitive groups. RUFFINI defined the second type to contain all per-<br />

mutations which did not belong to the first type and for which there existed some<br />

non-trivial subset <strong>of</strong> roots S such that, in modern notation, σ (S) = S or σ (S) ∩ S = ∅<br />

for any interchangement σ belonging to the permutation. Such transitive groups were<br />

later termed imprimitive. <strong>The</strong> last type consisted <strong>of</strong> any permutation not belonging to<br />

any <strong>of</strong> the previous types, and thus corresponds to primitive groups.<br />

Building on this classification <strong>of</strong> all permutations into the five types (table 5.1),<br />

RUFFINI introduced his other key concept <strong>of</strong> degree <strong>of</strong> equivalence (Italian: “grado di<br />

uguaglianza”) <strong>of</strong> a given function f <strong>of</strong> the n roots <strong>of</strong> an equation as the number <strong>of</strong> dif-<br />

ferent permutations not altering the formal value <strong>of</strong> f . Denoting the degree <strong>of</strong> equiva-<br />

lence by p, RUFFINI stated the result <strong>of</strong> LAGRANGE (see section 5.2) that p must divide<br />

n!.<br />

77 (Burkhardt, 1892, 133).<br />

78 In modern notation: With f the given function <strong>of</strong> n quantities, a permutazione to RUFFINI was a set<br />

G ⊆ Σn such that f ◦ σ = f for all σ ∈ G.<br />

79 RUFFINI’S simple permutations correspond to the modern concept <strong>of</strong> cyclic permutation groups.<br />

80 (Ruffini, 1799, 163).<br />

81 I.e. there exists two arrangements a and b such that σ (a) �= b for all σ in the set <strong>of</strong> interchangements.


5.5. RUFFINI’s pro<strong>of</strong>s <strong>of</strong> the insolubility <strong>of</strong> the quintic 87<br />

Possible numbers <strong>of</strong> values. RUFFINI at this point turned towards the fifth degree<br />

equation. By an extensive and laborious study, helped by his classification, RUFFINI<br />

was able to establish that if n = 5 the degree <strong>of</strong> equivalence p could not assume any<br />

<strong>of</strong> the values<br />

15, 30, or 40.<br />

Since the number <strong>of</strong> different values <strong>of</strong> the function f could be obtained by dividing n!<br />

by p, he had therefore demonstrated that no function f <strong>of</strong> the five roots <strong>of</strong> the quintic<br />

could exist which assumed<br />

5!<br />

15<br />

= 8, 5!<br />

30<br />

= 4, or 5!<br />

40<br />

different values under permutations <strong>of</strong> the five roots.<br />

Although still embedded in the Lagrangian approach to permutations, RUFFINI’S<br />

main result can be viewed as a determination <strong>of</strong> the index (corresponding to his degree<br />

<strong>of</strong> equivalence, p) <strong>of</strong> all subgroups in Σ5.<br />

Degrees <strong>of</strong> radical extractions. In order to prove the impossibility <strong>of</strong> solving the<br />

quintic algebraically, RUFFINI assumed without pro<strong>of</strong> that any radical occurring in<br />

a supposed solution would be rationally expressible in the roots <strong>of</strong> the equation. He<br />

never verified this hypothesis, which ABEL later independently formulated and proved.<br />

Based on the assumption and the result that no function <strong>of</strong> the roots x1, . . . , x5 could<br />

have 3, 4, or 8 values, RUFFINI could prove the insolubility by a nice and short argu-<br />

ment which ran as follows.<br />

He first considered a situation in which among two functions Z and M <strong>of</strong> x1, . . . , x5<br />

there existed a relationship <strong>of</strong> the form<br />

Z 5 − M = 0<br />

corresponding to the extraction <strong>of</strong> a fifth root <strong>of</strong> a rational function. <strong>The</strong> situation was<br />

drawn from the study <strong>of</strong> a possible solution to the quintic equation where it corre-<br />

sponded to the inner-most root extraction being a fifth root. By implicitly assuming<br />

that Z was altered by some interchangement Q which left M unaltered, RUFFINI first<br />

observed that Q would have to be a 5-cycle. If Z was unaltered by a non-identity in-<br />

terchangement P, it would also be unaltered by Q −1 PQ which belonged to the same<br />

permutation. By reference to a result, which he had previously established by exam-<br />

ining each <strong>of</strong> the different types <strong>of</strong> permutations, RUFFINI found (art. 273) that Q<br />

under these conditions would belong to the same permutation as Q −1 PQ and there-<br />

fore could not alter Z, contradicting the assumptions made about Q. Thus, no such<br />

non-identity interchangement P could exist, and the 120 values <strong>of</strong> Z corresponding to<br />

different arrangements <strong>of</strong> x1, . . . , x5 were necessarily distinct. Consequently, the first<br />

radical to be extracted could not be a fifth root, and since no function <strong>of</strong> the five roots<br />

having three or four values existed, it could not be a third or a fourth root, neither.<br />

<strong>The</strong>refore, it had to be a square root.<br />

= 3


88 Chapter 5. Towards unsolvable equations<br />

At this point, RUFFINI focused on the second radical to be extracted and the above<br />

argument applied equally well to rule out the case <strong>of</strong> a fifth root. Similarly, it could<br />

not be a square root or a fourth root since these would lead to functions having four<br />

(2 × 2) or eight (2 × 4) values, which were proved to be non-existent. RUFFINI had<br />

thus established that any supposed solution to the quintic equation would have to<br />

begin with the extraction <strong>of</strong> a square root followed by the extraction <strong>of</strong> a third root.<br />

However, as he laboriously proved by considering each case in turn, the six-valued<br />

function obtained by these two radical extractions did not become three-valued after<br />

the initial square root had been adjoined.<br />

<strong>The</strong> pro<strong>of</strong> which RUFFINI gave for the insolubility <strong>of</strong> the quintic was thus based on<br />

three central parts:<br />

1. <strong>The</strong> classification <strong>of</strong> permutations into types (table 5.1)<br />

2. A demonstration, based on (1), that no function <strong>of</strong> the five roots <strong>of</strong> the general<br />

quintic could have 3, 4, or 8 values under permutations <strong>of</strong> the roots.<br />

3. A study <strong>of</strong> the two inner-most (first) radical extractions <strong>of</strong> a supposed solution<br />

to the quintic, in which the result <strong>of</strong> (2) was used to reach a contraction.<br />

<strong>The</strong> mere extent <strong>of</strong> the classification and the caution necessary to include all cases 82<br />

combined with RUFFINI’S intellectual debt to LAGRANGE may serve to view RUFFINI’S<br />

work as filling in some <strong>of</strong> the “infinite labor” described by WARING and LAGRANGE<br />

in expressing their doubts about the solubility <strong>of</strong> higher degree equations (see section<br />

5.4.1 above). However, RUFFINI’S investigations led to the complete reverse result:<br />

that the solution <strong>of</strong> the quintic was impossible.<br />

One <strong>of</strong> RUFFINI’S friends and critical readers, P. ABBATI (1768–1842), gave sev-<br />

eral improvements <strong>of</strong> RUFFINI’S initial pro<strong>of</strong>. <strong>The</strong> most important one was that he<br />

replaced the laborious arguments based on thorough consideration <strong>of</strong> particular cases<br />

by arguments <strong>of</strong> a more general character. 83 <strong>The</strong>se more general arguments greatly<br />

simplified RUFFINI’S pro<strong>of</strong>s that no function <strong>of</strong> the five roots <strong>of</strong> the quintic could have<br />

3, 4, or 8 different values. ABBATI was convinced <strong>of</strong> the validity <strong>of</strong> RUFFINI’S result<br />

but wanted to simplify its pro<strong>of</strong>, and RUFFINI incorporated his improvements into<br />

subsequent pro<strong>of</strong>s, from 1802 and henceforth.<br />

Others, however, were not so convinced <strong>of</strong> the general validity <strong>of</strong> RUFFINI’S re-<br />

sults. Mathematicians belonging to the “old generation” were somewhat stunned<br />

by the non-constructive nature <strong>of</strong> the pro<strong>of</strong>s, which they described as “vagueness”.<br />

For instance, the mathematician G. F. MALFATTI (1731–1807) severely criticized RUF-<br />

FINI’S result since it contradicted a general solution which he, himself, previously had<br />

82 According to (Burkhardt, 1892, 135), RUFFINI actually missed the subgroup generated by the cycles<br />

(12345) and (132).<br />

83 (ibid., 140).


5.5. RUFFINI’s pro<strong>of</strong>s <strong>of</strong> the insolubility <strong>of</strong> the quintic 89<br />

given. 84 RUFFINI responded with another publication <strong>of</strong> a version <strong>of</strong> his pro<strong>of</strong> answer-<br />

ing to MALFATTI’S criticism; but before the discussion advanced further, MALFATTI<br />

died.<br />

5.5.3 RUFFINI’s final pro<strong>of</strong><br />

In his fifth, and final, publication <strong>of</strong> his insolubility theorem 1813, RUFFINI recapitu-<br />

lated important parts <strong>of</strong> LAGRANGE’S theory, in which he emphasized the distinction<br />

between numerical and formal equality, before giving the refined version <strong>of</strong> his pro<strong>of</strong>.<br />

According to (Burkhardt, 1892, 155–156), the pro<strong>of</strong> can be dissected into the following<br />

parts comparable to the parts <strong>of</strong> the 1799 pro<strong>of</strong> (see point 3 above):<br />

1. If two functions y and P <strong>of</strong> the roots x1, . . . , x5 <strong>of</strong> the quintic are related by<br />

y p − P = 0<br />

(for any p) and P remains unaltered by the cyclic permutation (12345), there<br />

must exist a value y1 <strong>of</strong> y which in turn changes into y2, y3, y4, and y5. Conse-<br />

quently,<br />

where β is a fifth root <strong>of</strong> unity.<br />

y k = β k y1<br />

2. If P is furthermore unaltered by the cyclic permutation (123), then y1 must<br />

change into γy1 where γ is a third root <strong>of</strong> unity.<br />

3. <strong>The</strong> permutation (13452) is comprised <strong>of</strong> the two cycles (12345) (123) and y must<br />

remain unaltered. <strong>The</strong>refore, β 5 γ 5 = 1 which in turn implies that γ = 1, demon-<br />

strating that y cannot be altered by any <strong>of</strong> the permutations (123), (234), (345),<br />

(451), or (512). By combining these 3-cycles the 5-cycle (12345) can be obtained,<br />

and thus y cannot be altered by the 5-cycle, neither.<br />

4. Consequently, it is impossible by sequential root extractions to describe func-<br />

tions which have more than two values, and the insolubility is demonstrated.<br />

5.5.4 Reactions to RUFFINI’s pro<strong>of</strong>s<br />

In a paper published 1845, 85 P. L. WANTZEL (1814–1848) gave a fusion argument in-<br />

corporating the permutation theoretic arguments <strong>of</strong> RUFFINI’S final pro<strong>of</strong> into the setting<br />

<strong>of</strong> ABEL’S pro<strong>of</strong>. 86<br />

84 (Malfatti, 1804).<br />

85 (Wantzel, 1845).<br />

86 See also (Burkhardt, 1892, 156).


90 Chapter 5. Towards unsolvable equations<br />

RUFFINI corresponded with CAUCHY, who in 1816 was a promising young Parisian<br />

ingenieur. 87 CAUCHY praised RUFFINI’S research on the number <strong>of</strong> values which a<br />

function could acquire when its arguments were permuted, a topic CAUCHY, himself,<br />

had investigated in an treatise published the year before 1815 with due reference to<br />

RUFFINI (see below). Following this exchange <strong>of</strong> letters CAUCHY wrote RUFFINI an-<br />

other letter in September 1821, in which he acknowledged RUFFINI’S progress in the<br />

important field <strong>of</strong> solubility <strong>of</strong> algebraic equations:<br />

“I must admit that I am anxious to justify myself in your eyes on a point which<br />

can easily be clarified. Your memoir on the general solution <strong>of</strong> equations is a work<br />

which has always appeared to me to deserve to keep the attention <strong>of</strong> geometers.<br />

In my opinion, it completely demonstrates the algebraic insolubility <strong>of</strong> the general<br />

equations <strong>of</strong> degrees above the fourth. <strong>The</strong> reason that I had not lectured on it<br />

[the insolubility ] in my course in analysis, and it must be said that these courses<br />

are meant for students at the École Royale Polytechnique, is that I would have<br />

deviated too much from the topics set forth in the curriculum <strong>of</strong> the École.” 88<br />

At least by 1821, the validity <strong>of</strong> RUFFINI’S claim that the general quintic could<br />

not be solved by radicals was propounded, not only by a somewhat obscure Italian<br />

mathematician and the allusions <strong>of</strong> GAUSS, but also one <strong>of</strong> the most promising and<br />

ambitious French mathematicians <strong>of</strong> the early nineteenth century. However, it should<br />

take further publications, notably by the young ABEL, before this validity would be<br />

accepted by the broad international community <strong>of</strong> mathematicians.<br />

5.6 CAUCHY’ theory <strong>of</strong> permutations and a new pro<strong>of</strong> <strong>of</strong><br />

RUFFINI’s theorem<br />

In November <strong>of</strong> 1812, CAUCHY handed in a memoir on symmetric functions to the In-<br />

stitut de France which was published three years later as two separate papers in the<br />

Journal d’École Polytechnique. 89 <strong>The</strong> first <strong>of</strong> the two papers is <strong>of</strong> special interest in the<br />

history <strong>of</strong> solubility <strong>of</strong> polynomial equations. It bears the long but precise title Mé-<br />

moire sur le nombre des valuers qu’une fonction peut acquérir, lorsqu’on y permute de toutes<br />

manières possibles les quantités qu’elle renferme. 90 Although CAUCHY’S issue was not<br />

the solubility-question, his paper was to become extremely important for subsequent<br />

research. It was primarily concerned with a more general version <strong>of</strong> RUFFINI’S result<br />

87 (Ruffini, 1915–1954, vol. 3, 82–83).<br />

88 “Je suis impatient, je l’avous, de me justifier à Vos yeux sur un point qui peut être facilement éclairi.<br />

Votre mémoire sur la résolution générale des équations est un travail qui m’a toujours paru digne de<br />

fixer l’attention des géomètres, et qui, à mon avis, démontre complètement l’insolubilité algébrique<br />

des équations générales d’un dégré supérieur au quatrième. Si je n’en ai pas parlé dans mon cours<br />

d’analyse, c’est que, ce cours étant destiné aux élèves d’École Royale Polytechnique, je ne devois<br />

pas trop m’écarter des matières indiquées dans les programmes de l’école.” (ibid., vol. 3, 88–89).<br />

89 (A.-L. Cauchy, 1815a; A.-L. Cauchy, 1815b).<br />

90 (A.-L. Cauchy, 1815a).


5.6. CAUCHY’ theory <strong>of</strong> permutations and a new pro<strong>of</strong> <strong>of</strong> RUFFINI’s theorem 91<br />

Figure 5.5: AUGUSTIN-LOUIS CAUCHY (1789–1857)<br />

that no function <strong>of</strong> five quantities could have three or four different values when its ar-<br />

guments were permuted (see above). Before going into this particular result, however,<br />

CAUCHY devised the terminology and notation which he was going to use. Precisely<br />

in formulating exact and useful notation and terminology, CAUCHY advanced well<br />

beyond his predecessors and laid the foundations upon which the nineteenth-century<br />

theory <strong>of</strong> permutations would later build.<br />

Notational advances. With CAUCHY, the term “permutation” came to mean an ar-<br />

rangement <strong>of</strong> indices, thereby replacing the “arrangements” <strong>of</strong> which RUFFINI spoke.<br />

A “substitution” was subsequently defined to be a transition from one permutation<br />

to another (which is the modern meaning <strong>of</strong> “permutation”), and CAUCHY devised<br />

writing it as, for instance,<br />

� �<br />

1.2.4.3<br />

. (5.12)<br />

2.4.3.1<br />

CAUCHY’S convention was that in the expression K, to which the substitution (5.12)<br />

was to be applied, the index 2 was to replace the index 1, the index 4 to replace 2, 3<br />

should replace 4, and 1 should replace 3. More generally, CAUCHY wrote<br />

� A1<br />

A2<br />


92 Chapter 5. Towards unsolvable equations<br />

for the substitution which transformed the permutation A1 into A2 in the above-<br />

mentioned way. 91 He then defined ( A1) to be the product <strong>of</strong> two substitutions (A2<br />

A6 A3 )<br />

and ( A4 A5 ) if it gave the same result as the two applied sequentially,92 in which case<br />

CAUCHY wrote � A1<br />

A6<br />

�<br />

=<br />

� ��<br />

A2 A4<br />

Furthermore, he defined the identical substitution and powers <strong>of</strong> a substitution to have<br />

the meanings we still attribute to these concepts today. 93 <strong>The</strong> smallest integer n such<br />

that the n th power <strong>of</strong> a substitution was the identity substitution, CAUCHY called the<br />

degree <strong>of</strong> the substitution. 94 All these notational advances played a central part in<br />

formalizing the manipulations on permutations and were soon generally adopted.<br />

LAGRANGE’S <strong>The</strong>orem. In order to demonstrate LAGRANGE’S theorem, CAUCHY<br />

let K denote an arbitrary expression in n quantities,<br />

A3<br />

A5<br />

K = K (x1, . . . , xn) .<br />

With N = n!, he labelled the N different permutations <strong>of</strong> these n quantities<br />

A1, . . . , AN.<br />

<strong>The</strong> values which K would acquire when the corresponding substitutions <strong>of</strong> the form<br />

( A1 Au ) were applied were correspondingly labelled K1, . . . , KN,<br />

� �<br />

A1<br />

Ku = K for 1 ≤ u ≤ N.<br />

Au<br />

If these were all distinct, the expression K would obviously have N different values<br />

when its arguments were interchanged. In the contrary case, CAUCHY assumed that<br />

for M indices the values <strong>of</strong> K were equal<br />

�<br />

.<br />

Kα = K β = Kγ = . . . .<br />

<strong>The</strong> core <strong>of</strong> the pro<strong>of</strong> was CAUCHY’S realization that if the permutation A λ was fixed<br />

and the substitution ( Aα<br />

A β ) was applied to A λ giving Aµ, i.e.<br />

�<br />

Aα<br />

Aµ =<br />

A β<br />

�<br />

A λ,<br />

the corresponding values K λ and Kµ would be identical. Consequently, the different<br />

values <strong>of</strong> K came in bundles <strong>of</strong> M and CAUCHY had deduced that M had to divide<br />

n!. <strong>The</strong> central concept <strong>of</strong> degree <strong>of</strong> equivalence, which RUFFINI had introduced to mean<br />

91 (A.-L. Cauchy, 1815a, 67).<br />

92 (ibid., 73).<br />

93 (ibid., 73, 74)<br />

94 (ibid., 76). ABEL was later to change this term to the now standard order.


5.6. CAUCHY’ theory <strong>of</strong> permutations and a new pro<strong>of</strong> <strong>of</strong> RUFFINI’s theorem 93<br />

the number <strong>of</strong> substitutions which left the given function unaltered, was renamed the<br />

indicative divisor (French: “diviseur indicatif”) by CAUCHY and was exactly what he<br />

had denoted by M. Terming the number <strong>of</strong> different values <strong>of</strong> K under all possible<br />

substitutions the index <strong>of</strong> the function K and denoting it by R, CAUCHY had obtained<br />

the formula<br />

n! = R × M. (5.13)<br />

<strong>The</strong> RUFFINI-CAUCHY <strong>The</strong>orem. After explicitly providing the function <strong>of</strong> n quan-<br />

tities a1, . . . , an given by<br />

� �<br />

∏ ai − aj 1≤i


94 Chapter 5. Towards unsolvable equations<br />

Each circle <strong>of</strong> permutations is represented by a row in the following table:<br />

A2 = ( As<br />

At )A1, . . . , Am = ( As<br />

At )Am−1, A1 = ( As<br />

At )Am,<br />

Am+2 = ( As<br />

At )Am+1, . . . , A2m = ( As<br />

At )A2m−1, Am+1 = ( As<br />

At )A2m,<br />

. . . . . . . . . . . .<br />

AN−m+2 = . . . , . . . , AN = ( As<br />

At )AN−1, AN−m+1 = ( As<br />

At )AN−m.<br />

<strong>The</strong>se permutations can be reordered when (5.14) is taken into account:<br />

A1, A2 = ( As<br />

At )A1, . . . , Am = ( As<br />

At )m−1 A1,<br />

Am+1, Am+2 = ( As<br />

At )Am+1, . . . , A2m = ( As<br />

At )m−1 Am+1,<br />

. . . . . . . . . . . .<br />

AN−m+1, AN−m+2 = ( As<br />

At )AN−m+1, . . . , AN = ( As<br />

At )m−1 AN−m+1.<br />

<strong>The</strong> notation ( As<br />

At )A1 indicates that the substitution ( As)<br />

be applied to the permutation<br />

At<br />

A1.<br />

Table 5.2: <strong>The</strong> N m<br />

circles formed by applying (As<br />

At ) to A1, . . . , AN.<br />

power <strong>of</strong> ( As<br />

At ) to Ax. Consequently, the substitution ( Ax)<br />

was equal to a power <strong>of</strong> (As<br />

Ay At ).<br />

If m were a prime, the converse would also be true, since if<br />

� �k � �<br />

As Ax<br />

=<br />

and (k, m) = 1, there existed α, β such that αk + βm = 1, i.e.<br />

� � � �αk+βm � �αk As As<br />

As<br />

=<br />

= =<br />

At<br />

At<br />

At<br />

Ay<br />

At<br />

� Ax<br />

<strong>The</strong> details <strong>of</strong> this argument were left out by CAUCHY, but were later provided by<br />

ABEL. 98 Since Ax and Ay corresponded to the same value <strong>of</strong> K, the function would<br />

not change if the substitution ( Ax)<br />

were applied. Consequently, K would also remain<br />

Ay<br />

unaltered when the substitution ( As)<br />

was applied, and the number <strong>of</strong> different values<br />

At<br />

<strong>of</strong> K, which CAUCHY had denoted M, could not be greater than N m , whereby he had<br />

reached a contradiction. Setting m = p, CAUCHY had obtained the desired result.<br />

In the second part, CAUCHY demonstrated by decomposing p-cycles into 3-cycles<br />

that if the value <strong>of</strong> K remained unaltered by all substitutions <strong>of</strong> degree p it would also<br />

be unaltered by any circular substitution <strong>of</strong> order 3. <strong>The</strong> important step was obtained<br />

Ay<br />

� α<br />

.<br />

by realizing that the product <strong>of</strong> the two circular substitutions <strong>of</strong> order p<br />

� � � �<br />

αβγδ . . . ζη βγδε . . . ηα<br />

and<br />

βγδε . . . ηα γαβδ . . . ζη<br />

98 (N. H. <strong>Abel</strong>, 1826a).<br />

(5.15)


5.7. Some algebraic tools used by GAUSS 95<br />

was the 3-cycle<br />

� �<br />

αβγ<br />

. (5.16)<br />

γαβ<br />

Thus, given any 3-cycle (5.16), the two p-cycles (5.15) could be formed. Under the hy-<br />

pothesis, these p-cycles left K unaltered, whereby the same was true <strong>of</strong> their product,<br />

i.e. the 3-cycle (5.16).<br />

In the third and final part <strong>of</strong> the pro<strong>of</strong> CAUCHY established that if the value <strong>of</strong><br />

K was unaltered by all 3-cycles, the function K would either be symmetric or have<br />

two different values. In his pro<strong>of</strong>, analogous to the second part described above, he<br />

decomposed the 3-cycle<br />

� �<br />

αβγ<br />

γαβ<br />

into the product <strong>of</strong> the two transpositions<br />

� �� �<br />

αβ βγ<br />

βα γβ<br />

which he wrote as (αβ) (βγ). This step <strong>of</strong> the pro<strong>of</strong> corresponds to proving that the<br />

alternating group An is generated by all 3-cycles.<br />

In the remaining part <strong>of</strong> the paper, CAUCHY demonstrated for functions <strong>of</strong> six<br />

arguments, if R < 5 the function would necessarily be symmetric or have two values.<br />

Generally, CAUCHY noted, for n > 4 no functions <strong>of</strong> n quantities were known which<br />

had less than n values without this number being either 1 or 2. After these two early<br />

papers on the theory <strong>of</strong> permutations, CAUCHY would let the topic rest for 30 years<br />

being preoccupied with his many other research themes and his teaching. When he<br />

finally returned to the theory <strong>of</strong> permutations in the 1840s, CAUCHY demonstrated the<br />

following generalization <strong>of</strong> his 1815 result: That no function <strong>of</strong> n quantities could take<br />

on less than n values without either being symmetric or taking on exactly 2 values. 99<br />

With his paper, 100 CAUCHY founded the theory <strong>of</strong> permutations by providing it<br />

with its principal objects: the permutations. He introduced terms and notation which<br />

enabled him to grasp the substitutions as objects abstracted from their action on the<br />

formal values <strong>of</strong> a function, and he provided an important theorem in this new theory<br />

which he based on an elegant, non-computational pro<strong>of</strong>.<br />

5.7 Some algebraic tools used by GAUSS<br />

GAUSS’ first pro<strong>of</strong> <strong>of</strong> the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra had, in a central way, de-<br />

pended on geometrical (topological) intuitions. In 1815, GAUSS published a second<br />

pro<strong>of</strong> <strong>of</strong> the theorem, 101 this time applying algebraic methods. In the process, GAUSS<br />

99 (Dahan, 1980, 281–282).<br />

100 (A.-L. Cauchy, 1815a).<br />

101 (C. F. Gauss, 1815). Eventually, GAUSS would publish two further pro<strong>of</strong>s (one in 1816 and one in<br />

1849) bringing his total to four.


96 Chapter 5. Towards unsolvable equations<br />

spelled out some <strong>of</strong> the most important algebraic tools <strong>of</strong> the early 19 th century; there-<br />

fore some <strong>of</strong> his tools are briefly sketched in the present context. During the pro<strong>of</strong>,<br />

GAUSS dealt with results such as the Euclidean algorithm applied to polynomials and<br />

the “elementariness” <strong>of</strong> the elementary symmetric functions, both <strong>of</strong> which will be-<br />

come immensely important in ABEL’S theory <strong>of</strong> algebraic solubility as described in<br />

subsequent chapters. Whether ABEL studied any <strong>of</strong> GAUSS’ pro<strong>of</strong>s <strong>of</strong> the fundamen-<br />

tal theorem <strong>of</strong> algebra is not clear; there are no explicit references to these pro<strong>of</strong>s in<br />

ABEL’S writings, nor is ABEL anywhere concerned with the existence <strong>of</strong> roots. 102 Thus,<br />

the similarity <strong>of</strong> methods in GAUSS’ pro<strong>of</strong> and ABEL’S subsequent algebraic research<br />

may equally well be attributed to their belonging to the same common framework and<br />

mathematical tradition.<br />

Explicitly stressing the connection to the procedure used to determine the greatest<br />

common divisor <strong>of</strong> integers, GAUSS applied the Euclidean algorithm to polynomials.<br />

Besides producing the greatest common divisor, the procedure also proved that two<br />

polynomials Y, Y ′ have no (non-trivial) common divisor if and only if there exists an-<br />

other pair <strong>of</strong> polynomials Z, Z ′ such that<br />

ZY + Z ′ Y ′ = 1.<br />

<strong>The</strong> second tool which GAUSS introduced concerned symmetric functions, and<br />

amounts to the central theorem on symmetric functions. By firstly decomposing any<br />

symmetric function <strong>of</strong> a, b, c, . . . in a sum <strong>of</strong> terms<br />

Ma α b β c γ . . .<br />

and secondly imposing an ordering on such terms, GAUSS was able to prove that any<br />

symmetric function could be realized as an entire function <strong>of</strong> the elementary symmet-<br />

ric functions.<br />

Besides these tools, GAUSS’ argument rested upon central properties <strong>of</strong> the quan-<br />

tity which he termed the determinant (today called the discriminant) <strong>of</strong> Y (x) = ∏ (x − x k),<br />

∏ i�=j<br />

� �<br />

xi − xj .<br />

GAUSS was able to demonstrate that the determinant vanishes if and only if Y and<br />

d<br />

dx Y have a common divisor, i.e. a common root.<br />

102 Without references, KLINE writes as if ABEL had given a pro<strong>of</strong> <strong>of</strong> the fundamental theorem <strong>of</strong> algebra<br />

(Kline, 1990, 599). I have not been able to identify such a pro<strong>of</strong>, nor have I any idea how KLINE<br />

had come to believe that ABEL had even worked on it.


Chapter 6<br />

ABEL on the algebraic insolubility <strong>of</strong><br />

the quintic: limiting the class <strong>of</strong><br />

solvable equations<br />

In spite <strong>of</strong> the efforts <strong>of</strong> P. RUFFINI (1765–1822) and C. F. GAUSS (1777–1855), the<br />

search for an algebraic solution <strong>of</strong> the quintic remained an attractive problem to a<br />

generation <strong>of</strong> young and aspiring mathematicians. In Norway, N. H. ABEL (1802–<br />

1829) thought he had solved it, but soon realized that he had been misled. In Germany,<br />

C. G. J. JACOBI (1804–1851) worked on the problem, 1 and in France E. GALOIS (1811–<br />

1832), too, thought he had found a solution, only to be disappointed. 2 All <strong>of</strong> them<br />

attacked the problem while they still attended pre-university education. <strong>The</strong> easy<br />

formulation and yet century-long history <strong>of</strong> the problem, and a general belief that its<br />

solution should be possible and not too difficult, made it appear as a good opening<br />

into doing creative mathematics.<br />

Inspired by the stimulation <strong>of</strong> his new, and young, mathematics teacher B. M.<br />

HOLMBOE (1795–1850), ABEL studied the masters and began to engage in creative<br />

mathematics <strong>of</strong> his own. In 1821, he thought he had produced a solution to the gen-<br />

eral fifth degree equation. In the incipient intellectual atmosphere <strong>of</strong> Christiania, few<br />

authorities capable <strong>of</strong> determining the validity <strong>of</strong> ABEL’S reasoning could be found.<br />

But more importantly, the scientific milieu <strong>of</strong> Norway was still without a means <strong>of</strong><br />

publication <strong>of</strong> technical mathematical results deserving international recognition. For<br />

these reasons, pr<strong>of</strong>essor C. HANSTEEN (1784–1873) sent ABEL’S manuscript to pr<strong>of</strong>es-<br />

sor C. F. DEGEN (1766–1825) in Copenhagen for evaluation and possibly publication<br />

in the transactions <strong>of</strong> the Royal Danish Academy <strong>of</strong> Sciences and Letters. <strong>The</strong> accompa-<br />

nying letter which HANSTEEN must have written and the paper, itself, are no longer<br />

preserved. Our only primary source <strong>of</strong> information is the letter which DEGEN wrote<br />

back to HANSTEEN, in which he asked for an elaborated version <strong>of</strong> the argument and<br />

1 (G. L. Dirichlet, 1852, 4).<br />

2 (Toti Rigatelli, 1996, 33).<br />

97


98 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

an application to a specific numerical example.<br />

“As for the talented Mr. <strong>Abel</strong>, I will be happy to present his treatise to the<br />

Royal Academy <strong>of</strong> Science. It shows, even if the goal has not been reached, an extraordinary<br />

head and extraordinary insights, especially for someone his age. Nevertheless,<br />

I excuse myself to require the condition that Mr. A. sends an elaborated<br />

deduction <strong>of</strong> his result together with a numerical example, taken from, for instance,<br />

an equation such as x 5 − 2x 4 + 3x 2 − 4x + 5 = 0. I believe that this will be a rather<br />

necessary lapis lydius [Lydian stone] for him, as I recall what happened to Meier<br />

Hirsche 3 and his ενρηκα [Eureka]; item [furthermore] I would, since the latter part<br />

<strong>of</strong> the communicated manuscript would not be easily readable to the majority <strong>of</strong><br />

the members <strong>of</strong> the Academy, ask for another copy <strong>of</strong> it.” 4<br />

We have no indication that ABEL ever produced an elaborated deduction; appar-<br />

ently the numerical examples worked their part — as the probes <strong>of</strong> truth — as DEGEN<br />

had suggested and led ABEL to a radically new insight. In 1824, he published, at his<br />

own expense, a short work in French entitled Mémoire sur les équations algébriques ou<br />

l’on démontre l’impossibilité de la résolution de l’équation générale du cinquième degré. 5 It<br />

demonstrated the impossibility <strong>of</strong> solving the general equation <strong>of</strong> the fifth degree by<br />

algebraic means — ABEL had left the last essential requirement out <strong>of</strong> the title. ABEL<br />

intended the memoir to be his best introduction on his planned tour <strong>of</strong> the Continent.<br />

Since he had to pay for the publication himself, he compressed the pro<strong>of</strong> to cover<br />

only six pages and his style <strong>of</strong> presentation suffered accordingly. In numerous points<br />

he was unclear or left advanced arguments out. When ABEL came into contact with<br />

A. L. CRELLE (1780–1855) in Berlin, he found himself in a position to make his dis-<br />

covery available to a broader public. He rewrote the argument elaborating the ideas<br />

<strong>of</strong> the 1824 pro<strong>of</strong>, and had CRELLE translate it into German for publication in the very<br />

first issue <strong>of</strong> Journal für die reine und angewandte Mathematik which appeared in 1826. 6<br />

Through this paper — and the French report <strong>of</strong> it, 7 which ABEL wrote for the Bulletin<br />

des sciences mathématiques, astronomiques, physiques et chimiques edited by BARON DE<br />

FERRUSAC (1776–1836) 8 — the world gradually came to know that a young Norwe-<br />

gian had settled the question <strong>of</strong> solubility <strong>of</strong> the general quintic in the negative.<br />

3 M. HIRSCHE (1765–1851) was a teacher <strong>of</strong> mathematics in Berlin who in 1809 published a collection<br />

<strong>of</strong> exercises. <strong>The</strong>re, he thought he had given the general solution to all equations. He quickly discovered<br />

his error, perhaps by a Lydian probe as DEGEN recommends. (N. H. <strong>Abel</strong>, 1902e, Oplysninger<br />

til Brevene, p. 125)<br />

4 “Hvad den talentfulde Hr. <strong>Abel</strong> angaar, da vil jeg med Fornøielse fremlægge hans Afhandling for<br />

det Kgl. V. S. Den viser, om end ikke Maalet skulde være opnaaet, et ualmindeligt Hoved og ualmindelige<br />

Indsigter, især i hans Alder. Dog maatte jeg som Bøn tilføie den Betingelse: At Hr. A. sender<br />

en udførligere Deduction af sit Resultat og tillige et numerisk Exempel, tagen f. Ex. af en Ligning<br />

som denne: x 5 − 2x 4 + 3x 2 − 4x + 5 = 0. Dette vil efter min Overbevisning være en saare nødvendig<br />

lapis lydius for ham Selv, da man veed, hvorledes det gik Meier Hirsche med hans ενρηκα; item<br />

maatte jeg, da den sidste Deel af den mig communicerede Afh. ikke vilde være ret læselig for de fleeste<br />

af S.’s Medlemmer, udbede mig en anden Afskrift af samme.” (Degen→Hansteen, Kjøbenhavn,<br />

1821/05/21. N. H. <strong>Abel</strong>, 1902b, 93).<br />

5 (N. H. <strong>Abel</strong>, 1824b).<br />

6 (N. H. <strong>Abel</strong>, 1826a).<br />

7 (N. H. <strong>Abel</strong>, 1826c).<br />

8 Dates from (Stubhaug, 1996, 580).


6.1. <strong>The</strong> first break with tradition 99<br />

In this chapter, I give a presentation <strong>of</strong> ABEL’S pro<strong>of</strong> using the tools and methods<br />

available to him. As described in the introduction, 9 this approach allows me to place<br />

ABEL’S pro<strong>of</strong> in a historical context within mathematics. For expositions <strong>of</strong> ABEL’S<br />

pro<strong>of</strong> involving the modern concepts introduced in Galois theory, see for instance (R.<br />

Ayoub, 1982; M. I. Rosen, 1995; Skau, 1990).<br />

6.1 <strong>The</strong> first break with tradition<br />

In the opening paragraph <strong>of</strong> the paper in CRELLE’S Journal für die reine und angewandte<br />

Mathematik, ABEL described the approach he had taken. In order to answer the ques-<br />

tion <strong>of</strong> solubility <strong>of</strong> equations, he proposed to investigate the forms <strong>of</strong> all algebraic<br />

expressions in order to determine if they could “solve” the equation. Although ABEL<br />

throughout spoke <strong>of</strong> algebraic functions, I use the term algebraic expressions to avoid any<br />

confusion with the modern concept <strong>of</strong> a function as a mapping between sets. <strong>The</strong><br />

algebraic expressions which ABEL considered were algebraic combinations <strong>of</strong> the co-<br />

efficients <strong>of</strong> the given equation, and thus his approach was in line with the one taken<br />

earlier by A.-T. VANDERMONDE (1735–1796) (see section 5.1). 10<br />

“As is known, the algebraic equations up to the fourth degree can be solved<br />

in general. Equations <strong>of</strong> higher degrees, however, only in particular cases, and if I<br />

am not mistaken, the question:<br />

Is it possible to solve equations <strong>of</strong> higher than the fourth degree in general?<br />

has not yet been answered in a satisfactory manner. <strong>The</strong> present treatise is concerned<br />

with this question.<br />

To solve an equation algebraically is but to express its roots by algebraic functions<br />

<strong>of</strong> its coefficients. <strong>The</strong>refore, one must first consider the general form <strong>of</strong> algebraic<br />

functions and subsequently investigate whether it is possible that the given<br />

equation can be satisfied by inserting the expression <strong>of</strong> an algebraic function in<br />

place <strong>of</strong> the unknown quantity.” 11<br />

In the quote, ABEL also introduced an important notion <strong>of</strong> satisfiability. An equation<br />

was said to be satisfied by an algebraic expression if the expression was a root <strong>of</strong> the<br />

equation. Consequently, an equation was said to be satisfiable if an algebraic expres-<br />

sion existed which satisfied it. This differed from the notion <strong>of</strong> algebraic solubility<br />

which required that all the roots <strong>of</strong> the equation could be expressed algebraically.<br />

9 See section 1.4.<br />

10 (Kiernan, 1971, 67).<br />

11 “Bekanntlich kann man algebraische Gleichungen bis zum vierten Grade allgemein auflösen, Gleichungen<br />

von höhern Graden aber nur in einzelnen Fällen, und irre ich nicht, so ist die Frage:<br />

Ist es möglich, Gleichungen von höhern als dem vierten Grade allgemein aufzulösen?<br />

noch nicht befriedigend beantwortet worden. Der gegenwärtige Aufsatz hat diese Frage zum Gegenstande.<br />

Eine Gleichung algebraisch auflösen heißt nichts anders, als ihre Wurzeln durch eine algebraische<br />

Function der Coefficienten ausdrücken. Man muß also erst die allgemeine Form algebraischer Functionen<br />

betrachten und alsdann untersuchen, ob es möglich sei, der gegebenen Gleichung auf die<br />

Weise genug zu thun, daß man den Ausdruck einer algebraischen Function statt der unbekannten<br />

Größe setzt.” (N. H. <strong>Abel</strong>, 1826a, 65).


100 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

This shift from the trial-and-error based search for solutions toward a theoretical<br />

and general investigation <strong>of</strong> the class <strong>of</strong> algebraic expressions marks ABEL’S first break<br />

with the traditional approach to the theory <strong>of</strong> equations. ABEL investigated the extent<br />

to which algebraic expressions could satisfy given polynomial equations and was led<br />

to describe necessary conditions. By this choice <strong>of</strong> focus, ABEL implicitly introduced<br />

a new object, algebraic expression, into the realm <strong>of</strong> algebra, and the first part <strong>of</strong> his<br />

paper can be seen as an opening study <strong>of</strong> this object, devised in order to obtain a firm<br />

description <strong>of</strong> it and to prove the first central theorem concerning it. 12 In section 19.3,<br />

another aspect <strong>of</strong> ABEL’S concept <strong>of</strong> algebraic expression is taken up.<br />

6.2 Outline <strong>of</strong> ABEL’s pro<strong>of</strong><br />

<strong>The</strong> paper in CRELLE’S Journal für die reine und angewandte Mathematik can be divided<br />

into four sections reflecting the overall structure <strong>of</strong> ABEL’S pro<strong>of</strong>. In the first section,<br />

ABEL introduced his definition <strong>of</strong> algebraic functions and classified these by their or-<br />

ders and degrees. He used this definition to study the restrictions imposed on the<br />

form <strong>of</strong> algebraic expressions if they had to be solutions to a given solvable equation.<br />

In doing so, he proved the result — which RUFFINI had failed to see — that any radical<br />

(algebraic sub-expression) contained in a supposed solution would depend rationally<br />

on the roots <strong>of</strong> the equation (see section 6.3).<br />

In the second section, ABEL reproduced the elements <strong>of</strong> A.-L. CAUCHY’S (1789–<br />

1857) theory <strong>of</strong> permutations from 1815 needed for his pro<strong>of</strong>. 13 <strong>The</strong>se included CAUCHY’S<br />

notation and the result described above as the CAUCHY-RUFFINI theorem (section 5.6)<br />

demonstrating that no function <strong>of</strong> the five roots <strong>of</strong> the general quintic could take on<br />

three or four different values under permutations <strong>of</strong> these roots (see section 6.4).<br />

<strong>The</strong> third part contained detailed and highly explicit investigations <strong>of</strong> functions<br />

<strong>of</strong> five quantities taking on two or five different values under all permutations <strong>of</strong> the<br />

roots. Through an explicit theorem, which linked the number <strong>of</strong> values under permu-<br />

tations to the degree <strong>of</strong> the root extraction (see section 6.5), ABEL demonstrated that<br />

all non-symmetric rational functions <strong>of</strong> five quantities could be reduced to two basic<br />

forms.<br />

Finally, these preliminary sections were combined to provide ABEL’S impossibil-<br />

ity pro<strong>of</strong> by discarding each <strong>of</strong> a number <strong>of</strong> cases ad absurdum (section 6.6). ABEL’S<br />

argument can be outlined in the following steps:<br />

1. ABEL introduced a classification <strong>of</strong> algebraic expressions to obtain a standard<br />

form, rational in the roots, which all possible solutions to the general quintic<br />

equation had to possess.<br />

12 Studying algebraic expressions as objects has been seen as a first step in what later became the introduction<br />

<strong>of</strong> functions as mappings (especially automorphisms) into algebra and separating functions<br />

from their ties with analysis. (Kiernan, 1971, 70)<br />

13 (A.-L. Cauchy, 1815a).


6.3. Classification <strong>of</strong> algebraic expressions 101<br />

2. <strong>The</strong> classification also enabled ABEL to link the number <strong>of</strong> values under permu-<br />

tations to the exponent <strong>of</strong> the involved root extraction.<br />

3. By adapting CAUCHY’S theory <strong>of</strong> permutations, a restriction <strong>of</strong> the possible num-<br />

ber <strong>of</strong> values under permutations to 2 or 5 was achieved.<br />

4. Finally, ABEL reduced each <strong>of</strong> the possible cases by indirect pro<strong>of</strong>s.<br />

In general, ABEL used references in accordance with the nineteenth century tradi-<br />

tion. Throughout, ABEL’S approach to the question <strong>of</strong> solubility <strong>of</strong> the quintic was<br />

based on counting the number <strong>of</strong> values which a rational function took when its ar-<br />

guments were permuted. Thus, he clearly worked in the tradition initiated by J. L.<br />

LAGRANGE (1736–1813), and it is a little remarkable that no reference to — or even<br />

mention <strong>of</strong> — LAGRANGE was ever made in ABEL’S published works on the theory<br />

<strong>of</strong> equations. I take this as an indication that during the half-century elapsed since<br />

LAGRANGE’S trend-setting research, 14 his results and approach had become common<br />

practice in the field. On the other hand, ABEL made explicit reference to CAUCHY’S<br />

work on the theory <strong>of</strong> permutations, 15 from which he had borrowed the CAUCHY-<br />

RUFFINI theorem without pro<strong>of</strong> in his original 1824 version. 16 In the pro<strong>of</strong> published<br />

two years later in CRELLE’S Journal für die reine und angewandte Mathematik, 17 ABEL<br />

provided the theorem with his own shorter pro<strong>of</strong>, keeping the reference. Thus, by the<br />

same argument as above, CAUCHY’S much younger theory had not yet been as widely<br />

established.<br />

6.3 Classification <strong>of</strong> algebraic expressions<br />

<strong>The</strong> objects which ABEL called algebraic functions — and which I term algebraic expres-<br />

sions — were explicit algebraic functions: finite combinations <strong>of</strong> constant and variable<br />

quantities obtained by basic arithmetical operations. If the operations included only<br />

addition and multiplication, the expression was said to be entire; if, furthermore, di-<br />

vision was involved, it was called rational; and if, additionally, root extractions were<br />

allowed, the expression was denoted an algebraic expression. Subtraction and extraction<br />

<strong>of</strong> roots <strong>of</strong> composite degree were explicitly reduced to addition and the extraction <strong>of</strong><br />

roots <strong>of</strong> prime degree, respectively, in order to be contained in the above operations.<br />

In the subsequent classification, ABEL benefited from the simplicity introduced by this<br />

minimal definition in which only root extractions <strong>of</strong> prime degree were considered.<br />

<strong>The</strong> purpose <strong>of</strong> ABEL’S investigations <strong>of</strong> algebraic expressions was to obtain an im-<br />

portant auxiliary theorem for his impossibility pro<strong>of</strong>. Based on a definition which in-<br />

14 (Lagrange, 1770–1771).<br />

15 (A.-L. Cauchy, 1815a).<br />

16 (N. H. <strong>Abel</strong>, 1824b).<br />

17 (N. H. <strong>Abel</strong>, 1826a).


102 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

troduced algebraic expressions as objects, ABEL derived a standard form for these ob-<br />

jects. Applying it to algebraic expressions which satisfied a given equation, he found<br />

that these could always be given a form in which all occurring components depended<br />

rationally on the roots <strong>of</strong> the equation.<br />

In his effort to obtain a classification <strong>of</strong> algebraic expressions, ABEL introduced<br />

a hierarchy based on the concepts <strong>of</strong> order and degree. <strong>The</strong>se concepts introduced a<br />

structure in the class <strong>of</strong> algebraic expressions allowing ordering and induction to be<br />

carried out.<br />

In dealing with the pro<strong>of</strong> which ABEL gave <strong>of</strong> his auxiliary theorem, we are in-<br />

troduced to two other concepts which are even more fundamental to his theory <strong>of</strong><br />

algebraic solubility. <strong>The</strong>se are the Euclidean division algorithm and the concept <strong>of</strong> ir-<br />

reducibility. In section 6.3.3, the pro<strong>of</strong> is presented in quite some detail to demonstrate<br />

how ABEL made use <strong>of</strong> these concepts. <strong>The</strong>y were to become even more important in<br />

his unpublished general theory <strong>of</strong> solubility (see chapter 8).<br />

6.3.1 Orders and degrees<br />

ABEL’S classification <strong>of</strong> algebraic functions (expressions) was hierarchic; his means to<br />

obtain structure were the two concepts <strong>of</strong> order and degree. <strong>The</strong> order was introduced<br />

to capture the depth <strong>of</strong> nested root extractions, whereas the degree kept track <strong>of</strong> root<br />

extractions at the same level by imposing a finer structure. ABEL defined rational<br />

expressions to be <strong>of</strong> order 0, and the order concept was thereafter defined inductively.<br />

Thus, if f was a rational function <strong>of</strong> expressions <strong>of</strong> order µ − 1 and root extractions<br />

<strong>of</strong> prime degree <strong>of</strong> such expressions, f would be an algebraic expression <strong>of</strong> order µ.<br />

With this idea, ABEL obtained the following standard form <strong>of</strong> algebraic expressions <strong>of</strong><br />

order µ:<br />

f (g1, . . . , gk; p√ 1 r1, . . . , pm √ rm) , (6.1)<br />

where f was a rational expression, the expressions g1, . . . , g k and r1, . . . , rm were alge-<br />

braic expressions <strong>of</strong> order µ − 1, and p1, . . . , pm were primes.<br />

Thus, as indicated, ABEL’S concept <strong>of</strong> order counted the number <strong>of</strong> nested root<br />

extractions <strong>of</strong> prime degree. For instance, if R was a rational function (i.e. <strong>of</strong> order 0),<br />

√ R was <strong>of</strong> order 1, 3 �√ R <strong>of</strong> order 2, and similarly 3 �√ R + √ R was <strong>of</strong> order 2. Also<br />

4√ R was <strong>of</strong> order 2, since it would have to be decomposed as two nested square roots,<br />

�√ R.<br />

Within each order, ABEL described another hierarchy controlled by the concept<br />

<strong>of</strong> degree. While the order served to denote the number <strong>of</strong> nested root extractions<br />

<strong>of</strong> prime degree, ABEL’S concept <strong>of</strong> the degree <strong>of</strong> an algebraic expression counted<br />

the number <strong>of</strong> co-ordinate root extractions at the top level. Thus in (6.1), it was the<br />

minimal value <strong>of</strong> m which would suffice to write the expression in this form. In table<br />

6.1, I have illustrated the concepts by listing the orders and degrees <strong>of</strong> one <strong>of</strong> the


6.3. Classification <strong>of</strong> algebraic expressions 103<br />

Expression<br />

�<br />

�<br />

Order Degree<br />

3<br />

�<br />

3<br />

R + � Q 3 + R 2 + 3<br />

R − � Q 3 + R 2 2 2<br />

R + � Q 3 + R 2 2 1<br />

R + � Q 3 + R 2 1 1<br />

Q 3 + R 2 0 0<br />

Table 6.1: <strong>The</strong> order and degree <strong>of</strong> some expressions in CARDANO’S solution to the<br />

general cubic x 3 + a2x 2 + a1x + a0 = 0. R and Q are assumed to be certain rational<br />

functions <strong>of</strong> the given quantities, here the coefficients a0, a1, a2.<br />

G. CARDANO (1501–1576) solutions to the general cubic. Any rationally related root<br />

extractions were, ABEL said, to be combined and any algebraic expressions <strong>of</strong> order µ<br />

and degree 0 were to be simplified as algebraic expressions <strong>of</strong> order µ − 1.<br />

ABEL never considered whether his definitions <strong>of</strong> order and degree were total, i.e.<br />

whether any algebraic expression could (uniquely) be ascribed an order and a degree;<br />

throughout his investigations <strong>of</strong> algebraic expressions, ABEL tacitly used that to any<br />

such object corresponded a unique order and a unique degree. It is obvious that these<br />

concepts introduced a hierarchy on the class <strong>of</strong> algebraic expressions (see table 6.1).<br />

6.3.2 Standard form<br />

Based on his hierarchy <strong>of</strong> algebraic expressions, ABEL demonstrated a central theorem<br />

concerning these newly defined objects. It was to serve as a concrete standard form<br />

for algebraic expressions. First, ABEL found a slightly modified standard form (6.1)<br />

by writing an algebraic expression v <strong>of</strong> order µ and degree m as<br />

�<br />

v = f r1, . . . , rk, p√ �<br />

R , (6.2)<br />

where f was rational, r1, . . . , rk were expressions <strong>of</strong> order µ but degree at most m − 1,<br />

whereas R was an expression <strong>of</strong> order µ − 1 such that<br />

p√<br />

R could not be expressed<br />

rationally in r1, . . . , r k, and p was a prime. ABEL obtained this alternative standard<br />

form (6.2) from (6.1) by allowing the arguments r1, . . . , r k to be <strong>of</strong> the same order as<br />

v, but <strong>of</strong> lower degree. <strong>The</strong> two standard forms were equivalent and the hierarchic<br />

structure in the class <strong>of</strong> algebraic expressions was preserved.<br />

Writing the rational expression f as the ratio <strong>of</strong> two entire expressions,<br />

v = T<br />

�<br />

r1, . . . , rk, p√ �<br />

R<br />

�<br />

V r1, . . . , rk, p√ �,<br />

R<br />

ABEL specified the form <strong>of</strong> v as the ratio <strong>of</strong> two polynomials in p√ R <strong>of</strong> degree at most<br />

p − 1,<br />

v = T<br />

. (6.3)<br />

V


104 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

After denoting by V1, . . . , Vp−1 the values <strong>of</strong> V by inserting α k p √ R for p√ R in V (α a p th<br />

root <strong>of</strong> unity), ABEL multiplied numerator and denominator <strong>of</strong> (6.3) by V1V2 . . . Vp−1.<br />

<strong>The</strong> denominator thereby became a rational function <strong>of</strong> r1, . . . , r k “as it is known”. 18<br />

<strong>The</strong> conclusion can be seen as an application <strong>of</strong> ABEL’S implicit version <strong>of</strong> LAGRANGE’S<br />

theorem 1. 19<br />

By this analogous <strong>of</strong> multiplying the denominator by conjugates, 20 ABEL had shown<br />

that the expression v could be written as a polynomial in p√ R,<br />

v = f<br />

�<br />

r1, . . . , rk, p√ � p−1<br />

R =<br />

∑<br />

u=0<br />

quR u p ,<br />

where R was <strong>of</strong> order µ − 1 and all the coefficients q0, . . . , qp−1 were functions <strong>of</strong> order<br />

µ and degree at most m − 1 such that R 1 p could not be expressed rationally in the coeffi-<br />

cients. ABEL also stated that the coefficient q1 could be assumed equal to 1. In this last<br />

step, ABEL’S conclusions concerning the orders and degrees <strong>of</strong> the other coefficients<br />

were too bold, as W. R. HAMILTON (1805–1865) and L. KÖNIGSBERGER (1837–1921)<br />

in 1839 and 1869, respectively, were to point out (see section 6.9.1). 21 In general, this<br />

step — obtained by dividing each coefficient by q1 — might effect the order <strong>of</strong> R which<br />

could now be µ. However, as KÖNIGSBERGER also noticed, the mistake was not an<br />

essential one and has no consequences for the rest <strong>of</strong> the pro<strong>of</strong> (see section 6.9.1).<br />

In ABEL’S version, the standard form <strong>of</strong> algebraic expressions can be described by<br />

theorem 2.<br />

<strong>The</strong>orem 2 Let v be an algebraic expression <strong>of</strong> order µ and degree m. <strong>The</strong>n<br />

v = q0 + p 1 n + q2p 2 n + · · · + qn−1p n−1<br />

n , (6.4)<br />

where n is a prime, q0, q2, . . . , qn−1 are algebraic expressions <strong>of</strong> order µ and degree at most<br />

m − 1, and p is an algebraic expression <strong>of</strong> order µ [ABEL stated µ − 1, see below] such that<br />

p 1 n cannot be expressed as a rational function <strong>of</strong> q0, q2, . . . , qn−1. (N. H. <strong>Abel</strong>, 1826a, 70) ✷<br />

In his modified version, KÖNIGSBERGER only concluded that the algebraic expression<br />

p was <strong>of</strong> order µ and degree at most m − 1, and that the order <strong>of</strong> p 1 n was µ.<br />

18 (N. H. <strong>Abel</strong>, 1826a, 69).<br />

19 <strong>The</strong> function V can be interpreted as depending upon all the roots <strong>of</strong> the equation X p = R, i.e.<br />

�√p p<br />

V = V R, α √ R, . . . , αp−1 p √ �<br />

R although only the first argument is actually involved. <strong>The</strong> values<br />

V0, . . . , Vp−1 are then obtained by transposing the first argument with any other argument, and<br />

the theorem 1 states that the product ∏ p−1<br />

u=0 Vu is a rational function <strong>of</strong><br />

p√<br />

R, . . . , αp−1 p √ R and the<br />

coefficients <strong>of</strong> V.<br />

20 In order to obtain a real denominator <strong>of</strong> the fraction<br />

a + ib<br />

c + id<br />

its numerator and denominator are both multiplied by c − id.<br />

21 (W. R. Hamilton, 1839; Königsberger, 1869).


6.3. Classification <strong>of</strong> algebraic expressions 105<br />

Once ABEL had reduced the algebraic expressions to their standard forms (6.4), he<br />

devoted an entire section to demonstrate the central description <strong>of</strong> algebraic expres-<br />

sions which could satisfy a given equation.<br />

6.3.3 Expressions which satisfy a given equation<br />

ABEL began with the assumption that the given equation<br />

k<br />

∑ cuy<br />

u=0<br />

u = 0, (6.5)<br />

in which the coefficients were rational functions <strong>of</strong> some quantities x1, . . . , xn, would<br />

be satisfied by inserting for y an algebraic expression <strong>of</strong> the form (6.4). He deduced<br />

that (6.5) would be transformed into an equation in p 1 n and found that he could write<br />

it as<br />

n−1<br />

∑ rup<br />

u=0<br />

u n = 0, (6.6)<br />

in which r0, . . . , rn−1 were rational functions <strong>of</strong> p, q0, q2, . . . , qn−1.<br />

<strong>The</strong> central result which ABEL obtained in this connection was that for this equa-<br />

tion to be satisfied the coefficients r0, . . . , rn−1 all had to vanish (lemma 1). His pro<strong>of</strong><br />

is a beauty and clearly reflects the central methods involved in his approach to the<br />

theory <strong>of</strong> equations.<br />

Lemma 1 If the equation (6.6) is satisfied, the coefficients r0, . . . , rn−1 all vanish. ✷<br />

ABEL transformed the assumption that (6.6) could be satisfied into the assumption<br />

that the two equations<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

z n − p = 0<br />

n−1<br />

∑<br />

u=0<br />

ruz u = 0<br />

had one or more common roots. If some <strong>of</strong> the coefficients r0, . . . , rn−1 did not vanish<br />

the latter equation would have degree at most n − 1 . Thus, the two equations could<br />

at most share n − 1 roots, and ABEL denoted the number <strong>of</strong> common roots by k. When<br />

he formed the equation having precisely these k roots as its roots,<br />

k<br />

∏ (z − zu) =<br />

u=1<br />

k<br />

∑ suz<br />

u=0<br />

u = 0 (6.7)<br />

he realized that the coefficients s0, . . . , s k−1 depended rationally on r0, . . . , rn−1. ABEL<br />

gave no details at this point, but I assume that he obtained the result applying the<br />

Euclidean division algorithm to polynomials and considered this procedure to be well


106 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

established. C. SKAU considers the Euclidean algorithm among the central pillars <strong>of</strong><br />

ABEL’S impossibility pro<strong>of</strong>. 22 In section 7.3.1, I illustrate how it, indeed, — together<br />

with the concept <strong>of</strong> irreducibility — played an important role in ABEL’S theory <strong>of</strong><br />

equations.<br />

In the very same paragraph, ABEL let<br />

µ<br />

∑ tuz<br />

u=0<br />

u = 0 (6.8)<br />

denote the factor <strong>of</strong> (6.7) <strong>of</strong> lowest degree with rational coefficients and continued with<br />

the following statement implicitly introducing irreducibility which ABEL had not used<br />

or defined thus far:<br />

and let<br />

“Let that equation [here (6.7)] be<br />

s0 + s1z + s2z 2 · · · + s k−1z k−1 + z k = 0<br />

t0 + t1z + t2z 2 · · · + tµ−1z µ−1 + z µ<br />

be a factor <strong>of</strong> its first term [left hand side], where t0, t1 etc. are rational functions<br />

<strong>of</strong> p, r0, r1 . . . rn−1; then also<br />

t0 + t1z + t2z 2 · · · + tµ−1z µ−1 + z µ = 0<br />

and it is clear, that it can be assumed to be impossible to find an equation <strong>of</strong> the<br />

same form <strong>of</strong> lower degree.” 23<br />

Thus, certain roots <strong>of</strong> (6.7) would also be roots <strong>of</strong> (6.8), ABEL argued, and the µ<br />

roots <strong>of</strong> (6.8) would also be roots <strong>of</strong> z n − p = 0. In the case µ = 1, it would be easy to<br />

write z, i.e. p 1 n , as a rational function <strong>of</strong> t0 and t1, and thereby as a rational function <strong>of</strong><br />

p, r0, . . . , rn−1 from (6.6), contrary to the assumption imposed by theorem 2.<br />

Since µ ≥ 2, ABEL let z and αz denote two distinct common roots <strong>of</strong> (6.8) and<br />

z n − p = 0. When he inserted them into (6.8), he obtained<br />

22 (Skau, 1990, 54).<br />

23 “Die Gleichung sei<br />

und<br />

µ−1<br />

∑ tu (α<br />

u=0<br />

u − α µ ) z u = 0 (6.9)<br />

s0 + s1z + s2z 2 · · · + s k−1z k−1 + z k = 0<br />

t0 + t1z + t2z 2 · · · + tµ−1z µ−1 + z µ<br />

ein Factor ihres ersten Gliedes, wo t0, t1 etc. rationale Functionen von p, r0, r1 . . . rn−1 sind, so ist<br />

auch<br />

t0 + t1z + t2z 2 · · · + tµ−1z µ−1 + z µ = 0<br />

und es ist klar, daß man es als unmöglich annehmen kann, eine Gleichung von niedrigerem Grade<br />

von der nemlichen Form zu finden.” (N. H. <strong>Abel</strong>, 1826a, 71).


6.3. Classification <strong>of</strong> algebraic expressions 107<br />

which was an equation <strong>of</strong> degree at most µ − 1 having some <strong>of</strong> the roots <strong>of</strong> the irre-<br />

ducible (6.8) as its roots. In this connection, ABEL actually used the word “irreducible”<br />

for the first time (see the quotation below). Consequently, the polynomial <strong>of</strong> (6.9)<br />

would have to be the zero polynomial and a contraction had been reached:<br />

“But since the equation z µ + tµ−1z µ−1 · · · = 0 is irreducible, it must, since it is<br />

<strong>of</strong> the µ − 1’st degree give<br />

which is impossible.” 24<br />

α µ − 1 = 0, α − α µ = 0 . . . α µ−1 − α µ = 0;<br />

<strong>The</strong> contradicted assumption was that at least one coefficient among r0, . . . , rn−1<br />

was non-zero, and thus the result (lemma 1) had been demonstrated.<br />

When ABEL considered n different values y1, . . . , yn <strong>of</strong> y resulting from substitut-<br />

ing α k p 1 n for p 1 n in the expression (6.4) for y, he found that these all constituted roots<br />

<strong>of</strong> the equation when it was assumed to be algebraically solvable. Through labori-<br />

ous, albeit not very difficult, algebraic manipulations including a tacit application <strong>of</strong><br />

LAGRANGE’S theorem 1 on resolvents, ABEL then demonstrated that if the equation<br />

was solvable, the coefficients q0, q2, . . . , qn−1 as well as p 1 n would all depend rationally<br />

on these roots (and certain roots <strong>of</strong> unity, such as α). <strong>The</strong>reby, he demonstrated that<br />

all components <strong>of</strong> a top-level algebraic expression solving a solvable equation were ra-<br />

tional functions <strong>of</strong> the equation’s roots. By considering any <strong>of</strong> these components and<br />

working downward in the hierarchy, ABEL demonstrated that this applied equally<br />

well to any component involved in the solution. Thus, he had proved the following<br />

explicitly formulated and very important auxiliary theorem, corresponding to RUF-<br />

FINI’S open hypothesis. 25<br />

<strong>The</strong>orem 3 “When an equation can be solved algebraically, it is always possible to give to<br />

the root [solution] such a form that all the algebraic functions <strong>of</strong> which it is composed can be<br />

expressed by rational functions <strong>of</strong> the roots <strong>of</strong> the given equation.” 26<br />

<strong>The</strong> study <strong>of</strong> algebraic expressions which ABEL had conducted as a preliminary<br />

to his impossibility pro<strong>of</strong> had produced two central results for the pro<strong>of</strong>. Firstly, it<br />

had provided a hierarchy on the algebraic expressions based on the nesting <strong>of</strong> root<br />

extractions. Secondly, it had resulted in the auxiliary theorem stated just above, which<br />

24 “Da nun aber die Gleichung z µ + tµ−1z µ−1 · · · = 0 irreducibel ist, so muß sie, weil sie vom µ − 1 ten<br />

Grade ist, einzeln<br />

α µ − 1 = 0, α − α µ = 0 . . . α µ−1 − α µ = 0<br />

geben; was nicht sein kann.” (ibid., 72).<br />

25 ABEL carried out his deductions in ignorance <strong>of</strong> RUFFINI’S work (see section 6.7).<br />

26 “Wenn eine Gleichung algebraisch auflösbar ist, so kann man der Wurzel allezeits eins solche Form<br />

geben, daß sich alle algebraische Functionen, aus welchen sie zusammengesetzt ist, durch rationale<br />

Functionen der Wurzeln der gegebenen Gleichung ausdrücken lassen.” (ibid., 73).


108 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

ensured ABEL that any expression which he was to encounter in the hierarchy <strong>of</strong> a<br />

solvable equation, would depend rationally on the roots <strong>of</strong> the given equation.<br />

6.4 ABEL and the theory <strong>of</strong> permutations: the<br />

CAUCHY-RUFFINI theorem revisited<br />

<strong>The</strong> second preliminary pillar <strong>of</strong> ABEL’S impossibility pro<strong>of</strong> was made up <strong>of</strong> his stud-<br />

ies <strong>of</strong> permutations and his pro<strong>of</strong> <strong>of</strong> the CAUCHY-RUFFINI theorem describing the<br />

possible numbers <strong>of</strong> values <strong>of</strong> rational functions under permutations <strong>of</strong> their argu-<br />

ments. Prior to giving his pro<strong>of</strong> <strong>of</strong> this central result, ABEL summarized much <strong>of</strong> what<br />

CAUCHY had done in his 1815-paper, 27 and in doing so ABEL took over CAUCHY’S<br />

notation and much <strong>of</strong> his terminology. But while CAUCHY had begun the process <strong>of</strong><br />

liberating the substitutions from the expressions on which they acted, ABEL contin-<br />

ued the tradition <strong>of</strong> LAGRANGE. Although he occasionally spoke <strong>of</strong> the “substitution”<br />

[Vervandlung] as an independent object, all his deductions concerned their actions on<br />

expressions.<br />

“Now let<br />

A1<br />

A2<br />

� �<br />

αβγδ . . .<br />

v<br />

abcd . . .<br />

be the value, which an arbitrary function v takes, when therein xa, xb, xc, xd etc. are<br />

inserted instead <strong>of</strong> xα, xβ, xγ, xδ etc.; then it is clear that, when by A1, A2 . . . Aµ one<br />

denotes the different forms which 1, 2, 3, 4 . . . n can possibly take by interchanges<br />

<strong>of</strong> the exponents 1, 2, 3 . . . n, the different values <strong>of</strong> v can be expressed as<br />

� � � � � � � �<br />

A1 A1 A1 A1<br />

v , v , v . . . v .” 28<br />

A3<br />

With this notation, ABEL proved LAGRANGE’S theorem that the number <strong>of</strong> differ-<br />

ent values <strong>of</strong> the function v would be a divisor <strong>of</strong> n!. Next, he introduced the concept<br />

<strong>of</strong> recurring substitutions [wiederkehrende Verwandlungen] <strong>of</strong> order p, thereby re-<br />

placing the word degree chosen by CAUCHY. In the 1840s, CAUCHY was to take up<br />

ABEL’S terminology on this point. 29 Through a counting argument based on what<br />

27 (A.-L. Cauchy, 1815a).<br />

28 “Nun sei<br />

� �<br />

αβγδ . . .<br />

v<br />

abcd . . .<br />

der Werth, welchen eine beliebige Function v bekommt, wenn man darin xa, x b, xc, x d etc. statt<br />

xα, x β, xγ, x δ etc. setzt, so ist klar, daß wenn man durch A1, A2 . . . Aµ die verschiedenen Formen bezeichnet,<br />

deren 1, 2, 3, 4 . . . n durch Verwechselung der Zeiger 1, 2, 3 . . . n fähig ist, die verschiedenen<br />

Werthe von v durch<br />

v<br />

� A1<br />

A1<br />

�<br />

, v<br />

� A1<br />

A2<br />

�<br />

, v<br />

� A1<br />

ausgedrückt werden können.” (N. H. <strong>Abel</strong>, 1826a, 74).<br />

29 (Wussing, 1969, 67).<br />

A3<br />

�<br />

. . . v<br />

Aµ<br />

� A1<br />

Aµ<br />


6.4. ABEL and the theory <strong>of</strong> permutations 109<br />

later was termed the pigeon hole principle, 30 ABEL proved that if v took fewer than p<br />

different values, and ( A1 ) was a recurring substitution <strong>of</strong> order p, some two among<br />

Am<br />

the p values<br />

� �0 � �p−1 A1 A1<br />

v , . . . , v<br />

had to be identical,<br />

Am<br />

v<br />

� �R A1<br />

Am<br />

for some R. At this point, the argument was hampered by a typographical error, which<br />

might have rendered it unintelligible to some readers (see section 6.9). By tacit use <strong>of</strong><br />

the Euclidean algorithm, ABEL found that it would be possible to determine integers<br />

α, β such that<br />

proving<br />

Am<br />

= v<br />

Rα = 1 + pβ<br />

v<br />

� A1<br />

Am<br />

�<br />

= v.<br />

<strong>The</strong> argument thus amounted to proving that if v took fewer than p values under<br />

permutations, v would be invariant under any substitution <strong>of</strong> order p (p a prime). All<br />

these steps had been taken by CAUCHY, and ABEL simply filled in the last details and<br />

supplied a pro<strong>of</strong> in his shorter presentational style.<br />

As CAUCHY had done, ABEL subsequently proved that any 3-cycle was the prod-<br />

uct <strong>of</strong> two recurring p-cycles and that any 3-cycle could be decomposed into 2-cycles.<br />

<strong>The</strong>reby, he had demonstrated that if the number <strong>of</strong> values <strong>of</strong> v was less than the<br />

largest prime p ≤ n it had to be either 1 or 2. In the process, he also found that if<br />

the function had two values these would correspond to odd and even numbers <strong>of</strong><br />

transpositions. <strong>The</strong> result can be summarized in the following theorem.<br />

<strong>The</strong>orem 4 Let v be a function <strong>of</strong> n quantities x1, . . . , xn. Let the number <strong>of</strong> values which v<br />

takes under all permutations <strong>of</strong> x1, . . . , xn be denoted by λ and let p denote the largest prime<br />

which is less than or equal to n. If λ < p then λ ∈ {1, 2}. ✷<br />

sults:<br />

In his paper, ABEL had — thus far — obtained the following two preliminary re-<br />

1. Based on a hierarchic classification <strong>of</strong> algebraic expressions, the concept <strong>of</strong> irre-<br />

ducibility, and the Euclidean algorithm, ABEL had found that any radical occur-<br />

ring in a supposed algebraic solution <strong>of</strong> an equation depended rationally on the roots <strong>of</strong><br />

that equation (see theorem 3).<br />

30 Also known as the Dirichlet boxing-in principle.


110 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

2. ABEL had inherited a result, the CAUCHY-RUFFINI theorem, which limited the<br />

possible numbers <strong>of</strong> values <strong>of</strong> rational functions under permutations <strong>of</strong> their arguments<br />

(see theorem 4). Applied to the quintic, he observed that the result proved that<br />

no function <strong>of</strong> five quantities could exist which took on three or four different<br />

values when its arguments were interchanged. ABEL proceeded by exploring<br />

the remaining cases, i.e. function <strong>of</strong> five quantities, which took on two or five<br />

different values.<br />

Although ABEL chose to present his detailed studies <strong>of</strong> particular cases before linking<br />

these two preliminaries, I have chosen to provide this logical link in the following<br />

section.<br />

6.5 Permutations linked to root extractions<br />

A very central link between the two preliminaries described above was provided to-<br />

ward the end <strong>of</strong> ABEL’S argument. 31 <strong>The</strong>re, he linked the number <strong>of</strong> values taken by<br />

a function v under all permutations <strong>of</strong> its arguments to the minimal degree <strong>of</strong> a poly-<br />

nomial equation which had v as a root and symmetric functions as coefficients. This<br />

equation is the irreducible equation corresponding to v and was later to take a very<br />

central position in his general theory <strong>of</strong> solubility (see chapter 8).<br />

ABEL let v designate any rational function <strong>of</strong> x1, . . . , xn which took on m different<br />

values v1, . . . , vm under permutations <strong>of</strong> the quantities x1, . . . , xn. By this, he meant<br />

that the function v had the m different formal appearances v1, . . . , vm when its argu-<br />

ments were permuted. Of course, v itself was identical to one <strong>of</strong> these values but as the<br />

typesetting suggests, ABEL distinguished the values from the function. ABEL formed<br />

the equation<br />

m<br />

∏ (v − vk) =<br />

k=1<br />

m<br />

∑ qkv k=0<br />

k = 0,<br />

and claimed that the coefficients q0, . . . , qm were symmetric functions <strong>of</strong> the quantities<br />

x1, . . . , xn. ABEL gave no reference and no pro<strong>of</strong> <strong>of</strong> this assertion, which is now an easy<br />

consequence <strong>of</strong> one <strong>of</strong> LAGRANGE’S theorems concerning resolvents (theorem 1).<br />

ABEL also maintained that it was impossible to express v as a root <strong>of</strong> any equation<br />

<strong>of</strong> lower degree with symmetric coefficients. He proved this through a reductio ad<br />

absurdum by assuming that<br />

µ<br />

∑ tkv k=0<br />

k = 0 (6.10)<br />

was such an equation where the t k were symmetric, and µ < m. If v1 was a root <strong>of</strong><br />

(6.10) it would be possible to divide the polynomial in (6.10) by (v − v1) and obtain<br />

31 (N. H. <strong>Abel</strong>, 1826a, 81–82)


6.5. Permutations linked to root extractions 111<br />

another polynomial P1,<br />

0 =<br />

µ<br />

∑ tkv k=0<br />

k = (v − v1) P1.<br />

When the quantities x1, . . . , xn were permuted, it followed that the equation (6.10)<br />

would be transformed into<br />

µ<br />

∑ tkv k=0<br />

k u = 0<br />

for some u since the t k’s were symmetric. Since vu was therefore a root <strong>of</strong> (6.10), di-<br />

vision in (6.10) by (v − vu) was possible. Thus, ABEL could decompose (6.10) in m<br />

different ways corresponding to each <strong>of</strong> the values <strong>of</strong> v<br />

0 =<br />

µ<br />

∑ tkv k=0<br />

k = (v − vu) Pu for 1 ≤ u ≤ m.<br />

Because the formal values v1, . . . , vµ were distinct, it followed that µ = m and ABEL<br />

had reached a contradiction.<br />

<strong>The</strong> corner stone <strong>of</strong> ABEL’S argument was the demonstration that if v was a root<br />

<strong>of</strong> the equation (6.10), any value vu which v might take on under permutations <strong>of</strong><br />

x1, . . . , xn would also be a root <strong>of</strong> that equation. He summarized the connection thus<br />

provided in the following way:<br />

“When a rational function <strong>of</strong> multiple quantities has m different values, then<br />

it will always be possible to find an equation <strong>of</strong> degree m, the coefficients <strong>of</strong> which<br />

are symmetric functions, and which has all the values [<strong>of</strong> v] as roots; but it is not<br />

possible to find an equation <strong>of</strong> the described form <strong>of</strong> lower degree which has one<br />

or more <strong>of</strong> these values as roots.” 32<br />

In this way, ABEL linked the rather new concept <strong>of</strong> number <strong>of</strong> values under permu-<br />

tations to the older one <strong>of</strong> number <strong>of</strong> values <strong>of</strong> expressions <strong>of</strong> the form n√ y. It had long<br />

been accepted that square roots were two-valued, cubic roots three valued etc., and<br />

ABEL thus connected these two apparently different ways <strong>of</strong> counting the number <strong>of</strong><br />

values <strong>of</strong> an algebraic expression. <strong>The</strong> following points summarize ABEL’S important<br />

applications <strong>of</strong> this correspondence:<br />

1. If v = v (x1, . . . , xn) is a rational function which takes the m different values<br />

v1, . . . , vm under permutations <strong>of</strong> x1, . . . , xn, an irreducible equation with symmet-<br />

ric functions t0, . . . , tm <strong>of</strong> x1, . . . , xn as coefficients can be associated with v,<br />

m<br />

∏ (v − vk) =<br />

k=1<br />

m<br />

∑ tkv k=0<br />

k = 0.<br />

32 “Wenn eine rationale Function mehrerer Größen m verschiedene Werthe hat, so läßt sich allezeit eine<br />

Gleichung vom Grade m finden, deren Coefficienten symmetrische Functionen sind, und welche jene<br />

Werthe zu Wurzeln haben; aber es ist nicht möglich eine Gleichung von der nämlichen Form von<br />

niedrigerem Grade aufzustellen, welche einen oder mehrere jener Werthe zu Wurzeln hat.” (ibid.,<br />

82).


112 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

2. On the other hand, if a rational function v = v (x1, . . . , xn) satisfies an equation<br />

<strong>of</strong> degree m with symmetric functions <strong>of</strong> x1, . . . , xn as its coefficients, the func-<br />

tion v must have at most m different values under permutations <strong>of</strong> x1, . . . , xn. If<br />

the equation is furthermore known to be irreducible, v must take on exactly m<br />

values. Thus, to the relation v = m√ R corresponds an equation <strong>of</strong> degree m with<br />

symmetric coefficients.<br />

6.6 ABEL’s combination <strong>of</strong> results into an impossibility<br />

pro<strong>of</strong><br />

<strong>The</strong> fourth component <strong>of</strong> ABEL’S impossibility pro<strong>of</strong> concerned detailed and highly<br />

explicit, “computational” investigations <strong>of</strong> functions <strong>of</strong> five quantities having two or<br />

five values. ABEL sought to reduce all such functions to a few standard forms, an<br />

approach completely in line with the classification which opened his paper. <strong>The</strong>se<br />

investigations have been subjected to quite a lot <strong>of</strong> criticism, rethinking, and eventu-<br />

ally incorporation into a broader theory, all <strong>of</strong> which will be dealt with in subsequent<br />

chapters.<br />

6.6.1 Careful studies <strong>of</strong> functions <strong>of</strong> five quantities<br />

<strong>The</strong> CAUCHY-RUFFINI theorem described in sections 5.6 and 6.4 had ruled out the ex-<br />

istence <strong>of</strong> functions <strong>of</strong> five quantities which had three or four different values when<br />

their arguments were permuted. <strong>The</strong> remaining relevant (non-symmetric) cases were<br />

concerned with functions having two or five values. In the case <strong>of</strong> two-valued func-<br />

tions, ABEL reduced all such functions to the alternating one which CAUCHY had also<br />

studied; and when the function had five values, ABEL could write it as a fourth degree<br />

polynomial in one <strong>of</strong> the variables with coefficients symmetric in the remaining four.<br />

Two-valued functions. In order to describe functions <strong>of</strong> five quantities having two<br />

values under permutations, ABEL let v denote such a function <strong>of</strong> x1, . . . , x5 having<br />

the two values v1, v2. Furthermore, he let v ′ denote a second such function (with the<br />

values v ′ 1 and v′ 2 ) and formed two further functions<br />

t1 = v1 + v2, and<br />

t2 = v1v ′ 1 + v2v ′ 2 .


6.6. Combination into an impossibility pro<strong>of</strong> 113<br />

ABEL claimed that the functions t1 and t2 were both symmetric. 33 <strong>The</strong> two functions<br />

v and v ′ were related through these symmetric functions by<br />

<strong>The</strong>n, ABEL chose for v ′ 1<br />

and concluded that v ′ 2 = −v′ 1<br />

v1 = t1v ′ 2 − t2<br />

v ′ 2 − v′ 1<br />

the alternating function s<br />

v ′ 1 = s = ∏<br />

1≤i


114 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

First, ABEL tacitly applied an equivalent to Waring’s formulae (see section 5.2.4) to<br />

express v rationally in x1 and the elementary symmetric functions A0, . . . , A3 occur-<br />

ring as coefficients in the equation<br />

0 =<br />

5<br />

∏ (x − xk) = x<br />

k=2<br />

4 + A3x 3 + A2x 2 + A1x + A0.<br />

To ABEL, the calculations to obtain this were straightforward and not worth mention-<br />

ing. When ABEL factorized the general quintic as<br />

0 =<br />

5<br />

∏<br />

k=1<br />

(x − x k) = (x − x1)<br />

4<br />

∑<br />

k=0<br />

A kx k =<br />

5<br />

∑ akx k=0<br />

k ,<br />

he found that the coefficients A0, . . . , A4 depended rationally on a0, . . . , a5. Conse-<br />

quently, v could also be expressed rationally in x1 and a0, . . . , a5 as<br />

v = t<br />

φ (x1) ,<br />

where both t and φ (x1) were entire functions <strong>of</strong> x1, a0, . . . , a5. By inserting the other<br />

roots x2, . . . , x5 for x1 in φ (x1), ABEL obtained another four entire functions in which<br />

the coefficients were symmetric functions <strong>of</strong> x1, . . . , x5. When ABEL multiplied both<br />

numerator and denominator by ∏ 5 k=2 φ (x k), 35 tacitly used LAGRANGE’S theorem (1)<br />

on resolvents, and reduced the degree according to the relationship imposed by the<br />

quintic equation, he obtained v in the desired form <strong>of</strong> a fourth degree polynomial in<br />

x1.<br />

Five-valued functions in general. In order to obtain a standard form <strong>of</strong> all functions<br />

<strong>of</strong> five quantities having five values, ABEL relied on an extensive investigation <strong>of</strong> par-<br />

ticular cases. Denoting by v any function <strong>of</strong> five quantities, which took on the five<br />

values v1, . . . , v5 when all its arguments were permuted, ABEL introduced an inde-<br />

terminate m and formed the function x m 1 v. When only x2, . . . , x5 were permuted, this<br />

function would attain its values from the list<br />

x m 1 v1, . . . , x m 1 v5. (6.12)<br />

ABEL let µ denote the number <strong>of</strong> different values <strong>of</strong> x m 1 v when x2, . . . , x5 were per-<br />

muted in all possible ways. He then considered the different cases corresponding to<br />

different values <strong>of</strong> µ in detail and either eliminated them through a reductio ad absur-<br />

dum or reduced them to the standard form (6.11). Throughout this procedure, it is<br />

important to keep in mind which quantities are permuted, and ABEL was not always<br />

very explicit.<br />

35 A similar argument resembling multiplying the denominator by its conjugate is described in section<br />

6.3.2.


6.6. Combination into an impossibility pro<strong>of</strong> 115<br />

<strong>The</strong> first case, in which µ = 5, was eliminated, ABEL said, since that assumption<br />

would require all the values (6.12) to be different. Considering transpositions <strong>of</strong> the<br />

form ( 1k<br />

k1 ), ABEL found that xm 1 v would take on another 20 different values, which<br />

would also be distinct from those in (6.12). 36 Thus in total, xm 1 v would take on 25 different<br />

values, and since 25 did not divide 5! = 120 a contradiction had been obtained.<br />

Secondly, ABEL assumed µ = 1 and found that the function v would only take on<br />

one value under all permutations <strong>of</strong> x2, . . . , x5 and thus the case had been reduced to<br />

the one above, giving v in the form (6.11).<br />

Thirdly, for µ = 4, the function x m 1 v would take on the different values xm 1 v1, . . . , x m 1 v4,<br />

and the function v would take on the values v1, . . . , v4 under permutations <strong>of</strong> x2, . . . , x5.<br />

Thus, the function<br />

v1 + v2 + v3 + v4<br />

was a symmetric function <strong>of</strong> x2, . . . , x5, and therefore <strong>of</strong> the form (6.11). Writing v5 as<br />

v5 = (v1 + · · · + v5) − (v1 + · · · + v4) ,<br />

ABEL concluded that the symmetric function v1 + · · · + v5 could be incorporated in<br />

the constant term <strong>of</strong> (6.11), and therefore v5 itself was <strong>of</strong> the form (6.11).<br />

<strong>The</strong>se first three cases were not very difficult to follow. However, the remaining<br />

two cases were subjected to much criticism from his contemporaries (see section 6.9).<br />

In a letter to the Swiss mathematician E. J. KÜLP (⋆1801), 37 who in a private corre-<br />

spondence had asked for clarifications, ABEL described a refined argument, which I<br />

have incorporated in the present description.<br />

<strong>The</strong> fourth case, in which µ = 2, reduced to the well known situation <strong>of</strong> a function<br />

having only two values under permutations. ABEL concluded that since x m 1<br />

v took<br />

on the two values x m 1 v1 and x m 1 v2 under all permutations <strong>of</strong> x1, . . . , x5, the function v<br />

would take on only two values, say v1 and v2, when only x2, . . . , x5 were permuted.<br />

Letting<br />

φ (x1) = v1 + v2, (6.13)<br />

ABEL found that φ (x1) was symmetric under permutations <strong>of</strong> x2, . . . , x5 and thus<br />

<strong>of</strong> the form (6.11). <strong>The</strong> expression φ (x1) had to take on the five different values<br />

φ (x1) , . . . , φ (x5) under all permutations <strong>of</strong> x1, . . . , x5 since only transpositions <strong>of</strong> the<br />

form ( 1k<br />

k1 ) effected the value <strong>of</strong> φ.<br />

36 To see this, it suffices to realize that any permutation σ <strong>of</strong> five quantities can be written as a product<br />

<strong>of</strong> a permutation ˜σ fixing the symbol 1 and a transposition τ <strong>of</strong> the form (1k). <strong>The</strong>n, if an application<br />

<strong>of</strong> σ to v gives vu, it follows that<br />

(x m 1 v) ◦ σ = (xm 1 ◦ ˜σ ◦ τ) (v ◦ σ) = xm ˜σ(k) vu.<br />

37 (<strong>Abel</strong>→Külp, Paris, 1826/11/01. In Hensel, 1903, 237–240).


116 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

In the published paper, ABEL involved himself in a difficult reductio ad absurdum<br />

to rule out this case. However, because the pro<strong>of</strong> given in the letter to KÜLP is more<br />

detailed, it is presented before the differences between the two pro<strong>of</strong>s are sketched.<br />

Besides the symmetric function (6.13), there is another obvious symmetric function<br />

under permutations <strong>of</strong> x2, . . . , x5,<br />

f (x1) = v1v2.<br />

<strong>The</strong> function f (x1) is <strong>of</strong> the form (6.11). ABEL introduced<br />

and found that it must divide<br />

(z − v1) (z − v2) = z 2 − φ (x1) z + f (x1) = R, (6.14)<br />

5<br />

5<br />

∏ (z − vk) = ∑ pkz k=1<br />

k=0<br />

k = R ′ ,<br />

in which p0, . . . , p5 were symmetric functions <strong>of</strong> x1, . . . , x5 by the theorem 1 on LA-<br />

GRANGE resolvents. Since R ′ was unaltered by transpositions ( 1u<br />

u1 ) it followed that all<br />

the polynomials derived from (6.14) through this transposition,<br />

z 2 − φ (xu) z + f (xu) = ρu for 1 ≤ u ≤ 5,<br />

would divide R ′ . However, as R ′ was a polynomial <strong>of</strong> the fifth degree, some polyno-<br />

mials among ρ1, . . . , ρ5 had to share a common factor. Assuming that ρ1 and ρ2 had a<br />

factor in common ABEL concluded<br />

z = f (x1) − f (x2)<br />

φ (x1) − φ (x2) .<br />

This value <strong>of</strong> z must be one <strong>of</strong> the values <strong>of</strong> v and thus the left hand side had five differ-<br />

ent values. However, the right hand side had 10 different values, and a contradiction<br />

had been reached, ruling out the case µ = 2.<br />

<strong>The</strong> published argument in Beweis der Unmöglichkeit 38 followed the one given in<br />

the letter to KÜLP until ABEL had demonstrated that<br />

φ (x1) = v1 + v2 =<br />

4<br />

∑ rkx k=0<br />

k 1<br />

and had recognized that φ had five different values under permutations <strong>of</strong><br />

x1, . . . , x5. Whereas the pro<strong>of</strong> in the letter then explicitly constructed the polynomi-<br />

als R and R ′ , the original argument was much more roundabout. Substituting any one<br />

x k <strong>of</strong> x2, . . . , x5 for x1, ABEL obtained the value φ (x k) as the sum <strong>of</strong> two <strong>of</strong> the five<br />

values <strong>of</strong> v.<br />

38 (N. H. <strong>Abel</strong>, 1826a).


6.6. Combination into an impossibility pro<strong>of</strong> 117<br />

“When x1 is sequentially interchanged with x2, x3, x4, x5 one obtains<br />

v1 + v2 = φ (x1)<br />

v2 + v3 = φ (x2)<br />

.<br />

vm−1 + vm = φ (xm−1)<br />

vm + v1 = φ (xm) ,<br />

where m is one <strong>of</strong> the numbers 2, 3, 4, 5.” 39<br />

<strong>The</strong> m here is not the indeterminate introduced earlier, but a number introduced<br />

for this particular purpose. It is unclear to me, and probably also a point <strong>of</strong> concern<br />

to ABEL’S contemporaries, how this set <strong>of</strong> equations could be put on the circular form<br />

above. But once it had been done (assuming it could be done) it was a simple matter<br />

<strong>of</strong> contradicting the different assumptions for m. If m = 2, it followed that φ (x1) =<br />

φ (x2) and φ could not have five values after all. If m = 3, ABEL deduced that<br />

2v1 = φ (x1) − φ (x2) + φ (x3) ,<br />

whereby a contradiction was reached because the left hand side had 5 values, whereas<br />

the right hand side had 5×4<br />

2 × 3 = 30 values. In a similar way, ABEL claimed he could<br />

prove that m = 4 or m = 5 could be ruled out as well, 40 which in turn proved that µ<br />

could not be equal to 2.<br />

ABEL’S argument presented in the paper depended on a rather obscure sequence<br />

<strong>of</strong> functions and was severely criticized. <strong>The</strong> pro<strong>of</strong> which ABEL gave in his letter to<br />

KÜLP avoided this central step and was much clearer. I conjecture that KÜLP had<br />

questioned the sequence <strong>of</strong> equations, and that ABEL had subsequently developed<br />

his new pro<strong>of</strong> which he presented as an answer; I have no indication that ABEL had<br />

possessed the pro<strong>of</strong> presented to KÜLP when he wrote his paper.<br />

<strong>The</strong> final case, µ = 3, was ruled out in the same way as µ = 2 above. ABEL found<br />

that if µ = 3, the function<br />

v1 + v2 + v3<br />

would be symmetric under permutations <strong>of</strong> x2, . . . , x5 and therefore<br />

v4 + v5 = (v1 + · · · + v5) − (v1 + · · · + v3)<br />

39 “Vertauscht man der Reihe nach x1 mit x2, x3, x4, x5, so erhält man<br />

v1 + v2 = φ (x1)<br />

v2 + v3 = φ (x2)<br />

.<br />

vm−1 + vm = φ (xm−1)<br />

vm + v1 = φ (xm) ,<br />

wo m eine der Zahlen 2, 3, 4, 5 ist.” (ibid., 80).<br />

40 HOLMBOE supplied the expressions (see section 6.9.1).


118 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

could be written in the form (6.11) as he had done in the case µ = 4 above. However,<br />

ABEL had just demonstrated in the case µ = 2 that no sum <strong>of</strong> two values <strong>of</strong> v could<br />

have five values under permutations <strong>of</strong> x1, . . . , x5, whereby he reached a contradiction.<br />

<strong>The</strong> core <strong>of</strong> ABEL’S description <strong>of</strong> functions <strong>of</strong> five quantities having five values<br />

under permutations <strong>of</strong> these consisted <strong>of</strong> two parts:<br />

1. A direct manipulation based on LAGRANGE’S theorem 1 on resolvents, resulting<br />

in a pro<strong>of</strong> that any function <strong>of</strong> five quantities x1, . . . , x5 which is unaltered by per-<br />

mutations <strong>of</strong> four <strong>of</strong> these, x2, . . . , x5, has the form <strong>of</strong> a fourth degree polynomial<br />

(6.11).<br />

2. A meticulous study <strong>of</strong> the particular cases in which any function <strong>of</strong> five quanti-<br />

ties which has five values under permutations <strong>of</strong> x1, . . . , x5 is either contradicted<br />

or proved to be <strong>of</strong> the form (6.11), too.<br />

At the conclusion <strong>of</strong> his investigations, ABEL had added a complete description<br />

<strong>of</strong> functions <strong>of</strong> five quantities having five values to the one he had obtained, in case<br />

the function had only two values. <strong>The</strong>reby, he had obtained workable standard forms<br />

for all non-symmetric rational functions which could be involved in a supposed solu-<br />

tion to the general quintic. All he lacked was to put the pieces together to obtain the<br />

impossibility pro<strong>of</strong>.<br />

6.6.2 <strong>The</strong> goal in sight<br />

To combine his preliminary results into a pro<strong>of</strong> <strong>of</strong> the algebraic insolubility <strong>of</strong> the<br />

general quintic<br />

x 5 + a4x 4 + a3x 3 + a2x 2 + a1x + a0 = 0, (6.15)<br />

ABEL assumed that an algebraic solution could be obtained. <strong>The</strong> auxiliary theorem 3<br />

obtained earlier ensured him that he could assume that any subexpression occurring<br />

therein would be a rational function <strong>of</strong> the roots x1, . . . , x5 <strong>of</strong> the equation (6.15). Since<br />

the quintic could not be solved by a rational expression alone, some root extraction<br />

had to occur. ABEL focused his attention on the algebraic expression <strong>of</strong> the first order<br />

in the supposed solution. Thus, he dissected the solution from the inside by focus-<br />

ing on this innermost non-rational function. According to ABEL’S classification, an<br />

algebraic expression <strong>of</strong> the first order contained only rational functions <strong>of</strong> the coeffi-<br />

cients a0, . . . , a4 and roots <strong>of</strong> the form m√ R where m was a prime and R was a rational<br />

function <strong>of</strong> a0, . . . , a4. Such roots would satisfy the equation<br />

v m − R = 0, (6.16)<br />

and v would have to be a rational function <strong>of</strong> the roots x1, . . . , x5. His earlier results<br />

showed that it was impossible to diminish the degree <strong>of</strong> the equation. <strong>The</strong>refore, the


6.6. Combination into an impossibility pro<strong>of</strong> 119<br />

link between root extractions and permutations ensured him that the function v would<br />

take on m values under all permutations <strong>of</strong> x1, . . . , x5. Since m was a prime and had<br />

to divide 5! by LAGRANGE’S theorem, ABEL argued, the only possibilities were that<br />

m equaled 2, 3, or 5. And since no function <strong>of</strong> five quantities could have three values<br />

under permutations by the CAUCHY-RUFFINI theorem, ABEL ruled out this possibility.<br />

<strong>The</strong> two remaining cases were subsequently both brought to contradictions.<br />

<strong>The</strong> innermost root extraction could not be a fifth root. In the simplest case, corre-<br />

sponding to m = 5, the function v had to have the form <strong>of</strong> a fourth degree polynomial,<br />

as ABEL had demonstrated:<br />

v = 5√ R =<br />

4<br />

∑ rkx k=0<br />

k 1 .<br />

Through a process <strong>of</strong> inversion <strong>of</strong> polynomials in which the quintic equation (6.15) was<br />

used to lower the degree, ABEL found that<br />

x1 =<br />

4<br />

∑ skR k=0<br />

k 5 ,<br />

where s0, . . . , s4 were symmetric functions <strong>of</strong> x1, . . . , x5. Furthermore, by use <strong>of</strong> basic<br />

properties <strong>of</strong> primitive roots <strong>of</strong> unity, he obtained<br />

s1R 1 5 = 1<br />

�<br />

x1 + α<br />

5<br />

4 x2 + α 3 x3 + α 2 �<br />

x4 + αx5 ,<br />

where α was a primitive fifth root <strong>of</strong> unity. <strong>The</strong> left hand side <strong>of</strong> the equation was a so-<br />

lution to a fifth degree equation, and thus had (at most) five different values, whereas<br />

the right hand side was formally altered by any permutation <strong>of</strong> x1, . . . , x5 and thus<br />

had 5! = 120 different values. This ruled out the case m = 5, and the innermost root<br />

extraction could not be a fifth root.<br />

<strong>The</strong> innermost root extraction could not be a square root, neither. ABEL brought<br />

the case m = 2 to a contradiction in a similar way, although it involved studying<br />

expressions <strong>of</strong> the second order as well. He knew that the root would have to be <strong>of</strong><br />

the form<br />

√ R = p + qs,<br />

and the other value under permutations would be<br />

− √ R = p − qs.<br />

Subtracting these two, ABEL concluded that √ R was <strong>of</strong> the form<br />

√ R = qs,


120 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

and he saw that any rational combination <strong>of</strong> such root extractions would continue to<br />

be <strong>of</strong> the same form. <strong>The</strong>refore, any algebraic expression <strong>of</strong> the first order contained<br />

in the solution would have to be <strong>of</strong> the form<br />

α + βs<br />

where α, β were symmetric functions. ABEL observed that such functions were not<br />

powerful enough to solve the general quintic (6.15), and found that such a solution<br />

would necessarily contain root extractions <strong>of</strong> the form<br />

m ′� α + βs,<br />

where m ′ was a prime and β �= 0. ABEL knew that such a root, say v, was a rational<br />

function <strong>of</strong> x1, . . . , x5. Among the values <strong>of</strong> v obtained by permuting x1, . . . , x5 he<br />

found that two were <strong>of</strong> particular interest,<br />

When these two were multiplied,<br />

v1 = m′� α + βs, and<br />

v2 = m′� α − βs.<br />

γ = v1v2 = m′�<br />

α 2 − β 2 s 2 ,<br />

the expression under the root sign was a symmetric function.<br />

At this point, ABEL again considered two individual cases: either γ, itself, was a<br />

symmetric function, or it was not. In case γ was a non-symmetric function, it would be<br />

a first order algebraic expression, and ABEL had proved that for such expressions the<br />

value <strong>of</strong> m ′ would have to be 2. This led to a contradiction, since v then had four values<br />

under permutations <strong>of</strong> x1, . . . , x5 because β �= 0. However, by the CAUCHY-RUFFINI<br />

theorem no such function could exist. Consequently, m′� α 2 − β 2 s 2 would have to be a<br />

symmetric function. By adding v1 and v2, ABEL obtained a function p,<br />

p = v1 + v2 = R 1<br />

m ′ + γ<br />

R R m′ −1<br />

m ′<br />

with R = α + βs. He studied the values <strong>of</strong> p resulting from substituting α k R 1<br />

m ′ for<br />

R 1<br />

m ′ and demonstrated that p had to have m ′ values under permutations <strong>of</strong> x1, . . . , x5.<br />

But since m ′ = 2 had been ruled out, he concluded that m ′ = 5, and the second root<br />

extraction counted from the inside had to be a fifth root. This time ABEL obtained<br />

�<br />

x1 + α 4 x2 + α 3 x3 + α 2 �<br />

x4 + αx5<br />

t1R 1 5 = 1<br />

5<br />

in which t1 was a symmetric function <strong>of</strong> the roots. <strong>The</strong> left hand side was the root <strong>of</strong> an<br />

irreducible equation <strong>of</strong> the tenth degree, 41 thus having 10 values under permutations.<br />

41 With y = t1R 1 5 and R = α + βs, the equation was � y 5 − t 5 1 α� 2 − t 10<br />

1 β2 s 2 = 0 in which the coefficients<br />

are symmetric functions.


6.6. Combination into an impossibility pro<strong>of</strong> 121<br />

<strong>The</strong> right hand side had a complete 120 values since none <strong>of</strong> the roots x1, . . . , x5 could<br />

be interchanged without altering the value <strong>of</strong> the expression. Thus, a contradiction<br />

had again been reached.<br />

<strong>The</strong> line <strong>of</strong> ABEL’S argument in knitting together his preliminary investigations<br />

can be divided into the following steps:<br />

1. <strong>The</strong> innermost root extraction in any supposed solution to the general quintic<br />

had to be either a fifth root (m = 5) or a square root (m = 2); any other possibili-<br />

ties were ruled out by the CAUCHY-RUFFINI theorem.<br />

2. <strong>The</strong> innermost root extraction could not be a fifth root (m �= 5) since this was<br />

brought to a contradiction by comparing the number <strong>of</strong> values <strong>of</strong> certain expres-<br />

sions.<br />

3. Thus, the innermost root extraction had to be a square root (m = 2). <strong>The</strong>n the<br />

second innermost root extraction was taken into consideration. Its degree has<br />

been denoted m ′ .<br />

4. <strong>The</strong> second innermost root extraction, too, had to be <strong>of</strong> degree either two (m ′ =<br />

2) or five (m ′ = 5).<br />

5. In case the second innermost root extraction was a square root, a function having<br />

four values under permutations would be obtained, from which a contradiction<br />

could be reached. Thus m ′ �= 2.<br />

6. <strong>The</strong>refore the second innermost root extraction had to be a fifth root, but this,<br />

too, was brought to a contradiction in a way similar to step 2 above.<br />

7. Consequently, no algebraic solution to the general quintic could exist, and the<br />

algebraic insolubility had been demonstrated.<br />

Apparently, the argument carried out applied to the quintic equation alone. How-<br />

ever, ABEL claimed that it also proved the insolubility <strong>of</strong> all general higher degree<br />

equations.<br />

“From this [the insolubility <strong>of</strong> the general quintic] it follows immediately that<br />

it is also impossible generally to solve equations <strong>of</strong> degrees above the fifth. <strong>The</strong>refore<br />

the equations which can be generally solved are only <strong>of</strong> the four first degrees.” 42<br />

Although he produced no further evidence, ABEL probably thought <strong>of</strong> a pro<strong>of</strong> by<br />

the following argument. If the roots <strong>of</strong> the general sixth degree equation<br />

x 6 + a5x 5 + a4x 4 + a3x 3 + a2x 2 + a1x + a0 = 0<br />

42 “Daraus folgt unmittelbar weiter, daß es ebenfalls unmöglich ist, Gleichungen von höheren als dem<br />

fünften Grade allgemein aufzulösen. Mithin sind die Gleichungen, welche sich algebraisch allgemein<br />

auflösen lassen, nur die von den vier ersten Graden.” (N. H. <strong>Abel</strong>, 1826a, 84).


122 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

could be expressed by any algebraic formula, this formula would also provide the so-<br />

lution to the general fifth degree equation by inserting a0 = 0 in that formula. Central<br />

to the argument is that the supposed general solution formula for sixth degree equa-<br />

tions not only produces a single root, but can somehow be made to produce all the<br />

roots <strong>of</strong> the equation. This was a recurring idea in ABEL’S work on the theory <strong>of</strong> equa-<br />

tions (see for instance theorem 10), which linked the concepts <strong>of</strong> satisfiability (a single<br />

root could be found) and solubility (all roots could be found).<br />

6.7 ABEL and RUFFINI<br />

According to ABEL and his commentators, ABEL was unaware <strong>of</strong> the pro<strong>of</strong>s published<br />

by RUFFINI when he published his pro<strong>of</strong>s <strong>of</strong> the impossibility result in 1824 and 1826. 43<br />

Since questions <strong>of</strong> priority are a frequently recurring theme in the history <strong>of</strong> mathe-<br />

matics, this independence <strong>of</strong> results is noticed by most biographers <strong>of</strong> ABEL. 44 It is my<br />

firm conviction — based on the mathematical contents <strong>of</strong> his pro<strong>of</strong> — that ABEL devel-<br />

oped his pro<strong>of</strong> independently <strong>of</strong> RUFFINI. However, the primary sources <strong>of</strong> informa-<br />

tion on ABEL’S independence <strong>of</strong> RUFFINI are limited. <strong>The</strong> only mention <strong>of</strong> RUFFINI<br />

made by ABEL is in his notebook entry on the theory <strong>of</strong> solubility (see chapter 8), in<br />

the introduction to which he described RUFFINI’S pro<strong>of</strong>:<br />

“<strong>The</strong> first person, and if I am not mistaken, the only one prior to me, who has<br />

tried to prove the impossibility <strong>of</strong> the algebraic solution <strong>of</strong> the general equations,<br />

is the geometer Ruffini; but his memoir is so complicated that it is very difficult<br />

to judge the validity <strong>of</strong> his reasoning. It seems to me that his reasoning is not<br />

always satisfying. I think that the pro<strong>of</strong> I gave in the first issue <strong>of</strong> this journal<br />

[CRELLE’S Journal] leaves nothing to be desired as to rigor, but it does not have all<br />

the simplicity <strong>of</strong> which it is susceptible. I have reached another pro<strong>of</strong> based on the<br />

same principles, but more simple, in trying to solve a more general problem.” 45<br />

<strong>The</strong> answer derived from “trying to solve a more general problem” was never<br />

made available in print, though. As is documented in chapter 8, such an answer was,<br />

indeed, indirectly obtainable from ABEL’S more general research on algebraic solu-<br />

bility which even produced an explicit example <strong>of</strong> a particular special equation which<br />

could not be solved.<br />

43 (N. H. <strong>Abel</strong>, 1824b; N. H. <strong>Abel</strong>, 1826a)<br />

44 See for instance (Bjerknes, 1885, 22–23), (Bjerknes, 1930, 23), (Ore, 1954, 89–90), (Ore, 1957, 125), and<br />

(Stubhaug, 1996, 352–353).<br />

45 “Le premier, et, si je ne me trompe, le seul qui avant moi ait cherché à démontrer l’impossibilité<br />

de la résolution algébrique des équations générales, est le géomètre Ruffini; mais son mémoire est<br />

tellement compliqué qu’il est très difficile de juger de la justesse de son raisonnement. Il me paraît<br />

que son raisonnement n’est pas toujours satisfaisant. Je crois que la démonstration que j’ai donnée<br />

dans le premier cahier de ce journal, ne laisse rien à désirer du côté de la rigueur; mais elle n’a pas<br />

toute la simplicité dont elle est susceptible. Je suis parvenu à une autre démonstration, fondée sur<br />

les mêmes principes, mais plus simple, en cherchant à résoudre un problème plus général.” (N. H.<br />

<strong>Abel</strong>, [1828] 1839, 218).


6.7. ABEL and RUFFINI 123<br />

<strong>The</strong> notebook has been dated to 1828 by P. L. M. SYLOW (1832–1918) — a date<br />

which implies that ABEL disclosed his knowledge <strong>of</strong> RUFFINI only after returning to<br />

Christiania. 46 It is most likely that ABEL learned about RUFFINI during his European<br />

tour, and two instances are <strong>of</strong> particular importance. During his stay in Vienna in<br />

April and May 1826, ABEL became acquainted with the local astronomers K. L. VON<br />

LITTROW (1811–1877) and A. VON BURG (1797–1882). In the first volume <strong>of</strong> their<br />

journal Zeitschrift für Physik und Mathematik, occurring while ABEL was in town, an<br />

anonymous paper on the theory <strong>of</strong> equations was published. 47 <strong>The</strong> author, 48 who was<br />

inspired by ABEL’S pro<strong>of</strong> and praised it highly, reviewed RUFFINI’S pro<strong>of</strong>. <strong>The</strong>refore it<br />

is not unlikely that ABEL learned <strong>of</strong> RUFFINI’S pro<strong>of</strong> from his Viennese connections. 49<br />

Once in Paris, ABEL took on the duty <strong>of</strong> writing unsigned reviews for FERRUSAC’S<br />

Bulletin des sciences mathématiques, astronomiques, physiques et chimiques <strong>of</strong> papers pub-<br />

lished in CRELLE’S Journal für die reine und angewandte Mathematik. We know from<br />

one <strong>of</strong> ABEL’S letters that he, himself, wrote the review <strong>of</strong> his Beweis der Unmöglichkeit<br />

which gave a short exposition <strong>of</strong> the flow <strong>of</strong> the pro<strong>of</strong>. 50 However, appended to the re-<br />

view was a short note by the editor, J. F. SAIGEY (1797–1871), 51 which drew attention<br />

to the works <strong>of</strong> RUFFINI. 52 SAIGEY mentioned CAUCHY’S favorable review <strong>of</strong> RUF-<br />

FINI’S treatise and made it clear that CAUCHY’S view was not universally accepted:<br />

“Other geometers have not understood this demonstration and some have<br />

made the justified remark that by proving too much, Ruffini could not prove anything<br />

in a satisfactory manner; to be sure it was not known how an equation <strong>of</strong><br />

the fifth degree, e.g., could not have transcendental roots, equivalent to infinite series<br />

<strong>of</strong> algebraic terms, since one demonstrates that every equation <strong>of</strong> odd degree<br />

necessarily has some root. By a more pr<strong>of</strong>ound analysis, M. <strong>Abel</strong> proves that such<br />

roots cannot exist algebraically; but he has not solved the question <strong>of</strong> the existence<br />

<strong>of</strong> transcendental roots in the negative.” 53<br />

Thus, at two instances in 1826, ABEL had been in close contact with journals, in<br />

which his result was linked to that <strong>of</strong> RUFFINI. A third possible source <strong>of</strong> information<br />

46 (L. Sylow, 1902, 16).<br />

47 (Anonymous, 1826).<br />

48 Or authors? Unlike the review in FERRUSAC’S Bulletin (see below), ABEL is not likely to be the<br />

author, himself.<br />

49 See (Ore, 1957, 125).<br />

50 (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. N. H. <strong>Abel</strong>, 1902a, 44). <strong>The</strong> paper reviewed is, <strong>of</strong> course, (N.<br />

H. <strong>Abel</strong>, 1826a).<br />

51 (Stubhaug, 1996, 589).<br />

52 (N. H. <strong>Abel</strong>, 1826c, 353–354).<br />

53 “D’autres géomètres avouent n’avoir pas compris cette démonstration, et il y en a qui ont fait la<br />

remarque très-juste que Ruffini en prouvant trop pourrait n’avoir rien prouvé d’une manière satisfaisante;<br />

en effet, on ne conçoit pas comment une équation du cinquième degré, par exemple,<br />

n’admettrait pas de racines transcendantes, qui équivalent à des séries infinies de termes algébriques,<br />

puisqu’on démontre que toute équation de degré impair a nécessairement une racine quelconque.<br />

M. <strong>Abel</strong>, au moyen d’une analyse plus pr<strong>of</strong>onde, vient de prouver que de telles racines ne<br />

peuvent exister algébriquement; mais il n’a pas résolu négativement la question de l’existence des<br />

racines transcendantes.” (Saigey in ibid., 354).


124 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

on RUFFINI’S research was, <strong>of</strong> course, CAUCHY whom ABEL met in Paris without any<br />

traceable interaction taking place. 54<br />

Although the primary information on how ABEL came to know <strong>of</strong> RUFFINI’S pro<strong>of</strong>s<br />

is rather sparse, I find further support for the assumption <strong>of</strong> independence in the<br />

mathematical technicalities as documented in the preceding sections. RUFFINI’S and<br />

ABEL’S differences in notation and approach to permutations, ABEL’S definition <strong>of</strong><br />

algebraic expressions and his careful pro<strong>of</strong> <strong>of</strong> the auxiliary theorems describing them<br />

all suggest to me that ABEL’S deduction was a tailored argument for the impossibility,<br />

independent <strong>of</strong> any earlier such pro<strong>of</strong>s. <strong>The</strong> common inspiration from LAGRANGE,<br />

which permeated both the works <strong>of</strong> RUFFINI and ABEL, should be evident enough to<br />

account for similarities in studying the blend <strong>of</strong> equations and permutations.<br />

6.8 Limiting the class <strong>of</strong> solvable equations<br />

At a conceptual level, ABEL’S pro<strong>of</strong> that the general quintic could not be solved al-<br />

gebraically was more than just another pro<strong>of</strong> added to the body <strong>of</strong> mathematics. For<br />

centuries, mathematical intuition had suggested that an algebraic solution to the fifth<br />

degree equation should exist but probably be difficult to find. ABEL had demonstrated<br />

that any supposed solution carried with it an internal contradiction and thus the re-<br />

sult not only made the belief in general algebraic solubility tremble, it completely<br />

destroyed it.<br />

In negating the existence <strong>of</strong> an algebraic solution, ABEL provided an instance <strong>of</strong><br />

a negative result — negative in the sense that it contradicted contemporary intuition.<br />

Similar counter intuitive results abound in mathematics in the period as a result <strong>of</strong> a<br />

fundamental transition toward concept based mathematics. 55<br />

<strong>The</strong> outspoken reactions <strong>of</strong> the mathematical community to ABEL’S impossibility<br />

pro<strong>of</strong>s can be divided in three. Some mathematicians, <strong>of</strong>ten belonging to the older<br />

generation or the loosely institutionalized amateur mathematicians, protested against<br />

the result and held both the statements and their pro<strong>of</strong>s to be flawed. To these math-<br />

ematicians, the break with their established intuition forced them into their rejection.<br />

Others accepted the result but provided refinements <strong>of</strong> the pro<strong>of</strong>s and their assump-<br />

tions. And yet others not only accepted the results but saw that the quintic only con-<br />

stituted one example <strong>of</strong> an unsolvable equation. <strong>The</strong>reby, the more general problem<br />

<strong>of</strong> algebraic solubility could be formulated.<br />

From the perspective <strong>of</strong> investigating the concept <strong>of</strong> solubility, the quintic helped<br />

distinguish the class <strong>of</strong> algebraically solvable equations within the class <strong>of</strong> all polyno-<br />

mial equations (see figure 6.1). On the other hand, in his research on the division <strong>of</strong> the<br />

circle, GAUSS had demonstrated that infinitely many equations existed which could<br />

54 For a discussion <strong>of</strong> the relationship between CAUCHY and ABEL, see chapter 12.<br />

55 See chapter 21.


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 125<br />

❅ ❅❘<br />

❄<br />

✲ Solvable equations ✛<br />

� ✻<br />

Polynomial equations<br />

�✒<br />

Figure 6.1: Limiting the class <strong>of</strong> solvable equations<br />

be solved algebraically (see section 5.3). <strong>The</strong>refore, the class <strong>of</strong> solvable equations<br />

did not collapse to a few low degree equations; soon, further solvable equations were<br />

found (see chapter 7). <strong>The</strong> search for a procedure useful in determining whether or<br />

not a given equation could be solved algebraically soon became an interesting project<br />

for mathematical research.<br />

In the following section 6.9, I deal with the first two classes <strong>of</strong> reactions: the global<br />

and local criticism, which was advanced by ABEL’S contemporaries. In chapter 8, I<br />

describe how ABEL worked on the general problem <strong>of</strong> solubility, which was realized<br />

to its full extent and attacked shortly afterwards by GALOIS (section 8.5).<br />

6.9 <strong>The</strong> reception <strong>of</strong> ABEL’s work on the quintic<br />

When RUFFINI published his pro<strong>of</strong> <strong>of</strong> the algebraic insolubility <strong>of</strong> the quintic in Ital-<br />

ian in 1799 the mathematical community <strong>of</strong> Europe paid little attention. Apart from a<br />

limited Italian discussion involving mathematicians outside the main stream such as<br />

P. ABBATI (1768–1842) and G. F. MALFATTI (1731–1807), only CAUCHY seems to have<br />

taken notice. Twenty-five years later, when ABEL published his pro<strong>of</strong> in a brand new<br />

German mathematical journal, history could have repeated itself. However, ABEL’S<br />

pro<strong>of</strong> soon became widespread knowledge and acquired a status within the math-<br />

ematical community <strong>of</strong> being rigorous and close to definitive. In this section, I trace<br />

some <strong>of</strong> the events which played a role in this development, scientific and non-scientific<br />

factors, in order to describe the influence which ABEL’S research had on the subse-<br />

quent evolution <strong>of</strong> the theory <strong>of</strong> equations.<br />

��✠<br />

❅❅■


126 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

Immediate reception. In his short lifetime, ABEL’S impossibility pro<strong>of</strong> was pub-<br />

lished on three occasions. ABEL’S first pro<strong>of</strong> — published as a pamphlet in 1824 —<br />

was, although written in French, only sparsely circulated. 56 According to HANSTEEN<br />

the copy which ABEL had sent to GAUSS in Göttingen was received with very little<br />

enthusiasm. 57 When ABEL met CRELLE in Berlin, they discussed the subject and be-<br />

cause CRELLE and others had a hard time following the arguments, ABEL elaborated<br />

his pro<strong>of</strong>. Under this procedure to clarify, ABEL produced his second pro<strong>of</strong> which<br />

CRELLE subsequently found worthy <strong>of</strong> publication, translated into German, and in-<br />

serted into the first issue <strong>of</strong> his Journal für die reine und angewandte Mathematik. 58 <strong>The</strong><br />

immediate impact <strong>of</strong> ABEL’S paper in the mathematical community was limited. <strong>The</strong><br />

following issue <strong>of</strong> CRELLE’S Journal carried a paper by the unknown mathematician L.<br />

OLIVIER on the form <strong>of</strong> roots <strong>of</strong> algebraic equations based on LAGRANGE’S research. 59<br />

In it, OLIVIER voiced reservations concerning the solubility <strong>of</strong> the general equations<br />

which indicate that he was not an intimate member <strong>of</strong> the circle around CRELLE and<br />

had not learned <strong>of</strong> ABEL’S result prior to its publication in the Journal für die reine und<br />

angewandte Mathematik. <strong>The</strong> reservation might even have been inserted by CRELLE<br />

who probably also translated OLIVIER’S paper into German.<br />

“Furthermore, a pro<strong>of</strong> <strong>of</strong> the impossibility <strong>of</strong> the solution <strong>of</strong> algebraic equations<br />

<strong>of</strong> higher degrees by radicals, should such a pro<strong>of</strong> be possible, would in no<br />

way contradict the results <strong>of</strong> the above investigations on the form <strong>of</strong> the radicals,<br />

whose generality has been claimed.” 60<br />

ABEL tried to improve the distribution <strong>of</strong> his pro<strong>of</strong> as well as the reputation <strong>of</strong><br />

CRELLE’S Journal für die reine und angewandte Mathematik by publishing — as a report<br />

on the paper in CRELLE’S Journal für die reine und angewandte Mathematik — a third<br />

version <strong>of</strong> his pro<strong>of</strong> in FERRUSAC’S Bulletin. 61 However, in his own lifetime, ABEL<br />

was mainly known for his later work on elliptic functions (see subsequent chapters)<br />

and the impossibility pro<strong>of</strong> remained less known. In corresponding with ABEL on the<br />

subject <strong>of</strong> elliptic functions, A.-M. LEGENDRE (1752–1833) urged ABEL to make public<br />

his researches on the solubility <strong>of</strong> equations which ABEL had announced in an earlier<br />

letter. 62<br />

“Sir, you have announced a very beautiful work on algebraic solutions which<br />

has the purpose <strong>of</strong> giving the solution <strong>of</strong> any given numerical equation, whenever<br />

56 (N. H. <strong>Abel</strong>, 1824b)<br />

57 See (Hansteen, 1862, 37) and (Stubhaug, 1996, 291).<br />

58 (N. H. <strong>Abel</strong>, 1826a)<br />

59 Of Mr. LOUIS OLIVIER little is known. He published a total <strong>of</strong> 11 articles in the first four volumes<br />

<strong>of</strong> CRELLE’S Journal für die reine und angewandte Mathematik 1826–1829. OLIVIER’S mathematics and<br />

his relations to the Berlin mathematicians are the themes <strong>of</strong> a separate article under preparation.<br />

60 “Uebrigens würde ein Beweis der Unmöglichkeit der Auflösung höheren algebraischer Gleichungen<br />

durch Wurzelgrößen, wenn ein solcher gelänge, keinesweges den Resultaten der obigen Untersuchungen<br />

über die Form der Wurzeln, deren Allgemeinheit behauptet wurde, widersprechen.”<br />

(Olivier, 1826a, 116).<br />

61 (N. H. <strong>Abel</strong>, 1826c).<br />

62 (<strong>Abel</strong>→Legendre, Christiania, 1828/11/25. N. H. <strong>Abel</strong>, 1881, 279)


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 127<br />

it can be developed in radicals, and to declare any equation unsolvable in this way<br />

[by radicals] which does not satisfy the required conditions; from this it follows<br />

as a necessary consequence that the general solution <strong>of</strong> the equations beyond the<br />

fourth degree is impossible. I invite you to publish this new theory as soon as you<br />

can; it would bring you much honour and generally be regarded as the biggest<br />

discovery remaining to be made in analysis.” 63<br />

<strong>The</strong> investigations to which LEGENDRE alluded were also described in one <strong>of</strong> ABEL’S<br />

letters to HOLMBOE, 64 and were partially presented in his notebooks (see chapter 8).<br />

<strong>The</strong> above citation seems also to indicate that LEGENDRE was unaware <strong>of</strong> ABEL’S impossibility<br />

pro<strong>of</strong> in CRELLE’S Journal für die reine und angewandte Mathematik 1826. 65<br />

HOLMBOE and the first edition <strong>of</strong> ABEL’S Œuvres. When the French mathematical<br />

community learned <strong>of</strong> ABEL’S death in 1829, the Academy sent baron J. F. T. MAU-<br />

RICE. 66 to condole with the Swedish envoy in Paris and suggest that the Swedish<br />

Crown Prince OSCAR undertook the publication <strong>of</strong> ABEL’S complete works. 67 In 1831,<br />

MAURICE repeated his suggestion and the editorship was delegated to ABEL’S teacher<br />

and friend HOLMBOE and the university in Christiania. 68 By 1836, HOLMBOE had<br />

completed his commentaries on the published works, but intended to include also<br />

selections from ABEL’S unpublished material in the Œuvres. 69 In his report to the<br />

Ministry <strong>of</strong> Ecclesiastic Affairs 70 in 1838 HOLMBOE declared that — except for a ma-<br />

nuscript which ABEL had handed in to the French Academy 71 — he had finished col-<br />

lecting and commenting upon ABEL’S unpublished works. 72 Two volumes containing<br />

most <strong>of</strong> ABEL’S published works (with the noticeable exclusion <strong>of</strong> ABEL’S Parisian<br />

manuscript) 73 and some <strong>of</strong> the unpublished material from his notebooks and Nach-<br />

lass appeared in 1839. Since most <strong>of</strong> ABEL’S papers had originally been published<br />

in French and most <strong>of</strong> his mature entries in the notebooks were in French, it had been<br />

decided that the Œuvres should be in French. An effort was made by HOLMBOE to dis-<br />

63 “Vous m’annoncez, Monsieur, un très beau travail sur les équations algébriques, qui a pour objet<br />

de donner la résolution de toute équation numérique proposée, lorsqu’elle peut être développée en<br />

radicaux, et de déclarer insoluble sous ce rapport, toute équation qui ne satisferait pas aux conditions<br />

exigées; d’où résulte comme conséquence nécessaire que la résolution générale des équations<br />

au delà du quatrième degré, est impossible. Je vous invite à publier le plutôt que vouz pourrez,<br />

cette nouvelle théorie; elle vous fera beaucoup d’honneur, et sera généralement regardée comme<br />

la plus grande dévouverte qui restait à faire dans l’analyse.” (Legendre→<strong>Abel</strong>, Paris, 1829/01/16.<br />

N. H. <strong>Abel</strong>, 1902a, 88–89).<br />

64 (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. ibid., 44–45).<br />

65 (Holmboe, 1829, 349).<br />

66 (Stubhaug, 1996, 587).<br />

67 After a turbulent period <strong>of</strong> trembling Danish monarchy and brief independence, Norway was integrated<br />

in the Swedish monarchy 1814.<br />

68 (Ore, 1957, 269)<br />

69 (N. H. <strong>Abel</strong>, 1902d, 49).<br />

70 As noted in chapter 2, the University was subsumed in the Ministry <strong>of</strong> Ecclesiastic Affairs.<br />

71 <strong>The</strong> search for ABEL’S Paris mémoire is a fascinating story in its own right. See (Brun, 1949; Brun,<br />

1953) and section 19.4, below.<br />

72 (N. H. <strong>Abel</strong>, 1902d, 51).<br />

73 ABEL’S Parisian manuscript known as the Paris mémoire is dealt with extensively in chapter 19.


128 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

tribute copies to prominent mathematicians. <strong>The</strong>refore, ten years after ABEL’S death,<br />

the mathematical community had the opportunity to follow his arguments through<br />

HOLMBOE’S careful annotations.<br />

“During the revision <strong>of</strong> <strong>Abel</strong>’s works it has been necessary for me to give numerous<br />

demonstrations and prove many theorems which are presented without<br />

pro<strong>of</strong> by the author or whose pro<strong>of</strong> is indicated so briefly that for many readers it<br />

is impossible and for almost all difficult to understand.” 74<br />

Apart from the elaborations <strong>of</strong> vaguely suggested arguments, HOLMBOE also cor-<br />

rected most <strong>of</strong> the numerous misprints which had occurred in ABEL’S works pub-<br />

lished in CRELLE’S Journal. <strong>The</strong>se, too, had served to make ABEL’S writings hard to<br />

understand. 75<br />

6.9.1 Local criticism <strong>of</strong> the quintic pro<strong>of</strong><br />

Inspired by I. LAKATOS’ (1922–1974) distinction between global and local counter ex-<br />

amples, the criticism which mathematicians in the first half <strong>of</strong> the 19 th century ex-<br />

pressed toward ABEL’S pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the quintic can be separated in two<br />

classes. 76 As noted, a handful <strong>of</strong> mathematicians continued to doubt or dispute the va-<br />

lidity <strong>of</strong> the result that the general quintic was algebraically unsolvable. <strong>The</strong>ir doubt<br />

was largely founded in an incomplete induction that equations were to be solvable;<br />

and their attitude toward ABEL’S pro<strong>of</strong> ranged from ignorance to indifference. <strong>The</strong><br />

importance <strong>of</strong> this global criticism is traced in section 6.9.2. On the other hand, ABEL’S<br />

pro<strong>of</strong> was scrutinized by some <strong>of</strong> his contemporaries. <strong>The</strong>ir local criticism picked out<br />

the vulnerable points <strong>of</strong> ABEL’S argument and sought to illuminate them or supply<br />

alternative pro<strong>of</strong>s. Central to these local criticisms was the fact that they were based<br />

on an acceptance <strong>of</strong> the overall validity <strong>of</strong> the result but sought to secure some unclear<br />

arguments. <strong>The</strong> central parts <strong>of</strong> ABEL’S argument which was thought in need <strong>of</strong> elab-<br />

oration were the classification <strong>of</strong> algebraic expressions, ABEL’S pro<strong>of</strong> <strong>of</strong> the CAUCHY-<br />

RUFFINI theorem, and in particular ABEL’S study <strong>of</strong> functions <strong>of</strong> five quantities having<br />

five values under permutations.<br />

EDMUND JACOB KÜLP. From ABEL’S correspondence with EDMUND JACOB KÜLP<br />

only ABEL’S reply to KÜLP’S questions has been preserved. 77 <strong>The</strong>refore, we know<br />

74 “Under Revisionen af <strong>Abel</strong>s Arbeider har det været mig nødvendigt at optegne en heel Deel Udviklinger<br />

og at bevise mange Sætninger, som hos Forfatteren ere anførte uden Beviis, eller hvis Beviis er<br />

saa kort antydet, at det for mange Læsere er umueligt og næsten for alle vanskeligt at fatte.” (N. H.<br />

<strong>Abel</strong>, 1902d, 50).<br />

75 (ibid., 49).<br />

76 <strong>The</strong> distinction is inspired by (Lakatos, 1976), However, as discussed in the introduction (section<br />

1.4.2), LAKATOS’ scheme <strong>of</strong> mathematical evolution by dialectical dynamics can only be applied<br />

through largely a-historical reconstructions.<br />

77 (<strong>Abel</strong>→Külp, Paris, 1826/11/01. In Hensel, 1903, 237–240)


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 129<br />

Figure 6.2: WILLIAM ROWAN HAMILTON (1805–1865)<br />

nothing <strong>of</strong> KÜLP’S attitude toward the validity <strong>of</strong> ABEL’S result. KÜLP’S criticism fo-<br />

cused on two individual parts <strong>of</strong> ABEL’S argument. <strong>The</strong> first question was concerned<br />

with a misprint which occurred in ABEL’S pro<strong>of</strong> <strong>of</strong> the CAUCHY-RUFFINI theorem.<br />

Due to the relatively new character <strong>of</strong> the theory <strong>of</strong> permutations and their notation,<br />

KÜLP apparently had trouble following ABEL’S argument and was halted by the mis-<br />

print. ABEL’S notation was apparently also a problem for KÜLP; in his answer, ABEL<br />

proved the claim that any 3-cycle could be decomposed as the product <strong>of</strong> two p-cycles<br />

by writing out the substitutions in detail. I mention these objections in order to illus-<br />

trate the difficulties, conceptual and technical, which nineteenth century mathemati-<br />

cians had in understanding and accepting ABEL’S pro<strong>of</strong>.<br />

KÜLP’S other objection concerned ABEL’S descriptive classification <strong>of</strong> rational func-<br />

tions <strong>of</strong> five quantities which have five values. Again, we do not have KÜLP’S formu-<br />

lation but only ABEL’S reply which ABEL posted from Paris less than a year after his<br />

paper had appeared in CRELLE’S Journal. <strong>The</strong> argument given in the letter differed<br />

substantially from the published one. As I have discussed above (in section 6.6.1),<br />

the original argument was, indeed, very hard to understand. If ABEL’S refined pro<strong>of</strong><br />

communicated to KÜLP had made it into print, ABEL’S conclusion might have been<br />

accepted at an earlier point.


130 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

WILLIAM ROWAN HAMILTON. On the British Isles, the debate over the solubility<br />

<strong>of</strong> the general quintic took a different turn. During the years 1832–35, G. B. JER-<br />

RARD (1804–1863) published his three volume work Mathematical Researches in which<br />

he claimed to have presented a general method <strong>of</strong> solving equations algebraically.<br />

At the 1835 meeting <strong>of</strong> the British Association in Dublin, WILLIAM ROWAN HAMIL-<br />

TON was appointed reporter on the paper and was thus led into the theory <strong>of</strong> equa-<br />

tions. 78 In May 1836, after having dismissed JERRARD’S claim for a general solution to<br />

higher degree equations in a paper in the Philosophical Magazine, 79 HAMILTON asked<br />

his friend J. W. LUBBOCK (1803–1865) to supply him with a copy <strong>of</strong> ABEL’S paper from<br />

CRELLE’S Journal für die reine und angewandte Mathematik. Approaching it in his very<br />

thorough and critical style, HAMILTON found it somewhat unsatisfactory, and began<br />

to write his own exposition <strong>of</strong> ABEL’S result. 80 <strong>The</strong> following year he presented his<br />

investigations to the Royal Irish Academy, in whose Transactions they were printed. 81<br />

In his study <strong>of</strong> ABEL’S pro<strong>of</strong>, HAMILTON noticed two “mistakes”, the first <strong>of</strong> which<br />

concerned ABEL’S classification <strong>of</strong> algebraic expressions (see theorem 2). After hav-<br />

ing translated ABEL’S pro<strong>of</strong> into his own notation, HAMILTON clearly expressed his<br />

objection: 82<br />

“Although the whole <strong>of</strong> the foregoing argument has been suggested by that<br />

<strong>of</strong> <strong>Abel</strong>, and may be said to be a commentary thereon; yet it will not fail to be<br />

perceived, that there are several considerable differences between the one method<br />

<strong>of</strong> pro<strong>of</strong> and the other. More particularly, in establishing the cardinal proposition<br />

that every radical in every irreducible expression for any one <strong>of</strong> the roots <strong>of</strong> any<br />

general equation is a rational function <strong>of</strong> those roots, it has appeared to the writer<br />

<strong>of</strong> this paper more satisfactory to begin by showing that the radicals <strong>of</strong> highest<br />

order will have that property, if those <strong>of</strong> lower orders have it, descending thus to<br />

radicals <strong>of</strong> the lowest order, and afterwards ascending again; than to attempt, as<br />

<strong>Abel</strong> has done, to prove the theorem, in the first instance, for radicals <strong>of</strong> the highest<br />

order. In fact, while following this last-mentioned method, <strong>Abel</strong> has been led to<br />

assume that the coefficient <strong>of</strong> the first power <strong>of</strong> some highest radical can always<br />

be rendered equal to unity, by introducing (generally) a new radical, which in the<br />

notation <strong>of</strong> the present paper may be expressed as follows:<br />

�<br />

�<br />

�⎧<br />

�⎨<br />

�<br />

�<br />

α (m)<br />

k<br />

⎩ ∑<br />

β (m)<br />

i


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 131<br />

is not, in general, free from the irrationalities <strong>of</strong> the same order introduced by the<br />

other radicals a (m)<br />

1 , . . . <strong>of</strong> that order; and consequently the new radical, to which<br />

this process conducts, is in general elevated to the order m + 1; a circumstance<br />

which <strong>Abel</strong> does not appear to have remarked, and which renders it difficult to<br />

judge <strong>of</strong> the validity <strong>of</strong> his subsequent reasoning.” 83<br />

To HAMILTON, the mistake made by ABEL had obscured the validity <strong>of</strong> ABEL’S<br />

subsequent reasoning, but the validity <strong>of</strong> the impossibility result, itself, was not ques-<br />

tioned since HAMILTON had provided it with a pro<strong>of</strong> not based on ABEL’S hierarchy.<br />

Later, KÖNIGSBERGER would prove that ABEL’S hierarchy <strong>of</strong> algebraic expressions<br />

could still be rescued (see below). By the end <strong>of</strong> the century, it was eventually realized<br />

that the hierarchic structure imposed on algebraic expressions was actually superfluous<br />

for the impossibility pro<strong>of</strong>. 84<br />

HAMILTON continued his scrutiny <strong>of</strong> ABEL’S pro<strong>of</strong> by attacking ABEL’S character-<br />

ization <strong>of</strong> functions <strong>of</strong> five quantities having five values under permutations:<br />

“And because the other chief obscurity in <strong>Abel</strong>’s argument (in the opinion <strong>of</strong><br />

the present writer) is connected with the pro<strong>of</strong> <strong>of</strong> the theorem, that a rational function<br />

<strong>of</strong> five independent variables cannot have five values and five only, unless it<br />

be symmetric relatively to four <strong>of</strong> its five elements; it has been thought advantageous,<br />

in this paper, as preliminary to the discussion <strong>of</strong> the forms <strong>of</strong> functions <strong>of</strong><br />

five arbitrary quantities, to establish certain auxiliary theorems respecting functions<br />

<strong>of</strong> fewer variables; which have served also to determine à priori all possible<br />

solutions (by radicals and rational functions) <strong>of</strong> all general algebraic equations<br />

below the fifth degree.” 85<br />

Thus, HAMILTON pointed his finger directly at the two weak points <strong>of</strong> ABEL’S ar-<br />

gument. For ABEL’S flawed pro<strong>of</strong> <strong>of</strong> the central auxiliary theorem — that all occurring<br />

radicals were rational functions <strong>of</strong> the roots — which he had proved by the hierar-<br />

chic structure <strong>of</strong> algebraic expressions, HAMILTON substituted an argument descend-<br />

ing and re-ascending the hierarchy <strong>of</strong> algebraic expressions. 86 <strong>The</strong> characterization<br />

<strong>of</strong> functions <strong>of</strong> five variables having five values under permutations was also car-<br />

ried out at length in an analysis which — following ABEL — reduced it to the study <strong>of</strong><br />

such functions when only four <strong>of</strong> the arguments were permuted. As ABEL had done,<br />

HAMILTON completed his analysis <strong>of</strong> these functions through an extensive investigation<br />

<strong>of</strong> particular classes. 87<br />

HAMILTON employed a detailed style <strong>of</strong> presentation and extensive use <strong>of</strong> low<br />

degree equations as examples; nevertheless, his exposition <strong>of</strong> ABEL’S result is not par-<br />

ticularly clear and easy to grasp. 88 <strong>The</strong> degree <strong>of</strong> detail and a complicated notation<br />

might also have obscured the main results to some <strong>of</strong> HAMILTON’S contemporaries.<br />

83 (ibid., 248); small-caps changed into italic..<br />

84 (J. Pierpont, 1896, 200).<br />

85 (W. R. Hamilton, 1839, 248–249); small-caps changed into italic..<br />

86 (ibid., 194–196).<br />

87 (ibid., 237–246).<br />

88 (Dickson, 1959, 179) calls it “a very complicated reconstruction <strong>of</strong> ABEL’S pro<strong>of</strong>”.


132 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

Neither HAMILTON’S exposition <strong>of</strong> ABEL’S pro<strong>of</strong> nor his more direct criticisms <strong>of</strong> JER-<br />

RARD’S works seemed to convince JERRARD <strong>of</strong> his mistake. 89 JERRARD continued to<br />

announce his claim in the Philosophical Magazine and in 1858 he published his Essay<br />

on the Resolution <strong>of</strong> Equations. By that time it was left to A. CAYLEY (1821–1895) and J.<br />

COCKLE to refute JERRARD’S claims. 90<br />

BERNT MICHAEL HOLMBOE. <strong>The</strong> French mathematical community mainly knew<br />

<strong>of</strong> ABEL’S work on the solubility <strong>of</strong> equations through BERNT MICHAEL HOLMBOE’S<br />

edition <strong>of</strong> ABEL’S collected works (see above). 91 HOLMBOE’S extensive annotations<br />

and elaborations were <strong>of</strong>ten supplying explicit calculations in places where ABEL had<br />

been brief. In terms <strong>of</strong> criticism and modification <strong>of</strong> ABEL’S pro<strong>of</strong>, HOLMBOE’S anno-<br />

tations center on three topics: irreducibility, functions <strong>of</strong> five quantities, and an explicit<br />

description <strong>of</strong> the process <strong>of</strong> inversion which ABEL had employed (see page 119).<br />

HOLMBOE opened with a short treatment <strong>of</strong> reducible and irreducible equations,<br />

in which he gave examples. He explicitly termed an equation irreducible when no<br />

root <strong>of</strong> the equation could be the root <strong>of</strong> an equation <strong>of</strong> “the same form”, but <strong>of</strong> lower<br />

degree. 92 This definition was implicit in ABEL’S paper; 93 it later took on a more explicit<br />

and very central role in ABEL’S theory <strong>of</strong> solubility (see chapters 7 and 8).<br />

Concerning ABEL’S investigations <strong>of</strong> functions <strong>of</strong> five quantities with five values,<br />

HOLMBOE’S annotations are <strong>of</strong> another character giving alternative pro<strong>of</strong>s <strong>of</strong> unclear<br />

points. Remaining faithful to ABEL’S approach in the case in which µ = 2 (see page<br />

115), HOLMBOE supplied expressions with 30 and 10 different values to rule out the<br />

cases m = 4 and m = 5 which ABEL had left out. Thus, HOLMBOE sought to complete<br />

ABEL’S deduction <strong>of</strong> a contradiction. But sensing the obscure nature <strong>of</strong> ABEL’S classi-<br />

fication <strong>of</strong> functions <strong>of</strong> five quantities with five values, HOLMBOE set out to derive his<br />

own. 94 HOLMBOE applied a general theorem, which he had proved in the Magazin for<br />

Naturvidenskaberne:<br />

“In the same way one can demonstrate that if u designates a given function<br />

<strong>of</strong> n quantities which takes on m different values when one interchanges these n<br />

quantities among themselves in all possible ways, the general form <strong>of</strong> the function<br />

<strong>of</strong> n quantities which by these mutual permutations can obtain m different values<br />

will be<br />

r0 + r1u + r2u 2 + · · · + rm−1u m−1 ,<br />

r0, r1, r2 . . . rm−1 being symmetric functions <strong>of</strong> the n quantities.” 95<br />

89 Actually, HAMILTON thought highly <strong>of</strong> JERRARD’S results, which he interpreted in a restricted<br />

frame. Although JERRARD’S claim for solving general equations could not be supported, the method<br />

which he had employed was nevertheless <strong>of</strong> great importance since it — if applied to the quintic —<br />

could reduce it to the normal trinomial form x 5 + px + q = 0.<br />

90 For instance (Cayley, 1861; Cockle, 1862; Cockle, 1863).<br />

91 (N. H. <strong>Abel</strong>, 1839).<br />

92 (Holmboe in ibid., 409).<br />

93 (N. H. <strong>Abel</strong>, 1826a, 71, 82). See quotation on page 106.<br />

94 (Holmboe in N. H. <strong>Abel</strong>, 1839, 411–413).


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 133<br />

HOLMBOE’S pro<strong>of</strong> implicitly involved LAGRANGE’S notion <strong>of</strong> semblables functions<br />

(functions which are altered in the same way by the same permutations), and argued<br />

directly that any function <strong>of</strong> five quantities, which took on five different values, must<br />

have the form <strong>of</strong> a fourth degree polynomial in which the coefficients were symmetric<br />

functions <strong>of</strong> x1, . . . , x5.<br />

<strong>The</strong> final noteworthy contribution by HOLMBOE to ABEL’S impossibility pro<strong>of</strong> was<br />

his calculations relating to the process described as inversion <strong>of</strong> polynomials. HOLM-<br />

BOE proved — through manipulations on power sums — that any fourth degree poly-<br />

nomial v in x<br />

could be inverted into<br />

v =<br />

x =<br />

4<br />

∑ rαx<br />

α=0<br />

α<br />

4<br />

∑ sαv<br />

α=0<br />

α . 96<br />

<strong>The</strong> pro<strong>of</strong> is a tour de force dealing with symmetric functions, much in the style <strong>of</strong> E.<br />

WARING (∼1736–1798), although in a clearer notational setting.<br />

In his commentary, HOLMBOE did not penetrate to the core <strong>of</strong> the problems spotted<br />

by HAMILTON. Instead, he elaborated many <strong>of</strong> ABEL’S arguments and manipulations<br />

and supplied pro<strong>of</strong>s <strong>of</strong> obscure passages. HOLMBOE’S only real reservation against<br />

ABEL’S pro<strong>of</strong> concerned the classification <strong>of</strong> functions with five values, and HOLM-<br />

BOE provided an alternative deduction using methods and concepts introduced by<br />

LAGRANGE and quite familiar to ABEL.<br />

KÖNIGSBERGER. While HOLMBOE’S elaboration <strong>of</strong> ABEL’S classification <strong>of</strong> func-<br />

tions with five quantities might have settled HAMILTON’S unease on this objection,<br />

it took longer before HAMILTON’S other reservation was lifted. <strong>The</strong> objection raised<br />

against ABEL’S classification <strong>of</strong> algebraic expressions was lifted in two steps: In 1869,<br />

KÖNIGSBERGER demonstrated how ABEL’S classification could be rescued by modi-<br />

fying the claims concerning the orders and degrees <strong>of</strong> the coefficients in the represen-<br />

tation<br />

v = q0 + p 1 n + q2p 2 n + · · · + qn−1p n−1<br />

n .<br />

(see page 104). 97 <strong>The</strong> slight modification which KÖNIGSBERGER introduced revali-<br />

dated ABEL’S hierarchy on algebraic expressions, and showed that ABEL’S “mistake”<br />

was <strong>of</strong> no real consequence to the pro<strong>of</strong>. KÖNIGSBERGER had been stimulated to make<br />

95 “De la même manière on peut démontrer que, si u signifie une fonction donnée de n quantités qui<br />

prend m valeurs différentes lorsqu’on échange ces n quantités entre elles de toutes les manières<br />

possibles, la forme générale de la fonction de n quantités qui par leurs permutations mutuelles peut<br />

obtenir m valuers différentes sera<br />

r0 + r1u + r2u 2 + · · · + rm−1u m−1 ,<br />

r0, r1, r2 . . . rm−1 étant des fonctions symétriques des n quantités.” (Holmboe in ibid., 413).<br />

97 (Königsberger, 1869).


134 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

his remedy public by the fact that such a classification was <strong>of</strong> importance by itself and<br />

the circumstance that ABEL’S flawed classification had been reproduced in J. A. SER-<br />

RET’S (1819–1885) textbook Cours d’algèbre. 98<br />

<strong>The</strong> two central arguments in ABEL’S pro<strong>of</strong>, to which HAMILTON raised his objec-<br />

tions, had also stimulated other mathematicians to give alternative pro<strong>of</strong>s and mod-<br />

ifications <strong>of</strong> ABEL’S deduction. <strong>The</strong> defective classification <strong>of</strong> algebraic expressions,<br />

which for ABEL served to demonstrate that any radical in a solution was a rational<br />

function <strong>of</strong> the roots, had made HAMILTON doubt the subsequent reasoning and he<br />

supplied another deduction. With KÖNIGSBERGER, the original deduction was res-<br />

cued by a slight modification, and the flaw was claimed — without detailed pro<strong>of</strong> —<br />

to be <strong>of</strong> no importance in the pro<strong>of</strong>. Subsequently, the classification <strong>of</strong> algebraic ex-<br />

pressions was disbanded altogether in the impossibility pro<strong>of</strong>. <strong>The</strong> other obscurity —<br />

ABEL’S classification <strong>of</strong> rational functions <strong>of</strong> five quantities which take on five dif-<br />

ferent values under permutations — had been noticed in private correspondence by<br />

KÜLP, and ABEL had presented him with another more transparent deduction which,<br />

unfortunately, remained unknown to the larger mathematical public. When HAMIL-<br />

TON noticed the weakness <strong>of</strong> the published argument, he recast the deduction within<br />

his own framework; HOLMBOE provided it with a more general pro<strong>of</strong> along the lines<br />

<strong>of</strong> other parts <strong>of</strong> ABEL’S reasoning.<br />

For various reasons — doubt and curiosity, debate over the validity <strong>of</strong> result, and<br />

concerns for the best presentation <strong>of</strong> ABEL’S work — these mathematicians took up<br />

weak parts <strong>of</strong> ABEL’S pro<strong>of</strong> and provided clearer arguments and pro<strong>of</strong>s. This local<br />

criticism served to establish the overall validity <strong>of</strong> the impossibility <strong>of</strong> algebraically<br />

solving the quintic by examining and improving the pro<strong>of</strong>.<br />

6.9.2 Dissemination <strong>of</strong> the knowledge that the quintic was<br />

unsolvable<br />

<strong>The</strong> controversy which raged on the British Isles concerning the insolubility <strong>of</strong> the gen-<br />

eral quintic equation seems to have been largely confined to there, 99 although HAMIL-<br />

TON was also called upon to refute the claim for solubility made by the Italian P. G.<br />

BADANO. 100 While HAMILTON’S penetrating local criticism <strong>of</strong> ABEL’S pro<strong>of</strong> was un-<br />

dertaken to resolve an ongoing debate over ABEL’S statement, the Continental incor-<br />

poration <strong>of</strong> ABEL’S result apparently followed another path. On the Franco-German<br />

scene, I am not aware <strong>of</strong> any global rejections <strong>of</strong> ABEL’S result. <strong>The</strong> style <strong>of</strong> later local re-<br />

workings <strong>of</strong> ABEL’S pro<strong>of</strong> left little clue as to what, besides refinement and aesthetics,<br />

had spurred the mathematician to reformulate the argument.<br />

98 (Königsberger, 1869, 168).<br />

99 Besides JERRARD, MACCULLAGH also transmitted a claim to have solved the general fifth degree<br />

equation and was refuted by HAMILTON (Hankins, 1980, 438, note 22).<br />

100 (W. R. Hamilton, 1843; W. R. Hamilton, 1844). See also (Hankins, 1980, 438, note 22). Personal data<br />

for Mr. BADANO have proved to be inaccessible.


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 135<br />

PIERRE LAURENT WANTZEL. In a short paper published in 1845, PIERRE LAURENT<br />

WANTZEL refined ABEL’S pro<strong>of</strong> by reversing the succession in which the radicals <strong>of</strong><br />

a supposed solution is studied like HAMILTON had done. 101 Although WANTZEL<br />

deemed ABEL’S pro<strong>of</strong> to be exact, he also found its presentation vague and compli-<br />

cated. Nevertheless, WANTZEL gave no detailed reasons for his evaluation.<br />

“Although his [ABEL’S ] pro<strong>of</strong> is basically exact it is presented in a way so<br />

complicated and so vague that it would not be generally permissible.” 102<br />

Through a fusion <strong>of</strong> ABEL’S pro<strong>of</strong> with the even vaguer and more insufficient pro<strong>of</strong><br />

by RUFFINI, WANTZEL arrived at a clear and precise pro<strong>of</strong> which he thought would<br />

“lift all doubts concerning this important part <strong>of</strong> the theory <strong>of</strong> equations”. 103 Unfortu-<br />

nately, he did not specify his “doubts”.<br />

In his fusion pro<strong>of</strong>, WANTZEL took over the most important <strong>of</strong> ABEL’S preliminary<br />

arguments: the classification <strong>of</strong> algebraic expressions by orders and degrees and the<br />

auxiliary theorem derived from it (see page 104). By studying any supposed solution<br />

<strong>of</strong> the general n th degree equation and permutations <strong>of</strong> the roots, WANTZEL deduced<br />

that the outermost root extraction would have to be a square root. 104 Continuing to<br />

the radical <strong>of</strong> second highest order, he found that it had to remain unaltered by any 3-<br />

cycle, and therefore by any 5-cycle. 105 At this point he reached a contradiction because<br />

the supposed solution would thus only have two different values under all permuta-<br />

tions <strong>of</strong> the five roots.<br />

WANTZEL’S pro<strong>of</strong> was published in the Nouvelles annales de mathematique and soon<br />

became the widely accepted simplification <strong>of</strong> ABEL’S pro<strong>of</strong>. It made no use <strong>of</strong> ABEL’S<br />

classification <strong>of</strong> functions <strong>of</strong> five quantities, and may thus be seen as an indirect lo-<br />

cal criticism <strong>of</strong> this classification. On the other hand, it builds directly upon ABEL’S<br />

classification <strong>of</strong> algebraic expressions.<br />

A. E. G. ANDERSSEN. In 1848, the Königlichen Friedrichs-Gymnasium in Breslau in-<br />

vited its “protectors, sons, and friends” to be present at the annual exams. Included<br />

with the invitation was a short essay written by one <strong>of</strong> the teachers; at the time, this<br />

was not uncommon practice for German Gymnasien. 106 In the essay, A. E. G. ANDER-<br />

SSEN 107 sought to illuminate the central arguments <strong>of</strong> ABEL’S impossibility pro<strong>of</strong>.<br />

Being largely a reproduction <strong>of</strong> ABEL’S argument with some elaboration <strong>of</strong> its briefest<br />

arguments, the interesting parts <strong>of</strong> ANDERSSEN’S essay are his evaluations <strong>of</strong> ABEL’S<br />

101 (Wantzel, 1845).<br />

102 “Quoique sa démonstration soit exacte au fond, elle est présentée sous une forme trop compliquée<br />

et tellement vague, qu’elle n’a pas été généralement admise.” (ibid., 57).<br />

103 (ibid., 57).<br />

104 (ibid., 62).<br />

105 (ibid., 63–64). See also CAUCHY’S pro<strong>of</strong> <strong>of</strong> the CAUCHY-RUFFINI theorem, section 5.6.<br />

106 (Anderssen, 1848).<br />

107 No further personal information concerning this Mr. ANDERSSEN has been accessible.


136 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

pro<strong>of</strong>. ANDERSSEN found the pro<strong>of</strong> to be simple, coherent, and not built upon calcu-<br />

lations but on arguments and deductions; at the same time, and possibly for the same<br />

reasons, he rated it as being difficult.<br />

“However simple this pro<strong>of</strong> is, first <strong>of</strong> all because a single idea serves throughout<br />

as a decisive criterion, secondly because the truths by which the application<br />

<strong>of</strong> the main idea is possible, are communicated not by artificial calculations but<br />

by conclusions and deductions, it nevertheless (and even therefore) demands the<br />

most thorough contemplation in order to be understood in its entire clarity. <strong>The</strong>refore<br />

it would not be a superfluous work to present the most important arguments<br />

<strong>of</strong> this instructive yet difficult pro<strong>of</strong> by examples and further elaborations in their<br />

true spirit and full power <strong>of</strong> pro<strong>of</strong>.” 108<br />

ABEL’S classification <strong>of</strong> algebraic expressions according to orders and degrees was<br />

reproduced in an overly simplified form, in which the concept <strong>of</strong> degree has com-<br />

pletely vanished. When it came to the classification <strong>of</strong> functions with five quantities,<br />

which HAMILTON had scrutinized, ANDERSSEN found it quite satisfactory:<br />

“Both these two theorems [no function <strong>of</strong> five quantities can have two or<br />

five different values under all possible interchanges <strong>of</strong> the quantities] have been<br />

proved in <strong>Abel</strong>’s treatise with a clarity which cannot be improved.” 109<br />

ANDERSSEN’S essay contained no criticism <strong>of</strong> parts <strong>of</strong> ABEL’S pro<strong>of</strong> nor any origi-<br />

nal modifications but only simple elaborations and some examples. However, its mere<br />

existence is evidence that ABEL’S result was becoming known to the broader circle <strong>of</strong><br />

German mathematicians.<br />

LEOPOLD KRONECKER. <strong>The</strong> introduction <strong>of</strong> ABEL’S work on the quintic equation<br />

into German academic circles is due to LEOPOLD KRONECKER. Much <strong>of</strong> KRONECKER’S<br />

work on algebra was inspired by ideas which he got reading ABEL and KRONECKER<br />

completed and rigorized many parts <strong>of</strong> ABEL’S research. In KRONECKER’S elegant<br />

pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the general fifth (and higher) degree equation, ABEL’S<br />

pro<strong>of</strong> found its final form. KRONECKER presented his simplified version <strong>of</strong> ABEL’S<br />

pro<strong>of</strong> in a paper read to the Akademie der Wissenschaften. 110 <strong>The</strong>re, he presented no<br />

criticism <strong>of</strong> ABEL’S pro<strong>of</strong> but simply put forward alternative deductions preferable to<br />

ABEL’S on account <strong>of</strong> their simplicity and general nature. <strong>The</strong> validity <strong>of</strong> the result<br />

108 “So einfach dieser Beweis ist, erstens weil ein einziger Gedanke durchgehends zum entscheidenden<br />

Kriterium dient, zweitens weil diejenigen Wahrheiten, kraft deren die Anwendung des Hauptgedankens<br />

möglich ist, nicht durch künstliche Rechnungen, sondern durch Urtheile und Schlüsse<br />

vermittelt werden; so erheischt er dennoch, ja eben deswegen das gesammeltste Nachdenken, um in<br />

seiner ganzen Klarheit begriffen zu werden. Es dürfte daher keine unnöthige Arbeit sein, die wichtigsten<br />

Argumente dieses eben so lehrreichen als schwierigen Beweises durch Beispiele und weitere<br />

Ausführung in ihrem wahren Sinne und in ihrer vollständigen Beweiskraft zur Anschauung zu<br />

bringen.” (Anderssen, 1848, 3).<br />

109 “Diese beiden Lehrsätze sind in <strong>Abel</strong>’s Abhandlung mit einer Klarheit bewiesen, welche durch<br />

Nichts erhöht werden kann.” (ibid., 14).<br />

110 (Kronecker, 1879).


6.9. Reception <strong>of</strong> ABEL’s work on the quintic 137<br />

was never questioned by KRONECKER; his improvements were local in the sense <strong>of</strong><br />

replacing some <strong>of</strong> ABEL’S arguments by more apt ones.<br />

Through a detailed reworking <strong>of</strong> ABEL’S classification, KRONECKER obtained a<br />

more precise formulation <strong>of</strong> ABEL’S auxiliary theorem on the rationality <strong>of</strong> all radicals<br />

occurring in any solution. KRONECKER let R, R ′ , R ′′ , etc. denote quantities, which<br />

were to be considered known, and spoke <strong>of</strong> the collection <strong>of</strong> these as “the quantities<br />

R”. Later, this R evolved into his general concept <strong>of</strong> domains <strong>of</strong> rationality.<br />

“In the described way the explicit algebraic function satisfying an equation<br />

Φ (x) = 0 can be expressed as an entire function <strong>of</strong> the quantities<br />

W1, W2, . . . Wµ<br />

the coefficients <strong>of</strong> which are rational functions <strong>of</strong> the quantities R; the quantities<br />

W are on the one hand entire integer functions <strong>of</strong> the roots <strong>of</strong> the equation<br />

Φ (x) = 0 and <strong>of</strong> roots <strong>of</strong> unity and on the other hand determined through a chain<br />

<strong>of</strong> equations<br />

W n β<br />

β<br />

= Gβ<br />

� �<br />

Wβ+1, Wβ+2, . . . Wµ<br />

(β = 1, 2, . . . µ)<br />

n1, n2, . . . being prime numbers and G1, G2, . . . Gµ designating entire functions <strong>of</strong><br />

the bracketed quantities W in which the coefficients are rational functions <strong>of</strong> the<br />

quantities R.” 111<br />

Although apparently formulated in a slightly different way, this theorem is very<br />

close to the one <strong>of</strong> ABEL’S auxiliary theorems ensuring the rationality <strong>of</strong> the involved<br />

radicals (theorem 3), and served KRONECKER as its equivalent. <strong>The</strong> improvements are<br />

mainly the introduction <strong>of</strong> the rationally known quantities R and the explicit mention<br />

<strong>of</strong> the roots <strong>of</strong> unity. To ABEL, the roots <strong>of</strong> unity had been “known” in the common<br />

language version <strong>of</strong> this word, because he knew enough <strong>of</strong> them to handle them as<br />

simple objects. Consequently, roots <strong>of</strong> unity were not explicitly mentioned. <strong>The</strong> pro-<br />

cess <strong>of</strong> attributing technical mathematical meaning to a common language term oc-<br />

curred frequently in the period as is evident, for example, in the way GALOIS’S notion<br />

<strong>of</strong> groups was transformed from meaning a “collection <strong>of</strong> objects” into a term with a<br />

highly technical meaning.<br />

111 “In der dargelegten Weise erhält die einer Gliechung Φ (x) = 0 genügende explicite algebraische<br />

Function als ganze Function von Grössen<br />

W1, W2, . . . Wµ<br />

dargestellt, deren Coëfficienten rationale Functionen der Grössen R sind, und die Grössen W sind<br />

einerseits ganze ganzzahlige Functionen von Wurzeln der Gleichung Φ (x) = 0 und von Wurzeln<br />

der Einheit andererseits durch eine Kette von Gleichungen<br />

W nβ β = G � �<br />

β Wβ+1, Wβ+2, . . . Wµ<br />

(β = 1, 2, . . . µ)<br />

bestimmt, in denen n1, n2, . . . Primzahlen und G1, G2, . . . Gµ ganze Functionen der eingeklammerten<br />

Grössen W bedeuten, deren Coëfficienten rationale Functionen der Grössen R sind.” (ibid., 77).


138 Chapter 6. Algebraic insolubility <strong>of</strong> the quintic<br />

<strong>The</strong> succeeding part <strong>of</strong> KRONECKER’S pro<strong>of</strong> concerned the substitution theoretic<br />

aspects <strong>of</strong> ABEL’S pro<strong>of</strong> and consisted <strong>of</strong> an extended version <strong>of</strong> the CAUCHY-RUFFINI<br />

theorem. For n > 4, KRONECKER let f designate a function <strong>of</strong> quantities x1, . . . , xn and<br />

studied the conjugate functions f1, . . . , fm; these functions were the analogous <strong>of</strong> what<br />

ABEL had called the different values <strong>of</strong> f under all permutations <strong>of</strong> x1, . . . , xn. KRO-<br />

NECKER derived the result that for any non-symmetric function f , some permutation<br />

would exist which altered the value <strong>of</strong> one <strong>of</strong> the conjugate functions. He could even<br />

demonstrate that if only the n!<br />

� �<br />

2 permutations, which left the product ∏i


6.10. Summary 139<br />

<strong>of</strong> ABEL’S argument but superior in rigour. <strong>The</strong> classification <strong>of</strong> algebraic expressions<br />

was also a concern <strong>of</strong> some later 19 th century mathematicians, until it was settled by<br />

KÖNIGSBERGER and KRONECKER.<br />

<strong>The</strong> fact that global criticism <strong>of</strong> ABEL’S impossibility pro<strong>of</strong> was limited can be taken<br />

as a sign that the mathematical community soon came to realize the overall validity <strong>of</strong><br />

the result. <strong>The</strong> change <strong>of</strong> attitude toward the problem, which had been facilitated by<br />

the statements <strong>of</strong> experts such as LAGRANGE and GAUSS (section 5.4) and the pro<strong>of</strong>s<br />

<strong>of</strong> RUFFINI, which at least were known in some circles in Paris, had been a prerequisite<br />

for the quick acceptance. However, local criticism was still conducted in an effort to<br />

make ABEL’S pro<strong>of</strong> clearer and more powerful. Central lemmata, on which doubt<br />

could be cast, were reexamined and new pro<strong>of</strong>s were given.<br />

6.10 Summary<br />

As described, ABEL’S pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the general quintic was a curious<br />

combination <strong>of</strong> general theorems and investigations <strong>of</strong> particular cases. Partly be-<br />

cause <strong>of</strong> the counter intuitive nature <strong>of</strong> the result and partly because <strong>of</strong> legitimate<br />

local objections to ABEL’S argument, the result was subsequently scrutinized. Inter-<br />

preted in terms <strong>of</strong> delineation <strong>of</strong> concepts, the algebraic insolubility <strong>of</strong> the general<br />

quintic distinguished the concepts <strong>of</strong> polynomial equations and algebraically solvable<br />

equations.


Chapter 7<br />

Particular classes <strong>of</strong> equations:<br />

enlarging the class <strong>of</strong> solvable<br />

equations<br />

If N. H. ABEL’S (1802–1829) pro<strong>of</strong> <strong>of</strong> the impossibility <strong>of</strong> solving the general quin-<br />

tic algebraically was hampered by its brevity and obscure arguments, his only other<br />

published work on the theory <strong>of</strong> equations was more mature, beautifully lucid, and<br />

rigorous. In the Mémoire sur une classe particulière d’équations résolubles algébriquement,<br />

written in 1828 and published the following year, ABEL abandoned one <strong>of</strong> the central<br />

pillars <strong>of</strong> the impossibility pro<strong>of</strong> — the theory <strong>of</strong> permutations — and provided a direct<br />

and affirmative pro<strong>of</strong> <strong>of</strong> the algebraic solubility <strong>of</strong> a particular class <strong>of</strong> equations. 1<br />

Focusing instead on the other pillar — the concepts <strong>of</strong> divisibility, irreducibility, and<br />

the Euclidean algorithm — this work illuminates central ideas in ABEL’S reasoning<br />

which permeate his entire work on the theory <strong>of</strong> equations.<br />

<strong>The</strong> 1829-paper has become a classic <strong>of</strong> mathematics for its pro<strong>of</strong> that the class <strong>of</strong><br />

equations, now called <strong>Abel</strong>ian and defined by certain properties <strong>of</strong> the roots, are always<br />

algebraically solvable. When contrasted with the contents <strong>of</strong> the impossibility pro<strong>of</strong>,<br />

this result highlights a feature <strong>of</strong> the new ways <strong>of</strong> asking questions — the mechanisms<br />

<strong>of</strong> limiting and enlarging class <strong>of</strong> objects which in the nineteenth century provided the<br />

background for a new, concept based approach to mathematics. However, the paper<br />

contains more information than just this main result; in this chapter I describe some <strong>of</strong><br />

the connections between this work and other parts <strong>of</strong> ABEL’S research as well as some<br />

<strong>of</strong> the very central concepts which ABEL put to use in it.<br />

1 (N. H. <strong>Abel</strong>, 1829c).<br />

141


142 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

7.1 Solubility <strong>of</strong> <strong>Abel</strong>ian equations<br />

<strong>The</strong> structure <strong>of</strong> ABEL’S Mémoire sur une classe particulière 2 is a descent from the gen-<br />

eral to the particular. At the outset, ABEL proposed to study irreducible equations in<br />

which one <strong>of</strong> the roots depended rationally on another one. <strong>The</strong> concept <strong>of</strong> irreducible<br />

equations took a central place in this research (see section 7.3). Part <strong>of</strong> the study was<br />

especially devoted to circular functions to which ABEL had been led by C. F. GAUSS’<br />

(1777–1855) work on the cyclotomic equation. Besides this application to circular func-<br />

tions, ABEL also worked on another application <strong>of</strong> the general theory to the division<br />

problem for elliptic functions. Likewise inspired by GAUSS’ Disquisitiones arithmeti-<br />

cae (see section 7.2), this application was, however, not contained in the paper but<br />

had been presented the previous year in a paper on elliptic functions. ABEL was led<br />

by these two applications to an even more general result — valid for a broader class<br />

<strong>of</strong> equations having rationally related roots. In this section, I outline ABEL’S results<br />

before turning to discussions <strong>of</strong> his inspirations and methods.<br />

7.1.1 Decomposition <strong>of</strong> the equation into lower degrees<br />

Throughout the paper, ABEL studied polynomial equations <strong>of</strong> degree µ,<br />

φ (x) = 0,<br />

in which two roots x1, x ′ were related by the rational function θ,<br />

x ′ = θ (x1) .<br />

<strong>The</strong> quantities which ABEL considered known in his deductions comprise all coeffi-<br />

cients occurring in φ or θ. From a modern perspective, it will become clear that he also<br />

considered any required roots <strong>of</strong> unity to be known.<br />

ABEL defined the equation φ (x) = 0 to be irreducible when none <strong>of</strong> its roots could<br />

be expressed by a similar equation <strong>of</strong> lower degree (see section 7.3).<br />

Employing the Euclidean division algorithm (see section 7.3) and the notation<br />

θ k (x1) for the k th iterated application <strong>of</strong> the rational function θ to x1, ABEL found<br />

that the set <strong>of</strong> roots <strong>of</strong> φ (x) = 0 split into sequences (chains). He deduced — using the<br />

irreducibility <strong>of</strong> φ (x) = 0 — that because the two roots x1, x ′ <strong>of</strong> the equation φ (x) = 0<br />

were rationally related, every iteration θ k (x1) would also be a root <strong>of</strong> φ (x) = 0. <strong>The</strong>re-<br />

fore, the entire set <strong>of</strong> roots <strong>of</strong> φ (x) = 0 could be collected in sequences <strong>of</strong> equal length,<br />

say n, and he wrote the roots as (µ = m × n), 3<br />

θ k (xu) for 0 ≤ k ≤ n − 1 and 1 ≤ u ≤ m. (7.1)<br />

2 (N. H. <strong>Abel</strong>, 1829c).<br />

3 For brevity, I have added to ABEL’S notation the convention θ 0 (x1) = x1.


7.1. Solubility <strong>of</strong> <strong>Abel</strong>ian equations 143<br />

After ABEL had divided the roots into sequences, he proceeded to reduce the so-<br />

lution <strong>of</strong> the equation <strong>of</strong> degree µ to equations <strong>of</strong> lower degrees. To the first sequence<br />

x1, θ (x1) , . . . , θ n−1 (x1), ABEL assigned an arbitrary rational and symmetric function<br />

y1 <strong>of</strong> these quantities. Since θ was also a rational function, y1 was actually a rational<br />

function <strong>of</strong> x1,<br />

y1 = f<br />

and using the symmetry <strong>of</strong> y1, ABEL found<br />

and more generally<br />

�<br />

x1, θ (x1) , . . . , θ n−1 �<br />

(x1) = F (x1) ,<br />

y ν 1<br />

y1 = 1<br />

n−1<br />

n<br />

∑<br />

k=0<br />

�<br />

F θ k �<br />

(x1)<br />

n−1<br />

1<br />

� �<br />

=<br />

n ∑ F θ<br />

k=0<br />

k ��ν (x1)<br />

(7.2)<br />

for any non-negative integer ν. In the same way as ABEL formed the function y1<br />

from x1, he formed an additional m − 1 functions y2, . . . , ym corresponding to the other<br />

chains,<br />

yu = f<br />

�<br />

xu, θ (xu) , . . . , θ n−1 �<br />

(xu) = F (xu) for 1 ≤ u ≤ m. (7.3)<br />

Each <strong>of</strong> these produced the equivalent <strong>of</strong> (7.2)<br />

ABEL added these (over u) as<br />

y ν u = 1<br />

n−1 � �<br />

n ∑ F θ<br />

k=0<br />

k ��ν (xu)<br />

rν =<br />

for 1 ≤ u ≤ m and ν ≥ 0.<br />

m<br />

∑ y<br />

u=1<br />

ν u for ν ≥ 0 (7.4)<br />

and obtained rational and symmetric functions <strong>of</strong> all the roots <strong>of</strong> φ (x) = 0. <strong>The</strong>se<br />

could, he noticed, therefore be expressed rationally in the coefficients <strong>of</strong> the known<br />

functions φ and θ. Once these power sums (7.4) were known, ABEL could determine<br />

any rational and symmetric function <strong>of</strong> y1, . . . , ym by the solution <strong>of</strong> an equation <strong>of</strong><br />

degree m by E. WARING’S (∼1736–1798) result (see section 5.2.4). In particular, ABEL<br />

found that each <strong>of</strong> the coefficients <strong>of</strong> the equation<br />

m<br />

∏ (y − yu) = 0 (7.5)<br />

u=1<br />

could be determined by solving an equation <strong>of</strong> the m th degree.<br />

A central topic <strong>of</strong> ABEL’S paper is the detailed study <strong>of</strong> this decomposition <strong>of</strong> the<br />

equation <strong>of</strong> degree µ = m × n into equations <strong>of</strong> degrees m and n. His next step


144 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

was to focus attention on the equation connected with the first sequence <strong>of</strong> roots<br />

x1, θ (x1) , . . . , θ n−1 (x1), i.e.<br />

n−1 �<br />

∏ x − θ<br />

k=0<br />

k �<br />

(x1) = 0. (7.6)<br />

ABEL proved that any coefficient ψ (x1) <strong>of</strong> this equation would depend rationally on<br />

y1 and known quantities <strong>of</strong> φ and θ by the following nice and typical argument.<br />

Denoting by ψ (x1) any coefficient <strong>of</strong> (7.6), ABEL formed the expressions<br />

tν =<br />

m<br />

∑ y<br />

u=1<br />

ν u · ψ (xu) for ν ≥ 0,<br />

which he proved to be rational and symmetric functions <strong>of</strong> all the roots <strong>of</strong> φ (x) = 0.<br />

<strong>The</strong>reby, tν could be expressed rationally in the known quantities.<br />

From a set <strong>of</strong> linear equations equivalent to the matrix equation<br />

⎡<br />

⎢<br />

⎣<br />

11 . . .1<br />

y1y2<br />

..<br />

y m−1<br />

1 y m−1<br />

2<br />

. . .ym<br />

. ...<br />

. . .y m−1<br />

m<br />

⎤ ⎡ ⎤ ⎡<br />

ψ (x1)<br />

⎥ ⎢ ⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢<br />

ψ (x2) ⎥ ⎢<br />

⎥ ⎢<br />

⎥ ⎢ ⎥ = ⎢<br />

⎥ ⎢<br />

⎦ ⎣ .<br />

⎥ ⎢<br />

⎦ ⎣<br />

ψ (xm)<br />

t0<br />

⎥ ,<br />

.<br />

⎥<br />

⎦<br />

t1<br />

tm−1<br />

ABEL deduced that ψ (x1) could be expressed as a rational function <strong>of</strong> y1, . . . , ym. His<br />

argument is based on the possibility <strong>of</strong> attributing a non-vanishing form to the equiva-<br />

lent <strong>of</strong> the determinant <strong>of</strong> the matrix. This was possible because y1 had — up to now —<br />

been an arbitrary symmetric function, and ABEL gave it the non-vanishing form<br />

y1 =<br />

n−1 �<br />

∏ α − θ<br />

k=0<br />

k �<br />

(x1) ,<br />

where α was unspecified. Furthermore, ABEL continued to show how each <strong>of</strong> the<br />

quantities y2, . . . , ym could be replaced by a rational function <strong>of</strong> y1, and how ψ (x1)<br />

could be expressed as a rational function <strong>of</strong> y1 alone. Thus, each coefficient ψ (x1) in<br />

the equation (7.6) could be determined rationally in y1; and y1 could be determined<br />

by solving an equation <strong>of</strong> degree m. ABEL summarized these results in an important<br />

theorem:<br />

<strong>The</strong>orem 5 “<strong>The</strong> equation under consideration φx = 0 can thus be decomposed into a num-<br />

ber m <strong>of</strong> equations <strong>of</strong> degree n in which the coefficients are rational functions <strong>of</strong> a fixed root <strong>of</strong><br />

a single equation <strong>of</strong> degree m, respectively.” 4<br />

4 “L’équation proposée φx = 0 peut donc être décomposée en un nombre de m d’équations du degré<br />

n; donc[!] les coëfficiens sont respectivement des fonctions rationnelles d’une même racine d’une<br />

seule équation du degré m.” (N. H. <strong>Abel</strong>, 1829c, 139). <strong>The</strong> misprint “donc” has been replaced by<br />

“dont” in both editions <strong>of</strong> ABEL’S Œuvres.<br />


7.1. Solubility <strong>of</strong> <strong>Abel</strong>ian equations 145<br />

Thus, the original problem <strong>of</strong> solving the equation φ (x) = 0 <strong>of</strong> degree µ had<br />

been reduced to solving certain equations, (7.5) and (7.6), <strong>of</strong> lower degrees. Gener-<br />

ally, the equation <strong>of</strong> degree m would not be solvable by radicals, but as ABEL went on<br />

to demonstrate, the m equations <strong>of</strong> degree n could always be solved algebraically.<br />

7.1.2 Algebraic solubility <strong>of</strong> <strong>Abel</strong>ian equations<br />

If all the roots <strong>of</strong> the equation φ (x) = 0 fell into the same orbit <strong>of</strong> θ (one chain), i.e. are<br />

<strong>of</strong> the form<br />

x1, θ (x1) , θ 2 (x1) , . . . , θ n−1 (x1) ,<br />

the situation was equivalent to assuming m = 1 above. In this case, ABEL let α denote<br />

a primitive µ th root <strong>of</strong> unity and formed the rational expression<br />

ψ (x) =<br />

�<br />

µ−1<br />

∑ α<br />

k=0<br />

k θ k �µ<br />

(x) . (7.7)<br />

Through direct calculations, he proved that<br />

�<br />

ψ θ k �<br />

(x) = ψ (x) for all k = 0, 1, . . . , µ − 1,<br />

which showed that ψ was a symmetric function <strong>of</strong> the roots <strong>of</strong> φ (x) = 0. Thus, ψ (x)<br />

could be expressed rationally in known quantities. Next, ABEL introduced µ radicals<br />

<strong>of</strong> (7.7),<br />

µ√ vu =<br />

µ−1<br />

∑ α<br />

k=0<br />

k uθ k (x) for 0 ≤ u ≤ µ − 1, (7.8)<br />

by attributing to αu the different µ th roots <strong>of</strong> unity 1, α, α 2 , . . . , α µ−1 . From these radi-<br />

cals, it was a routine procedure for ABEL to obtain the expression<br />

where A was a constant.<br />

θ k (x) = 1<br />

�<br />

µ−1<br />

−A +<br />

µ<br />

∑<br />

u=1<br />

α uk µ√ vu<br />

�<br />

, k = 0, 1, . . . , µ − 1, (7.9)<br />

<strong>The</strong> expression (7.9), however, contained µ − 1 extractions <strong>of</strong> roots with exponent<br />

µ which seemed to indicate that a total <strong>of</strong> µ µ−1 different values could be obtained al-<br />

though the degree <strong>of</strong> φ (x) = 0 was only µ. ABEL resolved this apparent contradiction,<br />

similar to one noticed by L. EULER (1707–1783) (see section 5.1), by an elegant argu-<br />

ment prototypic <strong>of</strong> his approach to the theory <strong>of</strong> equations. In the deduction, ABEL<br />

proved that all the root extractions depended on one <strong>of</strong> them by considering<br />

µ√ vk ( µ√ v1) µ−k =<br />

�<br />

µ−1<br />

∑ α<br />

u=0<br />

ku θ u �<br />

(x) ×<br />

�<br />

µ−1<br />

∑ α<br />

u=0<br />

u θ u �µ−k<br />

(x) .


146 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

Obviously, the form <strong>of</strong> the right hand side shows that this expression was a rational<br />

function <strong>of</strong> x. ABEL stated that it was unaffected by substituting θ m (x) for x and<br />

considered it so obvious that he did not provide the details. 5 Thus, the expression<br />

was a rational function <strong>of</strong> the coefficients <strong>of</strong> φ (x) = 0; ABEL denoted this function by<br />

a k,<br />

µ√<br />

vk = ak ( µ√ v1) k .<br />

v1<br />

ABEL stated the conclusion <strong>of</strong> this investigation by giving an algebraic formula for<br />

the root x, 6<br />

x = 1<br />

�<br />

−A +<br />

µ<br />

µ√ v1 + a2<br />

(<br />

v1<br />

µ√ v1) 2 + a3<br />

(<br />

v1<br />

µ√ v1) 3 + · · · + aµ−1<br />

(<br />

v1<br />

µ√ v1) µ−1<br />

�<br />

. (7.10)<br />

All the other roots were contained in this formula by giving µ√ v1 its µ different values<br />

α k µ √ v1. ABEL expressed the implications for solubility in two theorems capturing the<br />

essence <strong>of</strong> this research. If the set <strong>of</strong> roots fell into one “orbit” <strong>of</strong> the rational expres-<br />

sion, θ, ABEL found the equation to be solvable by radicals:<br />

<strong>The</strong>orem 6 “If the roots <strong>of</strong> an algebraic equation can be represented by:<br />

x, θx, θ 2 x, . . . θ µ−1 x,<br />

where θ µ x = x and θx denotes a rational function <strong>of</strong> x and known quantities, this equation<br />

will always be algebraically solvable.” 7<br />

Applying this result to the particular case <strong>of</strong> irreducible equations <strong>of</strong> prime degree,<br />

which always had only one chain, ABEL found that such equations were algebraically<br />

solvable:<br />

“If two roots <strong>of</strong> an irreducible equation, <strong>of</strong> which the degree is a prime number,<br />

have such a relation that one can express the one rationally in the other, this<br />

equation will be algebraically solvable.” 8<br />

Subsequently, ABEL refined the hypothesis that all the roots could be expressed as<br />

iterations <strong>of</strong> a rational function. That hypothesis had ensured algebraic solubility <strong>of</strong><br />

the equation, but the same conclusion could also be established for a broader class <strong>of</strong><br />

equations. Under the general assumption that every root <strong>of</strong> an equation, χ (x) = 0,<br />

5 <strong>The</strong> details can easily be provided by inserting into the right hand side and rearranging terms.<br />

6 (N. H. <strong>Abel</strong>, 1829c, 142).<br />

7 “Si les racines d’une équation algébrique peuvent être représentées par:<br />

x, θx, θ 2 x, . . . θ µ−1 x,<br />

où θ µ x = x et θx désigne une fonction rationelle de x et de quantités connues, cette équation sera<br />

toujours résoluble algébriquement.” (ibid., 142–143).<br />

8 “Si deux racines d’une équation irréductible, dont le degré est un nombre premier, sont dans un<br />

tel rapport, qu’on puisse exprimer l’une rationnellement par l’autre, cette équation sera résoluble<br />

algébriquement.” (ibid., 143).


7.1. Solubility <strong>of</strong> <strong>Abel</strong>ian equations 147<br />

could be expressed rationally in a single root x, ABEL went on to assume “commuta-<br />

tivity” <strong>of</strong> these rational dependencies, i.e. if θ (x) and θ1 (x) were any two roots <strong>of</strong> the<br />

equation χ (x) = 0, written as rational expressions in x, the assumption was that<br />

θ (θ1 (x)) = θ1 (θ (x)) .<br />

ABEL’S method <strong>of</strong> proving the algebraic solubility <strong>of</strong> χ (x) = 0 under this hypothesis<br />

was to reduce the situation to the one solved above. Since all roots were known ra-<br />

tionally once x was considered known, it sufficed to search for the root x. In order to<br />

study an irreducible equation, ABEL focused on the irreducible factor φ <strong>of</strong> χ having x<br />

as a root, repeating his concept <strong>of</strong> irreducibility (see section 7.3).<br />

“If the equation<br />

is not irreducible, let<br />

χx = 0<br />

φx = 0<br />

be the equation <strong>of</strong> lowest degree which the root x satisfies such that the coefficients<br />

<strong>of</strong> this equation contain nothing but known quantities.” 9<br />

Thus, ABEL assumed that φ (x) = 0 was the irreducible factor which had x as<br />

a root. By the deductions carried out above, the roots were thus expressed as (7.1),<br />

where for simplicity I write x0 for x:<br />

<strong>The</strong> coefficients <strong>of</strong> the equation<br />

θ k (xu) for 0 ≤ k ≤ n − 1 and 0 ≤ u ≤ m − 1.<br />

n−1 �<br />

∏ z − θ<br />

k=0<br />

k �<br />

(x0) = 0 (7.11)<br />

could all be expressed rationally in a single quantity y0 (above denoted y1) which was<br />

a root <strong>of</strong> an equation (7.5) <strong>of</strong> degree m. In a footnote, ABEL demonstrated that the<br />

latter equation was irreducible. <strong>The</strong>reby, he had reduced the determination <strong>of</strong> x to the<br />

solution <strong>of</strong> two equations <strong>of</strong> degrees n and m. Of these, he knew that the former was<br />

9 “Si l’équation<br />

n’est pas irréductible, soit<br />

χx = 0<br />

φx = 0<br />

l’équation la moins élevée, à laquelle puisse satisfaire la racine x, les coëfficiens de cette équation ne<br />

contenant que des quantités connues.” (ibid., 149–150).


148 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

algebraically solvable if y0 was considered known. Although the equation <strong>of</strong> degree<br />

m<br />

m−1<br />

∏ (z − yu) = 0 (7.12)<br />

u=0<br />

giving the coefficients <strong>of</strong> (7.11) would generally not be algebraically solvable, ABEL<br />

next proved that equation (7.12) ‘inherited’ the property <strong>of</strong> commutative rational de-<br />

pendence among its roots, which φ (x) = 0 possessed. Thus, a ‘descent’ down a string<br />

<strong>of</strong> equations was made possible.<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> this ‘inheritance’, the commutative rational dependence among<br />

the roots <strong>of</strong> (7.12), ran as follows. <strong>The</strong> hypothesis was that all the roots were given<br />

rationally in a single root, i.e.<br />

xu = θu (x0) for 0 ≤ u ≤ m − 1. (7.13)<br />

<strong>The</strong> expression for yu which in the previous argument was given by (7.3),<br />

�<br />

yu = f xu, θ (xu) , . . . , θ n−1 �<br />

(xu) for 0 ≤ u ≤ m − 1,<br />

under the current hypothesis became<br />

�<br />

y1 = f θ1 (x0) , θ (θ1 (x0)) , . . . , θ n−1 �<br />

(θ1 (x0)) .<br />

Combining this with the hypothesis <strong>of</strong> commutativity <strong>of</strong> the functions θ and θ1, ABEL<br />

found<br />

y1 = f<br />

�<br />

�<br />

θ1 (x0) , θ1 (θ (x0)) , . . . , θ1 θ n−1 ��<br />

(x0) .<br />

<strong>The</strong>refore, y1 was a rational and symmetric function <strong>of</strong> the sequence <strong>of</strong> roots (7.13) and<br />

could therefore be expressed rationally in y0 and known quantities. Obviously, ABEL<br />

could carry out the same argument for any other y2, . . . , ym−1. When he let λ (y0) and<br />

λ1 (y0) denote any two among the quantities y0, . . . , ym−1, he found that, without loss<br />

<strong>of</strong> generality,<br />

y1 = λ (y0) = f<br />

y2 = λ1 (y0) = f<br />

�<br />

θ1 (x0) , θ (θ1 (x0)) , . . . , θ n−1 �<br />

(θ1 (x0)) and<br />

�<br />

θ2 (x0) , θ (θ2 (x0)) , . . . , , θ n−1 �<br />

(θ2 (x0)) .<br />

Inserting θ2 (x) for x0 in λ (y0), which transformed y0 into y2, ABEL obtained 10<br />

λλ1 (y0) = λ (y2) = f<br />

while inserting θ1 (x) for x0 in λ1 (y0) produced<br />

λ1λ (y0) = λ1 (y1) = f<br />

�<br />

θ1θ2 (x0) , θθ1θ2 (x0) , . . . , θ n−1 �<br />

θ1θ2 (x0)<br />

�<br />

θ2θ1 (x0) , θθ2θ1 (x0) , . . . , θ n−1 �<br />

θ2θ1 (x0)<br />

10 Here I deviate from my usual notation by writing the composition <strong>of</strong> functions in multiplicative<br />

mode, i.e. θ1θ2 (x0) instead <strong>of</strong> θ1 (θ2 (x0)).<br />

,<br />

.


7.1. Solubility <strong>of</strong> <strong>Abel</strong>ian equations 149<br />

Since θ1θ2 (x0) = θ2θ1 (x0), ABEL concluded<br />

λλ1 (y0) = λ1λ (y0) ,<br />

and any two roots <strong>of</strong> the equation (7.12) would thus also commute. <strong>The</strong>refore, the equa-<br />

tion (7.12) determining the coefficients <strong>of</strong> (7.11) inherited this property from φ (x) = 0<br />

and could be treated in the same way as above. Since the degree was reduced by this<br />

argument, a chain <strong>of</strong> equations <strong>of</strong> strictly decreasing degrees could be constructed. At<br />

some point, where the procedure would have to terminate, the degree had to be 1 and<br />

the final equation would amount to a rational dependency.<br />

ABEL had thus established the following important theorem on the solubility <strong>of</strong><br />

this class <strong>of</strong> equations:<br />

<strong>The</strong>orem 7 <strong>The</strong> equation φ (x) = 0 is algebraically solvable if the following two requirements<br />

are met:<br />

1. All roots <strong>of</strong> φ (x) = 0 are rational expressions θ1 (x) , . . . , θµ (x) <strong>of</strong> one root<br />

2. <strong>The</strong> rational expressions satisfy a requirement <strong>of</strong> commutativity θ iθ j (x) = θ jθ i (x). 11 ✷<br />

Since the time <strong>of</strong> L. KRONECKER (1823–1891), equations with these properties have<br />

been named abelian [abelsche]; 12 in 1932, the Springer-Verlag decided to change the<br />

first letter into a capital: <strong>Abel</strong>ian. 13 Later, the term <strong>Abel</strong>ian was also adopted to denote<br />

groups corresponding to <strong>Abel</strong>ian equations, i.e. commutative groups.<br />

In two theorems, ABEL summarized the implications for the degrees <strong>of</strong> the equa-<br />

tions involved in the algebraic solution <strong>of</strong> the equation φ (x) = 0 in which two roots<br />

were rationally related. <strong>The</strong> following theorem completely describes these degrees:<br />

“Supposing that the degree µ <strong>of</strong> the equation φx = 0 is decomposed as follows:<br />

µ = ε v1<br />

1<br />

· εv2<br />

2<br />

· · · · · εvα<br />

α ,<br />

where ε1, ε2, . . . , εα are prime numbers, the determination <strong>of</strong> x can be effected with<br />

the help <strong>of</strong> the solution <strong>of</strong> v1 equations <strong>of</strong> degree ε1, v2 equations <strong>of</strong> degree ε2, etc.,<br />

and all these equations will be algebraically solvable.” 14<br />

11 It is remarkable and unfortunate that (Toti Rigatelli, 1994, 717) got the logic <strong>of</strong> ABEL’S reasoning<br />

wrong, reproducing the result as “he [ABEL in (N. H. <strong>Abel</strong>, 1829c)] showed that, in those equations<br />

which were solvable by radicals, all roots could be expressed as rational functions <strong>of</strong> any other root,<br />

and that these functions were permutable with respect to the four arithmetical operations. That is,<br />

if F1 and F2 are any two corresponding functional operations, then F1F2x = F2F1x.”<br />

12 (Kronecker, 1853, 6).<br />

13 <strong>The</strong> decision prompted a brief discussion among mathematicians, see<br />

the letters (Noether→Hasse, 1932.10.29 and 1932.12.09, described at<br />

������������������������������������������������������������)<br />

14 “Supposant le degré µ de l’équation φx = 0 décomposé comme il suit:<br />

µ = ε v 1<br />

1<br />

· εv2<br />

2<br />

· · · · · εvα<br />

α ,<br />

où ε1, ε2, . . . εα sont des nombres premiers, la détermination de x pourra s’effectuer à l’aide de la<br />

résolution de v1 équations du degré ε1, de v2 équations du degré ε2, etc., et toutes ces équations<br />

seront résolubles algébriquement.” (N. H. <strong>Abel</strong>, 1829c, 152).


150 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

<strong>The</strong> resemblance to GAUSS’ investigation <strong>of</strong> the cyclotomic equation is more than<br />

accidental. In multiple ways, GAUSS’ work was the direct inspiration for this research.<br />

Part <strong>of</strong> the purpose <strong>of</strong> ABEL’S paper was to reproduce GAUSS’S result in this more<br />

general framework.<br />

7.1.3 Application to circular functions and the link with GAUSS’<br />

Disquisitiones arithmeticae<br />

ABEL was led to the study <strong>of</strong> <strong>Abel</strong>ian equations by his in-depth studies <strong>of</strong> the divi-<br />

sion problem for elliptic functions (see section 7.2), which in turn were motivated by<br />

the division problem for circular functions treated by GAUSS in his Disquisitiones arith-<br />

meticae. 15 In the last part <strong>of</strong> the paper, ABEL incorporated GAUSS’ division <strong>of</strong> the circle<br />

into his broader theory <strong>of</strong> <strong>Abel</strong>ian equations by the following approach.<br />

<strong>The</strong> central result <strong>of</strong> the paper Mémoire sur une classe particulière was contained in<br />

the second theorem (here theorem 5) on the reduction <strong>of</strong> equations <strong>of</strong> degree m × n to<br />

m solvable equations <strong>of</strong> degree n and a single equation <strong>of</strong> degree m. Originating from<br />

this theorem, ABEL deduced more particular results in various directions. Assuming<br />

that all the known quantities (i.e. coefficients) <strong>of</strong> φ and θ were real numbers, he studied<br />

the constructions required for the solution <strong>of</strong> the equation φ (x) = 0. Considering real<br />

and imaginary parts <strong>of</strong> the radical µ√ v1 (7.8), ABEL found the (non-algebraic) solution<br />

formula<br />

x = 1<br />

µ<br />

�<br />

−A +<br />

µ−1 � √ �<br />

∑ fk + gk −1 (<br />

k=1<br />

√ �<br />

k<br />

ρ) cos k(δ+2mπ)<br />

µ<br />

+ √ −1 sin k(δ+2mπ)<br />

�<br />

µ<br />

�<br />

where the quantities ρ, A, f1, . . . , fµ−1, g1, . . . , gµ−1 were rational functions <strong>of</strong> cos 2π µ ,<br />

sin 2π µ<br />

and the coefficients <strong>of</strong> φ and θ. From this, he drew the following conclusion<br />

which was intimately linked to one <strong>of</strong> GAUSS’ results:<br />

<strong>The</strong>orem 8 “In order to solve the equation φx = 0 it suffices:<br />

1) to divide the circumference <strong>of</strong> the circle into µ equal parts,<br />

2) to divide an angle δ, which can then be constructed, into µ equal parts,<br />

3) to extract a square root <strong>of</strong> a single quantity ρ.” 16<br />

ABEL, himself, remarked that his result was an extension <strong>of</strong> one <strong>of</strong> the key results<br />

found in GAUSS’ Disquisitiones, stating the equivalent conclusion for cyclotomic equa-<br />

tions: That the solution <strong>of</strong> the equation x n = 1 could be reduced to the following three<br />

steps: 17<br />

15 (C. F. Gauss, 1801).<br />

16 “[Q]ue pour résoudre l’équation φx = 0, il suffit:<br />

1) de diviser la circonférence entière du cercle en µ parties égales,<br />

2) de diviser un angle δ, qu’on peut construire ensuite, en µ parties égales,<br />

3) d’extraire la racine carrée d’une seule quantité ρ.”<br />

(N. H. <strong>Abel</strong>, 1829c, 144).<br />

17 (C. F. Gauss, 1801, 454) and (C. F. Gauss, 1986, 450).


7.1. Solubility <strong>of</strong> <strong>Abel</strong>ian equations 151<br />

1) <strong>The</strong> division <strong>of</strong> the whole circle into n − 1 parts (n − 1 because the irreducible<br />

= 0),<br />

equation in GAUSS’ research was xn −1<br />

x−1<br />

2) <strong>The</strong> division into n − 1 parts <strong>of</strong> another arc which could be constructed after step<br />

1 had been completed, and<br />

3) <strong>The</strong> extraction <strong>of</strong> a square root.<br />

<strong>The</strong> final step, the extraction <strong>of</strong> a square root, could be assumed to equal the construc-<br />

tion <strong>of</strong> √ n, GAUSS claimed without providing any pro<strong>of</strong>. Later, ABEL adopted and<br />

proved the assertion.<br />

In the fifth section <strong>of</strong> the Mémoire sur une classe particulière, ABEL applied his theory<br />

directly to the cyclotomic equation and the circular functions related to the division <strong>of</strong><br />

the circle. By the addition formulae for cosine, ABEL could express cos ma rationally<br />

in cos a, and assuming θ (cos a) = cos ma and θ1 (cos a) = cos m ′ a, he obtained<br />

θθ1 (x) = θ � cos m ′ a � = cos � mm ′ a �<br />

= cos � m ′ ma � = θ1 (cos ma) = θ1θ (x) .<br />

From a previously established result (here theorem 7), ABEL found that cos 2π µ could<br />

be determined algebraically — which was a well known result.<br />

ABEL, however, did not stop his investigations <strong>of</strong> the circular functions at this<br />

point, as he might have done had he only been interested in the algebraic solubility <strong>of</strong><br />

the division. Assuming that µ = 2n + 1 was prime, ABEL studied the equation<br />

and used the rational dependency established above<br />

to write<br />

n �<br />

∏ X − cos<br />

k=1<br />

2kπ<br />

�<br />

= 0, (7.14)<br />

2n + 1<br />

θ (cos a) = cos ma<br />

θ k (cos a) = cos m k a.<br />

By an argument based on GAUSS’ primitive roots <strong>of</strong> the module 2n + 1, ABEL demon-<br />

strated that the roots <strong>of</strong> (7.14) were<br />

x, θ (x) , θ 2 (x) , . . . , θ n−1 (x) where θ n (x) = x.<br />

<strong>The</strong>refore, the equation (7.14) was algebraically solvable by ABEL’S third theorem<br />

(here theorem 6), and ABEL adapted theorem 8 to this particular equation, obtain-<br />

ing the same result as GAUSS had found. Furthermore, ABEL presented a pro<strong>of</strong> <strong>of</strong><br />

the result, which GAUSS had only announced, that the square root extracted in step 3<br />

could always be made to equal √ 2n + 1 (in ABEL’S variables).<br />

<strong>The</strong> contents <strong>of</strong> ABEL’S Mémoire sur une classe particulière can be summarized in the<br />

following five points depicting a descent from the general to the particular:


152 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

1. A general study <strong>of</strong> equations in which one root depended rationally on another.<br />

2. A restriction to irreducible equations and an application <strong>of</strong> the concept <strong>of</strong> ir-<br />

reducibility to prove that if x and θ (x) were roots <strong>of</strong> the irreducible equation<br />

φ (x) = 0, then so was θ k (x) for all integers k.<br />

3. A study <strong>of</strong> equations <strong>of</strong> degree µ = m × n in which the result was obtained<br />

that the solution <strong>of</strong> such equations could be reduced to solving m algebraically<br />

solvable equations <strong>of</strong> degree n and a single (generally unsolvable) equation <strong>of</strong><br />

degree m.<br />

4. An application <strong>of</strong> these — and other — results to the class <strong>of</strong> <strong>Abel</strong>ian equations,<br />

and a demonstration that these were always solvable by radicals.<br />

5. A further application <strong>of</strong> this result to the circular functions by which GAUSS’<br />

results on the cyclotomic equation were reproduced.<br />

ABEL had further ideas for applications <strong>of</strong> this new theory to elliptic functions,<br />

but these were not printed on this occasion (see below). In his research on <strong>Abel</strong>ian<br />

equations, KRONECKER much later came to the conclusion that “these general <strong>Abel</strong>ian<br />

equations in reality are nothing but cyclotomic equations.” 18 ABEL’S paper contains,<br />

however, more than just the solubility-result for <strong>Abel</strong>ian equations, and the general<br />

theory <strong>of</strong> the class <strong>of</strong> equations with rationally dependent roots sprung from — and<br />

had quite interesting implications for — ABEL’S approach to the theory <strong>of</strong> elliptic func-<br />

tions.<br />

7.2 Elliptic functions<br />

In his very first publication on elliptic functions entitled Recherches sur les fonctions<br />

elliptiques, 19 ABEL made several interesting innovations. 20 ABEL devoted a large por-<br />

tion <strong>of</strong> the first part <strong>of</strong> the Recherches to the inversion <strong>of</strong> elliptic integrals into elliptic<br />

functions, the extension <strong>of</strong> these functions into the complex domain, and the study<br />

<strong>of</strong> algebraic relations involving these functions. He derived addition formulae and<br />

studied the singularities <strong>of</strong> elliptic functions in order to address the central problem,<br />

which can be summarized in the following way:<br />

Problem 1 (Division Problem) Given an integer m and the value φ (mβ) <strong>of</strong> an elliptic<br />

function <strong>of</strong> the first kind, φ, at mβ, express φ (β) by radicals. ✷<br />

18 “[. . . ] so daß dise allgemeinen <strong>Abel</strong>schen Gleichungen im Wesentlichen nichts Anderes sind, als<br />

Kreistheilungs-Gleichungen.” (Kronecker, 1853, 11).<br />

19 (N. H. <strong>Abel</strong>, 1827b).<br />

20 <strong>The</strong> history <strong>of</strong> these elliptic functions and ABEL’S works on them is studied in much greater depth<br />

in part IV. For the present discussion, I am only concerned with the ideas behind ABEL’S result on<br />

the solubility <strong>of</strong> <strong>Abel</strong>ian equations.


7.2. Elliptic functions 153<br />

Figure 7.1: ABEL’S drawing <strong>of</strong> the lemniscate in one <strong>of</strong> his notebooks. (Stubhaug, 1996,<br />

270)<br />

ABEL’S inspirations for this problem were tw<strong>of</strong>old. <strong>The</strong> case in which m = 2 and φ<br />

was the lemniscate function useful in measuring the arc length <strong>of</strong> the lemniscate curve<br />

(see figure 7.1)<br />

φ (x) =<br />

� x<br />

0<br />

dx<br />

√ 1 − x 4<br />

had been settled in the eighteenth century by G. C. FAGNANO DEI TOSCHI (1682–<br />

1766). 21 In his study <strong>of</strong> the equivalent problem for circular functions GAUSS had<br />

expressed his conviction that his approach would apply equally well to other tran-<br />

scendentals, for instance the lemniscate integral (see the quotation in section 5.3.1, p.<br />

74).<br />

ABEL had learned <strong>of</strong> FAGNANO DEI TOSCHI’S work and the tradition in research<br />

on elliptic integrals through his studies <strong>of</strong> the much more advanced works on the<br />

subject by EULER and A.-M. LEGENDRE (1752–1833). 22 Complementary to his gen-<br />

eralization <strong>of</strong> FAGNANO DEI TOSCHI’S result to the bisection <strong>of</strong> elliptic functions <strong>of</strong><br />

the first kind, ABEL gave a detailed investigation <strong>of</strong> the division <strong>of</strong> such functions<br />

into 2n + 1 parts. Reformulated in the light <strong>of</strong> the addition formulae, which he had<br />

previously developed, ABEL obtained a different version <strong>of</strong> the problem, summarized<br />

in:<br />

Problem 2 (Division Problem) Given n, solve the equation<br />

φ ((2n + 1) β) = P2n+1 (φ (β))<br />

Q2n+1 (φ (β))<br />

which has degree (2n + 1) 2 . ✷<br />

ABEL’S central insight was that the equation <strong>of</strong> degree (2n + 1) 2 could be reduced<br />

to lower degree equations which were always solvable if the divisions <strong>of</strong> the periods<br />

21 (Houzel, 1986, 298).<br />

22 See chapter 15.


154 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

<strong>of</strong> the elliptic function were known. Addressing this division <strong>of</strong> the complete periods,<br />

ABEL demonstrated, directly inspired by GAUSS, that the roots<br />

φ 2<br />

� kω ′<br />

2n + 1<br />

�<br />

for 1 ≤ k ≤ n<br />

could be found by solving an equation <strong>of</strong> degree 2n + 2 which might not, however, be<br />

solvable by radicals.<br />

In the second part <strong>of</strong> the Recherches sur les fonctions elliptiques which appeared in<br />

1828, 23 ABEL applied the preceding investigation to the lemniscate integral. In com-<br />

plete correspondence with GAUSS’ result for the division <strong>of</strong> the circle, ABEL stated his<br />

result, using n in two different meanings:<br />

“<strong>The</strong> value <strong>of</strong> the function φ � �<br />

mω [the lemniscate function] can be expressed<br />

n<br />

by square roots whenever n is a number <strong>of</strong> the form 2n or 1 + 2n , the latter number<br />

being prime, or a product <strong>of</strong> multiple numbers <strong>of</strong> these two forms.” 24<br />

<strong>The</strong>refore, the division <strong>of</strong> the lemniscate into n equal parts could always be con-<br />

structed by ruler and compass if n was a number <strong>of</strong> the described form.<br />

In the Recherches sur les fonctions elliptiques, ABEL used direct methods to reduce the<br />

degrees and prove the solubility <strong>of</strong> the involved equations. However, as he soon re-<br />

alized, these properties depended on a deeper relation between the roots <strong>of</strong> the equa-<br />

tions, and in his letters he considered the division <strong>of</strong> the lemniscate as a by-product<br />

<strong>of</strong> his research in the theory <strong>of</strong> equations. 25 As ABEL indicated in the introduction to<br />

the Mémoire sur une classe particulière, he had planned to apply the theory concerning<br />

these equations to elliptic functions:<br />

“After having presented this theory in its generality, I will apply it to circular<br />

and elliptic functions.” 26<br />

Although no explicit application ever appeared in print (see section 7.2.1), it is not<br />

hard to see that for instance the equation<br />

n �<br />

∏ X − φ<br />

k=1<br />

2<br />

�<br />

kω ′ ��<br />

= 0<br />

2n + 1<br />

falls into the category studied in the general theory because <strong>of</strong> the rational dependency<br />

expressed by the addition formulae for φ.<br />

23 (N. H. <strong>Abel</strong>, 1828b).<br />

24 “La valeur de la fonction φ � �<br />

mω<br />

n peut être exprimée par des racines carrées toutes les fois que n<br />

est un nombre de la forme 2n ou 1 + 2n , le dernier nombre étant premier, ou même un produit de<br />

plusieurs nombres de ces deux formes.” (ibid., 168).<br />

25 (<strong>Abel</strong>→Holmboe, Paris, 1826/12. N. H. <strong>Abel</strong>, 1902a, 52) and (<strong>Abel</strong>→Holmboe, Berlin, 1827/03/04.<br />

ibid., 57).<br />

26 “Après avoir presenté généralement cette théorie, je l’appliquerai aux fonctions circulaires et elliptiques.”<br />

(N. H. <strong>Abel</strong>, 1829c, 132).


7.2. Elliptic functions 155<br />

7.2.1 <strong>The</strong> lost sections<br />

<strong>The</strong> paper Mémoire sur une classe particulière, 27 which was published in the second is-<br />

sue <strong>of</strong> the fourth volume <strong>of</strong> A. L. CRELLE’S (1780–1855) Journal appearing on March<br />

28 th 1829, i.e. a few days before ABEL’S death, was not complete. At the end <strong>of</strong> the<br />

published part, following the application to circular functions, CRELLE added a foot-<br />

note:<br />

“<strong>The</strong> author <strong>of</strong> this treatise will, on another occasion, present applications to<br />

elliptic functions.” 28<br />

At the end <strong>of</strong> ABEL’S manuscript for the Mémoire sur une classe particulière, 29 the<br />

opening page <strong>of</strong> a sixth — not printed — section entitled “Application aux fonctions<br />

elliptiques” can still be found (see figure 7.2). In the limited space <strong>of</strong> this one page,<br />

ABEL outlined the link with the Recherches sur les fonctions elliptiques. Its purpose was<br />

to facilitate the application <strong>of</strong> his newly developed theory to the division problem.<br />

From a letter to CRELLE — which ABEL wrote in October 1828 — it becomes clear that<br />

ABEL had sent a manuscript including the application to elliptic functions to CRELLE<br />

for publication in the Journal. 30 Because <strong>of</strong> his intense competition with C. G. J. JA-<br />

COBI (1804–1851) on elliptic functions, 31 ABEL urged CRELLE to rush publication <strong>of</strong> his<br />

sketch <strong>of</strong> his general theory <strong>of</strong> elliptic functions, the Précis d’une théorie des fonctions el-<br />

liptiques. 32 ABEL wanted CRELLE to delay the publication <strong>of</strong> the Mémoire sur une classe<br />

particulière, which had been scheduled for publication in the first issue <strong>of</strong> the fourth<br />

volume, and to leave out the part concerning the application to elliptic functions. 33<br />

CRELLE followed ABEL’S desire and published the Mémoire sur une classe particulière<br />

in the second issue. <strong>The</strong> Précis d’une théorie des fonctions elliptiques was published in<br />

the third and fourth issues <strong>of</strong> the fourth volume <strong>of</strong> the Journal, concluding a volume<br />

in which ABEL had published repeatedly on elliptic functions.<br />

Unfortunately, CRELLE’S correspondence and Nachlass appears to have been lost. 34<br />

<strong>The</strong>refore, little hope remains <strong>of</strong> finding the lost sections <strong>of</strong> ABEL’S paper. Neverthe-<br />

less, some information on their contents can be reconstructed from two sources: a<br />

notebook entry and the paper Précis d’une théorie des fonctions elliptiques.<br />

In one <strong>of</strong> ABEL’S notebooks, a brief list <strong>of</strong> contents <strong>of</strong> the manuscript Mémoire sur<br />

une classe particulière was found. 35 It was intended for ABEL’S own use and carried cor-<br />

rectly numbered references to central formulae and results, both in the published and<br />

27 (ibid.), described above.<br />

28 “L’auteur de ce mémoire donnera dans une autre occasion des applications aux fonctions elliptiques.”<br />

(ibid., 156, footnote).<br />

29 (<strong>Abel</strong>, MS:592, 64).<br />

30 (<strong>Abel</strong>→Crelle, Christiania, 1828/10/18. Biermann, 1967, 28–29)<br />

31 See part IV.<br />

32 (N. H. <strong>Abel</strong>, 1829d)<br />

33 This letter published in 1967 thus settles the speculation <strong>of</strong> SYLOW as to why these parts were not<br />

published (L. Sylow, 1902, 18).<br />

34 See footnote 54 on 33.<br />

35 (<strong>Abel</strong>, MS:351:C, 52). See also (N. H. <strong>Abel</strong>, 1881, II, 310–311) and (L. Sylow, 1902, 7–8).


156 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

Figure 7.2: <strong>The</strong> last page <strong>of</strong> ABEL’S manuscript for Mémoire sur une classe particulière<br />

(<strong>Abel</strong>, MS:592, 94) with the crossed out beginning <strong>of</strong> a sixth section. See also (N. H.<br />

<strong>Abel</strong>, 1881, II, 313).


7.3. <strong>The</strong> concept <strong>of</strong> irreducibility at work 157<br />

in the missing sections. Thus it must have been produced shortly before the manu-<br />

script was sent to CRELLE. Apparently, the manuscript contained two further sections<br />

besides the five published ones. In the sixth section, concerning the application to<br />

elliptic functions, ABEL listed the result that if<br />

was integral, the complex division<br />

m 2 + 2n + 1<br />

2µ + 1<br />

φ (m − αi) ω<br />

2n + 1<br />

could be effected by solving a µ th degree equation. 36 This is a generalized version<br />

<strong>of</strong> the division problem treated above. <strong>The</strong> seventh section concerned the transfor-<br />

mations <strong>of</strong> elliptic functions; a topic which also constituted a major part <strong>of</strong> ABEL’S<br />

competition with JACOBI. In the notebook, ABEL listed a number <strong>of</strong> formulae captur-<br />

ing central results. In order to produce a reliable interpretation, ABEL’S works in the<br />

transformation theory <strong>of</strong> elliptic functions have to be taken into consideration. 37<br />

7.3 <strong>The</strong> concept <strong>of</strong> irreducibility at work<br />

Of central importance to ABEL’S research in the theory <strong>of</strong> equations was his use <strong>of</strong><br />

the concepts <strong>of</strong> irreducibility and divisibility. In his Disquisitiones arithmeticae, GAUSS<br />

devoted a paragraph to the following result concerning the equation X = 0 which<br />

corresponded to the system Ω <strong>of</strong> imaginary n th roots <strong>of</strong> unity:<br />

“<strong>The</strong>ory <strong>of</strong> the roots <strong>of</strong> the equation x n − 1 = 0 (where n is assumed to be<br />

prime).<br />

Except for the root 1, the remaining roots contained in (Ω) are included in the<br />

equation X = x n−1 + x n−2 +etc.+x + 1 = 0.<br />

<strong>The</strong> function X cannot be decomposed into lower factors in which all the coefficients<br />

are rational.” 38<br />

GAUSS demonstrated the indecomposibility <strong>of</strong> X by an ad hoc argument and did<br />

not put it to central use later in the pro<strong>of</strong> (section 5.3). In ABEL’S impossibility pro<strong>of</strong>,<br />

numerous allusions to irreducibility had been made; however, they all served as sim-<br />

plifications and not as central concepts (see section 6.3.3). 39 By 1829 ABEL promoted<br />

the concept into a fundamental one on which theorems could be built. ABEL’S defini-<br />

tion <strong>of</strong> irreducibility was intended to capture the same property as GAUSS had demon-<br />

strated for X, although ABEL spoke <strong>of</strong> irreducible equations where GAUSS had spoken<br />

36 (<strong>Abel</strong>, MS:351:C, 52).<br />

37 This theme is taken up in part IV.<br />

38 “<strong>The</strong>oria radicum aequationis x n − 1 = 0 (ubi supponitur, n esse numerum primum).<br />

Omittendo radicem 1, reliquae (Ω) continentur in aequatione X = x n−1 + x n−2 + etc. +x + 1 = 0.<br />

Functio X resolvi nequit in factores inferiores, in quibus omnes coefficients sint rationales.” (C. F.<br />

Gauss, 1801, 417); English translation from (C. F. Gauss, 1986, 412).<br />

39 (N. H. <strong>Abel</strong>, 1826a, 71).


158 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

<strong>of</strong> indecomposable functions. This switch from polynomial functions to their associated<br />

equations was not uncommon, and is mainly a distinction in terms. ABEL gave his first<br />

definition <strong>of</strong> irreducibility in a footnote in the paper on <strong>Abel</strong>ian equations:<br />

“An equation φx = 0, in which the coefficients are rational functions <strong>of</strong> a<br />

certain number <strong>of</strong> known quantities a, b, c, . . . , is called irreducible when it is impossible<br />

to express any <strong>of</strong> its roots by an equation <strong>of</strong> lower degree, in which the<br />

coefficients are also rational functions <strong>of</strong> a, b, c, . . . .” 40<br />

<strong>The</strong> first — and highly useful — theorem which ABEL demonstrated with this defi-<br />

nition was that no equation could share a root with an irreducible one without having<br />

all the roots <strong>of</strong> the irreducible equation as roots. In the following section, I describe<br />

ABEL’S pro<strong>of</strong> and use <strong>of</strong> this important theorem.<br />

7.3.1 EUCLID’s division algorithm<br />

Formulated in the terminology <strong>of</strong> the Mémoire sur une classe particulière, the central<br />

theorem on irreducible equations was the following one expressing the property de-<br />

scribed above:<br />

<strong>The</strong>orem 9 “If one <strong>of</strong> the roots <strong>of</strong> an irreducible equation, φx = 0, satisfies another equation,<br />

f x = 0, where f x denotes a rational function <strong>of</strong> x and known quantities which are supposed<br />

contained in φx, this latter equation will also be satisfied if instead <strong>of</strong> x any other root <strong>of</strong> the<br />

equation φx = 0 is inserted.” 41<br />

ABEL gave a pro<strong>of</strong> <strong>of</strong> this theorem — again relegated to a footnote — which is a<br />

beautiful application <strong>of</strong> the division algorithm much along the lines <strong>of</strong> a modern ar-<br />

gument. Because f was a rational function, ABEL could write it as<br />

f = M<br />

, (7.15)<br />

N<br />

where M and N were entire functions <strong>of</strong> x. But, as ABEL noticed, “any [polynomial]<br />

function <strong>of</strong> x can always be put on the form P + Q · φx where P and Q are entire func-<br />

tions such that the degree <strong>of</strong> P is less than that <strong>of</strong> the function φx.” 42 This application<br />

<strong>of</strong> the division algorithm with remainder was well known to ABEL and received no<br />

40 “Une équation φx = 0, dont les coefficients sont des fonctions rationnelles d’un certain nombre<br />

de quantités connues a, b, c, . . . s’appelle irréductible, lorsqu’il est impossible d’exprimer aucune<br />

de ses racines par une équation moins élevée, dont les coefficiens soient également des fonctions<br />

rationnelles de a, b, c, . . . .” (N. H. <strong>Abel</strong>, 1829c, 132, footnote).<br />

41 “Si une des racines d’une équation irréductible φx = 0 satisfait à une autre équation f x = 0, où f x<br />

désigne une fonction rationnelle de x et des quantités connues qu’on suppose contenues dans φx;<br />

cette dernière équation se trouvera encore satisfaite en mettant au lieu de x une racine quelconque<br />

de l’équation φx = 0.” (ibid., 133).<br />

42 “mais une fonction de x peut toujours être mise sous la forme P + Q · φx, ou P et Q sont des fonctions<br />

entières, telles, que le degré de P soit moindre que celui de la fonction φx.” (ibid., 132–133, footnote).


7.3. <strong>The</strong> concept <strong>of</strong> irreducibility at work 159<br />

further comment. 43 By inserting into (7.15), ABEL found<br />

f (x) =<br />

P + Q · φ (x)<br />

. (7.16)<br />

N<br />

Next, he let x denote a common root <strong>of</strong> φ and f and concluded that x would also be<br />

a root <strong>of</strong> P = 0. However, if P were not identically zero, “this equation gives x as a<br />

root <strong>of</strong> an equation <strong>of</strong> degree less than that <strong>of</strong> φx = 0; which is a contradiction <strong>of</strong> the<br />

hypothesis. <strong>The</strong>refore, P = 0 and it follows that f x = φx · Q<br />

N .”44 Thus, it was obvious<br />

that f would vanish whenever φ did and, therefore, that any root <strong>of</strong> φ (x) = 0 would<br />

also be a root <strong>of</strong> f (x) = 0.<br />

ABEL put this important theorem to use in the very first description <strong>of</strong> the equa-<br />

tions treated in the Mémoire sur une classe particulière. If x ′ and x were two roots <strong>of</strong> the<br />

irreducible equation φ (x) = 0 among which a rational dependency existed,<br />

x ′ = θ (x) ,<br />

then every iterated application <strong>of</strong> θ to x would also be a root <strong>of</strong> this equation. ABEL’S<br />

demonstration followed directly from the theorem above. He argued that since it fol-<br />

lowed from the hypothesis that the equations<br />

φ (θ (x)) = 0 and φ (x) = 0<br />

had a root, x, in common, theorem 7.3 stated that for any root, y, <strong>of</strong> φ (x) = 0, θ (y)<br />

would also be a root <strong>of</strong> that equation. Once he had established this result, the argu-<br />

ment <strong>of</strong> ABEL’S paper was on its way, and the complex <strong>of</strong> conclusions described above<br />

could be obtained.<br />

ABEL turned the concept <strong>of</strong> irreducibility <strong>of</strong> equations, which had existed as an<br />

ad hoc tool before into a central foundation upon which a building <strong>of</strong> theorems could<br />

be established. 45 <strong>The</strong> irreducibility in ABEL’S sense was defined as minimality <strong>of</strong> the<br />

equation expressing the roots under the restriction that the coefficients must depend<br />

rationally on the same quantities as the original equation. From this definition, gen-<br />

eralizations were later made toward the general concept <strong>of</strong> domain <strong>of</strong> rationality. But<br />

working with this definition — and the division algorithm <strong>of</strong> EUCLID (∼295 B.C.) —<br />

ABEL demonstrated the important theorem 9 <strong>of</strong> divisibility, which in turn established<br />

the basic property <strong>of</strong> the class <strong>of</strong> equations studied in the Mémoire sur une classe particulière.<br />

46<br />

43 It had been explicitly employed in GAUSS’ second pro<strong>of</strong> <strong>of</strong> the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra (see<br />

section 5.7).<br />

44 “cette équation donnera x, comme racine d’une équation d’un degré moindre que celui de φx = 0;<br />

ce qui est contre l’hypothèse; donc P = 0 et par suite f x = φx · Q<br />

45<br />

N .” (ibid., 133,footnote).<br />

See also (L. Sylow, 1902, 23–24).<br />

46 (N. H. <strong>Abel</strong>, 1829c).


160 Chapter 7. Particular classes <strong>of</strong> solvable equations<br />

❅❅■<br />

✻<br />

✛ Solvable equations ✲<br />

��✠<br />

❄<br />

Polynomial equations<br />

Figure 7.3: Extending the class <strong>of</strong> solvable equations: <strong>Abel</strong>ian equations<br />

7.4 Enlarging the class <strong>of</strong> solvable equations<br />

ABEL considered the positive result demonstrating the solubility <strong>of</strong> certain equations<br />

as a counterpart to the insolubility <strong>of</strong> higher degree general equations. In the intro-<br />

duction to the Mémoire sur une classe particulière, ABEL wrote:<br />

“It is true that the algebraic equations are not generally solvable, but there is a<br />

particular class <strong>of</strong> each degree for which the algebraic solution is possible.” 47<br />

To this class <strong>of</strong> solvable equations belonged the equations <strong>of</strong> the form x n − 1 = 0<br />

studied by GAUSS and the generalizations <strong>of</strong> these obtained by ABEL in the paper.<br />

Only few other equations were explicitly known to be solvable, and ABEL’S result<br />

can thus be seen to provide a demonstration that the total class <strong>of</strong> solvable equations<br />

had a certain range. In the limitation-enlargement model suggested in section 6.8,<br />

the situation can be described by figure 7.3 and much <strong>of</strong> ABEL’S research to describe<br />

the precise extent <strong>of</strong> solubility can be interpreted in this context. In a letter to B. M.<br />

HOLMBOE (1795–1850) written during his stay in Paris, ABEL described the problem<br />

and his progress:<br />

“I am currently working on the theory <strong>of</strong> equations, which is my favorite<br />

theme, and have finally reached a point where I see a way to solve the following<br />

general problem: To determine the form <strong>of</strong> all algebraic equations which can<br />

be solved algebraically. I have found an infinitude <strong>of</strong> the fifth, sixth, seventh, etc.<br />

degree which had never been smelled before.” 48<br />

47 “Il est vrai que les équations algébriques ne sont pas résolubles généralement; mais il y en a une<br />

classe particulière de tous les degrés dont la résolution algébrique est possible.” (N. H. <strong>Abel</strong>, 1829c,<br />

131).<br />

� �✒<br />

❅ ❅❘


7.4. Enlarging the class <strong>of</strong> solvable equations 161<br />

Thus, ABEL’S two publications on the theory <strong>of</strong> equations which appeared during<br />

his lifetime contributed a negative, limiting result <strong>of</strong> insolubility <strong>of</strong> the general higher<br />

degree equations and a positive, enlarging result <strong>of</strong> the solubility <strong>of</strong> a certain class <strong>of</strong><br />

equations <strong>of</strong> all degrees. <strong>The</strong> program set out above in the letter to HOLMBOE was<br />

pursued by ABEL from his time in Paris, and traces <strong>of</strong> it can be found in his note-<br />

books. However, his correspondence also announced further and far-reaching results<br />

for which no detailed studies or pro<strong>of</strong>s have been recovered. <strong>The</strong> determination <strong>of</strong><br />

the exact extension <strong>of</strong> the concept <strong>of</strong> algebraic solubility was approached by ABEL<br />

through a theory largely based on the same tools as his published works, but never<br />

completed nor published in his lifetime. <strong>The</strong>refore, the solution to this fundamental<br />

problem is rightfully attributed to E. GALOIS (1811–1832). In the next chapter, ABEL’S<br />

steps toward a general theory <strong>of</strong> solubility are analyzed against the background <strong>of</strong> his<br />

other works and GALOIS’ contemporary ideas.<br />

48 “Jeg arbeider nu paa Ligningernes <strong>The</strong>orie, mit Yndlingsthema og er endelig kommen saa vidt at jeg<br />

seer Udvei til at løse følgende alm: Problem. Determiner la forme de toutes les équation algébriques<br />

qui peuvent être resolues algebriquement. Jeg har fundet en uendelig Mængde af 5te, 6te, 7de etc.<br />

Grad som man ikke har lugtet indtil nu.” (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. N. H. <strong>Abel</strong>, 1902a,<br />

44).


Chapter 8<br />

A grand theory in spe: algebraic<br />

solubility<br />

In his correspondence with A. L. CRELLE (1780–1855) and B. M. HOLMBOE (1795–<br />

1850), N. H. ABEL (1802–1829) announced numerous results in the theory <strong>of</strong> equa-<br />

tions beyond the impossibility <strong>of</strong> solving the quintic and the study <strong>of</strong> <strong>Abel</strong>ian equa-<br />

tions. Some concerned the form <strong>of</strong> solutions to algebraically solvable equations <strong>of</strong> the<br />

fifth degree, 1 others dealt with solubility results for broader classes <strong>of</strong> equations, 2 and<br />

yet others testify to ABEL’S general progress in his program <strong>of</strong> determining the form<br />

<strong>of</strong> solvable equations. 3 <strong>The</strong> information provided in the letters is complemented by<br />

a notebook entry dating from 1828 which has been included in both editions <strong>of</strong> the<br />

Œuvres under the title Sur la résolution algébrique des équations. 4 <strong>The</strong> entry begins as a<br />

manuscript almost ready for press, but after some introductory remarks, a few theo-<br />

rems and some deductions it turns from its initial thoroughness and clarity to nothing<br />

but bare calculations. Nevertheless, when considered together, these sources give an<br />

impression <strong>of</strong> the methods and extent <strong>of</strong> the general theory <strong>of</strong> algebraic solubility<br />

which ABEL set out to develop in the last years <strong>of</strong> his life.<br />

8.1 Inverting the approach once again<br />

<strong>The</strong> notebook manuscript dealt with the general form <strong>of</strong> algebraically solvable equa-<br />

tions. In one <strong>of</strong> the two lengthy introductions which ABEL wrote for this work the<br />

problem was clearly set out:<br />

“Given an equation <strong>of</strong> any given degree, to determine whether or not it could<br />

be satisfied algebraically.” 5<br />

1 (<strong>Abel</strong>→Crelle, Freyberg, 1826/03/14. N. H. <strong>Abel</strong>, 1902a, 21–22).<br />

2 (<strong>Abel</strong>→Crelle, Christiania, 1828/08/18. ibid., 72–73).<br />

3 (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. ibid., 44–45).<br />

4 (N. H. <strong>Abel</strong>, [1828] 1839).<br />

5 “Une équation d’un degré quelconque étant proposée, reconnaître si elle pourra être satisfaite algébriquement,<br />

ou non.” (N. H. <strong>Abel</strong>, 1881, vol. 2, 330).<br />

163


164 Chapter 8. A grand theory in spe<br />

ABEL’S initial step in solving this general problem was to reformulate it in the<br />

following program which he described in the introduction to the other version <strong>of</strong> the<br />

manuscript.<br />

“From this, the following two problems stem naturally whose complete solution<br />

comprises the entire theory <strong>of</strong> the algebraic solution <strong>of</strong> equations, namely:<br />

1) To find all equations <strong>of</strong> any determinate degree which are algebraically solvable.<br />

2) To decide whether or not a given equation is algebraically solvable.” 6<br />

Thus, the problem <strong>of</strong> determining the algebraic solubility <strong>of</strong> equations had been<br />

inverted once again. In principle, ABEL’S program amounted to listing — by some de-<br />

scriptive form — all equations <strong>of</strong> a certain degree which could be solved algebraically<br />

and then deducing whether any given equation was in this list. In pursuing this prob-<br />

lem, ABEL focused on a given algebraic expression, a radical, and sought to describe<br />

the irreducible equation which it satisfied. In doing so, the concept <strong>of</strong> irreducibility ac-<br />

quired its second importance in ABEL’S research as a means <strong>of</strong> obtaining the equation<br />

linked to a given radical. This shift from working with equations <strong>of</strong> which some rep-<br />

resentation was known, either as a fifth degree polynomial or as relations among its<br />

roots, to general equations which were only characterized by their external structure<br />

as being irreducible is what I consider a second inversion <strong>of</strong> approach.<br />

ABEL’S inversion was intimately connected to a general consideration on mathe-<br />

matical methodology. In his introduction, he described this inversion <strong>of</strong> approach in<br />

a much quoted paragraph:<br />

“To solve these equations [<strong>of</strong> the first four degrees], a uniform method was<br />

discovered which, it was thought, was applicable to an equation <strong>of</strong> any degree;<br />

but in spite <strong>of</strong> all the efforts <strong>of</strong> a Lagrange and other distinguished geometers, the<br />

proposed goal could not be reached. This led to the assumption that the solution <strong>of</strong><br />

the general equation was algebraically impossible; but this could not be decided<br />

since the adopted method had only been able to lead to reliable conclusions in<br />

the case in which the equations were solvable. In fact, one proposed to solve the<br />

equations without knowing if that was possible. In this case, one might come to<br />

the solution although that was not certain at all; but if by misfortune the solution<br />

was impossible, one might search an eternity without finding it. To infallibly reach<br />

anything in this matter, it is necessary to follow another route. One should give the<br />

problem such a form that it will always be possible to solve it, which can always<br />

be done for any problem. Instead <strong>of</strong> demanding a relation, <strong>of</strong> which the existence<br />

is unknown, one should ask whether such a relation is possible at all.” 7<br />

6 “De là dérivent naturellement les deux problèmes suivans, dont la solution complète comprend<br />

toute la théorie de la résolution algébrique des équations, savoir:<br />

1) Trouver toutes les équations d’un degré déterminé quelconque qui soient résolubles algébriquement.<br />

2) Juger si une équation donnée est résoluble algébriquement, ou non.”<br />

(N. H. <strong>Abel</strong>, [1828] 1839, 218–219).<br />

7 “On découvrit pour résoudre ces équations une méthode uniforme et qu’on croyait pouvoir appliquer<br />

à une équation d’un degré quelconque; mais malgré tous les efforts d’un Lagrange et d’autres


8.2. Construction <strong>of</strong> the irreducible equation 165<br />

As previously noted, ABEL’S belief that any problem could be converted into a<br />

solvable one was held by most mathematicians throughout the 19 th century. It be-<br />

came prominent in the so-called Hilbert Programme before the development <strong>of</strong> axiomat-<br />

ics stressed that decidability could only be asked and answered relatively to the (ax-<br />

iomatic) system in which the problem was embedded.<br />

With ABEL’S new driving question, modified from the ones motivating the impos-<br />

sibility pro<strong>of</strong> and the study <strong>of</strong> <strong>Abel</strong>ian equations, it was his intention to explore the<br />

grey area between the entire set <strong>of</strong> equations and the ones known to be solvable (see<br />

figures 6.1 and 7.3). ABEL’S hope was to delineate the border line between solvable<br />

and unsolvable equations by some external characteristic.<br />

8.2 <strong>The</strong> construction <strong>of</strong> the irreducible equation<br />

satisfied by a given expression<br />

<strong>The</strong> general program <strong>of</strong> ABEL’S research was to construct a list <strong>of</strong> all irreducible solv-<br />

able equations and subsequently match any given equation against this list. His at-<br />

tempt at implementing this scheme consisted <strong>of</strong> a construction <strong>of</strong> the irreducible equa-<br />

tion satisfied by a given algebraic expression. After establishing certain properties <strong>of</strong><br />

this equation from the expression which satisfies it, ABEL returned to the problem <strong>of</strong><br />

determining whether a given equation was solvable or not.<br />

<strong>The</strong> first part <strong>of</strong> ABEL’S notebook manuscript contained theorems and results pre-<br />

sented in a clear and deductive manner. <strong>The</strong>ir contents showed frequent similarities<br />

with the opening studies <strong>of</strong> the form <strong>of</strong> algebraic expressions satisfying an equation as<br />

carried out in the impossibility pro<strong>of</strong> (see section 6.3.3). If anything, the 1828 notebook<br />

lacked — by comparison to the impossibility pro<strong>of</strong> — the clear, albeit defective, classi-<br />

fication <strong>of</strong> algebraic expressions <strong>of</strong> which only reminiscences were given. <strong>The</strong> clas-<br />

sification established in the notebook was insufficient to cover some <strong>of</strong> the required<br />

deductions, and it is possible that ABEL, himself, had noticed this deficiency (see be-<br />

low).<br />

Basic concepts. In the opening section <strong>of</strong> the manuscript proper (following a lengthy<br />

introduction), ABEL outlined his own characterization <strong>of</strong> algebraic expressions which<br />

géomètres distingués on ne put parvenir au but proposé. Cela fit présumer que la résolution des<br />

équations générales était impossible algébriquement; mais c’est ce qu’on ne pouvait pas décider,<br />

attendu que la méthode adoptée n’aurait pu conduire à des conclusions certaines que dans le cas<br />

où les équations étaient résolubles. En effet on se proposait de résoudre les équations, sans savoir<br />

si cela était possible. Dans ce cas, on pourrait bien parvenir à la résolution, quoique cela ne fût<br />

nullement certain; mais si par malheur la résolution était impossible, on aurait pu la chercher une<br />

éternité, sans la trouver. Pour parvenir infailliblement à quelque chose dans cette matière, il faut<br />

donc prendre une autre route. On doit donner au problème une forme telle qu’il soit toujours possible<br />

de le résoudre, ce qu’on peut toujours faire d’une problème quelconque. Au lieu de demander<br />

une relation dont on ne sait pas si elle existe ou non, il faut demander si une telle relation est en effet<br />

possible.” (ibid., 217).


166 Chapter 8. A grand theory in spe<br />

could occur in the solution <strong>of</strong> a solvable equation. This characterization had been one<br />

<strong>of</strong> the points <strong>of</strong> objection to his impossibility pro<strong>of</strong> <strong>of</strong> 1826 (see section 6.9.1). However,<br />

only the objections raised by E. J. KÜLP (⋆1801) were known to ABEL and ABEL did<br />

not react directly to them in the notebook. <strong>The</strong> characterization which ABEL presented<br />

in the notebook was only a limited version <strong>of</strong> the one found in the impossibility pro<strong>of</strong>.<br />

In the notebook, ABEL described the radicals from the outer-most one inward in the<br />

following form<br />

1<br />

µ 1<br />

2<br />

µ 1<br />

y = P0 + P1 · R1<br />

+ P2 · R1<br />

+ · · · + Pµ 1−1 · R1<br />

, (8.1)<br />

in which P0, . . . , Pµ 1−1 and R1 were rational expressions in known quantities and the<br />

1 1<br />

µ 2 µ 3<br />

other radicals R2<br />

, R3<br />

, . . . . In relation to the route he had taken in the impossibility<br />

pro<strong>of</strong>, he abandoned the concept <strong>of</strong> degree <strong>of</strong> algebraic expressions and imposed only<br />

the hierarchy from the concept <strong>of</strong> order which counted the number <strong>of</strong> nested root<br />

extractions <strong>of</strong> prime degree.<br />

ABEL introduced three notational concepts which he used throughout the prelimi-<br />

nary part <strong>of</strong> the manuscript to simplify his notation:<br />

1. He chose to denote algebraic expressions by writing their order as subscripts, for<br />

µ 1 −1<br />

µ 1<br />

instance writing Am for an algebraic expression A <strong>of</strong> order m.<br />

2. With y being <strong>of</strong> the form (8.1) and φ (y) = 0 an equation satisfied by y, ABEL<br />

chose to write the equation as φ (y, m) = 0 if all the coefficients <strong>of</strong> φ (y) were<br />

algebraic expressions <strong>of</strong> order m. Furthermore, he denoted the degree <strong>of</strong> the<br />

equation by δφ (y, m).<br />

3. Most importantly, he introduced a symbol ∏ Am for the product <strong>of</strong> all values <strong>of</strong><br />

Am obtained from attributing to the outermost radical in Am, R 1 µ , all its possible<br />

values, R 1 µ , ωR 1 µ , . . . , ω µ−1 R 1 µ (ω a µ th root <strong>of</strong> unity). Thus, if<br />

Am =<br />

the new symbol denoted the expression<br />

∏ Am =<br />

µ−1<br />

∑ pkR k=0<br />

k µ ,<br />

�<br />

µ−1 µ−1<br />

∏ ∑ pkω u=0 k=0<br />

uk R k �<br />

µ .<br />

Using these concepts and a number <strong>of</strong> immediate consequences derived from them,<br />

ABEL constructed and characterized the irreducible equation associated with a given<br />

algebraic expression.


8.2. Construction <strong>of</strong> the irreducible equation 167<br />

Lemmata. In the first lemma, ABEL obtained a result which had played a central role<br />

in his impossibility pro<strong>of</strong>. It stated that if the equation<br />

µ−1<br />

∑<br />

u=0<br />

tuy<br />

u<br />

µ 1<br />

1<br />

= 0 (8.2)<br />

could be satisfied, in which the coefficients t0, . . . , tµ−1 were rational functions <strong>of</strong> ω,<br />

known quantities (i.e. coefficients <strong>of</strong> the equation φ (y) = 0), and lower order radicals,<br />

then all the coefficients had to vanish, i.e. t0 = t1 = · · · = tµ−1 = 0 (cmp. lemma 1).<br />

By and large, the pro<strong>of</strong> resembled the one given in 1826 (see section 6.3.3) but differed<br />

when ABEL had to eliminate the possibility <strong>of</strong> a first degree irreducible factor. Letting<br />

z = y 1<br />

µ 1 , ABEL assumed that the irreducible factor (corresponding to an irreducible<br />

equation) ∑ κ u=0 snz n divided (8.2) and excluded the possibility <strong>of</strong> k ≥ 2. <strong>The</strong> case <strong>of</strong> a<br />

first degree irreducible factor was briefly dismissed by the following argument:<br />

“Thus, it is necessary that k = 1, but that gives<br />

from which<br />

which is similarly impossible.” 8<br />

s0 + z = 0<br />

z = µ √<br />

1 y1 = −s0,<br />

As P. L. M. SYLOW (1832–1918) has remarked, the conclusion that z = −s0 is<br />

impossible is essentially correct, 9 it can be supported if an improved hierarchy is im-<br />

posed on the radicals. 10 Again, ABEL’S notebook does not contain all the technical<br />

details <strong>of</strong> his deductions.<br />

ABEL put forward another important proposition when he claimed that the roots<br />

<strong>of</strong> satisfiable equations come in “bundles”. He stated that if the equation<br />

was satisfied by an algebraic expression <strong>of</strong> order n<br />

8 “Il faut donc que k = 1, or cela donne<br />

d’où<br />

φ (y, m) = 0 (8.3)<br />

y =<br />

µ−1<br />

∑<br />

k=0<br />

s0 + z = 0<br />

p ky<br />

k<br />

µ 1<br />

1 ,<br />

z = µ √<br />

1 y1 = −s0,<br />

ce qui est de même impossible.” (N. H. <strong>Abel</strong>, [1828] 1839, 229).<br />

9 (Sylow in N. H. <strong>Abel</strong>, 1881, vol. 2, 332).<br />

10 As shown by (Holmboe in N. H. <strong>Abel</strong>, 1839, vol. 2, 289) and (Maser in <strong>Abel</strong> and Galois, 1889, 149).


168 Chapter 8. A grand theory in spe<br />

it would also be satisfied if ωu 1<br />

µ<br />

y1<br />

1<br />

µ<br />

were inserted for y1<br />

(ω a µth root <strong>of</strong> unity). He gave<br />

no explicit pro<strong>of</strong> <strong>of</strong> this result, which is a simple consequence <strong>of</strong> the vanishing <strong>of</strong> the<br />

coefficients <strong>of</strong> (8.2). 11 <strong>The</strong> result provided the important connection that any root <strong>of</strong><br />

∏ φ (y, m) = 0 would also be a root <strong>of</strong> φ (y, m) = 0.<br />

<strong>The</strong> manuscript also contains the fundamental characterization <strong>of</strong> irreducible equa-<br />

tions that no equation can share a root with an irreducible equation without the latter<br />

dividing the former (cmp. theorem 7.3). ABEL derived this along the lines described<br />

in section 7.3 but applied the terminology developed in the manuscript. By implicit<br />

application <strong>of</strong> the Euclidean division algorithm, ABEL demonstrated that if the equa-<br />

tions<br />

φ (y, m) = 0 and φ1 (y, n) = 0<br />

had a common root, φ was assumed to be irreducible, and n ≤ m, then<br />

φ1 (y, n) = φ (y, m) · f (y, m) .<br />

Properties <strong>of</strong> ∏ φ (y, m). In his subsequent argument, ABEL sought to describe the<br />

irreducible equation satisfied by a given algebraic expression. <strong>The</strong> central tool employed<br />

was his construction <strong>of</strong> this equation based on the construction <strong>of</strong> ∏ φ (y, m)<br />

and the demonstration <strong>of</strong> its properties. <strong>The</strong> construction which ABEL gave was<br />

mainly existential; it amounted to proving the existence <strong>of</strong> an equation having spe-<br />

cific useful properties.<br />

Continuing to build upon the fundamental result on irreducible equations, ABEL<br />

proved the following theorem.<br />

<strong>The</strong>orem 10 If<br />

then for some m ′<br />

φ1 (y, n) = f (y, m) · φ (y, m) ,<br />

φ1 (y, n) = f1<br />

� y, m ′ � · ∏ φ (y, m) . ✷<br />

ABEL’S pro<strong>of</strong> was elegant and made prototypical usage <strong>of</strong> the previously estab-<br />

lished theorems and the concept <strong>of</strong> the outer-most radical. Denoting by µ√ y1 the outer-<br />

most root extraction <strong>of</strong> φ (y, m) = 0, c.f. (8.1), this equation would also be satisfied<br />

if ω k µ √ y1 were substituted for µ√ y1 where ω was a µ th root <strong>of</strong> unity. Consequently,<br />

ω k µ √ y1 was a root <strong>of</strong> φ and, therefore, also <strong>of</strong> φ1. Thus, φ1 would have the different<br />

roots <strong>of</strong> φ corresponding to different values <strong>of</strong> k as roots. As ABEL noticed, if these<br />

factors corresponding to different values <strong>of</strong> k had no common factors (were relatively<br />

prime), their product would also be a factor <strong>of</strong> φ and the pro<strong>of</strong> had been completed.<br />

In the impossibility pro<strong>of</strong> <strong>of</strong> 1826, ABEL had stated this result, which translated into<br />

the notation <strong>of</strong> the manuscript concludes that “it is clear that the given equation must<br />

11 See for instance (Holmboe in N. H. <strong>Abel</strong>, 1839, vol. 2, 289).


8.2. Construction <strong>of</strong> the irreducible equation 169<br />

1<br />

µ<br />

be satisfied by all values <strong>of</strong> y which are obtained by attributing to y1<br />

1<br />

µ<br />

ωy1<br />

, ω2 1<br />

µ<br />

y1<br />

, . . . , ωn−1 1<br />

µ<br />

y<br />

all the values<br />

1 ”12 (see section 6.3.3). In 1826, it had been given no pro<strong>of</strong>, but<br />

in the notebook, ABEL provided the pro<strong>of</strong> as an easy and elegant application <strong>of</strong> the<br />

fundamental concepts and tools.<br />

ABEL proceeded by establishing a central link between the irreducibility <strong>of</strong> φ (y, m) =<br />

0 and that <strong>of</strong> ∏ φ (y, m) = 0.<br />

<strong>The</strong>orem 11 If the equation<br />

is irreducible, then so is the equation<br />

φ (y, m) = 0<br />

φ1 (y, m) = ∏ φ (y, m) = 0. ✷<br />

ABEL argued for this theorem by a reductio ad absurdum pro<strong>of</strong> against which SYLOW<br />

later raised well founded objections. ABEL assumed that φ1 was reducible and that<br />

φ2 (y, m ′ ) was an irreducible 13 factor <strong>of</strong> ∏ φ (y, m) = 0. Under these assumptions,<br />

φ2 and φ would have a common root since all the roots <strong>of</strong> ∏ φ were also roots <strong>of</strong><br />

φ. <strong>The</strong> assumed irreducibility <strong>of</strong> φ then enabled ABEL to conclude that because the<br />

irreducible φ and φ2 had a root in common, φ would be a factor <strong>of</strong> φ2,<br />

φ2<br />

This in turn implied (by theorem 10)<br />

φ2<br />

� y, m ′ � = f (y, m) · φ (y, m) .<br />

� y, m ′ � = f1<br />

� ′′<br />

y, m � · ∏ φ (y, m) . (8.4)<br />

� �� �<br />

=φ1(y,m) On the other hand, φ2 had been assumed to be an irreducible factor <strong>of</strong> φ1 implying<br />

deg φ2 < deg φ1, which contradicted (8.4).<br />

SYLOW’S objections concerned the properties <strong>of</strong> ∏ φ. Besides certain points, at<br />

which ABEL left out assumptions <strong>of</strong> irreducibility, SYLOW noticed that ABEL tacitly<br />

assumed that φ (y, m) did not have factors in which all the coefficients were rational<br />

expressions in inner radicals and known quantities. If such factors were involved, the<br />

equation ∏ φ (y, m) = 0 might turn out to be a power <strong>of</strong> an irreducible equation. 14<br />

SYLOW repaired ABEL’S argument by refining his hierarchy <strong>of</strong> algebraic expressions.<br />

12 “[. . . ] so ist klar, daß der gegebenen Gleichung durch alle die Werthe von y genug werden muß,<br />

welche man findet, wenn man der Größe p 1 n alle die Werthe αp 1 n , α 2 p 1 n , . . . , α n−1 p 1 n beilegt.” (N. H.<br />

<strong>Abel</strong>, 1826a, 72).<br />

13 Actually, ABEL did not, presumably inadvertently, state the condition <strong>of</strong> irreducibility <strong>of</strong> φ2.<br />

14 (Sylow in N. H. <strong>Abel</strong>, 1881, vol. 2, 332).


170 Chapter 8. A grand theory in spe<br />

Construction <strong>of</strong> the irreducible equation. With the first theorems and the lemmata<br />

described above, ABEL was in a position to give a construction <strong>of</strong> the irreducible equa-<br />

tion which a given algebraic expression satisfied. More importantly, this construction<br />

allowed him to demonstrate that central properties <strong>of</strong> this equation could be deduced<br />

from properties <strong>of</strong> the initially given algebraic expression. ABEL let<br />

am = f ( µm √ ym, µ √<br />

m−1 ym−1, . . . )<br />

denote a given algebraic expression and constructed the irreducible equation ψ (y) = 0<br />

which would have am as a root in the following way.<br />

Since am was to satisfy ψ (y) = 0, it would be necessary that y − am was a factor <strong>of</strong><br />

ψ. By the theorem 10, it followed that<br />

φ1 (y, m1) = ∏ (y − am)<br />

would also be a factor. Because y − am was a first degree polynomial and, therefore,<br />

irreducible, it followed that φ1 was also irreducible (by theorem 11). Consequently,<br />

φ1 was an irreducible factor <strong>of</strong> ψ (y) and the procedure could be repeated yielding a<br />

sequence <strong>of</strong> irreducible factors<br />

φn (y, mn) = ∏ φn−1 (y, mn−1) ,<br />

in which the radicals <strong>of</strong> am were sequentially removed by the analogue <strong>of</strong> multiplying<br />

with the complex conjugate (c.f. section 6.3.2).<br />

ABEL claimed that the sequence <strong>of</strong> positive integers m1, m2, . . . was decreasing but<br />

gave no explicit argument. However, by J. L. LAGRANGE’S (1736–1813) theorem (sec-<br />

tion 5.2.3) it is not hard to see that ∏ am is a rational function <strong>of</strong> ym and the inner<br />

radicals involved. <strong>The</strong>refore, the order <strong>of</strong> ∏ am is less than the order <strong>of</strong> am. Thus,<br />

at a certain point (after, say, u steps) the sequence m1, m2, . . . had to vanish, and an<br />

equation would be obtained in which all the coefficients were rationally known. This<br />

equation was the sought-for ψ (y) = 0,<br />

ψ (y) = φu (y, 0) = ∏ φu−1 (y, mu−1) .<br />

Directly from this construction, ABEL deduced his characterization <strong>of</strong> the irre-<br />

ducible equation satisfied by a given algebraic expression, laying the foundations for<br />

his further reasoning. He summarized the properties in the following four points: 15<br />

Proposition 1 <strong>The</strong> following four results link properties <strong>of</strong> the irreducible equation ψ (y) = 0<br />

satisfied by a given algebraic expression am to properties <strong>of</strong> the expression itself:<br />

1. <strong>The</strong> degree <strong>of</strong> ψ is the product <strong>of</strong> certain exponents <strong>of</strong> root extractions occurring in am.<br />

Among these exponents, the one <strong>of</strong> the outer-most root extractions is always present.<br />

15 (N. H. <strong>Abel</strong>, [1828] 1839, 232–233)


8.3. Refocusing on the equation 171<br />

2. <strong>The</strong> exponent <strong>of</strong> the outer-most root extraction divides the degree <strong>of</strong> ψ [actually contained<br />

in 1, above].<br />

3. If ψ can be algebraically satisfied, it is also algebraically solvable. All its roots are ob-<br />

tained by attributing to the root extractions y<br />

1<br />

µu<br />

mu<br />

all their possible values.<br />

4. If the degree <strong>of</strong> ψ is µ, the expression am may have µ, and no more than µ, values. ✷<br />

ABEL’S deduction <strong>of</strong> these properties was straightforward from considerations on<br />

the exponents <strong>of</strong> involved root extractions and the construction described. A formal<br />

consideration <strong>of</strong> the uniqueness <strong>of</strong> the irreducible equation constructed was not car-<br />

ried out, but must have seemed obvious to ABEL.<br />

8.3 Refocusing on the equation<br />

<strong>The</strong> first theorems and the construction <strong>of</strong> the irreducible equation connected to a<br />

given algebraic expression are fascinating pieces <strong>of</strong> mathematics revealing traces <strong>of</strong><br />

ABEL’S pr<strong>of</strong>ound ideas. Whereas the presentation <strong>of</strong> these fundamental results was<br />

lucid — and basically acceptable to present day mathematicians — ABEL’S following<br />

investigations in the notebook took another form. As he progressed farther from the<br />

well established results founded in the theory <strong>of</strong> LAGRANGE, his explanatory remarks<br />

and general narrative became ever more sparse until they finally ceased altogether.<br />

However, ABEL’S notebook is the only source illustrating how he planned to proceed,<br />

and I will try to reconstruct the central result <strong>of</strong> these investigations, which was never<br />

presented in a form intended for publication.<br />

Because ABEL’S argument, from this point onward, consists <strong>of</strong> little but equations,<br />

I have reconstructed how he could, with his tools and methods, have argued. In lim-<br />

iting myself to ABEL’S argument for the reduction <strong>of</strong> the general problem to <strong>Abel</strong>ian<br />

equations, I remain close to the sources. ABEL’S unfinished manuscript inspired math-<br />

ematicians <strong>of</strong> the nineteenth century — such as C. J. MALMSTEN (1814–1886), SYLOW,<br />

and L. KRONECKER (1823–1891) — to elaborate and extend the investigation; 16 recently,<br />

L GÅRDING and C. SKAU have taken up the problem anew. 17<br />

SYLOW has speculated that ABEL recognized the insufficiency <strong>of</strong> his description<br />

<strong>of</strong> algebraic expressions. In response to his realization, ABEL should, according to<br />

SYLOW, have abandoned his attempt at presenting a manuscript ready for printing<br />

and instead recorded his further findings in the order and form in which he came to<br />

them. 18 As also noted by SYLOW, this change in style <strong>of</strong> presentation was not uncom-<br />

mon. In his notebooks, ABEL frequently started out writing coherent manuscripts,<br />

16 (Malmsten, 1847), (Sylow, 1861), (L. Sylow, 1902, 18–22), and (Kronecker, 1856).<br />

17 (Gårding, 1992) and (Gårding and Skau, 1994).<br />

18 (L. Sylow, 1902, 19).


172 Chapter 8. A grand theory in spe<br />

which gradually turned into a sequence <strong>of</strong> formulae. 19 At a later time, when the ideas<br />

had matured and pro<strong>of</strong>s had been improved, the results emerged in another manu-<br />

script or in print.<br />

In the third section <strong>of</strong> the Sur la résolution algébrique des équations, which was enti-<br />

tled “On the form <strong>of</strong> algebraic expressions which can satisfy an irreducible equation<br />

<strong>of</strong> a given degree”, ABEL reverted his approach once again. In the impossibility pro<strong>of</strong>,<br />

he had fixed the equation (the general quintic) and sought to describe any algebraic<br />

expression which could satisfy it. In the opening part <strong>of</strong> the notebook manuscript, he<br />

had reversed this approach in order to describe the simplest equation which a given<br />

algebraic expression could satisfy; ABEL’S concept <strong>of</strong> simplicity was, <strong>of</strong> course, that <strong>of</strong><br />

irreducibility. But in this third section, ABEL once again fixed the equation<br />

φ (y) = 0 (8.5)<br />

<strong>of</strong> degree µ and tried to analyse the form <strong>of</strong> any algebraic expression am <strong>of</strong> order m<br />

which could satisfy it.<br />

ABEL’S attention restricted to equations <strong>of</strong> prime degree. ABEL’S ambition had<br />

been to treat — in all its generality — all degrees µ. From his correspondence there<br />

is some indication that he made some progress in solving this general problem. 20<br />

However, the notebook manuscript only contains conclusive arguments concerning<br />

the simpler case in which µ was a prime. <strong>The</strong> pivotal tools in ABEL’S investigations<br />

were the results on the constructed irreducible equation, summed up in the proposi-<br />

tion 1 above, and his penetrating knowledge <strong>of</strong> properties <strong>of</strong> <strong>Abel</strong>ian equations (see<br />

chapter 7).<br />

For ABEL, the first — and most important — consequence <strong>of</strong> assuming µ prime was<br />

to rewrite am in accordance with proposition 1:2. Writing s in place <strong>of</strong> ym above, he<br />

found<br />

am =<br />

µ−1<br />

∑ pks k=0<br />

k µ ,<br />

which follows from the fact that the exponent <strong>of</strong> the outer-most root extraction in am<br />

had to divide µ. <strong>The</strong> proposition 1:3 furthermore stated that the other roots <strong>of</strong> (8.5)<br />

could be obtained by inserting ω u s 1 µ for s 1 µ . ABEL denoted 21 these µ roots z0, . . . , zµ−1,<br />

zu =<br />

µ−1<br />

∑ pkω k=0<br />

uk s k µ for 0 ≤ u ≤ µ − 1.<br />

Since each <strong>of</strong> these was a root <strong>of</strong> the equation (8.5), they had to remain unaltered when<br />

all the root extractions in p0, . . . , pµ−1, s were given all their respective possible values,<br />

ABEL argued from proposition 1:3.<br />

19 (L. Sylow, 1902, 8).<br />

20 (<strong>Abel</strong>→Holmboe, Berlin, 1827/03/04. N. H. <strong>Abel</strong>, 1902a, 57).<br />

21 I have chosen to enumerate them starting from zero, whereas ABEL began with the number 1. <strong>The</strong><br />

benefit <strong>of</strong> my enumeration is simplicity <strong>of</strong> the subsequent formulae.


8.3. Refocusing on the equation 173<br />

<strong>The</strong> coefficients p0, . . . , pµ−1 depended rationally upon s. In the following, ABEL<br />

investigated the dependency <strong>of</strong> the coefficients p0, . . . , pµ−1 upon s. ABEL linked the<br />

choice <strong>of</strong> other root extractions 22 in the expressions for p0, . . . , pµ−1 to permutations<br />

<strong>of</strong> the roots z0, . . . , zµ−1 in a way resembling the auxiliary theorem 3 <strong>of</strong> the impos-<br />

sibility pro<strong>of</strong>. <strong>The</strong>re, ABEL had used results obtained from permuting the roots to<br />

demonstrate that any radical occurring in a supposed solution formula would have<br />

to be a rational function <strong>of</strong> the roots <strong>of</strong> the equation. A similar result was needed in<br />

this context which was more general than the quintic studied in 1826. Although his<br />

explicit calculations took another form, the underlying ideas <strong>of</strong> the reworking remain<br />

the same.<br />

ABEL based his argument on letting 23 ˆp0, . . . , ˆpµ−1, ˆs denote any set <strong>of</strong> values <strong>of</strong><br />

p0, . . . , pµ−1, s corresponding to choosing other roots <strong>of</strong> unity in the algebraic expres-<br />

sions for p0, . . . , pµ−1, s. <strong>The</strong> above argument ensuring that z0 was unaltered by other<br />

choices <strong>of</strong> root extractions, was summarized by ABEL as<br />

µ−1<br />

∑ pkω k=0<br />

uk s k µ =<br />

µ−1<br />

∑<br />

k=0<br />

ˆp k ˆω uk ˆs k µ for 0 ≤ u ≤ µ − 1.<br />

Through a simple interchange <strong>of</strong> the order <strong>of</strong> summation, ABEL found that the first<br />

coefficient p0 was unaltered if another root extraction ˆs <strong>of</strong> s was chosen. Turning his at-<br />

tention to the quantities s and ˆs, he then — by a sequence <strong>of</strong> formulae — demonstrated<br />

that there existed an integer ν such that these quantities were related by the equation<br />

ˆs = p µ ν s v . (8.6)<br />

In the course <strong>of</strong> his deductions, ABEL introduced the further simplification p1 = 1<br />

which earlier led him into the mistaken assumptions on the degrees and orders <strong>of</strong><br />

the coefficients in the impossibility pro<strong>of</strong> (see sections 6.3.2 and 6.9.1). In the present<br />

situation, it had no negative implications, though. With this simplification, the roots<br />

z0, . . . , zµ−1 could be expressed as<br />

zu = p0 + ω u s 1 µ +<br />

µ−1<br />

∑ pkω k=2<br />

uk s k µ for 0 ≤ u ≤ µ − 1.<br />

Summing over the roots and using basic properties <strong>of</strong> primitive roots <strong>of</strong> unity, ABEL<br />

obtained<br />

s 1 µ−1<br />

1<br />

µ =<br />

µ ∑<br />

k=0<br />

pus u µ−1<br />

1<br />

µ =<br />

µ<br />

∑<br />

k=0<br />

ω −k z k, and<br />

ω −ku z k for 2 ≤ u ≤ µ − 1.<br />

22 By “choosing another root extraction”, I mean (in a general setup) choosing α n√ y for n√ y where α is<br />

an n th root <strong>of</strong> unity.<br />

23 ABEL wrote s ′ , w ′ , p ′ 0 , . . . , p′ µ−1 for the quantities I have denoted ˆs, ˆω, ˆp0, . . . , ˆpµ−1. I have altered his<br />

notation to make powers such as ˆs k µ more readable.


174 Chapter 8. A grand theory in spe<br />

For any u > 1, ABEL had, therefore, explicitly demonstrated that pus was a rational<br />

function <strong>of</strong> the roots z0, . . . , zµ−1.<br />

<strong>The</strong> irreducible equation for s was <strong>Abel</strong>ian. <strong>The</strong> ultimate result <strong>of</strong> ABEL’S studies<br />

<strong>of</strong> the solubility <strong>of</strong> equations amounted to a characterization <strong>of</strong> the irreducible equation<br />

P = 0 which the quantity s satisfied. By arguments founded in C. F. GAUSS’ (1777–<br />

1855) theory <strong>of</strong> primitive roots, ABEL found that P = 0 had the property <strong>of</strong> having all<br />

its roots representable as the “orbit” <strong>of</strong> a rational function (see page 145) whereby the<br />

equation fell into the category studied in the Mémoire sur une classe particulière. 24<br />

Denoting the degree <strong>of</strong> the irreducible equation P = 0 by ν, ABEL could express its<br />

ν roots in one <strong>of</strong> the two forms<br />

s or p µ m k1 s m k 1 for 1 ≤ k1 ≤ ν − 1<br />

where m k1 ∈ {2, 3, . . . , µ − 1}. He deduced this from (8.6) described above, since<br />

choosing any other root extraction would give an ˆs <strong>of</strong> the form p µ<br />

θ sθ . Fixing some m, a<br />

sequence could be constructed, possibly renumbering the coefficients p0, . . . , p k1−1,<br />

s1 = p µ<br />

0 sm ,<br />

s2 = p µ<br />

1 sm 1 ,<br />

.<br />

s k1 = p µ<br />

k 1−1 sm k 1−1 .<br />

At some point, the sequence would stabilize because only finitely many different roots<br />

<strong>of</strong> P = 0 could be listed. Assuming this to have occurred after the k th<br />

1<br />

which point the value could be assumed to be s again, ABEL wrote<br />

s = sk1 = p µ<br />

k1−1sm k1−1 = smk k1−1 1<br />

∏ p<br />

u=0<br />

µmu<br />

k1−(u+1) .<br />

Dividing this equation by s and extracting the µ th root, he obtained the relation<br />

s mk1 −1<br />

µ<br />

k1−1 ∏ p<br />

u=0<br />

mu<br />

k = 1.<br />

1−u−1<br />

iteration, at<br />

Since the product was a rational function <strong>of</strong> s by the previous result, ABEL concluded<br />

that the exponent <strong>of</strong> s would have to be integral<br />

24 (N. H. <strong>Abel</strong>, 1829c)<br />

m k 1 − 1<br />

µ<br />

= integer,<br />

or m k 1 ≡ 1 (mod µ) .


8.3. Refocusing on the equation 175<br />

<strong>The</strong> central question <strong>of</strong> this part <strong>of</strong> the paper was whether k1 = ν, i.e. whether all<br />

the roots <strong>of</strong> P = 0 were found in the sequence above. ABEL answered this important<br />

question by a nice application <strong>of</strong> GAUSS’ primitive roots, although his presentation<br />

in the notebook becomes increasingly obscure (see figure 8.1). Eventually, nothing<br />

but a sequence <strong>of</strong> equations can be found. However, ABEL’S intended argument can<br />

be inferred and reconstructed. In the following, I add some explanation to ABEL’S<br />

equations based on arguments by HOLMBOE and SYLOW. 25<br />

In order to demonstrate that s<br />

1 1<br />

µ µ<br />

1 , . . . , sk<br />

were rational functions <strong>of</strong> s<br />

1 1 µ , ABEL let m<br />

denote a primitive root <strong>of</strong> the modulus µ and recast the procedure described above as<br />

1<br />

µ<br />

s1<br />

1<br />

µ<br />

s2<br />

= p0s mα<br />

µ ,<br />

m<br />

= p1s<br />

α<br />

µ<br />

1 ,<br />

1<br />

µ<br />

sk<br />

= pk−1s mα<br />

µ .<br />

At some point, say after the k th iteration, the procedure would stabilize and give<br />

s 1 m<br />

µ = s αk k−1<br />

µ × ∏ p<br />

u=0<br />

muα<br />

k−u−1 .<br />

By the same argument as above, ABEL could write<br />

m αk − 1<br />

µ<br />

.<br />

(8.7)<br />

= integer, (8.8)<br />

and he concluded that k divided µ − 1. This conclusion can be seen to impose a mini-<br />

mality condition upon k with respect to (8.8). However, in ABEL’S equations no men-<br />

tion <strong>of</strong> such a minimality requirement can be found. <strong>The</strong> congruence (8.8)<br />

led ABEL to introduce n such that<br />

m αk ≡ 1 (mod µ)<br />

αk = (µ − 1) n.<br />

In subsequent reasoning, ABEL repeatedly used the fact that (k, n) = 1 without going<br />

into details. However, it is a consequence <strong>of</strong> the minimality <strong>of</strong> k mentioned above.<br />

Through a sequence <strong>of</strong> deductions based on primitive roots and congruences inspired<br />

by GAUSS, ABEL could link a number β to the sequence (8.7) such that<br />

βk = µ − 1.<br />

25 (Holmboe in N. H. <strong>Abel</strong>, 1839, vol. 2, 288–293), (Sylow in N. H. <strong>Abel</strong>, 1881, vol. 2, 329–338), and (L.<br />

Sylow, 1902, 18–22).


176 Chapter 8. A grand theory in spe<br />

If any root existed outside the sequence (8.7), a sequence could be based on this<br />

root, and a similar deduction would produce another pair <strong>of</strong> integers β ′ , k ′ related by<br />

β ′ k ′ = µ − 1.<br />

However, as ABEL demonstrated, from two such sequences a third one corresponding<br />

to β ′′ = gcd (β, β ′ ) could also be constructed with the same property<br />

β ′′ k ′′ = µ − 1.<br />

ABEL knew that if β = β ′ , the two initial sequences were not distinct. If the two<br />

initial sequences were assumed to be maximal, a contraction was obtained, since the<br />

sequence corresponding to β ′′ was longer than both the initial sequences.<br />

Thus, ABEL had demonstrated that the assumption <strong>of</strong> a root existing outside the<br />

maximal sequence (8.7) led to a contradiction, and therefore all the roots were located<br />

in a single chain. Using the same notation as in the Mémoire sur une classe particulière,<br />

ABEL wrote the set <strong>of</strong> roots <strong>of</strong> P = 0 as<br />

s, θ (s) , θ 2 (s) , . . . , θ ν−1 (s) , where θ ν (s) = s,<br />

and the equation P = 0 was seen to be a specimen <strong>of</strong> the class <strong>of</strong> equations which have<br />

become known as <strong>Abel</strong>ian equations (see chapter 7).<br />

<strong>The</strong> first result <strong>of</strong> ABEL’S research had been to reduce the search for algebraic ex-<br />

pressions satisfying an arbitrary equation to the search for expressions satisfying an<br />

irreducible one. As SYLOW remarks, 26 the present investigation had led to the fur-<br />

ther restriction to studying only the possible solutions to irreducible <strong>Abel</strong>ian equations<br />

whose degree divided µ − 1. <strong>The</strong> desired complete characterization <strong>of</strong> expressions<br />

solving irreducible <strong>Abel</strong>ian equations was, however, not undertaken in the notebook<br />

study.<br />

8.4 Further ideas on the theory <strong>of</strong> equations<br />

Besides the described reduction to <strong>Abel</strong>ian equations, the notebook manuscript and<br />

ABEL’S letters contain other interesting results one <strong>of</strong> which addressed the form <strong>of</strong><br />

roots <strong>of</strong> solvable equations. This result can be seen as an elaboration and rigorization<br />

<strong>of</strong> one <strong>of</strong> L. EULER’S (1707–1783) claims.<br />

<strong>The</strong> form <strong>of</strong> roots <strong>of</strong> solvable equations: rigorizing EULER. At the end <strong>of</strong> the in-<br />

vestigation <strong>of</strong> possible solutions in the notebook, ABEL found that if an equation was<br />

solvable by radicals, its solution would be based on the relationship<br />

26 (L. Sylow, 1902, 21).<br />

1<br />

µ<br />

sk<br />

= A ν−1<br />

i ∏<br />

u=0<br />

m<br />

a<br />

kuα<br />

µ<br />

u<br />

for 0 ≤ k ≤ ν − 1,


8.4. Further ideas on the theory <strong>of</strong> equations 177<br />

Figure 8.1: One <strong>of</strong> the last pages from ABEL’S notebook manuscript on algebraic solubility<br />

(<strong>Abel</strong>, MS:696, 66). Reproduced from (N. H. <strong>Abel</strong>, 1902e, facsimile III)


178 Chapter 8. A grand theory in spe<br />

where a0, . . . , aν−1 were roots <strong>of</strong> an irreducible <strong>Abel</strong>ian equation <strong>of</strong> degree ν and the<br />

coefficients A i were rational expressions in s. <strong>The</strong> root z0 <strong>of</strong> the initial equation was in<br />

turn given from the sequence s<br />

1<br />

µ<br />

0<br />

, . . . , s<br />

z0 = p0 +<br />

1<br />

µ<br />

ν−1<br />

ν−1<br />

∑<br />

u=0 k=0<br />

by a relationship <strong>of</strong> the form<br />

ν−1<br />

∑<br />

m<br />

φu (sk) · s<br />

u<br />

µ<br />

k ,<br />

where φ0, . . . , φν−1 were rational functions. In a letter to CRELLE dated 1826, ABEL<br />

had announced a result for equations <strong>of</strong> the fifth degree which was a particular case<br />

<strong>of</strong> the above.<br />

“When an equation <strong>of</strong> the fifth degree, whose coefficients are rational numbers,<br />

is algebraically solvable, one can always give its roots the following form:<br />

x = c + A · a 1 5 · a 2 5<br />

1 · a 4 5<br />

2 · a 3 5<br />

3 + A1 · a 1 5<br />

1 · a 2 5<br />

2 · a 4 5<br />

3 · a 3 5<br />

where<br />

�<br />

a = m + n 1 + e2 +<br />

�<br />

a1 = m − n<br />

�<br />

a2 = m + n<br />

+ A2 · a 1 5<br />

2 · a 2 5<br />

3 · a 4 5 · a 3 5<br />

1 + A3 · a 1 5<br />

3 · a 2 5 · a 4 5<br />

1 · a 3 5<br />

2<br />

�<br />

h<br />

�<br />

1 + e2 + h<br />

�<br />

1 + e2 − h<br />

�<br />

�<br />

a3 = m − n 1 + e2 +<br />

A = K + K ′ a + K ′′ a2 + K ′′′ aa2,<br />

A1 = K + K ′ a1 + K ′′ a3 + K ′′′ a1a3,<br />

A2 = K + K ′ a2 + K ′′ a + K ′′′ aa2,<br />

A3 = K + K ′ a3 + K ′′ a1 + K ′′′ a1a3.<br />

h<br />

�<br />

1 + e2 + �<br />

1 + e2 �<br />

,<br />

�<br />

1 + e2 − �<br />

1 + e2 �<br />

,<br />

�<br />

1 + e2 + �<br />

1 + e2 �<br />

,<br />

�<br />

1 + e2 − �<br />

1 + e2 �<br />

,<br />

<strong>The</strong> quantities c, b [h], e, m, n, K, K ′ , K ′′ , K ′′′ are all rational numbers.<br />

In this way, however, the equation x 5 + ax + b = 0 cannot be solved as long as<br />

a and b are arbitrary quantities.” 27<br />

Probably from his realization that all quantities involved in the solution are ratio-<br />

nals, square roots <strong>of</strong> rationals, or fifth roots <strong>of</strong> rationals, ABEL concluded that there<br />

were values <strong>of</strong> a and b for which the equation x 5 + ax + b = 0 could not be solvable by<br />

27 “Wenn eine Gleichung des fünften Grades, deren Coëfficienten rationale Zahlen sind, algebraisch<br />

auflösbar ist, so kann man immer den Wurzeln folgende Gestalt geben:<br />

x = c + A · a 1 5 · a 2 5<br />

1 · a 4 5<br />

2 · a 3 5<br />

3 + A1 · a 1 5<br />

1 · a 2 5<br />

2 · a 4 5<br />

3 · a 3 5<br />

+ A2 · a 1 5<br />

2 · a 2 5<br />

3 · a 4 5 · a 3 5<br />

1 + A3 · a 1 5<br />

3 · a 2 5 · a 4 5<br />

1 · a 3 5<br />

2


8.4. Further ideas on the theory <strong>of</strong> equations 179<br />

radicals. In this way, the insolubility <strong>of</strong> fifth degree equations <strong>of</strong> the standard form 28<br />

x 5 + ax + b = 0 was demonstrated directly: If the equation had been solvable, ABEL<br />

possessed a solution formula, which he saw was not powerful enough to give the<br />

solution <strong>of</strong> arbitrary equations.<br />

In a letter to HOLMBOE from the same year, the result on the form <strong>of</strong> roots was<br />

given another twist.<br />

“Concerning equations <strong>of</strong> the 5 th degree I have found that whenever such an<br />

equation can be solved algebraically, the root must have the following form:<br />

x = A + 5√ R + 5√ R ′ + 5√ R ′′ + 5√ R ′′′<br />

where R, R ′ , R ′′ , R ′′′ are the 4 roots <strong>of</strong> an equation <strong>of</strong> the 4 th degree and have the<br />

property that they can be expressed with help <strong>of</strong> only square roots. — It has been<br />

a difficult task for me with respect to expressions and notation.” 29<br />

In this form, the statement is a refined version <strong>of</strong> EULER’S “conjecture” that the<br />

solution <strong>of</strong> the fifth degree equation should be <strong>of</strong> the form<br />

A + 5√ R + 5√ R ′ + 5√ R ′′ + 5√ R ′′′ (8.9)<br />

where R, R ′ , R ′′ , R ′′′ were solutions to an equation <strong>of</strong> the fourth degree (see section<br />

5.1).<br />

wo<br />

�<br />

a = m + n 1 + e2 +<br />

�<br />

a1 = m − n<br />

�<br />

a2 = m + n<br />

�<br />

a3 = m − n 1 + e2 +<br />

�<br />

�<br />

h 1 + e2 + �<br />

1 + e2 �<br />

,<br />

1 + e2 � �<br />

+ h 1 + e2 − �<br />

1 + e2 �<br />

,<br />

1 + e2 � �<br />

− h 1 + e2 + �<br />

1 + e2 �<br />

,<br />

� �<br />

h 1 + e2 − �<br />

1 + e2 �<br />

,<br />

A = K + K ′ a + K ′′ a2 + K ′′′ aa2,<br />

A1 = K + K ′ a1 + K ′′ a3 + K ′′′ a1a3,<br />

A2 = K + K ′ a2 + K ′′ a + K ′′′ aa2,<br />

A3 = K + K ′ a3 + K ′′ a1 + K ′′′ a1a3.<br />

Die Grössen c, b, e, m, n, K, K ′ , K ′′ , K ′′′ sind alle rationale Zahlen.<br />

Auf diese Weise lässt sich aber die Gleichung x 5 + ax + b = 0 nicht auflösen, so lange a und b<br />

beliebige Grössen sind.” (<strong>Abel</strong>→Crelle, Freyberg, 1826/03/14. N. H. <strong>Abel</strong>, 1902a, 21–22).<br />

28 If formulated in positive way, the researches <strong>of</strong> JERRARD (see section 6.9.1) demonstrated that every<br />

fifth degree equation could be transformed to this normal trinomial form. (W. R. Hamilton, 1839,<br />

251)<br />

29 “Med Hensyn til Ligninger af 5th Grad har jeg faaet at naar en saadan Ligning lader sig løse algebraisk<br />

maa Roden have følgende Form:<br />

x = A + 5√ R + 5√ R ′ + 5√ R ′′ + 5√ R ′′′<br />

hvor R, R ′ , R ′′ , R ′′′ ere de 4 Rødder af en Ligning af 4de Grad, og som lade sig udtrykke blot ved<br />

Hjelp af Qvadratrødder. — Det har været mig en vanskelig Opgave med Hensyn til Udtryk og<br />

Tegn.” (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. N. H. <strong>Abel</strong>, 1902a, 45).


180 Chapter 8. A grand theory in spe<br />

ABEL had turned the argument around and demonstrated that, although not all<br />

fifth degree equations were algebraically solvable, those which were all had solutions<br />

<strong>of</strong> the form (8.9). A particular instance <strong>of</strong> this result had already been obtained for<br />

<strong>Abel</strong>ian equations in the Mémoire sur une classe particulière as recorded in (7.10). As a<br />

result <strong>of</strong> this inversion <strong>of</strong> argument, EULER’S hypothesis can be seen as a bold conjec-<br />

ture which ABEL later turned into a pro<strong>of</strong> through a restriction on the class <strong>of</strong> objects<br />

dealt with. Where EULER had been concerned with the class <strong>of</strong> all fifth degree equa-<br />

tions, ABEL restricted (barred) his results on the form <strong>of</strong> roots to only those equations<br />

which were algebraically solvable.<br />

An extension <strong>of</strong> the class <strong>of</strong> <strong>Abel</strong>ian equations. In a later letter to CRELLE, writ-<br />

ten around the same time as the notebook entry, i.e. 1828, ABEL announced further<br />

results in the theory <strong>of</strong> equations. Generalizing the assumptions on the rational corre-<br />

spondences between roots <strong>of</strong> an irreducible equation sufficient to guarantee solubility,<br />

ABEL had found:<br />

“If three roots <strong>of</strong> an irreducible equation <strong>of</strong> a certain prime degree have such a<br />

relation between them that one can express one <strong>of</strong> the roots rationally in the two<br />

others, the equation under consideration will always be solvable by radicals.” 30<br />

As SYLOW has noticed, the assumption on the rational relationship among the<br />

three roots is not quite clear: <strong>The</strong> mathematical correct assumption is that all the roots<br />

<strong>of</strong> the equation can be expressed rationally if any two among them are considered<br />

known. 31 In the form <strong>of</strong> a corollary to his result, ABEL gave the result contained in<br />

the Mémoire sur une classe particulière that if two roots <strong>of</strong> an irreducible equation <strong>of</strong><br />

prime degree were rationally related, the equation would be algebraically solvable.<br />

Although this indicates that ABEL had, at the time <strong>of</strong> writing the Mémoire sur une<br />

classe particulière, the result on the solubility <strong>of</strong> irreducible equations <strong>of</strong> prime degree<br />

in which any root can be written as<br />

x i = θ i (x0, x1)<br />

at his disposal, he never made the more general result public in print.<br />

This class <strong>of</strong> equations, which ABEL saw contained the so-called <strong>Abel</strong>ian ones, was<br />

taken up by E. GALOIS (1811–1832) after whom they are now named. Within his the-<br />

ory (see chapter 8.5), GALOIS stated the theorem that it was a necessary and sufficient<br />

condition for algebraic solubility that “if some two <strong>of</strong> the roots <strong>of</strong> an irreducible equation<br />

<strong>of</strong> prime degree are considered known, the others can be expressed rationally.” 32<br />

30 “Si trois racines d’une équation quelconque irreductible d’un degré marqué par un nombre premier<br />

sont liées entre elles de la manière que l’on pourra exprimer l’une de ces racines rationellement en<br />

les deux autres, l’équation en question sera toujours resoluble à l’aide de radicaux.” (<strong>Abel</strong>→Crelle,<br />

Christiania, 1828/08/18. N. H. <strong>Abel</strong>, 1902a, 73).<br />

31 (L. Sylow, 1902, 17).<br />

32 “Théorème. Pour qu’une équation irréductible de degré premier soit soluble par radicaux, il faut et<br />

il suffit que deux quelconques des racines étant connues, les autres s’en déduisent rationnellement.”<br />

(Galois, 1831c, 69).


8.5. General resolution <strong>of</strong> the problem by E. GALOIS 181<br />

<strong>The</strong> major parts <strong>of</strong> ABEL’S research on equations which can be rendered intelligible<br />

have been presented above. Nevertheless, ABEL’S notebooks are filled with notes and<br />

scribbles for additional research which he never translated into a finished form suit-<br />

able for presentation. During the few remaining years <strong>of</strong> his life, ABEL became pre-<br />

occupied with other mathematical topics. Thus, we can only wonder what he might<br />

have achieved, had he returned to the theory <strong>of</strong> solubility per se.<br />

8.5 General resolution <strong>of</strong> the problem by E. GALOIS<br />

ABEL’S attempt at a general theory <strong>of</strong> the algebraic solubility <strong>of</strong> equations was not<br />

published until the first edition <strong>of</strong> the Œuvres 1839. Hence, it is most likely that GA-<br />

LOIS was unaware <strong>of</strong> ABEL’S general research when he wrote down his theory in the<br />

early 1830s. GALOIS knew the published works <strong>of</strong> LAGRANGE and A.-L. CAUCHY<br />

(1789–1857), and he had probably read ABEL’S two publications on the theory <strong>of</strong> equa-<br />

tions — the impossibility pro<strong>of</strong> <strong>of</strong> 1826 and the Mémoire sur une classe particulière pub-<br />

lished 1829 33 — as well as ABEL’S more widely known works on the theory <strong>of</strong> elliptic<br />

functions, the Recherches sur les fonctions elliptiques 34 and the Précis d’une théorie des<br />

fonctions elliptiques 35 . 36 GALOIS “vehemently denied” 37 dependence on ABEL as can<br />

be seen from the fragmentary Note sur <strong>Abel</strong>, 38 but undeniably they share many <strong>of</strong> their<br />

inspirations. In section 8.5.1, I briefly describe GALOIS’ unified theory before I com-<br />

ment upon the common inspiration and central problems shared in the works <strong>of</strong> ABEL<br />

and GALOIS (section 8.5.2).<br />

<strong>The</strong> turbulent life <strong>of</strong> EVARISTE GALOIS as well as the interplay between his life<br />

and the fate <strong>of</strong> his mathematics have been studied intensively. 39 GALOIS’ theory <strong>of</strong><br />

algebraic solubility was not made public to the mathematical community except for a<br />

small group <strong>of</strong> members <strong>of</strong> the Institut de France until J. LIOUVILLE (1809–1882) pub-<br />

lished selections from GALOIS’ mathematical manuscripts in the Journal de mathéma-<br />

tiques pures et appliquées in 1846. 40 Subsequently, many mathematicians in the sec-<br />

ond half <strong>of</strong> the nineteenth century invested great efforts in incorporating GALOIS’<br />

at times fragmentary and non-rigorous mathematics into the new standards <strong>of</strong> clar-<br />

ity and rigour. <strong>The</strong> process made mathematicians like KRONECKER return to ABEL’S<br />

works and manuscripts (see section 6.9.2), but was largely an enterprise <strong>of</strong> digesting<br />

GALOIS’ work. <strong>The</strong>refore, the reception <strong>of</strong> GALOIS’ theory is not the primary concern<br />

33 (N. H. <strong>Abel</strong>, 1826a; N. H. <strong>Abel</strong>, 1829c)<br />

34 (N. H. <strong>Abel</strong>, 1827b; N. H. <strong>Abel</strong>, 1828b)<br />

35 (N. H. <strong>Abel</strong>, 1829d)<br />

36 (Wussing, 1969, 75).<br />

37 (Kiernan, 1971, 90).<br />

38 (Galois, 1831b).<br />

39 For instance (Wussing, 1975), (Rothman, 1982), or (Toti Rigatelli, 1996).<br />

40 (Lützen, 1990, 559–580).


182 Chapter 8. A grand theory in spe<br />

Figure 8.2: EVARISTE GALOIS (1811–1832)<br />

in the present context, 41 which focuses on the differences and similarities between the<br />

almost concurrent works <strong>of</strong> ABEL and GALOIS.<br />

8.5.1 <strong>The</strong> emergence <strong>of</strong> a general theory <strong>of</strong> solubility<br />

In a sequence <strong>of</strong> manuscripts, GALOIS attacked the same two problems as ABEL had<br />

suggested in order to describe the extension <strong>of</strong> algebraic solubility (see section 8.1).<br />

ABEL had attempted to solve the first problem — that <strong>of</strong> finding all solvable equations<br />

<strong>of</strong> a given degree — in his notebook manuscript described in this chapter. ABEL’S sec-<br />

ond question concerning the determination <strong>of</strong> whether a given equation was algebra-<br />

ically solvable or not was the direct purpose <strong>of</strong> GALOIS’ theory. GALOIS intended to<br />

give characterizations <strong>of</strong> solubility which could, at least in principle, be used to decide<br />

the solubility <strong>of</strong> any given equation, but the machinery needed for actually determining<br />

the solubility <strong>of</strong> given equations was <strong>of</strong> lesser interest to him. 42<br />

<strong>The</strong> important feature <strong>of</strong> GALOIS’ theory was to associate a structure called a group<br />

to any given equation such that the question <strong>of</strong> solubility <strong>of</strong> equations could be trans-<br />

lated into questions concerning these structures. Although the concept <strong>of</strong> group only<br />

saw its first instances and was not a developed abstract concept in the works <strong>of</strong> GA-<br />

41 It has been dealt with extensively in the literature, for instance (J. Pierpont, 1898), (Kiernan, 1971),<br />

(Hirano, 1984), (Scholz, 1990), or (Martini, 1999).<br />

42 (Kiernan, 1971, 83).


8.5. General resolution <strong>of</strong> the problem by E. GALOIS 183<br />

LOIS, he was instrumental in bringing about the structural approach to mathematics,<br />

which came to dominate much <strong>of</strong> 20 th century mathematics. 43<br />

GALOIS’ work was, as he himself somewhat laconically remarked, 44 founded in<br />

the theory <strong>of</strong> permutations most <strong>of</strong> which he had taken over from CAUCHY. GALOIS<br />

considered an equation <strong>of</strong> degree m<br />

φ (x) = 0<br />

having the roots x1, . . . , xm, and claimed that the group <strong>of</strong> the equation G — later called<br />

the Galois group — could always be found, which had the following two properties:<br />

1. that every function <strong>of</strong> the roots x1, . . . , xm which was (numerically) invariant un-<br />

der the substitutions <strong>of</strong> G was rationally known, and conversely,<br />

2. that every rational function <strong>of</strong> the roots x1, . . . , xm was invariant under the sub-<br />

stitutions <strong>of</strong> G.<br />

GALOIS took over the concept <strong>of</strong> rationally known from LAGRANGE but changed<br />

the notion <strong>of</strong> invariant to stress numerical invariance instead <strong>of</strong> LAGRANGE’S formal in-<br />

variance in order to deal with special (i.e. non-general) equations. However, GALOIS’<br />

pro<strong>of</strong> <strong>of</strong> the existence <strong>of</strong> the group <strong>of</strong> the equation suffered from the unclear character<br />

<strong>of</strong> his concept <strong>of</strong> invariance. 45<br />

Although the concepts <strong>of</strong> permutation and substitution underwent some uncom-<br />

pleted changes in GALOIS’ manuscripts, he clearly perceived the multiplicative nature<br />

<strong>of</strong> substitutions — understood as transitions from one arrangement (permutation) to<br />

another — as well as the multiplicative closure <strong>of</strong> the GALOIS group.<br />

“It is clear in the group <strong>of</strong> permutations under consideration, the arrangement<br />

<strong>of</strong> letters is not important, but only the substitutions on the letters, by which we<br />

move from one permutation to another. Thus, if in similar group one has the<br />

substitutions S and T, one is also certain to have the substitution ST.” 46<br />

<strong>The</strong> second component <strong>of</strong> GALOIS’ theory addressed the reduction <strong>of</strong> the group<br />

<strong>of</strong> an equation by the adjunction <strong>of</strong> quantities to the set <strong>of</strong> rationally known quantities.<br />

By adjoining to the rationally known quantities a single root <strong>of</strong> an irreducible aux-<br />

iliary equation, GALOIS could decompose the group <strong>of</strong> the equation into a number,<br />

p, <strong>of</strong> subgroups. <strong>The</strong>se had the remarkable property that applying a substitution to<br />

43 <strong>The</strong>se aspects <strong>of</strong> GALOIS’ work have been studied by, for instance, (Wussing, 1969) and (Kiernan,<br />

1971).<br />

44 (Galois, 1830, 165).<br />

45 (Kiernan, 1971, 80–81).<br />

46 “Comme il s’agit toujours de questions où la disposition primitive des lettres n’influe en rien, dans<br />

les groupes que nous considérons, on devra avoir les mêmes substitutions quelle que soit la permutation<br />

d’où l’on sera parti. Donc si dans un pareil groupe on a les substitutions S et T, on est sûr<br />

d’avoir la substitution ST.” (Galois, 1831c, 47). I have extended the translation found in (Kiernan,<br />

1971, 80).


184 Chapter 8. A grand theory in spe<br />

permutations in one <strong>of</strong> the subgroups gave the permutations <strong>of</strong> another subgroup. 47<br />

When GALOIS adjoined the entire set <strong>of</strong> roots <strong>of</strong> the irreducible auxiliary equation, he<br />

obtained an even more remarkable result:<br />

“<strong>The</strong>orem. If one adjoins to an equation all the roots <strong>of</strong> an auxiliary equation,<br />

the groups in question in theorem II [i.e. the p subgroups mentioned above]<br />

will furthermore have the property that the substitutions are the same in each<br />

group.” 48<br />

Of this important theorem GALOIS gave no pro<strong>of</strong>, but hastily remarked “the pro<strong>of</strong><br />

will be found.” 49 <strong>The</strong> contents <strong>of</strong> the theorem is GALOIS’ characterization <strong>of</strong> the defin-<br />

ing property <strong>of</strong> what was be called normal subgroups, since GALOIS’ statement corresponds<br />

to saying that all the conjugate classes <strong>of</strong> a subgroup U are identical. 50<br />

<strong>The</strong> link between properties <strong>of</strong> the decomposition into normal subgroups <strong>of</strong> the<br />

group <strong>of</strong> the equation and the algebraic solubility <strong>of</strong> the equation was provided in the<br />

far-reaching fifth problem <strong>of</strong> the manuscript. Using modern concepts and terms, it<br />

can be summarized as follows. Assuming that the equation under consideration had<br />

the group G, and that p was the smallest prime divisor <strong>of</strong> the number <strong>of</strong> permutations<br />

in G, GALOIS argued that the equation could be reduced to another equation having<br />

a smaller group G ′ whenever a normal subgroup N existed in G with index p. Fur-<br />

thermore, the link with algebraic solubility was provided when GALOIS stated that the<br />

equation would be solvable in radicals precisely when its group could be decomposed<br />

into the trivial group by iterated applications <strong>of</strong> the preceding principle. 51<br />

GALOIS applied the general result on algebraic solubility in two ways to obtain<br />

important characterizations <strong>of</strong> solubility <strong>of</strong> equations. First, he sought criteria for sol-<br />

ubility <strong>of</strong> irreducible equations <strong>of</strong> prime degree and found the following:<br />

“Thus, for an irreducible equation <strong>of</strong> prime degree to be solvable by radicals it<br />

is necessary and sufficient that any function which is invariant under the substitutions<br />

x k x ak+b<br />

[a and b are integer constants] is rationally known.” 52<br />

47 (Galois, 1831c, 55).<br />

48 “Théorème. Si l’on adjoint à une équation toutes les racines d’une équation auxiliaire, les groupes<br />

dont il est question dans le théorème II jouiront de plus de cette propriété que les substitutions sont<br />

les mêmes dans chaque groupe.” (ibid., 57).<br />

49 “On trouvera la démonstration.” (ibid., 57).<br />

50 (Scholz, 1990, 384).<br />

51 (ibid., 384–385).<br />

52 “Ainsi, pour qu’une équation irréductible de degré premier soit soluble par radicaux, il faut et il<br />

suffit que toute fonction invariable par les substitutions<br />

x k x ak+b<br />

soit rationnellement connue.” (Galois, 1831c, 69).


8.5. General resolution <strong>of</strong> the problem by E. GALOIS 185<br />

Thus, GALOIS had characterized solvable irreducible equations <strong>of</strong> prime degree p<br />

by the necessary and sufficient requirement that their GALOIS group contained nothing<br />

but permutations corresponding to the linear congruences 53<br />

i → ai + b (mod p) where p ∤ a. (8.10)<br />

From this characterization <strong>of</strong> solubility, GALOIS deduced a second one which ABEL<br />

had also hit upon (see section 8.4), when he demonstrated his eighth proposition:<br />

“<strong>The</strong>orem. For an equation <strong>of</strong> prime degree to be solvable by radicals it is necessary<br />

and sufficient that any two <strong>of</strong> its roots being known, the others can be deduced<br />

rationally from them.” 54<br />

<strong>The</strong> character <strong>of</strong> GALOIS’ reasoning <strong>of</strong>ten left quite a lot to be desired. When LI-<br />

OUVILLE eventually published GALOIS’ manuscripts, he accompanied them with an<br />

evaluation <strong>of</strong> GALOIS’ clarity and rigour:<br />

“Clarity is indeed an absolute necessity. [. . . ] Galois too <strong>of</strong>ten neglected this<br />

precept.” 55<br />

In making GALOIS’ new ideas available to the mathematical community and in<br />

providing pro<strong>of</strong>s and elaborations <strong>of</strong> obscure points, mathematicians <strong>of</strong> the second<br />

half <strong>of</strong> the nineteenth century invested much effort in the theory <strong>of</strong> equations, per-<br />

mutations, and groups. Although GALOIS had found out how the solubility <strong>of</strong> a<br />

given equation could be determined by inspecting the decomposability <strong>of</strong> its asso-<br />

ciated group into a tower <strong>of</strong> normal subgroups, a number <strong>of</strong> points were left open<br />

for further research. To mathematicians around 1850, three problems were <strong>of</strong> primary<br />

concern: GALOIS’ construction <strong>of</strong> the group <strong>of</strong> an equation was considered to be unrig-<br />

orous, no characterization <strong>of</strong> the important solvable groups had been carried out, and a<br />

certain arbitrariness <strong>of</strong> the order <strong>of</strong> decomposition also remained. <strong>The</strong>se matters were<br />

cleared, one by one, until the theory ultimately found its mature form in the abstract<br />

field theoretic formulation <strong>of</strong> H. WEBER (1842–1913) and E. ARTIN (1898–1962). 56<br />

8.5.2 Common inspiration and common problems<br />

As mentioned earlier (p. 181), GALOIS and ABEL drew extensively on common sources.<br />

<strong>The</strong> ideas <strong>of</strong> invariance under permutations <strong>of</strong> the roots, founded in LAGRANGE’S<br />

work, 57 were important to both <strong>of</strong> them; and they both relied on the general theory<br />

53 (Scholz, 1990, 385).<br />

54 “Théorème. Pour qu’une équation de degré premier soit soluble par radicaux, il faut et il suffit que<br />

deux quelconques des racines étant connues, les autres s’en déduisent rationnellement.” (Galois,<br />

1831c, 69).<br />

55 (Liouville quoted from Kiernan, 1971, 77).<br />

56 (ibid.) and (Scholz, 1990, 392–398).<br />

57 (Lagrange, 1770–1771)


186 Chapter 8. A grand theory in spe<br />

<strong>of</strong> permutations and notations which CAUCHY had developed. 58 GALOIS’ investiga-<br />

tions, however, took a different approach from the one ABEL had employed even in<br />

his attempt at a general theory <strong>of</strong> solubility. GALOIS’ decisive step <strong>of</strong> relating a certain<br />

group to an equation and transforming the investigation <strong>of</strong> solubility <strong>of</strong> the equation<br />

into investigating properties <strong>of</strong> the group was as distant from ABEL as it was from<br />

LAGRANGE. But although ABEL’S research on algebraic solubility did not appear un-<br />

til 1839, it might have helped the mathematical community understand the purposes<br />

and intentions <strong>of</strong> GALOIS’ difficult manuscripts. As H. WUSSING has put it, ABEL<br />

would probably have been the only person with the capacity to immediately under-<br />

stand GALOIS’ works, but unfortunately ABEL had died before GALOIS ever wrote<br />

down his manuscripts. 59<br />

A common component <strong>of</strong> central importance to both ABEL and GALOIS was the<br />

concept <strong>of</strong> irreducibility (see section 7.3). In one <strong>of</strong> his manuscripts, GALOIS defined an<br />

equation to be reducible whenever it had rational divisors, and irreducible otherwise. 60<br />

This definition closely resembles the one given by ABEL, who had been more explicit<br />

about the rationality <strong>of</strong> the divisor, though. <strong>The</strong> first theorem on irreducible equations,<br />

which ABEL proved, can also be found in GALOIS’ manuscripts:<br />

“Lemma I. An irreducible equation cannot have any root in common with another<br />

rational equation without dividing it.” 61<br />

Of this lemma GALOIS gave no pro<strong>of</strong>, but he used the concept and lemma ex-<br />

tensively. Through GALOIS the concept <strong>of</strong> irreducibility in its present sense finally<br />

entered algebra as a central concept upon which deductions could be built.<br />

ABEL’S investigations had led to abstract and not easily applicable results con-<br />

cerning solubility. <strong>The</strong> same is true for GALOIS’ approach — even to a larger extent.<br />

ABEL’S positive criteria <strong>of</strong> solubility <strong>of</strong>, for instance, <strong>Abel</strong>ian equations concerned cer-<br />

tain relationships existing among the unknown roots <strong>of</strong> an equation. In case nothing<br />

but the coefficients <strong>of</strong> the equation was known, this approach had no chance <strong>of</strong> pro-<br />

ducing an answer to the question <strong>of</strong> the solubility <strong>of</strong> the equation. In GALOIS’ concept<br />

<strong>of</strong> the group <strong>of</strong> an equation, this non-constructive approach is carried to an extreme.<br />

GALOIS had tried to prove that such a group always existed, but did not address the<br />

question <strong>of</strong> how to construct it. He had presented his thoughts in a sequence <strong>of</strong> mem-<br />

oirs, one <strong>of</strong> which he had handed in to the Institut de France in January 1831. <strong>The</strong><br />

reviewers, S. F. LACROIX (1765–1843) and S.-D. POISSON (1781–1840), immediately<br />

noticed this “deficiency” and allowed it to play a role in their refusal:<br />

“[. . . ] it should be noted that [the theorem] does not contain, as the title would<br />

have the reader believe, the condition <strong>of</strong> solubility <strong>of</strong> equations by radicals. [. . . ]<br />

58 (A.-L. Cauchy, 1815a; A.-L. Cauchy, 1815b)<br />

59 (Wussing, 1975, 397).<br />

60 (Galois, 1831c, 45).<br />

61 “Lemme I. Une équation irréductible ne peut avoir aucune racine commune avec une équation<br />

rationnelle sans la diviser.” (ibid., 47).


8.5. General resolution <strong>of</strong> the problem by E. GALOIS 187<br />

This condition, if it exists, should have an external character, that can be tested<br />

by examining the coefficients <strong>of</strong> a given equation, or, at most, by solving other<br />

equations <strong>of</strong> lesser degree than that proposed. We made all possible efforts to<br />

understand M. Galois’ evidence. His thesis is neither clear enough, nor sufficiently<br />

developed to enable us to judge its rigour.” 62<br />

<strong>The</strong> interplay between the theory <strong>of</strong> equations and the flourishing theory <strong>of</strong> ellip-<br />

tic functions had been essential in ABEL’S approach (see section 7.2). <strong>The</strong> division <strong>of</strong><br />

elliptic functions had given rise to certain classes <strong>of</strong> equations described by relations<br />

among the roots, and ABEL had pursued his favorite subject, the theory <strong>of</strong> equations<br />

(see the quotation on page 160), in investigating the question <strong>of</strong> algebraic solubility <strong>of</strong><br />

these equations. Although not to the same extent engaged in research on elliptic func-<br />

tions, GALOIS also saw the modular equations <strong>of</strong> elliptic functions as an important<br />

application <strong>of</strong> and inspiration for his theory <strong>of</strong> solubility. After GALOIS was expelled<br />

from the École Normale in 1831, he <strong>of</strong>fered classes on, among other subjects <strong>of</strong> algebra,<br />

“elliptic functions treated as pure algebra”, 63 presumably dealing with the subject in<br />

a way similar to ABEL’S approach. In the 1831-manuscript, 64 GALOIS gave a general<br />

solution to the division problem concerning the division <strong>of</strong> an elliptic function <strong>of</strong> the<br />

first kind into p n equal parts, where p was a prime. <strong>The</strong> central step <strong>of</strong> the pro<strong>of</strong> was<br />

given by his result that any rational function which is unaltered by linear congruence<br />

substitutions <strong>of</strong> the form (8.10) is known. Just as ABEL had generalized his interest in<br />

elliptic functions into the integration theory <strong>of</strong> algebraic functions, GALOIS’ investi-<br />

gations took a similar turn, and a large part <strong>of</strong> his manuscripts concerned this theory.<br />

<strong>The</strong> creation <strong>of</strong> Galois <strong>The</strong>ory in many ways marked the transition into modern<br />

mathematics. <strong>The</strong> concept <strong>of</strong> group was implicitly introduced by GALOIS, and he ex-<br />

plicitly gave it its name; but more importantly, GALOIS’ revolutionary attitude toward<br />

explicit arguments in mathematics marked a transition from arguments based on ma-<br />

nipulations <strong>of</strong> formal expressions to more concept based deductions. To many nine-<br />

teenth century mathematicians this transition — together with the fragmentary and<br />

hasty character <strong>of</strong> GALOIS’ arguments — rendered the new results “vague”, faulty, or<br />

at least in need <strong>of</strong> elaboration and pro<strong>of</strong>. 65 <strong>The</strong> transition proved to be irreversible,<br />

though, and concept based mathematics was the mathematics <strong>of</strong> the future.<br />

62 (Toti Rigatelli, 1996, 90).<br />

63 (ibid., 79–80).<br />

64 (Galois, 1831a)<br />

65 (Kiernan, 1971, 59).


Part III<br />

Interlude: ABEL and the ‘new rigor’<br />

189


Chapter 9<br />

<strong>The</strong> nineteenth-century change in<br />

epistemic techniques<br />

Of the numerous transitions in epistemic techniques — changes in how mathematics<br />

was conducted — which took place in the 1820s, few were as far reaching as the ini-<br />

tiation <strong>of</strong> the movement aiming at rigorizing analysis through arithmetization. <strong>The</strong><br />

rigorization <strong>of</strong> analysis involved fundamental changes in the basic concepts <strong>of</strong> the<br />

discipline and also manifested itself on the technical level. <strong>The</strong> causal events leading<br />

to the rigorization are varied and span both external and internal factors. However,<br />

it is no coincidence that the rigorization was originally promoted in textbooks which<br />

were needed for the large-scale instruction in mathematics brought about by external<br />

events.<br />

Critical revision: A change in epistemic techniques. Central to the replacement<br />

<strong>of</strong> existing practice was the prominence given by leading research mathematicians<br />

to the rigorization program and the critical revision. J. L. LAGRANGE’S (1736–1813)<br />

textbooks marked a new awareness concerning the foundations <strong>of</strong> the calculus — later<br />

rigorization built upon the Lagrangian program.<br />

In the 1820s, A.-L. CAUCHY (1789–1857) presented his revision <strong>of</strong> the foundations<br />

<strong>of</strong> analysis which meant ingeniously revising the basic notions <strong>of</strong> the discipline. At<br />

the core <strong>of</strong> the change, CAUCHY discarded the eighteenth century conception <strong>of</strong> for-<br />

mal equality between expressions in favor <strong>of</strong> a new concept <strong>of</strong> arithmetical equality<br />

between functions. This change had implications for most <strong>of</strong> the other basic notions:<br />

limits, convergence, continuity, and differentiability to name but a few. For instance,<br />

CAUCHY was led by his new rigor to abandon attributing meaning to sums <strong>of</strong> di-<br />

vergent series and to promote tests <strong>of</strong> convergence into central positions within his<br />

theoretical framework.<br />

By the mid-1820s, N. H. ABEL (1802–1829) expressed severe concerns for the con-<br />

temporary state <strong>of</strong> the calculus: he felt that it lacked system and rigor. Simultaneously,<br />

ABEL revealed his interest in finding out how the previous generations could have ob-<br />

191


192 Chapter 9. <strong>The</strong> nineteenth-century change in epistemic techniques<br />

tained correct results from their “unrigorous” foundations. This question represents<br />

another aspect <strong>of</strong> the critical revision; an aspect which is intimately tied to the percep-<br />

tion <strong>of</strong> cumulativity in mathematics.<br />

Toward a concept based version <strong>of</strong> analysis. One <strong>of</strong> the main achievements <strong>of</strong> CAUCHY’S<br />

new rigor was the new internal relationship between definitions, theorems and pro<strong>of</strong>s.<br />

Not only did CAUCHY promote notions such as limits into the core concepts <strong>of</strong> analysis<br />

he also put his definitions into direct use in forming new concepts, e.g. convergence<br />

and continuity, and in proving theorems about these concepts. Thus, CAUCHY’S defi-<br />

nitions in certain senses continued existing trends but were more concrete and appli-<br />

cable in stating and demonstrating theorems. Furthermore, CAUCHY’S critical revi-<br />

sion forced him to restructure the network <strong>of</strong> definitions and theorems changing the<br />

internal fabric <strong>of</strong> the theory.<br />

Interpreted in the framework <strong>of</strong> the nineteenth century transition toward concept<br />

based mathematics, the rigorization <strong>of</strong> analysis thus provides an important example in<br />

which both the structure and the techniques <strong>of</strong> a discipline underwent deep changes.<br />

In the following chapters, the general transition is described and analyzed from<br />

the perspective <strong>of</strong> ABEL. ABEL’S impact on the rigorization program mainly consists<br />

<strong>of</strong> three themes:<br />

1. A very critical attitude which was mainly expressed in letters.<br />

2. A pro<strong>of</strong> <strong>of</strong> the binomial theorem which surpassed its predecessors in generality<br />

and rigor.<br />

3. A discussion on the existence <strong>of</strong> general criteria <strong>of</strong> convergence.<br />

All three themes fall in the changing standard <strong>of</strong> analysis which was brought about<br />

by CAUCHY’S revision <strong>of</strong> the discipline. <strong>The</strong>refore, important aspects <strong>of</strong> this context<br />

must first be described.


Chapter 10<br />

Toward rigorization <strong>of</strong> analysis<br />

From the time the calculus emerged in the 17 th century until the end <strong>of</strong> the 18 th cen-<br />

tury, mathematicians and philosophers were wary when confronted with questions<br />

concerning its foundations. To some extent ignoring foundational questions, mathe-<br />

maticians focused on creating new results which could be useful in answering inter-<br />

esting questions, for instance in the field <strong>of</strong> mathematical physics. To some mathe-<br />

maticians working toward the end <strong>of</strong> the 18 th century, rigorously founding the calcu-<br />

lus remained one <strong>of</strong> the few open problems; but one <strong>of</strong> relatively lesser importance<br />

than the development <strong>of</strong> new analytical results. 1 To others, primarily J. L. LAGRANGE<br />

(1736–1813), the foundations <strong>of</strong> the calculus became a prestigious mathematical research<br />

problem. 2<br />

<strong>The</strong> transformation <strong>of</strong> concepts, theorems, and pro<strong>of</strong>s in the process <strong>of</strong> rigorization<br />

in analysis have been subject to a variety <strong>of</strong> historical enquiries; in the following, em-<br />

phasis is given to establishing and illustrating certain ideas and developments which<br />

are <strong>of</strong> importance in subsequent chapters. 3<br />

10.1 EULER’s vision <strong>of</strong> analysis<br />

To understand the revision and the contents <strong>of</strong> the refocus on rigor, some aspects <strong>of</strong><br />

eighteenth century analysis are <strong>of</strong> key importance. In particular, the results and tech-<br />

niques <strong>of</strong> L. EULER (1707–1783) dominated the way mathematicians worked in the<br />

field for half a century.<br />

Focus on functions and formal equality. Beginning with his influential monograph<br />

Introductio in analysin infinitorum, 4 EULER promoted functions to become the basic ob-<br />

1 See for instance the quotations in section 3.3 frequently invoked to document a belief in the stagnation<br />

<strong>of</strong> the mathematical sciences.<br />

2 For LAGRANGE’S algebraic approach to the calculus, see e.g. (Grabiner, 1990); for its influence on<br />

CAUCHY, see (Grabiner, 1981b). <strong>The</strong> best general presentation <strong>of</strong> the development <strong>of</strong> analysis in the<br />

nineteenth century is, I think, (Bottazzini, 1986).<br />

3 For the evolution <strong>of</strong> rigorization in analysis, see e.g. (Bottazzini, 1986; Jahnke, 1999; Lützen, 1999).<br />

4 (L. Euler, 1748).<br />

193


194 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

jects <strong>of</strong> analysis. EULER’S definitions and use <strong>of</strong> functions have attracted the interest<br />

<strong>of</strong> historians <strong>of</strong> mathematics. 5 In the present context, the two most important aspects<br />

<strong>of</strong> EULER’S approach are:<br />

1. EULER’S variable quantities were universal in the sense that they would “com-<br />

prise all determinate values” including positive and negative, rational and irra-<br />

tional, and real and imaginary values.<br />

2. EULER defined a function <strong>of</strong> a variable quantity to be an “analytic expression<br />

composed in any way from the variable quantity and numbers or constant quan-<br />

tities”. <strong>The</strong> operations allowed to form analytic expressions were algebraic opera-<br />

tions, both finite and infinite.<br />

Together, these two aspects entail an important interpretation <strong>of</strong> the concept <strong>of</strong><br />

equality between functions. To EULER, two analytic expressions were considered<br />

equal if one could be transformed into the other by a sequence <strong>of</strong> (formal) manip-<br />

ulations. For instance, in developing methods for expanding rational functions into<br />

power series, EULER described — in the Introductio — a method by which the two ex-<br />

pressions<br />

1<br />

1 − x and<br />

∞<br />

∑ x<br />

n=0<br />

n<br />

should be considered equal because the latter could be obtained by (formally) carrying<br />

out the division. 6 Of course, EULER was aware that peculiar results would emerge if<br />

certain numerical values were inserted for x and the equality was believed to apply to<br />

this numerical case as well. <strong>The</strong> proper interpretation <strong>of</strong> the sum<br />

1 − 1 + 1 − 1 + . . .<br />

had been a controversial subject throughout the first half <strong>of</strong> the eighteenth century.<br />

To EULER, its sum would be 1 2 by the formal equality above. Generally, EULER chose<br />

to focus on the formal aspect <strong>of</strong> functional equalities ignoring the “paradoxes” which<br />

might occur if numerical values were inserted.<br />

To a modern reader, EULER’S disregard for numerical convergence may seem odd.<br />

However, it corresponds to a paradigm in analysis — the Euclidean paradigm — which<br />

focused on the fruitful manipulations <strong>of</strong> finite or infinite expressions; the un-problematic<br />

transition from one such representation to another constituted a cornerstone <strong>of</strong> EU-<br />

LER’S skillful investigations in analysis.<br />

5 See e.g. (Jahnke, 1999; Lützen, 1978; Youschkevitch, 1976).<br />

6 (L. Euler, 1748, §60–61).


10.1. EULER’s vision <strong>of</strong> analysis 195<br />

10.1.1 <strong>The</strong> binomial theorem<br />

In its various forms and various degrees <strong>of</strong> specialization, i.e. various restrictions on<br />

m and x, the binomial theorem asserts the equality<br />

(1 + x) m = 1 + m<br />

1<br />

m (m − 1)<br />

x + x<br />

1 · 2<br />

2 +<br />

m (m − 1) (m − 2)<br />

x<br />

1 · 2 · 3<br />

3 + . . . .<br />

<strong>The</strong> theorem became one <strong>of</strong> the pivotal points <strong>of</strong> analysis since it was first employed as<br />

a heuristic tool by I. NEWTON (1642–1727) to obtain series expansions for expressions<br />

such as (1 + x) 1 2 . For integral exponents (m ∈ N), the binomial theorem reduced to<br />

the well known — and firmly established — binomial formula<br />

(1 + x) m =<br />

m<br />

∑<br />

n=0<br />

� m<br />

n<br />

�<br />

x n for m ∈ N.<br />

NEWTON used extrapolation from the cases <strong>of</strong> integral exponents to obtain the equal-<br />

ity <strong>of</strong> the finite and infinite expressions in situations corresponding to fractional ex-<br />

ponents (e.g. m = 1 2 , above). In the eighteenth century, the binomial theorem was<br />

provided with various pro<strong>of</strong>s. 7<br />

To further illustrate the Eulerian paradigm in analysis, the role played by the bino-<br />

mial theorem within EULER’S structuring <strong>of</strong> analysis provides many interesting hints;<br />

furthermore, that theorem is <strong>of</strong> direct importance in understanding the way analysis<br />

was reorganized in the early nineteenth century.<br />

EULER’S first pro<strong>of</strong> <strong>of</strong> the binomial theorem: the link with Taylor series. In the In-<br />

troductio, EULER gave no general pro<strong>of</strong> <strong>of</strong> the binomial theorem but repeatedly used<br />

a particular version in which he let n → ∞ in the binomial formula. Later, he pre-<br />

sented two different pro<strong>of</strong>s <strong>of</strong> this highly important tool. <strong>The</strong> first pro<strong>of</strong>, published in<br />

his sequel textbook Institutiones calculi integralis, 8 highlighted the intimate connection<br />

between the binomial theorem and the Taylor expansion theorem which in modern<br />

notation stated that any function f (later with certain restrictions) could be expanded<br />

as<br />

f (x + a) = f (x) + f ′ (x)<br />

a +<br />

1<br />

f ′′ (x)<br />

1 · 2 a2 + . . . .<br />

For EULER, the binomial theorem was a rather easy consequence <strong>of</strong> the Taylor expan-<br />

sion provided the relation<br />

d<br />

dx xµ = µx µ−1<br />

had previously been established for exponents µ. However, as EULER later realized,<br />

the binomial theorem was central to the differentiation <strong>of</strong> such monomials if µ was<br />

not an integer. <strong>The</strong>refore, proving the binomial theorem from the Taylor series expan-<br />

sion had created a vicious circle in the argument. Nevertheless, pro<strong>of</strong>s <strong>of</strong> the binomial<br />

7 <strong>The</strong> history <strong>of</strong> the binomial theorem has attracted the interest <strong>of</strong> many scholars, see e.g. (Dhombres<br />

and Pensivy, 1988; Pensivy, 1994).<br />

8 (L. Euler, 1755, 276–279).


196 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

theorem from Taylor theorem recurred throughout the century and even into the nineteenth<br />

century. 9<br />

EULER’S second pro<strong>of</strong> <strong>of</strong> the binomial theorem based on functional equations. In<br />

his second pro<strong>of</strong> <strong>of</strong> the binomial theorem, published in 1775, 10 EULER devised his<br />

pro<strong>of</strong> following an outline which would recur in most subsequent “rigorous” pro<strong>of</strong>s.<br />

EULER introduced the notation<br />

[m] = 1 + m<br />

1<br />

m (m − 1)<br />

x + x<br />

1 · 2<br />

2 + . . .<br />

to denote the binomial series associated with the exponent m. Thus, proving the bino-<br />

mial theorem thus amounted to proving the equality [m] = (1 + x) m . <strong>The</strong> central step<br />

in the pro<strong>of</strong> was the realization that the brackets satisfied a functional equation 11<br />

[m + n] = [m] · [n] .<br />

EULER’S pro<strong>of</strong> <strong>of</strong> the functional equation was based on formally multiplying the corre-<br />

sponding infinite series. Once EULER had obtained the above functional equation and<br />

the binomial formula secured the equality [m] = (1 + x) m for integral m, he extended<br />

the domain for m by the computation<br />

[1] =<br />

�<br />

m · 1<br />

� � �m 1<br />

=<br />

m m<br />

⇒<br />

m,n∈N<br />

�<br />

n<br />

�<br />

= [n]<br />

m<br />

1 m = (1 + x) n m .<br />

Thus, EULER proved the binomial theorem for all fractional exponents and claimed —<br />

without giving any pro<strong>of</strong> — that it extended to all real exponents by way <strong>of</strong> continuity<br />

(see below). In summary, the central steps <strong>of</strong> EULER’S second pro<strong>of</strong> <strong>of</strong> the binomial<br />

theorem are:<br />

1. <strong>The</strong> binomial formula, [m] = (1 + x) m for m ∈ N.<br />

2. <strong>The</strong> functional equation [m + n] = [m] · [n] proved by manipulating the associ-<br />

ated power series.<br />

3. An extension to rational exponents.<br />

4. A further extension to real exponents by continuity arguments.<br />

In complete correspondence with his views on formal equality, EULER did not ven-<br />

ture into considerations <strong>of</strong> the convergence <strong>of</strong> the infinite expression contained in the<br />

binomial theorem. To him, the theorem simply stated a formal equivalence <strong>of</strong> two<br />

different representations <strong>of</strong> the same function (expression).<br />

9 On the pro<strong>of</strong> by WALLACE, see (Craik, 1999, 252–253).<br />

10 (L. Euler, 1775).<br />

11 For the history <strong>of</strong> functional equations, mainly with CAUCHY, see (J. Dhombres, 1992).


10.2. LAGRANGE’s new focus on rigor 197<br />

1797 Théorie des fonctions analytiques<br />

1806 Leçons sur le calcul des fonctions<br />

1813 Théorie des fonctions analytiques, nouvelle<br />

édition<br />

Table 10.1: LAGRANGE’s monographs on his algebraic analysis<br />

10.2 LAGRANGE’s new focus on rigor<br />

<strong>The</strong> Eulerian approach to analysis based on functions and series representations proved<br />

highly productive for mathematicians with right kinds <strong>of</strong> intuitions and understand-<br />

ing. Toward the end <strong>of</strong> the century, a number <strong>of</strong> events — in particular the external<br />

influence <strong>of</strong> mass instruction in mathematics and the change <strong>of</strong> generations — intro-<br />

duced a different view on the status <strong>of</strong> analysis. Its fruitfulness was admired but its<br />

lack <strong>of</strong> strict logical order was realized by some <strong>of</strong> its most distinguished practitioners.<br />

More so than anybody else, JOSEPH LOUIS LAGRANGE was instrumental in fertilizing<br />

the ground for a fundamental revision <strong>of</strong> the Eulerian paradigm.<br />

LAGRANGE presented his new algebraic theory <strong>of</strong> functions in three important<br />

monographs (see table 10.1). In what follows, references are made to the second edi-<br />

tion <strong>of</strong> the Théorie des fonctions analytiques which was the latest <strong>of</strong> the three and was<br />

included in LAGRANGE’S collected works. 12<br />

Importantly, LAGRANGE believed he could prove that any function could be ex-<br />

panded “by the theory <strong>of</strong> series” 13 into a series <strong>of</strong> the form<br />

f (x + i) = f (x) + ip (x) + i 2 q (x) + i 3 r (x) + . . . .<br />

<strong>The</strong> functions p, q, r, . . . were called the ‘derived’ functions <strong>of</strong> f , and it was the crux<br />

<strong>of</strong> the theory to show that they corresponded to the ordinary differentials obtained in<br />

the usual — less rigorous — way.<br />

Thus, at the very center <strong>of</strong> the Lagrangian system laid the expansion <strong>of</strong> a function<br />

into a power series. As J. V. GRABINER has convincingly described in her thesis, the<br />

expansion into power series was not an assumption in the Lagrangian system but was<br />

provided with an algebraic pro<strong>of</strong> using one <strong>of</strong> EULER’S ideas. 14 As a consequence, the<br />

general expansion <strong>of</strong> any function into power series was made into a general principle<br />

replacing the important tool for obtaining such expansions which EULER had used to<br />

such a high effect, the binomial theorem.<br />

LAGRANGE’S contribution to the rigorization <strong>of</strong> the calculus was at least tw<strong>of</strong>old:<br />

1. <strong>The</strong> mere fact that LAGRANGE — “a most illustrious mathematician” — devoted<br />

12 (Lagrange, 1813).<br />

13 (ibid., 7–8).<br />

14 (Grabiner, 1990, 93ff).


198 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

so much attention to the foundational questions raised the prestige <strong>of</strong> such ques-<br />

tions; rigorization became a legitimate mathematical research topic.<br />

2. Just as importantly, LAGRANGE’S work on rigorization provided a revolutionary<br />

new synthesis <strong>of</strong> the formal interpretation <strong>of</strong> series at the heart <strong>of</strong> the calculus.<br />

Furthermore, the binomial theorem and the expansion into power series (Taylor<br />

series) changed their internal relationship and dependency (see below).<br />

10.3 Early rigorization <strong>of</strong> theory <strong>of</strong> series<br />

In the first decades <strong>of</strong> the 19 th century, a number <strong>of</strong> mathematicians responded to<br />

the call for rigorization in the theory <strong>of</strong> infinite series. Two <strong>of</strong> the most interesting<br />

reactions to the state <strong>of</strong> rigor in the theory <strong>of</strong> infinite series were made by C. F. GAUSS<br />

(1777–1855) and B. BOLZANO (1781–1848).<br />

To GAUSS, BOLZANO, and their contemporaries, analysis was a conglomerate <strong>of</strong><br />

various methods, key results, and foundations. An interesting illustration <strong>of</strong> the con-<br />

currently existing approaches to the discipline can be found in the textbooks written<br />

by S. F. LACROIX (1765–1843) just before the turn <strong>of</strong> the century. 15 In three volumes,<br />

LACROIX presented much <strong>of</strong> the key material <strong>of</strong> analysis adapting various approaches<br />

and foundations to suit his needs.<br />

Both GAUSS and BOLZANO reacted inspired partly by philosophical arguments; in<br />

the following, some <strong>of</strong> the relevant aspects <strong>of</strong> their contributions are outlined.<br />

10.3.1 GAUSS’ hypergeometric series<br />

GAUSS’ main contribution to the rigorization <strong>of</strong> the theory <strong>of</strong> series consisted <strong>of</strong> a<br />

paper concerning the so-called hypergeometric series . 16 <strong>The</strong> paper was presented in<br />

1812 and published the following year. 17 To GAUSS, the hypergeometric series<br />

F (α, β, γ, x) = 1 +<br />

∞<br />

x<br />

∑<br />

n=1<br />

n n<br />

(α + m) (β + m)<br />

n! ∏<br />

m=0 (γ + m)<br />

constituted a preferred representation <strong>of</strong> a vast range <strong>of</strong> functions including logarith-<br />

mic, elliptic, and other transcendental functions. By studying this series in its gen-<br />

erality, GAUSS obtained knowledge <strong>of</strong> the functions which it could represent. GAUSS<br />

never conducted a full investigation <strong>of</strong> which functions it could represent but the study<br />

<strong>of</strong> the series remained an interesting topic, in particular for Göttingen mathematicians.<br />

GAUSS’ research represents an intermediate between the old direct and more special-<br />

ized representations and the modern concept based approach to analysis. 18 More im-<br />

15 (Lacroix, 1797; Lacroix, 1798; Lacroix, 1800).<br />

16 <strong>The</strong> name hypergeometric series is a later invention, see (Wussing, 1982, 299).<br />

17 (C. F. Gauss, 1813).<br />

18 See also chapter 21.


10.3. Early rigorization <strong>of</strong> theory <strong>of</strong> series 199<br />

portantly, GAUSS’ investigations <strong>of</strong> the hypergeometric series also contained a number<br />

<strong>of</strong> very interesting results. In particular, GAUSS’ attitude toward convergence <strong>of</strong> series<br />

and criteria for deciding the convergence is relevant to the present analyses.<br />

Concepts <strong>of</strong> convergence and a criterion <strong>of</strong> convergence. Before he could advance<br />

to deeper questions, GAUSS emphasized that the convergence or divergence <strong>of</strong> the<br />

hypergeometric series had to be investigated. In his paper on the hypergeometric<br />

series, GAUSS gave no explicit definition <strong>of</strong> convergence. However, by the following<br />

argument, GAUSS claimed that the series converged for |x| < 1 and diverged for |x| ><br />

1. As was customary, these requirements were stated verbally without the notation <strong>of</strong><br />

numerical values (see, e.g., quotation below).<br />

GAUSS compared the coefficients <strong>of</strong> two sequential powers <strong>of</strong> x, say the coefficients<br />

<strong>of</strong> x m and x m+1 , and found that their ratio<br />

1 + γ+1<br />

m<br />

1 + α+β<br />

m<br />

+ γ<br />

m 2<br />

+ αβ<br />

m 2<br />

approached the value 1 when m was taken to be increasingly large. GAUSS then con-<br />

cluded that for any complex value <strong>of</strong> x with |x| < 1, the series would be convergent “at<br />

least from some point onward” and lead to a determinate finite sum. In case |x| > 1,<br />

the series would necessarily diverge and it could not have a sum. 19 GAUSS summa-<br />

rized his position:<br />

“Since our function is defined as the sum <strong>of</strong> a series, it is obvious, that our<br />

investigations are naturally confined to the cases in which the series actually converges<br />

and that it is absurd to ask for the value <strong>of</strong> the series whenever x has a<br />

value greater than unity.” 20<br />

Of the cases with |x| = 1, GAUSS only investigated x = 1 and found that under the<br />

condition α + β − γ < 0, the series would have a finite sum. 21<br />

In the above context, GAUSS appears to have employed a concept <strong>of</strong> series conver-<br />

gence which corresponded to the partial sums approaching a finite limit. We are easily<br />

led to believe that GAUSS’ familiar looking notions such as convergent and sum meant<br />

the same to him as they do to us. However, another concept <strong>of</strong> convergence was also<br />

in use at GAUSS’ time and even appeared later in his manuscripts (see below). <strong>The</strong>re-<br />

fore, it is worth re-examining the evidence to see if it appears different with this added<br />

information.<br />

Originating with J. LE R. D’ALEMBERT (1717–1783) in the mid-eighteenth century,<br />

the term convergent was used by mathematicians within the formal paradigm to denote<br />

19 (ibid., 126).<br />

20 “Patet itaque, quatenus functio nostra tamquam summa seriei definita sit, disquisitionem natura<br />

sua restrictam esse ad casus eos, ubi series revera convergat, adeoque quaestionem ineptam esse,<br />

quinam sit valor seriei pro valore ipsius x unitate maiori.” (ibid., 126). For a German translation, see<br />

(C. F. Gauss, 1888, 10).<br />

21 (C. F. Gauss, 1813, 139, 142–143).


200 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

series in which the numerical value <strong>of</strong> the general term vanished monotonically, i.e.<br />

series ∑ an for which the sequence |an| was monotonically decreasing and approached<br />

zero. 22 <strong>The</strong> vanishing <strong>of</strong> terms, clearly contrasted to the convergence <strong>of</strong> the partial<br />

sums, can be found in an unpublished manuscript written by GAUSS probably after<br />

1831. 23<br />

“By convergence <strong>of</strong> an infinite series, I will simply understand nothing but the<br />

infinite approaching <strong>of</strong> its terms toward 0 when the series is infinitely continued.<br />

<strong>The</strong> convergence <strong>of</strong> a series in itself is thus to be distinguished from the convergence<br />

<strong>of</strong> its summation toward a finite limit; however, the latter implies the former<br />

but not the other way around.” 24<br />

Exactly which concept <strong>of</strong> convergence, GAUSS had in mind in his research on the<br />

hypergeometric series can seem unclear. From a modern perspective, we are tempted<br />

to assume that GAUSS interpreted convergence as convergence <strong>of</strong> the partial sums and<br />

interpret GAUSS’ comparison <strong>of</strong> subsequent terms as an implicit quotient criterion.<br />

However, GAUSS’ reasoning can equally well be interpreted within the older concept<br />

<strong>of</strong> D’ALEMBERT-convergence. 25<br />

In terms <strong>of</strong> the development described in the next chapter, GAUSS’ investigation<br />

on the hypergeometric series is important in three respects:<br />

1. GAUSS’ investigation was confined to a particular series, albeit one with three<br />

parameters which enabled GAUSS to model a number <strong>of</strong> transcendental func-<br />

tions using it.<br />

2. GAUSS insisted on establishing the convergence <strong>of</strong> the series before speaking <strong>of</strong><br />

its sum. He used an implicit theorem — apparently equivalent to the ratio test 26<br />

(see subsequent chapters) — to determine restrictions on the variable x.<br />

3. Despite aiming at “the rigorous methods <strong>of</strong> the ancient geometers” 27 , GAUSS’<br />

theory <strong>of</strong> infinite series as expressed in the paper on the hypergeometric series<br />

was rudimentary and not spelled out in much detail. For instance, it is not com-<br />

pletely clear precisely what his basic notions meant.<br />

22 (Grabiner, 1981b, 60).<br />

23 (Schneider, 1981, 55–56).<br />

24 “Ich werde unter Convergenz, einer unendlichen Reihe schlechthin beigelegt, nichts anders verstehen<br />

als die beim unendlichen Fortschreiten der Reihe eintretende unendliche Annäherung ihrer<br />

Glieder an die 0. Die Convergenz einer Reihe an sich ist also wohl zu unterscheiden von der Convergenz<br />

ihrer Summirung zu einem endlichen Grenzwerthe; letztere schliesst zwar die erstere ein,<br />

aber nicht umgekehrt.” (C. F. Gauss, Fa, Kapsel 46a, A1–A13, 400).<br />

25 In most <strong>of</strong> the (earlier) secondary literature, e.g. (Pringsheim, 1898–1904, 79), GAUSS’ emphasis on<br />

establishing the convergence <strong>of</strong> the hypergeometric series and his use <strong>of</strong> the quotient comparison<br />

have been taken as precursors <strong>of</strong> the rigorization program (see next chapter). SCHNEIDER has aptly<br />

interpreted GAUSS’ concept <strong>of</strong> convergence in terms <strong>of</strong> the sequence <strong>of</strong> terms (Schneider, 1981, 56).<br />

26 <strong>The</strong> ratio test is also sometimes called the quotient test but I will use the term ratio test, throughout.<br />

27 “Ostendemus autem, et quidem, in gratiam eorum, qui methodis rigorosis antiquorum geometrarum<br />

favent, omni rigore.” (C. F. Gauss, 1813, 139).


10.3. Early rigorization <strong>of</strong> theory <strong>of</strong> series 201<br />

To summarize the debate, we have to emphasize three dates. In 1812, GAUSS’<br />

presented his research on hypergeometric series in which his concept <strong>of</strong> convergence<br />

remains undefined; in 1821, A.-L. CAUCHY (1789–1857) promoted the convergence<br />

<strong>of</strong> the partial sums into the only acceptable definition <strong>of</strong> convergence; but as late as<br />

1831, GAUSS employed a D’ALEMBERT-like concept <strong>of</strong> convergence which entailed the<br />

vanishing <strong>of</strong> the terms and did not provide convergence <strong>of</strong> the series. To believe that<br />

GAUSS had anticipated CAUCHY’S notion <strong>of</strong> convergence and the ratio test in 1812<br />

thus seems to be the least efficient interpretation. GAUSS may very well have held the<br />

same conceptions about convergence in 1812 as he evidently did in 1831. Instead, it<br />

seems that until CAUCHY’S work, different notions <strong>of</strong> convergence were co-existing<br />

and the position <strong>of</strong> definitions and tests <strong>of</strong> convergence within the structure <strong>of</strong> the<br />

theory <strong>of</strong> series floated.<br />

10.3.2 BOLZANO’s rigorization <strong>of</strong> the binomial theorem<br />

Contrary to GAUSS, the Czech priest and mathematician BOLZANO did not have the<br />

ear <strong>of</strong> the international mathematical community although his ideas and visions for<br />

the foundation <strong>of</strong> the calculus reached even further than GAUSS’. To promote inter-<br />

est in his work, BOLZANO published critical investigations and new pro<strong>of</strong>s <strong>of</strong> key<br />

theorems <strong>of</strong> analysis. He hoped that mathematicians would pay more attention to a<br />

broader philosophical program which he was developing.<br />

In 1816 and in Prague, BOLZANO published a book entitled Der binomische Lehrsatz<br />

which is <strong>of</strong> particular relevance to the current purpose. 28 In that book, BOLZANO<br />

scrutinized existing derivations <strong>of</strong> the binomial theorem before going on to present his<br />

own pro<strong>of</strong>. As noted, N. H. ABEL (1802–1829) once praised BOLZANO’S cleverness<br />

(see p. 42); important aspects <strong>of</strong> ABEL’S criticism may well have their origins with<br />

BOLZANO.<br />

BOLZANO’S critical attitude. In the introduction <strong>of</strong> his book, BOLZANO reviewed<br />

the structures <strong>of</strong> previous pro<strong>of</strong>s <strong>of</strong> the binomial theorem. In the process, BOLZANO<br />

developed a penetrating criticism <strong>of</strong> the accepted methods <strong>of</strong> reasoning with infinite<br />

series. Soon, others would repeat BOLZANO’S criticism — at least, ABEL’S judgement<br />

<strong>of</strong> eighteenth century epistemic techniques in analysis resembled some <strong>of</strong> BOLZANO’S<br />

points.<br />

A number <strong>of</strong> interesting themes were raised in BOLZANO’S introduction. BOLZANO<br />

observed that the foundation <strong>of</strong> the entire “higher analysis” (calculus) rested on Tay-<br />

lor’s <strong>The</strong>orem and that this theorem in turn relied on the binomial theorem. Conse-<br />

quently, the obscure status <strong>of</strong> the pro<strong>of</strong> <strong>of</strong> the latter theorem had severe implications<br />

for the entire discipline.<br />

28 (Bolzano, 1816). BOLZANO’S titles are <strong>of</strong>ten very precise and very long; here the abridged version is<br />

used throughout.


202 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

Figure 10.1: BERNARD BOLZANO (1781–1848)<br />

Indeed, from B. TAYLOR’S (1685–1731) days, pro<strong>of</strong>s <strong>of</strong> Taylor’s <strong>The</strong>orem had relied<br />

on an analogy between repeated differences and the binomial formula. 29 However, the<br />

relevant step in the pro<strong>of</strong> <strong>of</strong> Taylor’s <strong>The</strong>orem seems to have been a limit process based<br />

on the binomial formula in which the exponent n increased to infinity and thus did not<br />

rely on the full binomial theorem. This distinction between the binomial theorem and<br />

the indicated limit process does not seem to have been undertaken by eighteenth and<br />

nineteenth century mathematicians, though.<br />

Next, BOLZANO criticized previous pro<strong>of</strong>s for operating with (completed) infinite<br />

series, i.e. working with series as if they were polynomials. Instead, he proposed a<br />

concept <strong>of</strong> numerical limit processes based on (variable) quantities (Größen) ω which<br />

could be assumed positive but less than any given value. He also described these<br />

quantities as “quantities which can be made as small as one desires.” 30 Importantly,<br />

BOLZANO’S ω was not a completed infinitesimal but a variable quantity which de-<br />

pended on a limit process.<br />

In continuation <strong>of</strong> the previous point, BOLZANO insisted that restrictions be im-<br />

posed on the binomial such that the series was (arithmetically) convergent. He claimed<br />

that previous pro<strong>of</strong>s had “proved to much” by not taking such restrictions on x into<br />

account and forbade application <strong>of</strong> the theorem outside the domains <strong>of</strong> convergence<br />

<strong>of</strong> the series. In the argument, BOLZANO employed a counter example which based<br />

29 See e.g. (Jahnke, 1999, 139–142).<br />

30 “Größen, welche so klein werden können, als man nur immer will.” (Bolzano, 1816, v).


10.3. Early rigorization <strong>of</strong> theory <strong>of</strong> series 203<br />

on uncritical use <strong>of</strong> the binomial theorem<br />

√ 1<br />

−1 = (1 − 2) 2 = 1 − 1 1 1<br />

· 2 − · 4 − · 8 − . . . (10.1)<br />

2 8 16<br />

exhibited the imaginary unit as an infinite sum <strong>of</strong> real numbers. 31<br />

Following his program, BOLZANO rejected all previous pro<strong>of</strong>s: NEWTON’S pro<strong>of</strong><br />

because it had been based on extrapolation and not “everything which corresponds<br />

to known truths is necessarily true”; 32 the pro<strong>of</strong>s based on the expansion into Taylor<br />

series because they introduced a vicious circle and the binomial theorem should be the<br />

more basic <strong>of</strong> the two theorems; and even EULER’S second pro<strong>of</strong> because it operated<br />

with completed infinite series and did not consider the convergence <strong>of</strong> the series.<br />

Revision <strong>of</strong> EULER’S pro<strong>of</strong>. Subsequent to his critical remarks, BOLZANO presented<br />

his own new rigorized pro<strong>of</strong> <strong>of</strong> the binomial theorem. He based it on the outline<br />

<strong>of</strong> EULER’S second pro<strong>of</strong> but replaced the way in which EULER handed infinite series<br />

with his new concept <strong>of</strong> numerical equality and limits. Overturning EULER’S manipu-<br />

lations <strong>of</strong> completed infinite series, BOLZANO worked with the partial sums and limit<br />

arguments. Expressed in EULER’S notation, BOLZANO proved by multiplying the first<br />

s terms <strong>of</strong> [m] with the first t terms <strong>of</strong> [n], that the first min (s, t) terms <strong>of</strong> the product<br />

corresponded to the first min (s, t) terms <strong>of</strong> [m + n]. 33 If we introduce the notation [m] t<br />

s<br />

to denote the sum <strong>of</strong> the terms ranging from s to t in the series [m], 34 the result can be<br />

expressed as<br />

�<br />

[m] s<br />

1 [n]t<br />

�min(s,t) 1 = [m + n]<br />

1<br />

min(s,t)<br />

1 .<br />

However, the terms after min (s, t) would not always be equal in the two expressions<br />

but BOLZANO found that the difference<br />

�<br />

[m] s<br />

1 [n]t<br />

�r+s 1<br />

min(s,t)<br />

− [m + n]s+t<br />

min(s,t)<br />

could be made smaller than any given positive value by taking min (s, t) sufficiently<br />

large provided that |x| < 1. 35 Thus, BOLZANO obtained his pro<strong>of</strong> <strong>of</strong> the functional<br />

equation<br />

under the important assumption |x| < 1.<br />

[m] · [n] = [m + n]<br />

Extension to real exponents. EULER’S (second) pro<strong>of</strong> <strong>of</strong> the binomial theorem had<br />

focused on rational exponents. At the end <strong>of</strong> the argument, he suggested that other<br />

(positive) exponents could also be considered:<br />

31 (ibid., vi).<br />

32 (ibid., xi).<br />

33 (ibid., §38).<br />

34 This notation has been adapted from (Hauch, 1997).<br />

35 (Bolzano, 1816, §40).


204 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

“[. . . ] and thus it shows that<br />

� �<br />

i<br />

= (1 + x)<br />

a<br />

i a ,<br />

which demonstrates that our theorem is true when for the exponent n any fraction<br />

i<br />

a is taken, from this its truth is evident for all positive numbers taken in place <strong>of</strong><br />

the exponent n.” 36<br />

BOLZANO was slightly more specific on binomial expansion for irrational expo-<br />

nents. As a consequence <strong>of</strong> the meaning <strong>of</strong> (1 + x) i for i an irrational number, BOLZANO<br />

claimed, (1 + x) i could be approached as closely as desired by (1 + x) m n for m, n inte-<br />

gers. Inserting m n for i everywhere in the series and letting m n<br />

approach i, the sum<br />

would approach (1 + x) i as closely as desired. Thus, BOLZANO alluded to his concept<br />

<strong>of</strong> continuity applied to the exponentiation and to power series in order to obtain the<br />

binomial theorem for all real exponents. 37<br />

10.4 New types <strong>of</strong> series<br />

<strong>The</strong> series discussed thus far have all been power series but in the early nineteenth<br />

century, this situation changed. Series which were not power series had emerged<br />

in various contexts in the eighteenth century but became very important in the first<br />

decades <strong>of</strong> the nineteenth century, mainly through investigations in the theory <strong>of</strong> heat<br />

conducted by J. B. J. FOURIER (1768–1830). 38<br />

FOURIER’S term-wise integration. From the first decade <strong>of</strong> the nineteenth century,<br />

FOURIER had begun representing physical phenomena — mainly heat conduction —<br />

by trigonometric series. In 1822, his investigations were published as a monograph. 39<br />

One <strong>of</strong> FOURIER’S central tricks was the term-wise integration <strong>of</strong> an infinite series em-<br />

ployed to obtain the Fourier coefficients in the following way. Assuming that a function<br />

φ (x) could be expanded as<br />

φ (x) =<br />

36 “[. . . ] atque hinc in genere manifestum fore<br />

∞<br />

∑ ai sin ix, (10.2)<br />

i=1<br />

� �<br />

i<br />

= (1 + x)<br />

a<br />

i a ,<br />

ita ut iam demonstratum sit theorema nostrum verum esse, si pro exponente n fractio quaecunque<br />

i<br />

a accipiatur, unde veritas iam est evicta pro omnibus numeris positivis loco exponentis n accipiendis.”<br />

(L. Euler, 1775, 215–216).<br />

37 (Bolzano, 1816, §46).<br />

38 FOURIER and his works leading to Fourier series have been widely studied, see e.g. (Bottazzini,<br />

1986; I. Grattan-Guinness, 1972).<br />

39 (Fourier, 1822).


10.4. New types <strong>of</strong> series 205<br />

Figure 10.2: JEAN BAPTISTE JOSEPH FOURIER (1768–1830)<br />

FOURIER multiplied both sides <strong>of</strong> the equation by sin nx and integrated from 0 to π,<br />

� π<br />

0<br />

φ (x) sin nx dx =<br />

∞ �<br />

∑ ai i=1<br />

sin ix sin nx dx,<br />

where the integration <strong>of</strong> the sum was carried out term-wise. By the orthogonality,<br />

⎧<br />

�<br />

⎨<br />

π<br />

if n = i,<br />

sin ix sin nx dx = 2<br />

⎩<br />

0otherwise,<br />

FOURIER found the coefficients <strong>of</strong> the expansion (10.2) to be<br />

π<br />

2 a � π<br />

i = φ (x) sin ix dx.<br />

0<br />

<strong>The</strong> interchange <strong>of</strong> summation and integration would soon become a point <strong>of</strong> objec-<br />

tion against FOURIER’S rigor.<br />

SIMÉON-DENIS POISSON’S peculiar example. Almost simultaneous with FOURIER’S<br />

first investigations, a problem arose which also involved series which were not power-<br />

series. <strong>The</strong> problem was raised by SIMÉON-DENIS POISSON in 1811 and was inten-<br />

sively debated for the next decades, in particular in the French journal Annales de<br />

mathématiques pures et appliquées. 40<br />

40 This is well described in (Jahnke, 1987, 105–117) and (Bottazzini, 1990, lx–lxiii).


206 Chapter 10. Toward rigorization <strong>of</strong> analysis<br />

It all began when POISSON noticed that a peculiar situation arose from letting m =<br />

1 3 and x = π in the binomial expansion <strong>of</strong> (2 cos x) m , 41<br />

(2 cos x) m =<br />

∞<br />

∑<br />

n=0<br />

∏ n−1<br />

k=0 (m − k)<br />

cos ((m − 2n) x) .<br />

n!<br />

In a short paper, POISSON observed that the left hand side had the three values<br />

− 3√ 2, 3√ �<br />

1 +<br />

2<br />

√ �<br />

−3<br />

, and<br />

2<br />

3√ �<br />

1 −<br />

2<br />

√ �<br />

−3<br />

2<br />

although the right hand side was a single-valued function,<br />

cos<br />

�<br />

π<br />

� ∞ ∏<br />

3 ∑<br />

n=0<br />

n−1<br />

k=0<br />

� �<br />

13 − k<br />

n!<br />

= 1<br />

2 × (1 + 1) 1 3 =<br />

Thus, the sum <strong>of</strong> the series on the right hand side corresponded to neither <strong>of</strong> the values<br />

<strong>of</strong> the expression on the left hand side but was the average <strong>of</strong> its two complex values.<br />

Poisson’s example is a particular example <strong>of</strong> the kind <strong>of</strong> strange relations which<br />

could result by interpreting formal equalities in situations outside the domain <strong>of</strong> nu-<br />

3√ 2<br />

2 .<br />

merical equality. In this sense, it is similar to the peculiar formal equality<br />

1<br />

2<br />

= 1 − 1 + 1 − 1 + . . .<br />

which had puzzled mathematicians in the eighteenth century. However, Poisson’s ex-<br />

ample was a convergent series and the problem was that it did not agree with its true<br />

value.<br />

Upon POISSON’S publication, mathematicians sought to understand how and why<br />

this peculiarity emerged, and the debate also spread to Berlin. In Berlin, A. L. CRELLE<br />

(1780–1855) and M. OHM (1792–1872) became interested in the explanation <strong>of</strong> this<br />

phenomenon — as did the mysterious L. OLIVIER who will appear prominently in<br />

chapter 13. 42 ABEL also became acquainted with the problem and it provoked him<br />

into producing a new pro<strong>of</strong> <strong>of</strong> the binomial theorem. Before attention is focused on<br />

ABEL’S work in the theory <strong>of</strong> series, the following chapter is devoted to his greatest<br />

inspiration in the field: CAUCHY’S Cours d’analyse.<br />

41 (Poisson, 1811). POISSON wrote x = 200 ◦ and thus adhered to the new radian system.<br />

42 See e.g. (Jahnke, 1987, 105–117).


Chapter 11<br />

CAUCHY’s new foundation for analysis<br />

Against the background <strong>of</strong> J. L. LAGRANGE’S (1736–1813) algebraic foundation for the<br />

calculus, another and radically different program <strong>of</strong> rigorization emerged when A.-L.<br />

CAUCHY (1789–1857) set out to write a textbook suitable for his courses at the École<br />

Polytechnique. In a sense, CAUCHY’S famous textbook Cours d’analyse continues the<br />

Lagrangian program as its subtitle Analyse algébrique testifies, 1 but its contents consti-<br />

tuted a remarkable break with the Lagrangian system. In the Cours d’analyse, CAUCHY<br />

reformulated and revised the theory <strong>of</strong> infinite series from his novel viewpoint based<br />

on a shift in the conception <strong>of</strong> equality (see below). 2 Later, CAUCHY continued the<br />

reworking <strong>of</strong> the foundations <strong>of</strong> the calculus in two further textbooks dealing with the<br />

differential and integral calculus (see table 11.1).<br />

<strong>The</strong> Lagrangian foundation for the calculus relied on a notion <strong>of</strong> equality between<br />

functions (expressions) which was largely formal and had been inherited from L. EU-<br />

LER (1707–1783) (see chapter 10). In CAUCHY’S hands, the concept <strong>of</strong> equality shifted<br />

toward focusing on numerical or arithmetical equality: to CAUCHY, two functions<br />

were only equal if they produced equal numerical results for equal numerical values<br />

<strong>of</strong> the arguments. By way <strong>of</strong> a few central examples, I will document how this change<br />

<strong>of</strong> approach was implemented and what its consequences were.<br />

11.1 Programmatic focus on arithmetical equality<br />

Generality <strong>of</strong> algebra. Describing the methods used in the Cours d’analyse, CAUCHY<br />

stressed the way in which he had fought to obtain the standard <strong>of</strong> rigor which is char-<br />

acteristic <strong>of</strong> geometry by denouncing arguments relying on the “generality <strong>of</strong> alge-<br />

bra”. Such arguments could be, CAUCHY admitted, suitable inductions for obtaining<br />

1 (A.-L. Cauchy, 1821b).<br />

2 CAUCHY’S Cours d’analyse marks a turning point in the history <strong>of</strong> the calculus and has been given<br />

due attention by historians <strong>of</strong> mathematics. <strong>The</strong> most comprehensive presentation is probably BOT-<br />

TAZZINI’S introduction to a photographic reproduction <strong>of</strong> the Cours d’analyse (Bottazzini, 1990). In<br />

particular, the section entitled <strong>The</strong> “Generality <strong>of</strong> Algebra” (ibid., xliv–xcvii) is <strong>of</strong> direct relevance to<br />

the present discussion.<br />

207


208 Chapter 11. CAUCHY’s new foundation for analysis<br />

1821 Cours d’analyse d’École Royale Polytechnique.<br />

Première partie. Analyse algébrique<br />

1823 Résumé des leçons données a l’École<br />

Royale Polytechnique sur le calcul infinitésimal<br />

1829 Leçons sur le calcul différentielle<br />

Table 11.1: CAUCHY’s textbooks on the calculus<br />

the truth but should never be allowed to act as exact pro<strong>of</strong>s. In particular, CAUCHY<br />

mentioned how arguments by the generality <strong>of</strong> algebra had been leading mathemati-<br />

cians into unfounded passages “from convergent to divergent series, from real quan-<br />

tities to imaginary expressions”. 3 CAUCHY continued,<br />

“Similarly, one should realize that they [arguments by the generality <strong>of</strong> algebra]<br />

tend to attribute to algebraic formulae an indefinite extension whereas in<br />

reality, the majority <strong>of</strong> these formulae only subsists under certain conditions and<br />

for certain values <strong>of</strong> the quantities which they contain.” 4<br />

Important examples <strong>of</strong> the problems which this requirement addressed was the<br />

relationship between formulae such as<br />

1<br />

1 − x and<br />

∞<br />

∑ x<br />

n=0<br />

n<br />

(11.1)<br />

which have been described above. Mathematicians unknowingly adhering to the for-<br />

mal concept <strong>of</strong> equality had been aware that counter-intuitive results could emerge<br />

if numerical values were inserted into the two expressions and their equality was ex-<br />

tended to cover numerical values. For instance,<br />

1<br />

1 − (−1)<br />

1<br />

= , but<br />

2<br />

∞<br />

∑ (−1)<br />

n=0<br />

n = 1 − 1 + 1 − 1 + . . .<br />

and the sum did certainly not represent the value 1 2 in any numerical sense.<br />

CAUCHY removed these anomalies by dismissing the concept <strong>of</strong> formal equality<br />

and carefully analyzing the conditions under which a numerical equality between ex-<br />

pressions such as (11.1) would hold. Thus, he found and emphasized that for numer-<br />

ical equality it was required that |x| < 1, and consequently the peculiar results were<br />

all explained away.<br />

3 (A.-L. Cauchy, 1821a, iii).<br />

4 “On doit même observer qu’elles tendent à faire attribuer aux formules algébriques une étendue<br />

indéfinie, tandis que, dans la réalité, la plupart de ces formules subsistent uniquement sous certaines<br />

conditions, et pour certaines valeurs des quantités qu’elles renferment.” (ibid., iii).


11.2. CAUCHY’s concepts <strong>of</strong> limits and infinitesimals 209<br />

11.2 CAUCHY’s concepts <strong>of</strong> limits and infinitesimals<br />

In the preliminaries, CAUCHY defined what he understood under the terms limit and<br />

infinitely small. <strong>The</strong> two concepts are closely related in CAUCHY’S book although their<br />

interrelation is not trivial. <strong>The</strong> proper interpretation <strong>of</strong> the concepts has sparked some<br />

controversy in the historical literature which has sometimes chosen to focus on one<br />

<strong>of</strong> the concepts and neglecting the other. To be fair to the source and to facilitate a<br />

discussion <strong>of</strong> N. H. ABEL’S (1802–1829) reading <strong>of</strong> it, both CAUCHY’S definitions are<br />

reproduced and translated here. 5<br />

First, CAUCHY defined his concept <strong>of</strong> limit:<br />

“Whenever the values successively attributed to one and the same variable<br />

approach a fixed value indefinitely in such a way that it eventually differs from it<br />

by as little as one desires, the latter is called the limit <strong>of</strong> all the others.” 6<br />

A few sentences later, CAUCHY gave his definition <strong>of</strong> infinitely small quantities which<br />

he based on the notion <strong>of</strong> limits:<br />

“Whenever the successive numerical values <strong>of</strong> one and the same variable decrease<br />

indefinitely in such a way that they become less than any given number,<br />

this variable becomes what is called infinitely small or an infinitely small quantity. A<br />

variable <strong>of</strong> this kind has zero for limit.” 7<br />

Various conceptions <strong>of</strong> limits and infinitesimals had previously been suggested as<br />

the foundations for the calculus and EULER called the calculus the “algebra <strong>of</strong> zeros”<br />

because <strong>of</strong> his prolific use <strong>of</strong> infinitesimals. However, CAUCHY gave a process-based<br />

definition <strong>of</strong> limits and — more importantly — showed how to work with it. Corre-<br />

spondingly, to CAUCHY, infinitesimals were variable quantities which were involved<br />

in limit processes and could be made as small as desired by particular choices <strong>of</strong> the<br />

variable <strong>of</strong> the limit process.<br />

It has puzzled certain historians <strong>of</strong> mathematics why CAUCHY simultaneously<br />

employed limits and retained the older concept <strong>of</strong> (completed) infinitesimals. 8 <strong>The</strong><br />

fact remains that CAUCHY employed both concepts in different pro<strong>of</strong>s and probably<br />

thought <strong>of</strong> them as equivalent but suited for different purposes. 9<br />

Importantly, CAUCHY sometimes used symbols to denote infinitely small quanti-<br />

ties which were really (according to the definition) variables which tended toward the<br />

5 A good interpretation is provided in (Grabiner, 1981b, 80–81).<br />

6 “Lorsque les valeurs successivement attribuées à une même variable s’approchent indéfiniment<br />

d’une valeur fixe, de manière à finir par en différer aussi peu que l’on voudra, cette dernière est<br />

appelée la limite de toutes les autres.” (A.-L. Cauchy, 1821a, 4).<br />

7 “Lorsque les valeurs numériques successives d’une même variable décroissent indéfiniment, de<br />

manière à s’abaisser au-dessous de tout nombre donné, cette variable devient ce qu’on nomme<br />

un infiniment petit ou une quantité infiniment petite. Une variable de cette espèce a zéro pour<br />

limite.” (ibid., 4).<br />

8 See discussion in (Lützen, 1999, 198–211).<br />

9 See e.g. (Grabiner, 1981b, 87).


210 Chapter 11. CAUCHY’s new foundation for analysis<br />

limit zero. However, by introducing symbols, the process under which the variable<br />

vanished was obscured and the order in which limit processes were conducted was<br />

not explicit.<br />

11.3 Divergent series have no sum<br />

When CAUCHY extended the procedure <strong>of</strong> analyzing the requirements for numerical<br />

equality involving series, he was led to a conclusion which he knew would be painful<br />

for his contemporaries to accept.<br />

“It is true that in order to always remain faithful to these principles, I see myself<br />

forced to accept multiple propositions which may appear a bit harsh at first<br />

sight. For instance, in the sixth chapter, I announce that a divergent series has no<br />

sum.” 10<br />

CAUCHY’S treatment <strong>of</strong> series began with series <strong>of</strong> positive real terms (section<br />

VI.2), was then extended to series <strong>of</strong> general real terms (section VI.3), before he went<br />

on to treat series with complex terms (chapter IX). In the sixth chapter, CAUCHY elab-<br />

orated his definition <strong>of</strong> convergence and his attitude toward divergent series.<br />

“Let<br />

sn = u0 + u1 + u2 + · · · + un−1<br />

be the sum <strong>of</strong> the first n terms where n designates any integer. If, for ever increasing<br />

values <strong>of</strong> n, the sum sn approaches a certain limit s indefinitely, the series is<br />

said to be convergent and the above mentioned limit is called the sum <strong>of</strong> the series.<br />

In the contrary case, if the sum sn does not approach any fixed limit when n<br />

increases indefinitely, the series is divergent and no longer has a sum.” 11<br />

As the quotations demonstrate, CAUCHY sought to limit the concept <strong>of</strong> “sum <strong>of</strong> a<br />

series” to apply only to convergent series. This position was radicalized by ABEL in<br />

his correspondence as we will see in section 12.3: ABEL wanted an outright ban on<br />

divergent series and saw them as the creation <strong>of</strong> the Devil. To CAUCHY, who was also<br />

a very creative mathematician outside fundamental issues, divergent series remained<br />

<strong>of</strong> interest in asymptotic mathematics; only in questions <strong>of</strong> foundational nature, they<br />

were not attributed any sum.<br />

10 “Il est vrai que, pour rester constamment fidèle à ces principes, je me suis vu forcé d’admettre plusieurs<br />

propositions qui paraîtront peut-être un peu dures au premier abord. Par exemple, j’énonce<br />

dans le chapitre VI, qu’un série divergente n’a pas de somme [. . . ]” (A.-L. Cauchy, 1821a, iv).<br />

11 “Soit<br />

sn = u0 + u1 + u2 + · · · + un−1<br />

la somme des n premiers termes, n désignant un nombre entier quelconque. Si, pour des valeurs<br />

de n toujours croissantes, la somme sn s’approche indéfiniment d’une certaine limite s; la série sera<br />

dite convergente, et la limite en question s’appellera la somme de la série. Au contraire, si, tandis<br />

que n croît indéfiniment, la somme sn ne s’approche d’aucune limite fixe, la série sera divergente, et<br />

n’aura plus de somme.” (ibid., 123).


11.3. Divergent series have no sum 211<br />

For the moment, focus will be given to another intrinsic aspect <strong>of</strong> CAUCHY’S defini-<br />

tion: the position <strong>of</strong> certain concepts and theorems within the theoretical framework.<br />

If only convergent series were allowed to have a sum, and knowledge <strong>of</strong> the value <strong>of</strong><br />

the sum was required to apply the definition to determine whether the series was con-<br />

vergent or not, a problem emerged. Phrased in modern notation, it is not practical to<br />

attempt establishing that some s exists such that<br />

|sn − s| → 0 as n → ∞<br />

without knowing which s could be a candidate; for complicated or general series, such<br />

candidates might not be available. <strong>The</strong>refore, CAUCHY’S theoretical complex — in<br />

an essential way — required a means <strong>of</strong> investigating convergence without any prior<br />

knowledge <strong>of</strong> the purported sum. This important problem was met by a number<br />

<strong>of</strong> criteria <strong>of</strong> convergence which were also among the chief innovations in the Cours<br />

d’analyse.<br />

After giving his definition <strong>of</strong> convergence, CAUCHY gave a first characterization <strong>of</strong><br />

his new concept, which was, however, <strong>of</strong> little use in practically establishing conver-<br />

gence (see below).<br />

“Thus, for the series u0 + u1 + u2 + · · · + un + . . . to be convergent, it is first<br />

necessary that the general term un decreases indefinitely when n grows. However,<br />

this condition does not suffice and it must also be so that for increasing values <strong>of</strong><br />

n the different sums<br />

i.e. the sums <strong>of</strong> the quantities<br />

un + un+1,<br />

un + un+1 + un+2,<br />

etc.,<br />

un, un+1, un+2, etc.<br />

taken starting from the first and to whatever number one may wish, eventually always<br />

produce numerical values less than any assignable limit. Conversely, whenever<br />

these two conditions are fulfilled, the convergence <strong>of</strong> the series is assured.” 12<br />

12 “Donc, pour que la série (1) soit convergente, il est d’abord nécessaire que le terme général un<br />

décroisse indéfiniment, tandis que n augmente; mais cette condition ne suffit pas, et il faut encore<br />

que, pour des valeurs croissantes de n, les différentes sommes<br />

c’est-à-dire, les sommes des quantités<br />

un + un+1,<br />

un + un+1 + un+2,<br />

&c. . .<br />

un, un+1, un+2, &c. . .<br />

prises, à partir de la première, en tel nombre que l’on voudra, finissent par obtenir constamment<br />

des valeurs numériques inférieures à toute limite assignable. Réciproquement, lorsque ces diverses<br />

conditions sont remplies, la convergence de la série est assurée.” (ibid., 125–126).


212 Chapter 11. CAUCHY’s new foundation for analysis<br />

In the second half <strong>of</strong> the nineteenth century, CAUCHY’S characterization <strong>of</strong> conver-<br />

gence by means <strong>of</strong> the so-called Cauchy criterion became even more important when it<br />

was realized, that no pro<strong>of</strong> <strong>of</strong> CAUCHY’S last assertion could be given. <strong>The</strong> solution<br />

devised by mathematicians such as J. W. R. DEDEKIND (1831–1916) and G. CANTOR<br />

(1845–1918) to secure the validity <strong>of</strong> CAUCHY’S second claim was to construct the real<br />

numbers in such ways that they possessed this property <strong>of</strong> completeness.<br />

11.4 Means <strong>of</strong> testing for convergence <strong>of</strong> series<br />

With his new emphasis on convergence, CAUCHY’S theory needed criteria <strong>of</strong> conver-<br />

gence which would operate simply from the general terms without any information<br />

about the sum. Based on previous observations, CAUCHY established three important<br />

such tests which are still all important today. First, he proved the root test to the effect<br />

that if n√ un has a limit k as n → ∞, the series ∑ un will be convergent if k < 1 and<br />

divergent if k > 1. In case k = 1, nothing could be said <strong>of</strong> the convergence by this<br />

criterion. CAUCHY then transformed the root test to obtain the ratio test (see below)<br />

and a logarithmic criterion by comparison with the harmonic series.<br />

It is a general feature <strong>of</strong> these criteria <strong>of</strong> convergence that there are cases for which<br />

they do not provide any information concerning convergence. As we will see in chap-<br />

ter 13, the search for a complete test <strong>of</strong> convergence which could always determine the<br />

convergence or divergence <strong>of</strong> series from its general term was also actively pursued<br />

in the 1820s.<br />

CAUCHY’S pro<strong>of</strong> <strong>of</strong> the ratio test. CAUCHY’S proved his criteria <strong>of</strong> convergence by<br />

an ingenious route albeit complicated. 13 First, he proved the root test directly from<br />

the assumptions and the previously established convergence <strong>of</strong> the geometric pro-<br />

gression. Next, he referred to a previously established theorem to the effect that if<br />

the sequences { n√ � �<br />

un+1<br />

un} and were both convergent, their limits would be equal.<br />

un<br />

Ultimately, this theorem provided the pro<strong>of</strong> <strong>of</strong> the ratio test. To get a grasp <strong>of</strong> the way<br />

CAUCHY reasoned with his concepts and the way in which he obtained his criteria,<br />

the details <strong>of</strong> his pro<strong>of</strong> are considered. 14 Later, after ABEL’S way <strong>of</strong> commencing the<br />

theory <strong>of</strong> series has been described, the role <strong>of</strong> the ratio test in the two theories can be<br />

discussed.<br />

CAUCHY proved the convergence part <strong>of</strong> the root test by letting U denote a number<br />

k < U < 1. <strong>The</strong>n, he observed that a large integer n had to exist such that for any larger<br />

number, say N ≥ n,<br />

N√ uN < U, i.e. uN < U N .<br />

13 (A.-L. Cauchy, 1821a, 132–135).<br />

14 Today, this theorem is standard textbook material in basic calculus courses. <strong>The</strong> modern pro<strong>of</strong><br />

closely resembles CAUCHY’S pro<strong>of</strong>.


11.4. Means <strong>of</strong> testing for convergence <strong>of</strong> series 213<br />

Thus, the tail <strong>of</strong> the series was term-wise less than a convergent geometric progression,<br />

and the convergence <strong>of</strong> the series ∑ un was concluded. Similarly, CAUCHY proved the<br />

divergence part <strong>of</strong> the root test by comparing with a divergent geometric progression.<br />

CAUCHY based his pro<strong>of</strong> <strong>of</strong> the ratio test on the following result which he had<br />

previously obtained.<br />

“2nd theorem. If the function f (x) is positive for large values <strong>of</strong> x and the<br />

ratio<br />

f (x + 1)<br />

f (x)<br />

converges toward the limit k when x increases indefinitely, the expression<br />

[ f (x)] 1 x<br />

will converge at the same time toward the same limit.” 15<br />

In his pro<strong>of</strong>, CAUCHY distinguished between two cases namely k finite or not; we<br />

shall here only be concerned with the case where k is finite. CAUCHY let ε denote an<br />

as yet unspecified number which was presumably very small. By assuming that for<br />

x ≥ h,<br />

f (x + 1)<br />

f (x)<br />

∈ [k − ε, k + ε] ,<br />

CAUCHY found by the theory <strong>of</strong> means which he developed in a note 16 that the geo-<br />

metric mean 17 <strong>of</strong><br />

f (h + 1)<br />

,<br />

f (h)<br />

f (h + 2)<br />

, . . . ,<br />

f (h + 1)<br />

f (h + n)<br />

f (h + n − 1) ,<br />

would also belong to this interval, i.e.<br />

�<br />

n f (h + n)<br />

= k + α, α ∈ [−ε, ε] .<br />

f (h)<br />

<strong>The</strong>n, CAUCHY found by inserting x = h + n<br />

which meant<br />

f (x) = f (h) · (k + α) x−h<br />

f (x) 1 x = f (h) 1 x · (k + α) 1− h x →<br />

x→∞ k + α.<br />

15 “2.e Théorème. Si, la fonction f (x) étant positive pour de très-grandes valeurs de x, le rapport<br />

f (x + 1)<br />

f (x)<br />

converge, tandis que x croit indéfiniment, vers la limite k, l’expression<br />

[ f (x)] 1 x<br />

convergera en même temps vers la même limite.” (ibid., 53–54).<br />

16 (ibid., note II).<br />

17 <strong>The</strong> geometric mean <strong>of</strong> the quantities a1, . . . , an was the quantity n� ∏ n k=1 a k.


214 Chapter 11. CAUCHY’s new foundation for analysis<br />

Thus, the limit <strong>of</strong> f (x) 1 x belonged to the same arbitrarily small interval as the limit<br />

<strong>of</strong><br />

f (x+1)<br />

f (x) , and thus the two limits were equal. Now, the ratio test followed by letting<br />

f (n) = un and observing that the limits <strong>of</strong> n√ un and u n+1<br />

un coincided.<br />

11.5 CAUCHY’s pro<strong>of</strong> <strong>of</strong> the binomial theorem<br />

CAUCHY agreed with his predecessors in considering the binomial theorem a corner<br />

stone <strong>of</strong> the calculus. His pro<strong>of</strong> <strong>of</strong> it relied on and promoted two <strong>of</strong> his new techniques<br />

and concepts in analysis: those <strong>of</strong> functional equations and continuous functions. 18 As<br />

EULER had done, CAUCHY considered the functional equation<br />

<strong>of</strong> which he knew that the binomial<br />

f (m) f (n) = f (m + n) (11.2)<br />

f (m) = (1 + x) m<br />

(11.3)<br />

was a continuous solution for all m provided x was fixed. On the other hand, the<br />

function defined by the infinite series<br />

∞<br />

∑<br />

k=0<br />

� �<br />

m<br />

x<br />

k<br />

k<br />

(11.4)<br />

was also a solution to the functional equation (11.2) under the assumptions that it<br />

converged and m was a rational number.<br />

To demonstrate that the series (11.4) satisfied the functional equation, CAUCHY<br />

had to be able to multiply infinite series. He invented a way <strong>of</strong> multiplying absolutely<br />

convergent series which rigorously established the convergence <strong>of</strong> the product. 19 Based<br />

on the argument which EULER had also used, CAUCHY then knew that the series (11.4)<br />

coincided with f (m) for all rational values <strong>of</strong> m. <strong>The</strong>refore, the general equality <strong>of</strong><br />

(11.3) and (11.4) would be proved if the series was a continuous function <strong>of</strong> m. In order<br />

to prove the continuity <strong>of</strong> the series (11.4), CAUCHY devised and proved a general<br />

theorem to the effect that a convergent sum <strong>of</strong> continuous functions was always a<br />

continuous function. Later, this theorem would arouse much controversy (see below).<br />

CAUCHY’S way <strong>of</strong> multiplying infinite series. In the Cours d’analyse, CAUCHY in-<br />

vented a way <strong>of</strong> multiplying two absolutely convergent series such that the product<br />

would be a new convergent series. As he had done throughout, CAUCHY developed<br />

his theory <strong>of</strong> infinite series in three steps:<br />

1. Series <strong>of</strong> real, positive terms (section VI.2)<br />

18 For CAUCHY’S theory <strong>of</strong> functional equations, see (J. Dhombres, 1992).<br />

19 (A.-L. Cauchy, 1821a, 157).


11.5. CAUCHY’s pro<strong>of</strong> <strong>of</strong> the binomial theorem 215<br />

2. Series <strong>of</strong> general real terms (section VI.3)<br />

3. Series <strong>of</strong> complex terms (chapter IX)<br />

In each <strong>of</strong> these three theories, CAUCHY developed a product theorem, 20 and CAUCHY’S<br />

pro<strong>of</strong>s will be relevant when compared with ABEL’S subsequent pro<strong>of</strong>s. <strong>The</strong> multipli-<br />

cation theorems applying to series <strong>of</strong> real terms (the first and second) are the most in-<br />

teresting for the present study. In his theory <strong>of</strong> series <strong>of</strong> positive terms, CAUCHY had<br />

stated and proved that if two series ∑ un and ∑ vn were convergent and converged<br />

toward s and s ′ , the series whose general term was<br />

wk = ∑<br />

n+m=k<br />

unvm<br />

(11.5)<br />

would be convergent and converge toward the product ss ′ . When he wanted to gener-<br />

alize this theorem to general series <strong>of</strong> real terms, CAUCHY imposed the restriction that<br />

each <strong>of</strong> the series ∑ un and ∑ vn was to be convergent when their terms were replaced<br />

by their absolute values, i.e. both factors were to be absolutely convergent, although the<br />

term and an elaborate concept was only invented some years later (see section 12.7).<br />

In the first case, in which all terms were positive quantities, CAUCHY proved the<br />

theorem by a direct argument. He let s ′′<br />

n designate the sum <strong>of</strong> the first n terms <strong>of</strong> the<br />

purported product series (11.5) and defined<br />

to obtain<br />

He had thus obtained<br />

⎧<br />

⎪⎨<br />

n − 1<br />

if n is odd<br />

m = 2<br />

⎪⎩ n − 2<br />

if n is even<br />

2<br />

n−1<br />

∑ wk <<br />

k=0<br />

n−1<br />

∑ wk ><br />

k=0<br />

�<br />

n−1<br />

∑ uk k=0<br />

�<br />

m<br />

∑ uk k=0<br />

sm+1s ′ m+1<br />

� � n−1<br />

∑<br />

vk k=0<br />

� �<br />

m<br />

∑ vk k=0<br />

�<br />

�<br />

< s′′<br />

n < sns ′ n,<br />

.<br />

and<br />

and by letting m grow beyond all bounds, the theorem was established.<br />

When he came to generalize this theorem to the case <strong>of</strong> general real terms, CAUCHY<br />

wanted to apply the simpler case <strong>of</strong> series with positive terms. With the same notation<br />

as above, he obtained the formula<br />

20 (ibid., 141–142,147–149,283–285).<br />

s ′ nsn − s ′′<br />

2n−2<br />

n =<br />

∑<br />

t=n<br />

∑<br />

m+k=t<br />

umv k, (11.6)


216 Chapter 11. CAUCHY’s new foundation for analysis<br />

and if the terms um and v k were positive, the difference (11.6) would converge to zero<br />

as a consequence <strong>of</strong> the theorem for series <strong>of</strong> positive terms. If the terms were not all<br />

positive, CAUCHY could still apply the theorem to the corresponding series <strong>of</strong> numer-<br />

ical values ρn = |un| and ρ ′ n = |vn|. <strong>The</strong>refore, in this case, the difference (11.6) would<br />

still converge to zero and yet majorize the series <strong>of</strong> the original terms. Thus, CAUCHY<br />

had established the convergence and the product equation for absolutely convergent<br />

series <strong>of</strong> real terms.<br />

CAUCHY’S concept <strong>of</strong> continuous functions. CAUCHY’S concept <strong>of</strong> continuous func-<br />

tions (see below) was among the main innovations <strong>of</strong> his new calculus. <strong>The</strong>re, in the<br />

definition and in the pro<strong>of</strong>s, his new foundation on an algebraic concept <strong>of</strong> limits<br />

played its most important role. In the eighteenth century, EULER had used the term<br />

continuous to indicate that the function was defined by the same analytic expression<br />

throughout its domain. 21 However, in the Cours d’analyse, CAUCHY took it upon him-<br />

self to completely redefine the concept <strong>of</strong> continuous functions in order to capture a<br />

different property <strong>of</strong> the functions: their continuous, unbroken variation.<br />

“Let f (x) be a function <strong>of</strong> the variable x and suppose that for every value <strong>of</strong><br />

x between two given boundaries this function always takes a unique and finite<br />

value. If, starting from a value <strong>of</strong> x contained between these boundaries, one attributes<br />

to x an infinitely small increment α, the function will receive the increment<br />

f (x + α) − f (x)<br />

which depends simultaneously on the new variable α and the value <strong>of</strong> x. Between<br />

the two boundaries assigned to the variable x, the function f (x) will be a continuous<br />

function <strong>of</strong> this variable if for every value <strong>of</strong> x between these boundaries, the<br />

numerical value <strong>of</strong> the difference<br />

f (x + α) − f (x)<br />

decreases indefinitely with that [numerical value] <strong>of</strong> α. In other words, the function<br />

f (x) remains continuous with respect to x between the given boundaries if, between<br />

these boundaries, an infinitely small increment <strong>of</strong> the variable produces an infinitely small<br />

increment <strong>of</strong> the function.” 22<br />

21 (L. Euler, 1748). See (Lützen, 1978) and (Youschkevitch, 1976).<br />

22 “Soit f (x) une fonction de la variable x, et supposons que, pour chaque valeur de x intermédiaire<br />

entre deux limites données, cette fonction admette constamment une valeur unique et finie. Si, en<br />

partant d’une valeur de x comprise entre ces limites, on attribue à la variable x un accroissement<br />

infiniment petit α, la fonction elle-même recevra pour accroissement la différence<br />

f (x + α) − f (x) ,<br />

qui dépendra en même temps de la nouvelle variable α et da la valeur de x. Cela posé, la fonction<br />

f (x) sera, entre les deux limites assignées à la variable x, fonction continue de cette variable, si,<br />

pour chaque valeur de x intermédiaire entre ces limites, la valeur numérique de la différence<br />

f (x + α) − f (x)<br />

décroit indéfiniment avec celle de α. En d’autres termes, la fonction f (x) restera continue par rapport<br />

à x entre les limites données, si, entre ces limites, un accroissement infiniment petit de la va-


11.5. CAUCHY’s pro<strong>of</strong> <strong>of</strong> the binomial theorem 217<br />

CAUCHY’S novel definition defines continuity not at a point (as is customary to-<br />

day, see below) but on an entire interval enclosed by two boundary points. 23 Thus,<br />

CAUCHY’S implicit choice <strong>of</strong> infinitesimal ω such that<br />

| f (x + α) − f (x)| = ω<br />

could seem to be independent <strong>of</strong> x on the interval and the definition would actually<br />

be <strong>of</strong> what is now known as a uniformly continuous function. <strong>The</strong> doubt over the proper<br />

interpretation is introduced by the fact CAUCHY’S use <strong>of</strong> the symbol ω to designate<br />

the infinitesimal: it “hides” the order in which the limit processes are to be carried out.<br />

CAUCHY and series <strong>of</strong> functions. In a famous theorem which provided an impor-<br />

tant step in CAUCHY’S pro<strong>of</strong> <strong>of</strong> the binomial theorem, CAUCHY sought to link the<br />

concepts <strong>of</strong> convergence and continuity. Since we will discuss the theorem in details<br />

in chapter 12, its entire wording and CAUCHY’S pro<strong>of</strong> <strong>of</strong> it are reproduced here.<br />

“1st theorem. Whenever the different terms <strong>of</strong> the series u0 + u1 + u2 + · · · + un +<br />

. . . are functions <strong>of</strong> one and the same variable x and continuous with respect to this variable<br />

in the neighborhood <strong>of</strong> a particular value for which the series is convergent, the sum<br />

s <strong>of</strong> the series is also a continuous function <strong>of</strong> x in the neighborhood <strong>of</strong> that particular<br />

value.” 24<br />

As was customary, CAUCHY actually presented the pro<strong>of</strong> before he made the theo-<br />

rem explicit. <strong>The</strong> pro<strong>of</strong> which he gave proceeded along the following lines. If the sum<br />

is split after n terms<br />

s = sn + rn =<br />

n−1<br />

∑ un +<br />

k=0<br />

∞<br />

∑ un, (11.7)<br />

k=n<br />

the partial sum sn is a polynomial and therefore continuous and the remainder rn can<br />

be made less than any given quantity by the convergence <strong>of</strong> the series. In consequence,<br />

the difference s (x + α) − s (x) could be made less than any assignable quantity and<br />

the sum was therefore continuous. As we are well aware today, with our common<br />

interpretations <strong>of</strong> the basic concepts <strong>of</strong> continuity, limits, and convergence, the theo-<br />

rem is false as stated. In section 14.1.2, its future history through the works <strong>of</strong> ABEL,<br />

P. L. VON SEIDEL (1821–1896) (and G. G. STOKES (1819–1903)) and CAUCHY again<br />

is outlined to understand how ABEL’S contribution to rigorization was accepted and<br />

interpreted.<br />

riable produit toujours un accroissement infiniment petit de la fonction elle-même.” (A.-L. Cauchy,<br />

1821a, 34–35).<br />

23 See also (Bottazzini, 1990, lxxxi–lxxxiii) and (Giusti, 1984).<br />

24 “1.er Théorème. Lorsque les différens termes de la série (1) sont des fonctions d’une même variable<br />

x, continues par rapport à cette variable dans le voisinage d’une valeur particulière pour laquelle la<br />

série est convergente, la somme s de la série est aussi, dans le voisinage de cette valeur particulière,<br />

fonction continue de x.” (A.-L. Cauchy, 1821a, 131–132).


218 Chapter 11. CAUCHY’s new foundation for analysis<br />

For now, we have to stress the importance this theorem played in CAUCHY’S pro<strong>of</strong><br />

<strong>of</strong> the binomial theorem. <strong>The</strong> argument was, that since both the binomial (1 + x) m and<br />

the infinite power series (11.4) satisfied the same functional equation for all rational m,<br />

and since they were both continuous functions <strong>of</strong> m, they would coincide for all real<br />

values <strong>of</strong> m. Thus, CAUCHY proved the binomial theorem for all real exponents (here<br />

designated m).<br />

In CAUCHY’S theory <strong>of</strong> rigorously restructuring calculus, the binomial theorem<br />

played an extremely important role. 25 It provided one <strong>of</strong> the basic, theoretical bricks<br />

which could be used to rebuild the existing theory <strong>of</strong> real analysis. Historians have<br />

argued that CAUCHY’S rigorization and his pro<strong>of</strong> <strong>of</strong> the binomial theorem constitute<br />

a veiled attack on Fourier series. 26 However, although CAUCHY was critical toward<br />

J. B. J. FOURIER’S (1768–1830) reasoning, I find such a hypothesis largely unnecessary<br />

as CAUCHY’S rigorization program makes good sense from its own premises. 27<br />

11.6 Early reception <strong>of</strong> CAUCHY’s new rigor<br />

CAUCHY’S Cours d’analyse dealt exclusively with the theory functions from the per-<br />

spective <strong>of</strong> infinite series. Later in the 1820s, he also published lectures pertaining to<br />

rigorously founding the theory <strong>of</strong> differentiation and integration. 28<br />

As a textbook for the École Polytechnique, the Cours d’analyse was not successful.<br />

Because <strong>of</strong> internal animosities among the teachers, CAUCHY’S textbook was never<br />

used as a textbook but it may have served as inspiration for students preparing for<br />

the entrance exams <strong>of</strong> the school. Among his fellow mathematicians, CAUCHY’S pro-<br />

gram also received mixed reactions. In Germany, A. L. CRELLE (1780–1855) men-<br />

tioned CAUCHY’S textbooks in very positive terms, 29 and a German translation <strong>of</strong> the<br />

Cours d’analyse appeared in 1828. 30 However, a distinct German reaction also existed<br />

which sought to continue the formal, algebraic approach to foundations <strong>of</strong> analysis in<br />

the so-called combinatorial school initiated by C. F. HINDENBURG (1741–1808), and M.<br />

OHM (1792–1872) pursued his own rigorization program. 31<br />

Thus, in the 1820s, the mathematical community could be divided into three camps<br />

reflecting their attitudes toward rigor:<br />

1. Some had picked up CAUCHY’S vision <strong>of</strong> a rigorization <strong>of</strong> analysis; both its<br />

theme and its tools. <strong>The</strong>y joined in the restriction to arithmetical equality and<br />

adopted CAUCHY’S redefinition <strong>of</strong> central concepts in terms <strong>of</strong> limits.<br />

25 (Grabiner, 1981b, 111).<br />

26 See e.g. (Bottazzini, 1986, 110).<br />

27 (Grabiner, 1981b, 111).<br />

28 (A.-L. Cauchy, 1823; A.-L. Cauchy, 1829).<br />

29 (Crelle, 1827; Crelle, 1828).<br />

30 (A. L. Cauchy, 1828).<br />

31 See (Jahnke, 1992).


11.6. Early reception <strong>of</strong> CAUCHY’s new rigor 219<br />

2. Others, primarily Germans, felt a similar need for a new rigorous foundation<br />

<strong>of</strong> the calculus. Often inspired by needs stimulated by the German educational<br />

reforms, they sought to distill a combinatorial theory from the approaches <strong>of</strong><br />

LAGRANGE and others.<br />

3. <strong>The</strong> rest, and the vast majority, were mostly concerned with contributing new<br />

mathematical knowledge. Although they probably also sometimes worried about<br />

the foundations <strong>of</strong> their discipline, they left it to the teachers and experts to<br />

straighten it out.<br />

In the 1820s, the need for rigorization had entered the agenda and at least two<br />

programs had been suggested for resolving the need. In the course <strong>of</strong> the century,<br />

the need became even more urgent and CAUCHY’S conception became the preferred<br />

solution — after it had undergone the interpretations <strong>of</strong> ABEL, G. P. L. DIRICHLET<br />

(1805–1859), G. F. B. RIEMANN (1826–1866), K. T. W. WEIERSTRASS (1815–1897) and<br />

others.


Chapter 12<br />

ABEL’s reading <strong>of</strong> CAUCHY’s new rigor<br />

and the binomial theorem<br />

N. H. ABEL (1802–1829) was one <strong>of</strong> the first converts to A.-L. CAUCHY’S (1789–1857)<br />

new program <strong>of</strong> rigorizing analysis. Alone and together with B. M. HOLMBOE (1795–<br />

1850), ABEL had studied L. EULER’S (1707–1783) works on analysis and other text-<br />

books on the calculus from the late eighteenth century. 1 But during his time in Berlin,<br />

ABEL learnt <strong>of</strong> CAUCHY’S textbooks Cours d’analyse and Resumé des leçons. 2 His ac-<br />

quaintance with these works left clear traces in his publications and letters. Although<br />

CAUCHY’S rigorization program comprised all <strong>of</strong> analysis (at least in principle), ABEL<br />

was particularly interested in the theory <strong>of</strong> series. ABEL’S interest in the theory <strong>of</strong><br />

series manifested itself in two publications, numerous remarks in letters, and an inter-<br />

esting draft in one <strong>of</strong> his notebooks.<br />

ABEL’S most prestigious contribution to the rigorization <strong>of</strong> analysis was a paper<br />

published in 1826 which contained a new pro<strong>of</strong> <strong>of</strong> the binomial theorem. 3 ABEL had<br />

previously employed a complete induction argument to deduce the binomial formula. 4<br />

<strong>The</strong> method <strong>of</strong> complete induction had previously been used by B. BOLZANO (1781–<br />

1848) in his Der binomische Lehrsatz, 5 and together with ABEL’S curious and flattering<br />

remark concerning BOLZANO (see page 42), this might suggest that ABEL was familiar<br />

with this work. However, there is no direct evidence to support this speculation.<br />

When it came to proving the binomial theorem, ABEL followed the path set out by<br />

CAUCHY’S Cours d’analyse, which he praised highly, recommending it to anybody who<br />

loved the rigor <strong>of</strong> mathematics. 6 However, in certain details, ABEL’S deduction dif-<br />

fered slightly from the guideline <strong>of</strong> the Cours d’analyse. In particular, ABEL presented<br />

his own way <strong>of</strong> deducing the important ratio tests and when he found that Cauchy’s<br />

1 See the sections 3.2 and 2.2, above.<br />

2 (A.-L. Cauchy, 1821a; A.-L. Cauchy, 1823).<br />

3 (N. H. <strong>Abel</strong>, 1826f).<br />

4 (N. H. <strong>Abel</strong>, 1826b).<br />

5 (Bolzano, 1816, §7–10).<br />

6 (N. H. <strong>Abel</strong>, 1826f, 313).<br />

221


222 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

<strong>The</strong>orem “suffered exceptions”, he replaced it with a tailored — but also problematic —<br />

theorem which he found sufficient to carry through his argument. <strong>The</strong> description and<br />

analysis <strong>of</strong> all these aspects and the differences between ABEL’S and CAUCHY’S con-<br />

cepts and pro<strong>of</strong>s are the purposes <strong>of</strong> the present chapter.<br />

<strong>The</strong> Cours d’analyse was certainly ABEL’S main inspiration for his research on the<br />

theory <strong>of</strong> series. From another one <strong>of</strong> his letters, 7 we learn that ABEL had bought<br />

and read the first nine issues <strong>of</strong> CAUCHY’S Exercises des mathématiques. Although they<br />

proved highly interesting in another context, 8 these installments contained nothing<br />

with an explicit bearing on the rigorization <strong>of</strong> the theory <strong>of</strong> series. Another one <strong>of</strong><br />

CAUCHY’S publications — the Resumé des leçons — did address infinite series and, as<br />

noticed, we know from one <strong>of</strong> his letters that ABEL was familiar with it. Continu-<br />

ing the quotation on page 31 taken from a letter to HOLMBOE, ABEL expressed his<br />

concerns over the state <strong>of</strong> analysis (see below) and wrote:<br />

“<strong>The</strong> Taylorian <strong>The</strong>orem, the foundation for all higher mathematics is equally<br />

ill founded. Only one rigorous pro<strong>of</strong> have I found and that is by Cauchy in his<br />

Resumé des leçons sur le calcul infinitesimal. <strong>The</strong>re, he proves that<br />

φ (x + α) = φx + αφ ′ x + α2<br />

2 φ′′ x + . . .<br />

whenever the series is convergent (but it is frequently used in all cases).” 9<br />

Thus, ABEL claimed that CAUCHY had proved in the Resumés that any convergent<br />

Taylor series expansion represents the function. Certainly, CAUCHY considered the<br />

theorem <strong>of</strong> B. TAYLOR (1685–1731) <strong>of</strong> major importance in his lectures and repeated<br />

his programmatic criticism <strong>of</strong> working with divergent series. However, the statement<br />

which ABEL attributed to CAUCHY is exactly what CAUCHY criticized by ways <strong>of</strong> a<br />

counter example in the Resumé: In the Résume, CAUCHY considered the function de-<br />

fined by f (x) = e<br />

1 −<br />

x2 whose derivatives all vanished at the origin and whose Maclau-<br />

rin series therefore was the zero function. This counter example and the use which<br />

CAUCHY made <strong>of</strong> it will be discussed further in section 21.3.<br />

Thus, it appears that ABEL misread or misunderstood CAUCHY. How could a<br />

smart mathematician who was — in the words <strong>of</strong> I. GRATTAN-GUINNESS — was “more<br />

Cauchyian than Cauchy” be led to such a statement? 10 A hint may be taken from<br />

7 See the quotation p. 306.<br />

8 See section 16.2.3.<br />

9 “Det Taylorske <strong>The</strong>orem, Grundlaget for hele den høiere Mathematik er ligesaa slet begrundet. Kun<br />

eet eneste strængt Beviis har jeg fundet og det er af Cauchy i hans Resumé des leçons sur le calcul<br />

infinitesimal. Han viser der at man har:<br />

φ (x + α) = φx + αφ ′ x + α2<br />

2 φ′′ x + . . .<br />

saa <strong>of</strong>te Rækken er convergente, (men man bruger den rask væk i alle Tilfælde).” (<strong>Abel</strong>→Holmboe,<br />

1826/01/16. N. H. <strong>Abel</strong>, 1902a, 16–17).<br />

10 (I. Grattan-Guinness, 1970b, 80).


12.1. ABEL’s critical attitude 223<br />

from a manuscript entitled Sur les séries which ABEL worked on and hoped to present<br />

in A. L. CRELLE’S (1780–1855) Journal. However, the publication never materialized<br />

and ABEL’S manuscript was left in the form <strong>of</strong> an interesting draft. It was eventu-<br />

ally published in the Œuvres. 11 In the draft, ABEL expanded an otherwise unspecified<br />

function<br />

and rearranged its terms to find<br />

f (x + ω) = a0 + a1 (x + ω) + a2 (x + ω) 2 + . . . (12.1)<br />

f (x + ω) = a0 + a1x + a2x 2 + · · · + (a1 + 2a2x + . . . ) ω + . . . .<br />

From this, ABEL concluded<br />

f (x + ω) = f (x) + f ′ (x)<br />

1 ω + f ′′ (x)<br />

2 ω2 + . . . (12.2)<br />

“if this series is convergent”. 12 <strong>The</strong> argument was followed by remarks to the effect<br />

that the series <strong>of</strong> (12.2) was indeed always convergent! This serves to illustrate that de-<br />

spite the extensive criticism which ABEL raised against the unrigorous reasoning with<br />

series, his own reasoning was constantly at risk <strong>of</strong> making the same mistakes. Fur-<br />

thermore, the example shows how the reordering <strong>of</strong> terms was an unrealized problem<br />

in the 1820s. This becomes interesting when we consider the emergence <strong>of</strong> a concept<br />

<strong>of</strong> absolute convergence (see below).<br />

12.1 ABEL’s critical attitude<br />

ABEL’S name is frequently mentioned in the same sentence as CAUCHY and K. T. W.<br />

WEIERSTRASS (1815–1897) when historians <strong>of</strong> mathematics attempt to pin-point the<br />

movement within mathematics known as rigorization or — more specifically — arith-<br />

metization. 13 And certainly, after an almost religious conversion, ABEL became an ar-<br />

dent follower <strong>of</strong> a version <strong>of</strong> CAUCHY’S new rigor; a version which ABEL to a large<br />

extent helped form, himself. On the other hand, rigorizing the calculus meant re-<br />

founding the entire domain <strong>of</strong> analysis on a completely new system, and ABEL’S<br />

mathematical contribution to the rigorization was limited to a single sub-discipline,<br />

the theory <strong>of</strong> infinite series. But <strong>of</strong> equal importance, ABEL’S written testimony <strong>of</strong> his<br />

conversion to Cauchyism and his hearted, public interpretation <strong>of</strong> some <strong>of</strong> its doctrines<br />

helped shape the movement in the nineteenth century. In this and the following chap-<br />

ter, ABEL’S critical attitude as well as his contributions to the theory <strong>of</strong> series will be<br />

investigated and analyzed.<br />

11 (N. H. <strong>Abel</strong>, [1827] 1881).<br />

12 (ibid., 204).<br />

13 See e.g. (Kline, 1990, 948).


224 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

ABEL’S critical attitude was expressed both in his letters and in the opening para-<br />

graphs as well as in the overall structure <strong>of</strong> his important paper on the binomial the-<br />

orem. ABEL attacked the usual way <strong>of</strong> reasoning about infinite series which he saw<br />

as an induction from the permissible ways <strong>of</strong> reasoning about finite series, i.e. poly-<br />

nomials. This distinction between finite and infinite arguments was one <strong>of</strong> the core<br />

components <strong>of</strong> ABEL’S critical attitude.<br />

A direct, outspoken ban on divergent series. When ABEL announced the contents<br />

<strong>of</strong> his binomial paper, he explicitly pointed to the fact that convergence <strong>of</strong> a series was<br />

a requirement to be established before anything like an equality between the series<br />

and other (possibly infinite) expressions were to be asserted. He wrote, combining his<br />

distinction between finite and infinite expressions and his criticism <strong>of</strong> divergent series,<br />

“This equation [the CAUCHY product <strong>of</strong> infinite series] is completely correct<br />

when both <strong>of</strong> the series<br />

u0 + u1 + . . . and v0 + v1 + . . .<br />

are finite. If they are infinite, they must firstly necessarily converge — because a divergent<br />

series has no sum — and then the series in the second term [the CAUCHY<br />

product] must also converge. Only with these restrictions is the statement above<br />

correct. If I am not mistaken, this restriction has hitherto not been taken into<br />

account.” 14<br />

Thus, even in the case <strong>of</strong> the CAUCHY multiplication <strong>of</strong> infinite series, ABEL found<br />

reason to criticize current practice on the same grounds as stated above; different rea-<br />

soning applied to finite and infinite series, and divergent series have no sum. When<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> the multiplication theorem has been described, a further analysis <strong>of</strong><br />

its relations to CAUCHY’S original version can be described (see section 12.8, below).<br />

In some <strong>of</strong> his letters, ABEL was even more outspoken about the status <strong>of</strong> divergent<br />

series and the implications they had had on the development <strong>of</strong> rigorous mathematics.<br />

ABEL’S notion <strong>of</strong> rigorous pro<strong>of</strong> and critical revision. In another frequently quoted<br />

letter, written shortly after leaving Berlin in 1826 and directed to C. HANSTEEN (1784–<br />

1873), ABEL spoke <strong>of</strong> turning more <strong>of</strong> his attention toward the study <strong>of</strong> analysis, and<br />

again commented on the status <strong>of</strong> the field.<br />

“I will commit all my strength to shedding some light on the immense darkness,<br />

which incontestably covers analysis. It [analysis] completely lacks all plan<br />

14 “Diese Gleichung ist vollkommen richtig, wenn die beiden Reihen<br />

u0 + u1 + . . . und v0 + v1 + . . .<br />

endlich sind. Sind sie aber unendlich, so müssen sie erstlich nothwendig convergiren, weil eine<br />

divergirende Reihe keine Summe hat, und dann muß auch die Reihe im zweiten Gliede ebenfalls<br />

convergiren. Nur mit dieser Einschränkung ist der obige Ausdruck richtig. Irre ich nicht, so ist diese<br />

Einschränkung bis jetzt nicht berücksichtigt worden.” (N. H. <strong>Abel</strong>, 1826f, 311).


12.1. ABEL’s critical attitude 225<br />

and coherence, so it is highly remarkable that it can be studied by so many — and<br />

now worst <strong>of</strong> all that it is not treated rigorously. [. . . ] Whenever one proceeds in<br />

the ordinary fashion, it is probably all right; but I have had to be very cautious because<br />

the theorems which have been accepted without rigorous pro<strong>of</strong> (i.e. without<br />

pro<strong>of</strong>) have struck such deep roots with me that I constantly run the risk <strong>of</strong> using<br />

them without further probing.” 15<br />

Here, we learn <strong>of</strong> another distinction which ABEL saw between ordinary pro<strong>of</strong>s and<br />

rigorous pro<strong>of</strong>s. He was acutely aware that a fundamental change in the techniques <strong>of</strong><br />

proving mathematical theorems was required. At the same time, ABEL also expressed<br />

the opinion that it would be interesting to investigate how unrigorous reasoning had<br />

led to correct results in almost all cases. This idea <strong>of</strong> critically revising the existing<br />

structure <strong>of</strong> a mathematical theory is as old as the rigorization movement and had<br />

first been stated in the Berlin Academy prize problem for 1784. 16 ABEL also suggested<br />

an answer to the question when he emphasized that analysis until most recently had<br />

only worked with power series and for these, the established methods <strong>of</strong> reasoning<br />

did not lead to false results. A completely different situation could arise if other series<br />

were included in the study, 17 ABEL remarked thereby alluding to both Fourier series<br />

and Poisson’s example described above.<br />

ABEL and the paradoxes <strong>of</strong> analysis. On two occasions, in letters to HOLMBOE and<br />

HANSTEEN written in 1826, ABEL described some <strong>of</strong> the paradoxes to which unrig-<br />

orous reasoning had led. 18 Of the two, the letter to HOLMBOE is the most detailed.<br />

<strong>The</strong>re, ABEL ridiculed anybody willing to claim equalities such as<br />

0 = 1 − 2 n + 3 n − 4 n + . . . . (12.3)<br />

ABEL gave no references as to where he had picked up this absurd equality but we<br />

get a hint from another much more famous example which he described. In the letter<br />

and in an intriguing footnote in the binomial paper — which is discussed below, see<br />

12.6 — ABEL called attention to the series<br />

x<br />

2 =<br />

∞ (−1)<br />

∑<br />

n=1<br />

n−1 sin nx<br />

n<br />

(12.4)<br />

15 “Alle mine Kræfter vil jeg anvende paa at bringe noget mere Lys i det uhyre Mørke som der uimodsigelig<br />

nu findes i Analysen. Den mangler saa ganske al Plan og System, saaat det virkelig er<br />

høist forunderlig at den kan studeres af saa mange og nu det værste at den aldeles ikke er stræng<br />

behandlet. [. . . ] Naar man blot gaaer almindelig tilværks saa gaaer det nok; men jeg har maattet<br />

være særdeles forsigtig, thi de engang uden strængt Bevis (c : uden Bevis) antagne Sætninger har<br />

slaaet saa dybe Rødder hos mig at jeg hvert Øjeblik staaer Fare at bruge dem uden nøiere Prøvelse.”<br />

(<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. N. H. <strong>Abel</strong>, 1902a, 22–23).<br />

16 (Grabiner, 1981b, 40–43), the passage on revision in the prize proposal is translated (ibid., 41).<br />

17 (<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. N. H. <strong>Abel</strong>, 1902a, 22–23).<br />

18 (<strong>Abel</strong>→Holmboe, 1826/01/16. In ibid., 13–19) and (<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. In<br />

ibid., 22–26).


226 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

which is the Fourier series corresponding to the function f (x) = x 2 on the interval<br />

]−π, π[. However, as ABEL noted in the letter, by inserting e.g. x = π he would be led<br />

to the absurd equality<br />

π<br />

2 =<br />

∞ (−1)<br />

∑<br />

n=1<br />

n−1 sin nπ<br />

= 0.<br />

n<br />

In his letter, ABEL applied this example to illustrate that although an equality held<br />

for x < π it could fail in the limit x = π. This criticism is thus an elaboration <strong>of</strong><br />

CAUCHY’S dismissal <strong>of</strong> the generality <strong>of</strong> algebra. ABEL continued,<br />

“Operations are applied to infinite series as if they were finite but is that permissible?<br />

I doubt it. — Where has it been proved that one obtains the differential<br />

<strong>of</strong> a series by differentiating each term? It is easy to present examples for which<br />

this it not true.” 19<br />

Here, ABEL indirectly criticized J. B. J. FOURIER’S (1768–1830) interchange <strong>of</strong> the limit<br />

processes involved in term-wise integration. Differentiating the series (12.4), ABEL<br />

obtained<br />

1<br />

2 =<br />

∞<br />

∑ (−1)<br />

n=1<br />

n−1 cos nx<br />

in which the series was divergent. Now, the example (12.3) can be seen to result if this<br />

procedure <strong>of</strong> differentiation is repeated and either x = π or x = π 2<br />

is inserted.<br />

ABEL’S reaction to Poisson’s example. A strong connection between ABEL’S research<br />

on the theory <strong>of</strong> series and Poisson’s example is clearly discernible from his letter to<br />

HOLMBOE. 20 <strong>The</strong>re, ABEL explained how he had undertaken to find the sum <strong>of</strong> the<br />

series<br />

m (m − 1)<br />

cos mx + m cos (m − 2) x + cos (m − 4) x + . . .<br />

2<br />

which was an important open problem at the time. ABEL mentioned that a large num-<br />

ber <strong>of</strong> mathematicians (including S.-D. POISSON (1781–1840) and CRELLE) had failed<br />

to solve the problem but that he, himself, had found a complete answer in the form<br />

m (m − 1)<br />

cos mx + m cos (m − 2) x + cos (m − 4) x + · · · = (2 + 2 cos 2x)<br />

2<br />

m 2 cos mkπ<br />

�<br />

in which m > −1, k an integer and k − 1 � �<br />

2 π < x < k + 1 �<br />

2 π; for m < −1, the series<br />

was divergent and this led to ABEL’S outburst against the use <strong>of</strong> divergent series:<br />

“Divergent series are the creations <strong>of</strong> the Devil and it is a shame that anybody<br />

dare construct a demonstration upon them.” 21<br />

19 “Man anvender alle Operationer paa uendelige Rækker som om de vare endelige, men er dette<br />

tilladt? Vel neppe. — Hvor staar det beviist at man faaer Differentialet af en uendelig Række ved at<br />

differentiere hvert Led? Det er let at anføre Exempler hvor dette ikke er rigtigt.” (<strong>Abel</strong>→Holmboe,<br />

1826/01/16. N. H. <strong>Abel</strong>, 1902a, 18).<br />

20 (<strong>Abel</strong>→Holmboe, 1826/01/16. In ibid., 13–19).


12.1. ABEL’s critical attitude 227<br />

Thus, ABEL’S criticism <strong>of</strong> divergent series was intimately tied to his interest in Poisson’s<br />

example and the other paradoxes in the theory <strong>of</strong> infinite series.<br />

<strong>The</strong> core components <strong>of</strong> ABEL’S criticism. ABEL’S critical position toward the con-<br />

temporary conceptions <strong>of</strong> rigor in analysis can thus be divided into three parts. <strong>The</strong><br />

first part was primarily rhetorical emphasizing his belief that the accepted standards<br />

were utterly insufficient and ABEL supported his argument by enlisting a number <strong>of</strong><br />

“paradoxes”, in particular (12.3).<br />

<strong>The</strong> second — and more substantial part — consisted <strong>of</strong> an attempt at locating the<br />

points where the customary reasoning was led astray. Among the critical points, ABEL<br />

repeated and radicalized CAUCHY’S dogma that divergent series have no sum and<br />

should not be treated in analysis. Beside his ban on divergent series, ABEL also re-<br />

peated CAUCHY’S concern for numerical equality and stressed that even if two expres-<br />

sions were numerically equal in the interior <strong>of</strong> an interval they needed not coincide in<br />

the endpoints. This led him to explicitly question specific practices such as the passing<br />

to the limit in power series and the term-wise differentiation (and integration).<br />

<strong>The</strong> third component <strong>of</strong> ABEL’S critical position was more constructive. In his<br />

letters and publications, he suggested that most <strong>of</strong> the unfortunate paradoxes and<br />

malpractices arose out <strong>of</strong> considering series which were not power series. This led<br />

him to focus attention on power series which he saw as some sort <strong>of</strong> safe haven where<br />

the commonly used methods would still apply.<br />

<strong>The</strong> structure <strong>of</strong> ABEL’S pro<strong>of</strong> <strong>of</strong> the binomial theorem. <strong>The</strong> path which ABEL took<br />

in his publication on the binomial theorem makes a lot <strong>of</strong> sense when seen from a per-<br />

spective integrating Poisson’s example and the components <strong>of</strong> ABEL’S criticism listed<br />

above. In the binomial paper, ABEL separated two distinct problems concerning the<br />

binomial theorem. First, he wanted to find the precise set <strong>of</strong> assumptions on m and x<br />

for which the binomial series<br />

∞<br />

∑<br />

n=0<br />

� �<br />

m<br />

x<br />

n<br />

n<br />

converged. In order to do so, he developed and revised some important theorems<br />

in the theory <strong>of</strong> series. And second, he wanted to investigate whether — in the cases<br />

where the series converged — the sum <strong>of</strong> the series agreed with (one <strong>of</strong> the values <strong>of</strong>)<br />

the binomial<br />

(1 + x) m .<br />

This approach was adapted to overcome the problems <strong>of</strong> multi-valued functions which<br />

lie at the core <strong>of</strong> the Poisson’s example.<br />

21 “Divergente Rækker ere i det Hele noget Fandenskab, og det er en Skam at man vover at grunde<br />

nogen Demonstration derpaa.” (<strong>Abel</strong>→Holmboe, 1826/01/16. ibid., 16).


228 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

12.2 Infinitesimals<br />

ABEL’S reading and rendering <strong>of</strong> CAUCHY’S fundamental concepts came to influence<br />

the development <strong>of</strong> analysis in the nineteenth century. At certain points, his interpre-<br />

tations were clearer and more specific and some <strong>of</strong> them would eventually coincide<br />

with the standardized interpretations laid down by men such as G. P. L. DIRICHLET<br />

(1805–1859) and WEIERSTRASS.<br />

One <strong>of</strong> the basic notions which ABEL introduced in his binomial paper was that <strong>of</strong><br />

infinitesimals. In a footnote, ABEL explained,<br />

“For brevity, in this paper ω denotes a quantity which can be less than any<br />

given arbitrarily small quantity.” 22<br />

Despite the awe which ABEL felt for CAUCHY’S work, this definition is not truly in<br />

accord with CAUCHY’S notion <strong>of</strong> infinitesimals. As discussed above, CAUCHY had<br />

interpreted infinitesimals as variables with limit zero but in ABEL’S definition, the<br />

infinitesimals seem to reenter as completed quantities less than any finite quantity<br />

but different from zero. <strong>The</strong> limit process has seemingly faded into the background.<br />

To illustrate this way <strong>of</strong> designating infinitesimals by symbols, we may reconsider<br />

CAUCHY’S pro<strong>of</strong> <strong>of</strong> the Cauchy <strong>The</strong>orem (see page 217) interpreted in ABEL’S notation.<br />

ABEL did not undertake this pro<strong>of</strong>, but the arguments are directly to those which he<br />

employed in proving the Lehrsatz IV (see below). By continuity <strong>of</strong> the finite polyno-<br />

mial sn,<br />

and by the convergence <strong>of</strong> s,<br />

<strong>The</strong>refore<br />

sn (x + α) − sn (x) = ω,<br />

rn (x + α) = ω, (12.5)<br />

rn (x) = ω.<br />

s (x + α) − s (x) = ω<br />

and the continuity has been “proved”. This way <strong>of</strong> designating infinitesimals by the<br />

same symbols regardless <strong>of</strong> the way in which they depend on other variables and<br />

infinitesimals hid and obscured the basic problems <strong>of</strong> the above argument. In the ar-<br />

gument, the n which appears in (12.5) must depend on α and ω and can be unbounded<br />

as α and ω vanish. However, it took quite some time and a detailed analysis <strong>of</strong> these<br />

dependencies to clear out the pro<strong>of</strong> (see section 14.1.2, below).<br />

22 “Die Kürze wegen soll in dieser Abhandlung unter ω eine Größe verstanden werden, die kleiner<br />

sein kann, als jede gegebene, noch so kleine Größe.” (N. H. <strong>Abel</strong>, 1826f, 313, footnote).


12.3. Convergence 229<br />

12.3 Convergence<br />

While ABEL’S definition and use <strong>of</strong> infinitesimals were not completely in the line <strong>of</strong><br />

CAUCHY’S new rigor, his concept <strong>of</strong> convergence — and the importance which he at-<br />

tributed to it — closely resembled CAUCHY’S.<br />

“Definition. An arbitrary<br />

v0 + v1 + v2 + · · · + vm etc.<br />

will be called convergent if the sum v0 + v1 + · · · + vm steadily approaches a certain<br />

limit for ever increasing values <strong>of</strong> m. This limit will be called the sum <strong>of</strong> the<br />

series. In the contrary case, the series is called divergent and therefore has no sum.<br />

From this definition follows that for a series to converge, it will be necessary and<br />

sufficient that the sum vm + vm+1 + · · · + vm+n steadily approach zero for ever<br />

increasing values <strong>of</strong> m whatever value n may have.” 23<br />

Just as CAUCHY had done, ABEL quickly related the convergence <strong>of</strong> a series to the<br />

Cauchy criterion and claimed that it constituted a necessary and sufficient condition<br />

for convergence. As described above, the assertion that convergence followed from<br />

the Cauchy criterion was later realized to be non-trivial, but in the 1820s it was con-<br />

sidered obvious. Although both CAUCHY and ABEL drew the connection between<br />

convergence and the Cauchy criterion, ABEL gave the criterion a much more central<br />

position in his theory <strong>of</strong> series as will be described below (see page 231).<br />

Immediately following his definition <strong>of</strong> convergence, ABEL made the rather curious<br />

remark that “in every arbitrary series, the general term vm will approach zero.” 24<br />

Judging from the context, an omission <strong>of</strong> the word “convergent” must have crept in at<br />

this point. 25<br />

An extended ratio test: Lehrsätze I&II. <strong>The</strong> first theorem <strong>of</strong> ABEL’S binomial paper<br />

is most remarkable because <strong>of</strong> its conceptual contents. Without pro<strong>of</strong> (see a modern-<br />

ized pro<strong>of</strong> in box 2), ABEL observed that for any series <strong>of</strong> positive terms<br />

23 “Erklärung. Eine beliebige Reihe<br />

∞<br />

∑ ρm<br />

m=0<br />

v0 + v1 + v2 + · · · + vm u.s.w.<br />

soll convergent heißen, wenn, für stets wachsende Werthe von m, die Summe v0 + v1 + · · · + vm sich<br />

immerfort eine gewisse Gränze nähert. Diese Grenze soll Summe der Reihe heißen. Im entgegengesetzten<br />

Falle soll die Reihe divergent heißen, und hat alsdann keine Summe. Aus dieser Erklärung<br />

folgt, daß, wenn eine Reihe convergiren soll, es nothwendig und hinreichend sein wird, daß, für<br />

stets wachsende Werthe von m, die Summe vm + vm+1 + · · · + vm+n sich Null immerfort nähert,<br />

welchen Werth auch n haben mag.” (ibid., 313).<br />

24 “In irgend einer beliebigen Reihe wird also das allgemeine Glied vm sich Null stets nähern.” (ibid.,<br />

313).<br />

25 This omission has therefore also been silently corrected in the French translation found in (N. H.<br />

<strong>Abel</strong>, 1839; N. H. <strong>Abel</strong>, 1881).


230 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

Pro<strong>of</strong> <strong>of</strong> Lehrsatz I In order to present a pro<strong>of</strong> <strong>of</strong> Lehrsatz I, we write ε = α−1<br />

2<br />

and chose n ∈ N such that<br />

<strong>The</strong>n, by iteration,<br />

ρm+1 ≥ (α − ε) ρm for all m ≥ n.<br />

ρm+1 ≥ (α − ε) m−n ρn.<br />

<strong>The</strong> choice <strong>of</strong> ε ensures that ρm+1 increases beyond all bounds. Now assume that the<br />

series in (12.6), ∑ εmρm, is to be convergent. <strong>The</strong>n, in particular, a k ∈ N exists such<br />

that<br />

If m ≥ max (k, n),<br />

meaning<br />

|εm+1ρm+1| < ε for all m ≥ k.<br />

ε > |εm+1| · ρm+1 ≥ |εm+1| (α − ε) m−n ρn<br />

|εm+1| ≤ ε<br />

ρn<br />

1<br />

m−n → 0 for m → ∞<br />

(α − ε)<br />

because α − ε > 1. In conclusion, if ∑ εmρm is be convergent, the sequence {εm} has<br />

to converge toward zero. And since this is not the case, the sum (12.6) cannot be<br />

convergent. ✷<br />

Box 2: Pro<strong>of</strong> <strong>of</strong> Lehrsatz I<br />

for which the ratio <strong>of</strong> consecutive terms converges toward α > 1,<br />

any linear combination<br />

ρm+1<br />

ρm<br />

→ α > 1,<br />

∞<br />

∑ εmρm<br />

m=0<br />

will be divergent, provided the sequence {εm} does not converge to zero.<br />

> 0<br />

(12.6)<br />

<strong>The</strong> contents <strong>of</strong> this Lehrsatz I is thus a generalization <strong>of</strong> one part <strong>of</strong> CAUCHY’S ra-<br />

tio test <strong>of</strong> convergence. It is remarkable from a conceptual viewpoint that ABEL’S first<br />

theorem would be one <strong>of</strong> divergence when his entire theory was so focused on conver-<br />

gent series. <strong>The</strong> Lehrsatz I is thus — as it stands — a negative demarcation criterion.<br />

This apparent imbalance was leveled by the second theorem. ABEL’S Lehrsatz II<br />

is a counterpart to the Lehrsatz I describing analogous — but this time sufficient —<br />

conditions for convergence. ABEL found that if<br />

ρm+1<br />

ρm<br />

→ α < 1


12.3. Convergence 231<br />

and {εm} was a sequence <strong>of</strong> terms which do not exceed 1 (ABEL was not explicit about<br />

requiring εm to be positive; if this assumption is not made, the requirement would be<br />

|εm| < 1), the series<br />

∞<br />

∑ εmρm<br />

m=0<br />

was necessarily convergent. Throughout, ABEL’S use <strong>of</strong> numerical values was spo-<br />

radic. At times, he noticed that numerical values had to be taken, at other times such<br />

a remark might be inferred from his language, and at yet other times — as will be<br />

shown with respect to Lehrsatz VI — it seems to have evaded his attention. For in-<br />

stance, in the example given above, ABEL actually required that εm did not surpass<br />

unity which could also mean −1 < εm < 1.<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> Lehrsatz II had similarities with the modernized pro<strong>of</strong> <strong>of</strong> Lehrsatz<br />

I given in box 2 although he chose to use the characterization <strong>of</strong> convergence by the<br />

Cauchy criterion given above. He argued that for m sufficiently large,<br />

and consequently<br />

m+n<br />

∑ ρk < ρm<br />

k=m<br />

ρ m+k < α k ρm<br />

n<br />

∑ α<br />

k=0<br />

k 1 − α<br />

= ρm<br />

n+1<br />

1 − α<br />

< ρm<br />

1 − α .<br />

Since εm < 1, a similar conclusion would hold for the series εmρm,<br />

m+n<br />

∑ εkρk <<br />

k=m<br />

ρm<br />

1 − α .<br />

(12.7)<br />

ABEL concluded the argument by observing that ρm → 0 followed from (12.7), and<br />

thus the convergence <strong>of</strong> the series was secured by the characterization.<br />

Although ABEL’S way to the two theorems might seem obvious to a modern reader,<br />

it is interesting to compare ABEL’S theorems including their pro<strong>of</strong>s with CAUCHY’S<br />

original deduction <strong>of</strong> the ratio test as given in the Cours d’analyse (see page 212, above).<br />

Such a comparison reveals that CAUCHY and ABEL devised different structural sys-<br />

tems for their theories <strong>of</strong> infinite series. Although the results were the same, the the-<br />

orems and pro<strong>of</strong>s played slightly different roles in the two systems. In CAUCHY’S<br />

theory, the characterization <strong>of</strong> convergent series by means <strong>of</strong> the Cauchy criterion was<br />

noticed but was <strong>of</strong> little use in obtaining the other — more important — tests <strong>of</strong> con-<br />

vergence. Instead, these tests were derived from the convergence <strong>of</strong> geometric pro-<br />

gressions which was proved directly. In ABEL’S theory, however, the Cauchy criterion<br />

was made into a central tool which he used to deduce his slightly generalized version<br />

<strong>of</strong> the ratio test (see figure 12.1). This made the pro<strong>of</strong> <strong>of</strong> the ratio test much simpler in<br />

ABEL’S framework than it had been in CAUCHY’S Cours d’analyse. In the manuscript<br />

Sur les séries, ABEL placed the Cauchy criterion in an equally central position.


232 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

CAUCHY’s structure ABEL’s structure<br />

Definition <strong>of</strong> convergence<br />

✟<br />

✟❍<br />

✟✟✙<br />

❍❍❍❥<br />

Cauchy Geometric<br />

criterion progression<br />

❄<br />

Root test<br />

❄<br />

Ratio test<br />

Definition <strong>of</strong> convergence<br />

❄<br />

Cauchy criterion<br />

❄<br />

Ratio test<br />

Figure 12.1: Comparison <strong>of</strong> CAUCHY’s and ABEL’s structures <strong>of</strong> the basic theory <strong>of</strong><br />

infinite series.<br />

An auxiliary theorem: Lehrsatz III. As his third theorem, 26 ABEL presented an aux-<br />

iliary result which — although not difficult — was put to great use in the pro<strong>of</strong>s to<br />

follow. He demonstrated that if {tn} denoted a sequence whose partial sums were<br />

bounded,<br />

m<br />

∑ tk < δ for all m ∈ N,<br />

k=0<br />

and {εn} denoted a decreasing sequence <strong>of</strong> positve terms, then<br />

rm =<br />

m<br />

∑ εktk < δε0 for all m ∈ N.<br />

k=0<br />

ABEL’S pro<strong>of</strong> consisted <strong>of</strong> a rather simple manipulation, in which he observed that<br />

with<br />

each term could be written as<br />

and, thus,<br />

rm =<br />

=<br />

m<br />

∑ εktk =<br />

k=0<br />

Since {εn} was decreasing,<br />

26 (N. H. <strong>Abel</strong>, 1826f, 314)<br />

pm =<br />

m<br />

∑ tk, k=0<br />

t k = p k − p k−1,<br />

m<br />

∑ εk (pk − pk−1) =<br />

k=0<br />

m−1<br />

∑ pk (εk − εk+1) + εmpm.<br />

k=0<br />

0 < ε k − ε k+1 < ε k < ε0,<br />

m<br />

∑ εkp k −<br />

k=0<br />

m−1<br />

∑ εk+1p k<br />

k=0


12.4. Continuity 233<br />

it followed that<br />

rm <<br />

and the inequality had been obtained.<br />

12.4 Continuity<br />

m−1<br />

∑ δ (εk − εk+1) + εmδ = δε0,<br />

k=0<br />

Just as had been the case with CAUCHY’S pro<strong>of</strong> <strong>of</strong> the binomial theorem, the concept<br />

<strong>of</strong> continuity played an important role in ABEL’S pro<strong>of</strong>. In his paper on the binomial<br />

theorem, ABEL gave rudiments <strong>of</strong> a different rendering <strong>of</strong> the theory <strong>of</strong> interaction<br />

between the concepts <strong>of</strong> continuity and convergence. ABEL’S definition <strong>of</strong> continuity<br />

seems to closely resemble CAUCHY’S (see page 216), although it may be noticed that<br />

ABEL’S definition is only formulated in the terminology <strong>of</strong> limits.<br />

“Definition. A function f (x) shall be called a continuous function <strong>of</strong> x between<br />

the boundaries x = 0 and x = b when for any arbitrary value <strong>of</strong> x between these<br />

limits, the quantity f (x − β) for ever decreasing values <strong>of</strong> β approach the limit<br />

f (x).” 27<br />

Although their definitions <strong>of</strong> continuity are almost identical, ABEL and CAUCHY<br />

attributed slightly different meaning to their concepts when they were employed. <strong>The</strong><br />

apparent ambiguity concerning the order in which quantification is to be made in<br />

CAUCHY’S definition was resolved in ABEL’S persistent insistence on point-wise def-<br />

initions. ABEL’S definition as stated seems just as susceptible to the ambiguity as<br />

CAUCHY’S, but ABEL throughout interpreted it to mean that a function is continuous<br />

at a point x ∈ [0, b] if f (x − β) → f (x) as β → 0.<br />

Combining continuity, convergence, and power series. <strong>The</strong> fourth and fifth theo-<br />

rems <strong>of</strong> ABEL’S binomial paper provided two important combinations <strong>of</strong> the concepts<br />

<strong>of</strong> continuity and convergence. <strong>The</strong> fourth theorem, Lehrsatz IV, stated and proved the<br />

continuity <strong>of</strong> a power series in the interior <strong>of</strong> its interval <strong>of</strong> convergence, while Lehrsatz<br />

V attempted to provide a rigorous replacement for what Cauchy’s <strong>The</strong>orem (see page<br />

217) had promised but not rigorously delivered. At this point, Lehrsatz IV together<br />

with its pro<strong>of</strong> will be described first, and Lehrsatz V will be postponed to be discussed<br />

in its proper context <strong>of</strong> ABEL’S famous Ausnahme or counter example to CAUCHY’S<br />

theorem (see section 12.6). At that point, the strong internal relations between the two<br />

theorems will also be described and explained.<br />

As his fourth theorem, ABEL stated and proved the following result which has<br />

become a classic <strong>of</strong> the theory <strong>of</strong> series and is <strong>of</strong>ten associated with ABEL’S name.<br />

27 “Erklärung. Eine Function f (x) soll stetige Function von x, zwischen den Grenzen x = 0, x = b<br />

heißen, wenn für einen beliebigen Werth von x, zwischen diesen Grenzen, die Größe f (x − β) sich<br />

für stets abnehmende Werthe von β, der Grenze f (x) nähert.” (ibid., 314).


234 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

“Lehrsatz IV. When the series<br />

f (α) = v0 + v1α + v2α 2 + · · · + vmα m + . . .<br />

converges for a certain value δ <strong>of</strong> α, it will also converge for every smaller value <strong>of</strong><br />

α. Furthermore it will be <strong>of</strong> the sort that f (α − β) approaches the limit f (α) for<br />

ever decreasing values <strong>of</strong> β provided α is less than or equal to δ.” 28<br />

In most modern presentations, the variable α is interpreted as a complex variable,<br />

and the theorem states the continuity <strong>of</strong> a power series in the interior <strong>of</strong> its disc <strong>of</strong><br />

convergence. However, in this part <strong>of</strong> the paper, ABEL was exclusively interested in<br />

power series with real terms.<br />

In order to facilitate comparison with ABEL’S pro<strong>of</strong> <strong>of</strong> the fifth theorem <strong>of</strong> the paper<br />

and in order to exemplify ABEL’S use <strong>of</strong> infinitesimals in his arguments, a presentation<br />

<strong>of</strong> ABEL’S pro<strong>of</strong> is worth giving. It should be remarked even before embarking on<br />

a tour <strong>of</strong> ABEL’S pro<strong>of</strong>, that at a number <strong>of</strong> points its argument diverges from the<br />

modernized version <strong>of</strong> the pro<strong>of</strong>; these occasions will be noticed below and elaborated<br />

in the following section.<br />

ABEL began his pro<strong>of</strong> by splitting the power series after m terms,<br />

φ (α) =<br />

<strong>The</strong>n, he rewrote the tail <strong>of</strong> the series as<br />

m−1<br />

∑ vnα<br />

n=0<br />

n ∞<br />

and ψ (α) = ∑ vnα<br />

n=m<br />

n .<br />

ψ (α) =<br />

∞<br />

∑<br />

n=m<br />

and obtained from Lehrsatz III the inequality<br />

ψ (α) <<br />

�<br />

α<br />

�n vnδ<br />

δ<br />

n<br />

�<br />

α<br />

�m · p (12.8)<br />

δ<br />

“where p denotes the largest number among the quantities vmδ m , vmδ m + vm+1δ m+1 ,<br />

vmδ m + vm+1δ m+1 + vm+2δ m+2 etc.” 29 This definition <strong>of</strong> p might seem strange or dan-<br />

gerous to the modern reader. When translated into modern notation, ABEL’S p corre-<br />

sponds to the following supremum<br />

28 “Lehrsatz IV. Wenn die Reihe<br />

p = sup<br />

k<br />

m+k<br />

∑<br />

n=m<br />

f (α) = v0 + v1α + v2α 2 + · · · + vmα m + . . .<br />

vnδ n . (12.9)<br />

für einen gewissen Werth δ von α convergirt, so wird sie auch für jeden kleineren Werth von α<br />

convergiren, und von der Art seyn, daß f (α − β), für stets abnehmende Werthe von β, sich der<br />

Grenze f (α) nähert, vorausgesetzt, daß α gleich oder kleiner ist als δ.” (N. H. <strong>Abel</strong>, 1826f, 314).<br />

29 “wenn p die größte der Größen vmδ m , vmδ m + vm+1δ m+1 , vmδ m + vm+1δ m+1 + vm+2δ m+2 u.s.w. bezeichnet.”<br />

(ibid., 315).


12.4. Continuity 235<br />

Actually, if taken literally, ABEL’S p would be the maximum <strong>of</strong> the sums in (12.9) — not<br />

the supremum — but this distinction is beyond the point because for ABEL, the entire<br />

discussion on the definition <strong>of</strong> p was a non-issue. <strong>The</strong> quantity p was simply (and<br />

un-problematically) defined to be the greatest one among a sequence <strong>of</strong> numbers. <strong>The</strong><br />

nature <strong>of</strong> the number p will be pursued in section 12.6, where Lehrsatz V will shed even<br />

more light on this non-issue. In the present case, the sequence <strong>of</strong> numbers is bounded<br />

since the series is assumed to converge at δ, but this was not explicitly remarked by<br />

ABEL.<br />

In order to follow ideas <strong>of</strong> ABEL’S pro<strong>of</strong>, it suffices to take either ABEL’S naïve<br />

definition <strong>of</strong> p or the modernized one expressed in (12.9). From the equality (12.8),<br />

ABEL then concluded, that<br />

“for any value <strong>of</strong> α which is less than or equal to δ, m can be taken sufficiently<br />

large that<br />

ψ (α) = ω.” 30<br />

Next, ABEL observed that φ (α) was an entire function <strong>of</strong> α, i.e. a polynomial, and<br />

thus β could be taken small enough that<br />

φ (α) − φ (α − β) = ω.<br />

By combining these two results, ABEL concluded<br />

f (α) − f (α − β) = ω.<br />

Here we encounter ABEL’S way <strong>of</strong> operating with infinitesimals. <strong>The</strong> internal de-<br />

pendencies among ω, m, and α have been completely obscured by the notation and<br />

the argumentative style.<br />

A simple observation, inspired by comparing ABEL’S pro<strong>of</strong> with modern exposi-<br />

tions <strong>of</strong> the calculus, concerns the use <strong>of</strong> infinitesimals. Today, infinitesimals have been<br />

completely abandoned from “rigorous” presentations <strong>of</strong> the calculus, and to a person<br />

trained within this program, ABEL’S usage <strong>of</strong> infinitesimals and even CAUCHY’S dual<br />

definitions involving both limits and infinitesimals can be repulsive. But to ABEL they<br />

were legitimate means <strong>of</strong> proving theorems.<br />

ABEL’S Lehrsatz IV and the paradoxes <strong>of</strong> analysis. In the binomial paper, the fourth<br />

theorem is given without further comments, but in his letters, ABEL had related it to<br />

one <strong>of</strong> the strongest ongoing discussions among analysts. As already described in<br />

30 “Mithin kann man für jeden Werth von α, der gleich oder kleiner ist, als δ, m groß genug annehmen,<br />

daß<br />

ist.” (ibid., 315).<br />

ψ (α) = ω


236 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

section 12.1, ABEL observed that a common practice for evaluating infinite sums, say<br />

∑ an had been to transform the series into a power series ∑ anx n , obtain an expression<br />

for the sum and then insert x = 1. Commenting on this practice, ABEL wrote,<br />

“This is probably right, but it appears to me that one cannot assume it without<br />

pro<strong>of</strong>; just because<br />

φ (x) = a0 + a1x + a2x 2 + . . .<br />

for all values <strong>of</strong> x less than 1, it is not thereby said that the same conclusion holds<br />

for x = 1.” 31<br />

In order to illustrate his point <strong>of</strong> criticism, ABEL remarked that his claim was cer-<br />

tainly to the point if the power series failed to converge for x = 1 in which case it had<br />

no sum. However, when he gave explicit examples, he took them from the emerging<br />

theory <strong>of</strong> trigonometric series and not from within the realm <strong>of</strong> power series. And<br />

there is very good reason why he did not give a power series as a counter example;<br />

his fourth theorem states that for power series, the procedure <strong>of</strong> passing to the limit<br />

(inserting x = 1) can only fail in case the series is divergent for x = 1. Thus, Lehrsatz IV<br />

is the assurance needed to justify this procedure for the class <strong>of</strong> power series provided<br />

the resulting series is assumed to converge.<br />

DIRICHLET’S modification <strong>of</strong> ABEL’S pro<strong>of</strong>. In 1862, J. LIOUVILLE (1809–1882) re-<br />

ported having discussed ABEL’S fourth theorem with his friend DIRICHLET, who had<br />

died just a few years before. LIOUVILLE had expressed his concern about the orig-<br />

inal pro<strong>of</strong> <strong>of</strong> ABEL’S very important theorem which he found difficult to present in<br />

courses and even to understand. On the spot, DIRICHLET gave an alternative pro<strong>of</strong> <strong>of</strong><br />

Lehrsatz IV, which LIOUVILLE felt would remove all such difficulties. It was this new<br />

pro<strong>of</strong> by DIRICHLET which LIOUVILLE reproduced in verbatim in a short note in his<br />

Journal de mathématiques pures et appliquées. 32 Similarly, a page in G. F. B. RIEMANN’S<br />

(1826–1866) Nachlass contains his reworking <strong>of</strong> ABEL’S pro<strong>of</strong> which also supports the<br />

impression that ABEL’S original pro<strong>of</strong> was not universally accepted. 33<br />

Before the differences between the ABEL’S and DIRICHLET’S pro<strong>of</strong>s are discussed<br />

and analyzed, a presentation <strong>of</strong> DIRICHLET’S new pro<strong>of</strong> is required. A modern recon-<br />

struction <strong>of</strong> DIRICHLET’S pro<strong>of</strong> is given in box 3.<br />

For the infinite series<br />

A =<br />

∞<br />

∑ am,<br />

m=0<br />

31 “Dette er vel rigtigt; men mig synes at man ikke kan antage det uden Beviis, thi fordi man beviser<br />

at<br />

φ (x) = a0 + a1x + a2x 2 + . . .<br />

for alle Værdier af x som er mindre end 1, saa er det ikke derfor sagt at det samme finder Sted for<br />

x = 1.” (<strong>Abel</strong>→Holmboe, 1826/01/16. N. H. <strong>Abel</strong>, 1902a, 17).<br />

32 (G. L. Dirichlet, 1862). <strong>The</strong> pro<strong>of</strong> is also described in (I. Grattan-Guinness, 1970b, 108).<br />

33 (Laugwitz, 1999, 207).


12.4. Continuity 237<br />

DIRICHLET introduced the partial sums<br />

sn =<br />

n<br />

∑ am.<br />

m=0<br />

He assumed that the numerical values <strong>of</strong> the partial sums were always bounded by a<br />

constant k and that they converge toward the limit A as n grows beyond all bounds.<br />

<strong>The</strong>n DIRICHLET introduced the associated power series<br />

S =<br />

∞<br />

∑ amρ<br />

m=0<br />

m ,<br />

where 0 < ρ < 1 and thus wanted to prove that S (ρ) → A when ρ → 1. He observed<br />

that since am = sm+1 − sm, it could be rewritten as<br />

into<br />

S = s0 +<br />

∞<br />

∑ (sm − sm−1) ρ<br />

m=1<br />

m . (12.10)<br />

DIRICHLET subsequently transformed this expression for the power series (12.10)<br />

S = (1 − ρ)<br />

∞<br />

∑<br />

m=0<br />

smρ m<br />

by the finite argument, i.e. by considering only the first n + 1 terms <strong>of</strong> (12.10)<br />

Sn+1 = s0 +<br />

n<br />

∑ (sm − sm−1) ρ<br />

m=1<br />

m n−1<br />

= (1 − ρ) ∑ smρ<br />

m=0<br />

m + snρ n<br />

(12.11)<br />

and the observation that snρ n “vanishes for n = ∞”, i.e. snρ n → 0 as n → ∞. Since the<br />

two expressions correspond for any finite n, their limits also correspond,<br />

S = lim<br />

n→∞ Sn+1 = (1 − ρ) lim<br />

∞<br />

sn = (1 − ρ) ∑<br />

n→∞<br />

m=0<br />

smρ m .<br />

Now, DIRICHLET wanted to prove that the expression (12.11) converged to A as<br />

ε = 1 − ρ converged to zero. To do so, he split the series (12.11) into two parts<br />

S = (1 − ρ)<br />

n−1<br />

∑ smρ<br />

m=0<br />

m ∞<br />

+ (1 − ρ) ∑<br />

m=n<br />

smρ m , (12.12)<br />

and observed that the first sum was bounded by εnk, since |smρ m | < |sm| < k. Thus,<br />

he claimed that it converged to zero with ε → 0, which is true, provided n is kept<br />

fixed. As for the second sum, DIRICHLET claimed it could be written as<br />

P (1 − ρ)<br />

∞<br />

∑<br />

m=n<br />

ρ m = Pρ n = P (1 − ε) n , (12.13)<br />

provided P be chosen as a number between the smallest and the largest among the<br />

quantities sn, sn+1, . . . . <strong>The</strong>re is no explicit explanation for this claim, but it could be<br />

obtained in a number <strong>of</strong> ways, either from CAUCHY’S theory <strong>of</strong> means or as an easy<br />

consequence <strong>of</strong> the intermediate value theorem <strong>of</strong> the integral calculus. Because the<br />

partial sums sn, sn+1, . . . all converge to A, DIRICHLET had proved his claim that S<br />

converges to A when ρ converges to 1.


238 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

Comparison <strong>of</strong> ABEL’S and DIRICHLET’S pro<strong>of</strong>s. It is interesting to note how the<br />

attitude toward infinitesimals and limit arguments evolved over the first 40 years after<br />

CAUCHY’S Cours d’analyse and we can get an indication <strong>of</strong> this by comparing ABEL’S<br />

and DIRICHLET’S pro<strong>of</strong>s. Contrary to ABEL’S pro<strong>of</strong>, DIRICHLET completely avoided<br />

the use <strong>of</strong> infinitesimals in his pro<strong>of</strong>. Instead, he argued completely within the process<br />

based interpretation <strong>of</strong> limits when he reduced the series to finitely many terms, ma-<br />

nipulated the polynomials and applied the limit process n → ∞. However, DIRICH-<br />

LET’S notation still hid the order in which limit processes are to be sequenced.<br />

In the ultimate step <strong>of</strong> DIRICHLET’S pro<strong>of</strong>, two limit processes were involved; both<br />

expressions (12.12) and (12.13) involved both ε and n which were intended to converge<br />

toward zero and infinity, respectively. A modern reconstruction <strong>of</strong> the limit processes<br />

<strong>of</strong> DIRICHLET’S argument could proceed along the lines suggested by the pro<strong>of</strong> in box<br />

3. In the box, it is illustrated how the limit processes can be straightened by first fixing<br />

a value <strong>of</strong> n such that |sm − A| is sufficiently small for all m ≥ n and then specifying<br />

the ε that will make the power series differ from A by as little as had been required.<br />

<strong>The</strong> order <strong>of</strong> DIRICHLET’S pro<strong>of</strong> does not reflect the order in which the limit processes<br />

are to be carried out, and neither does his notation. <strong>The</strong>refore, we are still faced with a<br />

line <strong>of</strong> argument in which limit processes are not as clearly identified and sequentially<br />

ordered as it is required today.<br />

12.5 ABEL’s “exception”<br />

<strong>The</strong> “exception” in the binomial paper. In pro<strong>of</strong>s <strong>of</strong> the binomial theorem which<br />

follow EULER’S method <strong>of</strong> extending the binomial formula through the use <strong>of</strong> the<br />

functional equation, some argument based on continuity has to be applied to get from<br />

rational to real exponents. To meet this demand in his pro<strong>of</strong>, CAUCHY deduced and<br />

stated the so-called Cauchy’s <strong>The</strong>orem (see section 11.5). At the corresponding point <strong>of</strong><br />

his binomial paper, ABEL discarded CAUCHY’S version <strong>of</strong> the theorem because he had<br />

discovered that it “suffered exceptions”: 34<br />

“Remark. In the above-mentioned work <strong>of</strong> Mr. Cauchy (on page 131) the following<br />

theorem can be found:<br />

»Whenever the different terms <strong>of</strong> the series<br />

u0 + u1 + u2 + u3 + . . . etc.<br />

are functions <strong>of</strong> one and the same variable quantity and moreover continuous<br />

functions with regard to this variable in the vicinity <strong>of</strong> a particular value for which<br />

the series is convergent, then the sum s <strong>of</strong> the series will also be a continuous<br />

function <strong>of</strong> x in the vicinity <strong>of</strong> that particular value.«<br />

34 For CAUCHY’S original formulation, which is authentically translated in ABEL’S paper, see page<br />

217.


12.5. ABEL’s “exception” 239<br />

A modern reconstruction <strong>of</strong> DIRICHLET’s pro<strong>of</strong>. Let δ > 0 be given and choose (by<br />

convergence) n such that<br />

<strong>The</strong>n chose ε1 > 0 and ε2 > 0 such that<br />

|sm − A| < δ<br />

for all m ≥ n.<br />

6<br />

nεk < δ<br />

3 for ε < ε1, and<br />

|ρ n − 1| < δ<br />

for all ε = 1 − ρ < ε2.<br />

3k<br />

<strong>The</strong>n, if ε < min {ε1, ε2}, the equation (12.12) becomes<br />

n−1<br />

|S − A| ≤ (1 − ρ) ∑ |sm| ρ m �<br />

�<br />

�<br />

+ �(1<br />

− ρ)<br />

�<br />

m=0<br />

∞<br />

∑<br />

m=n<br />

In the first sum, the interesting reconstructed inequality is<br />

(1 − ρ)<br />

n−1<br />

∑<br />

m=0<br />

When the second sum is rewritten as<br />

(1 − ρ)<br />

|sm| ρ m ≤ ε<br />

n−1<br />

∑<br />

m=0<br />

smρ m − A<br />

|sm| ≤ εnk < δ<br />

3 .<br />

∞<br />

∑ smρ<br />

m=n<br />

m ∞<br />

− A = P (1 − ρ) ∑ ρ<br />

m=n<br />

m − A = Pρ n − A<br />

�<br />

�<br />

where P ∈ infm≥n sm, supm≥n sm , the inequalities <strong>of</strong> interest obtained from the requirements<br />

are<br />

|Pρ n − A| ≤ |Pρ n − P| + |P − A|<br />

≤ k × δ δ<br />

+ 2 ×<br />

3k 6<br />

= 2δ<br />

3 .<br />

Combining the inequalities show that for ε < min {ε1, ε2},<br />

|S − A| ≤ δ 2δ<br />

+<br />

3 3<br />

Thus, as this modern reconstruction illustrates, it can be proved along the lines <strong>of</strong><br />

DIRICHLET’S pro<strong>of</strong> that for ε → 0 (i.e. ρ = 1 − ε → 1), the power series S (ρ) converges<br />

to A. <strong>The</strong> interrelation among the limit processes is contained in the specification <strong>of</strong> ε1<br />

and ε2. ✷<br />

= δ.<br />

Box 3: A modern reconstruction <strong>of</strong> DIRICHLET’s pro<strong>of</strong>.<br />

�<br />

�<br />

�<br />

�<br />

� .


240 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

❜<br />

✟ ✟✟✟✟✟✟✟✟✟<br />

� �<br />

❜<br />

✟ ✟✟✟✟✟✟✟✟✟<br />

� �<br />

❜<br />

✟ ✟✟✟✟✟✟✟✟✟<br />

� �<br />

−3π −2π −π π 2π 3π<br />

❜<br />

Figure 12.2: Graphical representation <strong>of</strong> ABEL’s “Exception”, ∑ ∞ n=1<br />

<strong>The</strong> series<br />

π 2<br />

− π 2<br />

❜<br />

❜<br />

(−1) n−1 sin nx<br />

n .<br />

However, it appears to me that this theorem admits [or suffers] exceptions.<br />

For instance, the series<br />

sin φ − 1 1<br />

sin 2φ + sin 3φ − . . . etc.<br />

2 3<br />

is discontinuous for every value (2m + 1) π <strong>of</strong> x where m is an integer. As is well<br />

known, a multitude <strong>of</strong> series with similar properties exist.” 35<br />

∞ (−1)<br />

∑<br />

n=1<br />

n−1 sin nx<br />

n<br />

(12.14)<br />

is a particularly simple trigonometric series: it is the Fourier series expansion <strong>of</strong> the<br />

function f (x) = x 2 on the interval ]−π, π[ (see figure 12.2). As such, it can be found<br />

in FOURIER’S works, for instance in the Théorie analytique de la chaleur, and even as<br />

a side result in one <strong>of</strong> EULER’S papers. 36 Possibly because EULER’S and FOURIER’S<br />

arguments for the convergence <strong>of</strong> the series (12.14) might have been wanting from<br />

the perspective <strong>of</strong> the new rigor, ABEL explicitly derived it as a result <strong>of</strong> some <strong>of</strong> the<br />

formulae proved in the binomial paper (see page 259, below).<br />

Aspects <strong>of</strong> ABEL’S “exception”. For subsequent reference, a few points concerning<br />

ABEL’S “exception” must be brought to attention. First, the exception was one <strong>of</strong> the<br />

35 “Anmerkung. In der oben angeführten Schrift des Herrn Cauchy (Seite 131) findet man folgende<br />

Lehrsatz:<br />

»Wenn die verschiedenen Glieder der Reihe<br />

u0 + u1 + u2 + u3 + . . . u.s.w.<br />

Functionen einer und derselben veränderlichen Größe sind, und zwar stetige Functionen, in Beziehung<br />

auf diese Veränderliche, in der Nähe eines besonderen Werthes, für welchen die Reihe<br />

convergirt, so ist auch die Summe s der Reihe, in der Nähe jenes besonderen Werthes, eine stetige<br />

Function von x.«<br />

Es scheint mir aber, daß dieser Lehrsatz Ausnahmen leidet. So ist z. B. die Reihe<br />

sin φ − 1 1<br />

sin 2φ + sin 3φ − . . . u.s.w.<br />

2 3<br />

unstetig für jeden Werth (2m + 1) π von x, wo m eine ganze Zahl ist. Bekanntlich giebt es eine<br />

Menge von Reihen mit ähnlichen Eigenschaften.” (N. H. <strong>Abel</strong>, 1826f, 316, footnote).<br />

36 (Fourier, 1822, 182, 241) and (L. Euler, 1754, 584); see also (I. Grattan-Guinness, 1970b, 84–85).


12.6. A curious reaction: Lehrsatz V 241<br />

“new series” which — according to ABEL’S opinion (see page 250, below) — had only<br />

recently entered analysis and brought so many paradoxes with it. Second, from a<br />

modern perspective, it is curious that ABEL called the series an “exception” and not a<br />

counter example or even a paradox. Although the binomial paper was translated from<br />

a French manuscript by CRELLE (see page 30), this choice <strong>of</strong> words appears not to have<br />

been merely accidental. Furthermore, it appears to have mattered to ABEL that the<br />

exception was not “singular” — if required, a multitude <strong>of</strong> similar exceptions could<br />

be devised. <strong>The</strong>se points will enter into the argument in chapter 21. Finally, it should<br />

be observed that the “exception” was actually a recurring item in ABEL’S works on<br />

rigorization. Above (see page 225), it has been described how ABEL employed the<br />

same series to criticize the practice <strong>of</strong> differentiating a series by differentiating each<br />

term. Similarly, it appeared in one <strong>of</strong> ABEL’S drafts when he wanted to probe the<br />

limits <strong>of</strong> the theorem — Lehrsatz V — which was his tailored replacement for Cauchy’s<br />

<strong>The</strong>orem. 37<br />

12.6 A curious reaction: Lehrsatz V<br />

ABEL’S fifth theorem: a revision <strong>of</strong> CAUCHY’S theorem. <strong>The</strong> fifth theorem <strong>of</strong> ABEL’S<br />

binomial paper plays a central role in a story to be told in a subsequent chapter. For<br />

the present, the theorem is mainly <strong>of</strong> interest because it, like Lehrsatz IV, provides an<br />

important combination <strong>of</strong> the three concepts currently under consideration: conver-<br />

gence, continuity, and power series. In his fifth theorem, ABEL found that the binomial<br />

series was a continuous function; a result which was inherently important in the ap-<br />

proach to the binomial theorem chosen by CAUCHY and adapted by ABEL. Again, the<br />

statement <strong>of</strong> the theorem is worth quoting, 38<br />

“Lehrsatz V. Let<br />

v0 + v1δ + v2δ 2 + . . . etc.,<br />

be a convergent series in which v0, v1, v2, . . . are continuous functions <strong>of</strong> one and<br />

the same variable quantity x between the boundaries x = a and x = b. <strong>The</strong>n the<br />

series<br />

f (x) = v0 + v1α + v2α 2 + . . . ,<br />

where α < δ is convergent and a continuous function <strong>of</strong> x between the same<br />

boundaries.” 39<br />

37 (N. H. <strong>Abel</strong>, [1827] 1881, 202); see page 245, below.<br />

38 <strong>The</strong> statement <strong>of</strong> the theorem as given in the paper is sloppy in a couple <strong>of</strong> respects. First, the β<br />

introduced at the end should obviously be a δ, and the convergence <strong>of</strong> the initial series<br />

∞<br />

∑ vnδ<br />

n=0<br />

n<br />

must obviously also be assumed. Both corrections have been made by the editors in ABEL’S collected<br />

works (N. H. <strong>Abel</strong>, 1881, I, 223–224).


242 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> Lehrsatz V. ABEL’S proved the fifth theorem by an approach closely<br />

resembling his pro<strong>of</strong> <strong>of</strong> the preceding theorem. He first did away with the first claim<br />

<strong>of</strong> convergence by referring to the fourth theorem. <strong>The</strong> fourth theorem also told him<br />

that the sum function was continuous with respect to α; and he then moved on to<br />

prove the continuity <strong>of</strong> the sum function considered to be a function <strong>of</strong> x. As was<br />

common practice, ABEL split the sum function into two<br />

φ (x) =<br />

n−1<br />

∑ vm (x) α<br />

m=0<br />

m ∞<br />

and ψ (x) = ∑ vm (x) α<br />

m=n<br />

m .<br />

Just as he had done for the fourth theorem, he rewrote ψ (x) as<br />

ψ (x) =<br />

∞<br />

∑<br />

m=n<br />

�<br />

α<br />

�m vm (x) δ<br />

δ<br />

m<br />

and introduced a quantity θ (x) to denote “the greatest among the quantities vmδ m ,<br />

vmδ m + vm+1δ m+1 , vmδ m + vm+1δ m+1 + vm+2δ m+2 .” 40 Thus, θ (x) in the fifth theorem<br />

took the place <strong>of</strong> the quantity p used in the pro<strong>of</strong> <strong>of</strong> the fourth theorem (see also be-<br />

low). In his further argument, ABEL used Lehrsatz III to write<br />

ψ (x) <<br />

�<br />

α<br />

�m θ (x)<br />

δ<br />

just as he had done previously and his pro<strong>of</strong> <strong>of</strong> the continuity <strong>of</strong> f (x) followed ex-<br />

actly the same arguments as had been used in the fourth theorem. Again, infinitesi-<br />

mals were used instead <strong>of</strong> explicit limit processes when ABEL claimed that m could be<br />

chosen such that ψ (x) = ω, which allowed him to write<br />

f (x) − f (x − β) = φ (x) − φ (x − β) + ω.<br />

<strong>The</strong>n it was a simple matter to observe that φ was a polynomial and therefore β could<br />

be chosen small enough that<br />

and the theorem had been proved.<br />

39 “Lehrsatz V. Es sei<br />

φ (x) − φ (x − β) = ω<br />

v0 + v1δ + v2δ 2 + . . . u.s.w.,<br />

eine Reihe, in welcher v0, v1, v2 continuierliche Functionen einer und derselben veränderlichen Größe<br />

x sind, zwischen den Grenzen x = a und x = b, so ist die Reihe<br />

f (x) = v0 + v1α + v2α 2 + . . . ,<br />

wo α < β [α < δ], convergent und eine stetige Function von x, zwischen denselben Grenzen.” (N.<br />

H. <strong>Abel</strong>, 1826f, 315).<br />

40 “wenn man durch θ (x) die größte unter den Größen vmδ m , vmδ m + vm+1δ m+1 , vmδ m + vm+1δ m+1 +<br />

vm+2δ m+2 u.s.w. bezeichnet.” (ibid., 315).


12.6. A curious reaction: Lehrsatz V 243<br />

An objection to Lehrsatz V: is θ uniformly bounded? <strong>The</strong> major problem with<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> the fifth theorem is closely tied to its connection with his pro<strong>of</strong> <strong>of</strong> the<br />

preceding theorem. In his pro<strong>of</strong> <strong>of</strong> Lehrsatz IV, ABEL had introduced the quantity p to<br />

denote the largest quantity among the partial tails <strong>of</strong> the series. In the pro<strong>of</strong> <strong>of</strong> Lehrsatz<br />

IV, ABEL’S reasoning can be ‘saved’ by the observation that since the series ∑ vmδ m is<br />

convergent, its tails are bounded. <strong>The</strong>refore, an upper bound — if not an outright max-<br />

imum — will exist which can be used for p. Such an argument is nowhere to be found<br />

in ABEL’S pro<strong>of</strong>, and there are two points indicating that it was neither at his disposal<br />

nor <strong>of</strong> his concern. First, ABEL spoke <strong>of</strong> “the largest among” an infinite collection <strong>of</strong><br />

quantities, i.e. <strong>of</strong> a maximum. If he had had anything but a naïve intuition about this<br />

step in his argument he might well have expressed himself differently using phrases<br />

analogous to “bounded by”. Second, in the pro<strong>of</strong> <strong>of</strong> the fifth theorem which is mod-<br />

elled precisely over the pro<strong>of</strong> <strong>of</strong> the fourth theorem, this exact step in the argument<br />

falls apart.<br />

When, in the pro<strong>of</strong> <strong>of</strong> Lehrsatz V, ABEL introduced θ (x) analogous to the quantity<br />

p above<br />

θ (x) = largest quantity among<br />

n+k<br />

∑ vm (x) δ<br />

m=n<br />

m for k ≥ 0, (12.15)<br />

it seems to be a point-wise definition <strong>of</strong> the function θ (x). ABEL clearly thought <strong>of</strong><br />

θ (x) as a quantity which, given x represented the largest among an infinite collection<br />

<strong>of</strong> quantities each depending on x. When, in the pro<strong>of</strong>, ABEL claimed that<br />

ψ (x) = ψ (x − β) = ω,<br />

he implicitly used a supposed property <strong>of</strong> the function θ (x) — that the choice <strong>of</strong> n could<br />

be made uniformly throughout a small region surrounding x. However, in ABEL’S<br />

argument, there is no way <strong>of</strong> assuring that θ satisfies this requirement.<br />

As P. L. M. SYLOW (1832–1918) has remarked, 41 ABEL’S argument is sound if one<br />

further restriction is imposed on the convergence. If a constant M exists which uni-<br />

formly bounds the general term around x0<br />

|vm (x) δ m | ≤ M for all m and for all x ∈ � x0 − x ′ , x0 + x ′′� ,<br />

ABEL’S reasoning can be applied by observing that both<br />

|ψ (x)| and |ψ (x − β)| will be less than M<br />

� �<br />

αδ m<br />

1 − α δ<br />

However, as observed, this is a reconstruction and certainly not part <strong>of</strong> ABEL’S argu-<br />

ment.<br />

41 (N. H. <strong>Abel</strong>, 1881, II, 303).<br />

.


244 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

Another pro<strong>of</strong> <strong>of</strong> the Lehrsatz V from ABEL’S notebook. In ABEL’S notebooks, re-<br />

sults similar to the Lehrsatz V can be found treated in the manuscript Sur les séries<br />

which was presumably written in 1827. 42 Thus, ABEL returned to the theorem after<br />

the binomial paper had been published and attacked it from a slightly different per-<br />

spective. In his notes on the Sur les séries, M. S. LIE (1842–1899) interprets this fact<br />

as clear evidence that ABEL had become dissatisfied with the version printed in the<br />

Journal. 43 <strong>The</strong> manuscript was never completed for printing and its contents exhibit<br />

the characteristics <strong>of</strong> a draft. In particular, the precise assumptions and some <strong>of</strong> the<br />

notation are not made explicit and interpretation is slightly difficult. I interpret the<br />

relevant part <strong>of</strong> the manuscript as presenting two new deductions <strong>of</strong> the Lehrsatz V.<br />

In his manuscript, ABEL dealt with a function f (y) introduced as a power series<br />

in x with coefficients which vary continuously in y,<br />

f (y) =<br />

∞<br />

∑ φn (y) x<br />

n=0<br />

n<br />

(12.16)<br />

and assumed that the series was convergent for x < α and all values <strong>of</strong> y near β.<br />

ABEL’S first “pro<strong>of</strong>” <strong>of</strong> the continuity <strong>of</strong> f at y = β consisted in interchanging the<br />

limit processes, and he wrote it as<br />

lim f (y) =<br />

y=β−ω<br />

∞ �<br />

∑<br />

n=0<br />

lim<br />

y=β−ω φn (y)<br />

�<br />

x n =<br />

∞<br />

∑ Anx<br />

n=0<br />

n = R<br />

with the convention An = lim y=β−ω φn (y). <strong>The</strong> practice <strong>of</strong> interchanging limit pro-<br />

cesses had been among the points which attracted ABEL’S criticism, in particular when<br />

it came to term-wise differentiation (see above). Accordingly, ABEL did not simply in-<br />

terchange the two processes but made the additional restriction that the series R had<br />

to be convergent.<br />

However, as ABEL’S own “exception” could have illustrated, even the convergence<br />

<strong>of</strong> the resulting series was sufficient to warrant the general interchange <strong>of</strong> limits. Thus,<br />

ABEL had to make some use <strong>of</strong> the particular form <strong>of</strong> the series (12.16). However, ABEL<br />

made no remarks on this argument and due to the style <strong>of</strong> the notebook, its role and<br />

status — was it an observation? a theorem? a hypothesis? — remains my suggestive<br />

interpretation.<br />

ABEL’S second “deduction” is much more interesting and is presented here based<br />

on ABEL’S original argument and LIE’S reconstruction <strong>of</strong> it. 44 In this deduction, ABEL<br />

studied the differences between the corresponding terms <strong>of</strong> the series for f (β − ω)<br />

and f (β).<br />

42 (N. H. <strong>Abel</strong>, [1827] 1881, 201–202).<br />

43 (N. H. <strong>Abel</strong>, 1881, II, 326).<br />

44 (ibid., II, 326).<br />

(φn (β − ω) − An) x n .


12.6. A curious reaction: Lehrsatz V 245<br />

Assuming that x1 was a value such that x < x1 < α, ABEL introduced a bound by<br />

assuming that the m’th term was the maximum <strong>of</strong> these differences,<br />

(φm (β − ω) − Am) x m 1<br />

= max<br />

n≥0 {(φn (β − ω) − An) x n 1 } . (12.17)<br />

This step resembles the introduction <strong>of</strong> the problematic quantity θ (x) in the bino-<br />

mial paper and the existence (i.e. finiteness) <strong>of</strong> such a maximum was apparently un-<br />

problematic to ABEL. Accordingly, LIE has suggested the same method <strong>of</strong> saving<br />

ABEL’S argument as SYLOW had done for the Lehrsatz V, i.e. by turning its existence<br />

into an explicit assumption (see above). ABEL concluded that<br />

f (β − ω) − R = ξ<br />

1 − x x 1<br />

(φm (β − ω) − Am) x m 1<br />

for some ξ ∈ [−1, 1]. When he let ω vanish, ABEL observed that the term<br />

φm (β − ω) − Am<br />

also vanished by the continuity <strong>of</strong> φm. <strong>The</strong>refore, ABEL concluded, the function f was<br />

continuous.<br />

As described, ABEL’S two notebook pro<strong>of</strong>s <strong>of</strong> the Lehrsatz V are slightly different<br />

from the printed version. However, they share the same structure and many <strong>of</strong> the<br />

methods which they apply, in particular concerning the belief in the existence <strong>of</strong> uni-<br />

form bounds (12.15 and 12.17). It is tempting to speculate with LIE that ABEL had<br />

realized that his original pro<strong>of</strong> <strong>of</strong> Lehrsatz V was problematic — perhaps seizing on<br />

the same objection as SYLOW did and proposing the solution which amounts to uni-<br />

form convergence (see above). However, despite the new pro<strong>of</strong>s, ABEL’S treatment<br />

<strong>of</strong> Lehrsatz V continued to suffer from essentially the same problems and such an in-<br />

terpretation is not compelling. If ABEL had become uneasy about his pro<strong>of</strong>, it was<br />

probably for another reason or perhaps he just wanted another pro<strong>of</strong> <strong>of</strong> a well estab-<br />

lished result?<br />

Probing the extent <strong>of</strong> Lehrsatz V. Following his new pro<strong>of</strong> <strong>of</strong> Lehrsatz V in the note-<br />

book, ABEL observed that the theorem demonstrated the continuity <strong>of</strong> the function<br />

f (y) =<br />

∞<br />

x<br />

∑<br />

n=1<br />

n sin ny<br />

n<br />

for all x < 1, although for x = 1, the function — which was the “exception” <strong>of</strong> his<br />

binomial paper — had certain discontinuities. Under similar assumptions, the series<br />

corresponding to x = 1 could also fail to be divergent, altogether, ABEL observed and<br />

exemplified. <strong>The</strong>se remarks again illustrate ABEL’S repeated criticism <strong>of</strong> the unwar-<br />

ranted passage to the limit in series.


246 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

12.7 From power series to absolute convergence<br />

As indicated in his letter to HANSTEEN and in the general approach <strong>of</strong> the binomial<br />

paper, ABEL put a lot <strong>of</strong> emphasis on power series in his attempt to rebuild the theory<br />

<strong>of</strong> series. In particular, ABEL’S replacement for the invalidated Cauchy <strong>The</strong>orem was<br />

based on a particular kind <strong>of</strong> series which were power series in one variable with<br />

coefficients which were continuous functions <strong>of</strong> another variable. Well into the second<br />

half <strong>of</strong> the nineteenth century, this particular argument was found to be better recast<br />

within a concept <strong>of</strong> absolute convergence which had emerged over the century.<br />

Emergence <strong>of</strong> a concept <strong>of</strong> absolute convergence. During the 19 th century, numer-<br />

ical (absolute) values <strong>of</strong> real numbers and the moduli <strong>of</strong> complex numbers entered<br />

ever more explicitly in arguments <strong>of</strong> analysis. As described in the examples from<br />

CAUCHY’S Cours d’analyse and ABEL’S binomial paper, the only way mathematicians<br />

could describe numerical values in the first decades <strong>of</strong> the century was through verbal<br />

formulations. Generally, CAUCHY was quite careful about these in stating his theo-<br />

rems on series; but proper concern for numerical values was <strong>of</strong>ten lacking in ABEL’S<br />

formulations. 45 It appears that notation such as |x| was only invented by WEIER-<br />

STRASS in unpublished papers <strong>of</strong> the 1840s and did not become customary until the<br />

1870s. 46<br />

In the first decades <strong>of</strong> the 19 th century, series <strong>of</strong> numerical values mostly entered<br />

the picture in connection with the multiplication theorem. In the Cours d’analyse, when<br />

CAUCHY generalized his multiplication theorem for series <strong>of</strong> positive terms to more<br />

arbitrary series, he based his argument on the assumption <strong>of</strong> convergence <strong>of</strong> the series<br />

<strong>of</strong> absolute terms. However, despite its use in proving theorems, CAUCHY’S implicit<br />

concept <strong>of</strong> absolute convergence still lacked most <strong>of</strong> its later structural position.<br />

Immediately following the pro<strong>of</strong> <strong>of</strong> the multiplication theorem, CAUCHY took an<br />

interesting step in investigating the consequences <strong>of</strong> relaxing the assumptions. He<br />

proved, based on squaring the series<br />

∞<br />

∑<br />

n=1<br />

(−1) n−1<br />

√ , (12.18)<br />

n<br />

that the assumption <strong>of</strong> absolute convergence was indeed necessary: Because the terms<br />

<strong>of</strong> the alternating series (12.18) are decreasing in absolute value, the series was conver-<br />

gent as CAUCHY had proved. 47 However, it was not absolutely convergent and when<br />

CAUCHY produced the square <strong>of</strong> the series, he obtained another divergent series. 48<br />

Here, CAUCHY took a rather modern step <strong>of</strong> using a counter example to a fictitious<br />

45 See e.g. the editors’ remark in (Lakatos, 1976, 134).<br />

46 (K. Weierstrass, 1876, 78) and e.g. (K. Weierstrass, [1841] 1894).<br />

47 (A.-L. Cauchy, 1821a, 144).<br />

48 (ibid., 149–150). <strong>The</strong> divergence <strong>of</strong> the series ∑ 1<br />

√ n can be obtained by comparing with the harmonic<br />

series.


12.7. From power series to absolute convergence 247<br />

more general theorem in order to illustrate that the requirements <strong>of</strong> his own theorem<br />

were necessary. However, as F. MERTENS (1840–1927) was later to show, 49 if scru-<br />

tinized more carefully, the example illustrated that the multiplication theorem could<br />

fail if both factors were non-absolutely convergent. This led P. L. WANTZEL (1814–<br />

1848) to prove a version <strong>of</strong> the multiplication theorem which only assumed absolute<br />

convergence <strong>of</strong> one <strong>of</strong> the factors (the other factor just being assumed convergent).<br />

<strong>The</strong> real beginning <strong>of</strong> a concept <strong>of</strong> absolute convergence came with DIRICHLET’S<br />

paper on primes in arithmetic progression which was published in 1837. 50 In that<br />

paper, DIRICHLET introduced a separation <strong>of</strong> convergent series into two classes based<br />

on the convergence <strong>of</strong> the series which resulted when the terms were replaced by<br />

the absolute values: Either the series <strong>of</strong> absolute values remained bounded or it was<br />

unbounded. For series <strong>of</strong> the first class (absolutely convergent series), DIRICHLET<br />

stated that their convergence and sum remained unaffected if the order <strong>of</strong> terms was<br />

altered. In particular, in double (and multiple) sums, the order <strong>of</strong> summation wold<br />

not effect the result. Dirichlet observed that these properties — which were certainly<br />

nice and expected — could fail to hold for series <strong>of</strong> the second class, and he gave two<br />

examples <strong>of</strong> what could happen: a convergent series could either become divergent<br />

or alter its sum if its terms were rearranged. 51<br />

For his Habilitation in 1854, RIEMANN presented a paper on the representability <strong>of</strong><br />

functions by trigonometric series. 52 <strong>The</strong> paper is a milestone in the theory <strong>of</strong> trigono-<br />

metric series and the theory <strong>of</strong> integrals and also contains interesting remarks on the<br />

concept <strong>of</strong> absolute and non-absolute convergence. In the historical preface, RIEMANN<br />

outlined the previous developments in the field and claimed that DIRICHLET’S impor-<br />

tant 1829 paper on the convergence <strong>of</strong> trigonometric series was directly inspired by<br />

DIRICHLET’S discovery <strong>of</strong> the distinction between absolute and non-absolute conver-<br />

gence. 53 RIEMANN expressed his belief that the prevalence <strong>of</strong> power series in analysis<br />

was the reason why those concepts had not previously been separated. 54 <strong>The</strong>re are no<br />

obvious traces <strong>of</strong> the alleged inspiration visible in DIRICHLET’S paper <strong>of</strong> 1829 but —<br />

as mentioned — the distinction became very explicit in a paper with a different topic<br />

in 1837.<br />

RIEMANN advanced a step beyond DIRICHLET’S observation <strong>of</strong> the differences be-<br />

tween absolutely and non-absolutely (conditionally) convergent series when he de-<br />

scribed a very simple method by which the partial sums <strong>of</strong> a conditionally convergent<br />

series could be made to approach any given value by proper rearrangement <strong>of</strong> the<br />

terms <strong>of</strong> the series. Central to RIEMANN’S argument was the realization that if a se-<br />

ries ∑ an was conditionally convergent, the series <strong>of</strong> its positive and negative terms<br />

49 (Mertens, 1875).<br />

50 (G. L. Dirichlet, 1837), see also e.g. (I. Grattan-Guinness, 1970b, 94–95).<br />

51 See also (ibid., 94–95).<br />

52 (B. Riemann, 1854).<br />

53 (ibid., 235) DIRICHLET’S paper is (G. L. Dirichlet, 1829).<br />

54 (B. Riemann, 1854, 235). See below.


248 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

would both have to converge to infinity. This could be used to prescribe a procedure<br />

by which examples such as those given by DIRICHLET could be constructed. As D.<br />

LAUGWITZ (1932–2000) remarks, 55 the Riemann arrangement theorem is not a particu-<br />

larly deep mathematical result in its own right but exemplifies the new approach to<br />

the concepts <strong>of</strong> convergence which was developing in the mid-nineteenth century.<br />

P. D. G. DU BOIS-REYMOND (1831–1889) on a generalization <strong>of</strong> Lehrsatz V. A con-<br />

cept <strong>of</strong> absolutely convergent series was thus being established in the nineteenth cen-<br />

tury and it was gradually emerging as a very central and useful tool in doing analysis.<br />

In 1871, DU BOIS-REYMOND published a short note which treated CAUCHY’S theorem<br />

on the continuity <strong>of</strong> an infinite sum <strong>of</strong> continuous functions which had also been the<br />

subject <strong>of</strong> ABEL’S Lehrsatz V. 56 DU BOIS-REYMOND was an active participant in the<br />

restructuring <strong>of</strong> analysis which — flooding from WEIERSTRASS’ lectures in Berlin —<br />

took place in the last part <strong>of</strong> the nineteenth century. In the note which falls into this<br />

Weierstrassian tradition <strong>of</strong> rigorizing analysis, DU BOIS-REYMOND proved a theorem<br />

to the following effect.<br />

<strong>The</strong>orem 12 (DU BOIS-REYMOND) If a series<br />

is considered, for which the series<br />

∞<br />

∑ wn (x) µn<br />

n=1<br />

∞<br />

∑ µn<br />

n=1<br />

(12.19)<br />

converges absolutely and for which the functions wn are continuous functions on the interval<br />

[a, b] and all the wn as well as limn→∞ wn remain finite on that interval, then the series (12.19)<br />

is a continuous function <strong>of</strong> x on the interval [a, b]. ✷<br />

As well as providing a pro<strong>of</strong> <strong>of</strong> this theorem, DU BOIS-REYMOND also gave ex-<br />

amples in which the theorem would warrant continuity and examples in which no<br />

such conclusion could be drawn from the theorem. <strong>The</strong> theorem which DU BOIS-<br />

REYMOND presented was a generalization <strong>of</strong> ABEL’S fifth theorem; the latter could be<br />

deduced from the former (see box 4). It is interesting to compare the two theorems<br />

and the motivating problems which inspired them.<br />

<strong>The</strong> focus on power series: Comparing theorems. ABEL’S fifth theorem sought to<br />

avoid the over-generality <strong>of</strong> Cauchy’s <strong>The</strong>orem by focusing on some particular form<br />

<strong>of</strong> convergence similar to the convergence <strong>of</strong> power series. Compared to Cauchy’s<br />

55 (Laugwitz, 1999, 211).<br />

56 (Bois-Reymond, 1871); DU BOIS-REYMOND referred to both CAUCHY and ABEL.


12.7. From power series to absolute convergence 249<br />

ABEL’s Lehrsatz V derived from DU BOIS-REYMOND’s theorem Actually, ABEL’S<br />

fifth theorem is a consequence <strong>of</strong> DU BOIS-REYMOND’S theorem.<br />

Corollary 1 Assume that<br />

∞<br />

∑ vn (x) δ<br />

n=1<br />

n<br />

(12.20)<br />

is convergent for some δ > 0 and that the functions vn are continuous functions <strong>of</strong> x on some<br />

interval I. <strong>The</strong>n, for any 0 < α < δ, the function<br />

f (x) =<br />

∞<br />

∑ vn (x) α<br />

n=1<br />

n<br />

(12.21)<br />

is a continuous function <strong>of</strong> x on the interval I. ✷<br />

PROOF We wish to use DU BOIS-REYMOND’S theorem to prove the theorem stated<br />

above. For this, we write<br />

and denote<br />

f (x) =<br />

∞<br />

∑ vn (x) δ<br />

n=1<br />

n � α<br />

�n δ<br />

wn (x) = vn (x) δ n and<br />

�<br />

α<br />

�n µn = .<br />

δ<br />

Now, we are ready to test the requirements <strong>of</strong> DU BOIS-REYMOND’S theorem. <strong>The</strong><br />

first requirement, that ∑ µn converges absolutely is obviously satisfied since α < δ.<br />

Secondly, the functions wn are obviously finite and continuous since this was required<br />

<strong>of</strong> vn. Lastly, we have to show that limn→∞ wn (x) is also finite. However, this is<br />

an easy consequence to draw from the convergence <strong>of</strong> (12.20) which ensures us that<br />

limn→∞ wn (x) = 0 for all x ∈ I. Thus, the continuity <strong>of</strong> (12.21) follows from DU<br />

BOIS-REYMOND’S theorem. �<br />

Box 4: ABEL’s Lehrsatz V derived from DU BOIS-REYMOND’s theorem


250 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

<strong>The</strong>orem, one further assumption was introduced in ABEL’S fifth theorem to the effect<br />

that the series<br />

f (x, δ) =<br />

∞<br />

∑ vn (x) δ<br />

n=1<br />

n<br />

(12.22)<br />

was convergent for some δ > 0; and the conclusion <strong>of</strong> continuity only applied to<br />

f (x, α) where 0 < α < δ, i.e. on the interior <strong>of</strong> the interval <strong>of</strong> convergence. Apparently,<br />

this further requirement did away with ABEL’S own exception to Cauchy’s <strong>The</strong>orem; the<br />

series<br />

∞ (−1)<br />

∑<br />

n=1<br />

n sin nx<br />

n<br />

could not easily be transformed into a power series with radius <strong>of</strong> convergence sharply<br />

greater than one. However, it was hardly this barring <strong>of</strong> known exceptions which<br />

prompted ABEL to formulate his fifth theorem in the way he did. Instead, two other<br />

factors are most likely to have contributed to the formulation <strong>of</strong> the theorem. First,<br />

ABEL’S fifth theorem was closely modelled over Lehrsatz IV, and their pro<strong>of</strong>s were<br />

almost identical. At the crucial step <strong>of</strong> the pro<strong>of</strong>, a generalization was made from<br />

the constant p to the function θ (x) which were both designed to serve as uniform<br />

bounds. Second, the focus on power series was introduced by ABEL as a heuristic<br />

which “saved” ordinary intuition and many previously established results.<br />

In the letter to HANSTEEN quoted above, ABEL expressed his concern over a dark-<br />

ness which he saw persisting in the field <strong>of</strong> mathematical analysis. He even described<br />

some ideas concerning the reason for the relatively few paradoxes which these obscure<br />

and ill-founded procedures had created.<br />

“In my opinion, it [the reason for the few paradoxes] lies in the fact that the<br />

functions which analysis has dealt with have mostly been expressible by powers.<br />

As soon as others [functions] enter, which certainly does not happen <strong>of</strong>ten, it <strong>of</strong>ten<br />

does not go well and from false conclusions a string <strong>of</strong> connected false theorems<br />

flow.” 57<br />

According to ABEL, it was the introduction <strong>of</strong> new kinds <strong>of</strong> series which had pro-<br />

duced problems for theorems which implicitly relied on properties <strong>of</strong> power series<br />

although they were <strong>of</strong>ten expressed so as to apply to all series. As RIEMANN’S sim-<br />

ilar remarks seems to indicate, this was a generally held — and valid — belief in the<br />

nineteenth century.<br />

Half a century after ABEL’S solution to the problem raised by his counter example<br />

to Cauchy’s <strong>The</strong>orem, DU BOIS-REYMOND devised another answer to the same prob-<br />

lem. Set in a different time and inspired by the system <strong>of</strong> analysis which WEIERSTRASS<br />

57 “Efter mine Tanker ligger den deri at de Functioner som Analysen hidentil har beskjæftiget sig med<br />

mestendels lade sig udtrykke ved Potenser. — Saasnart der komme andre imellem hvilket rigtig nok<br />

ikke <strong>of</strong>te er Tilfældet saa gaaer det gjerne ikke godt og af falske Slutninger opstaae da en Mængde<br />

med hinanden forbundne urigtige Sætninger.” (<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. N. H. <strong>Abel</strong>,<br />

1902a, 22–23).


12.8. Product theorems <strong>of</strong> infinite series 251<br />

taught in Berlin, DU BOIS-REYMOND’S solution differed from ABEL’S at a conceptual<br />

level. When compared with ABEL’S theorem and its pro<strong>of</strong>, DU BOIS-REYMOND’S the-<br />

orem differed in three respects. First, when he focused on the purpose which ABEL’S<br />

use <strong>of</strong> power series had served in the original pro<strong>of</strong>, DU BOIS-REYMOND could relax<br />

the assumptions and only assume that the series ∑ µn converged (absolutely). Second,<br />

during the semi-century, a more rigid concept <strong>of</strong> absolute convergence had emerged<br />

which made the use <strong>of</strong> numerical values in series explicit and consistent. In the<br />

process, absolute convergence had become a concept about which theorems could be<br />

proved. Finally, in DU BOIS-REYMOND’S pro<strong>of</strong>, the uniformity requirement discussed<br />

above was explicitly taken into account by the assumption that limn→∞ wn (x) remain<br />

finite.<br />

<strong>The</strong> example <strong>of</strong> DU BOIS-REYMOND’S revision <strong>of</strong> Cauchy’s <strong>The</strong>orem serves to illus-<br />

trate how the concept <strong>of</strong> absolute convergence became a very central and powerful<br />

concept in the theory <strong>of</strong> series. Formulating theorems using absolute convergence <strong>of</strong>-<br />

ten led to more functional assumptions which were directly usable in the pro<strong>of</strong>s. In<br />

this example, this was contrasted with the older focus on formal assumptions which<br />

stressed the particular formal appearance <strong>of</strong> the objects — here the particular form <strong>of</strong><br />

the series (12.22) — under consideration.<br />

12.8 Product theorems <strong>of</strong> infinite series<br />

Besides the theorems discussed above, another basic, important theorem on infinite<br />

series received renewed interest in ABEL’S paper on the binomial theorem; and again,<br />

this theorem goes back to CAUCHY’S Cours d’analyse.<br />

In the binomial paper, ABEL’S sixth and final preliminary theorem dealt with the<br />

product <strong>of</strong> two infinite series. In its presentation, it reveals an intriguing transition in<br />

the understanding <strong>of</strong> the concept <strong>of</strong> absolute convergence and reads as follows,<br />

“Lehrsatz VI. When by ρ0, ρ1,ρ2 etc., ρ ′ 0 , ρ′ 1 , ρ′ 2 etc one designates the numerical<br />

values <strong>of</strong> the respective terms <strong>of</strong> two convergent series<br />

then the series<br />

v0 + v1 + v2 + . . . = p and<br />

v ′ 0 + v ′ 1 + v ′ 2 + . . . = p ′ ,<br />

ρ0 + ρ1 + ρ2 + . . . and<br />

ρ ′ 0 + ρ ′ 1 + ρ ′ 2 + . . .<br />

are likewise convergent. Similarly, the series<br />

whose general term is<br />

r0 + r1 + r2 + · · · + rm<br />

rm = v0v ′ m + v1v ′ m−1 + v2v ′ m−2 + · · · + vmv ′ 0,


252 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

and whose sum is<br />

will also be convergent.” 58<br />

(v0 + v1 + v2 + . . . ) × � v ′ 0 + v ′ 1 + v ′ 2 + . . . �<br />

As it appears, the theorem consists <strong>of</strong> two halves each contributing a distinct con-<br />

clusion:<br />

1. Convergence <strong>of</strong> terms implies convergence <strong>of</strong> numerical terms.<br />

2. Convergence <strong>of</strong> the Cauchy product toward the correct sum.<br />

<strong>The</strong> first <strong>of</strong> these two conclusions is wrong, and it is difficult to explain the mishap<br />

in ABEL’S presentation. In the collected works, it has been corrected without com-<br />

ments by replacing “so sind die Reihen [. . . ] ebenfalls noch convergent” with “si les<br />

séries [. . . ] sont de même convergente”. 59 In the pro<strong>of</strong>, ABEL does not give argu-<br />

ments for the first part <strong>of</strong> the supposed theorem; instead it is used as an assumption<br />

in proving the Cauchy product theorem.<br />

Thus, based on ABEL’S pro<strong>of</strong>, SYLOW and LIE attributed the mishap to a slip <strong>of</strong> the<br />

pen or perhaps a slight incompetence on the part <strong>of</strong> the translator. This is probably<br />

the best interpretation available, but a little more may perhaps be inferred from the<br />

fact that such a misprint found its way into a pr<strong>of</strong>essional journal — as did a number<br />

<strong>of</strong> others. First, this fact suggests that conceptual handling <strong>of</strong> series and series <strong>of</strong> nu-<br />

merical values was not very well established among the mathematical class to which<br />

CRELLE belonged. And secondly, we may also be tempted to infer something on the<br />

58 “Lehrsatz VI. Bezeichnet man durch ρ0, ρ1, ρ2 u.s.w., ρ ′ 0 , ρ′ 1 , ρ′ 2 u.s.w. die Zahlenwerthe der resp. Glieder<br />

zweier convergenten Reihen<br />

so sind die Reihen<br />

v0 + v1 + v2 + . . . = p und<br />

v ′ 0 + v′ 1 + v′ 2 + . . . = p′ ,<br />

ρ0 + ρ1 + ρ2 + . . . und<br />

ρ ′ 0 + ρ′ 1 + ρ′ 2<br />

ebenfalls noch convergent, und auch die Reihe<br />

deren allgemeines Glied<br />

und deren Summe<br />

+ . . .<br />

r0 + r1 + r2 + · · · + rm<br />

rm = v0v ′ m + v1v ′ m−1 + v2v ′ m−2 + · · · + vmv ′ 0 ,<br />

(v0 + v1 + v2 + . . . ) × � v ′ 0 + v′ 1 + v′ 2<br />

ist, wird convergent seyn.” (N. H. <strong>Abel</strong>, 1826f, 316–317).<br />

59 (N. H. <strong>Abel</strong>, 1881, 225).<br />

+ . . . �


12.8. Product theorems <strong>of</strong> infinite series 253<br />

standards <strong>of</strong> the newly established journal, which was hampered by some similar and<br />

less grave misprints in the first years, although its standards <strong>of</strong> technical printing were<br />

quite high.<br />

ABEL’S pro<strong>of</strong> <strong>of</strong> Cauchy product theorem followed a path similar to those taken by<br />

CAUCHY in his pro<strong>of</strong>s (see section 11.5). ABEL let pm and p ′ m denote the partial sums<br />

<strong>of</strong> the factors p and p ′ and wrote<br />

After introducing the notation<br />

he found<br />

2m<br />

∑ rk = pmp<br />

k=0<br />

′ m−1<br />

m+ ∑<br />

k=0<br />

u =<br />

p kv ′ 2m−k<br />

� �� �<br />

=t<br />

+<br />

∞<br />

∑ ρk and u<br />

k=0<br />

′ ∞<br />

= ∑ ρ<br />

k=0<br />

′ k ,<br />

m−1<br />

∑ v2m−kp k=0<br />

′ k<br />

� �� �<br />

=t ′<br />

m−1<br />

t < u ∑ ρ<br />

k=0<br />

′ 2m−k and t′ m−1<br />

′ < u ∑ ρ2m−k k=0<br />

. (12.23)<br />

“without reference to the sign”, i.e. for the numerical values <strong>of</strong> t and t ′ . ABEL then<br />

employed the Cauchy sequence characterization <strong>of</strong> convergence (for its prominent posi-<br />

tion in the <strong>Abel</strong>ian framework, see above) to ensure that since the series ∑ ρ k and ∑ ρ ′ k<br />

were convergent, the sums<br />

m−1<br />

∑ ρ<br />

k=0<br />

′ 2m−k and<br />

m−1<br />

∑ ρ2m−k k=0<br />

would both tend to zero as m grew to infinity. Thus, ABEL claimed, by setting m equal<br />

to infinity, the equation (12.23) became<br />

∞<br />

∑ rk =<br />

k=0<br />

�<br />

∞<br />

∑ vk k=0<br />

�<br />

×<br />

�<br />

∞<br />

∑ v<br />

k=0<br />

′ �<br />

k .<br />

At this point, the theorem was proved, but ABEL continued his argument by gen-<br />

eralizing the theorem through the use <strong>of</strong> power series. ABEL now abandoned the<br />

assumptions that both factors had to be absolutely convergent in favor <strong>of</strong> the assump-<br />

tion that both the factors and the Cauchy product were (simply) convergent. In his<br />

notes on ABEL’S binomial paper, SYLOW wrote <strong>of</strong> this generalization: “<strong>The</strong> theorem<br />

VI is due to Cauchy but the new form which it is given [. . . ] originates with ABEL.” 60<br />

<strong>The</strong> generalized version <strong>of</strong> the Cauchy product theorem can thus be stated as follows.<br />

<strong>The</strong>orem 13 (Generalized Cauchy product theorem) If the three series<br />

∞<br />

∑ tk, k=0<br />

∞<br />

∑ t<br />

k=0<br />

′ k , and<br />

∞<br />

∑<br />

k=0<br />

∑<br />

tnt<br />

n+m=k<br />

′ m<br />

60 “Le théorème VI est dú à Cauchy, mais la forme nouvelle qu’il a reçue page 226 appartient à <strong>Abel</strong>.”<br />

(ibid., II, 303).


254 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

are all convergent, then<br />

� � �<br />

∞<br />

∞<br />

∑ tk × ∑ t<br />

k=0<br />

k=0<br />

′ �<br />

k =<br />

∞<br />

∑<br />

∑<br />

k=0 n+m=k<br />

tnt ′ m.<br />

ABEL proved this theorem through elegant application <strong>of</strong> the previously established<br />

theorems. First he simply assumed that {tk} and � t ′ �<br />

k were two sequences<br />

converging to zero. <strong>The</strong>n, by his Lehrsatz II, the two power series<br />

∞<br />

∑ tkα k=0<br />

k and<br />

∞<br />

∑ t<br />

k=0<br />

′ kαk would be convergent for 0 < α < 1, even if numerical values were taken. Thus, be the<br />

version <strong>of</strong> the Cauchy product theorem expressed in Lehrsatz VI,<br />

�<br />

∞<br />

∑ tkα k=0<br />

k<br />

� �<br />

∞<br />

× ∑ t<br />

k=0<br />

′ kαk �<br />

=<br />

�<br />

∞<br />

∑<br />

k=0<br />

∑ tnt<br />

n+m=k<br />

′ m<br />

and by letting α → 1, Lehrsatz IV supplied the desired conclusion.<br />

12.9 ABEL’s pro<strong>of</strong> <strong>of</strong> the binomial theorem<br />

After having established his six preliminary theorems, ABEL proceeded to the bino-<br />

mial theorem. Throughout the binomial paper, ABEL never equated the expressions<br />

(1 + x) m and<br />

∞<br />

∑<br />

µ=0<br />

∏ µ−1<br />

k=0<br />

(m − k)<br />

x<br />

µ!<br />

µ<br />

directly because the former could be multi-valued whereas the latter had only a single<br />

value as a function <strong>of</strong> x (see page 227). Instead, ABEL began his argument by asking<br />

for which values <strong>of</strong> m and x the binomial series<br />

φ (x) =<br />

∞<br />

∑ mµx<br />

µ=0<br />

µ<br />

converged, where he let mµ represented the binomial coefficient<br />

mµ =<br />

m (m − 1) (m − 2) . . . (m − µ + 1)<br />

µ!<br />

�<br />

α k ,<br />

= ∏µ−1 s=0 (m − s) .<br />

µ!<br />

Transformation into real series. ABEL wanted to include complex values <strong>of</strong> m and<br />

x, which he introduced by letting<br />

x = a + ib and m = k + ik ′ ,<br />


12.9. ABEL’s pro<strong>of</strong> <strong>of</strong> the binomial theorem 255<br />

where the notation i has been adopted for ABEL’S<br />

binomial coefficients in polar form,<br />

which meant<br />

δµ<br />

m − µ + 1<br />

µ<br />

= δµ<br />

� cos γµ + i sin γµ<br />

√ −1. ABEL wrote the factors <strong>of</strong> the<br />

� �<br />

cos γµ + i sin γµ ,<br />

� k + ik<br />

= ′ − µ + 1<br />

,<br />

µ<br />

and for each given µ, the values <strong>of</strong> δµ and γµ could be found. When these factors were<br />

multiplied to produce the binomial coefficients, ABEL found<br />

� �<br />

�<br />

mµ =<br />

With the conventions<br />

�<br />

µ<br />

∏ δn<br />

n=1<br />

×<br />

cos<br />

x = α (cos φ + i sin φ) , λµ =<br />

�<br />

µ<br />

∑ γn<br />

n=1<br />

+ i sin<br />

�<br />

µ<br />

∑ γn<br />

n=1<br />

µ<br />

∏ δn, and θµ = µφ +<br />

n=1<br />

��<br />

.<br />

µ<br />

∑ γn,<br />

n=1<br />

ABEL had thus decomposed the general term <strong>of</strong> the binomial series into the form<br />

mµx µ = λµ<br />

� � µ<br />

cos θµ + i sin θµ α ,<br />

thereby reducing the binomial series to its real and imaginary parts,<br />

φ (x) =1 +<br />

∞<br />

∑<br />

µ=1<br />

λµα µ cos θµ +i<br />

� �� �<br />

=p<br />

∞<br />

∑<br />

µ=1<br />

λµα µ sin θµ . (12.24)<br />

� �� �<br />

=q<br />

Convergence <strong>of</strong> the binomial series. Having obtained the decomposition <strong>of</strong> the bi-<br />

nomial series into real and imaginary parts (12.24), ABEL claimed that it converged if<br />

α < 1 and diverged if α > 1. In order to prove this claim, he applied his version <strong>of</strong> the<br />

ratio test, observing that because<br />

�<br />

� �2 � �<br />

k − µ k ′ 2<br />

δµ+1 =<br />

+ → 1 as µ → ∞,<br />

µ + 1 µ + 1<br />

the ratio <strong>of</strong> consecutive terms converged to α,<br />

λµ+1α µ+1<br />

λµα µ<br />

= δµ+1α → α for µ → ∞.<br />

ABEL took care <strong>of</strong> the trigonometric factors <strong>of</strong> the general terms, cos θµ and sin θµ<br />

by applying his own version <strong>of</strong> the ratio test as expressed in his Lehrsätze I&II. How-<br />

ever, he did not provide any details <strong>of</strong> the argument. In the simplest case, α < 1, the<br />

absolute convergence <strong>of</strong> both the series p and q can be obtained directly from Lehrsatz


256 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

Pro<strong>of</strong> that the trigonometric coefficients cannot approach zero First, we consider<br />

the series<br />

<strong>The</strong> calculation<br />

p − 1 =<br />

∞<br />

∑ λµα<br />

µ=1<br />

µ cos θµ.<br />

cos θµ+1 = cos � �<br />

θµ + φ + γµ+1<br />

= cos φ cos � � � �<br />

θµ + γµ+1 − sin φ sin θµ + γµ+1<br />

= cos φ � �<br />

cos θµ cos γµ+1 − sin θµ sin γµ+1<br />

− sin φ � �<br />

sin θµ cos γµ+1 + cos θµ sin γµ+1<br />

shows that if � �<br />

cos θµ is convergent, T = lim cos θµ, then<br />

�<br />

T = −T cos φ + sin φ 1 − T2 because lim cos γµ = −1 and lim sin γµ = 0.<br />

Consequently, if � �<br />

cos θµ is convergent, its limit has to be given by the equation<br />

�<br />

1 − cos φ<br />

T = ±<br />

,<br />

2<br />

which is only zero if cos φ = 1, i.e. if x is on the real axis.<br />

Similarly, for the series<br />

q =<br />

∞<br />

∑ λµα<br />

µ=1<br />

µ sin θµ,<br />

the situation is completely analogous and the conclusion remains the same.<br />

Box 5: Pro<strong>of</strong> that the trigonometric coefficients cannot approach zero<br />

II, because the trigonometric coefficients never surpass 1, numerically. On the other<br />

hand, if α > 1, the divergence <strong>of</strong> the series p and q rested on the observation that nei-<br />

ther <strong>of</strong> the trigonometric coefficients approached zero. <strong>The</strong> details <strong>of</strong> this observation<br />

which unless x is real are provided in box 5.<br />

In the case α < 1, ABEL proceeded to utilize his previously established theorems.<br />

First, he showed by Lehrsatz VI, that provided φ (m), φ (n), and the Cauchy product<br />

φ (m) φ (n) were all convergent series, the product was equal to φ (m + n). And since<br />

φ (m + n) was assumed to be convergent, ABEL had showed that φ (m) was a solution<br />

to the functional equation.<br />

In order to express everything in real variables, ABEL next introduced<br />

φ (m) = p + qi = r (cos s + i sin s)


12.9. ABEL’s pro<strong>of</strong> <strong>of</strong> the binomial theorem 257<br />

which he wrote as<br />

φ � k + k ′ i � = f � k, k ′� � cos ψ � k, k ′� + i sin ψ � k, k ′�� .<br />

With n = l + l ′ i, ABEL found the analogous <strong>of</strong> the above and proceeded to express<br />

φ (m + n) in the same way to find internal relations <strong>of</strong> f and ψ. He expressed the<br />

functional relation φ (m + n) = φ (m) φ (n) in terms <strong>of</strong> the real arguments<br />

f � k + l, k ′ + l ′� = f � k, k ′� f � l, l ′� , and<br />

ψ � k + l, k ′ + l ′� = 2Mπ + ψ � k, k ′� + ψ � l, l ′� ,<br />

where M denoted an integer. Now, ABEL wanted to find the functions f and ψ which<br />

satisfied these equations. He first proved that f was a continuous function, basically<br />

because it was composed <strong>of</strong> continuous functions. Similarly, he claimed ψ could be<br />

assumed to be continuous by choosing a constant value for M.<br />

ABEL then obtained the equation<br />

ψ � k, k ′ + l ′� + ψ � l, k ′ + l ′� = 2Mπ + ψ � 0, k ′� + ψ � 0, l ′� + ψ � k + l, k ′ + l ′� .<br />

This equation helped him determine the way ψ depended upon its first argument.<br />

ABEL let θ (k) = ψ (k, k ′ + l ′ ) which rendered the equation as<br />

θ (k) + θ (l) = a + θ (k + l) (12.25)<br />

and he proceeded to solve this equation. He did so by first proving directly that for<br />

integer ρ and any k,<br />

In particular, for k = 1, ABEL found the solution<br />

ρθ (k) = (ρ − 1) a + θ (ρk) . (12.26)<br />

θ (ρ) = ρ (θ (1) − a) + a<br />

for integer values <strong>of</strong> ρ. He then proved that this result extended first to rational values<br />

<strong>of</strong> ρ and then to any positive or negative real value <strong>of</strong> ρ by the continuity <strong>of</strong> θ. His<br />

extension to rational values was classical: By (12.26),<br />

� �<br />

µ<br />

ρθ = (ρ − 1) a + θ (µ) = (ρ − 1) a + µ (θ (1) − a) + a, i.e.<br />

ρ<br />

� �<br />

µ<br />

θ = a +<br />

ρ<br />

µ<br />

(θ (1) − a) .<br />

ρ<br />

ABEL next investigated the second argument <strong>of</strong> ψ by similar methods, and he found<br />

ψ � k, k ′� = βk + β ′ k ′ − 2Mπ.


258 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

Having solved the functional equation <strong>of</strong> the angular function ψ, ABEL next re-<br />

duced the functional equation <strong>of</strong> the modular function f to the same equation. He<br />

observed that if<br />

the equation<br />

was reduced to<br />

which he had just solved to find<br />

<strong>The</strong>refore, he found the solution<br />

f � k, k ′� = e F(k,k′ ) ,<br />

f � k + l, k ′ + l ′� = f � k, k ′� f � l, l ′�<br />

F � k + l, k ′ + l ′� = F � k, k ′� + F � l, l ′�<br />

F � k, k ′� = δk + δ ′ k ′ .<br />

φ � k + k ′ i � = e δk+δ′ k ′ � cos � βk + β ′ k ′ � + i sin � βk + β ′ k ′��<br />

and reduced it to its real and imaginary terms. Still, the constants β, β ′ , δ, δ ′ were not<br />

more precisely determined. ABEL returned to this issue and provided formulae for<br />

determining these constants,<br />

α sin φ<br />

β = arctan<br />

1 + α cos φ and<br />

δ = 1<br />

2 log<br />

�<br />

1 + 2α cos φ + α 2�<br />

.<br />

Finally, ABEL employed his Lehrsatz IV to treat the case α = 1 as the limit case <strong>of</strong><br />

the previously considered cases. He summarized the results <strong>of</strong> the entire investigation<br />

in the following way:<br />

<strong>The</strong>orem 14 I. Whenever the series<br />

1 +<br />

m + ni<br />

1<br />

is convergent, it has the sum<br />

�<br />

(1 + a) 2 + b 2� m 2<br />

(a + bi) +<br />

b<br />

−n arctan<br />

e 1+a ×<br />

�<br />

cos<br />

�<br />

m arctan b<br />

1 + a<br />

(m + ni) (m − 1 + ni)<br />

1 · 2<br />

+ i sin<br />

(a + bi) 2 + . . .<br />

�<br />

(1 + a) 2 + b 2��<br />

n<br />

+<br />

2 log<br />

�<br />

m arctan b n<br />

+<br />

1 + a 2 log<br />

�<br />

(1 + a) 2 + b 2���<br />

II. <strong>The</strong> series is convergent for every value <strong>of</strong> m and n whenever the quantity √ a 2 + b 2 is less<br />

than one. If √ a 2 + b 2 is equal to one, the series is convergent for every value <strong>of</strong> m comprised<br />

between −1 and +∞ if one does not simultaneously have α = −1. If α = −1, m must be<br />

positive. In every other case, the series is divergent. 61<br />

✷<br />

61 (N. H. <strong>Abel</strong>, 1826f, 333–334)


12.9. ABEL’s pro<strong>of</strong> <strong>of</strong> the binomial theorem 259<br />

This characterization contained the complete two-part answer to the questions which<br />

ABEL had raised: the sum <strong>of</strong> the binomial series when it is convergent and the condi-<br />

tions <strong>of</strong> its convergence. <strong>The</strong> cumbersome form <strong>of</strong> the sum <strong>of</strong> the binomial series arises<br />

partly from the fact that ABEL expressed its complex variables separated into real and<br />

imaginary parts, and partly from the answer it gives to the problem <strong>of</strong> multivalued<br />

answers: ABEL’S expression for the sum <strong>of</strong> the series only has a single value because<br />

the bracket is a positive number and the extraction <strong>of</strong> roots <strong>of</strong> positive numbers results<br />

in a canonical, positive value.<br />

An example relating to ABEL’S “exception”. At the very end <strong>of</strong> the paper, ABEL<br />

used the results which he had found to carry out the summation <strong>of</strong> certain interesting<br />

series. In particular, the first example is <strong>of</strong> interest in connection with ABEL’S famous<br />

exception.<br />

In the first example, 62 ABEL proposed to sum the series<br />

α sin φ − 1<br />

2 α2 sin 2φ + 1<br />

3 α3 sin 3φ + . . .<br />

which he found was convergent for |α| < 1 where it converged toward the value β<br />

above,<br />

β = arctan<br />

α sin φ<br />

1 + α cos φ =<br />

∞<br />

∑<br />

n=1<br />

(−1) n−1 sin nφ<br />

α<br />

n<br />

n .<br />

To determine the value for α = 1, it sufficed to let α approach the limit 1 provided the<br />

resulting series remained convergent (Lehrsatz IV). Thus, for φ between −π and π,<br />

1<br />

φ = arctan<br />

2<br />

sin φ<br />

1 + cos φ = ∑ (−1)n−1 sin nφ<br />

.<br />

n<br />

For φ = ±π, the situation was different because the series vanished and the expression<br />

for β degenerated. ABEL observed:<br />

“It follows, that the function<br />

sin φ − 1 1<br />

sin 2φ + sin 3φ − . . .<br />

2 3<br />

has the remarkable property <strong>of</strong> being discontinuous for the values φ = π and<br />

φ = −π.” 63<br />

Thus, in this case, ABEL used the same object as in the exception for another purpose.<br />

This time, he wanted to illustrate the same point as in the notebook (see section 12.6):<br />

that although the series <strong>of</strong> the form ∑ vm (x) α m was continuous for α < 1, it needed<br />

not be continuous for α = 1.<br />

62 (ibid., 336–337).<br />

63 “Hieraus folgt, daß die Function:<br />

sin φ − 1 1<br />

sin 2φ + sin 3φ − u.s.w.<br />

2 3<br />

die merkwürdige Eigenschaft hat, für die Werthe φ = π und φ = −π unstetig zu seyn.” (ibid.,<br />

336–337).


260 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

<strong>The</strong> solution to Poisson’s example. <strong>The</strong> ultimate result in ABEL’S binomial paper<br />

concerned the series which had probably inspired him to work on the binomial theo-<br />

rem in the first place. As a result <strong>of</strong> applying his characterization <strong>of</strong> the convergence<br />

<strong>of</strong> the binomial series, ABEL found precise conditions for the convergence <strong>of</strong> the bino-<br />

mial series corresponding to (2 cos x) m . <strong>The</strong> result — an analogy <strong>of</strong> which ABEL also<br />

communicated to HOLMBOE in a letter (see page 226) — was that the identity<br />

(2 cos x) m = cos mx + m<br />

1<br />

cos (m − 2) x + m (m − 1)<br />

1 · 2<br />

cos (m − 4) x + . . .<br />

was valid when m was positive and x belonged to the interval � − π 2 , π 2<br />

� . Thus, ABEL<br />

ruled out validity <strong>of</strong> this formula in the situation <strong>of</strong> Poisson’s example which had in-<br />

volved setting x = π. For general values <strong>of</strong> x, ABEL obtained an identity which in-<br />

volved a correction term,<br />

� (2ρ−1)π<br />

2<br />

(2 cos x) m cos 2ρmπ =<br />

∞<br />

∑<br />

k=0<br />

� �<br />

m<br />

cos (m − 2k) x<br />

k<br />

, (2ρ+1)π<br />

�<br />

2 . Thus, ABEL’S resolution to Poisson’s example consisted <strong>of</strong><br />

for x ∈<br />

two steps deriving from his general pro<strong>of</strong> <strong>of</strong> the binomial theorem. First, he divided<br />

the values <strong>of</strong> x into smaller intervals in which the value <strong>of</strong> cos x had a constant sign.<br />

And second, he considered all expressions as single valued and introduced an addi-<br />

tional term to provide the correction.<br />

12.10 Aspects <strong>of</strong> ABEL’s binomial paper<br />

Having presented and investigated the contents <strong>of</strong> ABEL’S binomial paper, I believe<br />

that three slightly broader aspects <strong>of</strong> it also merit attention: ABEL’S use <strong>of</strong> complex<br />

numbers, his use <strong>of</strong> functional equations, and the style <strong>of</strong> the binomial paper.<br />

12.10.1 ABEL’s understanding <strong>of</strong> complex numbers<br />

Compared with CAUCHY’S Cours d’analyse, ABEL’S pro<strong>of</strong> <strong>of</strong> the binomial theorem ex-<br />

celled by including complex values <strong>of</strong> the exponent. For complex values <strong>of</strong> the ar-<br />

gument x, CAUCHY had reduced the study <strong>of</strong> the binomial series corresponding to<br />

(1 + x) m to the study <strong>of</strong> two real series by writing out the real and imaginary parts.<br />

By diligent use <strong>of</strong> polar representations <strong>of</strong> complex numbers, ABEL succeeded in re-<br />

ducing the functional equations for complex exponents m to the simple, additive one.<br />

Thus, in the binomial paper, ABEL worked with complex numbers which he always<br />

reduced to pairs <strong>of</strong> reals either as real and imaginary parts or in polar representation.<br />

In his inversion <strong>of</strong> elliptic integrals into elliptic functions (see chapter 16, below), ABEL<br />

also worked with complex numbers as arguments <strong>of</strong> functions. Again, complex num-<br />

bers were reduced to real and imaginary parts. From the rather scarce evidence, it


12.10. Aspects <strong>of</strong> ABEL’s binomial paper 261<br />

seems justified to say that ABEL held a strictly algebraic view <strong>of</strong> complex numbers<br />

and that he considered these numbers rather unproblematic.<br />

Because some <strong>of</strong> ABEL’S initial theorems — in particular Lehrsatz IV — dealt with<br />

power series, they have subsequently been interpreted as pertaining to complex vari-<br />

ables. However, there is no absolute indication that this was the interpretation which<br />

ABEL held. ABEL’S original formulations were not very explicit about these issues<br />

and <strong>of</strong>ten neglected taking numerical values into consideration. However, I find very<br />

little reason to suspect that ABEL would develop his theorems for complex variables<br />

and afterwards go through rather cumbersome arguments to reduce complex series to<br />

series <strong>of</strong> real terms.<br />

Going through ABEL’S loans from the Christiania University library, V. BRUN (1885–<br />

1978) discovered that ABEL had — in 1822 — borrowed the volume <strong>of</strong> the transactions<br />

<strong>of</strong> the Danish Academy in which his fellow Norwegian C. WESSEL (1745–1818) had<br />

published his geometric interpretation <strong>of</strong> complex numbers as directed line segments<br />

in the plane. 64 However, as Ø. ORE (1899–1968) also pointed out, ABEL was prob-<br />

ably much more interested in a paper on equations which C. F. DEGEN (1766–1825)<br />

published in the same volume. 65 Even if ABEL read WESSEL’S paper — which seems a<br />

reasonable assumption given the limited amount <strong>of</strong> Danish mathematical literature —<br />

he certainly never did anything to adopt its idea or promote it in any other way. This<br />

just supports K. ANDERSEN’S (⋆1941) hypothesis that the geometrical interpretation<br />

<strong>of</strong> complex numbers was not a hot topic in the first decades <strong>of</strong> the nineteenth century. 66<br />

12.10.2 ABEL on functional equations<br />

As had been the case in both EULER’S and CAUCHY’S pro<strong>of</strong> <strong>of</strong> the binomial theorem,<br />

functional equations played a central part in ABEL’S pro<strong>of</strong>. In his Cours d’analyse,<br />

CAUCHY had developed the topic into a theory <strong>of</strong> its own and he studied multiple<br />

types <strong>of</strong> functional equations. 67<br />

In a paper published in 1827 in CRELLE’S Journal, 68 ABEL presented some results<br />

on functional equations, which when applied to the functional equation<br />

φ (x) + φ (y) = φ (x + y) (12.27)<br />

gave the solution φ (x) = Ax as the unique (continuous) solution to this equation.<br />

It was precisely this functional equation which was central to ABEL’S pro<strong>of</strong> <strong>of</strong> the<br />

binomial theorem. Whereas EULER and CAUCHY had used the equation<br />

f (x) f (y) = f (x + y)<br />

64 (Brun, 1962, 110–111). For an analysis <strong>of</strong> WESSEL’S work, see (K. Andersen, 1999).<br />

65 (C. F. Degen, 1799).<br />

66 (K. Andersen, 1999, 94).<br />

67 (J. Dhombres, 1992).<br />

68 (N. H. <strong>Abel</strong>, 1827c).


262 Chapter 12. ABEL’s reading <strong>of</strong> CAUCHY’s new rigor and the binomial theorem<br />

as the foundation for their pro<strong>of</strong>s, ABEL chose to focus on the other equation (12.27)<br />

because it was better suited for his investigations <strong>of</strong> complex exponents. When he had<br />

to address the multiplicative functional equation expressing the modulus <strong>of</strong> the series,<br />

he transformed it into the form (12.27) by way <strong>of</strong> exponentiation.<br />

<strong>The</strong> additive functional equation (12.27) had also been studied by CAUCHY in his<br />

Cours d’analyse, and it can be interpreted as testimony to ABEL’S familiarity with that<br />

book that he was able to replace the basic tool <strong>of</strong> the previous pro<strong>of</strong>s with one more<br />

suited for his slightly more general situation.<br />

Two general problems concerning functional equations. As mentioned, ABEL pub-<br />

lished two papers in 1826 and 1827 in which he addressed questions concerning func-<br />

tional equations. <strong>The</strong> problems which he attacked were within the immediate scope<br />

<strong>of</strong> CAUCHY’S approach to the theory although they may appear a little odd. Thus,<br />

ABEL’S results can be seen as contributing to the early growth <strong>of</strong> the theory <strong>of</strong> func-<br />

tional equations.<br />

In paper published in 1826, 69 ABEL dealt with functions f such that f (z, f (x, y))<br />

was symmetric in x, y, z. For such functions, ABEL obtained the characterization:<br />

“Whenever a function f (x, y) <strong>of</strong> two independent variable quantities x and<br />

y has the property that f (z, f (x, y)) is a symmetric function <strong>of</strong> x, y, z, there will<br />

always be a function ψ for which<br />

ψ ( f (x, y)) = ψx + ψy.” 70<br />

Furthermore, ABEL found that the stipulated function ψ could be determined by the<br />

differential equation<br />

ψ (x) = ψ ′ � ∂ f<br />

∂x (x, y)<br />

(y)<br />

dx.<br />

∂ f<br />

∂y (x, y)<br />

In the process, ABEL also integrated the equation<br />

to find<br />

∂r (x, y)<br />

φ (y) =<br />

∂x<br />

��<br />

r = ψ<br />

�<br />

φ (x) dx +<br />

∂r (x, y)<br />

φ (x)<br />

∂y<br />

�<br />

φ (y) dy ,<br />

a result which will resurface in section 16.2.2 where ABEL’S deduction <strong>of</strong> the addition<br />

theorems <strong>of</strong> elliptic functions are described.<br />

69 (N. H. <strong>Abel</strong>, 1826e).<br />

70 “Sobald eine Function f (x, y) zweier unabhängig veränderlichen Größen x und y die Eigenschaft<br />

hat, daß f (z, f (x, y)) eine symmetrische Function von x, y und z ist, so muß es allemal eine Function<br />

ψ geben, für welche<br />

ist.” (ibid., 13).<br />

ψ f (x, y) = ψ (x) + ψ (y)


12.10. Aspects <strong>of</strong> ABEL’s binomial paper 263<br />

<strong>The</strong> following year, ABEL treated another problem concerning the form <strong>of</strong> func-<br />

tions satisfying a functional relation. 71 In that paper, he studied the form <strong>of</strong> functions<br />

φ which satisfy<br />

φ (x) + φ (y) = ψ (x f (y) + y f (x)) . (12.28)<br />

ABEL found his main result which stated that the most general way <strong>of</strong> satisfying the<br />

equation (12.28) was with<br />

φ (x) = φ ′ �<br />

(0) f (0)<br />

dx<br />

f (x) f ′ � �<br />

x<br />

and ψ (x) = φ (0) + φ .<br />

(0) x f (0)<br />

One very simple application <strong>of</strong> this result is particularly interesting and a reconstruction<br />

<strong>of</strong> it is given below. If we let f (x) = 1, we obtain α = f (0) = 1 and α ′ = f ′ (0) =<br />

0. <strong>The</strong>refore, the function φ which satisfies the equation (ψ = φ)<br />

φ (x) + φ (y) = φ (x + y) (12.29)<br />

must satisfy the requirement<br />

�<br />

dx<br />

φ (x) = φ (0) = xφ (0) + C,<br />

1<br />

where C = 0 by (12.29). Consequently, ABEL’S result leads to the result that the only<br />

(continuous) solution to the functional equation<br />

φ (x) + φ (y) = φ (x + y)<br />

is the linear function φ (x) = Ax for some constant A. As observed, this paper —<br />

which was published in 1827 — therefore contains a generalization <strong>of</strong> the additive<br />

functional equation (12.27) which had been so important to his pro<strong>of</strong> <strong>of</strong> the binomial<br />

theorem. However, in the binomial paper, ABEL probably relied directly on CAUCHY’S<br />

Cours d’analyse.<br />

12.10.3 Concepts and calculations in the binomial paper<br />

ABEL’S paper on the binomial series shows a remarkable blend <strong>of</strong> concepts and ex-<br />

plicit calculations. A superficial, textual analysis <strong>of</strong> ABEL’S paper reveals a division<br />

<strong>of</strong> the paper: First, six preliminary theorems were presented which were applicable<br />

to classes <strong>of</strong> series. <strong>The</strong>se were cast in a strictly Euclidean presentational style with<br />

definitions <strong>of</strong> convergence and continuity, statements <strong>of</strong> the six theorems and pro<strong>of</strong>s<br />

following each theorem. Second, detailed and explicit considerations <strong>of</strong> the conver-<br />

gence <strong>of</strong> the binomial series as well as formulae for its sum were given. <strong>The</strong>se in-<br />

vestigations relied extensively on explicit manipulations <strong>of</strong> the formulae in forms as<br />

described above. When analyzed from this perspective, ABEL’S binomial paper shows<br />

traits <strong>of</strong> both concept based and formula based mathematics as discussed in chapter<br />

21, below. Thus, the binomial paper is an example <strong>of</strong> the transitional status <strong>of</strong> ABEL’S<br />

mathematics which exhibited similarities with both paradigms.<br />

71 (N. H. <strong>Abel</strong>, 1827c).


Chapter 13<br />

ABEL and OLIVIER on convergence<br />

tests<br />

Besides his pro<strong>of</strong> <strong>of</strong> the binomial theorem, N. H. ABEL’S (1802–1829) only other pub-<br />

lication on analysis — in which his main interest was the theory <strong>of</strong> infinite series — is a<br />

curious little argument against another mathematician named L. OLIVIER. 1 In the first<br />

issue <strong>of</strong> the second volume <strong>of</strong> the Journal für die reine und angewandte Mathematik, A. L.<br />

CRELLE (1780–1855) had accepted a paper by OLIVIER in which the latter claimed to<br />

have obtained a general — yet very simple — test <strong>of</strong> convergence. Despite OLIVIER’S<br />

claims, the test was not generally applicable and in the next (i.e. third) volume <strong>of</strong> the<br />

Journal, ABEL published his refutation which consisted <strong>of</strong> a counter example to the<br />

original claim made by OLIVIER as well as a pro<strong>of</strong> that no such criterion could ever be<br />

found: it was an utopian dream. 2<br />

13.1 OLIVIER’s theorem<br />

In the article entitled Remarques sur les séries infinies et leur convergence, 3 OLIVIER worked<br />

with a distinction between convergent, indeterminate, and divergent series (see below)<br />

and stated a criterion to distinguish convergent and non-convergent series. This cri-<br />

terion, which OLIVIER called a “criterion <strong>of</strong> convergence <strong>of</strong> infinite series” 4 was the<br />

following:<br />

“Thus, if one finds that in an infinite series the product <strong>of</strong> the n th term – or the<br />

n th group <strong>of</strong> terms which keep the same sign — by n is zero for n = ∞, one can<br />

regard this single circumstance as a sign that the series is convergent. Reciprocally,<br />

the series cannot be convergent unless the product n · an is zero for n = ∞.” 5<br />

1 Very little is known about LOUIS OLIVIER. However, based on his pattern <strong>of</strong> publication, I believe<br />

that he was a non-pr<strong>of</strong>essional mathematician with ties to Berlin. I hope to be able to present my<br />

analyses <strong>of</strong> OLIVIER’S mathematical production in the near future.<br />

2 For an analysis <strong>of</strong> OLIVIER’S criterion and ABEL’S response, see also (I. Grattan-Guinness, 1970b,<br />

139–143) and (Goar, 1999).<br />

3 (Olivier, 1827).<br />

4 “criterium de la convergence des séries infinies” (ibid., 34).<br />

265


266 Chapter 13. ABEL and OLIVIER on convergence tests<br />

OLIVIER’S curious inserted remark — that the n th group <strong>of</strong> terms with the same sign<br />

could be considered instead <strong>of</strong> an — probably derived from the fact that within such a<br />

group, reordering <strong>of</strong> the terms could not effect the convergence or sum <strong>of</strong> the series.<br />

However, both in OLIVIER’S paper and in the present analysis, the important case<br />

arises by considering individual terms.<br />

For the historian, OLIVIER’S theorem requires an interpretation which is by no<br />

means easy or unambiguous; for instance, what does it mean that “this circumstance<br />

is a sign”? <strong>The</strong> main question in interpreting the theorem lies in this phrase, since<br />

it could be read to mean that if nan → 0 then the series will always be convergent.<br />

This was certainly the way it was interpreted by some <strong>of</strong> OLIVIER’S readers; however,<br />

the phrasing is sufficiently weak to call for further investigation. In the following,<br />

OLIVIER’S argument is outlined in order to illustrate how mathematicians in the early<br />

19 th century still argued about infinite series. <strong>The</strong>n, to supplement the theorem and<br />

its pro<strong>of</strong>, a consideration <strong>of</strong> the examples to which it was applied is necessary before<br />

a weighed interpretation <strong>of</strong> the theorem can be given.<br />

13.1.1 OLIVIER’s first pro<strong>of</strong><br />

OLIVIER gave two arguments which illustrate how he came to believe in his theorem.<br />

<strong>The</strong> first argument was given immediately before the theorem was stated, whereas<br />

the second one was prompted by ABEL’S objection to the theorem and printed as a<br />

response to ABEL’S note. 6<br />

In 1827, OLIVIER divided infinite series into three categories: convergent, indeter-<br />

minate, and divergent. His definitions and the ensuing pro<strong>of</strong>s are difficult to represent<br />

fairly, because his concepts are different and vague and his style <strong>of</strong> reasoning is rather<br />

verbal and leaves few hints on the unclear points. OLIVIER’S definition <strong>of</strong> convergent<br />

series consisted <strong>of</strong> two requirements:<br />

“One calls a series convergent which has the following two properties, namely:<br />

that one finds its numerical value ever more exactly when one calculates successively<br />

more terms and that by continuing the calculation indefinitely, one can approach<br />

the true value <strong>of</strong> the entire series to any degree one wishes.”” 7<br />

In OLIVIER’S definition, we see a curious and obscure mixture <strong>of</strong> the old and the<br />

new concepts <strong>of</strong> convergence. At the same time, OLIVIER speaks <strong>of</strong> numerical approx-<br />

5 “Donc si l’on trouve, que dans une série infinie, le produit du n me terme, ou du n me des groupes<br />

de termes qui conservent le même signe, par n, est zéro, pour n = ∞, on peut regarder cette seule<br />

circonstance comme une marque, que la série est convergente; et réciproquement, la série ne peut<br />

pas être convergente, si le produit n.an n’est pas nul pour n = ∞.” (Olivier, 1827, 34).<br />

6 (Olivier, 1828)<br />

7 “On appelle convergente une série, qui a les deux propriétés suivantes, savoir: qu’on trouve sa valeur<br />

numérique d’autant plus exactement, qu’on calcule successivement plusieurs termes, et qu’en<br />

continuant indéfiniment ce calcul, on peut se rapprocher de la vraie valeur de la série totale à tel<br />

degré qu’on voudra.” (Olivier, 1827, 31).


13.1. OLIVIER’s theorem 267<br />

imation in language similar to A.-L. CAUCHY’S (1789–1857) and <strong>of</strong> the “true value”<br />

<strong>of</strong> the series which resembles the Eulerian formal equality between functions.<br />

OLIVIER separated non-convergent series into indeterminate and divergent ones:<br />

“On the contrary, one calls a series indeterminate if continuing the calculation<br />

<strong>of</strong> terms does not make it approach anything.<br />

And one calls a series divergent in which the successive terms, added together,<br />

produces results which differ more and more from the true value <strong>of</strong> the series.” 8<br />

This distinction between two types on non-convergent series was probably inspired by<br />

the discussion <strong>of</strong> Poisson’s example in which OLIVIER also participated without making<br />

any noticeable contributions. 9<br />

OLIVIER proceeded to express the two criteria <strong>of</strong> his definition <strong>of</strong> convergence in a<br />

slightly different form. First, he observed, the terms <strong>of</strong> the series (or groups <strong>of</strong> terms<br />

with the same sign) had to constantly decrease. Second, the sum <strong>of</strong> terms after the n th<br />

term, i.e. the tail <strong>of</strong> the series, had to be zero for n = ∞. He gave similar translations<br />

<strong>of</strong> the concepts <strong>of</strong> indeterminate and divergent series.<br />

To obtain his theorem, OLIVIER first investigated the first condition concerning the<br />

vanishing <strong>of</strong> the terms. He stated that this condition would always be satisfied if the<br />

ratio <strong>of</strong> consecutive terms was always less than one. Thus, he apparently missed out<br />

on cases in which an+1 an < 1 but lim an+1 an<br />

= 1. For the second condition to also be<br />

fulfilled, OLIVIER noted that it would be necessary and sufficient that nan → 0 for<br />

n → ∞.<br />

Under hypothesis nan → 0 and the further assumption that the terms an vanish as<br />

n increases, OLIVIER claimed that<br />

if the series was written as<br />

R ≤ nan<br />

a1 + a2 + · · · + an + R.<br />

And thus, the vanishing <strong>of</strong> the tail R followed. How OLIVIER came to this [false] belief<br />

will be clearer below.<br />

On the other hand, the tail R could not vanish without nan also vanishing, OLIVIER<br />

claimed. Using the constantly decreasing nature <strong>of</strong> the terms, OLIVIER found<br />

na2n ≤ R ≤ nan<br />

where R suddenly meant “the sum <strong>of</strong> n terms which follows after the n th term.” 10<br />

Consequently, if nan vanished, so did R and the convergence <strong>of</strong> the series was secured.<br />

8 “Au contraire, on appelle indéterminée une série, qui ne donne aucun rapprochement, en continuant<br />

le calcul des termes.<br />

Et on appelle divergente une série, dont les termes suivants, ajoutés aux précédents, ne donnent<br />

que des résultats, qui s’éloignent plus en plus de vraie valeur de la série.” (ibid., 31).<br />

9 (Olivier, 1826b). For a contemporary evaluation, see ([Saigey], 1826, 112).<br />

10 “[. . . ] R, ou la somme des n termes qui suivent le n me terme.” (Olivier, 1827, 34).


268 Chapter 13. ABEL and OLIVIER on convergence tests<br />

Figure 13.1: OLIVIER’S geometrical argument<br />

Thus, OLIVIER’S pro<strong>of</strong> rested on some uncertain principles. Firstly, OLIVIER’S di-<br />

vision <strong>of</strong> series into three classes and his definition <strong>of</strong> convergent series was less sharp<br />

and less useful than other existing concepts. Secondly, he assumed that in any (nu-<br />

merically) convergent series, the general terms were monotonically decreasing. As<br />

indicated, this assumption was central to his pro<strong>of</strong>. Finally, OLIVIER switched from<br />

considering R as the tail <strong>of</strong> the series having infinitely many terms to cutting it <strong>of</strong>f after<br />

n terms. This transition between infinite and finite objects represented limit processes<br />

which were never spelled out.<br />

13.1.2 OLIVIER’s second pro<strong>of</strong><br />

Reacting to ABEL’S criticism (see below), OLIVIER gave a short indication <strong>of</strong> the in-<br />

tuition behind his original pro<strong>of</strong>. 11 <strong>The</strong>re, he showed by way <strong>of</strong> a geometrical figure<br />

(see figure 13.1) how he had reasoned. OLIVIER led Bb denote the term an and ob-<br />

served that the area <strong>of</strong> the parallelogram Bbuv which represents the value nan would<br />

be greater than the smaller parallelograms, e.g. Bc, Cd, etc. <strong>The</strong> inequality was ob-<br />

tained by the constant decreasing <strong>of</strong> the terms an. <strong>The</strong> equality between the parallelo-<br />

gram Bbvu and nan represented OLIVIER’S different uses <strong>of</strong> infinite values for n. Thus,<br />

the above interpretation <strong>of</strong> OLIVIER’S first pro<strong>of</strong> seems to be confirmed.<br />

OLIVIER reacted explicitly to ABEL’S criticism by observing that although his the-<br />

orem seemed to be well founded and the deduced examples were correct, ABEL had<br />

nevertheless observed that it was not “generally applicable”. OLIVIER saw ABEL’S<br />

criticism as an indication <strong>of</strong> the care which should be observed when dealing with<br />

infinite quantities and locate the mistake to his working indifferently with finite and<br />

infinite quantities.<br />

Finally, OLIVIER revised his theorem in order to make it correct by substituting the<br />

11 (Olivier, 1828).


13.2. ABEL’s counter example 269<br />

assumption<br />

n<br />

� �<br />

∞<br />

∑ amn → 0<br />

m=1<br />

for the original nan → 0. Here, we see how OLIVIER approached the convergence <strong>of</strong><br />

Cauchy sequences.<br />

13.1.3 <strong>The</strong> application to examples<br />

When OLIVIER came to apply his theorem to particular series, he did so as a complete<br />

test <strong>of</strong> convergence, i.e. as the bi-implication<br />

nan → 0 ⇔ ∑ an convergent.<br />

<strong>The</strong>re are, however, no definite tests <strong>of</strong> this interpretation, because OLIVIER did not<br />

apply his criterion to any divergent series for which nan → 0.<br />

One <strong>of</strong> the most interesting examples, which OLIVIER did treat, was the binomial<br />

theorem. After long manipulations <strong>of</strong> the binomial coefficients, OLIVIER used his the-<br />

orem to state that the binomial series<br />

(1 + c) m m (m − 1)<br />

= 1 + mc + c<br />

1 · 2<br />

2 + . . .<br />

was convergent for any exponent if c < 1 and for m ≥ 0 if c = 1. <strong>The</strong> binomial series<br />

was divergent for c > 1 or for c = 1 if m < 0. In this way, OLIVIER obtained the<br />

convergence <strong>of</strong> the binomial series for real arguments and exponents.<br />

13.2 ABEL’s counter example<br />

Soon after the publication <strong>of</strong> OLIVIER’S paper, ABEL responded with a short note in<br />

the Journal. 12 <strong>The</strong>re, ABEL commented on OLIVIER’S theorem in the following words,<br />

“<strong>The</strong> latter part <strong>of</strong> this theorem is very true but the first [part] does not seem<br />

to be so. For example, the series<br />

1 1 1<br />

1<br />

+ + + · · · + + . . .<br />

2 log 2 3 log 3 4 log 4 n log n<br />

is divergent although nan = 1<br />

log n is zero for n = ∞.”13<br />

Here, ABEL politely suggested that the first part <strong>of</strong> OLIVIER’S theorem “did not<br />

seem to be true”. In one <strong>of</strong> his notebooks, ABEL’S draft for the paper can be found;<br />

there he was more dramatic, remarking “Thus, Mr. Olivier is seriously mistaken.” 14 In<br />

12 (N. H. <strong>Abel</strong>, 1828a).<br />

13 “La dernière partie de ce théorème est très juste, mais la première ne semble pas l’être. Par exemple<br />

la série<br />

1 1 1<br />

1<br />

+ + + · · · + + . . .<br />

2 log 2 3 log 3 4 log 4 n log n<br />

est divergente, quoique nan = 1<br />

log n soit zéro pour n = ∞.” (ibid., 79).<br />

14 “Donc M. Olivier s’est trompé sérieusement.” (N. H. <strong>Abel</strong>, [1827] 1881, II, 199).


270 Chapter 13. ABEL and OLIVIER on convergence tests<br />

section 21.3, ABEL’S comments on OLIVIER’S theorem will serve as one among a class<br />

<strong>of</strong> cases where counter examples were employed for different ends and with differing<br />

confidence in the early nineteenth century.<br />

Central to ABEL’S pro<strong>of</strong> was the inequality<br />

log (1 + x) < x (13.1)<br />

which he claimed was valid for all positive x. For x ≥ 1, it was obvious to ABEL and<br />

he gave no argument. It can be easily obtained by observing that<br />

x − log (1 + x)<br />

is increasing for x ≥ 1 and positive for x = 1. For x < 1, ABEL gave an argument<br />

employing the expansion <strong>of</strong> the logarithm into power series as<br />

log (1 + x) =<br />

∞<br />

∑<br />

n=1<br />

(−1) n−1<br />

n<br />

x n = x −<br />

∞<br />

∑<br />

n=1<br />

�<br />

1 x<br />

−<br />

2n 2n + 1<br />

�<br />

x 2n .<br />

From this, he observed that the parentheses were always positive which produced the<br />

desired inequality. Here, ABEL thus rearranged the terms <strong>of</strong> the logarithmic series<br />

without further ado. 15<br />

ABEL employed the inequality (13.1) for x = 1 n<br />

1<br />

n<br />

or written differently<br />

> log<br />

�<br />

1 + 1<br />

�<br />

n<br />

= log<br />

log (1 + n)<br />

log n<br />

<<br />

n + 1<br />

n<br />

to produce<br />

= log (n + 1) − log n,<br />

�<br />

1 + 1<br />

�<br />

.<br />

n log n<br />

Taking logarithms and using the inequality (13.1) again, ABEL obtained<br />

� � �<br />

log (1 + n)<br />

log log (1 + n) − log log n = log<br />

< log 1 +<br />

log n<br />

1<br />

�<br />

<<br />

n log n<br />

ABEL had thus produced the inequality<br />

which when summed from 2 to n gave<br />

log log (1 + n) < log log n + 1<br />

n log n ,<br />

log log (1 + n) < log log 2 +<br />

n<br />

∑<br />

k=2<br />

1<br />

k log k .<br />

1<br />

n log n .<br />

Since the left hand side obviously became infinite for n = ∞, the series on the right<br />

→ 0.<br />

hand side was divergent contradicting OLIVIER’S theorem since nan = 1<br />

log n<br />

ABEL’S conclusion was again remarkably reserved and apparently underplayed,<br />

“<strong>The</strong> theorem announced in the above citation is thus at fault in this case.” 16<br />

15 For more on the history <strong>of</strong> absolute convergence, see section 12.7.<br />

16 “Le théorème énoncé dans l’endroit cité est donc en défaut dans ce cas.” (N. H. <strong>Abel</strong>, 1828a, (400)).


13.3. ABEL’s general refutation 271<br />

13.3 ABEL’s general refutation<br />

After he had given his counter example to OLIVIER’S theorem, ABEL might have been<br />

expected to leave the matter. However, he had more to say on the issue. Creating some<br />

procedures to obtain from one divergent series another one which diverged much<br />

slower, ABEL could prove that the quest which OLIVIER had undertaken was bound<br />

to result in frustration.<br />

ABEL observed that if the series<br />

∞<br />

∑ an<br />

n=0<br />

was divergent, then so was the series where each term had been divided by the re-<br />

spective partial sums,<br />

∞<br />

an<br />

∑ sn n=1<br />

=<br />

∞<br />

∑<br />

n=1<br />

an<br />

∑ n−1<br />

k=0 an<br />

.<br />

<strong>The</strong> pro<strong>of</strong> was easily obtained from arguments resembling the pro<strong>of</strong> above. ABEL<br />

observed for n ≥ 1, inequality (13.1) produced<br />

and therefore, by summation,<br />

log sn − log sn−1 < an−1<br />

,<br />

sn−1<br />

log sn − log a0 <<br />

Since sn → ∞, the left hand side diverged, which implied the divergence <strong>of</strong> the right<br />

n<br />

∑<br />

k=1<br />

hand side. We may express this result in modern language as lemma 2.<br />

Lemma 2 If ∑ ∞ n=0 an is a divergent series <strong>of</strong> positive terms, then the series defined as<br />

∞<br />

an<br />

∑ sn n=1<br />

will be divergent as well. ✷<br />

Now, in order to generally refute the theorem proposed by OLIVIER, ABEL as-<br />

sumed that a function φ taking integer arguments existed such that the series ∑ an<br />

was convergent if and only if φ (n) an → 0 as n → ∞. <strong>The</strong> series obtained as<br />

∞<br />

∑<br />

n=1<br />

1<br />

φ (n)<br />

ak .<br />

sk (13.2)<br />

would then produce a general counter example to this generalized theorem. <strong>The</strong> series<br />

(13.2) was divergent by the criterion, because φ (n) an = 1. On the other hand, when<br />

the procedure <strong>of</strong> obtaining a derived divergent series was applied to (13.2), a new<br />

series was obtained<br />

∞<br />

∑<br />

n=2<br />

1<br />

φ (n) ∑ n−1<br />

k=1<br />

1<br />

φ(k)<br />

. (13.3)


272 Chapter 13. ABEL and OLIVIER on convergence tests<br />

Since the series (13.2) was divergent, the series (13.3) would have to be divergent as<br />

well (by the procedure above). On the other hand, the generalized criterion, when<br />

applied to (13.3) gave<br />

φ (n) an =<br />

1<br />

∑ n−1<br />

k=1<br />

,<br />

1<br />

φ(k)<br />

which by the very divergence <strong>of</strong> (13.2) converged to zero for n → ∞. Thus, the se-<br />

ries (13.3) produced a general counter example to ABEL’S generalization <strong>of</strong> OLIVIER’S<br />

proposed convergence criterion.<br />

By this very elegant pro<strong>of</strong>, ABEL turned OLIVIER’S proposed criterion against itself<br />

and it imploded. Thus, ABEL proved that no simple test <strong>of</strong> convergence <strong>of</strong> series could<br />

be devised. Interpreted as a question <strong>of</strong> delineation <strong>of</strong> concepts, ABEL’S result thus<br />

meant that the extent <strong>of</strong> the concept <strong>of</strong> convergent series was not easily determined by<br />

external criteria.<br />

13.4 More characterizations and tests <strong>of</strong> convergence<br />

In its published form, ABEL’S answer to OLIVIER’S paper was a negative one, in the<br />

sense that it refused a proposed theorem. However, in his notebooks, ABEL elabo-<br />

rated some <strong>of</strong> the ideas found therein to such a degree as to produce new, positive<br />

knowledge in the form <strong>of</strong> new characterizations and tests <strong>of</strong> convergence. 17<br />

In his notebook draft, ABEL obtained his own version <strong>of</strong> a limit comparison theorem<br />

which provided a necessary criterion for convergence. He claimed that if ∑ φ (n)<br />

was a divergent series, and ∑ an was a convergent one, it would be necessary that<br />

“the smallest among the limits <strong>of</strong><br />

an<br />

φ(n) be zero.”18 ABEL’S pro<strong>of</strong> was indirect: Under<br />

the contrary assumption, he wrote un = pnφ (n) where pn ≥ α. <strong>The</strong>n<br />

∑ un > ∑ αφ (n) = α ∑ φ (n) → ∞.<br />

From this, ABEL obtained the second part <strong>of</strong> OLIVIER’S theorem which he had not<br />

objected to: Because ∑ 1 n was known to be divergent, if ∑ an was to be convergent, it<br />

would be necessary that nan vanished as n became infinite.<br />

Pairs <strong>of</strong> convergent and divergent series. Also in the notebook, we find a general-<br />

ization <strong>of</strong> the lemma 2 to the effect that the divergence <strong>of</strong> ∑ an implied the divergence<br />

<strong>of</strong> the series ∑ an<br />

s α n<br />

where 0 ≤ α ≤ 1 (lemma 2 results from setting α = 1). A converse to<br />

this result was also obtained when ABEL proved that if the series ∑ an was divergent,<br />

then the series<br />

17 (N. H. <strong>Abel</strong>, [1827] 1881).<br />

18 (N. H. <strong>Abel</strong>, 1881, II, 198).<br />

∞<br />

an<br />

∑<br />

n=1 s 1+α<br />

n


13.4. More characterizations and tests <strong>of</strong> convergence 273<br />

would be convergent if α > 0. Thus, from a divergent series, ABEL had prescribed<br />

means <strong>of</strong> obtaining two derived series, one <strong>of</strong> which was divergent, the other conver-<br />

gent.<br />

In another section <strong>of</strong> the note, ABEL devised another way <strong>of</strong> obtaining a divergent<br />

series, which would lead him to a new test <strong>of</strong> convergence. ABEL found that for any<br />

continuous function φ (n) which increased without bounds for n → ∞, the series <strong>of</strong><br />

derived terms,<br />

would be divergent.<br />

∞<br />

∑ φ<br />

n=1<br />

′ (n) , (13.4)<br />

ABEL’S pro<strong>of</strong> proceeded from the Taylor series expansion <strong>of</strong> φ (to the second term<br />

and with remainder),<br />

φ (n + 1) = φ (n) + φ ′ (n) + φ′′ (n + θ)<br />

, for some 0 < θ < 1.<br />

2<br />

At this point, ABEL’S draft style made the precise assumptions <strong>of</strong> the ensuing de-<br />

ductions difficult to interpret. However, if ABEL’S requirements interpreted to mean<br />

φ ′′ (n) < 0, we obtain what was his next line,<br />

φ (n + 1) − φ (n) < φ ′ (n) .<br />

<strong>The</strong>n, the divergence <strong>of</strong> (13.4) followed by summation,<br />

φ ′ (n) > φ (n + 1) − φ (0) → ∞ as n → ∞.<br />

Subsequently, ABEL applied this procedure to prove the divergence <strong>of</strong> the series<br />

∞<br />

∑<br />

n=2<br />

1<br />

n ∏ m k=1 logk for m integral,<br />

n<br />

where log k n = log log k−1 n. ABEL did so by defining<br />

and differentiating it to obtain<br />

φ ′ m (n) =<br />

φm (n) = log m (n + a)<br />

1<br />

(n + a) ∏ m−1<br />

k=1 logk (n + a) .<br />

As a consequence <strong>of</strong> the theorem stated above, the series (corresponding to a = 0)<br />

was divergent.<br />

∞<br />

∑ φ<br />

n=2<br />

′ m (n) =<br />

∞<br />

∑<br />

n=2<br />

1<br />

n ∏ m−1<br />

k=1 logk n


274 Chapter 13. ABEL and OLIVIER on convergence tests<br />

On the other hand, ABEL next turned to a function intimately related to the one<br />

studied above, 19<br />

ψ (n) = φm (n) 1−α<br />

.<br />

1 − α<br />

This time, ABEL’S calculations produced the inequality<br />

ψ (n + 1) − ψ (n) > ψ ′ (n + 1)<br />

corresponding to the fact that ψ ′ was a decreasing function. Consequently, through a<br />

number <strong>of</strong> calculations, ABEL was led to a series<br />

∞<br />

∑<br />

n=1<br />

1<br />

n log m (n) α+1 ∏ m−1<br />

k=1 logk n<br />

which was convergent if α > 0 and another one (corresponding to α = −1)<br />

which was divergent.<br />

∞<br />

∑<br />

n=1<br />

1<br />

n ∏ m−1<br />

k=1 logk n<br />

A logarithmic test <strong>of</strong> convergence. <strong>The</strong>se methods <strong>of</strong> constructing convergent and<br />

divergent series led ABEL to a new test <strong>of</strong> convergence. <strong>The</strong> underlying idea <strong>of</strong> ABEL’S<br />

argument starts from the two series, one convergent and the other divergent, and<br />

compares a given series with these two typical ones. He found by simple arguments<br />

based on the results above, that if<br />

�<br />

log<br />

lim<br />

log m+1 n<br />

1<br />

unn ∏ m−1<br />

k=1 logk n<br />

�<br />

> 1, (13.5)<br />

the series ∑ un was convergent. ABEL’S criterion also indicated, that if the limit in<br />

(13.5) was < 1, the series ∑ un would be divergent. In its polished form, ABEL’S<br />

criterion thus became the following:<br />

<strong>The</strong>orem 15 For a series <strong>of</strong> positive terms ∑ un, the limit<br />

�<br />

1 log un<br />

k = lim<br />

n→∞<br />

d<br />

dn logm �<br />

n<br />

log m+1 n<br />

is considered. If k > 1, the series will be convergent; if k < 1, it will be divergent; and if k = 1,<br />

nothing can be said <strong>of</strong> the convergence or divergence <strong>of</strong> the series by this test. ✷<br />

This result was later rediscovered by J. L. F. BERTRAND (1822–1900). 20<br />

19 ABEL actually also denoted this function by φ, but to avoid confusion, I have chosen to label it ψ.<br />

20 (Bertrand, 1842).


13.4. More characterizations and tests <strong>of</strong> convergence 275<br />

ABEL’S continued interest in the theory <strong>of</strong> series and — in particular — in obtain-<br />

ing new tests <strong>of</strong> convergence for series are indications <strong>of</strong> a continued interest in this<br />

topic. Due to CAUCHY’S re-founding <strong>of</strong> analysis, tests <strong>of</strong> convergence were becoming<br />

increasingly important, and a number <strong>of</strong> new tests were discovered in the nineteenth<br />

century. 21<br />

21 See e.g. (I. Grattan-Guinness, 1970b, 131–151).


Chapter 14<br />

Reception <strong>of</strong> ABEL’s contribution to<br />

rigorization<br />

In the course <strong>of</strong> the nineteenth century, analysis underwent an elaborate program <strong>of</strong><br />

rigorization which effected both the techniques, results, and questions <strong>of</strong> the dis-<br />

cipline. Basic notions such as real numbers, continuous functions, integrals, and<br />

trigonometric functions were revised and deeply changed as reflections <strong>of</strong> a funda-<br />

mental transition in the ways mathematicians thought about their subject. <strong>The</strong> chang-<br />

ing concepts and attitudes have been studied intensively by historians. 1 In the twen-<br />

tieth century, N. H. ABEL’S (1802–1829) critical attitude and his part in the revision <strong>of</strong><br />

analysis have received some interest but in the first decades after his death, these were<br />

not issues which attracted the most interest to his mathematics. <strong>The</strong> following section<br />

briefly discusses the reception <strong>of</strong> ABEL’S work on rigorization and certain aspects <strong>of</strong><br />

the subsequent development.<br />

14.1 Reception <strong>of</strong> ABEL’s rigorization<br />

Because <strong>of</strong> the overwhelming development <strong>of</strong> analysis in the nineteenth century and<br />

ABEL’S apparently limited direct impact, the reception <strong>of</strong> ABEL’S contribution to the<br />

rigorization movement is only briefly described from two different perspectives. 2<br />

14.1.1 Binomial theorem<br />

<strong>The</strong> most immediate reaction to ABEL’S binomial paper was actually a non-reaction.<br />

In 1829 and 1830, A. L. CRELLE (1780–1855) published two papers on the binomial theorem<br />

demonstrating that the subject had not been closed by ABEL’S paper <strong>of</strong> 1826. 3<br />

CRELLE’S pro<strong>of</strong>s were based on his previous research on the so-called analytical facul-<br />

1 See e.g. (Bottazzini, 1986; Hawkins, 1970).<br />

2 I hope to subsequently substantiate the analysis <strong>of</strong> the reception <strong>of</strong> this part <strong>of</strong> ABEL’S research<br />

through detailed studies <strong>of</strong> the works <strong>of</strong> selected, later mathematicians.<br />

3 (A. L. Crelle, 1829a; A. L. Crelle, 1830).<br />

277


278 Chapter 14. Reception <strong>of</strong> ABEL’s contribution to rigorization<br />

ties and were only partially within the Cauchyian approach. CRELLE only considered<br />

real arguments and exponents and divided his research into two parts correspond-<br />

ing to the two papers. First, he showed by formal arguments from the trivial identity<br />

1 = 1 that the binomial and its series had to be identical. <strong>The</strong> argument involved finite<br />

differences which had been such a key component <strong>of</strong> his research within the German<br />

combinatorial school. Second, CRELLE investigated the convergence <strong>of</strong> the binomial<br />

series dividing into separate cases corresponding to various assumptions on a and k<br />

(he wrote his binomial as (1 + a) k ). In each case, CRELLE considered the remainder<br />

terms <strong>of</strong> the series and established conditions <strong>of</strong> convergence or divergence.<br />

Concerning CRELLE’S publications on the binomial theorem, two remarks can be<br />

made. First, the fact that CRELLE published on a particular case <strong>of</strong> the binomial the-<br />

orem (real arguments and exponents) after ABEL’S more general result testifies to the<br />

debate between the Cauchyian program and the German algebraic school. CRELLE<br />

wrote in his introduction that he considered his pro<strong>of</strong> to fulfill all requirements in-<br />

cluding being truly rigorous and general and simultaneously clear and elementary.<br />

<strong>The</strong>se positive attributes were obtained through the use <strong>of</strong> algebraic manipulations. 4<br />

Second, CRELLE did not initially consider or even mention the necessity <strong>of</strong> conver-<br />

gence <strong>of</strong> the binomial series. ABEL’S critical attitude may have provoked CRELLE to<br />

take up the issue in the second paper. Thus, at least in Germany, A.-L. CAUCHY’S<br />

(1789–1857) new program <strong>of</strong> numerical equality was not immediately accepted — not<br />

even in CRELLE’S Journal after ABEL’S publication and the translation <strong>of</strong> the Cours<br />

d’analyse. 5<br />

Later in the nineteenth and the twentieth century, when the German combinato-<br />

rial school eventually lost ground, ABEL’S pro<strong>of</strong> <strong>of</strong> the binomial theorem was recog-<br />

nized as the first rigorous and general pro<strong>of</strong>. 6 <strong>The</strong> local criticism and scrutiny did not<br />

severely impair the evaluation <strong>of</strong> ABEL’S pro<strong>of</strong> — primarily because the fundamen-<br />

tal notions and knowledge <strong>of</strong> power series developed immensely over the nineteenth<br />

century.<br />

14.1.2 From ABEL’s “exception” to uniform convergence<br />

As already indicated and cited, one <strong>of</strong> the major historical interests in ABEL’S contri-<br />

bution to the rigorization movement was the “exception” which he presented against<br />

Cauchy’s <strong>The</strong>orem (see section 12.6). 7 <strong>The</strong> Fourier series representation <strong>of</strong> the function<br />

f (x) = x 2 on an interval such as ]−π, π[ provided an example that not every convergent<br />

sum <strong>of</strong> continuous functions was itself a continuous function as the Fourier series<br />

was periodically discontinuous at the end-points <strong>of</strong> the interval.<br />

4 (A. L. Crelle, 1829a, 305).<br />

5 (A. L. Cauchy, 1828).<br />

6 See e.g. (Stolz, 1904).<br />

7 This history is particularly well described in (Bottazzini, 1986). LAKATOS’ reconstruction (Lakatos,<br />

1976) is also extremely interesting.


14.1. Reception <strong>of</strong> ABEL’s rigorization 279<br />

ABEL’S “exception” met with little response in the 1820s and 1830s. Actually, it<br />

does not seem to have been quoted as influential in the first half <strong>of</strong> the 19 th century. In<br />

the 1840s, however, other and probably independent events put Cauchy’s <strong>The</strong>orem (see<br />

page 217) back on the agenda. Independently and simultaneously in 1847, the math-<br />

ematicians G. G. STOKES (1819–1903) and P. L. VON SEIDEL (1821–1896) published<br />

investigations <strong>of</strong> the conditions under which a convergent sum <strong>of</strong> continuous func-<br />

tions would not result in a continuous function. 8 In both cases, their research led to<br />

the realization that a particular mode <strong>of</strong> convergence was involved and SEIDEL gave<br />

it the name <strong>of</strong> “arbitrarily slow convergence”; 9 STOKES developed a refined hierarchy<br />

<strong>of</strong> modes <strong>of</strong> convergence on intervals. 10 Of the two publications, I find SEIDEL’S par-<br />

ticularly interesting because it was set up in the form <strong>of</strong> a pro<strong>of</strong> analysis and stressed<br />

the importance <strong>of</strong> keeping focus on the relations between limit processes. SEIDEL<br />

even proposed notational advances which would help clarify the interdependence <strong>of</strong><br />

nested limit processes. Such thoughts were important in completely separating limit<br />

processes from infinitesimals (see below).<br />

CAUCHY’S eventual reaction. Apparently, even these researches <strong>of</strong> British and Ger-<br />

man mathematicians did not directly prompt any reaction from the French mathemati-<br />

cians, in particular CAUCHY. Eventually, CAUCHY did address the Cauchy <strong>The</strong>orem<br />

again in an address to the Paris Academy <strong>of</strong> 1853. 11 In a paper, prompted by remarks<br />

made by French colleagues earlier that year, 12 CAUCHY described how the theorem <strong>of</strong><br />

the Cours d’analyse could be amended so that it no longer suffered any exceptions. <strong>The</strong><br />

fix which he proposed was the uniform convergence <strong>of</strong> the series in a form similar to<br />

the modern requirement. CAUCHY refined the assumptions <strong>of</strong> the theorem by requir-<br />

ing that a number N existed such that the difference |sm (x) − sn (x)| was less than ε<br />

for all values <strong>of</strong> x in the interval I under consideration when m, n ≥ N. CAUCHY’S<br />

requirement can easily be read as the modern definition <strong>of</strong> uniform convergence on<br />

the interval I,<br />

∀ε > 0 ∃N > 0 ∀m, n ≥ N ∀x ∈ I : |sm (x) − sn (x)| < ε.<br />

With this stricter assumption, the original pro<strong>of</strong> <strong>of</strong> the theorem carried through even<br />

without a more elaborate notation to handle the two limit processes. In the paper, 13<br />

CAUCHY considered the series<br />

∞<br />

sin nx<br />

∑ n n=1<br />

8 (Seidel, 1847; Stokes, 1847). GRATTAN-GUINNESS has pointed to the works <strong>of</strong> BJØRLING and considered<br />

him the “fourth man” in this development besides STOKES, SEIDEL, and CAUCHY (who was<br />

also involved, see below). See (I. Grattan-Guinness, 1986).<br />

9 (Seidel, 1847, 37).<br />

10 See (Stokes, 1847).<br />

11 (A.-L. Cauchy, 1853).<br />

12 (Briot and Bouquet, 1853a; Briot and Bouquet, 1853b).<br />

13 (A.-L. Cauchy, 1853, 31).


280 Chapter 14. Reception <strong>of</strong> ABEL’s contribution to rigorization<br />

which represents the function f (x) = π−x<br />

2<br />

on the interval ]0, 2π[. Thus, CAUCHY<br />

made use <strong>of</strong> a function similar to <strong>Abel</strong>’s exception and spoke <strong>of</strong> his job as removing the<br />

possibility <strong>of</strong> such exceptions. However, judging from his subsequent pro<strong>of</strong> revision,<br />

CAUCHY seems to have adapted a post-1820 use <strong>of</strong> counter examples (see chapter 21).<br />

CAUCHY’S attitude toward the status <strong>of</strong> the Cauchy <strong>The</strong>orem have been debated<br />

among historians <strong>of</strong> mathematics and different conclusions have been reached, in particular<br />

depending on which parts <strong>of</strong> CAUCHY’S 1853-paper have been emphasized. 14<br />

With the advent <strong>of</strong> non-standard analysis in the twentieth century, 15 interpretations <strong>of</strong><br />

various radicalism have proposed which render CAUCHY’S use <strong>of</strong> infinitesimals cor-<br />

rect within an enlarged system <strong>of</strong> real numbers. Although it is difficult to argue that<br />

CAUCHY was a modern non-standard analyst, 16 the debate over re-interpretations <strong>of</strong><br />

his works inspires consideration <strong>of</strong> the developments which led the modern concep-<br />

tion <strong>of</strong> the basic notions to be fixed they way they were in the century before non-<br />

standard analysis.<br />

<strong>The</strong> notions <strong>of</strong> convergence and continuity. ABEL’S reading <strong>of</strong> CAUCHY’S original<br />

definitions <strong>of</strong> continuity and convergence became standardized in the course <strong>of</strong> the<br />

nineteenth century century, noticeably through the works and teachings <strong>of</strong> G. P. L.<br />

DIRICHLET (1805–1859) and K. T. W. WEIERSTRASS (1815–1897). Whereas CAUCHY<br />

had possibly been unclear or ambiguous about his definitions, ABEL certainly read<br />

point-wise convergence and point-wise continuity into them. 17 This reading was en-<br />

forced by the path subsequently taken in analytical research, in particular concerning<br />

trigonometric series. For example, a multiplicity <strong>of</strong> different modes <strong>of</strong> convergence<br />

were introduced as the century unfolded. Some <strong>of</strong> the new modes <strong>of</strong> convergence such<br />

as absolute convergence and the ones introduced by SEIDEL and STOKES have been<br />

touched upon above. Similar to the creation <strong>of</strong> the concept <strong>of</strong> uniform convergence, a<br />

deliberate distinction between point-wise and uniform continuity was eventually intro-<br />

duced by H. E. HEINE (1821–1881) in 1872. 18 Toward the end <strong>of</strong> the nineteenth cen-<br />

tury, the concepts <strong>of</strong> convergence were even complemented (or stretched) by concepts<br />

<strong>of</strong> summability which could also treat non-convergent series. In the form <strong>of</strong> summabil-<br />

ity introduced by G. F. FROBENIUS (1849–1917), a class <strong>of</strong> non-convergent series such<br />

as ∑ (−1) n x n was ascribed a sum in a new sense. FROBENIUS considered the class <strong>of</strong><br />

series for which the average <strong>of</strong> the partial sums converged and called this limit the<br />

sum <strong>of</strong> the series. His central result generalized ABEL’S Lehrsatz IV and proved that<br />

his new definition <strong>of</strong> sum was a conservative extension <strong>of</strong> the Cauchyian definition.<br />

<strong>The</strong> variety <strong>of</strong> different concepts was the result <strong>of</strong> the explicit and precise formulation<br />

14 See e.g. the different interpretations in (Giusti, 1984), (Laugwitz, 1987), and (Bottazzini, 1990, xci).<br />

15 See e.g. (Cleave, 1971; Laugwitz, 1988–89).<br />

16 SPALT has come close to making this claim in (Spalt, 1981), though. More recently, he has retracted<br />

and debated this claim, see (Spalt, 2002, 326).<br />

17 See chapter 12.<br />

18 See (Dugac, 1989, 91–94).


14.2. Conclusion 281<br />

<strong>of</strong> basic notions in the Cauchyian sense to which ABEL had also contributed.<br />

14.2 Conclusion<br />

As described, ABEL’S contribution to the rigorization movement in analysis consisted<br />

<strong>of</strong> three aspects. Firstly, ABEL’S initial publication in the field on a general and rig-<br />

orous pro<strong>of</strong> <strong>of</strong> the binomial theorem was an impressive and early adoption <strong>of</strong> the<br />

Cauchyian program in the theory <strong>of</strong> series. In six theorems, ABEL presented results per-<br />

taining to series which were tailored for his pro<strong>of</strong> <strong>of</strong> the binomial theorem. Compared<br />

with CAUCHY’S original pro<strong>of</strong>, ABEL generalized the binomial theorem to include<br />

complex exponents by solving a slightly different functional equation. However, the<br />

six introductory theorems and the definitions with which they operated also revealed<br />

a development from CAUCHY’S Cours d’analyse. Most importantly, ABEL consistently<br />

read CAUCHY’S definitions <strong>of</strong> convergence and continuity as point-wise definitions.<br />

<strong>The</strong>se interpretations led him to an exception to Cauchy’s theorem and he subsequently<br />

replaced the effected theorem with his own version. Later, the exception would lead to<br />

the concept <strong>of</strong> uniform convergence. Secondly, ABEL partook in the debate over criteria<br />

<strong>of</strong> convergence which had become very important in CAUCHY’S reformulation <strong>of</strong> the<br />

theory <strong>of</strong> series. By means <strong>of</strong> a counter example and a very general argument, ABEL<br />

showed that no criteria <strong>of</strong> a particular form could complete delineate the concept <strong>of</strong><br />

convergent series. <strong>The</strong> methods which he employed to this end also led him to a new<br />

test <strong>of</strong> convergence which, however, he did not publish. Eventually, ABEL’S private<br />

criticism and scrutiny <strong>of</strong> the existing methods in analysis was expressed in his let-<br />

ters but only indirectly in his publications. This critical attitude may have influenced<br />

some <strong>of</strong> his contemporaries but — I believe — it was generally not considered among<br />

his most important contributions until more historical enquiries made it a central in-<br />

dication <strong>of</strong> the historical development <strong>of</strong> rigor in analysis.<br />

ABEL’S contribution to the rigorization movement has simultaneously been inter-<br />

preted in terms <strong>of</strong> changing epistemic standards. It has been described how ABEL was<br />

aware that a new set <strong>of</strong> standards were being deployed and that arguments should be<br />

modified to conform to these new norms. Moreover, ABEL advocated a critical re-<br />

vision which aimed at investigating how true results and only few paradoxes could<br />

arise from standards <strong>of</strong> argument which were no longer deemed to be rigorous. <strong>The</strong>se<br />

aspects as well as ABEL’S famous exception and the growth <strong>of</strong> new concepts will be-<br />

come important issues when the transition <strong>of</strong> paradigms is discussed in chapter 21.<br />

<strong>The</strong>re, the notion <strong>of</strong> critical revision is invoked as an explanation <strong>of</strong> the apparent cu-<br />

mulative nature <strong>of</strong> mathematics and the a rather literal interpretation <strong>of</strong> the exception<br />

is suggested.<br />

In conclusion, I find it fair to say that ABEL’S work on the rigorization <strong>of</strong> the theory<br />

<strong>of</strong> series was an interlude — albeit a quite passionate one. ABEL’S contributions were


282 Chapter 14. Reception <strong>of</strong> ABEL’s contribution to rigorization<br />

slightly marginal and not universally appreciated. Mathematicians who subsequently<br />

adhered to the rigorization program may have included ABEL among their heroes and<br />

have certainly adopted some <strong>of</strong> his notions and results but ABEL’S direct role was far<br />

from the role <strong>of</strong> the key initiator <strong>of</strong> the rigorization — CAUCHY.


Part IV<br />

Elliptic functions and the Paris<br />

mémoire<br />

283


Chapter 15<br />

Elliptic integrals and functions:<br />

Chronology and topics<br />

After the calculus was invented in the seventeenth century, it was quickly applied<br />

to classical problems concerning curves. One <strong>of</strong> the main achievements <strong>of</strong> the new<br />

tool was the ability to treat curves which had previously been outside the reach <strong>of</strong><br />

geometry. For instance, the quadrature <strong>of</strong> the hyperbola provided an analytical way<br />

<strong>of</strong> describing and treating logarithmic functions. After the calculus had conquered<br />

such basic curves as the logarithmic and trigonometric ones, the determination <strong>of</strong> the<br />

length <strong>of</strong> an ellipse became a major obstacle on the path to generality.<br />

In the eighteenth century, L. EULER’S (1707–1783) new vision <strong>of</strong> the calculus as<br />

founded upon functions also transformed the way in which curves were approached. 1<br />

<strong>The</strong> way in which the elliptic transcendentals enter into the realm <strong>of</strong> analysis touches<br />

upon a number <strong>of</strong> points which will be described below and — primarily — connected<br />

to N. H. ABEL’S (1802–1829) work:<br />

1. During the eighteenth century, a need came to be felt to include higher transcen-<br />

dentals (i.e. functions different from the algebraic, logarithmic, and trigonomet-<br />

ric ones) into analysis on a par with the well established elementary functions. In<br />

order to accept these new objects into analysis, they had to undergo a process <strong>of</strong><br />

becoming known. This process manifested itself in various ways, e.g. in the search<br />

for acceptable analytical representations <strong>of</strong> the new objects.<br />

2. Because the new functions were in some senses generalizations <strong>of</strong> the elementary<br />

functions, their study opened possibilities <strong>of</strong> generalization <strong>of</strong> existing results.<br />

In the process, insights into the new objects were obtained which also helped<br />

making them known.<br />

3. Ultimately, the consensus on how to introduce particular new higher transcen-<br />

dentals — as primitive functions <strong>of</strong> algebraic differentials — led to a research pro-<br />

1 For EULER’S approach to analysis, see section 10.1.<br />

285


286 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

gram aimed at describing in a general way properties <strong>of</strong> larger classes <strong>of</strong> tran-<br />

scendentals.<br />

By the end <strong>of</strong> the 1820s, the theory <strong>of</strong> elliptic (and even higher) transcendentals was<br />

establishing itself as one <strong>of</strong> the major research fields in mathematics in the century.<br />

Thus, much effort was put into the field and many connections with and implications<br />

for other fields were discovered.<br />

Evidently, the above themes represent aspects <strong>of</strong> the fundamental change toward<br />

concept based mathematics. Based on a description <strong>of</strong> the background and contents <strong>of</strong><br />

ABEL’S work with these objects, the transition becomes quite evident and susceptible<br />

<strong>of</strong> further qualification.<br />

<strong>The</strong> theory <strong>of</strong> transcendental functions constitutes a major part <strong>of</strong> ABEL’S works<br />

and provide material for further analysis <strong>of</strong> some <strong>of</strong> the characterizations <strong>of</strong> his work.<br />

For instance, it will become clear that he employed one standard <strong>of</strong> rigor — differ-<br />

ent from the sharp manifest <strong>of</strong> his foundational research on infinite series — studying<br />

these new objects, even when the theory <strong>of</strong> infinite series was involved. Infinite series<br />

(and products) were used to represent the new objects <strong>of</strong> analysis by better established<br />

ones. This process <strong>of</strong> coming to know new objects can be traced in ABEL’S works and<br />

merits attention.<br />

Furthermore, interesting aspects <strong>of</strong> ABEL’S general inclination toward algebraic<br />

methods is evident in many <strong>of</strong> his researches on transcendental objects. In order to<br />

analyze this algebraic approach to the theory <strong>of</strong> transcendental objects, an understand-<br />

ing <strong>of</strong> the questions which ABEL wanted to answer is required. Only when questions<br />

and methods are viewed together can sense be made <strong>of</strong> the statement that “ABEL’S<br />

approach was algebraic”.<br />

15.1 Elliptic transcendentals before the nineteenth<br />

century<br />

Ellipses were known to and treated by mathematicians since the times <strong>of</strong> the Greeks.<br />

<strong>The</strong>y knew that the ellipse can be obtained by a central projection from a circle and de-<br />

duced numerous results concerning these objects in their mathematical investigations<br />

<strong>of</strong> conic sections. However, with the advent <strong>of</strong> symbolic notation, the calculus, and I.<br />

NEWTON’S (1642–1727) theory <strong>of</strong> gravitation, the study <strong>of</strong> curves such as the ellipse<br />

changed.<br />

15.1.1 Rectification <strong>of</strong> the ellipse<br />

When the creators and early practitioners <strong>of</strong> the calculus sought to promote their new<br />

tool, they <strong>of</strong>ten attacked problems which belonged to the classical realm <strong>of</strong> analysis <strong>of</strong><br />

curves. Traditionally, the study <strong>of</strong> curves included such problems as the construction


15.1. Elliptic transcendentals before the nineteenth century 287<br />

<strong>of</strong> points on the curve, its rectification and quadrature and the determination <strong>of</strong> its<br />

tangents, centres <strong>of</strong> curvature, involutes, and evolutes. A number <strong>of</strong> these properties<br />

together constituted the knowledge required for a curve to be known and it was one<br />

<strong>of</strong> the greatest achievements <strong>of</strong> the calculus to devise a method for obtaining most <strong>of</strong><br />

them from a single defining equation. 2<br />

A variety <strong>of</strong> different curves was treated in the first decades <strong>of</strong> the calculus, and<br />

many problems were reduced to “simpler” problems, primarily to the construction <strong>of</strong><br />

points on algebraic curves, the rectification <strong>of</strong> the circle (inverse trigonometric func-<br />

tions) and the quadrature <strong>of</strong> the hyperbola (logarithmic functions). A classification <strong>of</strong><br />

construction problems was established based on the simpler problems to which the<br />

solution <strong>of</strong> the given problem could be reduced.<br />

Despite the efforts <strong>of</strong> the mathematicians, certain problems defied the accepted<br />

known means <strong>of</strong> solution; for instance, when asked to compute the arc lengths <strong>of</strong><br />

ellipses and some other curves, mathematicians found that the known simple integrals<br />

did not suffice.<br />

In the year 1675, G. W. LEIBNIZ (1646–1716) directed two enquiries concerning the<br />

rectification <strong>of</strong> the ellipse to the British mathematicians J. GREGORY (1638–1675) and<br />

NEWTON. LEIBNIZ received the answer that the British mathematicians could only<br />

compute the length <strong>of</strong> an ellipse by approximation, i.e. with the help <strong>of</strong> infinite series,<br />

and did not possess any closed expression for the length. LEIBNIZ, himself, at the time<br />

believed that he could reduce this problem (and the rectification <strong>of</strong> the hyperbola) to<br />

the quadratures <strong>of</strong> the circle and the hyperbola. Later, LEIBNIZ realized that he had<br />

been misled by a computational error. 3<br />

As can be seen from box 6, the rectification <strong>of</strong> the ellipse involved the computation<br />

<strong>of</strong> an integral <strong>of</strong> the form � R(x) dx<br />

√ P4(x) in which R and P4 were polynomials such that<br />

deg P4 ≤ 4. Such integrals (with the relaxed assumption that R be only a rational function)<br />

were soon called elliptic integrals by G. C. FAGNANO DEI TOSCHI (1682–1766). 4<br />

LEIBNIZ’ question reflects the search for simpler, finite representations <strong>of</strong> elliptic inte-<br />

grals.<br />

In a paper written in 1732 but not published until six years later, 5 EULER deduced<br />

a series representation <strong>of</strong> a quarter <strong>of</strong> the circumference <strong>of</strong> an ellipse. Based on a figure<br />

(see figure 15.1) in which M represented a point on the ellipse with center C and semi-<br />

axes CA = a and CB = b, EULER expressed the differential <strong>of</strong> the arc-length �AM<br />

as<br />

ds = b2√ b 2 + t 2 + nt 2 dt<br />

(b 2 + t 2 ) 3 2<br />

2 For information on these aspects <strong>of</strong> curves, see e.g. (Loria, 1902). For a general discussion on the<br />

conceptions <strong>of</strong> curves before the advent <strong>of</strong> the calculus, see (H. J. M. Bos, 2001).<br />

3 (H<strong>of</strong>mann, 1949, 75,118).<br />

4 (Natucci, 1971, 516). For more on FAGNANO DEI TOSCHI’S work on elliptic integrals, see below.<br />

5 (L. Euler, 1732a).


288 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

Rectification <strong>of</strong> the ellipse Consider the ellipse with major axis 2a and minor axis<br />

2b given by the Cartesian equation<br />

Obviously, from this equation<br />

x 2<br />

a<br />

b<br />

y2<br />

+ = 1.<br />

2 2<br />

x = a cos θ and y = b sin θ<br />

which means<br />

dx<br />

dy<br />

= −a sin θ and = b cos θ.<br />

dθ dθ<br />

To compute the arc length, we find<br />

�<br />

s (θ) =<br />

� �<br />

dx<br />

dθ<br />

� �<br />

= b<br />

� 2<br />

+<br />

� �2 �<br />

dy �<br />

dθ = a<br />

dθ<br />

2 sin2 θ + b2 cos2 θ dθ<br />

1 − k 2 sin 2 θ dθ with k 2 = b2 − a 2<br />

This is precisely the form <strong>of</strong> A.-M. LEGENDRE’S (1752–1833) second kind <strong>of</strong> elliptic<br />

integrals with the modulus k (denoted F (θ, k) in table 15.2).<br />

In order to reduce the integral to the form � R(x) dx<br />

, we substitute z = sin θ and get<br />

and thus<br />

dz = cos θ dθ =<br />

√ P4(x)<br />

b 2<br />

�<br />

1 − sin 2 θ dθ ⇒ dθ =<br />

� �<br />

1 − k2 sin2 �<br />

θ dθ =<br />

√ 1 − k2z2 �<br />

√ dz =<br />

1 − z2 Box 6: Rectification <strong>of</strong> the ellipse<br />

.<br />

1<br />

√ 1 − z 2 dz<br />

1 − k 2 z 2<br />

� (1 − k 2 z 2 ) (1 − z 2 ) dz.<br />

Figure 15.1: EULER’S rectification <strong>of</strong> an ellipse by infinite series (L. Euler, 1732a, 2)


15.2. <strong>The</strong> lemniscate 289<br />

in which the semi-axes were related by<br />

a 2 = (n + 1) b 2 .<br />

EULER then expanded the square root by use <strong>of</strong> the binomial theorem<br />

�<br />

(b 2 + t 2 ) + nt 2 = �<br />

b 2 + t 2 +<br />

∞<br />

∑<br />

µ=1<br />

(−1) µ−1 ∏ µ−1<br />

k=1 (2k − 1)<br />

∏ µ<br />

k=1 (2k)<br />

n µ t 2µ<br />

(b2 + t2 ) 2µ−1<br />

2<br />

Thus, EULER obtained the differential in the form (A0, A1, . . . were specified constants)<br />

∞<br />

2<br />

ds = b ∑<br />

µ=0<br />

Aµn µ t 2µ dt<br />

(b2 + t2 .<br />

µ+1<br />

)<br />

which he next integrated term-wise from 0 to ∞ to obtain the rectification <strong>of</strong> a quarter<br />

<strong>of</strong> the ellipse in the form<br />

�AMB = π<br />

2<br />

15.2 <strong>The</strong> lemniscate<br />

∞<br />

∑<br />

µ=0<br />

∏ µ−1<br />

k=0 (2k + 1)2<br />

∏ µ<br />

k=1 (2k)2 (2µ − 1) n µ .<br />

Another curve which received the attention <strong>of</strong> mathematicians starting with the broth-<br />

ers JAKOB I BERNOULLI (1654–1705) and JOHANN I BERNOULLI (1667–1748) was the<br />

so-called lemniscate. 6 <strong>The</strong> curve was defined by the Cartesian equation<br />

�<br />

x 2 + y 2� 2 �<br />

2 = a x 2 − y 2�<br />

,<br />

and both brothers recognized that the arc length <strong>of</strong> the curve depended on an integral<br />

<strong>of</strong> the form �<br />

(see box 7).<br />

dz<br />

√ 1 − z 4<br />

In Italy, the autodidact nobleman FAGNANO DEI TOSCHI took up the study <strong>of</strong> the<br />

lemniscate. 7 By a set <strong>of</strong> theorems, FAGNANO DEI TOSCHI was able to prove that the<br />

division <strong>of</strong> the quadrant <strong>of</strong> the lemniscate into k parts could be constructed by ruler<br />

and compass if k was <strong>of</strong> one <strong>of</strong> the forms 2 × 2 m , 3 × 2 m , or 5 × 2 m . By elimination <strong>of</strong><br />

the intermediate variable x in the substitutions<br />

x =<br />

�<br />

1 − √ 1 − z 4<br />

FAGNANO DEI TOSCHI obtained that<br />

dz<br />

√ 1 − z 4<br />

z<br />

and x =<br />

du<br />

= 2√<br />

1 − u4 √ 2u<br />

√ 1 − u 4 ,<br />

and he had obtained the duplication <strong>of</strong> any segment <strong>of</strong> the lemniscate arc.<br />

6 See (H. J. M. Bos, 1974). Mostly, discovery <strong>of</strong> the curve is attributed to BERNOULLI alone as he holds<br />

priority <strong>of</strong> publication and gave the curve its name.<br />

7 Unfortunately, I not have had access to FAGNANO DEI TOSCHI’S original works. Instead, the short<br />

outline is based on (R. Ayoub, 1984; Siegel, 1959).<br />

.


290 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

Rectification <strong>of</strong> the lemniscate To find the arc length <strong>of</strong> the lemniscate given by the<br />

polar equation<br />

we compute<br />

i.e.<br />

r 2 = a 2 cos 2θ,<br />

2r dr = −2a 2 sin 2θ dθ,<br />

ds 2 = r 2 dθ 2 + dr 2 = a 2 cos2 2θ + sin2 2θ<br />

dθ<br />

cos 2θ<br />

2 ,<br />

a dθ a dθ<br />

ds = √ = �<br />

cos 2θ<br />

�<br />

s (θ) = a<br />

dθ<br />

�<br />

1 − 2 sin2 θ .<br />

1 − 2 sin 2 θ ,<br />

This is an example <strong>of</strong> an elliptic integral <strong>of</strong> the first kind. With the substitution z =<br />

sin θ, we find (see box 6)<br />

�<br />

s = a<br />

1 dz<br />

√ √<br />

1 − 2z2 1 − z2 �<br />

= a<br />

dz<br />

√ 1 − 3z 2 + 2z 4 .<br />

Box 7: Rectification <strong>of</strong> the lemniscate<br />

15.2.1 Addition <strong>of</strong> lemniscatic arcs<br />

EULER took his inspiration directly from the works <strong>of</strong> FAGNANO DEI TOSCHI. In 1750,<br />

after FAGNANO DEI TOSCHI had published his collected works, the author sent a copy<br />

to the Berlin Academy <strong>of</strong> Sciences <strong>of</strong> which he was a member. <strong>The</strong> following year,<br />

on 23 December 1751, the work came into the hands <strong>of</strong> EULER who was given the<br />

assignment <strong>of</strong> commenting upon it. 8 C. G. J. JACOBI (1804–1851) has called this date<br />

the birthday <strong>of</strong> elliptic functions.<br />

In the process <strong>of</strong> preparing an answer for FAGNANO DEI TOSCHI, EULER became<br />

very interested in the topic <strong>of</strong> lemniscate integrals. EULER commented on his new<br />

research in a letter to C. GOLDBACH (1690–1764):<br />

“Recently, I have come across a curious integration. Just as the integral <strong>of</strong><br />

√ is yy + xx = cc + 2xy<br />

1−yy √ 1 − cc, the integral <strong>of</strong> the<br />

the equation dx<br />

√ 1−xx = dy<br />

equation dx<br />

√ 1−x 4<br />

= dy<br />

√ 1−x 4 is<br />

�<br />

yy + xx = cc + 2xy 1 − c4 − ccxxyy.” 9<br />

8 (Siegel, 1959).<br />

9 “Neulich bin ich auch auf curieuse Integrationen verfallen. Dann gleich wie von dieser Äquation<br />

√ dx =<br />

(1−xx)<br />

dy<br />

√ (1−yy) das integrale ist yy + xx = cc + 2xy � (1 − cc), also ist von dieser Äquation


15.2. <strong>The</strong> lemniscate 291<br />

Figure 15.2: LEONHARD EULER (1707–1783)<br />

In the letter, EULER applied the theorem to demonstrate how the difference be-<br />

tween two segments <strong>of</strong> the arc <strong>of</strong> an ellipse could be rectified. <strong>The</strong>re is no pro<strong>of</strong> <strong>of</strong> the<br />

theorem in the letter but EULER published a pro<strong>of</strong> in 1756/56. 10 <strong>The</strong> pro<strong>of</strong> progressed<br />

by direct differentiation <strong>of</strong> the purported integral<br />

to obtain<br />

y 2 + x 2 = c 2 �<br />

+ 2xy 1 − c4 − c 2 x 2 y 2<br />

dx<br />

√ 1 − x 4 =<br />

dy<br />

� 1 − y 4 .<br />

(15.1)<br />

<strong>The</strong> theorem founded a particular branch <strong>of</strong> the theory <strong>of</strong> elliptic integrals as it con-<br />

tained the so-called addition theorem for lemniscate integrals. If x and y were related<br />

by (15.1), the equation<br />

� dx<br />

(1−x4 ) =<br />

dy<br />

�<br />

(1−y4 )<br />

� x<br />

0<br />

das integrale:<br />

dt<br />

√<br />

1 − t4 =<br />

� y<br />

0<br />

dt<br />

√ + C<br />

1 − t4 �<br />

yy + xx = cc + 2xy (1 − c4 ) − ccxxyy.”<br />

(Euler→Goldbach, 1752. Euler and Goldbach, 1965, 347–348); also (Fuss, 1968, I, 567).<br />

10 (L. Euler, 1756/57).


292 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

holds where C is constant (independent <strong>of</strong> x and y). However, if we let y = 0, we find<br />

x 2 = c 2 , and thus<br />

C =<br />

� c<br />

0<br />

dt<br />

√ 1 − t 4 .<br />

In this form, the addition theorem for lemniscate integrals is apparent,<br />

� x<br />

0<br />

dt<br />

√<br />

1 − t4 +<br />

� y dt<br />

√<br />

0 1 − t4 =<br />

� z<br />

0<br />

dt<br />

√ , if<br />

1 − t4 x 2 + y 2 + z 2 x 2 y 2 = z 2 �<br />

+ 2xy<br />

15.2.2 EULER’s rectification <strong>of</strong> the lemniscate<br />

1 − z 4 .<br />

At least twice, EULER deduced expressions for the arc length <strong>of</strong> the lemniscate by<br />

infinite series. In the third part <strong>of</strong> his trilogy, the Institutiones calculi integralis, 11 EULER<br />

found the relation<br />

� 1<br />

0<br />

x m+1 dx<br />

√ 1 − x 2<br />

�<br />

m 1<br />

=<br />

m + 1 0<br />

xm−1 dx<br />

√ . (15.2)<br />

1 − x2 Subsequently, 12 EULER expressed the length <strong>of</strong> the first quadrant <strong>of</strong> the lemniscate<br />

based on the relation<br />

� 1<br />

Using the binomial theorem, he wrote<br />

0<br />

�<br />

1 + x 2� − 1 2<br />

0<br />

dx<br />

√<br />

1 − x4 =<br />

� 1 �<br />

1 + x<br />

0<br />

2� − 1 2<br />

dx<br />

√ 1 − x 2 .<br />

= 1 − 1<br />

2 x2 1 · 3<br />

+<br />

2 · 4 x4 1 · 3 · 5<br />

−<br />

2 · 4 · 6 x6 + . . .<br />

and when the term-wise integration was carried out, EULER found the expression<br />

� �<br />

1 dx π<br />

√ = 1 −<br />

1 − x4 2<br />

12<br />

22 + 1232 2242 − 123252 2242 �<br />

+ . . .<br />

62 using the relation (15.2), above.<br />

<strong>The</strong> deduction <strong>of</strong> the result in the Institutiones is less complicated than the similar<br />

result for the ellipse given by EULER in 1732 and described above. However, the basic<br />

tools <strong>of</strong> the two approaches are the same: expansion by use <strong>of</strong> the binomial theorem<br />

and term-wise integration <strong>of</strong> the power series which was thus obtained.<br />

15.3 LEGENDRE’s theory <strong>of</strong> elliptic integrals<br />

Toward the end <strong>of</strong> the eighteenth century, LEGENDRE gave the theory <strong>of</strong> elliptic inte-<br />

grals a new twist with his contributions. In a number <strong>of</strong> lengthy papers and mono-<br />

11 (L. Euler, 1768, XI, 208).<br />

12 (ibid., XI, 211).


15.3. LEGENDRE’s theory <strong>of</strong> elliptic integrals 293<br />

1793 Mémoire sur les transcendantes elliptiques<br />

1811,1817,1816 Exercises de calcul intégral sur divers<br />

ordres de transcendantes et sur les<br />

quadratures<br />

1825,1826,1828 Traité des fonctions ellipitiques et des<br />

intégrales eulériennes<br />

Table 15.1: LEGENDRE’s publications on elliptic transcendentals<br />

Figure 15.3: ADRIEN-MARIE LEGENDRE (1752–1833)<br />

graphs, LEGENDRE developed his theory <strong>of</strong> elliptic and other higher integrals. 13 Even-<br />

tually, in the 1820s toward the end <strong>of</strong> his life, LEGENDRE decided to publish his re-<br />

search on elliptic integrals in the form <strong>of</strong> a number <strong>of</strong> monographs. <strong>The</strong> first two vol-<br />

umes <strong>of</strong> the Traité des fonctions elliptiques et des intégrales eulériennes, laid the foundation<br />

and presented the state <strong>of</strong> the art in the field.<br />

Generally, LEGENDRE’S theory <strong>of</strong> elliptic integrals concerned the transformation<br />

and numerical approximation <strong>of</strong> these integrals. An important position in LEGEN-<br />

DRE’S approach to the theory was taken by the classification <strong>of</strong> such integrals into<br />

a small number <strong>of</strong> canonical forms. LEGENDRE worked with three canonical forms<br />

which he termed the elliptic functions <strong>of</strong> the first, second, and third kinds [espèce] (see<br />

table 15.2 and box 8). Later, ABEL reserved the word elliptic function for the inverse<br />

13 (A. M. Legendre, 1811–1817; A.-M. Legendre, 1793).


294 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

Kind Integral in x Integral in φ Symbol<br />

�<br />

�<br />

dx<br />

dφ<br />

First √ √ F (φ, k)<br />

(1−x2 )(1−k2 x2 )<br />

1−k2 sin2 φ<br />

Second � � 1−k2x2 1−x2 �<br />

dx<br />

�<br />

1 − k2 sin2 φ dφ E (φ, k)<br />

�<br />

� dφ<br />

Third<br />

Π (φ, n, k)<br />

dx<br />

(1+nx 2 ) √ (1−x 2 )(1−k 2 x 2 )<br />

(1+n sin 2 φ) √ 1−k 2 sin 2 φ<br />

<strong>The</strong> variables φ and x are related by x = sin φ.<br />

In accordance with 18 th century practice, the integrations<br />

are to be performed from 0 to the x and φ, respectively.<br />

Table 15.2: LEGENDRE’s classification <strong>of</strong> elliptic integrals<br />

function <strong>of</strong> the integrals and to avoid confusion, I will <strong>of</strong>ten refer to LEGENDRE’S el-<br />

liptic functions as elliptic integrals.<br />

LEGENDRE introduced the notation<br />

∆ (φ) = ∆ (φ, k) =<br />

�<br />

1 − k 2 sin 2 φ.<br />

LEGENDRE’S approach based on reducing elliptic integrals to a number <strong>of</strong> basic<br />

forms was not entirely new. In the 1780s, J. L. LAGRANGE (1736–1813) had made<br />

similar attempts reducing all elliptic integrals to the basic form<br />

�<br />

M � x 2� dx<br />

� (1 ± p 2 x 2 ) (1 ± q 2 x 2 )<br />

in which M was a rational function. 14 Later, ABEL took up the reduction and sug-<br />

gested another set <strong>of</strong> basic forms. Thus, from the last decades <strong>of</strong> the 18 th century and<br />

into the 1820s, the theory <strong>of</strong> elliptic integrals was still in its formative process and the<br />

basic representations had not been decided upon yet. With the advent <strong>of</strong> LEGENDRE’S<br />

Traité des fonctions elliptiques, its author hoped to settle the foundations once and for<br />

all and his categorization into three kinds <strong>of</strong> integrals was successful. Soon thereafter,<br />

however, the theory changed dramatically with the novel ideas <strong>of</strong> ABEL and JACOBI,<br />

and instead it was JACOBI’S Fundamenta nova which became the foundation <strong>of</strong> the theory<br />

and established its notation. 15<br />

Complete integrals and the reduction program. LEGENDRE introduced the notation<br />

E1 , F1 , Π1 for the complete integrals which corresponded to φ = π 2 . <strong>The</strong>se particular<br />

numbers (functions <strong>of</strong> the modulus k) received some special attention and a number<br />

<strong>of</strong> remarkable relations were discovered among them. LEGENDRE developed the com-<br />

plete integrals E 1 and F 1 into series. LEGENDRE also devoted quite some effort to the<br />

problems <strong>of</strong> comparison <strong>of</strong> elliptic integrals <strong>of</strong> the three kinds.<br />

14 (Lagrange, 1784–1785, 264).<br />

15 (C. G. J. Jacobi, 1829). See also chapter 20.


15.3. LEGENDRE’s theory <strong>of</strong> elliptic integrals 295<br />

LEGENDRE’S reduction <strong>of</strong> elliptic integrals In his reduction, LEGENDRE first considered<br />

the integral � P dx<br />

R in which P was a polynomial and R was the square root <strong>of</strong><br />

a fourth degree polynomial,<br />

�<br />

�<br />

�<br />

R =<br />

� 4<br />

∑<br />

m=0<br />

αmx m .<br />

Obviously, such integrals could be studied by studying the simpler ones <strong>of</strong> the form<br />

� x k dx<br />

R which LEGENDRE denoted Πk (this is distinct from the Π which denotes integrals<br />

<strong>of</strong> the third kind). LEGENDRE found<br />

x k−3 �<br />

R =<br />

=<br />

�<br />

d x k−3 �<br />

�<br />

R = (k − 3)<br />

4 �<br />

∑<br />

m=0<br />

k − 3 + m<br />

2<br />

x k−4 �<br />

R dx +<br />

x k−3 R ′ dx (15.3)<br />

�<br />

αmΠ m+k−4 . (15.4)<br />

This meant, that for any k ≥ 4, the integral Π k depended algebraically on the previous<br />

integrals Π 0 , Π 1 , . . . , Π k−1 . By writing out (15.4), LEGENDRE observed that also the<br />

integral Π 3 only depended algebraically on Π 0 , Π 1 , and Π 2 . <strong>The</strong>refore, the integral<br />

� P dx<br />

R in which P was a polynomial could be reduced algebraically to the integrals Π 0 ,<br />

Π 1 , and Π 2 . Furthermore, knowledge <strong>of</strong> Π 0 and Π 2 would also entail knowledge <strong>of</strong><br />

Π1 by a linear transformation.<br />

Thus, elliptic integrals � P(x) dx<br />

R in which P was a polynomial had been taken care <strong>of</strong><br />

and had been reduced to the two integrals � dx<br />

R and � x2 dx<br />

R . For more general, rational<br />

functions P, LEGENDRE expanded into partial fractions, considered �<br />

dx , and<br />

(1+nx)R<br />

applied a similar line <strong>of</strong> argument.<br />

Initially, LEGENDRE chose and ordered his basic forms according to the conic sec-<br />

tions whose rectification they described. Thus, the integral E = � ∆ dφ was considered<br />

the most basic as it described the rectification <strong>of</strong> the ellipse. <strong>The</strong> second class was ini-<br />

tially represented by Υ = ∆ tan φ − � ∆ dφ + b 2 � dφ<br />

∆<br />

hyperbola. This class was, however, soon replaced by F = � dφ<br />

∆<br />

represented by the integral Π = � dφ<br />

(1+n sin 2 φ)∆<br />

involved a third parameter, n.<br />

which represented the arc <strong>of</strong> the<br />

. <strong>The</strong> third class was<br />

which, contrary to the two first kinds,<br />

Box 8: LEGENDRE’S reduction <strong>of</strong> elliptic integrals


296 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

In order to facilitate computation <strong>of</strong> numerical values, LEGENDRE developed a the-<br />

ory by which a sequence <strong>of</strong> moduli could be constructed which allowed the numerical<br />

approximation <strong>of</strong> elliptic integrals to be reduced.<br />

15.4 Left in the drawer: GAUSS on elliptic functions<br />

When C. F. GAUSS (1777–1855) was informed <strong>of</strong> ABEL’S first publication on elliptic<br />

functions, the Recherches, his answer was extraordinary. 16 GAUSS was impressed with<br />

ABEL’S work and was happy to see that the young Norwegian had relieved him <strong>of</strong> the<br />

obligation to publish a third <strong>of</strong> his own knowledge concerning these elliptic functions.<br />

Furthermore, GAUSS was surprised to see that ABEL had followed almost exactly the<br />

same route as he, himself, had taken to the point where their symbols were the same.<br />

As GAUSS never published any <strong>of</strong> the monographs on elliptic functions which he had<br />

intended, historians have had to look in his diary and in some <strong>of</strong> his manuscripts for<br />

hints concerning his results and methods.<br />

From 1797, GAUSS’ diary documents an increasing interest in the lemniscate inte-<br />

gral. 17 Despite a lasting interest and many connections to other parts <strong>of</strong> his research,<br />

GAUSS never published on the theory <strong>of</strong> lemniscate integrals. Thus, GAUSS’ ideas<br />

only indirectly influenced the development <strong>of</strong> the theory <strong>of</strong> elliptic functions in the<br />

1820s. In order to illustrate how GAUSS arrived at some <strong>of</strong> his insights and to under-<br />

stand his remarks on ABEL’S Recherches, a brief discussion <strong>of</strong> important points in his<br />

manuscripts and in his mathematical diary is given. Emphasis is here put on the in-<br />

version <strong>of</strong> the lemniscate integral into GAUSS’ lemniscate function, the periods <strong>of</strong> the<br />

lemniscate function, and GAUSS’ representation <strong>of</strong> the lemniscate function by various<br />

infinite expressions.<br />

Whereas GAUSS’ diary obviously provides a strict chronological frame, his manu-<br />

scripts are less clearly ordered. <strong>The</strong> manuscripts contained in the Werke have been<br />

compiled and put into an order which fit the editor. <strong>The</strong>refore, and because GAUSS’<br />

ideas had no direct impact on his immediate successors, I have taken the liberty to<br />

treat his production in its thematic contexts.<br />

<strong>The</strong> role <strong>of</strong> GAUSS’ knowledge. As mentioned, GAUSS deferred publication on the<br />

subject <strong>of</strong> lemniscate functions. According to SCHLESINGER, GAUSS had hoped to<br />

publish on his research on higher transcendentals in a form which would combine his<br />

three greatest interests in the field: the lemniscate function, the arithmetic-geometric<br />

means and the hypergeometric series. 18 GAUSS did publish on the hypergeometric<br />

series, 19 and there is a brief description <strong>of</strong> arithmetic-geometric means in his work De-<br />

16 (Gauss→Bessel, 1828.03.30. In Gauss and Bessel, 1880).<br />

17 (C. F. Gauss, 1981; J. J. Gray, 1984).<br />

18 (Schlesinger, 1922–1933, 27).<br />

19 (C. F. Gauss, 1813).


15.5. Chronology <strong>of</strong> ABEL’s work on elliptic transcendentals 297<br />

terminatio attractionis. 20 Thus, when GAUSS spoke <strong>of</strong> the third <strong>of</strong> his research which<br />

ABEL had anticipated, he probably referred to the theory <strong>of</strong> lemniscate functions. Con-<br />

cerning the lemniscate function, GAUSS had performed the inversion, extended to<br />

complex variables, found the resulting function to be doubly periodic, expressed its<br />

addition formulae, and obtained infinite representations <strong>of</strong> it. <strong>The</strong> division problem <strong>of</strong><br />

the lemniscate had played an important role (together with the arithmetic-geometric<br />

means) in motivating his research. Concerning all these results and aspects, GAUSS<br />

was certainly correct in observing the similarity with ABEL’S approach. It is possible,<br />

but not a necessary assumption, that rumors <strong>of</strong> GAUSS’ investigations and their meth-<br />

ods had spread to ABEL through H. C. SCHUMACHER (1784–1873) and C. F. DEGEN<br />

(1766–1825); certain <strong>of</strong> GAUSS’ letters to SCHUMACHER suggest that GAUSS for a short<br />

while considered it a possibility. 21<br />

15.5 Chronology <strong>of</strong> ABEL’s work on elliptic<br />

transcendentals<br />

Little is known <strong>of</strong> ABEL’S first encounters with elliptic functions. Presumably, ABEL<br />

took up DEGEN’S suggestion in the letter to C. HANSTEEN (1784–1873) and began<br />

studying the higher transcendentals possibly through the works <strong>of</strong> EULER and LEG-<br />

ENDRE. A letter from his stay in Copenhagen 1823 indicates that he had shown DEGEN<br />

a small paper in which “inverse functions <strong>of</strong> elliptic transcendentals” played a role. 22<br />

In both editions <strong>of</strong> ABEL’S Œuvres, a number <strong>of</strong> manuscripts are included which were<br />

among the papers destroyed in a fire in B. M. HOLMBOE’S (1795–1850) house in 1849. 23<br />

According to HOLMBOE, the manuscripts date from before ABEL’S European tour, i.e.<br />

they were written before 1825. 24 Apparently based on these manuscripts and the let-<br />

ter from Copenhagen (as there are no other primary sources), 25 some historians have<br />

credited ABEL with possessing the key results and methods around 1823.<br />

It was, however, not until during and after the European tour that ABEL developed<br />

and published his research on elliptic functions and higher transcendentals which<br />

would merit so much attention. ABEL’S mature research on the topics can be seper-<br />

ated into three categories. His first publication on the subject was the Recherches, which<br />

introduced the crucial idea <strong>of</strong> inverting elliptic integrals <strong>of</strong> the first kind into elliptic<br />

functions and established the latter as doubly periodic functions <strong>of</strong> a complex variable.<br />

Simultanously with the publication <strong>of</strong> ABEL’S Recherches, the German mathemati-<br />

cian CARL GUSTAV JACOB JACOBI announced some results on the transformation <strong>of</strong><br />

20 (C. F. Gauss, 1818).<br />

21 (Schlesinger, 1922–1933, 167). I hope to have more to say on this at a later stage in connection with<br />

future research on the mathematical milieu in Copenhagen in the early nineteenth century.<br />

22 (<strong>Abel</strong>→Holmboe, Kjøbenhavn, 1823/08/04. N. H. <strong>Abel</strong>, 1902a, 5).<br />

23 (N. H. <strong>Abel</strong>, 1881, II, 324) and (Stubhaug, 1996, 560).<br />

24 (Holmboe in N. H. <strong>Abel</strong>, 1839, i)<br />

25 (<strong>Abel</strong>→Holmboe, Kjøbenhavn, 1823/08/04. In N. H. <strong>Abel</strong>, 1902a, 4–8).


298 Chapter 15. Elliptic integrals and functions: Chronology and topics<br />

VII Propriétés remarquables de la fonction<br />

y = φx déterminée par l’équation f y.dx −<br />

f x � (a − y) (a1 − y) (a2 − y) . . . (am − y) =<br />

0, f étant une fonction quelconque de y<br />

qui ne devient pas nulle ou infinie lorsque<br />

y = a, a1, a2, . . . , am<br />

VIII+IX Sur une propriété remarquable d’une classe très<br />

étendue de fonctions transcendantes<br />

X Sur la comparaison des fonctions transcendantes<br />

XIII Théorie des transcendantes elliptiques<br />

Table 15.3: ABEL’S early unpublished works on elliptic integrals and related topics.<br />

<strong>The</strong> manuscripts are no longer extant but HOLMBOE dated them all to the period before<br />

ABEL’S European tour. <strong>The</strong> roman numerals indicate the position <strong>of</strong> the manuscript<br />

in (N. H. <strong>Abel</strong>, 1881, II).<br />

elliptic integrals which were astounding to ABEL. JACOBI had obtained results which<br />

were special cases <strong>of</strong> ABEL’S own findings, and ABEL was surprised by the sudden<br />

element <strong>of</strong> competition. For a period <strong>of</strong> time, ABEL devoted himself to explaining and<br />

elaborating the results <strong>of</strong> JACOBI within his own framework and this constituted a<br />

second topic in his research on elliptic functions.<br />

ABEL’S last approach to elliptic functions was the most general. Applying the<br />

theory which he developed in the Paris memoir concerning integration <strong>of</strong> algebraic<br />

differentials (see chapter 19) to the special case <strong>of</strong> elliptic functions, ABEL could sketch<br />

a very general approach to elliptic functions which — based on functions <strong>of</strong> the first<br />

kind — introduced all kinds <strong>of</strong> elliptic functions.<br />

<strong>The</strong>se aspects — which far from exhaust the discipline <strong>of</strong> elliptic and higher tran-<br />

scendentals in the early nineteenth century — will be addressed in the subsequent<br />

chapters. A complete description <strong>of</strong> the history <strong>of</strong> these transcendental objects is way<br />

outside the scope <strong>of</strong> the present work as their study was one <strong>of</strong> the most important<br />

and widely studied mathematical topics in the period. Instead, selections have been<br />

made to illustrate how new objects were being introduced and how new tools — and<br />

primarily algebraic tools — were being put to use in the investigation <strong>of</strong> these new<br />

objects.


Chapter 16<br />

<strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

As noted, N. H. ABEL (1802–1829) probably started developing his interest in ellip-<br />

tic integrals as a direct response to C. F. DEGEN’S (1766–1825) suggestion. Before his<br />

European tour, he was already well acquainted with the comprehensive treatment the<br />

theory had been given by A.-M. LEGENDRE (1752–1833) and he had begun develop-<br />

ing his own ideas. Soon, the theory <strong>of</strong> elliptic functions would occupy most <strong>of</strong> his<br />

resources. As already stated, ABEL’S first publication on the theory <strong>of</strong> elliptic tran-<br />

scendentals was his Recherches sur les fonctions elliptiques which appeared in two parts<br />

in A. L. CRELLE’S (1780–1855) Journal in 1827 and 1828. 1<br />

16.1 <strong>The</strong> importance <strong>of</strong> the lemniscate<br />

ABEL’S Recherches were designed to address particular questions pertaining to elliptic<br />

integrals <strong>of</strong> the first kind because these integrals had “the most remarkable and simple<br />

properties”. 2 Later, ABEL’S penetrating knowledge <strong>of</strong> elliptic integrals <strong>of</strong> the first kind<br />

would also allow him to attack all elliptic functions from a general perspective (see<br />

chapter 20).<br />

In the Recherches, ABEL studied integrals <strong>of</strong> the form<br />

�<br />

dx<br />

� (1 − c 2 x 2 ) (1 + e 2 x 2 )<br />

(16.1)<br />

which do not immediately belong to either <strong>of</strong> the kinds classified by LEGENDRE. ABEL<br />

argued that his choice <strong>of</strong> representation made the obtained formulae more simple and<br />

stressed that in (16.1), the integrand was more symmetric than in LEGENDRE’S stan-<br />

dard form.<br />

<strong>The</strong> simplicity <strong>of</strong> central formulae which ABEL emphasized is particularly clear in<br />

one specific application <strong>of</strong> the theory which ABEL developed in the second part <strong>of</strong> the<br />

Recherches. 3 <strong>The</strong>re, ABEL solved the division problem for the lemniscate integral which<br />

1 (N. H. <strong>Abel</strong>, 1827b; N. H. <strong>Abel</strong>, 1828b).<br />

2 (N. H. <strong>Abel</strong>, 1827b, 102).<br />

3 (N. H. <strong>Abel</strong>, 1828b).<br />

299


300 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

Figure 16.1: Stamp depicting Gauss and the construction <strong>of</strong> the regular 17-gon.<br />

is undoubtedly the simplest integral <strong>of</strong> the form (16.1) corresponding to e = c = 1. 4<br />

Thus, there is a suggestion that ABEL’S choice <strong>of</strong> representation <strong>of</strong> the integrals is a<br />

direct reflection <strong>of</strong> one <strong>of</strong> the main purposes <strong>of</strong> the Recherches, the adaption <strong>of</strong> C. F.<br />

GAUSS’ (1777–1855) methods from the Disquisitiones arithmeticae to the division <strong>of</strong> the<br />

lemniscate. 5<br />

16.2 Inversion in the Recherches<br />

<strong>The</strong> issue <strong>of</strong> CRELLE’S Journal which contained ABEL’S inversion <strong>of</strong> elliptic integrals<br />

into elliptic functions was published on 20 September 1827; 6 the date is <strong>of</strong> importance<br />

in analyzing the internal relations between ABEL’S and C. G. J. JACOBI’S (1804–1851)<br />

approaches (see below and section 18.1, below).<br />

In the introduction to the Recherches, ABEL described his idea:<br />

“In this memoir, I propose to study the inverse function, i.e. the function φα<br />

determined by the equations<br />

4 See (Glaisher, 1902).<br />

5 See also chapter 7.<br />

6 (N. H. <strong>Abel</strong>, 1881, II, 305).<br />

�<br />

dθ<br />

α = �<br />

1 − c2 sin 2 θ and<br />

sin θ = φ (α) = x.” 7


16.2. Inversion in the Recherches 301<br />

Thus, from ABEL’S own description <strong>of</strong> it, the idea <strong>of</strong> considering the inverse func-<br />

tions appears quite natural. However, by this simple step, an entirely new class <strong>of</strong><br />

functions was introduced, and they certainly looked different from anything known<br />

so far.<br />

With ABEL’S choice <strong>of</strong> representation <strong>of</strong> the integral, the inversion became<br />

α =<br />

� x<br />

0<br />

dx<br />

� (1 − c 2 x 2 ) (1 + e 2 x 2 ) � φ (α) = x. (16.2)<br />

First, ABEL introduced a special name for the integral from 0 to 1 c :<br />

“By thus letting<br />

ω<br />

2 =<br />

� 1<br />

c<br />

0<br />

∂x<br />

√ [(1 − c 2 x 2 ) (1 + e 2 x 2 )] ,<br />

it is evident that φ (α) is positive and increasing from α = 0 to α = ω<br />

2 .”8<br />

This remark seems to indicate that ABEL was well aware that for the inversion to be<br />

meaningful, the integral had to be a strictly monotonous function.<br />

ABEL’S next step consisted in the observation that the integral was an odd function<br />

<strong>of</strong> x, and thus φ (−α) = −φ (α). At this point, ABEL had thus obtained the function φ<br />

�<br />

.<br />

for a segment <strong>of</strong> the real axis � − ω 2 , ω 2<br />

16.2.1 Going complex<br />

ABEL’S study <strong>of</strong> the inverse functions <strong>of</strong> elliptic integrals relied importantly on the<br />

extension <strong>of</strong> these inverse functions to allow for imaginary and complex arguments.<br />

As discussed below, this aspect is extremely interesting in connection with the creation<br />

<strong>of</strong> a (rigorous) theory <strong>of</strong> complex integration.<br />

In analogy with the substitution <strong>of</strong> −x for x used above, ABEL observed:<br />

“By inserting into (1) xi instead <strong>of</strong> x (where i for short represents the imaginary<br />

quantity √ −1) and designating the value <strong>of</strong> α by βi, it gives<br />

xi = φ (βi) and β =<br />

� x<br />

0<br />

∂x<br />

√ [(1 + c 2 x 2 ) (1 − e 2 x 2 )] .” 9<br />

7 “Je me propose, dans ce mémoire, de considérer la fonction inverse, c’est-à-dire la fonction φα,<br />

déterminée par les équations<br />

(N. H. <strong>Abel</strong>, 1827b, 102).<br />

8 “En faisant donc<br />

�<br />

α =<br />

sin θ = φ (α) = x.”<br />

ω<br />

2 =<br />

� 1<br />

c<br />

0<br />

∂θ<br />

√ � 1 − c 2 sin 2 θ � et<br />

∂x<br />

√ [(1 − c 2 x 2 ) (1 + e 2 x 2 )] ,<br />

il est évident, que φα e[s]t positif et va en augmentant depuis α = 0 jusqu’à α = ω 2 [. . . ]” (ibid., 104).


302 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

<strong>The</strong> formal substitution <strong>of</strong> an imaginary value thus seemed to preserve the form<br />

<strong>of</strong> the function with the sole exception that the roles <strong>of</strong> the quantities c 2 and e 2 were<br />

interchanged. To ABEL this was fully sufficient, as he simple wrote:<br />

“thus, one sees by supposing c instead <strong>of</strong> e and e instead <strong>of</strong> c,<br />

φ (αi)<br />

i<br />

changes into φ (α) .” 10<br />

Thus, when I let φ (c,e) (α) denote the function in (16.2), ABEL’S formal imaginary<br />

substitution gave<br />

φ (c,e) (αi) = iφ (e,c) (α)<br />

and he had found the function φ (c,e) for a section <strong>of</strong> the imaginary axis � − ¯ω 2 , ¯ω �<br />

2 i, 11 in<br />

which<br />

¯ω<br />

2 =<br />

� 1<br />

e<br />

0<br />

16.2.2 Addition theorems<br />

dx<br />

� (1 + c 2 x 2 ) (1 − e 2 x 2 ) .<br />

Of central importance to ABEL’S approach to the inversion was the use which he made<br />

<strong>of</strong> addition formulae for elliptic functions.<br />

Auxiliary functions f and F. ABEL introduced two auxiliary functions which he<br />

named f and F, derived from φ (α), which played central parts in his deductions and<br />

were treated analogous to φ,<br />

f (α) =<br />

�<br />

1 − c2φ2 �<br />

(α) and F (α) = 1 + e2φ2 (α).<br />

Obviously, the product <strong>of</strong> these functions equals φ ′ (α) and the functions f and F were,<br />

themselves, doubly periodic functions corresponding to JACOBI’S cn and dn, respec-<br />

tively, which will be introduced and discussed later.<br />

9 “En mettant dans (1.) xi au lieu de x (ou i, pour abréger, représente la quantité imaginaire √ − 1) et<br />

désignant la valeur de α par βi, il viendra<br />

xi = φ (βi) et β =<br />

� x<br />

0<br />

∂x<br />

√ [(1 + c 2 x 2 ) (1 − e 2 x 2 )] .”<br />

(N. H. <strong>Abel</strong>, 1827b, 104).<br />

10 “[. . . ] donc on voit, qu’en supposant c au lieu de e et e au lieu de c,<br />

φ (αi)<br />

i<br />

se changera en φα.”<br />

(ibid., 104).<br />

11 I write [a, b] i as a short-hand for the segment <strong>of</strong> the imaginary axis which can also be written as<br />

{xi : x ∈ [a, b]}.


16.2. Inversion in the Recherches 303<br />

ABEL’S derivation <strong>of</strong> the addition formulae. ABEL’S way <strong>of</strong> obtaining addition for-<br />

mulae for elliptic functions resembles L. EULER’S (1707–1783) argument (see section<br />

15.2.1) because both proceeded from a suggested formula. In ABEL’S case, he sought<br />

to establish the identity<br />

φ (α + β) =<br />

φ (α) f (β) F (β) + φ (β) f (α) F (α)<br />

1 + e 2 c 2 φ 2 (α) φ 2 (β)<br />

and similar formulae for the auxiliary functions f and F,<br />

f (α + β) = f (α) f (β) − c2 φ (α) φ (β) F (α) F (β)<br />

1 + e 2 c 2 φ 2 (α) φ 2 (β)<br />

(16.3)<br />

and (16.4)<br />

F (α + β) = F (α) F (β) + e2φ (α) φ (β) f (α) f (β)<br />

1 + e2c2φ2 (α) φ2 . (16.5)<br />

(β)<br />

ABEL denoted the right hand side <strong>of</strong> (16.3) by r = r (α, β) and proceeded to dif-<br />

ferentiate r with respect to α. <strong>The</strong> expression which he obtained after inserting the<br />

values <strong>of</strong> f and F proved to be symmetric in α and β. <strong>The</strong>refore, and because r itself<br />

was symmetric in α and β, ABEL concluded that<br />

∂r<br />

∂α<br />

∂r<br />

= . (16.6)<br />

∂β<br />

This differential equation, ABEL claimed, 12 showed that r was a function <strong>of</strong> α + β,<br />

r = ψ (α + β) . (16.7)<br />

Upon inserting β = 0, ABEL immediately recognized ψ = φ, and the addition formula<br />

for φ had been obtained.<br />

To understand how ABEL concluded that the solution to the differential equation<br />

(16.6) must be <strong>of</strong> the form (16.7), we can get a hint from one <strong>of</strong> his earlier papers,<br />

published in the Journal. 13 In that paper, ABEL had established that the solution <strong>of</strong> the<br />

equation 14<br />

has the solution<br />

� �<br />

∂r<br />

σ (y) =<br />

∂x<br />

��<br />

r = ψ<br />

�<br />

σ (x) dx +<br />

� �<br />

∂r<br />

σ (x)<br />

∂y<br />

�<br />

σ (y) dy<br />

where ψ was arbitrary. 15 To apply to the situation <strong>of</strong> the addition formulae, take σ = 1<br />

to obtain (16.7).<br />

12 <strong>The</strong> validity <strong>of</strong> this claim will be discussed below.<br />

13 (N. H. <strong>Abel</strong>, 1826e).<br />

14 ABEL’S use <strong>of</strong> d has been replaced by ∂; and ABEL wrote φ where I have substituted σ.<br />

15 (ibid., 12–13).


304 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

Periods <strong>of</strong> φ. With the addition formulae in place, ABEL inserted β = ± ω 2 and β = ¯ω 2<br />

to find by direct computation<br />

�<br />

φ α ± ω<br />

�<br />

= ±φ<br />

2<br />

Similarly, for the auxiliary functions<br />

�<br />

f α ± ω<br />

�<br />

2<br />

�<br />

F α ± ω<br />

�<br />

2<br />

�<br />

ω<br />

� �<br />

f (α)<br />

and φ α ±<br />

2 F (α) ¯ω<br />

2 i<br />

� �<br />

¯ω<br />

= ±φ<br />

2 i<br />

�<br />

F (α)<br />

f (α) .<br />

�<br />

φ � φ (α)<br />

�<br />

ω and f<br />

F (α) 2 � �<br />

and F α ±<br />

F (α) ¯ω<br />

2 i<br />

�<br />

= ∓ F � ω 2<br />

= F � ω 2<br />

�<br />

α ± ¯ω<br />

2 i<br />

�<br />

= f � ¯ω 2 i �<br />

f (α) ;<br />

= ∓ f � ¯ω<br />

2 i �<br />

φ � φ (α)<br />

�<br />

¯ω<br />

2 i f (α) .<br />

When he combined these and inserted e.g. α = α + ω 2 and β = ω 2 , ABEL found<br />

�<br />

φ (α + ω) = φ α + ω<br />

2<br />

= φ<br />

�<br />

ω<br />

� −<br />

2<br />

F( ω 2 )<br />

φ( ω 2 )<br />

ω<br />

�<br />

+ = φ<br />

2<br />

F( ω 2 )<br />

F(α)<br />

φ(α)<br />

F(α)<br />

�<br />

ω<br />

�<br />

� �<br />

ω f α + 2<br />

2 F � α + ω �<br />

2<br />

= −φ (α) .<br />

In other words, φ (α + 2ω) = φ (α), and ABEL had discovered that 2ω was a period <strong>of</strong><br />

φ. Similarly, 2 ¯ωi was also found to be a period <strong>of</strong> φ.<br />

<strong>The</strong> value <strong>of</strong> φ for any complex value α + βi <strong>of</strong> its argument could thus be found,<br />

ABEL emphasized, from the values φ (α) , f (α) , F (α) and φ (iβ) , f (iβ) , F (iβ). Fur-<br />

thermore, if<br />

α + βi = � mω ± α ′� + � n ¯ω ± β ′� i<br />

such that α ′ ∈ � 0, ω � � �<br />

2 and β ′ ¯ω ∈ 0, 2 , the values <strong>of</strong> these six functions could be obtained<br />

from the values <strong>of</strong> φ (α ′ ) , f (α ′ ) , F (α ′ ) and φ (β ′ i) , f (β ′ i) , F (β ′ i) by formulae<br />

such as<br />

φ (α) = φ � mω ± α ′� = ± (−1) m φ � α ′� .<br />

Consequently, the value <strong>of</strong> φ (and <strong>of</strong> f and F) at any complex argument was determined<br />

by the values <strong>of</strong> φ (α) (and f and F) in which α ∈ � 0, ω � � �<br />

¯ω<br />

2 or α ∈ 0, 2 i.<br />

ABEL’S extension <strong>of</strong> the elliptic function φ to the entire complex plane may thus be<br />

summarized in the following steps (see figure 16.2):<br />

1. <strong>The</strong> elliptic function φ (α) was obtained by inversion <strong>of</strong> the elliptic integral on a<br />

�<br />

. Because the function was odd, it was simultane-<br />

segment <strong>of</strong> the real axis � 0, ω 2<br />

ously found for α ∈ � − ω 2 , 0� .<br />

2. By a formal, imaginary substitution the function φ (iβ) was found for β ∈ � 0, ¯ω �<br />

2 i<br />

and consequently for β ∈ � − ¯ω 2 , 0� i. <strong>The</strong> value <strong>of</strong> φ (c,e) (iβ) was obtained from<br />

the inversion <strong>of</strong> a related elliptic function φ (e,c) (β) on a segment <strong>of</strong> the real axis.


16.2. Inversion in the Recherches 305<br />

¡ !<br />

2<br />

¹!<br />

2 i<br />

¡ ¹!<br />

2 i<br />

Figure 16.2: ABEL’S extension to the complex rectangle<br />

3. <strong>The</strong> important addition formulae was demonstrated by differentiation.<br />

4. <strong>The</strong> two periods <strong>of</strong> φ, 2ω and 2 ¯ω, were direct results <strong>of</strong> the addition formulae<br />

which also provided a way <strong>of</strong> reducing φ (α + iβ) to the values φ (α ′ ) and φ (β ′ i)<br />

in which α ′ ∈ � 0, ω � � �<br />

2 and β ′ ¯ω ∈ 0, 2 i.<br />

Zeros and poles <strong>of</strong> φ. Solution <strong>of</strong> φ (x) = φ (a). After having established the ad-<br />

dition formulae, ABEL proceeded to investigate the singular points <strong>of</strong> φ, i.e. its zeros<br />

and poles. He found that every zero <strong>of</strong> φ was <strong>of</strong> the form<br />

mω + n ¯ωi for m, n ∈ Z<br />

and that every pole <strong>of</strong> φ was <strong>of</strong> the form<br />

�<br />

m + 1<br />

� �<br />

ω + n +<br />

2<br />

1<br />

�<br />

¯ωi for m, n ∈ Z.<br />

2<br />

ABEL applied a formula — which he had derived directly from the addition formu-<br />

lae — to the equation<br />

φ (x) − φ (y) = 0<br />

and concluded that the complete solution to this equation was <strong>of</strong> the form<br />

x = (−1) m+n y + mω + n ¯ωi for m, n ∈ Z. (16.8)<br />

This determination <strong>of</strong> all the roots <strong>of</strong> the (transcendental) equation φ (x) − φ (a) = 0<br />

would soon become very important for ABEL’S main objective, the solution <strong>of</strong> the<br />

division problem (see below).<br />

C<br />

!<br />

2


306 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

16.2.3 <strong>The</strong> question <strong>of</strong> complex integration<br />

Before going into the division problem, we pause to discuss and comment on ABEL’S<br />

inversion and extension <strong>of</strong> elliptic functions to include complex variables.<br />

In a letter to B. M. HOLMBOE (1795–1850), ABEL praised A.-L. CAUCHY (1789–<br />

1857) highly and described how he had struggled to understand the nine issues <strong>of</strong> the<br />

Exercises de mathématiques which he had bought and studied:<br />

“Cauchy is mad and nothing is to be obtained from him although at present,<br />

he is the mathematician who knows how mathematics should be conducted. His<br />

things are excellent but he writes very obscurely. At first, I almost did not understand<br />

a thing <strong>of</strong> his works but now it is better. He is having a series <strong>of</strong> papers<br />

printed under the title Exercises des mathématiques. I buy them and read them carefully.<br />

Nine issues have appeared since the beginning <strong>of</strong> this year.” 16<br />

A large part <strong>of</strong> CAUCHY’S Exercises de mathématiques from the year 1826 concerns<br />

the introduction <strong>of</strong> CAUCHY’S revolutionary new idea <strong>of</strong> residues. <strong>The</strong> year before, in<br />

1825, CAUCHY had laid the foundation for his theory <strong>of</strong> complex integration with a<br />

brochure entitled Mémoire sur les intégrales définies, prises entre des limites imaginaries. 17<br />

Although ABEL never referred directly to the memoir, P. L. M. SYLOW (1832–1918)<br />

has found evidence in certain calculations in one <strong>of</strong> ABEL’S notebooks from the time<br />

in Paris that ABEL knew <strong>of</strong> it. 18<br />

ABEL’S concept <strong>of</strong> functions. In the Recherches, ABEL spoke <strong>of</strong> finding the values <strong>of</strong><br />

the function φ (α) for given α. Thus, ABEL’S deductions do not seem to be defining<br />

this function but rather to be manipulations which make the value <strong>of</strong> the function<br />

available to the observer. This concept <strong>of</strong> functions resembles EULER’S (see section<br />

10.1) in the sense that the function is tacitly supposed to exist for all values <strong>of</strong> the<br />

variable although it is only strictly meaningful for a subset <strong>of</strong> the arguments, in this<br />

case a segment <strong>of</strong> the real axis. 19<br />

ABEL on integration between imaginary limits. As noted, SYLOW found some sug-<br />

gestion in ABEL’S Notebook A — which dates from 1826 — that ABEL actually thought<br />

<strong>of</strong> his elliptic functions as defined by complex integration. For instance, one finds in<br />

that notebook the formula<br />

f (x + yi) =<br />

� (x+yi)<br />

0<br />

dp<br />

� (1 − p 2 ) (1 − c 2 p 2 )<br />

16 “Cauchy er fou, og der er ingen Udkomme med ham, omendskjøndt han er den Mathematiker som<br />

for nærværende Tid veed hvorledes Mathematiken skal behandles. Hans Sager ere fortræffelige<br />

men han skriver meget utydelig. I Førstningen forstod jeg næsten ikke et Gran af hans Arbeider nu<br />

gaar det bedre. Han lader nu trykke en Række Afhandlinger under titel Exercises des Mathematiques.<br />

Jeg kjøber og læser dem flittig. 9 Hefter ere udkomne fra dette Aars Begyndelse.” (<strong>Abel</strong>→Holmboe,<br />

Paris, 1826/10/24. N. H. <strong>Abel</strong>, 1902a, 43).<br />

17 (A.-L. Cauchy, 1825).<br />

18 (N. H. <strong>Abel</strong>, 1881, II, 284).<br />

19 See also (Nørgaard, 1990).


16.2. Inversion in the Recherches 307<br />

which is nowhere found in the published version in the Recherches. 20 Later in the same<br />

notebook, ABEL wrote<br />

and deduced the differential equations<br />

� �<br />

dp<br />

=<br />

dx<br />

�<br />

φ x + y √ �<br />

−1 = p + q √ −1<br />

� �<br />

dq<br />

dy<br />

and<br />

� �<br />

dp<br />

= −<br />

dy<br />

� �<br />

dq<br />

dx<br />

which are the important Cauchy-Riemann equations. 21 Thus, as SYLOW concludes, 22<br />

there is good reason to believe that ABEL had studied CAUCHY’S works on integra-<br />

tion between imaginary limits. However, there is still no direct indication that ABEL<br />

allowed any <strong>of</strong> these studies or considerations to have an impact on the way he pre-<br />

sented his inversion <strong>of</strong> elliptic integrals.<br />

Complex integration or formal substitution in the Recherches? As the evidence<br />

seems to be inconclusive, the interpretation <strong>of</strong> ABEL’S inversion must be left to the his-<br />

torian and depends on the temper <strong>of</strong> the interpretor. I believe that ABEL’S inversion<br />

was formal in the sense that he employed a formal, imaginary substitution to obtain<br />

the extension to imaginary arguments. Whether or not, he found any reassurance <strong>of</strong><br />

his method in CAUCHY’S theory <strong>of</strong> integration remains an undecidable question.<br />

16.2.4 GAUSS’ unpublished results on lemniscate functions<br />

<strong>The</strong> idea <strong>of</strong> inverting elliptic integrals into elliptic functions did not belong uniquely<br />

to ABEL. Actually, contrary to beliefs expressed throughout the secondary literature,<br />

the idea had occurred to LEGENDRE. 23 What LEGENDRE did not fully realize, though,<br />

was that the inverted functions should most naturally be considered as functions <strong>of</strong> a<br />

complex variable. This idea is most frequently attributed to GAUSS in whose drawer it<br />

remained, however. We may learn a bit more <strong>of</strong> the idea <strong>of</strong> inverting elliptic integrals<br />

by considering extracts from GAUSS’ unpublished works and by comparing with the<br />

approach taken by JACOBI after ABEL’S inversion had been published.<br />

In what appears to be GAUSS’ first manuscript on the lemniscate function, we get<br />

an impression <strong>of</strong> his approach. GAUSS wrote:<br />

“We designate the value <strong>of</strong> the integral from x = 0 to x = 1 by 1<br />

2 ¯ω. We denote<br />

the variable x <strong>of</strong> the respective integral by the sign sin lemn and its complementary<br />

integral to 1<br />

2 ¯ω by cos lemn. Thus,<br />

�<br />

� � �<br />

dx<br />

1 dx<br />

sin lemn √ = x, cos lemn ¯ω − √ = x.”<br />

1 − x4 2 1 − x4 24<br />

20 (<strong>Abel</strong>, MS:351:A, 64).<br />

21 (ibid., 100).<br />

22 (N. H. <strong>Abel</strong>, 1881, II, 284).<br />

23 See (Krazer, 1909, 55) and (J. J. Gray, 1984, 103).


308 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

Thus, the function sin lemnα produces the upper limit <strong>of</strong> the lemniscate integral<br />

whose value is α. This is the inverse function <strong>of</strong> the lemniscate integral and the direct<br />

counterpart to (special case <strong>of</strong>) the elliptic function φ which ABEL later, independently,<br />

introduced in his Recherches.<br />

Was GAUSS’ sin lemn a complex function? Clearly, GAUSS had taken the step <strong>of</strong><br />

considering the inverse function <strong>of</strong> the lemniscate integral. He invested extensive<br />

effort in developing representations by infinite series, infinite products, and ratios <strong>of</strong><br />

infinite series. In the literature, GAUSS is universally credited with the discovery <strong>of</strong> the<br />

doubly periodic nature <strong>of</strong> the lemniscate function. 25 This claim is generally supported<br />

by GAUSS’ consideration <strong>of</strong> the degree <strong>of</strong> the division problem for the lemniscate.<br />

GAUSS found, and noted in his diary, that the division problem for the lemniscate into<br />

n parts led to an equation <strong>of</strong> degree n 2 . 26 This may well have been GAUSS’ motivation<br />

for considering complex values <strong>of</strong> the argument. In a manuscript, in which GAUSS<br />

wrote the lemniscate function as the ratio <strong>of</strong> two infinite products<br />

he stated formulae such as<br />

�<br />

4√<br />

2P φ + 1<br />

2 ω<br />

�<br />

which amounted to<br />

sin lemnφ =<br />

P (φ)<br />

Q (φ) ,<br />

�<br />

= pφ and p φ + 1<br />

2 ω<br />

�<br />

= − 4√ 2P (φ)<br />

P (φ + ω) = 1<br />

�<br />

4√ p φ +<br />

2 1<br />

2 ω<br />

�<br />

= −P (φ) .<br />

This demonstrated the periodic nature <strong>of</strong> P and a similar result was obtained for Q.<br />

More interestingly, GAUSS also wrote<br />

which would indicate that<br />

P (iψω) = ie πψ2<br />

P (ψω) and Q (iψω) = e πψ2<br />

Q (ψω)<br />

sin lemn (iψ) = i sin lemn (ψ)<br />

and therefore produce the second period <strong>of</strong> sin lemnφ. GAUSS’ manuscripts also con-<br />

tain numerous formulae expressing the addition and multiplication <strong>of</strong> the lemniscate<br />

function. 27<br />

24 “Valorem huius integralis ab x = 0 usque ad x = 1 semper per 1 2 ¯ω designamus. Variabilem x<br />

respectu integralis per signum sin lemn denotamus, respectu vero complementi integralis ad 1 2 ¯ω<br />

per cos lemn. Ita ut<br />

�<br />

sin lemn<br />

dx<br />

√ = x,<br />

1 − x4 � �<br />

1<br />

cos lemn ¯ω −<br />

2<br />

�<br />

dx<br />

√ = x.”<br />

1 − x4 (C. F. Gauss, 1863–1933, III, 404).<br />

25 (Schlesinger, 1922–1933) and see (J. J. Gray, 1984, 102–103).<br />

26 (ibid., 102).<br />

27 [Ref]


16.2. Inversion in the Recherches 309<br />

16.2.5 JACOBI’s inversion in the Fundamenta nova<br />

As will be described in section 18.1, a third inversion <strong>of</strong> elliptic integrals was per-<br />

formed by CARL GUSTAV JACOB JACOBI who in 1829 published the first book entirely<br />

devoted to the study <strong>of</strong> the new elliptic functions. 28 As will also be illustrated in<br />

section 18.1, JACOBI’S main objective with his research on elliptic integrals and func-<br />

tions was the development <strong>of</strong> transformation theory. After having devised the first set<br />

<strong>of</strong> theorems concerning the transformation <strong>of</strong> elliptic integrals, JACOBI presented his<br />

version <strong>of</strong> the inversion:<br />

“Letting � φ<br />

0<br />

dφ<br />

√ 1−k 2 sin 2 φ = u, geometers have accustomed themselves to call<br />

the angle φ the amplitude <strong>of</strong> the function u. In the following, this angle is denoted<br />

by amplu or shorter by<br />

Thus, if<br />

then<br />

φ = amu.<br />

u =<br />

� x<br />

0<br />

dx<br />

� (1 − x 2 ) (1 − k 2 x 2 )<br />

x = sin amu.” 29<br />

JACOBI then introduced the complete integrals already stressed by LEGENDRE,<br />

K =<br />

K ′ =<br />

28 (C. G. J. Jacobi, 1829).<br />

29 “Posito � φ<br />

0<br />

� 1<br />

0<br />

� π 2<br />

0<br />

� π<br />

dx<br />

2<br />

� =<br />

(1 − x2 ) (1 − k2x2 ) 0<br />

�<br />

dφ<br />

1 − k ′ k ′ sin2 φ<br />

where k ′ k ′ + kk = 1.<br />

�<br />

dφ<br />

1 − k2 sin2 φ<br />

and<br />

dφ<br />

√ 1−k 2 sin 2 φ = u, angulum φ amplitudinem functionis u vocare geometrae consueverunt.<br />

Hunc igitur angulum in sequentibus denotabimus per amplu seu brevius per:<br />

Ita, ubi<br />

erit:<br />

(ibid., 81).<br />

φ = amu.<br />

u =<br />

� x<br />

0<br />

x = sin amu.”<br />

dx<br />

� (1 − x 2 ) (1 − k 2 x 2 ) ,


310 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

Next, JACOBI stated the addition formulae which were presented as well known re-<br />

sults concerning elliptic integrals. Only then were complex values <strong>of</strong> the variable in-<br />

troduced through a substitution sin φ = i tan ψ in the integrand:<br />

�<br />

dφ<br />

1 − k2 sin2 = �<br />

i dψ<br />

φ cos2 ψ + k2 sin2 = �<br />

i dψ<br />

ψ 1 − k ′ k ′ sin2 .<br />

ψ<br />

Finally, JACOBI obtained the doubly periodic nature <strong>of</strong> the function sin amu from the<br />

addition formulae.<br />

As described, JACOBI’S inversion is quite similar to ABEL’S approach. Based on<br />

two different elliptic integrals (corresponding to the complementary moduli k and k ′ ),<br />

JACOBI could obtain the value <strong>of</strong> sin amiu on the imaginary axis. <strong>The</strong>n, by the addition<br />

formulae which were apparently assumed to be valid for these complex values <strong>of</strong> u<br />

and v, the two independent periods were deduced. JACOBI was aware that the doubly<br />

periodic nature was a new and important feature <strong>of</strong> these new functions:<br />

“elliptic functions have two periods, one real and one imaginary whenever<br />

the modulus k is real. Both [periods] will be imaginary when the modulus itself is<br />

imaginary. We call this the principle <strong>of</strong> double periodicities.” 30<br />

JACOBI’S book became the corner stone <strong>of</strong> the research on elliptic functions in the<br />

following generation, and his notation and ways <strong>of</strong> introducing elliptic functions be-<br />

came standard for a while until he changed it by introducing elliptic functions by<br />

certain infinite series (see chapter 20). In that respect, JACOBI’S works surpassed LEG-<br />

ENDRE’S effort to update his monographs with the newest developments by ABEL and<br />

JACOBI which resulted in a supplement to his Traité des fonctions elliptiques published<br />

in 1828. 31<br />

16.2.6 Comparison: An earlier idea on inversion<br />

As indicated, ABEL’S inversion in the Recherches was the first inversion <strong>of</strong> elliptic inte-<br />

grals into elliptic functions <strong>of</strong> a complex variable to appear in print. However, prior to<br />

his departure on the European tour, ABEL had written a manuscript which also dealt<br />

with the inversion <strong>of</strong> functions and which is interesting in the discussion <strong>of</strong> whether<br />

ABEL used complex integration or not.<br />

<strong>The</strong> result. In a manuscript which bears the lengthy but very accurate title Propriétés<br />

remarquables de la fonction y = φx déterminée par l’équation f y.dx − f x � (a − y) (a1 − y) (a2 − y) . . . (am<br />

0, f étant une fonction quelconque de y qui ne devient pas nulle ou infinie lorsque y =<br />

30 “functiones ellipticas duplici gaudere periodo, altera reali, altera imaginaria, siquidem modulus k<br />

est realis. Utraque fit imaginaria, ubi modulus et ipse est imaginarius. Quod principium duplicis<br />

periodi nuncupabimus.” (C. G. J. Jacobi, 1829, 87).<br />

31 (A. M. Legendre, 1825–1828, III).


16.2. Inversion in the Recherches 311<br />

a, a1, a2, . . . , am, 32 ABEL considered a problem also bearing on the inversion <strong>of</strong> ellip-<br />

tic integrals. <strong>The</strong>re, he studied the function y = φ (x) given by a differential equation<br />

(which occurs in the title <strong>of</strong> the paper)<br />

dy<br />

dx =<br />

�<br />

ψ (y)<br />

, in which ψ (y) =<br />

f (y)<br />

m<br />

∏ (ak − y)<br />

k=1<br />

and f (a1) , . . . , f (am) were finite and non-zero. Of such functions φ (x), ABEL proved<br />

that they have m(m−1)<br />

2<br />

(possibly non-distinct) periods 2 (α k − αm) determined by<br />

� ak<br />

αk =<br />

0<br />

f (y) dy<br />

� ψ (y) .<br />

One <strong>of</strong> ABEL’S applications <strong>of</strong> the result. To recognize the connection with the in-<br />

version <strong>of</strong> elliptic integrals, consider first the case (given by ABEL) <strong>of</strong> trigonometric<br />

functions,<br />

i.e. � �<br />

f (y) dy<br />

� =<br />

ψ (y)<br />

This gave<br />

f (y) = 1 and ψ (y) = (1 − y) (1 + y)<br />

α1 =<br />

� 1<br />

0<br />

dy<br />

� 1 − y 2<br />

dy<br />

� 1 − y 2<br />

= arcsin y.<br />

= π<br />

2 and α2 = −π<br />

2 ,<br />

φ (x + 2nπ) = φ (x) .<br />

A speculative application <strong>of</strong> the same result. <strong>The</strong>re is no explicit restrictions on the<br />

roots a1, . . . , am mentioned by ABEL. If we, extending ABEL’S example, allow the roots<br />

to be imaginary and consider the lemniscate integral<br />

we find periods<br />

f (y) = 1 and ψ (y) = 1 − y 4 = (±1 − y) (±i − y) ,<br />

� 1 dy<br />

α1 = −α2 = � , and<br />

0 1 − y4 � i dy<br />

α3 = −α4 = �<br />

0 1 − y4 = iα1.<br />

In the last integral, the imaginary integration could be performed via the formal sub-<br />

stitution iy = z which ABEL employed in the Recherches. 33 Thus, the two periods <strong>of</strong><br />

the elliptic functions were immediate generalizations <strong>of</strong> the results obtained in the<br />

manuscript.<br />

32 (N. H. <strong>Abel</strong>, [1825] 1839a).<br />

33 (N. H. <strong>Abel</strong>, 1827b, 104). See above.


312 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

ABEL’S deduction. ABEL’S way <strong>of</strong> obtaining the described results was to expand the<br />

function φ in a Taylor series<br />

φ (x + v) =φ (x)<br />

� �� �<br />

=y<br />

+<br />

∞<br />

∑ v<br />

k=1<br />

2k Q2k +<br />

�<br />

ψ (y)<br />

With y = a k and α k equal to the corresponding value <strong>of</strong> x,<br />

ABEL found<br />

� ak<br />

αk =<br />

0<br />

φ (α k + v) = a k +<br />

f (y) dy<br />

� ψ (y) ,<br />

∞<br />

∑<br />

k=0<br />

∞<br />

∑ v<br />

k=1<br />

2k Q2k and thus in this case, φ (α k + v) was an even function <strong>of</strong> v,<br />

By inserting v ′ = α k − v, ABEL obtained<br />

<strong>The</strong>refore,<br />

φ (α k + v) = φ (α k − v) .<br />

φ � 2α k − v ′� = φ � v ′� .<br />

φ (2α k − 2αm + v) = φ (v) ,<br />

and the function was therefore periodic. In general<br />

�<br />

�<br />

φ<br />

v + 2<br />

m<br />

∑<br />

k,k ′ =1<br />

n k,k ′ (α k − α k ′)<br />

In particular, by taking for v a zero <strong>of</strong> φ, the values<br />

were also zeros <strong>of</strong> φ.<br />

v + 2<br />

16.2.7 Conclusion<br />

m<br />

∑<br />

k=1<br />

n kα k for n1, . . . , nm ∈ Z with<br />

= φ (v) .<br />

v 2k+1 Q 2k+1.<br />

m<br />

∑ nk = 0<br />

k=1<br />

Because ABEL’S general inversion — which admittedly did not explicitly concern el-<br />

liptic integrals — was written before he embarked on the European tour in 1825, it<br />

cannot rely on any knowledge <strong>of</strong> the new theory <strong>of</strong> complex integration which was<br />

presented that same year. Also admittedly, the manuscript does not contain any com-<br />

plex integrals or complex periods but I find the suggested application <strong>of</strong> the result<br />

plausible. I do so, because I read ABEL’S inversion <strong>of</strong> elliptic integrals in the Recherches<br />

rather literally and see in it a formal substitution without any justification in complex<br />

integration. This theme will surface again when the need and means <strong>of</strong> representa-<br />

tions for elliptic functions are discussed in chapter 17.


16.3. <strong>The</strong> division problem 313<br />

16.3 <strong>The</strong> division problem<br />

As already indicated, one <strong>of</strong> the main objectives <strong>of</strong> ABEL’S Recherches was the so-called<br />

division problems which encompass deducing and solving the equations which deter-<br />

mine the division <strong>of</strong> the function φ (mα) into m parts, i.e. the equations which deter-<br />

mine φ (α) from φ (mα).<br />

ABEL’S starting point for these investigations was the addition formula (16.3).<br />

Based on these formula, he found expressions such as<br />

φ ((n + 1) β) = −φ ((n − 1) β) +<br />

2φ (nβ) f (β) F (β)<br />

1 + c 2 e 2 φ 2 (nβ) φ 2 (β) .<br />

Consequently, ABEL observed that the functions φ (nβ), f (nβ), and F (nβ) depended<br />

rationally on φ (β), f (β), and F (β) and he wrote, e.g.<br />

φ (nβ) = Pn<br />

Qn<br />

in which Pn and Qn were polynomial functions <strong>of</strong> φ (β), f (β), and F (β). ABEL then let<br />

x = φ (α), y = f (β), z = F (β) and manipulated the equation to obtain the relations<br />

in which<br />

Qn+1 = Qn−1Rn and<br />

Pn+1 = −Pn−1Rn + 2yzPnQnQn−1<br />

Rn = Q 2 n + e 2 c 2 x 2 P 2 n.<br />

Obviously, Rn was a polynomial function in x 2 , and from the basic formulae<br />

i.e.<br />

φ (β) = P1<br />

Q1<br />

and φ (2β) =<br />

2φ (β) f (β) F (β)<br />

1 + e 2 c 2 φ 4 (β)<br />

Q0 = 1, Q1 = 1, P0 = 0, P1 = x,<br />

P2<br />

= ,<br />

Q2<br />

ABEL found that Qn was always an entire function <strong>of</strong> x2 . Furthermore, P2n<br />

xyz and P2n+1 x<br />

were also entire functions <strong>of</strong> x2 . ABEL merely provided the first few particular cases<br />

but the argument is easily completed by induction:<br />

P2n+2 −P2n<br />

=<br />

xyz xyz Q2 �<br />

P2n+1 c2e2x 2P2nP2n+1 2n+1 −<br />

x yz<br />

= −P2n<br />

xyz Q2 P2n+1<br />

2n+1 −<br />

x<br />

+ 2Q2n+1Q2n<br />

�<br />

c 2 e 2 x 2 2 P2n P2n+1<br />

× x + 2Q2n+1Q2n<br />

xyz x<br />

and all parts are seen to be entire functions <strong>of</strong> x 2 by the induction hypothesis. Simi-<br />

larly,<br />

P2n+1<br />

x<br />

−P2n−1<br />

= R2n +<br />

x<br />

2yzP2nQ2nQ2n−1<br />

x<br />

= −P2n−1<br />

R2n + 2y<br />

x<br />

2 2 P2n<br />

z<br />

xyz Q2nQ2n−1<br />

and the same conclusion holds because y 2 = 1 − c 2 x 2 and z 2 = 1 + e 2 x 2 .<br />

�<br />


314 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

Bisection by algebraic means. <strong>The</strong> bisection <strong>of</strong> the function φ was the easiest example<br />

<strong>of</strong> the program which ABEL was developing. In order to express x = φ � �<br />

α<br />

2 by<br />

φ (α), ABEL employed the addition formula for f and F (16.4 and 16.5) to get<br />

�<br />

α α<br />

�<br />

f (α) = f + =<br />

2 2<br />

y2 − c2x2z2 1 + e2c2 �<br />

1 − c2x2 =<br />

x4 � − c2x2 � 1 + e2x2� 1 + e2c2x 4<br />

and<br />

�<br />

α α<br />

�<br />

F (α) = F + =<br />

2 2<br />

z2 + e2x2y2 1 + e2c2 �<br />

1 + e2x2 =<br />

x4 � + e2x2 � 1 − c2x2� 1 + e2c2x 4<br />

.<br />

When ABEL formed the ratio<br />

he found that<br />

F (α) − 1<br />

1 + f (α) =<br />

φ<br />

2e 2 x 2 (1−c 2 x 2 )<br />

1+e 2 c 2 x 4<br />

2(1−c 2 x 2 )<br />

1+e 2 c 2 x 4<br />

�<br />

α<br />

�<br />

=<br />

2<br />

1<br />

�<br />

F (α) − 1<br />

e 1 + f (α) .<br />

= e 2 x 2 ,<br />

Thus, ABEL had proved that if the value <strong>of</strong> φ (α) was known, φ � �<br />

α<br />

2 could be expressed<br />

algebraically and actually using only the extraction <strong>of</strong> square roots.<br />

Division into an odd number <strong>of</strong> parts. In order to complete the program, ABEL had<br />

to find a way <strong>of</strong> algebraically obtaining φ � �<br />

α<br />

2n+1 from φ (α) and this proved both more<br />

difficult and much more fruitful. In chapter 7, we have already met the Mémoire sur<br />

une classe particulière where ABEL later introduced <strong>Abel</strong>ian equations as generalizations<br />

<strong>of</strong> these investigations. 34<br />

ABEL wanted to solve the equation which he wrote as<br />

φ (α) = P2n+1<br />

Q2n+1<br />

and similar equations for f (α) and F (α). Although ABEL’S argument was greatly<br />

simplified by his slightly more general approach <strong>of</strong> the memoir on <strong>Abel</strong>ian equations,<br />

the original version <strong>of</strong> the Recherches is worth a brief description to facilitate a com-<br />

parison. I have suppressed most <strong>of</strong> the computational technicalities in the following<br />

typical step <strong>of</strong> the pro<strong>of</strong>.<br />

ABEL let θ denote an imaginary (2n + 1)’th root <strong>of</strong> unity and introduced three aux-<br />

iliary functions,<br />

ψ (β) =<br />

34 (N. H. <strong>Abel</strong>, 1829c).<br />

n<br />

∑ θ<br />

µ=−n<br />

µ �<br />

φ1 β +<br />

φ1 (β) =<br />

2µ ¯ωi<br />

2n + 1<br />

n<br />

∑<br />

m=−n<br />

�<br />

φ β + 2mω<br />

�<br />

,<br />

2n + 1<br />

�<br />

, and ψ1 (β) =<br />

n<br />

∑ θ<br />

µ=−n<br />

µ φ1<br />

�<br />

β −<br />

�<br />

2µ ¯ωi<br />

.<br />

2n + 1


16.3. <strong>The</strong> division problem 315<br />

Based on a direct application <strong>of</strong> the addition formulae, ABEL obtained an expression<br />

� �<br />

2µ ¯ωi<br />

φ1 β ± = Rµ ± R<br />

2n + 1<br />

′ �<br />

µ (1 − c2x2 ) (1 + e2x2 )<br />

in which Rµ and R ′ µ were rational functions <strong>of</strong> x = φ (β). Consequently, ABEL found<br />

that both the functions<br />

ψ (β) ψ1 (β) = λ (β) and ψ (β) 2n+1 + ψ1 (β) 2n+1 = λ1 (β) (16.9)<br />

were rational in x. By direct calculation, he also found that both functions (16.9) were<br />

invariant if another root β + kω+k′ ¯ωi<br />

2n+1<br />

<strong>of</strong> the equation<br />

φ ((2n + 1) β) = P2n+1<br />

Q2n+1<br />

(16.10)<br />

was substituted for β. Thus, ABEL knew that λ (β) and λ1 (β) were rational in the<br />

coefficients <strong>of</strong> (16.10), in particular in the quantity φ ((2n + 1) β). When he solved the<br />

system <strong>of</strong> equations (16.9), ABEL found<br />

ψ (β) = 2n+1<br />

�<br />

�<br />

�<br />

�λ1 (β)<br />

2 +<br />

�<br />

λ1 (β) 2<br />

4<br />

− λ (β). (16.11)<br />

From these, ABEL obtained φ1 (β) and then φ (β) by similar arguments.<br />

However, as ABEL observed, the formula for φ (β) which he had obtained also<br />

contained the quantities<br />

� �<br />

ω<br />

φ<br />

2n + 1<br />

� �<br />

¯ωi<br />

and φ .<br />

2n + 1<br />

Thus, in order to completely solve the problem, these two quantities should also be<br />

determined and ABEL demonstrated how the equation P2n+1 = 0 which determined<br />

these could be reduced to equations <strong>of</strong> lower degrees, one <strong>of</strong> degree 2n + 2 and 2n + 2<br />

equations <strong>of</strong> degree n. Furthermore, ABEL also proved that the equations <strong>of</strong> degree<br />

n were always solvable by radicals. In the process, ABEL employed tools similar to<br />

those described above as well as some knowledge <strong>of</strong> primitive roots <strong>of</strong> an integer. Im-<br />

portantly, ABEL knew qualitatively how the roots were interrelated (by 16.8) and used<br />

this knowledge to investigate the system <strong>of</strong> roots and prove its reduction to equations<br />

<strong>of</strong> lower degree, some <strong>of</strong> which were proved to be solvable by radicals.<br />

16.3.1 Division <strong>of</strong> the lemniscate<br />

<strong>The</strong> culmination <strong>of</strong> ABEL’S research into the division problem was his application <strong>of</strong><br />

the theory to the case <strong>of</strong> the lemniscate. <strong>The</strong> symmetry <strong>of</strong> ABEL’S representation <strong>of</strong><br />

the elliptic integrals became evident when he chose e = c = 1 to obtain the lemniscate<br />

integral<br />

φ (α) = x, α =<br />

� x<br />

0<br />

dx<br />

√ 1 − x 4 .


316 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

In light <strong>of</strong> the previous result, ABEL’S investigations mainly concerned the division<br />

<strong>of</strong> the complete integral into an odd number <strong>of</strong> equal segments. He did so by first<br />

carrying out the division <strong>of</strong> the complete integral into 4v + 1 parts. ABEL’S argument<br />

was explicitly designed to make the case accessible with the Gaussian approach to the<br />

solution <strong>of</strong> cyclotomic equations. A brief outline <strong>of</strong> ABEL’S reasoning will provide a<br />

few interesting comparisons with GAUSS’ approach and the more general solution <strong>of</strong><br />

the problem found in ABEL’S paper on <strong>Abel</strong>ian equations.<br />

ABEL first assumed that 4v + 1 was the sum <strong>of</strong> two squares, 4v + 1 = α 2 + β 2 and<br />

found that α + β had to be an odd integer. In this case, he found an equation which he<br />

wrote as<br />

φ ((α + βi) δ) = x T<br />

S<br />

(16.12)<br />

with α = mδ, β = µδ, x = φ (δ), and T and S two entire functions <strong>of</strong> x4 . ABEL’S<br />

real objective was the considerations pertaining to δ = ω<br />

�<br />

α+βi for which he obviously<br />

found that x = φ (δ) = φ was a root <strong>of</strong> the equation T = 0. It thus became his<br />

� ω<br />

α+βi<br />

objective to solve this equation.<br />

Expressing the roots <strong>of</strong> T = 0. First, by his very powerful determination <strong>of</strong> the roots<br />

<strong>of</strong> φ (ξ) = 0, ABEL found that all the roots <strong>of</strong> T = 0 were related by<br />

(α + βi) δ = mω + µ ¯ωi = (m + µi) ω<br />

since ω = ¯ω for the lemniscate integral. Thus, any root was contained in the formula<br />

�<br />

m + µi<br />

x = φ<br />

α + βi ω<br />

�<br />

if m and µ were allowed to assume all integral values. However, in order to count<br />

each root only once, ABEL found that the set <strong>of</strong> roots <strong>of</strong> T = 0 could be listed as<br />

� �<br />

ρω<br />

φ<br />

α + βi<br />

for − α2 + β 2 − 1<br />

2<br />

≤ ρ ≤ α2 + β 2 − 1<br />

2<br />

(16.13)<br />

and he argued by an application <strong>of</strong> the Euclidean algorithm for integers: ABEL let λ, λ ′<br />

be determined by the equation<br />

and t denote an integer in order to obtain<br />

αλ ′ − βλ = 1,<br />

µ + β (µλ + tα) −α<br />

� �� �<br />

=k<br />

� µλ ′ + tβ �<br />

� �� �<br />

=k ′<br />

= 0.<br />

<strong>The</strong>n, with ρ = m + αk − βk ′ , ABEL obtained<br />

m + µi<br />

α + βi<br />

= ρ<br />

α + βi − k − k′ i


16.3. <strong>The</strong> division problem 317<br />

and therefore,<br />

�<br />

m + µi<br />

φ<br />

α + βi ω<br />

�<br />

Because <strong>of</strong> the relation<br />

the bounds <strong>of</strong> (16.13) were obtained.<br />

�<br />

ρ<br />

= φ<br />

α + βi ω − kω − k′ �<br />

iω = (−1) k+k′<br />

� �<br />

ρ<br />

φ .<br />

α + βi<br />

ρ = m + µ � λα + λ ′ β � �<br />

+ t α 2 + β 2�<br />

,<br />

<strong>The</strong> expressed roots <strong>of</strong> T = 0 are all distinct. To realize that all the roots <strong>of</strong> the<br />

equation T = 0 were contained in the list (16.13), ABEL first observed that none <strong>of</strong><br />

the roots corresponding to different values <strong>of</strong> ρ could be identical. He did so using<br />

the same techniques as above. <strong>The</strong>n, ABEL found that T had no multiple roots by<br />

observing that a multiple root <strong>of</strong> T would be a common root <strong>of</strong> T = 0 and T ′ = 0.<br />

However, any root <strong>of</strong> T ′ = 0 would also be a root <strong>of</strong> S = 0 which was forbidden by<br />

assuming that the rational function T S<br />

(16.12) was expressed in its reduced form. As a<br />

consequence, ABEL found that all the roots <strong>of</strong> the original equation reduced to roots<br />

<strong>of</strong> an equation R = 0 <strong>of</strong> degree 2v in which the roots were<br />

φ 2<br />

� ω<br />

α + βi<br />

�<br />

, φ 2<br />

� 2ω<br />

α + βi<br />

�<br />

, . . . , φ 2<br />

� 2vω<br />

α + βi<br />

This equation could, ABEL observed, “easily be solved by the method <strong>of</strong> GAUSS.” 35<br />

Actually, ABEL solved it using the same approach as in the general division problem,<br />

i.e. the approach which led to (16.11), above.<br />

Geometrical division <strong>of</strong> the lemniscate. Having solved the equation R = 0, ABEL<br />

thus had access to the values<br />

� �<br />

kω<br />

φ for k = 1, 2, . . . , 2v,<br />

α + βi<br />

and he now proceeded to obtain the value φ � �<br />

ω<br />

4v+1 by the following brief argument.<br />

By the addition theorem, ABEL expressed<br />

� � � �<br />

mv mv 2mαω<br />

φ + = φ<br />

α + βi α − βi 4v + 1<br />

� � � �<br />

mv<br />

mw<br />

in terms <strong>of</strong> φ α+βi and φ α−βi where the latter could be obtained from the former<br />

“by changing i into −i”. 36 Since 2α and 4v + 1 were relatively prime, ABEL could write<br />

any integral n as<br />

n = 2mα − (4v + 1) t,<br />

35 “Cela posé, on peut aisément résoudre l’équation R = 0, à l’aide de la méthode de M. Gauss.”<br />

(N. H. <strong>Abel</strong>, 1828b, 165).<br />

36 (ibid., 166).<br />

�<br />

.


318 Chapter 16. <strong>The</strong> idea <strong>of</strong> inverting elliptic integrals<br />

i.e.<br />

φ<br />

� � �<br />

2mαω nω<br />

= φ<br />

4v + 1 4v + 1<br />

�<br />

+ tω = (−1) t � �<br />

nω<br />

φ<br />

4v + 1<br />

which for n = 1 made φ � �<br />

ω<br />

4n+1 accessible (known).<br />

Throughout, ABEL had explicitly only considered the division into a prime number<br />

<strong>of</strong> parts. When ABEL then made the further assumption that 4v + 1 = 1 + 2 n , he found<br />

by considering the expressions obtained that all the root extractions reduced to square<br />

roots. In particular, ABEL had to use that the solution <strong>of</strong> the equation θ2n−1 = 1 could<br />

be reduced to square roots; this result is precisely the main result <strong>of</strong> GAUSS’ research<br />

on the division <strong>of</strong> the circle. Combining this result with the case <strong>of</strong> bisection and<br />

the general integral multiplication, ABEL could summarize his investigations on the<br />

division <strong>of</strong> the lemniscate:<br />

“<strong>The</strong> value <strong>of</strong> the function φ � �<br />

mω can be expressed by square roots whenever<br />

n<br />

n is a number <strong>of</strong> the form 2n or a prime number <strong>of</strong> the form 1 + 2n or a product <strong>of</strong><br />

multiple numbers <strong>of</strong> these two forms.” 37<br />

Two aspects <strong>of</strong> ABEL’S result merit attention. First, ABEL’S argument hinges on<br />

the factors 2 m0, 2 n 1 + 1, . . . , 2 n k + 1 <strong>of</strong> n to be relatively prime because he wanted to<br />

decompose any number m ′ into its residues modulo these factors,<br />

�<br />

m0 m1<br />

φ +<br />

2n0 2n1 + 1 + · · · + mk 2n � �<br />

= φ<br />

k + 1<br />

m ′<br />

2 n0 (2 n 1 + 1) . . . (2 n k + 1)<br />

If two <strong>of</strong> the Fermat primes were identical, the decomposition would no longer be pos-<br />

sible. 38 ABEL’S deductions immediately leading to the stated theorem contain tacitly<br />

the distinctness <strong>of</strong> the Fermat primes, but it could have been explicitly included in the<br />

statement.<br />

Second, the result states a sufficient condition <strong>of</strong> geometrical constructibility and<br />

says absolutely nothing <strong>of</strong> the necessity <strong>of</strong> this condition. <strong>The</strong> same can be said <strong>of</strong><br />

GAUSS’ stated result on the division <strong>of</strong> the circle. However, GAUSS also stated that the<br />

division <strong>of</strong> the circle would lead to precisely his equations and would therefore not<br />

be possible with ruler and compass unless the number <strong>of</strong> parts were <strong>of</strong> the prescribed<br />

form. Similarly, ABEL could have stated that division <strong>of</strong> the lemniscate was only pos-<br />

sible if n was a product <strong>of</strong> a power <strong>of</strong> 2 and distinct Fermat primes. 39 However, the<br />

pro<strong>of</strong> <strong>of</strong> such a statement would go beyond the types <strong>of</strong> questions which ABEL asked<br />

concerning these classes <strong>of</strong> equations.<br />

37 “La valeur de la fonction φ � �<br />

mω<br />

n peut être exprimée par des racines carrées toutes les fois que n<br />

est un nombre de la forme 2n ou un nombre premier de la forme 1 + 2n , ou même un produit de<br />

plusieurs nombres de ces deux formes.” (N. H. <strong>Abel</strong>, 1828b, 168).<br />

38 Consider writing e.g. 3<br />

39<br />

a+b<br />

25 as 5 with a, b ∈ Z.<br />

For a pro<strong>of</strong> using more modern techniques, see (M. Rosen, 1981).<br />

�<br />

.


16.4. Perspectives on inversion 319<br />

16.4 Perspectives on inversion<br />

<strong>The</strong> present chapter has documented ABEL’S inversion <strong>of</strong> elliptic integrals into elliptic<br />

functions and the intimately related step <strong>of</strong> extending the resulting function to allow<br />

complex values <strong>of</strong> the argument.<br />

This extension to the complex domain which laid the foundation for ABEL’S re-<br />

search on elliptic functions was based on a formal substitution. Although a rigor-<br />

ous foundation for complex integration was being undertaken in the period, the ev-<br />

idence suggests that ABEL based his inversion on an imaginary substitution in the<br />

Eulerian tradition which presumed the existence <strong>of</strong> the function and sought to dis-<br />

cover/construct its values.<br />

<strong>The</strong> step toward considering the inverse function <strong>of</strong> general elliptic integrals (<strong>of</strong><br />

the first kind) was probably motivated by the division problem for the lemniscate<br />

to which ABEL was led by GAUSS’ remark in the Disquisitiones arithmeticae. In the<br />

process <strong>of</strong> solving the division problem for the lemniscate, ABEL solved the associated<br />

equation using a technical and rather ad hoc approach. Later, in the Mémoire sur une<br />

classe particulière (see chapter 7), ABEL approached the same set <strong>of</strong> problems but from a<br />

more conceptual approach in which he had clearly grasped the very essential property<br />

<strong>of</strong> the equations.<br />

ABEL’S tailored deductions relied extensively on manipulations <strong>of</strong> formulae; in<br />

particular, ABEL’S addition theorems and his characterization <strong>of</strong> the roots <strong>of</strong> the equa-<br />

tion φ (α) = φ (β) served him as important tools. When it came to the investigations<br />

on the solubility <strong>of</strong> the division problems, number theoretic arguments inspired by<br />

GAUSS’ Disquisitiones arithmeticae were also used.


Chapter 17<br />

Steps in the process <strong>of</strong> coming to<br />

“know” elliptic functions<br />

As already described in the previous chapter, N. H. ABEL (1802–1829) introduced<br />

his elliptic functions by means <strong>of</strong> a formal inversion <strong>of</strong> the elliptic integrals. In or-<br />

der to make these new objects known, however, this definition appears to have been<br />

quite insufficient. ABEL, himself, was not explicit about the problem — but a com-<br />

parison <strong>of</strong> ABEL’S approach with A.-M. LEGENDRE’S (1752–1833) highly numerical<br />

approach suggests that ABEL’S purely formal definition based on the formal inver-<br />

sion was lacking in certain respects. For instance, in ABEL’S approach, it would be<br />

difficult to compute particular values <strong>of</strong> ABEL’S elliptic functions purely from the def-<br />

inition. <strong>The</strong>refore, he developed means <strong>of</strong> representing his new functions by existing<br />

objects and the objects which he chose were — not surprisingly — infinite series and<br />

products.<br />

17.1 Infinite representations<br />

In the first part <strong>of</strong> the Recherches, 1 a large portion <strong>of</strong> the text is occupied with highly<br />

technical and formula-based manipulations which aim at describing the elliptic func-<br />

tion φ (and the derived functions f and F) in infinite products and series. Two charac-<br />

teristic and interesting examples pertaining to the expansion in doubly infinite sums<br />

are discussed below.<br />

Based on the multiplication problem described above, ABEL had found that<br />

φ ((2n + 1) β) = P2n+1<br />

Q2n+1<br />

(17.1)<br />

in which deg P2n+1 = (2n + 1) 2 and deg Q2n+1 = (2n + 1) 2 − 1. Furthermore, P 2n+1<br />

x<br />

1 (N. H. <strong>Abel</strong>, 1827b).<br />

321


322 Chapter 17. Steps in the process <strong>of</strong> coming to “know” elliptic functions<br />

was a polynomial in x 2 and so was Q2n+1. Thus, with<br />

P2n+1 (x) = ax (2n+1)2 + · · · + bx and<br />

Q2n+1 (x) = cx (2n+1)2 −1 + · · · + d,<br />

ABEL found that the sum <strong>of</strong> the roots <strong>of</strong> the equation (17.1) would equal c aφ ((2n + 1) β).<br />

Furthermore, he already knew the complete solution <strong>of</strong> the equation (17.1), and thus<br />

he obtained<br />

φ ((2n + 1) β) = A<br />

n<br />

∑<br />

m=−n<br />

n<br />

∑<br />

µ=−n<br />

Similarly for the products <strong>of</strong> the roots <strong>of</strong> (17.1),<br />

φ ((2n + 1) β) = B<br />

n<br />

∏<br />

m=−n<br />

(−1) m+µ �<br />

φ β +<br />

n<br />

∏<br />

µ=−n<br />

�<br />

φ β +<br />

�<br />

mω + µ ¯ωi<br />

. (17.2)<br />

2n + 1<br />

�<br />

mω + µ ¯ωi<br />

.<br />

2n + 1<br />

<strong>The</strong>se formulae invited the limit process n → ∞, and it is interesting to see how ABEL<br />

carried it out.<br />

17.1.1 Determination <strong>of</strong> the coefficient A<br />

In order to determine the constant A <strong>of</strong> (17.2), ABEL wanted to insert a particular value<br />

for β and he chose β = ω 2 + ¯ω 2 i. However, this is a singular value (a pole) <strong>of</strong> φ, and<br />

ABEL applied a limit argument in the following form. First, using the relation<br />

�<br />

φ α + ω ¯ω<br />

+<br />

2 2 i<br />

�<br />

= − i 1<br />

ec φ (α)<br />

derivable from the addition formulae, ABEL found that for (m, µ) �= (0, 0),<br />

�<br />

mω + µ ¯ωi ω ¯ω<br />

φ<br />

+ +<br />

2n + 1 2 2 i<br />

� �<br />

mω + µ ¯ωi ω ¯ω<br />

+ φ − + +<br />

2n + 1 2 2 i<br />

�<br />

= 0.<br />

(17.3)<br />

Consequently, the sum reduced to the term corresponding to (m, µ) = (0, 0). For the<br />

last term, ABEL found that<br />

A = lim<br />

β→ ω 2 + ¯ω 2 i<br />

φ ((2n + 1) β)<br />

φ (β)<br />

=<br />

by (17.3) lim<br />

α→0<br />

φ (α)<br />

φ ((2n + 1) α)<br />

φ<br />

= lim<br />

α→0<br />

� (2n + 1) � ω<br />

2 + ¯ω 2 i + α��<br />

φ � ω<br />

2 + ¯ω 2 i + α�<br />

= 1<br />

2n + 1<br />

where tacit applications <strong>of</strong> the differentiability <strong>of</strong> φ and <strong>of</strong> the Rule <strong>of</strong> l’Hospital are<br />

involved.


17.1. Infinite representations 323<br />

17.1.2 Infinite sums<br />

In order to express φ (α) by an infinite series, ABEL set β = α<br />

2n+1 . Thus,<br />

φ (α) = φ ((2n + 1) β) = 1<br />

2n + 1<br />

n<br />

∑<br />

n<br />

∑ (−1)<br />

m=−n µ=−n<br />

m+µ φ<br />

�<br />

β +<br />

�<br />

mω + µ ¯ωi<br />

.<br />

2n + 1<br />

Following a string <strong>of</strong> manipulations designed to group the terms <strong>of</strong> the right hand<br />

side, ABEL reduced the problem to the search for the limit <strong>of</strong> the double sum<br />

with<br />

n−1 n−1<br />

∑ ∑<br />

m=0 µ=0<br />

ψ (m, µ) = 1<br />

2n + 1<br />

(−1) m+µ ψ (m, µ) (17.4)<br />

2φ � α<br />

φ 2 � α<br />

2n+1<br />

2n+1<br />

in which ζ (x) = f (x) F (x) and εµ =<br />

In order to find the limit <strong>of</strong> (17.4), ABEL remarked,<br />

and<br />

� ζ<br />

� − φ 2<br />

�<br />

m + 1<br />

2<br />

� �<br />

εµ<br />

2n+1<br />

�<br />

εµ<br />

2n+1<br />

�<br />

�<br />

ω +<br />

�<br />

µ + 1<br />

�<br />

2<br />

“one attempts to put the preceding quantity [here (17.4)] on the form<br />

P + v,<br />

in which P is independent <strong>of</strong> n and v is a quantity which has the limit zero, because<br />

then the quantity P is exactly the limit which is sought.” 2<br />

ABEL had a candidate in mind for the expression P when he defined<br />

θ (m, µ) = 2α<br />

α 2 − ε 2 µ<br />

ψ (m, µ) − θ (m, µ) =<br />

2α<br />

(2n + 1)<br />

2 Rµ.<br />

His candidate was the double sum ∑ ∞ m=0 ∑ ∞ µ=0 (−1) m+µ θ (m, µ) and he proceeded in<br />

the following way in obtaining this limit.<br />

For each value <strong>of</strong> m, ABEL argued, the difference was<br />

n−1<br />

∑<br />

µ=0<br />

(−1) µ n−1<br />

(ψ (m, µ) − θ (m, µ)) = 2α ∑<br />

µ=0<br />

2 “il faut essayer de mettre la quantité précédente sous la forme<br />

P + v,<br />

(−1) µ Rµ<br />

,<br />

2<br />

(2n + 1)<br />

où P est indépendant de n, et v une quantité qui a zéro pour limité, car alors la quantité P sera<br />

précisément la limite dont il s’agit.” (N. H. <strong>Abel</strong>, 1827b, 156).<br />

¯ωi.


324 Chapter 17. Steps in the process <strong>of</strong> coming to “know” elliptic functions<br />

and he claimed and proved that the right hand side was “<strong>of</strong> the form v<br />

2n+1 ”. In the<br />

process, ABEL made use <strong>of</strong> the result that<br />

φ (α) = α + Aα 3 + . . .<br />

because φ was an odd function and φ ′ (0) = f (0) F (0) = 1.<br />

ABEL’S next step concerned the sum <strong>of</strong> θ. He found that<br />

and therefore, for each m,<br />

∞<br />

∑ (−1)<br />

µ=n<br />

µ θ (m, µ) = v<br />

2n + 1<br />

n−1<br />

∑ (−1)<br />

µ=0<br />

µ ψ (m, µ) =<br />

When he summed these, ABEL found<br />

n−1 n−1<br />

∑ ∑ (−1)<br />

m=0 µ=0<br />

m+µ n−1<br />

ψ (m, µ) = ∑ (−1)<br />

m=0<br />

m<br />

and, as ABEL wrote,<br />

n−1<br />

∑<br />

m=0<br />

vm<br />

2n + 1<br />

∞<br />

∑ (−1)<br />

µ=0<br />

µ θ (m, µ) + vm<br />

2n + 1 .<br />

�<br />

∞<br />

∑ (−1)<br />

µ=0<br />

µ �<br />

n−1<br />

θ (m, µ) + ∑<br />

m=0<br />

= nv<br />

2n + 1<br />

= v<br />

2 ,<br />

vm<br />

2n + 1<br />

in which “v is a quantity which has zero for its limit”. 3 Consequently, ABEL had found<br />

a way <strong>of</strong> expressing φ (α) as a double infinite sum, e.g.<br />

φ (α) = 1<br />

ec<br />

∞<br />

∑<br />

∞<br />

∑ (−1) m+µ<br />

�<br />

(2µ+1) ¯ω<br />

m=0 µ=0<br />

(α−(m+ 1 2)ω) 2 +(µ+ 1 2) 2 −<br />

¯ω 2<br />

(2µ+1) ¯ω<br />

(α+(m+ 1 2)ω) 2 +(µ+ 1 2) 2 ¯ω 2<br />

�<br />

.<br />

(17.5)<br />

ABEL did not stop after having obtained the expansion in infinite series (17.5).<br />

Instead, he used similar methods to search for expressions for φ (α) involving infinite<br />

products. ABEL also invested an effort in obtaining expressions involving only one<br />

infinite series and transcendental objects in the terms. Among the formulae which he<br />

obtained, the following was probably the simplest:<br />

φ (α) = 2<br />

∞<br />

π<br />

ec ¯ω ∑ (−1)<br />

k=0<br />

k ε2k+1 − ε−(2k+1) r2k+1 − r−(2k+1) where<br />

�<br />

ε = exp α π<br />

�<br />

�<br />

ωπ<br />

�<br />

and r = exp .<br />

¯ω<br />

2 ¯ω<br />

Thus, as described, ABEL used his multiplication formulae to obtain rather simple<br />

infinite representations for elliptic functions. When he passed to the infinite limit, his<br />

arguments did not conform to the strict standards <strong>of</strong> rigor, which he had advocated<br />

in the theory <strong>of</strong> series. Further perspectives on ABEL’S motivations for searching for<br />

infinite expressions and his methods <strong>of</strong> obtaining them are described and discussed<br />

in the subsequent sections.<br />

3 (N. H. <strong>Abel</strong>, 1827b, 161).


17.2. Elliptic functions as ratios <strong>of</strong> power series 325<br />

17.2 Elliptic functions as ratios <strong>of</strong> power series<br />

In the Précis, 4 listed among the established facts <strong>of</strong> elliptic functions, ABEL observed<br />

that elliptic functions <strong>of</strong> the first kind (which he now denoted λ, see below) were<br />

expressible as the ratio <strong>of</strong> two convergent power series,<br />

With the notation<br />

λ (θ) = θ + a1θ3 + a2θ5 + . . .<br />

1 + b2θ4 + b3θ6 . (17.6)<br />

+ . . .<br />

φ (θ) = θ + a1θ 3 + a2θ 5 + . . . and<br />

f (θ) = 1 + b2θ 4 + b3θ 6 + . . . ,<br />

(17.7)<br />

ABEL claimed that the functions φ and f (which are not to be confused with the func-<br />

tions <strong>of</strong> the same names in the Recherches) satisfied the linked functional equations<br />

φ � θ ′ + θ � φ � θ ′ − θ � = φ 2 (θ) f 2 � θ ′� − φ 2 � θ ′� f 2 (θ) and<br />

f � θ ′ + θ � f � θ ′ − θ � = f 2 (θ) f 2 � θ ′� − c 2 φ 2 (θ) φ 2 � θ ′� .<br />

(17.8)<br />

<strong>The</strong> sign <strong>of</strong> the first equation is actually wrong, see below. ABEL also mentioned the<br />

same result in his letter to LEGENDRE but he never published any demonstration <strong>of</strong><br />

it. 5<br />

In order to see how ABEL came to this expression, the following reconstruction<br />

may be suggested based on ABEL’S sparse hints.<br />

In his comments published in the second volume <strong>of</strong> the Œuvres, M. S. LIE (1842–<br />

1899) has presented a reconstruction <strong>of</strong> ABEL’S reasoning based on the same sources,<br />

papers and manuscripts. LIE indicated how ABEL’S manuscript notes — given the<br />

power series expansion (17.7) — could be interpreted as steps toward determining the<br />

remaining coefficients a1, a2, . . . , b2, b3, . . . . However, I interpret the notes slightly dif-<br />

ferently and infer from them a suggestion <strong>of</strong> how ABEL came to claim the series ex-<br />

pansion (17.7), itself, by use <strong>of</strong> the expansion in two Maclaurin series. I am confident<br />

that the following reconstruction is close to ABEL’S original argument.<br />

Derivation <strong>of</strong> functional equations. In the Précis, ABEL had presented the following<br />

consequence <strong>of</strong> the addition formulae,<br />

Supposing that λ (θ) was written as<br />

λ � θ ′ + θ � λ � θ ′ − θ � = λ2 (θ ′ ) − λ 2 (θ)<br />

1 − c 2 λ 2 (θ) λ 2 (θ ′ ) .<br />

λ (θ) =<br />

φ (θ)<br />

f (θ)<br />

4 (N. H. <strong>Abel</strong>, 1829d).<br />

5 (<strong>Abel</strong>→Legendre, Christiania, 1828/11/25. N. H. <strong>Abel</strong>, 1902a, 82).<br />

(17.9)


326 Chapter 17. Steps in the process <strong>of</strong> coming to “know” elliptic functions<br />

for some functions φ and f , ABEL obtained<br />

φ (θ ′ + θ) φ (θ ′ − θ)<br />

f (θ ′ + θ) f (θ ′ − θ) =<br />

φ2 (θ ′ )<br />

f 2 (θ ′ ) − φ2 (θ)<br />

f 2 (θ)<br />

1 − c2 φ2 (θ)φ2 (θ ′ )<br />

f 2 (θ) f 2 (θ ′ )<br />

= φ2 (θ ′ ) f 2 (θ) − φ 2 (θ) f 2 (θ ′ )<br />

f 2 (θ) f 2 (θ ′ ) − c 2 φ 2 (θ) φ 2 (θ ′ )<br />

and by comparing numerators and denominators, the functional equations (17.8) re-<br />

sult with the exception <strong>of</strong> the change <strong>of</strong> sign in the numerator. 6 This is evidently a<br />

mistake in ABEL’S paper — repeated in the Œuvres — as can be seen by setting θ = 0<br />

and observing that because λ (0) = 0, φ (0) must also be zero.<br />

Except for this small error, this part <strong>of</strong> the argument is completely straight-forward<br />

and fits well with ABEL’S other manipulations and the formulae which he presented<br />

in the Précis. In order to obtain the series expansions, we may get a clue from one <strong>of</strong><br />

ABEL’S notebooks in which he proceeded to differentiate the equations (17.8). 7<br />

Coefficients <strong>of</strong> φ and f . First, we observe from the second functional equation by<br />

letting θ = θ ′ = 0 that<br />

f 2 (0) = f 4 (0) , i.e. f 2 (0) = 1.<br />

To be precise, f (0) = 0 was also a possibility but that would produce<br />

for all x which is not a relevant case.<br />

f 2 (x) = 0<br />

In the functional equations, letting θ = 0 gives<br />

φ � θ ′� φ � −θ ′� = −φ 2 � θ ′� f 2 (0) , i.e. φ � −θ ′� = −φ � θ ′� ,<br />

and φ is therefore an odd function. Thus, the coefficients <strong>of</strong> all odd powers in a power<br />

series expansion must be zero. <strong>The</strong> same argument applied to the second functional<br />

equation produces<br />

f � θ ′� f � −θ ′� = f 2 � θ ′�<br />

and therefore, f is an even function.<br />

Now, a few more coefficients may be determined, for instance f (0) and f ′′ (0)<br />

which can be found by differentiating the relation<br />

f (2θ) = f 4 (θ) − c 2 φ 4 (θ)<br />

twice and letting θ = 0 to obtain f 3 (0) = 1 and by the previous result, f (0) = 1.<br />

To determine f ′′ (0) we differentiate again and insert θ = 0 to obtain<br />

f ′′ (0) = 0.<br />

6 As mentioned above.<br />

7 (<strong>Abel</strong>, MS:351:C, 193), see also (N. H. <strong>Abel</strong>, 1881, II, 319).


17.2. Elliptic functions as ratios <strong>of</strong> power series 327<br />

<strong>The</strong> final coefficient which ABEL specified was φ ′ (0). It may be found directly by<br />

differentiation <strong>of</strong> (17.9)<br />

φ ′ (0) = d<br />

�<br />

�<br />

λ (x) f (x) �<br />

dx<br />

� x=0<br />

= λ ′ (0) f (0) = 1.<br />

Thus, summarizing the results, ABEL could claim by the use <strong>of</strong> an expansion in<br />

two Maclaurin series and detailed studies <strong>of</strong> the differential quotients that<br />

with the further information<br />

λ (θ) = ∑∞n=0 anθ2n+1 ∑ ∞ n=0 bnθ2n a0 = 1, b0 = 1, and b1 = 0.<br />

In the letter to LEGENDRE, ABEL claimed that the coefficients were always poly-<br />

nomial functions <strong>of</strong> the modulus c 2 . This claim can also be seen to be an easy conse-<br />

quence <strong>of</strong> the defining functional equations and can be proved by induction.<br />

Convergence <strong>of</strong> φ and f . Furthermore, ABEL claimed — both in the Précis and in<br />

the letter — that the two series <strong>of</strong> (17.6) were always convergent. This is perhaps the<br />

most difficult <strong>of</strong> ABEL’S claims to obtain from the approach presented above. Fur-<br />

thermore, there are no hints <strong>of</strong> ABEL’S reasoning left neither in the notebooks, nor in<br />

his publications. He may have obtained some sort <strong>of</strong> indication <strong>of</strong> convergence from<br />

the approach described above, or he may simply have stated the convergence <strong>of</strong> the<br />

Maclaurin series as an unproven fact. It is, <strong>of</strong> course, also possible if not probable that<br />

ABEL had actually grasped an implicit concept <strong>of</strong> meromorphic functions as general-<br />

ized rational functions.<br />

Later, the expression <strong>of</strong> elliptic functions as the ratio <strong>of</strong> convergent power series<br />

was exactly the point which provoked K. T. W. WEIERSTRASS’ (1815–1897) interest<br />

in research mathematics and — according to WEIERSTRASS, himself — convinced him<br />

that he wanted to become a mathematician. 8 WEIERSTRASS’ approach centered on<br />

differential equations and contained a theorem explicitly stating the convergence <strong>of</strong><br />

the power series equivalent to φ and f .<br />

Further coefficients <strong>of</strong> φ and f . Besides the coefficients which were determined<br />

in the paper, ABEL’S manuscript also contains the series including the coefficients<br />

a1 = − 1+c2<br />

6 and b2 = − c2<br />

12 . On the same sheet <strong>of</strong> the notebook, the differential quotients<br />

f ′ (x) = c2<br />

3 x3 , f ′′ (x) = −c2x2 , and f ′′′ (x) = −2c2x can also be found which are<br />

only correct provided terms with higher powers <strong>of</strong> x have been neglected, e.g. by con-<br />

sidering the local behaviour for small values <strong>of</strong> x. A further differentiation produces<br />

f (4) (0) = −2c 2<br />

8 (Weierstrass→Lie, Berlin, 1882/04/10. N. H. <strong>Abel</strong>, 1902b, 104).


328 Chapter 17. Steps in the process <strong>of</strong> coming to “know” elliptic functions<br />

and<br />

b2 = f (4) (0)<br />

c2<br />

= −2c2 = −<br />

4! 4! 12 .<br />

Thus, the calculations appear to be probes <strong>of</strong> ABEL’S result. <strong>The</strong> notebook does not<br />

contradict my interpretation, I believe, that ABEL obtained his representation <strong>of</strong> λ<br />

by means <strong>of</strong> two Maclaurin series. However, as LIE’S interpretation also illustrates,<br />

ABEL’S publications on the subject are too few and his notes to difficult to interpret to<br />

present any final opinion on the debate, in particular concerning the convergence <strong>of</strong><br />

the series.<br />

17.3 Characterization <strong>of</strong> ABEL’s representations<br />

Having presented an outline and discussion <strong>of</strong> the technical details <strong>of</strong> ABEL’S deriva-<br />

tion <strong>of</strong> the representations, a few themes are summarized which will also be relevant<br />

to the discussion in subsequent chapters in part IV and in section 21.2.<br />

17.3.1 ABEL’s style <strong>of</strong> reasoning.<br />

Characteristic <strong>of</strong> ABEL’S style in the Recherches, the derivation <strong>of</strong> the infinite represen-<br />

tations such as (17.5) relies extensively on manipulations <strong>of</strong> formulae and is repeated<br />

afterwards for the functions f and F, although the arguments are highly similar. On a<br />

related theme, ABEL introduced symbols to denote most <strong>of</strong> his auxiliary and interme-<br />

diate calculations. <strong>The</strong>se facts are textual evidence <strong>of</strong> the formula-manipulating style<br />

in which ABEL’S Recherches is mainly written.<br />

Brief comparison with L. EULER (1707–1783). As described, ABEL’S way <strong>of</strong> obtain-<br />

ing infinite expressions for the elliptic function φ closely resembles EULER’S methods.<br />

ABEL transformed an expression for φ (α) into a form which contained a number (re-<br />

lated to) n terms and then proceeded to split this expression into the sum <strong>of</strong> a part<br />

independent <strong>of</strong> n and a part which vanished with increasing values <strong>of</strong> n.<br />

This is comparable to EULER’S derivation <strong>of</strong> the power series expansion <strong>of</strong> the ex-<br />

ponential function in the Introductio ad analysin infinitorum. 9 <strong>The</strong>re, EULER — cloaked<br />

in his language <strong>of</strong> infinitesimals — considered<br />

�<br />

1 + x<br />

n<br />

� n<br />

, (17.10)<br />

expanded it by the binomial formula, and let n grow to infinity to get an infinite num-<br />

ber <strong>of</strong> terms. Simultaneously, this turned the expression into exp x.<br />

Compared to EULER’S expansion, ABEL’S original formula was valid for any value<br />

<strong>of</strong> n. <strong>The</strong> number n was only an auxiliary quantity later to disappear when he found<br />

limit expressions which were independent <strong>of</strong> n.<br />

9 (L. Euler, 1748, 85–87).


17.3. Characterization <strong>of</strong> ABEL’s representations 329<br />

Convergence <strong>of</strong> infinite expressions. As LIE commented in the notes, ABEL’S “meth-<br />

ods in obtaining expressions for the functions φ (α), f (α), F (α) in series and infinite<br />

products do not seem to us to be satisfactory in all details.” 10 In particular, it is re-<br />

markable that the ardent rigorist <strong>of</strong> the binomial paper never even mentioned the<br />

word convergence in the Recherches. In the Précis, however, the convergence <strong>of</strong> the se-<br />

ries (17.6) was stated as a fact without further explanation.<br />

ABEL’S way <strong>of</strong> obtaining series expansions <strong>of</strong> elliptic functions essentially let the<br />

number <strong>of</strong> terms in a finite expression grow to infinity. ABEL’S concern for the con-<br />

vergence <strong>of</strong> the process was dealt with by the method <strong>of</strong> writing the expression as one<br />

part which was independent <strong>of</strong> n and another part which vanished with increasing n.<br />

Thus, it appears, two different standards <strong>of</strong> rigor in the theory <strong>of</strong> series were in-<br />

volved in ABEL’S research using infinite series. In foundational issues, a strict adher-<br />

ence to A.-L. CAUCHY’S (1789–1857) program and the associated theoretical complex<br />

was advocated by ABEL. However, when it came to research on new groundbreaking<br />

objects, ABEL used the methods which he had learned from EULER and was content<br />

with observing that his results were sound.<br />

17.3.2 <strong>The</strong> need for multiple representations<br />

A final aspect which is revealing <strong>of</strong> the role played by representations <strong>of</strong> elliptic func-<br />

tions in ABEL’S works is the necessity <strong>of</strong> obtaining multiple representations. Under-<br />

standably, functions introduced in such an indirect way as the inversion <strong>of</strong> a non-<br />

elementary integral needed some other means <strong>of</strong> numerical determination and this is<br />

one <strong>of</strong> the roles played by representations in ABEL’S theory.<br />

Convergence. As already noticed repeatedly, ABEL was not very explicit about the<br />

convergence <strong>of</strong> his infinite representations. This may have been a reason for not rely-<br />

ing on any single representation but deriving multiple and various representations in<br />

the hope that at least some <strong>of</strong> them would prove adequate in particular instances.<br />

Applications. A connected motivation for multiple representations could also be the<br />

ambition <strong>of</strong> multiple applications (within pure mathematics). In the following chapter,<br />

an instance where infinite representations play a central role in the pro<strong>of</strong> <strong>of</strong> a theorem<br />

will be described. ABEL and his contemporaries would have hoped and expected to<br />

find many similar instances.<br />

Aesthetics. <strong>The</strong> third aspect <strong>of</strong> the discussion is less technical and more <strong>of</strong> a personal<br />

and contextual nature which can only be appreciated in a broader time scale. In section<br />

21.2, where the discussion <strong>of</strong> representations is taken up again, it will also become<br />

10 (N. H. <strong>Abel</strong>, 1881, II, 306).


330 Chapter 17. Steps in the process <strong>of</strong> coming to “know” elliptic functions<br />

clear that the preferred definitions and representation vary over time and between<br />

mathematical traditions.<br />

In ABEL’S theory <strong>of</strong> elliptic functions, the objects were introduced as inversions <strong>of</strong><br />

elliptic integrals. <strong>The</strong> largely Eulerian tradition to which ABEL’S Recherches generally<br />

belongs emphasized representations <strong>of</strong> transcendental functions by infinite (power)<br />

series and infinite products. On the way to obtaining these representations, ABEL also<br />

came across the functional equations (17.8) and proved that elliptic functions were<br />

doubly periodic and could be written as the ratio <strong>of</strong> power series. All these represen-<br />

tations and key results were later assumed as definitions <strong>of</strong> the concept elliptic function<br />

depending on the setting and context in which they were introduced.<br />

17.4 Conclusion<br />

One major achievement <strong>of</strong> the search for representations was, <strong>of</strong> course, that based on<br />

formulae such as (17.5), approximations to φ (α) could be computed with any degree<br />

<strong>of</strong> accuracy. Another, and equally interesting — but less anticipated — result was that<br />

infinite expressions, themselves, could play a role in the development <strong>of</strong> the theory. In<br />

the Recherches, this aspect remained little cultivated but in subsequent papers, ABEL<br />

occasionally applied infinite expressions, even to answer algebraic questions (see next<br />

chapter).<br />

In the Recherches, ABEL did more than solve the division problem for the lemnis-<br />

cate. While the lemniscate provided a clear question to which he produced a clear<br />

answer, the other part <strong>of</strong> the paper dealt with problems <strong>of</strong> more intrinsic nature and<br />

provided answers which are now hardly recognizable as answers because the ques-<br />

tions which they answer have faded in importance.


Chapter 18<br />

Tools in ABEL’s research on elliptic<br />

transcendentals<br />

In order to illustrate how N. H. ABEL (1802–1829) worked with elliptic transcenden-<br />

tals, the two most important topics are described in some details below. In order to<br />

make the presentation coherent, special emphasis is given to some <strong>of</strong> ABEL’S papers<br />

which illustrate important points concerning the types <strong>of</strong> tools involved by ABEL in<br />

his research on elliptic transcendentals.<br />

18.1 Transformation theory<br />

With A.-M. LEGENDRE’S (1752–1833) systematization <strong>of</strong> the theory <strong>of</strong> elliptic inte-<br />

grals, the transformations which he devised to reduce one integral to the simplest one<br />

in its class were powerful and important tools. ABEL approached the theory <strong>of</strong> trans-<br />

formations in his Recherches but his main contributions were spelled out in a number<br />

<strong>of</strong> articles written as a reaction to results announced by C. G. J. JACOBI (1804–1851) in<br />

the journal Astronomische Nachrichten.<br />

Competition with JACOBI. An important theme in a large part <strong>of</strong> the ABEL-related<br />

research has been his rivalry with the contemporary German mathematician JACOBI. 1<br />

As noted in connection with the idea <strong>of</strong> inverting elliptic integrals into elliptic func-<br />

tions, ABEL realized that he was in the middle <strong>of</strong> a rivalry when he became aware <strong>of</strong><br />

JACOBI’S published announcements in H. C. SCHUMACHER’S (1784–1873) Astronomis-<br />

che Nachrichten. 2 <strong>The</strong> present section describes ABEL’S contribution to the theory in<br />

which JACOBI had initially been the most interested. By studying ABEL’S techni-<br />

cal machinery in some details, his means <strong>of</strong> dealing with elliptic functions become<br />

clearer. Most interestingly, ABEL employed largely algebraic methods in his research<br />

1 See also chapter 2 and section 16.2.5.<br />

2 In a letter to SCHUMACHER, HANSTEEN described how ABEL, upon learning <strong>of</strong> these papers, became<br />

very pale and had to rush to the nearest bakery for a dram; see (Holst, 1902, 89) and (E.<br />

Andersen, 1975, 104).<br />

331


332 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

Figure 18.1: CARL GUSTAV JACOB JACOBI (1804–1851)<br />

but also put the machinery <strong>of</strong> infinite representations to good use in obtaining an im-<br />

portant theorem in transformation theory. After a spell <strong>of</strong> shorter papers which com-<br />

plemented his two major papers, the Recherches and the Précis, 3 ABEL left the theory<br />

<strong>of</strong> transformations alone — and he soon died, <strong>of</strong> course. Transformation theory was<br />

not ABEL’S major purpose but it was a topic which provoked the interest <strong>of</strong> some <strong>of</strong><br />

his most important contemporaries.<br />

In his announcements in the Astronomische Nachrichten, JACOBI communicated only<br />

the results and not the methods which he had employed to deduce them. <strong>The</strong> first<br />

biographies <strong>of</strong> ABEL have invested an effort in calling attention to ABEL’S priority<br />

in proving the results. In particular, C. A. BJERKNES (1825–1903) responded to L.<br />

KÖNIGSBERGER’S (1837–1921) account <strong>of</strong> the discovery <strong>of</strong> elliptic functions by arguing<br />

that JACOBI only obtained his pro<strong>of</strong>s after learning <strong>of</strong> ABEL’S inversion. 4 However,<br />

since the turn <strong>of</strong> the twentieth century, historians and mathematicians have agreed<br />

that although JACOBI initially obtained his results through an ingenious but unrig-<br />

orous heuristic, he probably developed the inversion <strong>of</strong> elliptic integrals on his own,<br />

possibly inspired by reading through ABEL’S Recherches. 5 To indicate the hectic char-<br />

acter <strong>of</strong> the events, some key dates <strong>of</strong> the rivalry have been presented in table 18.1.<br />

3 (N. H. <strong>Abel</strong>, 1827b; N. H. <strong>Abel</strong>, 1828b) and (N. H. <strong>Abel</strong>, 1829d), respectively.<br />

4 (Bjerknes, 1880; Koenigsberger, 1879).<br />

5 (Mittag-Leffler, 1907; Ore, 1954; Pieper, 1998).


18.1. Transformation theory 333<br />

<strong>The</strong> central question <strong>of</strong> transformation theory. ABEL picked up the theme <strong>of</strong> trans-<br />

formations <strong>of</strong> elliptic integrals from LEGENDRE. In ABEL’S words, the central problem<br />

was posed with variations on multiple occasions, for instance in the following way:<br />

“To find all the possible cases in which one can satisfy the differential equation<br />

dy<br />

�<br />

(1 − y2 ) � 1 − c2 = a<br />

1y2� dx<br />

� (1 − x 2 ) (1 − c 2 x 2 )<br />

by an algebraic equation between the variables x and y and supposing the moduli c<br />

and c1 less than unity and the coefficient a either real or imaginary.” 6<br />

Expressed in the notation <strong>of</strong> LEGENDRE’S differentials, the above question asks<br />

for every possible way <strong>of</strong> transforming x algebraically into y in such a way that the<br />

integral with modulus c1 transformed into the integral with modulus c.<br />

18.1.1 ABEL’s response to JACOBI’s announcements<br />

ABEL first published in the Astronomische Nachrichten in 1828; 7 in a lengthy paper, he<br />

demonstrated how the theory which he had developed in his Recherches could answer<br />

a question raised by JACOBI. In the paper — which is entitled Solution d’un problème<br />

général concernant la transformation des fonctions elliptiques — , 8 ABEL began by describ-<br />

ing key results concerning the inverse <strong>of</strong> elliptic integrals deduced in the Recherches.<br />

With the notation<br />

and the inversion<br />

∆ (x) =<br />

�<br />

θ =<br />

0<br />

�<br />

(1 − c 2 x 2 ) (1 − e 2 x 2 )<br />

dx<br />

and x = λ (θ) ,<br />

∆ (x)<br />

ABEL presented the highlights <strong>of</strong> the Recherches in two theorems which summarize<br />

(16.3) and (16.8), respectively. First, he expressed the addition theorem for the elliptic<br />

function <strong>of</strong> the first kind λ,<br />

λ � θ ± θ ′� = λ (θ) ∆ (θ′ ) ± λ (θ ′ ) ∆ (θ)<br />

1 − c2e2λ2 (θ) λ2 (θ ′ .<br />

)<br />

Second, ABEL described the conditions on the arguments which ensured that the func-<br />

tion took identical values,<br />

λ (θ) = λ � θ ′� if and only if θ ′ = (−1) m+m′<br />

θ + mω + m ′ ω ′<br />

6 “Trouver tous les cas possibles où l’on pourra satisfaire à l’équation différentielle:<br />

dy<br />

�<br />

(1 − y2 ) � 1 − c2 = a<br />

1y2� dx<br />

� (1 − x 2 ) (1 − c 2 x 2 )<br />

(18.1)<br />

par une équation algébrique entre les variables x et y, en supposant les modules c et c1 moindre que<br />

l’unité et le coeffcient a réel ou imaginaire.” (N. H. <strong>Abel</strong>, 1829a, 33).<br />

7 (N. H. <strong>Abel</strong>, 1828d). <strong>The</strong> paper is dated 27 May 1828.<br />

8 (ibid.).


334 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

in which the semi-periods were determined by<br />

� �<br />

ω 1<br />

= θ and<br />

2 c<br />

ω′<br />

2<br />

= θ<br />

� �<br />

1<br />

.<br />

e<br />

ABEL expressed the central properties <strong>of</strong> this theorem:<br />

“This theorem is generally valid no matter whether the quantities e and c are<br />

real or imaginary. In the paper cited above [Recherches], I have proved it in the<br />

case where e 2 is negative and c 2 is positive. [. . . ] <strong>The</strong> quantities ω, ω ′ always have<br />

an imaginary ratio. Otherwise, they have the same role in the theory <strong>of</strong> elliptic<br />

functions as the number π has in the theory <strong>of</strong> circular functions.” 9<br />

Now, in order to address the transformation problem, ABEL observed that the<br />

method <strong>of</strong> indeterminate coefficients could be applied. This method amounted to ap-<br />

proaching the problem by introducing two power series with indeterminate coeffi-<br />

cients and using the defining equations to obtain relations among the terms. How-<br />

ever, as ABEL critically remarked, this method would lead to extremely cumbersome<br />

calculations, and ABEL proposed a simpler and more direct one. Below, this method<br />

is briefly described.<br />

Rational transformations. With the notation and basic results set up, ABEL turned to<br />

a question which he proposed and ascribed great importance for the theory <strong>of</strong> elliptic<br />

functions. He was interested in finding all the possible ways in which the differential<br />

equation<br />

dy<br />

��1 − c2 1y2� � 1 − e2 = ±a<br />

1y2� dx<br />

� (1 − c 2 x 2 ) (1 − e 2 x 2 )<br />

(18.2)<br />

could be satisfied in which y was an algebraic function <strong>of</strong> x. In the paper, ABEL limited<br />

his considerations to rational functions y = ψ (x) because the general question “at first<br />

seems too difficult”. 10<br />

ABEL’S first result in this situation was an algebraic one, not so different from re-<br />

sults obtained in his paper on <strong>Abel</strong>ian equations. 11 By a string <strong>of</strong> manipulations, ABEL<br />

found that if the equation (18.2) was satisfied, the roots <strong>of</strong> the equation ψ (x) = y had<br />

the remarkable property <strong>of</strong> being related in a very specific way: if λ (θ) represented<br />

one <strong>of</strong> the roots, any other root <strong>of</strong> the equation would be representable as λ (θ + α)<br />

where α was a constant, i.e.<br />

y = ψ (λ (θ)) = ψ (λ (θ + α)) .<br />

9 “Ce théorème a lieu généralement, quelles que soient les quantités e et c, réelles ou imagainaires.<br />

Je l’ai démontré pour le cas où e 2 est négatif et c 2 positif dans le mémoire cité plus haut [. . . ]. Les<br />

quantités ω, ω ′ sont toujours dans un rapport imaginaire. Elles jouent d’ailleurs dans la théorie des<br />

fonctions elliptiques le même rôle que le nombre π dans celle des fonctions circulaires.” (N. H. <strong>Abel</strong>,<br />

1828d, 366).<br />

10 (ibid., 365).<br />

11 See chapter 7.


18.1. Transformation theory 335<br />

Obviously, only finitely many roots could exist, and the value <strong>of</strong> α could be obtained<br />

from (18.1),<br />

α = µω + µ ′ ω ′<br />

(18.3)<br />

in which µ, µ ′ were rational constants. However, the degree <strong>of</strong> the equation y = ψ (x)<br />

might surpass the number <strong>of</strong> different values produced in this fashion and a new<br />

group corresponding to a new value α2 <strong>of</strong> α might be necessary. ABEL found that a<br />

certain number ν (which he did not describe in any detail) existed such that all the<br />

roots <strong>of</strong> the equation y = ψ (x) would be representable by the different values <strong>of</strong> the<br />

expression<br />

λ<br />

�<br />

θ +<br />

�<br />

ν<br />

∑ knαn<br />

n=1<br />

when k1, . . . , kν took all integer values. However, the different values could also be<br />

represented as (possibly changing the values <strong>of</strong> α1, α2, . . . )<br />

λ (θ) , λ (θ + α1) , . . . , λ (θ + αm−1)<br />

in which α1, . . . , αm−1 were still rational linear combinations <strong>of</strong> ω and ω ′ (as in 18.3).<br />

ABEL then wrote the rational function ψ (x) = p(x)<br />

q(x)<br />

obtained 12<br />

m−1<br />

p (x) − q (x) y = A<br />

∏<br />

n=0<br />

with no common divisors and<br />

(x − λ (θ + αn)) .<br />

<strong>The</strong> constant A was <strong>of</strong> the form A = f − gy with f , g constants. ABEL’S next step<br />

was to find an expression for A and he did so by first imposing a limiting assumption<br />

and gradually relaxing it. First, ABEL considered the case in which both p and q were<br />

polynomials <strong>of</strong> the first degree. In this case (e.g. by Euclidean division),<br />

and ABEL found<br />

y = f ′ + f x<br />

g ′ + gx ,<br />

dy = f g′ − f ′ g<br />

(g ′ dx.<br />

2<br />

+ gx)<br />

When he inserted this into the original differential equation, its dependence on dy<br />

disappeared. Consequently, ABEL could conclude that the differential equation in this<br />

case implied either <strong>of</strong> the three solutions<br />

I.y = ax, c 2 1<br />

II.y = a 1<br />

ec<br />

= c2<br />

a 2 ,e2 1<br />

x , c2 c2<br />

1 =<br />

III.y = m 1 − x√ ec<br />

1 + x √ ec , c1 = 1<br />

m<br />

12 Here I write α0 = 0 for brevity.<br />

a 2 ,e2 1<br />

e2<br />

= , or<br />

a2 = e2<br />

a<br />

√ c − √ e<br />

√ c + √ e ,e1 = 1<br />

m<br />

2 , or<br />

√ √<br />

c + e<br />

√ √ , a =<br />

c − e im<br />

(c − e) .<br />

2


336 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

Next, ABEL returned to the results which he had previously established and now<br />

let f ′ and g ′ denote the coefficients <strong>of</strong> x m−1 in p and q. <strong>The</strong>n, by comparing the coeffi-<br />

cients, ABEL found<br />

f ′ − g ′ y = − ( f − gy)<br />

Consequently, when he isolated y, ABEL found<br />

m−1<br />

∑<br />

n=0<br />

y = f ′ + f φ (θ)<br />

g ′ + gφ (θ)<br />

λ (θ + αn) .<br />

� �� �<br />

=φ(θ)<br />

which would serve to determine y as a function <strong>of</strong> x in all but those particular cases<br />

where φ (θ) reduced to a constant.<br />

In order for y to be a rational function <strong>of</strong> x, ABEL observed that the new function<br />

φ (θ) must likewise be rational in x. ABEL set out to investigate the circumstances un-<br />

der which this would be the case. By computing and combining values <strong>of</strong> the elliptic<br />

function λ, he found that φ (θ) was always a rational function <strong>of</strong> x and that it was given<br />

by<br />

φ (θ) = (1 − k) x + k′′ − k ′<br />

ec<br />

1<br />

x +<br />

m−1<br />

∑<br />

n=1<br />

2x∆ (αn)<br />

1 − e 2 c 2 λ 2 (αn) x 2<br />

in which k, k ′ , k ′′ were constants which were either zero or one. Based on this repre-<br />

sentation, ABEL separated three cases corresponding to various combinations <strong>of</strong> the<br />

values <strong>of</strong> k, k ′ , k ′′ .<br />

In the end, after his technical manipulations in the various cases, ABEL found that<br />

the differential equation under consideration<br />

dy<br />

�<br />

(1 − y2 ) � 1 − e2 = ±<br />

1y2� was satisfied precisely if a, e1, and y were given by<br />

a = k<br />

n−1<br />

∏<br />

m=0<br />

n−1<br />

∏<br />

m=1<br />

�<br />

m<br />

λ<br />

n ω<br />

�<br />

e1 = e n<br />

, y = k<br />

�<br />

n−1<br />

∏ λ<br />

m=0<br />

a dx<br />

� (1 − x 2 ) (1 − e 2 x 2 ) = ±a dθ<br />

n−1 �<br />

∏ λ θ +<br />

m=0<br />

mω<br />

�<br />

, and<br />

n<br />

�<br />

2m + 1<br />

2n ω<br />

��2 where n was an arbitrary integer and the constants k and ω were given by<br />

1 = k<br />

�<br />

2m + 1<br />

λ<br />

2n ω<br />

�<br />

dx<br />

� .<br />

(1 − x2 ) (1 − e2x2 )<br />

and ω<br />

2 =<br />

� 1<br />

0<br />

Following this characterization <strong>of</strong> the solutions to the problem, ABEL first trans-<br />

lated the result into the trigonometric language employed by LEGENDRE and then<br />

announced a number <strong>of</strong> “remarkable theorems on elliptic functions” 13 which are <strong>of</strong><br />

less relevance in the present context.<br />

13 (N. H. <strong>Abel</strong>, 1828d, 385).


18.1. Transformation theory 337<br />

Summary: an algebraic pro<strong>of</strong>. As described, ABEL’S deduction consisted <strong>of</strong> five<br />

steps: First, ABEL set up his notation and definitions and introduced important results<br />

from the Recherches. Second, ABEL found that if λ (θ (x)) is a root, i.e. if y =<br />

ψ (λ (θ (x))), then any other root has the form λ (θ (x) + α). Next, the constant α could<br />

be determined and a general representation <strong>of</strong> the roots can be given — possibly in-<br />

volving multiple “orbits” corresponding to α1, α2, . . . . Fourth, the relation between y<br />

and x could be spelled out. Eventually, it could be necessary to consider a number <strong>of</strong><br />

cases in order to describe these relations and deduce formulae <strong>of</strong> particular interest.<br />

A few broader points should also be observed. First, ABEL’S pro<strong>of</strong> made central<br />

use <strong>of</strong> the properties <strong>of</strong> elliptic functions which had been deduced in the Recherches.<br />

In particular, the solution <strong>of</strong> the equation λ (x) = λ (y) — which originated from the<br />

double periodicity <strong>of</strong> the function λ — became very instrumental in the present con-<br />

text just as it had been in the solution <strong>of</strong> the division problem (see section 16.3). Sec-<br />

ond, the approach which ABEL took may well be described as an algebraic one; it relied<br />

on algebraic tools such as specific knowledge <strong>of</strong> the roots <strong>of</strong> certain polynomial equa-<br />

tions, division <strong>of</strong> polynomials, and considerations <strong>of</strong> the rationality <strong>of</strong> certain func-<br />

tions. <strong>The</strong>se are tools which were also present in ABEL’S purely algebraic researches<br />

on solubility (see part II). However, ABEL also adopted another approach to the same<br />

question.<br />

Counting the possible numbers <strong>of</strong> transformations. In another paper — this time<br />

published in CRELLE’S Journal and motivated by another <strong>of</strong> JACOBI’S papers — , 14<br />

ABEL gave the theory <strong>of</strong> transformation a slightly different turn. He continued the<br />

path laid out in the Astronomische Nachrichten and made frequent references to the<br />

paper described above, but in the Journal, ABEL wanted to count and enumerate the<br />

different transformations. ABEL considered a rational transformation <strong>of</strong> x into y <strong>of</strong><br />

a certain prime degree 2n + 1 and found — by employing algebraic tools similar to<br />

those described above — that 12 (2n + 2) different transformations corresponding to<br />

12 (n + 1) different values <strong>of</strong> the transformed modulus were generally possible. ABEL<br />

remarked that for certain particular values <strong>of</strong> the modulus c, the number <strong>of</strong> transfor-<br />

mations might degenerate. This notion <strong>of</strong> arguments carried out “in general” will be<br />

discussed further in section 19.3 and chapter 21.<br />

ABEL’S determination <strong>of</strong> the number <strong>of</strong> transformations spurred a reaction from<br />

LEGENDRE who believed that it was at odds with JACOBI’S determination <strong>of</strong> the de-<br />

gree <strong>of</strong> the so-called modular equation. JACOBI had claimed that for a transformation <strong>of</strong><br />

prime degree 2n + 1, 2n + 2 values <strong>of</strong> the transformed modulus were possible. Thus,<br />

ABEL’S value was six times JACOBI’S number <strong>of</strong> transformations. However, as ABEL<br />

argued in a letter to LEGENDRE, JACOBI had indeed solved an equation with 2n + 2<br />

different roots but each root <strong>of</strong> this equation could also produce five other values for<br />

14 (N. H. <strong>Abel</strong>, 1828e); JACOBI’S paper is (C. G. J. Jacobi, 1828).


338 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

the transformed modulus and ABEL’S result was valid. 15<br />

18.1.2 An additional note<br />

In a subsequent issue <strong>of</strong> the Astronomische Nachrichten, ABEL inserted a note which<br />

added a different deduction <strong>of</strong> the main result <strong>of</strong> the previous one. 16 Whereas ABEL<br />

had initially employed direct and detailed manipulations to obtain the characteriza-<br />

tion <strong>of</strong> transformations, he now used one <strong>of</strong> the infinite representations which he had<br />

also obtained in the Recherches.<br />

From the Recherches, ABEL imported the expansion <strong>of</strong> the auxiliary function f in an<br />

infinite product and manipulated it into the current setting in which it produced<br />

�<br />

λ (α) = A × ψ α π<br />

�<br />

¯ω<br />

×<br />

∞ � �<br />

∏ ψ<br />

n=1<br />

(nω + α) π<br />

¯ω<br />

� �<br />

ψ<br />

(nω − α) π<br />

¯ω<br />

with A a constant and<br />

1 − e−2x<br />

ψ (x) = .<br />

1 + e−2x Furthermore, the periods ω and ¯ω were also related in ABEL’S usual way to the mod-<br />

ulus c. ABEL now inserted α = θ + m n for m = 0, 1, . . . , n − 1 and multiplied these<br />

expressions together to obtain the central formula<br />

with<br />

n−1 �<br />

∏ λ θ +<br />

m=0<br />

mω<br />

�<br />

n<br />

��<br />

= A n �<br />

ψ δ π<br />

� ∞ � �<br />

∏ ψ (ω1 + δ)<br />

¯ω1 m=0<br />

π<br />

�<br />

ψ (ω1 − δ)<br />

¯ω1<br />

π<br />

�<br />

¯ω1<br />

δ = ¯ω1 ω1<br />

θ and<br />

¯ω ¯ω1<br />

= 1<br />

n<br />

ω<br />

. (18.4)<br />

¯ω<br />

<strong>The</strong> important idea which ABEL utilized now was to relate this formula to the sim-<br />

ilar one for the transformed function λ ′ corresponding to the modulus c1, the periods<br />

¯ω and ¯ω1, and the constant A1. ABEL found that<br />

λ ′<br />

�<br />

¯ω1<br />

¯ω θ<br />

�<br />

= A1<br />

An n−1 �<br />

∏ λ<br />

m=0<br />

θ + mω<br />

n<br />

whenever the modulus c1 was such that the periods were related by (18.4).<br />

This time, we see how the knowledge <strong>of</strong> an infinite representation <strong>of</strong> the associated<br />

function f helped ABEL make statements about the transformation <strong>of</strong> elliptic func-<br />

tions. This is particularly interesting because it illustrates a different approach from<br />

the more algebraic one which he had initially taken. With direct access to a represen-<br />

tation <strong>of</strong> the function f , ABEL could employ a mixture <strong>of</strong> finite and infinite results to<br />

obtain he characterization <strong>of</strong> the conditions <strong>of</strong> rational transformations.<br />

15 (<strong>Abel</strong>→Legendre, Christiania, 1828/11/25. N. H. <strong>Abel</strong>, 1902a, 79).<br />

16 (N. H. <strong>Abel</strong>, 1829a).<br />


18.2. Integration in logarithmic terms 339<br />

18.2 Integration in logarithmic terms<br />

Another problem which figures significantly in ABEL’S approach to and research on<br />

higher transcendentals was the question <strong>of</strong> integration in more elementary forms. In<br />

chapter 15, it was described how mathematicians attacked the study <strong>of</strong> elliptic inte-<br />

grals although these were non-elementary. One <strong>of</strong> the approaches adopted was to<br />

relate a number <strong>of</strong> elliptic integrals by elementary functions or to investigate situa-<br />

tions in which the integration could indeed be effected in elementary (or finite) terms.<br />

A similar idea was pursued by ABEL in his investigations on what he called “theory<br />

<strong>of</strong> integration”. Thus, ABEL’S understanding <strong>of</strong> this notion differed from the present<br />

one in the sense that it was highly formal or algebraic and did not concern a numeri-<br />

cal interpretation <strong>of</strong> the integral. Such an interpretation was, <strong>of</strong> course, part <strong>of</strong> A.-L.<br />

CAUCHY’S (1789–1857) complete program <strong>of</strong> rigorization and became very important<br />

in the 19 th century mainly in the efforts to answer the challenges raised by Fourier<br />

series.<br />

Reminiscences <strong>of</strong> the Collegium mémoire. <strong>The</strong> first evidence <strong>of</strong> ABEL’S interest in<br />

the theory <strong>of</strong> integration (in finite terms) originates from descriptions <strong>of</strong> a paper which<br />

is no longer extant. As was already described in section 2.3, ABEL had hoped to em-<br />

bark on his European Tour shortly after the application was sent to the Collegium <strong>of</strong><br />

the University in 1824. Before that time, in March 1823, ABEL presented a manuscript<br />

to the Collegium academicum through C. HANSTEEN (1784–1873). It concerned “a gen-<br />

eral presentation <strong>of</strong> the possibility <strong>of</strong> integrating all possible differential formulae” 17 .<br />

<strong>The</strong> manuscript was given to pr<strong>of</strong>essors HANSTEEN and S. RASMUSSEN (1768–1850)<br />

for their pr<strong>of</strong>essional evaluation. <strong>The</strong>ir review was positive but no means <strong>of</strong> publish-<br />

ing the paper were at hand and it was subsequently lost. However, from ABEL’S pub-<br />

lished research, we may get an impression <strong>of</strong> what it could have contained. ABEL’S<br />

notebooks contain a number <strong>of</strong> entries related to the question <strong>of</strong> integration in finite<br />

terms; in particular, a manuscript for a large memoir on the theory <strong>of</strong> elliptic transcen-<br />

dentals from this perspective has been included in the Œuvres. 18 Nevertheless, the<br />

present description focuses on his main publication on the subject which occurred in<br />

the Journal in 1826. 19<br />

<strong>The</strong> local context <strong>of</strong> ABEL’S work on integration in finite terms was mainly related<br />

to the same theme as his research in the Paris memoir (see chapter 19, below). How-<br />

ever, it also included such issues as the reduction <strong>of</strong> all elliptic integrals to four basic<br />

kinds which ABEL undertook in his manuscripts and which had been a corner stone<br />

<strong>of</strong> LEGENDRE’S theory <strong>of</strong> elliptic integrals. 20 In the 1840s, mainly through the works<br />

17 “en almindelig Fremstilling af Muligheden at integrere alle mulige Differential-Formler” (N. H.<br />

<strong>Abel</strong>, 1902d, 4).<br />

18 (N. H. <strong>Abel</strong>, [1825] 1839b).<br />

19 (N. H. <strong>Abel</strong>, 1826d).<br />

20 (N. H. <strong>Abel</strong>, [1825] 1839b, 101); for LEGENDRE’S theory, see section 15.3.


340 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

<strong>of</strong> J. LIOUVILLE (1809–1882), the theory <strong>of</strong> integration in finite terms established itself<br />

as an independent theory investigated for its own results. In this context, the theory<br />

and ABEL’S contribution to it have been well described in J. LÜTZEN’S biography <strong>of</strong><br />

LIOUVILLE. 21 Referring to LÜTZEN’S description, a presentation <strong>of</strong> ABEL’S argument<br />

and a brief discussion <strong>of</strong> relevant points <strong>of</strong> ABEL’S contribution are included below.<br />

18.2.1 Characterization by continued fractions<br />

In the paper from 1826, 22 ABEL investigated conditions under which the integral<br />

�<br />

ρ dx<br />

√R<br />

(18.5)<br />

could be reduced to the logarithmic expression<br />

log p + q√R p − q √ . (18.6)<br />

R<br />

<strong>The</strong> article first dealt with this question <strong>of</strong> reduction, but as ABEL ultimately noticed,<br />

the answer obtained was actually the answer to a more general question. ABEL noted<br />

that in case the integral (18.5) could be represented by logarithmic functions in any<br />

way, it would always have a representation <strong>of</strong> the form (18.6). ABEL promised a pro<strong>of</strong><br />

<strong>of</strong> this assertion but never published one; it was eventually given by P. L. CHEBYSHEV<br />

(1821–1894). 23<br />

A non-empty class. ABEL found by direct differentiation that for<br />

he would have<br />

Writing dz in the form<br />

dz =<br />

z = log p + q√ R<br />

p − q √ R ,<br />

pq dR + 2 (p dq − q dp) R<br />

.<br />

(p 2 − q 2 R) √ R<br />

M dx<br />

dz =<br />

N √ R with<br />

M = pq dR<br />

�<br />

+ 2 p<br />

dx dq<br />

�<br />

− qdp R and (18.7)<br />

dx dx<br />

N = p 2 − q 2 R, (18.8)<br />

he had thus found that for such values <strong>of</strong> M and N,<br />

� M dx<br />

N √ R = log p + q√ R<br />

p − q √ R .<br />

ABEL concluded:<br />

21 (Lützen, 1990, chapter IX).<br />

22 (N. H. <strong>Abel</strong>, 1826d).<br />

23 (Chebyshev, 1853).


18.2. Integration in logarithmic terms 341<br />

ρ dx<br />

“From this it follows, that in the differential √ an infinitude <strong>of</strong> rational func-<br />

R<br />

tions ρ can be found which make this differential integrable by<br />

�<br />

logarithms; furthermore,<br />

this is done by an expression <strong>of</strong> the form log .” 24<br />

� p+q √ R<br />

p−q √ R<br />

Thus, ABEL had proved that the class <strong>of</strong> differentials which were integrable by loga-<br />

rithms was non-empty.<br />

Delineation <strong>of</strong> the class. It was the converse <strong>of</strong> this result, that ABEL really wanted<br />

to investigate in the paper published in CRELLE’S Journal. He formulated the prob-<br />

ρ dx<br />

lem <strong>of</strong> determining all differentials <strong>of</strong> the form √ which could be integrated in the<br />

R<br />

logarithmic form<br />

log p + q√R p − q √ . (18.9)<br />

R<br />

This problem can be interpreted as another instance <strong>of</strong> a problem <strong>of</strong> delineation for a class<br />

<strong>of</strong> objects, in this case the class <strong>of</strong> objects integrable in the logarithmic form (18.9). 25<br />

With ρ = M N an entire function, ABEL took similar steps as above (18.7 and 18.8)<br />

and found the relation<br />

dp dN<br />

M 2 dx − p N dx<br />

= .<br />

N q<br />

<strong>The</strong>n followed a series <strong>of</strong> reductions to obtain a simple description <strong>of</strong> the relations<br />

between M, N, and R. Because M N was an entire function <strong>of</strong> x, it followed from this<br />

equation that p dN<br />

N dx was also an entire function <strong>of</strong> x,<br />

N =<br />

Reduction into partial fractions implied that<br />

dN<br />

N dx =<br />

n<br />

∏ (x + ak) k=0<br />

mk .<br />

n<br />

∑<br />

k=0<br />

m k<br />

x + a k<br />

and because p dN<br />

N dx was to be entire, ABEL could write<br />

p = p1<br />

n<br />

∏ (x + ak) k=0<br />

in which p1 was an entire function. As a consequence <strong>of</strong> the relation N = p 2 − q 2 R,<br />

ABEL found<br />

n<br />

∏ (x + ak) k=0<br />

mk=p 2 n<br />

1 ∏ (a + ak) k=0<br />

2<br />

� �� �<br />

=N<br />

24 “Daraus folgt, daß sich in dem Differential<br />

� �� �<br />

=p 2<br />

−q 2 R<br />

ρ dx<br />

√ R , für die rationale Function ρ unzählige Formen fin-<br />

den lassen, die dieses Differential durch Logarithmen integrabel machen, und zwar durch einen<br />

Ausdruck von der Form log<br />

25 See chapter 21.<br />

� p+q √ R<br />

p−q √ R<br />

�<br />

.” (N. H. <strong>Abel</strong>, 1826d, 186).


342 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

and because R did not contain any square factors and p and q could be assumed rela-<br />

tively prime, ABEL found m0 = m1 = · · · = mn = 1 and obtained the factorization<br />

R = R1<br />

n<br />

∏ (x + ak) = R1N.<br />

k=0<br />

with R1 an entire function. With this result, ABEL had found a reduced characteriza-<br />

tion in the form<br />

p 2 1 N − q2 R1 = 1 and M<br />

N<br />

= p1q dR<br />

dx<br />

�<br />

+ 2 p dq<br />

dx<br />

�<br />

− qdp R1.<br />

dx<br />

Considerations <strong>of</strong> degrees. ABEL’S next step was to investigate the consequences <strong>of</strong><br />

the first part <strong>of</strong> the characterization obtained above,<br />

p 2 1 N − q2 R1 = 1. (18.10)<br />

As he remarked, the equation could be solved by the method <strong>of</strong> indeterminate coeffi-<br />

cients but this approach would be extremely cumbersome and not lead to any general<br />

conclusion. Instead, he proposed a different approach. Before embarking on his novel<br />

approach, ABEL introduced the notations δP to denote the degree <strong>of</strong> the (rational)<br />

function P and EP to denote the entire part <strong>of</strong> P, i.e.<br />

u = Eu + u ′ with δu ′ < 0.<br />

Judging from the detailed introduction <strong>of</strong> these concepts, ABEL did not assume them<br />

to be familiar to his readers. Concerning these new concepts, ABEL easily proved the<br />

following lemma:<br />

Lemma 3 If the functions u, v, z are related by<br />

and δz < δv, then<br />

u 2 = v 2 + z<br />

Eu = ±Ev. ✷<br />

ABEL now returned to the equation p 2 1 N − q2 R1 = 1 and applied his new result<br />

(lemma 3). ABEL immediately obtained<br />

�<br />

δ p 2 1N �<br />

and consequently<br />

�<br />

= δ q 2 �<br />

R1<br />

2δp1 + δN = 2δq + δR1, i.e.<br />

δ (NR1) = 2 (δq + δR1 − δp1)


18.2. Integration in logarithmic terms 343<br />

and because NR1 = R, ABEL had found that the highest power in R had to be an even<br />

number. ABEL wrote δN = n − m and δR1 = n + m and generalized the study <strong>of</strong> the<br />

equation (18.10) to the equation<br />

in which v was an entire function with δv < δN+δR 1<br />

2<br />

ABEL then wrote<br />

p 2 1 N − q2 R1 = v (18.11)<br />

= n.<br />

R1 = Nt + t ′ with t = E R1<br />

N and δt′ < 0.<br />

As a consequence <strong>of</strong> the assumptions, ABEL found that δt = 2m and he wrote t in the<br />

form<br />

t = t 2 1 + t′ 1<br />

in which δt ′ 1 < m. With these conventions, ABEL had remodelled the equation (18.11)<br />

into<br />

After rewriting the equation as<br />

ABEL observed that<br />

and he applied lemma 3 to obtain<br />

v = p 2 1N − q2 � Nt + t ′� =<br />

�<br />

= p 2 1 − q2t 2 �<br />

1 N − q 2 � t ′ 1N + t′� .<br />

�<br />

p 2 1 − q2 �<br />

t N − q 2 t ′<br />

� �2 p1<br />

= t<br />

q<br />

2 v<br />

1 +<br />

Nq2 + t′ t′<br />

1 +<br />

N ,<br />

�<br />

v<br />

δ<br />

Nq2 + t′ �<br />

t′<br />

1 + < m = δt1,<br />

N<br />

E<br />

� �<br />

p1<br />

= ±Et1 = ±t1.<br />

q<br />

This meant, that ABEL had found a relation between p1 and q <strong>of</strong> the form<br />

p1 = t1q + β in which δβ < δq.<br />

Through a sequence <strong>of</strong> similar, very explicit manipulations, ABEL transformed the<br />

equation (18.11) into the form<br />

in which<br />

s1β 2 − 2r1ββ1 − sβ 2 1 = v (18.12)<br />

δr1 = 1<br />

2 δR = n, δβ1 < δβ, δs < n, and δs1 < n.


344 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

His investigations now turned toward solving the equation (18.12). ABEL did so by<br />

observing that the process used above could be iterated producing a sequence <strong>of</strong> rela-<br />

tions similar to (18.12). After n − 1 iterations, he found the relation<br />

snβ 2 n−1 − 2rnβn−1βn − sn−1β 2 n = (−1) n−1 v,<br />

in which δβn < δβn−1.<br />

Because the sequence <strong>of</strong> degrees was decreasing, it would eventually produce δβm =<br />

0, i.e. βm = 0, and the final relation would then become<br />

smβ 2 m−1 = (−1)m−1 v.<br />

Using this information, ABEL ascended the chain <strong>of</strong> β, β1, . . . , βm in the reverse order<br />

each time finding expressions for βn−1 <strong>of</strong> the form<br />

βn−1 = 2µnβn + βn+1.<br />

By solving these relations for the first term <strong>of</strong> the chain β, ABEL found an expression<br />

for β<br />

β 1 as a finite continued fraction.<br />

In order to answer the question <strong>of</strong> logarithmic integration <strong>of</strong> the original differen-<br />

tial, ABEL next investigated the consequences for the radical √ R. He found from his<br />

earlier results that by assuming m infinite, the expansion <strong>of</strong> √ R would be<br />

√<br />

R = t1 +<br />

2µ +<br />

1<br />

1<br />

2µ 1+ 1<br />

2µ 2 +...<br />

Here, ABEL noticed in a footnote that the equality <strong>of</strong> √ R and its continued fraction<br />

should not be interpreted as a numerical equality except in those situations where the<br />

continued fraction has a value.<br />

Finally, ABEL translated an earlier assumption that one among the quantities s1, s2, . . .<br />

should be independent <strong>of</strong> x into the property that the continued fraction for √ R<br />

should be periodic. <strong>The</strong> assumption on s1, s2, . . . had been introduced to ensure the<br />

solubility <strong>of</strong> the equations, and it thus amounted to a criterion for the possibility <strong>of</strong> in-<br />

tegrating<br />

ρ dx<br />

√ R in logarithmic terms. ABEL summarized his investigations as a complete<br />

criterion <strong>of</strong> logarithmic integrability stating that for polynomials ρ, the integration<br />

� ρ dx<br />

√R = log y + √ R<br />

y − √ R<br />

.<br />

(18.13)<br />

could be effected if and only if the expansion <strong>of</strong> √ R into continued fractions was<br />

periodic. In the affirmative case, the function y was determined by the first period <strong>of</strong><br />

the continued fraction for √ R.


18.3. Conclusion 345<br />

Summary. ABEL’S investigations concerning integration on the logarithmic form (18.13)<br />

serves to illustrate some interesting aspects. First, ABEL’S interest in the problem <strong>of</strong> in-<br />

tegrating differentials in logarithmic forms reveals the position <strong>of</strong> his research within<br />

a tradition <strong>of</strong> reducing complicated integrals to simpler ones. At the same time, the<br />

way he attacked the problem was rather novel. In his approach, ABEL applied the<br />

program which he had presented in his notebook research on solubility <strong>of</strong> equations<br />

and did not search “by divination” for an integration <strong>of</strong> the specific form (see section<br />

8.1). Instead, he took upon himself to establish the precise conditions under which the<br />

integration would be possible. Second, ABEL employed highly algebraic tools involv-<br />

ing polynomials, degrees <strong>of</strong> rational functions, and considerations <strong>of</strong> dependencies<br />

among quantities to reach his conclusion. At the point, where his argument came to<br />

involve an infinite representation in the form <strong>of</strong> an expansion <strong>of</strong> √ R into a continued<br />

fraction, he stressed that the equality should be interpreted as a formal one suited for<br />

determining the involved quantities.<br />

18.3 Conclusion<br />

In the present chapter, two major examples <strong>of</strong> ABEL’S work with elliptic transcenden-<br />

tals have been described in order to illustrate the tools which he employed. In partic-<br />

ular, ABEL’S recurring use <strong>of</strong> algebraic methods has been documented. This algebraic<br />

approach to the theory <strong>of</strong> higher transcendentals was a general theme in ABEL’S ap-<br />

proach and it will be described further in the following chapter. At points where he<br />

involved infinite expressions, they were <strong>of</strong>ten regarded as algebraic equalities — the<br />

precise conditions for convergence were rarely addressed. However, as noted, infinite<br />

representations could sometimes improve the deductions considerably.


346 Chapter 18. Tools in ABEL’s research on elliptic transcendentals<br />

1827.06.12 JACOBI dated his first letter to SCHUMACHER<br />

1827.08.02 JACOBI dated his second letter to SCHUMACHER<br />

1827.09 Extracts from JACOBI’S two letters to SCHUMACHER were published<br />

in the Astronomische Nachrichten (C. G. J. Jacobi, 1827b).<br />

1827.09.20 <strong>The</strong> issue <strong>of</strong> A. L. CRELLE’S (1780–1855) Journal containing the<br />

first part <strong>of</strong> ABEL’S Recherches appeared.<br />

1827.11.18 JACOBI dated his Demonstratio which was published in the Astronomische<br />

Nachrichten (C. G. J. Jacobi, 1827a).<br />

1828.01.25 JACOBI dated his one page addition to ABEL’S Recherches.<br />

1828.02.12 ABEL sent the second part <strong>of</strong> the Recherches to CRELLE.<br />

1828.04.02 JACOBI dated his first letter inserted in CRELLE’S Journal.<br />

1828.05.26 <strong>The</strong> issue <strong>of</strong> CRELLE’S Journal with the second part <strong>of</strong> ABEL’S<br />

Recherches and its note reacting to JACOBI was published<br />

1828.07.21 JACOBI dated his second letter inserted in CRELLE’S Journal.<br />

1828.10.03 JACOBI dated his third letter inserted in CRELLE’S Journal.<br />

1828.12.03 <strong>The</strong> issue <strong>of</strong> CRELLE’S Journal containing ABEL’S investigation<br />

on the number <strong>of</strong> transformations appeared.<br />

1829.01.11 JACOBI dated his fourth and final letter inserted in CRELLE’S<br />

Journal.<br />

Table 18.1: Important dates in the ABEL-JACOBI-rivalry


Chapter 19<br />

<strong>The</strong> Paris memoir<br />

N. H. ABEL’S (1802–1829) most famous result was first communicated in a paper<br />

which he delivered to the Parisian Academy <strong>of</strong> ScienceAcadémie des Sciences in 1826.<br />

<strong>The</strong> result which ABEL obtained in the so-called Paris memoir was a rather technical<br />

one which dealt with the integration <strong>of</strong> algebraic differentials. 1 In its original, it was<br />

formulated in the typical style <strong>of</strong> ABEL, his predecessors, and many <strong>of</strong> his contem-<br />

poraries but during the century ABEL’S result was recast in a quite different language<br />

and in another mathematical structure. 2 In modern mathematics, ABEL’S result is typ-<br />

ically considered a part <strong>of</strong> algebraic geometry; readers who wish to see a presentation<br />

<strong>of</strong> the result from such a modern perspective can consult e.g. (Shafarevich, 1974).<br />

Besides presenting the main results, the present rendering <strong>of</strong> ABEL’S Paris memoir<br />

aims at describing the central tools which ABEL employed in his reasoning. <strong>The</strong> fo-<br />

cus on tools facilitates a continuation <strong>of</strong> the comparison with the methods involved in<br />

ABEL’S purely algebraic works and also serves to support the discussion <strong>of</strong> the imme-<br />

diate reception <strong>of</strong> ABEL’S results taken up in section 19.5. ABEL’S arguments in the<br />

Paris memoir were conducted in a style heavily dependent on explicit manipulations<br />

<strong>of</strong> formulae. In section 19.5.1, his subsequent announcements <strong>of</strong> the main results and<br />

the much clearer sketches <strong>of</strong> pro<strong>of</strong> contained therein are described and discussed.<br />

19.1 ABEL’s approach to the Paris memoir<br />

ABEL’S Paris memoir represents a pivotal point in his mathematical production. Many<br />

investigations which ABEL had previously undertaken for their own sake and brought<br />

to interesting conclusions were surpassed by the main result <strong>of</strong> the Paris memoir. <strong>The</strong><br />

Paris memoir was three-fold monumental: it was the culmination <strong>of</strong> a line <strong>of</strong> research<br />

which ABEL had undertaken for years, it contained the result which brought him<br />

widespread fame in the nineteenth century, and yet it provided this result with an<br />

incredibly long and cumbersome pro<strong>of</strong>.<br />

1 ABEL’S result is also discussed in e.g. (Cooke, 1989; J. Gray, 1992).<br />

2 For a brief discussion <strong>of</strong> styles, see chapter 21.<br />

347


348 Chapter 19. <strong>The</strong> Paris memoir<br />

19.1.1 Set to work in Paris<br />

ABEL arrived in Paris on July 10, 1826. For quite some time, he had worked on the<br />

study <strong>of</strong> functions whose differentials satisfy certain algebraic conditions. In a letter<br />

to C. HANSTEEN (1784–1873) written shortly after his arrival, ABEL explained how he<br />

had postponed introducing himself to the Institut de France until his mastery <strong>of</strong> the<br />

French language had improved. He continued the letter:<br />

“Furthermore, and in particular, I want to complete the memoir I am working<br />

on and which I intend to present to the Institute. When it is complete, which<br />

will soon be the case, I will go there. <strong>The</strong> memoir has come out very well and<br />

contains many new things which I believe merit attention. It is the first draft <strong>of</strong> a<br />

theory <strong>of</strong> an infinitude <strong>of</strong> transcendental functions. — I nourish the hope that the<br />

Academy [Académie des Sciences] will have it printed in the Mémoires des savants<br />

étrangers.” 3<br />

Finally, on October 30, 1826, ABEL presented his memoir to the Institut de France.<br />

Three days later ABEL sent an article for publication in J. D. GERGONNE’S (1771–1859)<br />

Annales in which he presented his research on simultaneous solutions to two poly-<br />

nomial equations. 4 This paper contained an elaboration <strong>of</strong> one <strong>of</strong> the main tools <strong>of</strong><br />

ABEL’S Paris memoir; it is discussed in section 19.3, below. When he left France on 29<br />

December 1826, 5 ABEL had still not received any reaction from the Institut concerning<br />

his memoir and, in fact, he was never to receive one. ABEL’S Paris memoir was mis-<br />

placed before G. LIBRI (1803–1869) was commissioned with its printing. It eventually<br />

occurred in the Mémoires présentés par divers savants in 1841. <strong>The</strong> fate and reception <strong>of</strong><br />

ABEL’S Paris memoir are briefly described in section 19.4.<br />

19.1.2 Tools in ABEL’s toolbox<br />

<strong>The</strong> Paris memoir — when seen together with some <strong>of</strong> ABEL’S other publications de-<br />

scribed in chapter 18 — provides new insights into the toolbox <strong>of</strong> the creative, young<br />

mathematician. As could be expected <strong>of</strong> a mathematician devoted to algebra, the<br />

compartments for results concerning polynomials and equations are remarkably well<br />

equipped.<br />

1. ABEL used algebraic deductions and EUCLID’S (∼295 B.C.) algorithm in ways<br />

similar to those which he had already employed in his research on algebraic<br />

3 “Desuden vil jeg først og fremst have en Afhandling færdig som jeg arbeider paa og som jeg vil<br />

forelægge Institutet. Naar denne, hvilket snart skeer, er færdig gaar jeg derhen. Denne Afhandling<br />

er lykkets mig særdeles godt, og indeholder meget nyt og som jeg troer værdig Opmærksomhed.<br />

C’est la prémière ébauche d’une théorie d’une infinité de fonctions transcendantes. — Jeg har det<br />

Haab at Academiet vil lade den trykke i Mémoires des savants étrangers.” (<strong>Abel</strong>→Hansteen, Paris,<br />

1826/08/12. N. H. <strong>Abel</strong>, 1902a, 40).<br />

4 (N. H. <strong>Abel</strong>, 1827a).<br />

5 (Lange-<strong>Niels</strong>en, 1927, 65).


19.1. ABEL’s approach to the Paris memoir 349<br />

solubility. One central example was a tacitly employed result which has been<br />

presented in lemma 4, below.<br />

2. Another algebraic result was employed by ABEL to the effect that for any real<br />

polynomial without multiple roots, such as p (x) = ∏ n k=1 (x − xk), �<br />

0,if α ≤ n − 2 and<br />

n x<br />

∑<br />

k=1<br />

α k<br />

p ′ (xk) =<br />

1,if α = n − 1.<br />

This result, which is a consequence <strong>of</strong> the so-called Lagrange interpolation is dis-<br />

cussed in section 19.3.<br />

3. ABEL also borrowed results concerning primitive roots and congruences from<br />

C. F. GAUSS’ (1777–1855) Disquisitiones arithmeticae. Again, we have already<br />

noted how well acquainted ABEL was with this book by GAUSS.<br />

4. Finally, ABEL used expansions into series <strong>of</strong> decreasing powers to great effect.<br />

Such expansions had been forcefully employed in the 17 th century, and ABEL<br />

certainly considered the procedure well established. In ABEL’S work it was com-<br />

bined with a particular emphasis on the coefficient <strong>of</strong> x −1 , the coefficient which<br />

elsewhere, with A.-L. CAUCHY (1789–1857), became known as the residue.<br />

Most <strong>of</strong> these issues are addressed in some details in section 19.3 after ABEL’S use<br />

<strong>of</strong> them in the Paris memoir has been described.<br />

19.1.3 <strong>The</strong> presentational style <strong>of</strong> the Paris memoir<br />

When compared to the other works in ABEL’S corpus, the style <strong>of</strong> the Paris memoir<br />

stands out in a number <strong>of</strong> respects. When compared to the subsequent partial an-<br />

nouncements <strong>of</strong> results contained in the Paris memoir (see section 19.5.1), a pattern be-<br />

comes discernible. At the textual level, ABEL’S papers fell between two traditions, one<br />

mainly based on algebraic manipulations and derivations <strong>of</strong> formulae and an emerg-<br />

ing one returning to the Euclidean norm <strong>of</strong> definitions, theorem statements, and pro<strong>of</strong>s<br />

(see chapter 21). In this continuum <strong>of</strong> styles, the Paris memoir belongs to the manip-<br />

ulation based tradition with its long, tedious, and very explicit derivations <strong>of</strong> explicit<br />

formulae. <strong>The</strong> theorems which I have extracted (Main <strong>The</strong>orems I and II, 16 and 17)<br />

are reconstructions, and reformulating ABEL’S main results in the “If . . . , then . . . ”<br />

structure <strong>of</strong> modern theorem-based mathematics is by no means an easy and trivial<br />

task; the translation from ABEL’S explicit manipulative style to the structure <strong>of</strong> theo-<br />

rems is not a bijection, it requires interpretation.<br />

19.1.4 ABEL’s notational innovations<br />

1. ABEL used a summation shorthand<br />

ΣFx = Fx1 + Fx2 + · · · + Fxn


350 Chapter 19. <strong>The</strong> Paris memoir<br />

An important lemma<br />

Lemma 4 Let χ (y) = 0 be an irreducible equation <strong>of</strong> degree n and let θ (y) be an equation<br />

<strong>of</strong> degree n − 1. <strong>The</strong>n y can be expressed rationally in χ, θ. ✷<br />

PROOF (PROOF OF LEMMA 4) By the Euclidean algorithm, there exist polynomials q, r<br />

such that<br />

χ = qθ + r<br />

where deg r < deg θ. Since χ is irreducible and deg q = deg χ − deg θ = 1 > 0,<br />

deg r > 0. Thus, there exist numbers s, t such that<br />

Consequently,<br />

y =<br />

q (y) = sy + t.<br />

q (y) − t<br />

s<br />

=<br />

χ(y)−r<br />

θ(y)<br />

− t<br />

,<br />

s<br />

and y has been expressed rationally in χ, θ. �<br />

Box 9: An important lemma<br />

which apparently was innovative with him. As is evident from even this exam-<br />

ple, both the index over which the summation is to be performed and the upper<br />

summation limit are implicit in the shorthand version.<br />

2. For a rational function Fx, ABEL let ΠFx denote the coefficient <strong>of</strong> 1 x in the series<br />

expansion <strong>of</strong> Fx in decreasing powers <strong>of</strong> x. Designating the ‘same’ object as the<br />

residue which CAUCHY studied from an emerging perspective <strong>of</strong> his calculus <strong>of</strong><br />

residues, ABEL’S Π corresponds to CAUCHY’S E.<br />

3. ABEL also introduced the operation h on algebraic functions which represented<br />

a general degree <strong>of</strong> algebraic functions.<br />

4. In the ultimate example <strong>of</strong> hyperelliptic integrals, ABEL introduced the notation<br />

EA and εA for A any real number to denote the integer and remaining part, A =<br />

EA + εA (EA ∈ Z and 0 ≤ εA < 1). It is worth remarking that ABEL did not — in<br />

the Paris memoir considered as a whole — apply this notation consistently. Until<br />

the ultimate section, he preferred the verbal formulation ‘the greatest integer<br />

contained in the number’.<br />

All these notational innovations enabled ABEL to comprehend, master, and manip-<br />

ulate objects in a precise way which had hitherto been difficult to obtain.


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 351<br />

19.2 <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong><br />

ABEL’S objective in the memoir was to study integrals <strong>of</strong> the form<br />

�<br />

f (x, y) dx<br />

in which x and y were related by some algebraic equation (19.1) and f was a rational<br />

function. Such integrals provided a way <strong>of</strong> generalizing elliptic integrals; any elliptic in-<br />

tegral could be written in the form above. However, it was not through a direct study<br />

<strong>of</strong> one such integral that something new was to be learned, but by studying relations —<br />

arising from the additional equation (19.3) — among a number <strong>of</strong> such integrals.<br />

<strong>The</strong> contents <strong>of</strong> the Paris memoir can be structured into several results and their<br />

primary applications:<br />

1. <strong>The</strong> establishment <strong>of</strong> Main <strong>The</strong>orem I on integration <strong>of</strong> certain sums <strong>of</strong> algebraic<br />

differentials by elementary functions,<br />

2. <strong>The</strong> establishment <strong>of</strong> Main <strong>The</strong>orem II on the number <strong>of</strong> independent integrals <strong>of</strong><br />

algebraic differentials, and<br />

3. Application <strong>of</strong> Main <strong>The</strong>orem II to the simplest case, the case <strong>of</strong> hyperelliptic<br />

integrals.<br />

In the following, the results and methods <strong>of</strong> first two <strong>of</strong> these three parts will be<br />

described; as will the other instances where ABEL presented his findings on related<br />

issues. ABEL’S reasoning is rather cumbersome and not completely flawless. Some<br />

<strong>of</strong> the subsequent objections and comments — primarily by P. L. M. SYLOW (1832–<br />

1918) — are referred to in the course <strong>of</strong> the presentation. However, despite the reser-<br />

vations, ABEL’S original argument is presented to illustrate how he ingeniously used<br />

the tools at his disposal. Even if the contents and purpose <strong>of</strong> ABEL’S arguments can<br />

seem to evade attention, his various tools and the contents <strong>of</strong> the Paris memoir are<br />

subsequently summarized.<br />

To various degrees <strong>of</strong> authenticity, ABEL’S argument has been described from the<br />

viewpoint <strong>of</strong> the application to hyperelliptic integrals, see e.g. (Brill and Noether, 1894;<br />

Cooke, 1989). However, as will be discussed in section 19.5.1, the chronology and in-<br />

ternal logical structure suggests that the results <strong>of</strong> the Paris memoir were indeed prior<br />

to and to some extent independent <strong>of</strong> the applications to this (afterwards) immensely<br />

important special case.


352 Chapter 19. <strong>The</strong> Paris memoir<br />

19.2.1 Main <strong>The</strong>orem I<br />

In the Paris memoir, 6 ABEL dealt with two quantities x and y related through an irre-<br />

ducible polynomial equation such as<br />

χ (y) = 0, (19.1)<br />

where χ is a polynomial in y whose coefficients are polynomial functions <strong>of</strong> x,<br />

χ (y) =<br />

n<br />

∑ pk (x) y<br />

k=0<br />

k . (19.2)<br />

This relation (19.2) implicitly introduced n functions y (1) , . . . , y (n) <strong>of</strong> x corresponding<br />

to the n roots the equation would have for any particular value <strong>of</strong> x.<br />

Introducing another (later to be specialized) equation in y whose degree was one<br />

less than χ,<br />

ABEL formed the product<br />

θ (y) =<br />

r =<br />

n−1<br />

∑ qk (x) y<br />

k=0<br />

k = 0, (19.3)<br />

n �<br />

∏ θ y<br />

k=1<br />

(k)�<br />

, (19.4)<br />

by inserting the n different solutions <strong>of</strong> (19.2) into (19.3) and multiplying the n results.<br />

<strong>The</strong> coefficients q0, . . . , qn−1 could contain some indeterminate quantities a1, . . . , aN,<br />

and r was found to be an entire function <strong>of</strong> x and a1, . . . , aN by methods “imported”<br />

from the theory <strong>of</strong> equations.<br />

In order to focus attention, ABEL split the product r into parts dependent on and<br />

independent <strong>of</strong> the indeterminate quantities<br />

r = F0 (x) F (x) , (19.5)<br />

where only F depended on a1, . . . , aN. ABEL then considered the equation<br />

F (x) = 0, (19.6)<br />

which would provide expressions for its roots x1, . . . , xµ in terms <strong>of</strong> a1, . . . , aN,<br />

F (x) =<br />

µ<br />

∏ (x − xk) .<br />

k=1<br />

<strong>The</strong>se roots would become very important in the ensuing deductions. 7<br />

6 (N. H. <strong>Abel</strong>, [1826] 1841).<br />

7 At this point it might be fruitful to summarize ABEL’S results this far. He knew that r, defined from<br />

the polynomials θ and χ, was an entire function <strong>of</strong> x and the indeterminates a1, . . . , aN. Any root<br />

<strong>of</strong> the equation r (x) = 0 corresponded to a value y for which θ (y) = 0 by obvious inspection.<br />

However, r (x) = 0 also meant either F0 (x) = 0 or F (x) = 0. <strong>The</strong> former case represented an<br />

equation independent <strong>of</strong> the indeterminates a1, . . . , aN, whereas the latter introduced a relationship<br />

between x and the indeterminates. Thus, with given indeterminates, r (x) = 0 would mean either<br />

that x belonged to the set � �<br />

x1, . . . , xµ or that F0 (x) = 0. To each xk in � �<br />

x1, . . . , xµ corresponded a<br />

value <strong>of</strong> y, which ABEL termed yk, such that θ (yk) = 0.


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 353<br />

After these algebraic operations, ABEL now for the first time employed the calculus<br />

in differentiating the equation (19.6) above. ABEL wrote the differentiation as<br />

F ′ (x) dx + ∂F (x) = 0, (19.7)<br />

where F ′ (x) represents the differential <strong>of</strong> F with respect to x and ∂F (x) represents<br />

the differential <strong>of</strong> F with respect to all the indeterminates. This relationship was a<br />

fundamental one, and ABEL immediately put it to use. He introduced the differential<br />

dv =<br />

µ<br />

∑ f (xk, yk) dxk, k=1<br />

where f was a rational function. This differential was the real object <strong>of</strong> concern in these<br />

investigations. Through a sequence <strong>of</strong> deductions employing the theory <strong>of</strong> equations<br />

(see example below), ABEL reasoned that dv was a rational function <strong>of</strong> the parameters<br />

a1, . . . , aN. <strong>The</strong>refore, its integral v would have to be expressible by algebraic and<br />

logarithmic functions <strong>of</strong> these parameters 8 ,<br />

v =<br />

µ �<br />

∑<br />

k=1<br />

f (x k, y k) dx k = algebraic and logarithmic terms.<br />

This result is what I have termed Main <strong>The</strong>orem I.<br />

<strong>The</strong>orem 16 (Main <strong>The</strong>orem I) Under the present assumptions, the sum<br />

µ �<br />

∑<br />

k=1<br />

f (x k, y k) dx k<br />

can be expressed by algebraic and logarithmic functions <strong>of</strong> the parameters a1, . . . , aN. ✷<br />

An example <strong>of</strong> ABEL’S use <strong>of</strong> the theory <strong>of</strong> equations. In order to see that dv was<br />

indeed a rational function <strong>of</strong> the parameters, ABEL first claimed that the simultaneous<br />

equations (19.1) and (19.3) expressed y k as a rational function <strong>of</strong> x k, 9<br />

y k = ρ (x k) .<br />

Rearranging the equation (19.7) then produced<br />

f (x, y) dx = −<br />

f (x, ρ (x))<br />

F ′ ∂F (x) = φ2 (x) .<br />

(x)<br />

Of this function φ2, ABEL observed that it was obviously rational in x and the param-<br />

eters. Thus, dv could be rewritten as<br />

dv =<br />

µ<br />

∑ φ2 (xk) (19.8)<br />

k=1<br />

8 <strong>The</strong> integral <strong>of</strong> any rational function was <strong>of</strong> course expressible by rational and logarithmic terms.<br />

9 In order to see that ρ is rational as claimed, please observe that deg θ = deg χ − 1. See the pro<strong>of</strong> <strong>of</strong><br />

lemma 4 in box 9.


354 Chapter 19. <strong>The</strong> Paris memoir<br />

where φ2 was a rational function <strong>of</strong> the parameters and its explicit argument 10 . How-<br />

ever, because the right hand side <strong>of</strong> (19.8) was both rational and symmetric in the roots<br />

x1, . . . , xµ <strong>of</strong> the equation (19.6), dv could be expressed rationally in the coefficients <strong>of</strong><br />

F by a basic theorem which ABEL knew from his theory <strong>of</strong> equations (see section 5.2.4).<br />

However, the coefficients <strong>of</strong> F were supposed to depend rationally on the parameters,<br />

and the claim had been demonstrated.<br />

19.2.2 An explicit expression for v<br />

ABEL summarized the results <strong>of</strong> the Main <strong>The</strong>orem I and introduced the way forward<br />

with these words:<br />

“Previously, we have demonstrated how it is always possible to form the rational<br />

differential dv. However, as the indicated method will generally be very<br />

long and nearly impractical for slightly complicated functions, I will give another<br />

[method] by which one will immediately obtain the expression <strong>of</strong> the function v<br />

in all possible cases.” 11<br />

<strong>The</strong> expression for v, which ABEL obtained in “all the possible cases” was <strong>of</strong> the<br />

following form (we shall comment on the particulars and the notation below)<br />

v = C − Πφ (x) +<br />

α<br />

∑<br />

ν=1<br />

ν dν−1φ1 (x)<br />

dxν−1 . (19.9)<br />

First, we will pay some attention to ABEL’S arguments in order to illustrate how they<br />

relate to the deduction <strong>of</strong> Main <strong>The</strong>orem I and how they introduce tools which would<br />

become important in deducing Main <strong>The</strong>orem II (see below).<br />

A sequence <strong>of</strong> manipulations — in which ABEL again made important use <strong>of</strong> his<br />

knowledge from the theory <strong>of</strong> algebraic equations — led ABEL to express the sought<br />

after differential in the following form<br />

dv = − ∑<br />

R1 (x)<br />

A · F ′ (x) · ∏ α k=1 (x − β k) ν k<br />

, (19.10)<br />

where R1 was an entire function and A was a constant. <strong>The</strong> next step was the reduction<br />

<strong>of</strong> this expression. ABEL first revised the notation and wrote<br />

dv = −<br />

µ<br />

∑<br />

k=1<br />

R2 (x k)<br />

F ′ (x k) −<br />

µ<br />

∑<br />

k=1<br />

R3 (xk) θ1 (xk) · F ′ , (19.11)<br />

(xk) 10 ABEL chose to denote this function φ2, although no φ1 had been introduced at this point. This might<br />

suggest that the logical order in ABEL’S head <strong>of</strong> these sections had been reversed in the written<br />

version. See below.<br />

11 “Nous avons montré dans ce qui précède comment on peut toujours former la différentielle rationelle<br />

dv; mais comme la méthode indiquée sera en général très-longue, et pour des fonctions un peu<br />

composées, presque impractible, je vais en donner une autre, par laquelle on obtiendra immédiatement<br />

l’expression de la fonction v dans tous les cas possibles.” (N. H. <strong>Abel</strong>, [1826] 1841, 150).


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 355<br />

where the auxiliary θ1 had been introduced by<br />

and<br />

θ1 (x) = A<br />

α<br />

∏<br />

k=1<br />

(x − β k) ν k ,<br />

R1 (x) = θ1 (x) R2 (x) + R3 (x) , with<br />

deg R3 < deg θ1. (19.12)<br />

<strong>The</strong> introduction <strong>of</strong> the additional auxiliary functions R2 and R3 was made to ease the<br />

study <strong>of</strong> the quotient R 1<br />

θ 1 , which because <strong>of</strong> (19.10) was at the centre <strong>of</strong> ABEL’S interest.<br />

Implicitly using the method <strong>of</strong> Lagrange interpolation, 12 ABEL reduced the first term<br />

<strong>of</strong> (19.11),<br />

µ<br />

R2 (xk) ∑ F<br />

k=1<br />

′ (xk) = Π R2 (x)<br />

F (x)<br />

where the symbol Π was introduced in the following way:<br />

“Thus, in designating by ΠF1x the coefficient <strong>of</strong> 1<br />

x in the development <strong>of</strong> any<br />

function F1x according to decreasing powers <strong>of</strong> x, one will get [. . . ]” 13<br />

Because <strong>of</strong> (19.12), the reduction could be written<br />

µ<br />

R2 (xk) ∑ F<br />

k=1<br />

′ (xk) = Π R1 (x)<br />

θ1 (x) F (x) .<br />

ABEL’S reduction <strong>of</strong> the remaining term <strong>of</strong> (19.11) was marred by a faulty calculation<br />

which has been noticed and elaborated by SYLOW in his notes in the Œuvres. 14<br />

ABEL claimed — by an argument based on expansion into partial fractions — that<br />

R3 (x)<br />

θ1 (x) =<br />

α<br />

d<br />

∑ vk k=1<br />

vk−1 dβvk−1 �<br />

θ (v k)<br />

1<br />

R3 (β)<br />

(β) · (x − β)<br />

�<br />

β=β k<br />

. (19.13)<br />

However, during his deductions concerning partial fractions he had mistakenly placed<br />

the factor (x − β) in the denominator instead <strong>of</strong> in the numerator. 15 This flaw perme-<br />

ated the ensuing calculations, and a simple counter example could serve to demon-<br />

strate that the result claimed in (19.13) is only valid in some very particular cases.<br />

12 See below.<br />

13 “En désignant donc par ΠF1x le coefficient de 1 x dans le développement d’une fonction quelconque<br />

F1x, suivant les puissances descendantes de x, on aura [. . . ]” (ibid., 155).<br />

14 (Sylow in N. H. <strong>Abel</strong>, 1881, II, 295–296).<br />

15 <strong>The</strong> mistake is indeed ABEL’S which can be seen from the fact that it occurs in the manuscript <strong>of</strong> the<br />

Paris memoir.


356 Chapter 19. <strong>The</strong> Paris memoir<br />

ABEL and the method <strong>of</strong> Lagrange interpolation. Without any specific reference,<br />

ABEL employed a result to the effect that<br />

n<br />

∑<br />

k=1<br />

p (xk) χ ′ (xk) =<br />

�<br />

0if deg p < n − 1<br />

1if deg p = n − 1<br />

for any normed polynomial p where x1, . . . , xn are the roots <strong>of</strong> the polynomial equation<br />

χ (x) = 0, i.e.<br />

χ (x) =<br />

n<br />

∏ (x − xk) . (19.14)<br />

k=1<br />

<strong>The</strong> tool behind this result is known today as Lagrange interpolation, and — in various<br />

forms — it played central roles in ABEL’S arguments in the Paris memoir. Lagrange<br />

interpolation is used to demonstrate that for any polynomial such as (19.14),<br />

1<br />

χ (x) =<br />

n<br />

∑<br />

k=1<br />

1<br />

(x − xk) χ ′ . (19.15)<br />

(xk) Expansion into partial fractions using Lagrange interpolation. <strong>The</strong> method <strong>of</strong> ex-<br />

panding a quotient <strong>of</strong> polynomials into partial fractions was well established in the<br />

18 th century once it was known that the denominator could be decomposed into a<br />

product <strong>of</strong> linear and quadratic terms. <strong>The</strong> generality <strong>of</strong> the method thus rested es-<br />

sentially on the Fundamental <strong>The</strong>orem <strong>of</strong> Algebra, and the pro<strong>of</strong> <strong>of</strong> the latter theorem was<br />

<strong>of</strong>ten seen mainly as a prerequisite in rigorously founding this established practice (cf.<br />

GAUSS).<br />

<strong>The</strong> central trick in expanding a quotient into partial fractions is closely related to<br />

the method <strong>of</strong> Lagrange interpolation. If the polynomials are<br />

f2 (x) =<br />

f1 (x) and<br />

n<br />

∏ (x − xk) ,<br />

k=1<br />

where the roots <strong>of</strong> f2 are distinct, Lagrange interpolation (19.15) yields<br />

f1 (x)<br />

f2 (x) =<br />

n<br />

∑<br />

k=1<br />

f1 (x)<br />

(x − x k) f ′ 2 (x k) .<br />

Thus, when applied in the integral calculus, this formula reduces the integration <strong>of</strong><br />

a fraction to the integration <strong>of</strong> n fractions, the denominator <strong>of</strong> each <strong>of</strong> which only<br />

contains a first degree polynomial.


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 357<br />

ABEL’S result is faulty when the denominators have multiple roots. For situations<br />

in which the polynomial f2 has multiple roots, the procedure can be extended to ac-<br />

commodate this case as well. <strong>The</strong> flawed result which ABEL used (see above) was for<br />

a general rational function<br />

with<br />

the fraction could be expanded as<br />

where<br />

A m k =<br />

f2 (x) =<br />

f1 (x)<br />

f2 (x) =<br />

f1 (x)<br />

f2 (x)<br />

n<br />

∏ (x − xk) k=1<br />

mk ,<br />

n mk A<br />

∑ ∑<br />

k=1 m=1<br />

m k<br />

(x − xk) m<br />

d m k−1−m pk<br />

Γ (m k + 1 − m) dβ m k−1−m ,<br />

pk = Γ (mk − 1) f1 (βk) .<br />

(βk) f (m k)<br />

2<br />

However, ABEL’S formulae for the coefficients were wrong; they should have been<br />

A m k =<br />

1 d<br />

Γ (mk + 1 − m)<br />

mk−m dβm � m �<br />

k (x − β) R3 (x)<br />

,<br />

k−m θ1 (x)<br />

as SYLOW pointed out. 16 Incidentally, this formula was very similar to ABEL’S starting<br />

point:<br />

where<br />

““[. . . ] one will get<br />

for x = β; [. . . ]” 17<br />

A1 = dν−1 p<br />

Γν · dβ ν−1 , A2 =<br />

p = (x − β)ν R3x<br />

θ1x<br />

16 (Sylow in N. H. <strong>Abel</strong>, 1881, II, 295).<br />

17 “[. . . ] on aura<br />

où<br />

A1 = dν−1 p<br />

Γν · dβ ν−1 , A2 =<br />

p = (x − β)ν R3x<br />

θ1x<br />

pour x = β; [. . . ]” (N. H. <strong>Abel</strong>, [1826] 1841, 156).<br />

d ν−2 p<br />

Γ (v − 1) dβ ν−2 , . . . , Aν = p,<br />

d ν−2 p<br />

Γ (v − 1) dβ ν−2 , . . . , Aν = p,<br />

β=β k


358 Chapter 19. <strong>The</strong> Paris memoir<br />

<strong>The</strong> ensuing step was, however, unwarranted as ABEL claimed that<br />

p = Γ (ν + 1) R3 (β)<br />

θ (ν)<br />

.<br />

1 (β)<br />

One can guess how ABEL came to the latter belief by applying the rule <strong>of</strong> G.-F.-A. DE<br />

L’HOSPITAL (1661–1704) v times to the definition <strong>of</strong> p as both numerator and denomi-<br />

nator vanish. In the above presentation, the problem which ABEL’S deduction suffered<br />

from is hidden in the notation. First <strong>of</strong> all, ABEL’S way <strong>of</strong> suppressing the subscript<br />

k has made the β and x appear symbolically similar, although x is a true variable<br />

whereas β1, . . . , βα are the roots <strong>of</strong> a certain polynomial. This distinction is at the core<br />

<strong>of</strong> SYLOW’S objection to ABEL’S argument. 18 However, with a minor adjustment to<br />

the definitions, ABEL’S final product (19.9) <strong>of</strong> the argument could be allowed.<br />

19.2.3 Main <strong>The</strong>orem II<br />

After the first four sections <strong>of</strong> the Paris memoir, ABEL had thus obtained a formula<br />

which was essentially (apart from the corrections indicated above) the following,<br />

v =<br />

µ<br />

µ �<br />

∑ ψ (xk) = ∑<br />

k=1<br />

k=1<br />

f (x k, y k) dx k = C − Πφ (x) +<br />

α<br />

∑<br />

ν=1<br />

ν dν−1φ1 (x)<br />

dxν−1 .<br />

This expression allowed him to commence a study <strong>of</strong> the number <strong>of</strong> free parameters<br />

which would eventually lead to the second main theorem — the celebrated <strong>Abel</strong>ian<br />

<strong>The</strong>orem. To follow his argument, we need to backtrack a little to properly understand<br />

the use <strong>of</strong> the eliminant equation r = 0 (see page 352).<br />

ABEL’S trick was first to study the consequences <strong>of</strong> one further assumption con-<br />

cerning the factor F0 <strong>of</strong> r containing the indeterminate quantities. ABEL assumed that<br />

F0 had α distinct zeros,<br />

α<br />

F0 (x) = ∏ (x − βk) k=1<br />

µ k ;<br />

an assumption which — provided F0 is not a constant — introduced α linear interrelations<br />

among the coefficients q0, . . . , qn−1 <strong>of</strong> the auxiliary polynomial θ (y) (19.3). 19<br />

ABEL found the fact that the coefficients <strong>of</strong> θ (y) formed a non-independent set to be —<br />

in general — a contraction <strong>of</strong> the original hypothesis which required nothing <strong>of</strong> the co-<br />

efficients q0, . . . , qn−1. Consequently, he concluded that F0 (x) had to be a constant and<br />

r (x) could not — in general — contain any factor independent <strong>of</strong> the auxiliary quanti-<br />

ties.<br />

Under this assumption, ABEL proceeded to describe the various functions involved.<br />

<strong>The</strong> important outcome <strong>of</strong> these investigations was that a certain function f2 (x) in-<br />

troduced much earlier in the investigations, reduced to unity, thereby providing the<br />

result that f (x, y) χ ′ (y) equated the entire function f1 (x, y).<br />

18 (Sylow in N. H. <strong>Abel</strong>, 1881, II, 295).<br />

19 See box 19.2.3.


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 359<br />

Linear interdependence <strong>of</strong> q0, . . . , qn−1 In order to see how ABEL obtained the linear<br />

interrelations among the coefficients <strong>of</strong> θ (y), we notice from combining the factoriza-<br />

tion (19.5) with the definition <strong>of</strong> r (19.4),<br />

r (x) = F (x) · F0 (x)<br />

=<br />

n<br />

∏ θ<br />

k=1<br />

�<br />

y (k) (x)<br />

�<br />

=<br />

n n−1<br />

∏ ∑<br />

k=1 m=0<br />

�<br />

qm (x) y (k) �m (x) .<br />

Thus, since r (β1) = F (β1) · 0 = 0, some k must exist for which θ<br />

n−1 �<br />

∑ qm (x) y<br />

m=0<br />

(k) �m (x) = 0.<br />

�<br />

y (k) �<br />

(β1) = 0, i.e.<br />

This relation is a linear interdependence among the q0, . . . , qn−1 in which the<br />

here serving as coefficients, are functions <strong>of</strong> x.<br />

Box 10: Linear interdependence <strong>of</strong> q0, . . . , qn−1<br />

�<br />

y (k)� m<br />

,<br />

ABEL’S way <strong>of</strong> proving f (x, y) χ ′ (y) to be an entire function. ABEL’S investiga-<br />

tions leading to the result that f (x, y) χ ′ (y) was an entire function progressed along<br />

complicated and tedious arguments. At the very outset <strong>of</strong> the paper, ABEL had split<br />

the rational function <strong>of</strong> x in the following way<br />

f (x, y) χ ′ (y) = f1 (x, y)<br />

, (19.16)<br />

f2 (x)<br />

where f2 (x) was an entire function <strong>of</strong> x independent <strong>of</strong> y. By combining this with the<br />

important partial differentiation (19.7), ABEL found<br />

f (x, y) dx =<br />

f1 (x, y)<br />

f2 (x) · χ ′ (y) dx<br />

1<br />

= −<br />

F0 (x) · F ′ (x) · f2 (x)<br />

n<br />

∑<br />

k=1<br />

f1 (x, y k)<br />

χ ′ (y k)<br />

r<br />

θ (y k) ∂θ (y k) .<br />

Later, during his investigations leading to the Main <strong>The</strong>orem I, ABEL had studied the<br />

function<br />

which he had chosen to write as<br />

f1 (x, y)<br />

f1 (x, y)<br />

r<br />

∂θ (y)<br />

θ (y)<br />

r<br />

θ (y) ∂θ (y) = R′ (y) + R (x) y n−1 ,<br />

where R ′ (y) indicated an entire function <strong>of</strong> x and y in which no powers <strong>of</strong> y beyond<br />

the (n − 2)’nd occur, and R (x) was an entire function <strong>of</strong> x independent <strong>of</strong> y. By use <strong>of</strong>


360 Chapter 19. <strong>The</strong> Paris memoir<br />

the method <strong>of</strong> Lagrange interpolation described above and the factorization <strong>of</strong> r (19.5),<br />

ABEL found<br />

R (x) = F0 (x) · F (x) ·<br />

n<br />

∑<br />

k=1<br />

f1 (x, y k)<br />

χ ′ (y k)<br />

∂θ (y k)<br />

θ (y k) .<br />

Under the present assumption F0 (x) = 1, ABEL concluded that<br />

φ1 (x) =<br />

1<br />

f (m)<br />

n<br />

∑<br />

2 (x) k=1<br />

f1 (x, y k)<br />

χ ′ (y k) log θ (y k) ,<br />

and consequently, the sum <strong>of</strong> integrals took the form (cmp. equation 19.9)<br />

�<br />

f1 (x, y) dx<br />

∑ f2 (x) χ ′ (y) = C − Π ∑ f1 (x, y)<br />

f2 (x) χ ′ (y)<br />

+ ∑ m dm−1<br />

dβ m−1<br />

log θ (y)<br />

�<br />

1<br />

f (m)<br />

2 (β) ∑ f1 (β, B)<br />

χ ′ �<br />

log θ (B)<br />

(B)<br />

. (19.17)<br />

After inspecting the results <strong>of</strong> certain simple assumptions concerning f2, ABEL argued:<br />

“In the equation (19.17), the right-hand-side is in general a function <strong>of</strong> the<br />

quantities a, a ′ , a ′′ , etc. If one supposes this function equal to a constant, certain<br />

relations among these quantities thus generally result; but there are also certain<br />

cases for which the right-hand-side reduces to a constant no matter what the values<br />

<strong>of</strong> the quantities a, a ′ , a ′′ , etc. are. We investigate this case:<br />

From this it is evident that the function f2x must be constant, because in the<br />

contrary case the right-hand-side necessarily contains the quantities a, a ′ , a ′′ . . . ,<br />

with respect to the arbitrary values <strong>of</strong> these quantities.” 20<br />

ABEL’S argument here seems a little roundabout; in the cause <strong>of</strong> argument he in-<br />

troduced an important assumption — that v now reduces to a constant. He argued<br />

that unless f2 then also reduced to a constant, the right hand side <strong>of</strong> (19.17) would<br />

involve the auxiliary quantities, whereas the left hand side — which was nothing but<br />

v = ∑ � f (x, y) dx by (19.16) — was now a constant. Thus, unless f2 was constant, cer-<br />

tain relations among the indeterminates a, a ′ , a ′′ , . . . would result — a contradiction. In<br />

the end, ABEL had obtained a representation <strong>of</strong> the constant v in the following form<br />

�<br />

f1 (x, y) dx<br />

∑ χ ′ (y)<br />

= C − ∑ Π f1 (x, y)<br />

χ ′ (y)<br />

log θ (y) .<br />

If we pause for a second to consider what the consequences <strong>of</strong> this new hypothesis,<br />

the constancy <strong>of</strong> v, are, we can suggest some motivation for this at first sight rather un-<br />

natural assumption which will be elaborated during the uncovering <strong>of</strong> the remaining<br />

20 “Dans le formule (43) [here (19.17)], le second membre est en général une fonction des quantités<br />

a, a ′ , a ′′ , etc. Si on le suppose égal à une constante, il en résultera donc en général certaines relations<br />

entre ces quantités; mais il y a aussi certaines cas pour lesquels le second membre se réduit à une<br />

constante, quelles que soient d’ailleurs les valeurs des quantités a, a ′ , a ′′ , etc. Cherchons ces cas:<br />

D’abord il est évident que la fonction f2x doit être constante, car dans le cas contraire le second<br />

membre contiendrait nécessairement les quantités a, a ′ , a ′′ . . . , vu les valeurs arbitraires de ces quantités.”<br />

(N. H. <strong>Abel</strong>, [1826] 1841, 161).


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 361<br />

details. If v is a constant, say v = 0, the basic objects <strong>of</strong> the inquiry satisfy<br />

�<br />

0 = ∑<br />

f (x k, y k) dx k.<br />

<strong>The</strong> end product <strong>of</strong> the present section <strong>of</strong> the paper is the <strong>Abel</strong>ian <strong>The</strong>orem (Main <strong>The</strong>o-<br />

rem II) states something about exactly such sums <strong>of</strong> related integrals.<br />

ABEL next expanded the relevant function, whose derivative was rational in x by<br />

(19.8), according to decreasing powers <strong>of</strong> x,<br />

∑ f1 (x, y)<br />

χ ′ (y)<br />

log θ (y) = R log x +<br />

∞<br />

∑ Akx k=0<br />

µ0−k<br />

, (19.18)<br />

where R was “a function <strong>of</strong> x independent <strong>of</strong> a, a ′ , a ′′ , etc.,” 21 A0, A1, . . . were independent<br />

<strong>of</strong> x, and µ0 designated an integer. 22<br />

If expression (19.18) were to express a constant (independent <strong>of</strong> the indeterminates<br />

a, a ′ , a ′′ , . . . ), ABEL observed that since these quantities occurred in A0, A1, . . . , the sec-<br />

ond term corresponding to these had to vanish. And since he was only concerned<br />

with the coefficient <strong>of</strong> 1 x , his conclusion was that µ0 < −1. He expressed this using a<br />

newly introduced notational advance in the following sentence:<br />

“This done, in designating by the symbol hR the highest exponent <strong>of</strong> x in the<br />

development <strong>of</strong> any function R <strong>of</strong> this quantity following decreasing powers, it is<br />

evident that µ0 will be equal to the largest integer contained in [less than or equal<br />

to] the numbers<br />

h f1 (x, y ′ )<br />

χ ′ y ′<br />

, h f1 (x, y ′′ )<br />

χ ′ y ′′<br />

�<br />

f1<br />

, . . . h<br />

x, y (n)�<br />

χ ′ y (n)<br />

.<br />

It is necessary that all these numbers must be less than the unit taken negatively.” 23<br />

21 “R étant une fonction de x indépendante de a, a ′ , a ′′ , etc.” (ibid., 161).<br />

22 <strong>The</strong> version printed in the Savants étrangers read at this point<br />

⎧<br />

⎨<br />

R log x =<br />

⎩<br />

+Aµ0<br />

A0x µ0 + A1x µ0−1 + . . .<br />

+ Aµ0+1<br />

x<br />

+ Aµ0+2<br />

x 2<br />

+ . . .<br />

In the collected works (Sylow in N. H. <strong>Abel</strong>, 1881, II, 296), SYLOW commented: “C’est évidemment<br />

une faute d’écriture, ou d’<strong>Abel</strong> ou de Libri.” After the original manuscript has been recovered, it<br />

has become evident that the misprint is indeed due to LIBRI.<br />

23 “Cela posé, en désignant par le symbole hR le plus haut exposant de x dans le développement d’une<br />

fonction quelconque R de cette quantité, suivant les puissances descendantes, il est clair que µ0 sera<br />

égal au nombre entier le plus grand contenu dans les nombres:<br />

h f1 (x, y ′ )<br />

χ ′ y ′ , h f1 (x, y ′′ )<br />

χ ′ y ′′ �<br />

f1 x, y<br />

, . . . h<br />

(n)�<br />

χ ′ y (n)<br />

;<br />

il faut donc que tous ces nombres soient inférieurs à l’unité prise négativement.” (N. H. <strong>Abel</strong>, [1826]<br />

1841, 161).


362 Chapter 19. <strong>The</strong> Paris memoir<br />

Thus, for the individual terms <strong>of</strong> the sum, which were in general just algebraic func-<br />

tions <strong>of</strong> x, the ‘degree’ hR needed not be an integer, whence the conclusion<br />

h f1 (x, y k)<br />

χ ′ (y k)<br />

< −1. (19.19)<br />

<strong>The</strong> investigation now turned to algebraic manipulations <strong>of</strong> these new symbols.<br />

Determination <strong>of</strong> the most general form <strong>of</strong> the function f1 (x, y). ABEL put his new<br />

tool to immediate use. From the general formula<br />

h R1<br />

R2<br />

he derived from (19.19) the inequalities<br />

= hR1 − hR2,<br />

h f1 (x, y k) < hχ ′ (y k) − 1,<br />

which he claimed made “it easy to deduce the most general form <strong>of</strong> the function<br />

f1 (x, y) in each particular case.” 24 ABEL’S argument — but not its result — has been<br />

found unrigorous at this point, see e.g. SYLOW’S notes, the paper by ELLIOT, and be-<br />

low. 25 However, it is worth following the steps <strong>of</strong> his argument to see how he went<br />

about it.<br />

Because<br />

ABEL obtained<br />

χ ′ (yk) = ∏ (yk − ym) ,<br />

m�=k<br />

hχ ′ (yk) = ∑ h (yk − ym) ,<br />

m�=k<br />

and when the y1, . . . , y k were ordered according to decreasing degrees,<br />

hy k ≥ hym if k ≤ m,<br />

he found “in general, except for certain particular cases which he did not consider:” 26<br />

h (y k − ym) = hy min(k,m).<br />

<strong>The</strong> analogy with the ordinary degree operator makes the above-mentioned par-<br />

ticular cases easy to illustrate. For instance, if we have two monic polynomials <strong>of</strong> the<br />

same degree, the degree <strong>of</strong> their difference is strictly less than either <strong>of</strong> the original<br />

degrees,<br />

��<br />

deg x 2 � �<br />

+ x − 1 − x 2 ��<br />

− x + 2 = deg (2x − 3) = 1.<br />

24 “De ces inégalités on déduira facilement dans chaque cas particulier la forme la plus générale de la<br />

fonction f1 (x, y).” (N. H. <strong>Abel</strong>, [1826] 1841, 162).<br />

25 (Sylow in N. H. <strong>Abel</strong>, 1881, II, 296–297) and (Elliot, 1876, 404–406).<br />

26 “Alors on aura, en général, excepté quelques cas particuliers que je me dispense de considérer:”<br />

(N. H. <strong>Abel</strong>, [1826] 1841, 162).


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 363<br />

Later, in chapter 21, we shall have more to say on this notion <strong>of</strong> ‘in general, except for<br />

certain particular cases.’<br />

Writing f1 (x, y), which was by construction an entire function <strong>of</strong> y, in the way<br />

ABEL concluded<br />

Based on the identity<br />

this evolved into<br />

f1 (x, y) =<br />

n−1<br />

∑ tm (x) y<br />

m=0<br />

m ,<br />

htmy m k < hχ′ (y k) − 1 for k = 1, . . . , n and m = 0, . . . , n − 1.<br />

h (tmy m ) = ht k + mhy,<br />

htm + mhy < hχ ′ (y k) − 1.<br />

ABEL now combined the information contained in the equations () obtaining<br />

hχ ′ (y k) − mhy k − 1 = (n − m − k) hy k +<br />

k−1<br />

∑ hyu − 1.<br />

u=1<br />

Since y1, . . . , yn were assumed to be ordered according to decreasing degrees, the min-<br />

imal value among these (over k = 1, . . . , n) was obtained for k = n − m, resulting in<br />

the value<br />

� ′<br />

min hχ (yk) − mhyk − 1<br />

k=1,...,n<br />

� = hχ ′ n−m−1<br />

(yn−m) − mhyn−m − 1 =<br />

<strong>The</strong>refore, because htm was an integer,<br />

where 0 < εn−m−1 ≤ 1.<br />

htm =<br />

∑<br />

u=1<br />

hyu − 1.<br />

n−m−1<br />

∑ hyu − 2 + εn−m−1, (19.20)<br />

u=1<br />

Grouping <strong>of</strong> roots according to their degree. <strong>The</strong> next step in ABEL’S analysis was<br />

to write<br />

from which he obtained<br />

hy1 = m1<br />

µ1<br />

with (m1, µ1) = 1<br />

hy1 = hy2 = · · · = hyµ 1 = m1<br />

by an argument involving tools from the theory <strong>of</strong> equations. In his investigations<br />

on algebraic solubility <strong>of</strong> equations, ABEL had proved (see e.g. chapter 8) that if an<br />

equation (here χ (y) = 0) was satisfied by an expression such as y = Ax<br />

µ1<br />

m1 µ 1 , an entire


364 Chapter 19. <strong>The</strong> Paris memoir<br />

sequence <strong>of</strong> distinct roots could be obtained by substituting for x 1<br />

µ 1 in y the result <strong>of</strong><br />

x 1<br />

µ 1 multiplied with the different µ1’th roots <strong>of</strong> unity. <strong>The</strong>refore, the roots y1, . . . , yn<br />

fell into sequences with equal degrees,<br />

τ’th sequence: hy kτ+1 = hy kτ+2 = · · · = hy kτ+nτµτ<br />

n =<br />

ε<br />

∑ nτµτ.<br />

τ=1<br />

mτ<br />

= , (mτ, µτ) = 1,<br />

µτ<br />

When focusing his attention on an root ym belonging to the τ’th sequence<br />

ABEL found from (19.20)<br />

m = kτ + 1 + β, with 0 ≤ β ≤ kτ+1 − kτ,<br />

htn−m = ht n−kτ−β−1 =<br />

=<br />

=<br />

=<br />

τ−1<br />

∑<br />

α=1<br />

k α+1−kα<br />

∑<br />

τ=1<br />

kτ+β<br />

∑ hyu − 2 + εkτ+β u=1<br />

β<br />

∑ hykτ+τ − 2 + εkτ+β τ=1<br />

hy kα+τ +<br />

τ−1<br />

∑ (kα+1 − kα)<br />

α=1<br />

mα<br />

+ β<br />

µα<br />

mτ<br />

µτ<br />

τ−1<br />

∑ nαmα + β<br />

α=1<br />

mτ<br />

µτ<br />

− 2 + ε kτ+β<br />

+ ε kτ+β − 2 (19.21)<br />

Number-theoretic arguments to determine the form <strong>of</strong> f1 (x, y). <strong>The</strong> product ε kτ+β ·<br />

µτ, for which ABEL introduced a special symbol 27 A τ,β, was found to be the least<br />

positive number ζ for which<br />

µτ | βmτ + ζ. (19.22)<br />

In order for us to see that A τ,β has the prescribed property, it suffices to first observe<br />

from (19.21) that<br />

β mτ<br />

µτ<br />

+ ε kτ+β = βmτ + A τ,β<br />

µτ<br />

is an integer because tm is an entire function. Next, if we assume<br />

we obtain<br />

βmτ + ζ<br />

µτ<br />

= K < K ′ = βmτ + Aτ,β ,<br />

µτ<br />

� K ′ − K � µτ = A τ,β − ζ = µτε kτ+β − ζ<br />

27 Actually ABEL would write A (γ)<br />

β for this quantity, but I have chosen to move indices into subscripts.


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 365<br />

and thus<br />

ε kτ+β = � K ′ − K � + ζ<br />

which is a contradiction. Thus K = K ′ and A τ,β is the smallest number such that<br />

(19.22) holds. This condition <strong>of</strong> minimality, which ABEL just noticed as a matter <strong>of</strong><br />

µτ<br />

> 1<br />

fact, was soon (see below) invoked and found to be <strong>of</strong> great use.<br />

found<br />

When ABEL spelled out the results obtained above for the first sequence τ = 1, he<br />

ht n−β−1 = −2 + β m1<br />

µ1<br />

+ A 1,β<br />

µ1<br />

= −2 + βm1 + A1,β .<br />

µ1<br />

For ‘small’ β (starting with β = 0), the right hand side is obviously negative, because<br />

A 1,β < µ1 by the definition <strong>of</strong> ε kτ+β<br />

A 1,β<br />

µ1<br />

= ε β ≤ 1<br />

and the other term vanishes for β = 0. Consequently, some β ′ ≥ 0 existed such that<br />

which obviously meant that<br />

ht n−β−1 < 0 for β = 0, . . . β ′<br />

(19.23)<br />

t n−β−1 = 0 for β = 0, . . . , β ′ . (19.24)<br />

<strong>The</strong> general form <strong>of</strong> the function f1 (x, y) in the light <strong>of</strong> these results thus became<br />

f1 (x, y) =<br />

n−β ′ −1<br />

∑ tk (x) y<br />

k=0<br />

k<br />

in which β ′ was the largest integer less than µ 1<br />

m + 1,<br />

1<br />

� �<br />

µ1<br />

β ′ =<br />

m1<br />

+ 1<br />

.<br />

(19.25)<br />

ABEL’S way <strong>of</strong> obtaining this ultimate description <strong>of</strong> β ′ was found by SYLOW to miss<br />

certain particular cases. 28<br />

Based on the expression (19.25) which ABEL had obtained for the function f1 (x, y),<br />

he concluded:<br />

“A function like f1 (x, y) always exists when β does not surpass n − 1.” 29<br />

It is difficult to see exactly what ABEL meant by this phrase, which seems to infer<br />

that the existence <strong>of</strong> the function was deduced from the representation (19.25). How-<br />

ever, on a logical basis, the existence <strong>of</strong> the function f1 (x, y) had been presupposed in<br />

the decomposition <strong>of</strong> f (x, y) χ ′ (y), see (19.16).<br />

28 (Sylow in N. H. <strong>Abel</strong>, 1881, II, 298).<br />

29 “Une fonction telle que f1 (x, y) existe donc toujours à moins que β ′ ne surpasse n − 1.” (N. H. <strong>Abel</strong>,<br />

[1826] 1841, 166).


366 Chapter 19. <strong>The</strong> Paris memoir<br />

Investigation <strong>of</strong> the complementary case β ′ ≥ n. In order to study what happened<br />

if β ′ ≥ n, ABEL wrote<br />

µ1<br />

m1<br />

+ 1 = n + ε, ε ≥ 0<br />

and obtained for the inverse fraction only two possibilities<br />

m1<br />

µ1<br />

= 1 m1<br />

or<br />

n − 1 µ1<br />

= 1<br />

n .<br />

In both cases, ABEL claimed, the integral � f (x, y) dx would be expressible in alge-<br />

braic and logarithmic terms. His argument proceeded from claiming that the equation<br />

χ (y) = 0 was linear in x,<br />

χ (y) = P (y) + xQ (y) .<br />

If this was the case, the integrand in � f (x, y) dx was quickly seen to be a rational<br />

function, and the result was thus well known. However, as SYLOW has observed, 30<br />

in the case left unnoticed above, the conclusion <strong>of</strong> linearity does not hold, and the<br />

deduction thus suffers from this incompleteness.<br />

Back on track: the case β ′ ≤ n − 1. Returning to the more complicated case, ABEL<br />

noticed: “Thus, except for this case [β ′ ≥ n], the function f1 (x, y) always exists” 31 and<br />

he went on to elaborate the consequences <strong>of</strong> the hypothesis β ′ ≤ n − 1. He began by<br />

reducing the study <strong>of</strong> the equation<br />

to the study <strong>of</strong> the individual terms<br />

� n−β<br />

∑<br />

∑<br />

′ −1<br />

m=0 tmym χ ′ dx = C (19.26)<br />

(y)<br />

�<br />

xkym dx<br />

∑ χ ′ (y) .<br />

His next and decisive step was to begin considering the htm + 1 coefficients in the<br />

polynomial tm. He found that the function f1 (x, y) contained<br />

n−β ′ −1<br />

∑ (htm + 1) =<br />

m=0<br />

n−β ′ −1<br />

∑ htm + n − β<br />

m=0<br />

′ =<br />

n−2<br />

∑ htm + n − 1 (19.27)<br />

m=0<br />

coefficients and chose to designate this number <strong>of</strong> coefficients by γ. Once this number<br />

had been introduced, it became ABEL’S first objective to derive other general formu-<br />

lae for it and to study certain particular cases. Once these investigations had been<br />

concluded, ABEL again returned to (19.26), remarking that it was even valid in certain<br />

cases not included in the deduction:<br />

30 (Sylow in N. H. <strong>Abel</strong>, 1881, II, 298).<br />

31 “Excepté ce cas donc, la fonction f1 (x, y) existe toujours” (N. H. <strong>Abel</strong>, [1826] 1841, 167).


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 367<br />

“<strong>The</strong> formula (59) [here (19.26)] is generally valid for all values <strong>of</strong> the quanti-<br />

ties a, a ′ , a ′′ , . . . whenever the function r does not have a factor <strong>of</strong> the form F0x; in<br />

that case, it is also valid if F0x and χ′ y<br />

f1(x,y)<br />

vanish for the same value <strong>of</strong> x.”32<br />

Algebraic manipulations pertaining to the number γ. ABEL’S tedious manipula-<br />

tions <strong>of</strong> the expression for γ made critical use <strong>of</strong> a result in the line <strong>of</strong> GAUSS’ theory<br />

<strong>of</strong> moduli and primitive roots — referred to by ABEL’S as “the theory <strong>of</strong> numbers” 33 —<br />

as presented in GAUSS’ Disquisitiones arithmeticae. 34 ABEL found that because for the<br />

τ’th sequence mτ and µτ were relatively prime, and A τ,β ≡ −βmτ (mod µτ),<br />

nτµτ−1<br />

∑<br />

β=0<br />

A τ,β = nτ<br />

� �<br />

µτ−1<br />

∑ k = nτ<br />

k=0<br />

µτ (µτ − 1)<br />

.<br />

2<br />

By combining this with other previously established formulae, ABEL found<br />

γ = 1 +<br />

�<br />

ε τ−1<br />

∑ nτµτ<br />

τ=1<br />

∑ nvmv +<br />

v=1<br />

mτnτ − 1<br />

2<br />

�<br />

−<br />

ε<br />

∑<br />

τ=1<br />

nτ (mτ + 1)<br />

. (19.28)<br />

2<br />

At this point, two particular cases were noticed mainly as examples <strong>of</strong> how to calculate<br />

with the formula (see below).<br />

ABEL’S examples <strong>of</strong> calculating γ. After deducing the formula (19.28) by algebraic<br />

and number theoretic manipulations, ABEL gave some examples <strong>of</strong> how it could be<br />

used in particular cases to determine γ.<br />

1. If all the roots y1, . . . , yn have the same degree (ε = 1)<br />

the expression for γ reduced to<br />

hy1 = · · · = hyn = m1<br />

,<br />

m1n1 − 1<br />

γ = 1 + n1µ1<br />

2<br />

µ1<br />

− n1 (m1 + 1)<br />

.<br />

2<br />

If, furthermore, µ1 = n = n1, corresponding to the situation in which all the<br />

roots y1, . . . , yn involve n’th roots <strong>of</strong> x, it reduced further into<br />

γ = (n − 1) m1 − 1<br />

.<br />

2<br />

32 “La formule (59) a généralement lieu pour des valeurs quelconques des quantités a, a ′ , a ′′ , . . . toutes<br />

les fois que la fonction r n’a pas un facteur de la forme F0x; mais dans ce cas elle a encore lieu, sinon<br />

χ ′ y<br />

f1(x,y) 33<br />

F0x et s’évanouissement pour une même valeur de x.” (ibid., 169).<br />

(ibid., 168)<br />

34 (C. F. Gauss, 1801). See also the discussion in section 5.3.


368 Chapter 19. <strong>The</strong> Paris memoir<br />

2. If all the roots y1, . . . , yn have integer degrees,<br />

and<br />

the formula became (ε = n)<br />

γ = 1 +<br />

= 1 +<br />

�<br />

n τ−1<br />

∑<br />

τ=1<br />

n<br />

∑<br />

hy1, . . . , hyn ∈ Z,<br />

n1 = · · · = nε = 1,<br />

∑ mν +<br />

ν=1<br />

mτ − 1<br />

2<br />

τ−1<br />

∑<br />

τ=1 ν=1<br />

�<br />

mν − n = 1 − n +<br />

−<br />

n<br />

mτ + 1<br />

∑ 2 τ=1<br />

n<br />

∑ (n − τ) mτ.<br />

τ=1<br />

It is interesting to consider the usefulness <strong>of</strong> these examples. It appears that the<br />

examples were both chosen because the assumptions made therein corresponded to<br />

particularly interesting cases and because they illustrate cases, in which the rather<br />

complicated formula (19.28) — which looked even more complicated in ABEL’S nota-<br />

tion than in my modern one — reduced to extremely simple forms. <strong>The</strong> first class <strong>of</strong><br />

equations considered (in which all degrees were equal) contains equations such as<br />

χ (x, y) = y n − p (x) = 0<br />

in which p is a polynomial. <strong>The</strong> second assumption (all roots have integer degrees)<br />

applies to equations <strong>of</strong> the form<br />

χ (x, y) = ∏ k<br />

(y − p k (x)) = 0.<br />

<strong>The</strong> indeterminates a, a ′ , a ′′ , . . . . ABEL chose to designate by α the number <strong>of</strong> inde-<br />

terminates a, a ′ , a ′′ , . . . and ventured to investigate the relationships between the roots<br />

x1, . . . , xµ and the indeterminates a1, . . . , aα. To the α indeterminates corresponded α<br />

equations<br />

θ (yτ) = 0 for τ = 1, . . . , α<br />

which were linear in the indeterminates (see box 19.2.3, above). <strong>The</strong>se equations “in<br />

general” served to express the indeterminates rationally in x1, . . . , xα and y1, . . . , yα.<br />

Only in cases <strong>of</strong> multiple roots would the equations not suffice. In such cases ABEL<br />

involved the calculus which could be used to produce a set <strong>of</strong> α independent equations<br />

for determining a1, . . . , aα. When ABEL divided F (x) by ∏ α τ=1 (x − xτ), he obtained<br />

another equation<br />

F1 (x) =<br />

∏ α τ=1<br />

F (x)<br />

= 0<br />

(x − xτ)


19.2. <strong>The</strong> contents <strong>of</strong> ABEL’s Paris result and its pro<strong>of</strong> 369<br />

which was <strong>of</strong> degree µ − α, has as its roots xα+1, . . . , xµ and whose coefficients were<br />

rational functions <strong>of</strong> x1, . . . , xα and y1, . . . , yα. Thus, ABEL concluded from Main <strong>The</strong>o-<br />

rem I that any sum such as ∑ α k=1 ψ k (x k) could be expressed by a known function v and<br />

a similar sum <strong>of</strong> functions,<br />

α<br />

∑ ψk (xk) = v −<br />

k=1<br />

µ<br />

∑<br />

k=α+1<br />

ψk (xk) . (19.29)<br />

<strong>The</strong> relation γ = µ − α. <strong>The</strong> expression (19.29) at first might seem like a mere rep-<br />

etition <strong>of</strong> the Main <strong>The</strong>orem I, but as ABEL stressed, the number <strong>of</strong> terms on the right<br />

hand side (µ − α) shows remarkable features. <strong>The</strong> stress put on the number γ is cer-<br />

tainly one <strong>of</strong> the important aspects <strong>of</strong> ABEL’S paper, and it has received widespread<br />

mathematical interest not least after G. F. B. RIEMANN (1826–1866) transformed it<br />

into a coherent concept <strong>of</strong> genus, see section 19.3. For now, we focus our attention<br />

completely on ABEL’S argument and the inner logic <strong>of</strong> the paper.<br />

As he had done above, ABEL again divided his argument by whether r has a factor<br />

independent <strong>of</strong> the indeterminates or not. He started with the latter case, F0 (x) = 1<br />

for which he found that all the coefficient functions q0, . . . , qn−1 were arbitrary and<br />

their hq k + 1 coefficients had to correspond to the indeterminates,<br />

α =<br />

n−1<br />

n<br />

∑ (hqk + 1) = ∑ hqk + n − 1.<br />

k=1<br />

k=1<br />

Obtaining a corresponding formula for the other case, in which F0 (x) �= 1, proved<br />

much more tedious. In general, ABEL claimed, the equation<br />

r (x) = F0 (x) F (x) (19.30)<br />

would impose hF0 conditions, but particular forms for y can eliminate some <strong>of</strong> these<br />

conditions. 35 If the number <strong>of</strong> conditions imposed by (19.30) is hF0 − A, the number<br />

<strong>of</strong> indeterminates could be counted as<br />

α =<br />

n−1<br />

∑ (hqk + 1) − (hF0 − A) . (19.31)<br />

k=1<br />

On the other hand, ABEL easily obtained from the definitions<br />

and<br />

hr = h<br />

hr = hF0 + hF = hF0 + µ<br />

� �<br />

n<br />

n<br />

∏ θ (yk) = ∑ hθ (yk) ,<br />

k=1<br />

k=1<br />

35 Here, a remarkably clear juxtaposition <strong>of</strong> “in general” and “particular cases” was given.


370 Chapter 19. <strong>The</strong> Paris memoir<br />

which allowed him to rewrite (19.31) as<br />

α =<br />

=<br />

n−1<br />

∑ hqk + n − 1 − (hr − µ) + A<br />

k=1<br />

n−1<br />

∑ hqk −<br />

k=1<br />

n<br />

∑ hθ (yk) + n − 1 + A + µ. (19.32)<br />

k=1<br />

<strong>The</strong>refore, ABEL turned to algebraically manipulating the “degrees” hθ (y k) along the<br />

same lines as had been followed in describing γ above.<br />

Obviously, from the inequality<br />

and the formula<br />

ABEL obtained<br />

hθ (y) ≥ h (qmy m ) for each m = 0, . . . , n − 1,<br />

h (qmy m ) = hqm + mhy,<br />

hθ (y k) ≥ hqm + mhy k for k = 1, . . . , n.<br />

Designating by ρτ the index <strong>of</strong> the maximal value <strong>of</strong> h (qmy m ) within the τ’th sequence<br />

<strong>of</strong> roots, ABEL employed the same machinery which had served him before, although<br />

this time in a slightly different notational dressing. Summing the excesses ε τ,k within<br />

the τ’th sequence,<br />

nτµτ−1<br />

∑ ετ,k = Cτ,<br />

k=1<br />

ABEL found by the manipulations and number theoretic results<br />

or less specifically (Cτ ≥ 0)<br />

µ − α ≥ γ − A +<br />

µ − α ≥ γ − A.<br />

ε<br />

∑ Cτ,<br />

τ=1<br />

However, the inequality in (19.33) was actually an equality,<br />

µ − α = γ − A, (19.33)<br />

as ABEL deduced by another tedious sequence <strong>of</strong> manipulations.<br />

Specializing the relation expressed (19.33) to the other case (in which F0 (x) = 1)<br />

led to the result that if r did not contain any factors independent <strong>of</strong> the indeterminate<br />

quantities, then<br />

µ − α = γ.<br />

ABEL concluded these investigations by considering situations in which the coef-<br />

ficients q0, . . . , qn−1 were subjected to some kinds <strong>of</strong> conditions. Again arguing very<br />

“generally”, ABEL could claim that the result (19.33) would not generally be substan-<br />

tially altered, although the constant A should reflect the additional conditions.<br />

In the paper’s eighth section, ABEL applied the results obtained thus far to calcu-<br />

late γ in an example for which n = 13.


19.3. Additional, tentative remarks on ABEL’s tools 371<br />

µ − α independent <strong>of</strong> α. In the previous argument, ABEL had reached an expression<br />

such as<br />

hqm = hqρ 1 + Mm for m = 0, . . . , n − 1<br />

in which Mm is independent <strong>of</strong> hqρ 1 . Thus, counting the number <strong>of</strong> coefficients in<br />

q0, . . . , qn−1 yields as an upper limit for α<br />

ABEL could therefore write<br />

α ≤ nhqρ 1 +<br />

n−1<br />

∑ Mm.<br />

m=0<br />

α = nhqρ 1 + M,<br />

in which M was independent <strong>of</strong> hqρ 1 . <strong>The</strong>refore, taken together with (19.32), µ − α<br />

was found to be independent <strong>of</strong> the value <strong>of</strong> α.<br />

II:<br />

ABEL summarized these results in an announcement <strong>of</strong> the central Main <strong>The</strong>orem<br />

“<strong>The</strong> equation (74) [here (19.29)] enables us to express a sum <strong>of</strong> any number <strong>of</strong><br />

given function <strong>of</strong> the form ψx by a sum <strong>of</strong> a particular number <strong>of</strong> functions. <strong>The</strong><br />

last number can always be supposed equal to µ which — in general — will be its<br />

smallest value.” 36<br />

In summary, the Main <strong>The</strong>orem II can be expressed as follows:<br />

<strong>The</strong>orem 17 (Main <strong>The</strong>orem II) Under the present assumptions,<br />

τ � xk<br />

∑<br />

k=1<br />

f (x k, y k) dx k =<br />

γ � zk<br />

∑<br />

k=1<br />

f (x k, y k) dx k + v<br />

in which γ is independent <strong>of</strong> τ and x1, . . . , xτ, z1, . . . , zγ are given algebraically in x1, . . . , xτ<br />

and v is an algebraic and logarithmic function. ✷<br />

As his final act before turning toward the applications <strong>of</strong> this result (see discussion<br />

in section 19.5.2, below), ABEL generalized the theorem to apply to linear combina-<br />

tions <strong>of</strong> integrals with rational coefficients.<br />

19.3 Additional, tentative remarks on ABEL’s tools<br />

As mentioned in section 19.1.2 and described at length above, ABEL’S Paris memoir<br />

is full <strong>of</strong> interesting applications <strong>of</strong> algebraic tools. In the present section, some brief<br />

perspectives on five <strong>of</strong> the most important tools are <strong>of</strong>fered.<br />

36 “L’équation (74) nous met donc en état d’exprimer une somme d’un nombre quelconque de fonctions<br />

données, de la forme ψx, par une somme d’un nombre déterminé de fonctions. Le dernier<br />

nombre peut toujours être supposé égal à γ , qui, en général, sera sa plus petite valeur.” (N. H.<br />

<strong>Abel</strong>, [1826] 1841, 185).


372 Chapter 19. <strong>The</strong> Paris memoir<br />

<strong>The</strong> theory <strong>of</strong> residues and the expansion in decreasing power series. One <strong>of</strong> ABEL’S<br />

in the expansion <strong>of</strong> the function<br />

tools was to focus attention on the coefficient <strong>of</strong> 1 x<br />

f (x) according to decreasing powers <strong>of</strong> x (see page 355). Judged by ABEL’S way <strong>of</strong><br />

introducing the operation which he denoted Π f x, the expansion <strong>of</strong> the function f<br />

into series <strong>of</strong> decreasing powers was not considered problematic; my best guess is<br />

that it was considered a formal operation or a well established fact. However, simul-<br />

taneously, CAUCHY was attributing new importance to the same object although in a<br />

completely different theoretical environment when he laid the foundations for his new<br />

calculus <strong>of</strong> residues in a series <strong>of</strong> papers in the Exercises de mathématiques. 37 From the<br />

quote on page 306, we know that ABEL bought and studied these installments. Thus,<br />

this mathematical object acquired importance in two distinct theories possibly from<br />

two very distinct approaches. However, when we include another <strong>of</strong> ABEL’S tools, we<br />

may get the impression that ABEL’S Π might stand closer to CAUCHY’S residues.<br />

Lagrange interpolation and ABEL’S Annales-paper. ABEL’S made repeated use <strong>of</strong> a<br />

result which derives from the process <strong>of</strong> Lagrange interpolation. In its original form,<br />

Lagrange interpolation concerned the problem <strong>of</strong> fitting a polynomial <strong>of</strong> degree n − 1<br />

through n specified points in the plane {(xk, f (xk))} n<br />

k=1 . J. L. LAGRANGE (1736–1813)<br />

had attacked this problem in 1795 and presented the polynomial<br />

Pn (x) =<br />

n<br />

∑ f (xk) k=1<br />

as the solution; 38 the function ω (x) was defined by<br />

ω (x) =<br />

and its derivative consequently satisfied<br />

ω (x)<br />

(x − x k) ω ′ (x k)<br />

n<br />

∏ (x − xk) k=1<br />

ω ′ (xk) = ∏ (xk − xm) .<br />

m�=k<br />

This result can easily be proved and it immediately leads to the applications which<br />

ABEL made <strong>of</strong> it in the Paris memoir. Interestingly, CAUCHY pursued the same re-<br />

sult with his new theory <strong>of</strong> residues and explicitly referred to the process as Lagrange<br />

interpolation. 39 In a short paper which ABEL published in GERGONNE’S Annales de<br />

mathématiques, 40 he deduced the same result by elimination in two polynomial equa-<br />

tions. <strong>The</strong>se remarks are meant to indicate that Lagrange interpolation was a powerful<br />

tool which was being investigated from numerous perspectives in the early nineteenth<br />

37 See e.g. (A.-L. Cauchy, 1826).<br />

38 (Kolmogorov and Yushkevich, 1992–1998, III, 262).<br />

39 (A.-L. Cauchy, 1826, 35–37).<br />

40 (N. H. <strong>Abel</strong>, 1827a).


19.3. Additional, tentative remarks on ABEL’s tools 373<br />

century. I believe that the problem <strong>of</strong> elimination can be considered sufficient inspira-<br />

tion for ABEL to treat Lagrange interpolation but the very central role which it together<br />

with the operator Π played in the Paris memoir (see above) may also suggest that ABEL<br />

had actually picked up his tools from CAUCHY’S new theory <strong>of</strong> residues although he<br />

did not adapt the full theory.<br />

<strong>The</strong> degree-operator. ABEL’S arguments about the number <strong>of</strong> independent integrals<br />

γ relied extensively on ways <strong>of</strong> determining the number <strong>of</strong> independent (or free) coef-<br />

ficients. In turn, their determination was based on an extended degree operator which<br />

could apply to the implicitly defined algebraic functions with which he dealt. ABEL<br />

introduced the fractional degree operator hR as the highest exponent in a develop-<br />

ment <strong>of</strong> R according to decreasing powers. As was the case with the residue-operator,<br />

no explicit considerations as to the validity <strong>of</strong> this definition are presented and it ap-<br />

pears to be a well known, formal trick. ABEL presented the basic rules for the de-<br />

gree operator and when he wanted to apply it to differences between to expressions<br />

among y1, . . . , yn, he insisted that these expressions be ordered according to their de-<br />

gree. However, as observed on page 362, particular cases could still arise in which the<br />

identity<br />

h (ym − y k) = max {hym, hy k} (19.34)<br />

did not hold. Such cases were apparently peculiar situations <strong>of</strong> little interest, and<br />

ABEL dismissed them by claiming that the equation (19.34) would hold “in general”.<br />

In his comments, 41 SYLOW made a considerable effort in clarifying ABEL’S arguments<br />

and in particular in revising his deduction <strong>of</strong> the properties <strong>of</strong> γ by making explicit<br />

some <strong>of</strong> the assumptions which ABEL had not made when he simply argued “in gen-<br />

eral”. <strong>The</strong> notion <strong>of</strong> equations being “generally valid” will be addressed further in<br />

chapter 21 where it was be interpreted and explained based on the notion <strong>of</strong> formula<br />

based mathematics.<br />

<strong>The</strong> genus. <strong>The</strong> number which ABEL denoted γ expressed the number <strong>of</strong> indepen-<br />

dent integrals related to a particular algebraic differential. As we have seen, ABEL’S<br />

deduction <strong>of</strong> the invariance <strong>of</strong> the number γ was cumbersome and hampered by cer-<br />

tain points where it was not completely clear and rigorous. Furthermore, although his<br />

arguments were highly explicit they did not immediately produce a way <strong>of</strong> generally<br />

computing the number γ. In the subsequent decades, it became an extremely promi-<br />

nent mathematical problem to rigorously establish the basis for ABEL’S theorems and<br />

to investigate the number γ further. Eventually, RIEMANN presented an approach<br />

based on multi-sheeted surfaces and introduced the name genus and the symbol p for<br />

ABEL’S γ. 42 <strong>The</strong> further description <strong>of</strong> RIEMANN’S theory is, unfortunately, way be-<br />

41 (N. H. <strong>Abel</strong>, 1881, II).<br />

42 (B. Riemann, 1857).


374 Chapter 19. <strong>The</strong> Paris memoir<br />

Figure 19.1: GEORG FRIEDRICH BERNHARD RIEMANN (1826–1866)<br />

yond the present scope. 43 However, in the present context, it serves to demonstrate<br />

that ABEL’S ideas were pursued and rigorized over the ensuing decades.<br />

Birth <strong>of</strong> a new concept: algebraic functions. <strong>The</strong> final aspect which will be tenta-<br />

tively discussed here concerns the introduction <strong>of</strong> a new concept <strong>of</strong> implicitly defined<br />

algebraic functions. In his research on the solubility <strong>of</strong> equations, ABEL had — for obvi-<br />

ous reasons — only been interested in investigating explicit algebraic functions which<br />

could serve as solutions for equations. However, in the Paris memoir, implicit algebraic<br />

functions were the primary objects <strong>of</strong> study and ABEL developed a number <strong>of</strong> tools<br />

for their investigation. Later, implicitly given algebraic functions (or just algebraic<br />

functions) became very important objects <strong>of</strong> mathematics and — as BRILL remarks —<br />

ABEL may be seen as the initiator <strong>of</strong> the theory <strong>of</strong> algebraic functions; in particular, he<br />

proved a very important theorem in the theory. 44 Later, J. LIOUVILLE (1809–1882) also<br />

adopted — following ABEL but departing from his French precursors — the implicitly<br />

given algebraic functions into his theory <strong>of</strong> integration in finite terms. 45 In summary,<br />

we are faced with an introduction <strong>of</strong> a new set <strong>of</strong> objects, and ABEL’S highly formula<br />

based investigations <strong>of</strong> these objects and developments <strong>of</strong> tools for their study can be<br />

interpreted as a way <strong>of</strong> getting to know and express results about this new branch.<br />

43 One rather recent, very brief sketch is given in (Houzel, 1986, 310–313).<br />

44 (Brill and Noether, 1894, 212).<br />

45 (Lützen, 1990, 370).


19.4. <strong>The</strong> fate <strong>of</strong> the Paris memoir 375<br />

19.4 <strong>The</strong> fate <strong>of</strong> the Paris memoir<br />

ABEL’S Paris memoir had a strange and adventurous fate which had certain influences<br />

on ABEL’S career. 46 After ABEL had delivered his memoir to the Académie des Sciences,<br />

the Academy commissioned A.-M. LEGENDRE (1752–1833) and CAUCHY to present a<br />

report on it. <strong>The</strong> manuscript soon landed on CAUCHY’S desk, where it withered until<br />

1829. After having learned <strong>of</strong> ABEL’S untimely death from a communication on 22<br />

June 1829 by LEGENDRE, 47 CAUCHY finally took the time to write the report which<br />

was dated 29 June 1829. 48 CAUCHY explicitly noticed in his report how ABEL treated<br />

implicitly defined algebraic functions and that this was a necessary requirement for<br />

his theorems to be true. 49 This is further indication that the introduction <strong>of</strong> implicitly<br />

given algebraic functions was not completely obvious and standard.<br />

Manuscript lost and found. Based on CAUCHY’S positive report, the Académie des<br />

Sciences decided to include the ABEL’S Paris memoir in the Mémoires présentés par divers<br />

savants once a new copy had been prepared. Furthermore, ABEL was — posthumously<br />

and jointly with C. G. J. JACOBI (1804–1851) — awarded the Grand prix <strong>of</strong> the Académie<br />

des Sciences in 1830. However, the manuscript was misplaced — possibly as a result<br />

<strong>of</strong> the turbulent events in Paris in 1830 — as a Norwegian enquiry would realize in<br />

1832 when B. M. HOLMBOE (1795–1850) was beginning to prepare the first edition <strong>of</strong><br />

ABEL’S Œuvres. Consequently, the Œuvres appeared in 1839 without the Paris memoir.<br />

Apparently, this provoked some reaction from the [Académie des Sciences]Académie<br />

which commissioned LIBRI with the job <strong>of</strong> seeing it through print and, as noted, it was<br />

published in 1841.<br />

While preparing the second edition <strong>of</strong> ABEL’S collected works, M. S. LIE (1842–<br />

1899) in 1874 obtained permission to consult the original manuscript supposedly held<br />

in the archives <strong>of</strong> the Académie des Sciences. 50 However, the manuscript was again<br />

nowhere to be found in the archives and again had to be considered lost. Conse-<br />

quently, the Paris memoir was included in the second edition <strong>of</strong> ABEL’S collected works<br />

but with the version printed by the French Academy in 1841 as the source. In the twen-<br />

tieth century, a copy <strong>of</strong> the manuscript was first located in Rome by P. HEEGAARD<br />

(1871–1948) in 1942 before V. BRUN (1885–1978) 51 succeeded in finding the majority<br />

<strong>of</strong> ABEL’S original manuscript in Florence ten years later. Recently, in 2000, the fi-<br />

nal eight missing pages have been found and the whereabouts <strong>of</strong> the entire original<br />

manuscript <strong>of</strong> ABEL’S Paris memoir are known for the first time in 150 years.<br />

46 <strong>The</strong> fate <strong>of</strong> the Paris memoir is described in most biographies <strong>of</strong> ABEL. Additionally, information<br />

for the present sketch is drawn from papers including (Brun, 1949; Brun, 1953; Lange-<strong>Niels</strong>en, 1927;<br />

Lange-<strong>Niels</strong>en, 1929).<br />

47 (Lange-<strong>Niels</strong>en, 1929, 14).<br />

48 (Lange-<strong>Niels</strong>en, 1927, 67).<br />

49 CAUCHY’S report is reproduced in (ibid., 69).<br />

50 (N. H. <strong>Abel</strong>, 1881, II, 294)<br />

51 Information from (Scriba, 1980).


376 Chapter 19. <strong>The</strong> Paris memoir<br />

19.5 Reception <strong>of</strong> the Paris memoir<br />

Some aspects <strong>of</strong> the reception <strong>of</strong> ABEL’S Paris memoir have already been touched upon<br />

above when it was noted how ABEL introduced a new class <strong>of</strong> objects into mathemat-<br />

ical study. In the present section, some attention is paid to the primary application <strong>of</strong><br />

the Paris memoir which dealt with the so-called hyperelliptic integrals .<br />

19.5.1 ABEL’s announcements <strong>of</strong> the Paris memoir<br />

In response to the lacking reaction from the Parisian Académie des Sciences, ABEL went<br />

ahead and made public some <strong>of</strong> his results originally contained in the Paris memoir.<br />

He did so in two papers published in A. L. CRELLE’S (1780–1855) Journal in 1828 and<br />

1829. 52 <strong>The</strong> first <strong>of</strong> these contained the application <strong>of</strong> the Paris memoir to hyperelliptic<br />

integrals and will be treated in some detail below. <strong>The</strong> second one entitled “Démon-<br />

stration d’une propriété générale d’une certaine classe de fonctions transcendantes” had been<br />

signed by ABEL January 6, 1829 and was published in the second issue <strong>of</strong> the fourth<br />

volume <strong>of</strong> the Journal which appeared just days before ABEL’S death. In that paper,<br />

written in haste by an already ill-taken mathematician, ABEL presented the first main<br />

theorem <strong>of</strong> the Paris memoir and gave a short and direct pro<strong>of</strong> based on the theorem <strong>of</strong><br />

LAGRANGE (theorem 1) which had already served him well in the theory <strong>of</strong> equations.<br />

At the end <strong>of</strong> the two page pro<strong>of</strong>, ABEL promised to present multiple applications <strong>of</strong><br />

Main <strong>The</strong>orem I on a later occasion.<br />

19.5.2 Application to hyperelliptic integrals<br />

<strong>The</strong> application to hyperelliptic integrals has been seen by some historians <strong>of</strong> mathe-<br />

matics as the true motivation for ABEL’S deduction <strong>of</strong> the <strong>Abel</strong>ian <strong>The</strong>orem. 53 Certainly,<br />

the so-called hyperelliptic integrals figured prominently among ABEL’S examples in<br />

the Paris memoir when he applied the results to integrals <strong>of</strong> the form<br />

� � � �<br />

f x, φn (x) dx<br />

in which f was a rational function and Pn was a polynomial <strong>of</strong> degree n. <strong>The</strong>se in-<br />

tegrals are separated from the greater class considered in the Paris memoir by corre-<br />

sponding to an equation <strong>of</strong> the form χ (x, y) = y 2 − φn (x).<br />

For ordinary elliptic integrals (n = 3 or n = 4), ABEL’S calculations in the Paris<br />

memoir showed that γ = 1 and thus, any sum <strong>of</strong> similar elliptic integrals could be<br />

reduced to a single elliptic integral <strong>of</strong> the same form and possibly some � algebraic � and<br />

n−1<br />

logarithmic terms. For hyperelliptic integrals, ABEL found that γ = 2 where ⌊x⌋<br />

denotes the integer value <strong>of</strong> x. Thus, for instance, if n = 6, any number <strong>of</strong> similar<br />

52 (N. H. <strong>Abel</strong>, 1828c; N. H. <strong>Abel</strong>, 1829b).<br />

53 See (Brill and Noether, 1894, 210–211) and the discussion in (Cooke, 1989).


19.6. Conclusion 377<br />

hyperelliptic integrals could be reduced to two such integrals and simpler terms. This<br />

was also one <strong>of</strong> the main results <strong>of</strong> the paper Remarques sur quelques propriétés générales<br />

d’une certaine sorte de fonctions transcendantes which ABEL published in CRELLE’S Journal<br />

in 1828. 54<br />

ABEL’S result was picked up by JACOBI who praised it highly and suggested that<br />

it could be used as a form <strong>of</strong> generalized addition theorem for hyperelliptic functions .<br />

Thus, for hyperelliptic integrals, say, with X <strong>of</strong> degree 5,<br />

� x<br />

0<br />

dx<br />

√ X = Φ1 (x) and<br />

� x<br />

0<br />

x dx<br />

√X = Φ1 (x) ,<br />

the problem became to express the upper limits <strong>of</strong> the involved integrals as<br />

such that<br />

x = λ (u, v) and y = λ1 (u, v)<br />

Φ (x) + Φ (y) = u and Φ1 (x) + Φ1 (y) = v. 55<br />

This idea — which is a generalization <strong>of</strong> the addition theorems for trigonometric and<br />

elliptic functions — is called the (Jacobian) inversion problem and it attracted a great<br />

deal <strong>of</strong> attention in the 1830s and 1840s. Eventually, around 1850, it was solved inde-<br />

pendently by A. GÖPEL (1812–1847) and J. G. ROSENHAIN (1816–1887) who applied<br />

a generalization <strong>of</strong> JACOBI’S theory <strong>of</strong> theta functions for elliptic functions (see next<br />

chapter) to this case. 56 In the process, they were forced to accept that the functions<br />

were depended on two complex variables and had four periods. Thus, the search<br />

to generalize the theory <strong>of</strong> basic pattern <strong>of</strong> the theory <strong>of</strong> elliptic integrals even further<br />

forced mathematicians to face even more peculiar functions. <strong>The</strong> continuing extension<br />

<strong>of</strong> the theory <strong>of</strong> higher transcendentals posed new demands to the rigor and tools <strong>of</strong><br />

the mathematician and thereby influenced the general development <strong>of</strong> mathematics<br />

in important ways.<br />

19.6 Conclusion<br />

In the present chapter, ABEL’S Paris memoir has been described in some details to facil-<br />

itate a discussion <strong>of</strong> the tools which ABEL employed. Combining the characterization<br />

<strong>of</strong> the tools employed here with those described in the previous chapter, we are led to<br />

see that algebraic methods were extremely important in ABEL’S research on transcen-<br />

dentals. <strong>The</strong>se methods range from results which he had originally used to study the<br />

solubility <strong>of</strong> equations to newly developed tools which, nevertheless, also <strong>of</strong>ten were<br />

based in polynomials, equations, considerations <strong>of</strong> rational dependence, and the like.<br />

54 (N. H. <strong>Abel</strong>, 1828c).<br />

55 (C. G. J. Jacobi, 1832b, 400); see also (Houzel, 1986, 311).<br />

56 (Göpel, 1847; Rosenhain, 1851).


378 Chapter 19. <strong>The</strong> Paris memoir<br />

As described, ABEL’S deductions in the Paris memoir were extremely cumbersome<br />

and not always universally permitted. Generally, ABEL’S arguments in the memoir fell<br />

well inside the formula based paradigm: they were concerned with formulae, were car-<br />

ried by manipulations <strong>of</strong> formulae, and arguments which were not universally true<br />

but only true “in general” were acceptable. In chapter 21, these features will be shown<br />

to be intimately connected with an “old” paradigm which was gradually being re-<br />

placed in the period. Thus, and because it was not available for 15 years, it is not<br />

surprising to find that ABEL’S pro<strong>of</strong> in the Paris memoir was not so widely accepted<br />

and imitated as the statements or results to which they led.


Chapter 20<br />

General approaches to elliptic<br />

functions<br />

<strong>The</strong> present chapter serves to present the ideas which matured the theory <strong>of</strong> elliptic<br />

functions in the nineteenth century. In particular, attention is called to the variations<br />

in the ways <strong>of</strong> introducing elliptic functions because these variations are indicators <strong>of</strong><br />

the changing conceptions <strong>of</strong> rigorous foundations. Moreover, it is interesting to follow,<br />

how results are turned into definitions to meet the changing standards <strong>of</strong> rigorous<br />

definitions.<br />

20.1 ABEL’s version <strong>of</strong> a general theory <strong>of</strong> elliptic<br />

functions<br />

In his ultimate publication, the Précis d’une théorie des fonctions elliptiques, 1 N. H. ABEL<br />

(1802–1829) addressed the theory <strong>of</strong> elliptic functions from a more general perspective<br />

than the Recherches. Since the Recherches which dealt exclusively with elliptic func-<br />

tions <strong>of</strong> the first kind, ABEL had published a number <strong>of</strong> smaller papers on the theory<br />

<strong>of</strong> elliptic functions, some investigations on integration in finite form, and various<br />

announcements <strong>of</strong> his Paris memoir. ABEL had also been working on a continuation<br />

<strong>of</strong> the Recherches which he postponed to devote his energy to completing the Précis.<br />

<strong>The</strong> second Recherches mémoire was eventually published by M. G. MITTAG-LEFFLER<br />

(1846–1927) in 1902. 2 However, to focus on the most general <strong>of</strong> ABEL’S approaches, we<br />

shall limit the discussion to the Précis. In the Précis, these threads were woven together<br />

to present an exposition <strong>of</strong> the entire theory <strong>of</strong> elliptic functions which simultaneously<br />

addressed all three kinds.<br />

1 (N. H. <strong>Abel</strong>, 1829d).<br />

2 (N. H. <strong>Abel</strong>, 1902c).<br />

379


380 Chapter 20. General approaches to elliptic functions<br />

Reiterating established knowledge. ABEL began his Précis — which he privately<br />

called the “knockout <strong>of</strong> C. G. J. JACOBI’ (1804–1851)’ — by iterating some <strong>of</strong> his pre-<br />

vious results pertaining to elliptic functions <strong>of</strong> the first kind. With the radical<br />

�<br />

∆ (x, c) = (1 − x2 ) (1 − c2x2 ),<br />

he introduced his three kinds <strong>of</strong> elliptic integrals as<br />

�<br />

dx<br />

˜ω (x, c) =<br />

∆ (x) ,<br />

� x 2 dx<br />

˜ω0 (x, c) = , and<br />

∆ (x)<br />

�<br />

dx<br />

Π (x, c, a) = � � .<br />

∆ (x, c)<br />

1 − x2<br />

a 2<br />

One <strong>of</strong> the major tricks <strong>of</strong> the Précis was that once the elliptic integral <strong>of</strong> the first kind<br />

had been inverted<br />

θ =<br />

� λ(θ)<br />

0<br />

dx<br />

∆ (x, c) ,<br />

this elliptic function could be used to describe the elliptic integrals <strong>of</strong> the second and<br />

third kind,<br />

�<br />

˜ω0 (x, c) =<br />

λ 2 �<br />

(θ) dθ and Π (x, c, a) =<br />

dθ<br />

1 − λ2 θ<br />

a 2<br />

Thus, ABEL had essentially reduced the problem <strong>of</strong> inverting all three kinds <strong>of</strong> inte-<br />

grals and he could combine the study <strong>of</strong> all elliptic functions in knowledge about the<br />

elliptic functions <strong>of</strong> the first kind.<br />

Obviously, among the key results which ABEL iterated for the elliptic function λ<br />

was its two periods and the complete solution <strong>of</strong> the equation λ (θ ′ ) = λ (θ) which<br />

we have already encountered multiple times. Moreover, ABEL also presented various<br />

infinite representations <strong>of</strong> λ and investigated the conditions <strong>of</strong> transformations. Thus,<br />

all the key components <strong>of</strong> his previous approaches were included in the Précis, albeit<br />

in a more coherent and lucid form.<br />

General properties <strong>of</strong> elliptic functions. <strong>The</strong> major new purpose <strong>of</strong> ABEL’S Précis<br />

was to investigate a new program <strong>of</strong> representation for elliptic functions. In the pro-<br />

cess, ABEL made important use <strong>of</strong> the insights which he had developed and presented<br />

in relation with his Paris memoir.<br />

With two polynomial functions f (even) and φ (odd), ABEL defined<br />

ψ (x) = f (x) 2 − φ (x) 2 ∆ (x) 2<br />

which was an even function and therefore could be split in factors as<br />

ψ (x) = A<br />

µ<br />

∏<br />

n=1<br />

�<br />

x 2 − x 2 �<br />

n .<br />

.<br />

(20.1)


20.2. Other ways <strong>of</strong> introducing elliptic functions in the nineteenth century 381<br />

In this situation, ABEL found by employing Lagrange interpolation that the sum <strong>of</strong> inte-<br />

grals <strong>of</strong> the third kind reduced to a logarithmic expression<br />

µ<br />

∑ Π (xn, a) = C −<br />

n=1<br />

a<br />

2∆ (a)<br />

f (a) + φ (a) ∆ (a)<br />

log . (20.2)<br />

f (a) − φ (a) ∆ (a)<br />

ABEL extended this property <strong>of</strong> elliptic integrals <strong>of</strong> the third kind to analogous re-<br />

sults for elliptic integrals <strong>of</strong> the first and second kind. For integrals <strong>of</strong> the second kind,<br />

ABEL observed that ˜ω (x) = lima→∞ Π (x, a) whereas the logarithmic term vanished<br />

under this limit process,<br />

µ<br />

∑<br />

n=1<br />

˜ω (xn) = C.<br />

For integrals <strong>of</strong> the second kind, ABEL considered the expansion <strong>of</strong> (20.2) according to<br />

increasing powers <strong>of</strong> 1 a<br />

and compared coefficients <strong>of</strong> 1<br />

a 2 to conclude<br />

µ<br />

∑<br />

n=1<br />

˜ω0 (xn) = C − p<br />

where p was an algebraic function <strong>of</strong> x1, . . . , xµ.<br />

Thus, ABEL used tools similar to those which he employed in the Paris memoir to<br />

deduce results which also bear similarities with the Main <strong>The</strong>orem I (theorem 16).<br />

A new program <strong>of</strong> representability. In the second chapter, ABEL suggested a very<br />

general question: He wanted to describe all integrals <strong>of</strong> algebraic differentials which<br />

could be expressed by algebraic, logarithmic, and elliptic functions. Thus, if com-<br />

pared with the Paris memoir, the elliptic functions were now accepted among the basic<br />

functions to which other higher transcendentals could be reduced. However, when he<br />

came to answer the question, he restricted himself to attack transformation problems<br />

and other relations among elliptic integrals. <strong>The</strong> more specific contents <strong>of</strong> his inves-<br />

tigations are considered outside the present scope, although its presentation would<br />

probably reiterate the description <strong>of</strong> ABEL’S methods and tools which have been sug-<br />

gested in the previous chapters.<br />

ABEL did not manage to complete his papers before he died. Nevertheless, his<br />

approach illustrated the fruitful influence which the results <strong>of</strong> Paris memoir could have<br />

on the theory <strong>of</strong> elliptic functions. However, because <strong>of</strong> his early death, it was left<br />

to ABEL’S contemporaries and competitors to outline the future development <strong>of</strong> the<br />

theory <strong>of</strong> elliptic functions.<br />

20.2 Other ways <strong>of</strong> introducing elliptic functions in the<br />

nineteenth century<br />

During the nineteenth century, the definitions <strong>of</strong> elliptic functions were turned upside<br />

down a number <strong>of</strong> times. In chapter 16, it has been described, how ABEL introduced


382 Chapter 20. General approaches to elliptic functions<br />

elliptic functions (<strong>of</strong> the first kind) by a formal inversion <strong>of</strong> the corresponding elliptic<br />

integral and an extension to the complex domain. However, his contemporaries and<br />

successors soon found other ways <strong>of</strong> introducing elliptic functions more preferably. In<br />

the present context, it suffices to consider three different approaches.<br />

JACOBI. With ABEL out <strong>of</strong> the competition, JACOBI’S influence on the theory <strong>of</strong> el-<br />

liptic functions toward the end <strong>of</strong> the 1820s was overwhelming. JACOBI’S notation<br />

and means <strong>of</strong> introducing the functions became standardized through a number <strong>of</strong><br />

publications starting with his Fundamenta nova which was the first monograph de-<br />

voted to the study <strong>of</strong> the elliptic functions. 3 In the Fundamenta nova, JACOBI de-<br />

fined elliptic functions as inverses <strong>of</strong> elliptic integrals but beginning with a course<br />

taught in 1838, the changed the foundation to the so-called theta-functions. <strong>The</strong>se<br />

were a particular set <strong>of</strong> four exponential series the simplest <strong>of</strong> which can be written<br />

as ϑ (x) = ∑n∈Z (−1) n qn2e2nix . 4 Based on these series, JACOBI could introduce and<br />

investigate all parts <strong>of</strong> the theory <strong>of</strong> elliptic functions. Thus, JACOBI introduced series<br />

as the basic objects upon which everything else should be built. Other definitions by<br />

series, e.g. as ratios <strong>of</strong> power series were also being suggested and adopted.<br />

J. LIOUVILLE (1809–1882). A completely different approach to the introduction <strong>of</strong><br />

elliptic functions was taken by LIOUVILLE who chose to develop an entire theory for<br />

doubly periodic functions. Of such functions, he was able to deduce a number <strong>of</strong> re-<br />

sults and eventually prove that they could be used to represent the inverses <strong>of</strong> elliptic<br />

integrals. 5 Thus, LIOUVILLE circumvented the approach <strong>of</strong> ABEL and JACOBI and in-<br />

vestigated the concept <strong>of</strong> functions defined by what ABEL had deduced as a property.<br />

<strong>The</strong> strength <strong>of</strong> the approach was that eventually, the classes <strong>of</strong> elliptic functions and<br />

doubly periodic, meromorphic functions were found to coincide.<br />

K. T. W. WEIERSTRASS (1815–1897). <strong>The</strong> final approach which I wish to mention<br />

was through differential equations. For instance, WEIERSTRASS introduced his func-<br />

tion ℘ (u) as the solution to the differential equation<br />

� �2 dy<br />

= 4y<br />

du<br />

3 − g2y − g3 for g2, g3 constants<br />

which had a pole at u = 0. 6 Subsequently, WEIERSTRASS found means <strong>of</strong> obtaining<br />

more direct representations <strong>of</strong> his elliptic functions, for instance in the form<br />

σ (u + v) σ (u − v)<br />

℘ (v) − ℘ (u) =<br />

σ2u · σ2v where the function σ could be expanded in an infinite product. 7<br />

3 (C. G. J. Jacobi, 1829).<br />

4 (Houzel, 1986, 304–305).<br />

5 (Lützen, 1990, 555).<br />

6 (Houzel, 1986, 298).<br />

7 (ibid., 306).


20.3. Conclusion 383<br />

All these four definitions ultimately define the same objects and it is a major sign<br />

<strong>of</strong> strength <strong>of</strong> the theory over the century that so many different approaches had been<br />

described. Mathematicians were free to chose which <strong>of</strong> the definitions satisfied their<br />

requirements for usability and rigor; subsequently, the other representations could be<br />

deduced.<br />

20.3 Conclusion<br />

In the present part, ABEL’S approach to the theory <strong>of</strong> elliptic functions has been de-<br />

scribed from a number <strong>of</strong> perspectives. Based on a presentation <strong>of</strong> aspects <strong>of</strong> the the-<br />

ory <strong>of</strong> elliptic integrals in the eighteenth century, it has been illustrated, how ABEL<br />

introduced elliptic functions as formal inverses <strong>of</strong> elliptic integrals and extended the<br />

resulting function to the complex domain. <strong>The</strong> main inspiration which ABEL drew<br />

from C. F. GAUSS’ (1777–1855) suggestion that the division problem from the circle<br />

could also be attacked for the lemniscate has also been described. With the formal in-<br />

version as his definition, ABEL sought for ways <strong>of</strong> representing his elliptic functions,<br />

and some aspects <strong>of</strong> this program such as the standards <strong>of</strong> rigor and the use and re-<br />

quirements <strong>of</strong> representations have been addressed. Subsequently, special attention<br />

has been devoted to illustrating certain aspects <strong>of</strong> the tools which ABEL applied in the<br />

theory <strong>of</strong> transcendentals. In particular, it has been documented, how ABEL made re-<br />

peated and prolific use <strong>of</strong> algebraic methods resembling those which he had employed<br />

in his research on the solubility <strong>of</strong> equations. Finally, the changing roles <strong>of</strong> definitions<br />

and representations have been briefly debated in the present chapter.<br />

In conclusion, ABEL’S research on elliptic functions and higher transcendentals<br />

was his most directly influential legacy. Culminating in the later formulation <strong>of</strong> JA-<br />

COBI, ABEL’S Paris memoir suggested a problem which attracted mathematicians for<br />

decades and influenced the development <strong>of</strong> mathematics in the entire nineteenth cen-<br />

tury. In the next and final chapter, aspects <strong>of</strong> ABEL’S own research on elliptic functions<br />

will serve to illustrate how he sometimes worked within the formula based paradigm.


Part V<br />

ABEL’s mathematics and the rise <strong>of</strong><br />

concepts<br />

385


Chapter 21<br />

ABEL’s mathematics and the rise <strong>of</strong><br />

concepts<br />

In the preceding parts, I have presented and discussed the main parts <strong>of</strong> N. H. ABEL’S<br />

(1802–1829) mathematical production ranging from the theory <strong>of</strong> equations over the<br />

installation <strong>of</strong> rigor in the theory <strong>of</strong> series to the exploding field <strong>of</strong> elliptic and higher<br />

transcendentals. In addition, in each part, I have simultaneously addressed three<br />

broader themes: the rise <strong>of</strong> new questions with new kinds <strong>of</strong> answers, the change<br />

in the standards <strong>of</strong> doing mathematics, and a change in the objects and methods <strong>of</strong><br />

mathematics.<br />

In this concluding part, I unify these themes by arguing that they are signs <strong>of</strong> a<br />

rise <strong>of</strong> concept based mathematics. In four steps, it will be argued that a large part <strong>of</strong><br />

the development in mathematics in the early nineteenth century can appropriately<br />

be analyzed from the perspective <strong>of</strong> a change in the objects with which mathematics<br />

dealt. Firstly, the change is introduced by defining and discussing the “paradigms”<br />

<strong>of</strong> formula based mathematics and concept based mathematics. Secondly, these concepts<br />

are employed to present a brief analysis <strong>of</strong> how the ways <strong>of</strong> introducing objects into<br />

mathematics developed in the early nineteenth century. Thirdly, the changing role<br />

<strong>of</strong> counter examples in the two paradigms is discussed in some details. Finally, some<br />

reservations to the analysis are presented before a conclusion is drawn. Due to the lim-<br />

ited scope <strong>of</strong> ABEL’S mathematical production, my frame <strong>of</strong> interpretation can only be<br />

preliminary and I hope to develop it further through subsequent research by involving<br />

the works <strong>of</strong> other mathematicians.<br />

21.1 From formulae to concepts<br />

It seems fair to state that in the early nineteenth century, mathematics changed quite<br />

dramatically. 1 Any comparative and contextualized reading <strong>of</strong> the works <strong>of</strong>, say, L.<br />

1 Whether or not the early nineteenth century is an apt periodization in the history <strong>of</strong> mathematics<br />

has been discussed, though. See e.g. (Mikulinsky, 1982; Otte, 1982).<br />

387


388 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

EULER (1707–1783) and K. T. W. WEIERSTRASS (1815–1897) or G. F. B. RIEMANN<br />

(1826–1866) will reveal that the problems, methods, and styles <strong>of</strong> analytical mathe-<br />

matics developed immensely and changed fundamentally during the century from<br />

the 1750s to the 1850s. In my view, a large part <strong>of</strong> the change in mathematics in the<br />

period can be understood by analyzing a fundamental change in the basic objects <strong>of</strong><br />

the combined discipline <strong>of</strong> algebra and analysis.<br />

My description <strong>of</strong> changes in mathematics will — dictated by the scope <strong>of</strong> the pre-<br />

vious parts — mostly deal with the development <strong>of</strong> the algebraic and analytic disci-<br />

plines. In section 21.4, the applicability <strong>of</strong> the analyses outside these disciplines is<br />

briefly addressed.<br />

It will be argued that mathematics in the eighteenth century was tied to formu-<br />

lae and that mathematicians worked within a framework which was — in essential<br />

ways — adapted to these objects. In the early nineteenth century, so it is argued, these<br />

basic objects were gradually replaced by concepts and the change was so fundamental<br />

that it influenced all layers <strong>of</strong> mathematical knowledge and knowledge production.<br />

To allow for a more precise discussion, tentative definitions <strong>of</strong> the two styles (paradigms)<br />

<strong>of</strong> mathematics are given below. I have adopted the excessively broad Kuhnian term<br />

“paradigm” to include the entire mental horizon <strong>of</strong> the group <strong>of</strong> mathematicians who<br />

worked in the tradition. At the same time, I have introduced catch-word characteri-<br />

zations <strong>of</strong> the paradigms by terming them formula based and concept based . Below, the<br />

paradigms and their relations will be discussed further and certain relevant aspects <strong>of</strong><br />

the preceding presentation <strong>of</strong> ABEL’S mathematics will be analyzed.<br />

21.1.1 <strong>The</strong> Eulerian paradigm <strong>of</strong> formula based mathematics<br />

By formula based mathematics, I mean to indicate a paradigm prevalent in the eighteenth<br />

century in which formulae were the carriers <strong>of</strong> mathematical knowledge. Formulae<br />

were both the results and the methods <strong>of</strong> mathematics, and mathematicians thought<br />

about and in terms <strong>of</strong> formulae. Mathematical results were derived through strings <strong>of</strong><br />

explicit, formal manipulations <strong>of</strong> representations (formulae) and were stated in terms<br />

<strong>of</strong> new formulae.<br />

<strong>The</strong> essential notion <strong>of</strong> formula can be thought <strong>of</strong> as representations <strong>of</strong> mathemat-<br />

ical objects by symbols. However, such interpretations tend to be anachronistic and<br />

beside the point because — as I shall argue — the formulae were the basic objects <strong>of</strong><br />

mathematics and only gradually became representations <strong>of</strong> other objects. 2<br />

In analysis, the primary occurrence <strong>of</strong> formulae was in the form <strong>of</strong> functions; the<br />

study <strong>of</strong> functions had been based on the study <strong>of</strong> their algebraic formulae. For these<br />

reasons, this paradigm could also have been named function based mathematics if it<br />

2 In the seventeenth and part <strong>of</strong> the eighteenth century, formulae had also been representations <strong>of</strong><br />

e.g. curves (see section 15.1). However, as described, they became the primary objects in EULER’S<br />

new version <strong>of</strong> analysis.


21.1. From formulae to concepts 389<br />

✬<br />

✫<br />

All explicit<br />

algebraic<br />

expressions<br />

✩✬<br />

=<br />

✪✫<br />

All algebraic<br />

expressions <strong>of</strong> the<br />

form (6.4).<br />

✩<br />

✪<br />

Figure 21.1: <strong>The</strong> equality <strong>of</strong> the concepts <strong>of</strong> explicit algebraic expressions and ABEL’s<br />

normal form.<br />

was only to apply in analysis. However, in the algebraic discipline such a description<br />

would be misguided precisely because not functions but formulae were at the centre<br />

<strong>of</strong> mathematical reasoning (see e.g. section 5.2). If any other name should be used for<br />

formula based mathematics, the term Eulerian paradigm might be well suited.<br />

For the present purpose, formula based mathematics is best thought <strong>of</strong> in terms <strong>of</strong><br />

e.g. EULER’S introduction and manipulation <strong>of</strong> various algebraic expressions in analy-<br />

sis. ABEL’S mathematics also frequently exhibits key characteristics <strong>of</strong> this paradigm,<br />

e.g. in his manipulations <strong>of</strong> formulae in the Recherches or the latter part <strong>of</strong> the bino-<br />

mial paper (see chapter 17 and 12, respectively). On both these occasions, ABEL based<br />

his deductions on sequences <strong>of</strong> step-wise manipulations <strong>of</strong> formulae to obtain results<br />

which were, themselves, formulae.<br />

21.1.2 A new paradigm <strong>of</strong> concept based mathematics<br />

<strong>The</strong> anti-thesis to formula based mathematics in the present context is termed concept<br />

based mathematics. In analogy with the formula based version, this paradigm empha-<br />

sized thought in and about concepts by which I mean classes <strong>of</strong> objects. <strong>The</strong> concept<br />

based mathematics deals primarily with defining, representing, and relating concepts.<br />

<strong>The</strong> collection <strong>of</strong> objects which fall under a concept is called the extension or domain <strong>of</strong><br />

the concept.<br />

Typically, concept based mathematics could be concerned with e.g. continuous func-<br />

tions, differentiable functions, or algebraically solvable equations. <strong>The</strong> mathematical theo-<br />

rems dealing with concepts would then contain results relating these, e.g. by pointing<br />

out their differences or their overlaps or by relating one concept to another. In a truly<br />

concept based approach to mathematics, even representations become theorems relat-<br />

ing concepts; ABEL’S deduction <strong>of</strong> the normal form for (explicit) algebraic expressions<br />

stated that the two concepts were identical (see figure 21.1, section6.3, and below).<br />

For concept based mathematics to be efficient, specific knowledge <strong>of</strong> the individual<br />

objects within a concept has to fade in importance. Individual objects would serve


390 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

important roles as examples and counter examples but the definitions <strong>of</strong> the concepts<br />

must possess qualities which make them useful and central in the investigation <strong>of</strong> the<br />

concept. Such investigations, in turn, can benefit from the shift <strong>of</strong> focus onto concepts<br />

and produce results which were impossible (or very difficult) if only individual objects<br />

were considered. Thus, in order to analyze concept based mathematics, the role <strong>of</strong><br />

definitions, representations, and arguments <strong>of</strong> relation and delineation <strong>of</strong> concepts<br />

become key points <strong>of</strong> enquiry.<br />

21.1.3 <strong>The</strong> shift from formula based to concept based mathematics<br />

<strong>The</strong> purpose <strong>of</strong> introducing the paradigms <strong>of</strong> formula based and concept based math-<br />

ematics is to characterize the development in the early nineteenth century as a tran-<br />

sition from the former to the latter. This transition manifests itself in various ways<br />

which interact with the changing basic objects <strong>of</strong> mathematics. <strong>The</strong> questions asked,<br />

the tools employed to answer these questions, and the types <strong>of</strong> answers which are<br />

possible and expected all change as consequences <strong>of</strong> this shift.<br />

In the first half <strong>of</strong> the nineteenth century, some mathematicians were aware that<br />

their style <strong>of</strong> mathematics differed essentially from the standards <strong>of</strong> their time. ABEL<br />

expressed how heavy loads <strong>of</strong> computations could hamper the progress <strong>of</strong> research, 3<br />

and E. GALOIS (1811–1832) described his own works as “analysis <strong>of</strong> analysis” which<br />

would reduce the hitherto dominating calculations to particular cases. 4 This aware-<br />

ness <strong>of</strong> the transition grew stronger during the century and towards the end <strong>of</strong> the<br />

nineteenth century, mathematicians became increasingly explicit about it. For in-<br />

stance, F. RUDIO (1856–1929) wrote:<br />

“<strong>The</strong> essential principle <strong>of</strong> the newer mathematical school, which is established<br />

by Gauss, Jacobi, and Dirichlet, is that whereas the older one sought to<br />

reach the goal by lengthy and complicated calculations (as even still in Gauss’ Disquisitiones)<br />

and deductions — it comprises an entire field by avoiding those and<br />

applying a genius method in a main idea and simultaneously presents the end<br />

result in its highest elegance by a single strike. While the former [the older approach]<br />

after a long sequence eventually reached a fertile ground by progressing<br />

from theorem to theorem, the latter [the new approach] immediately produces a<br />

formula in which the complete sphere <strong>of</strong> truths <strong>of</strong> an entire field is compactly contained<br />

and only ought to be extracted and expressed. In the old way, one could<br />

also — if need be — prove theorems; but only now can the true nature <strong>of</strong> the entire<br />

theory be seen, its internal gears and wheels.” 5<br />

3 (N. H. <strong>Abel</strong>, [1828] 1839, 217–218).<br />

4 (Galois, 1831c, 11).<br />

5 “Das wesentliche Princip der neueren mathematische Schule, die durch Gauss, Jacobi und Dirichlet<br />

begründet ist, ist im Gegensatz mit der älteren, dass während jene ältere durch langwierige und<br />

verwickelte Rechnung (wie selbst noch in Gauss’ Disquisitiones) und Deduktionen zum Zweck zu<br />

gelangen suchte, diese mit Vermeidung derselben durch Anwendung eines genialen Mittels in einer<br />

Hauptidee die Gesammtheit eines ganzen Gebietes umfasst und gleichsam durch einen einzigen<br />

Schlag das Endresultat in der höchsten Eleganz darstellt. Während jene, von Satz zu Satz fortschrei-


21.1. From formulae to concepts 391<br />

RUDIO clearly expressed the transition; the formula which he describes as the prod-<br />

uct <strong>of</strong> the new approach is not a formula in the present sense but rather a theorem.<br />

A number <strong>of</strong> similar statements can be found by other late-nineteenth century math-<br />

ematicians, e.g. C. F. KLEIN (1849–1925) who described how G. P. L. DIRICHLET<br />

(1805–1859) would avoid long computations in favor <strong>of</strong> acute logical analyses. 6<br />

This shift in the roles <strong>of</strong> formulae and concepts has been noticed and investigated<br />

from slightly different perspectives by historians <strong>of</strong> mathematics. In particular, it has<br />

been addressed by H. N. JAHNKE (⋆1948) and by D. LAUGWITZ (1932–2000), who<br />

pointed to the significant influence which RIEMANN had in bringing about the eventual<br />

change <strong>of</strong> paradigms. 7<br />

<strong>The</strong> basic objects <strong>of</strong> mathematics. <strong>The</strong> definitions <strong>of</strong> the paradigms suggest that the<br />

purported shift from formula based to concept based mathematics was a question <strong>of</strong><br />

the size <strong>of</strong> the domains <strong>of</strong> mathematical results. Interpreted purely as a change in<br />

domains, the new approach could be seen as consisting <strong>of</strong> results which are simul-<br />

taneously true for a number <strong>of</strong> objects <strong>of</strong> the old paradigm (formulae). However,<br />

there is more to the transition that this; it concerns a real and fundamental change<br />

from formulae to concepts as the basic objects <strong>of</strong> mathematics. In the one extreme, a<br />

manipulation <strong>of</strong> a particular algebraic formula might produce another algebraic for-<br />

mula which would then be a mathematical result. At the other end <strong>of</strong> the spectrum, a<br />

number <strong>of</strong> results developed in the nineteenth century pointed out the differences be-<br />

tween important concepts such as continuous and differentiable functions or proved<br />

that particular classes <strong>of</strong> functions could be represented in particular ways. <strong>The</strong> ability<br />

to state and prove results for abstractly defined classes <strong>of</strong> objects is one <strong>of</strong> the main as-<br />

pects <strong>of</strong> the rise <strong>of</strong> concept based mathematics. Similarly, the issues <strong>of</strong> relating concepts<br />

and representing concepts are two <strong>of</strong> the central topics in a fully fledged concept-based<br />

version <strong>of</strong> mathematics.<br />

<strong>The</strong> techniques and questions <strong>of</strong> mathematics. Connected to the transition in the<br />

basic objects <strong>of</strong> mathematics, the techniques and questions <strong>of</strong> mathematics also un-<br />

derwent fundamental changes. <strong>The</strong> types <strong>of</strong> questions asked and the methods for<br />

answering them were not the same in the two paradigms. In the formula based<br />

paradigm, mathematical texts could be made up <strong>of</strong> long sequences <strong>of</strong> manipulations<br />

which transformed one formula into others or answered particular questions by de-<br />

tend, nach einer langen Reihe endlich zu einigem fruchtbaren Boden gelangt, stellt diese gleich von<br />

vorn herein eine Formel hin, in welcher der vollständige Kreis der Wahrheiten eines ganzen Gebietes<br />

konzentriert enthalten ist und nur herausgelesen und ausgesprochen zu werden darf. Auf<br />

die frühere Art konnte man die Sätze zwar auch zur Not beweisen, aber jetzt sieht man erst das<br />

wahre Wesen der ganzen <strong>The</strong>orie, das eigentliche innere Getriebe und Räderwerk.” (F. Rudio, 1895,<br />

894–895).<br />

6 (Klein, 1967, 250).<br />

7 See e.g. (Jahnke, 1987) and (Laugwitz, 1999, 293–340). <strong>The</strong>se are both very interesting works dealing<br />

with discussions similar to the present one but from slightly different perspectives.


392 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

veloping formulae which “solved” them. At times, concept based mathematics could<br />

apply the manipulation <strong>of</strong> “representations” provided that some representation result<br />

made it relevant. But more typically and interestingly, in concept based mathematics,<br />

statements about the extension <strong>of</strong> a concept grew to become the key results <strong>of</strong> math-<br />

ematics. A particularly illustrative example <strong>of</strong> these questions has been presented<br />

in part II where ABEL’S research on the quintic first showed that not all polynomial<br />

equations were solvable, i.e. the concepts <strong>of</strong> polynomial equation and algebraically solv-<br />

able equations were distinct, although related. Later in his research, ABEL’S pro<strong>of</strong> <strong>of</strong><br />

the algebraic solubility <strong>of</strong> <strong>Abel</strong>ian equations was another almost prototypical concept<br />

based result, at least in its final formulation. This result showed that the concept <strong>of</strong><br />

<strong>Abel</strong>ian equations was contained in the concept <strong>of</strong> algebraically solvable equations. When<br />

he first encountered <strong>Abel</strong>ian equations in connection with the division problem (see<br />

section 16.3), ABEL’S argument relied extensively on his particular knowledge <strong>of</strong> the<br />

individual objects and was thus much more formula based.<br />

<strong>The</strong> styles <strong>of</strong> mathematics. Not surprisingly, the changing techniques <strong>of</strong> mathemat-<br />

ics manifested themselves at the textual level. Because formulae had been the carriers<br />

<strong>of</strong> knowledge and argument in the formula based paradigm, mathematical publica-<br />

tions relied extensively on the powers <strong>of</strong> formulae and mathematical texts could be<br />

dominated by strings <strong>of</strong> explicit manipulations <strong>of</strong> formulae. Eventually, a conclusion<br />

could be stated in the form <strong>of</strong> a theorem. In the concept based paradigm, a Euclidean<br />

style with its emphasis on definitions, theorems, and pro<strong>of</strong>s became the customary<br />

style <strong>of</strong> written mathematics. This presentational style emphasized the precise state-<br />

ment <strong>of</strong> assumptions and the internal relations between concepts and theorems.<br />

A revolution? When a change <strong>of</strong> paradigms is involved, the question <strong>of</strong> revolutions<br />

naturally arises. According to the recent debate, revolutions in mathematics appears<br />

not to be the most apt scheme <strong>of</strong> interpreting the history <strong>of</strong> the discipline. 8 In particu-<br />

lar, the requirements <strong>of</strong> incommensurability seems to prohibit revolutions appearing<br />

in mathematics because the truth status <strong>of</strong> mathematical statements apparently never<br />

changes. This also seems to apply to the change <strong>of</strong> paradigms discussed here. During<br />

and after the transitional period, mathematicians devoted an effort to reconstructing<br />

and re-interpreting the established knowledge to make it fit into the new system. This<br />

is particularly visible in analysis where A.-L. CAUCHY’S (1789–1857) deliberate redef-<br />

inition <strong>of</strong> basic notions and priorities changed the status <strong>of</strong> certain results and per-<br />

ceptions. As a result, men like ABEL sought to refound the theory in such a way that<br />

absolute truth was retained by making explicit the domains <strong>of</strong> validity for the state-<br />

ments. This process can be called critical revision and its general success precludes<br />

revolutions in mathematics. In section 21.3, the role <strong>of</strong> counter examples in the early<br />

nineteenth century is invoked to shed some light on this discussion.<br />

8 See primarily the articles in (Gillies, 1992).


21.2. Concepts and classes enter mathematics 393<br />

21.2 Concepts and classes enter mathematics<br />

As the basic objects <strong>of</strong> mathematics went from formulae to concepts, new methods<br />

and standards for introducing the objects were developed and the internal purpose <strong>of</strong><br />

mathematical research also changed.<br />

21.2.1 Defining concepts<br />

While an object in the formula based paradigm could be introduced by merely exhibit-<br />

ing its formula, the introduction <strong>of</strong> concepts into the concept based paradigm required<br />

more sophisticated methods. However, these methods were not necessarily new — a<br />

number <strong>of</strong> them had been around since the births <strong>of</strong> the Euclidean style <strong>of</strong> mathemat-<br />

ics and Aristotelian logic in Ancient Greece. In the present context, two aspects <strong>of</strong> the<br />

new importance given to definitions deserve special attention. First, genetic defini-<br />

tions and nominal definitions are discussed and their interactions described. Second,<br />

the introduction <strong>of</strong> concepts with special properties through careful definitions is em-<br />

phasized.<br />

Genetic and nominal definitions. Concepts were <strong>of</strong>ten introduced by either genetic<br />

or nominal definitions. A genetic definition consists <strong>of</strong> prescribing the way the concept<br />

is constructed from other, simpler concepts whereas a nominal definition simply as-<br />

sociates a name to something. Typical examples <strong>of</strong> a genetic definitions in the present<br />

material include EULER’S definition <strong>of</strong> functions and ABEL’S definition <strong>of</strong> explicit al-<br />

gebraic expressions (see sections 10.1 and 6.3, respectively). <strong>The</strong>se two examples also<br />

illustrate a very important difference in defining concepts: EULER’S definition was<br />

purely nominal whereas ABEL put his definition to essential use in obtaining his nor-<br />

mal form <strong>of</strong> explicit algebraic expressions. 9 Nominal definitions were being discussed<br />

in the early nineteenth century but the debate mainly centered on the ancient question<br />

whether or not nominal definitions implied the existence <strong>of</strong> any objects under the con-<br />

cept being defined. 10 <strong>The</strong> main objection against nominal definitions from a concept<br />

based paradigm could have been that they were not useful in obtaining knowledge <strong>of</strong><br />

the concept being defined. 11<br />

Definition by desired property. <strong>The</strong> ultimate way <strong>of</strong> associating knowledge through<br />

definitions would be to let properties serve as definitions. In a sense, this is the final<br />

lesson <strong>of</strong> I. LAKATOS’ (1922–1974) Pro<strong>of</strong>s and Refutations: the polyhedra which satisfy<br />

the Eulerian formula are collected as a concept and called Eulerian polyhedra and<br />

9 See e.g. (Laugwitz, 1999, 311) and section 6.3.<br />

10 Among the mathematicians involved were GERGONNE and OLIVIER. See e.g. (Otero, 1997, 74–81)<br />

and (Olivier, 1826c).<br />

11 In (Grabiner, 1981b), GRABINER has similarly emphasized the role which CAUCHY’S new definitions<br />

played for his foundation for the calculus.


394 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

<strong>of</strong> those, the Eulerian formula is trivially true. 12 Nevertheless, such definitions can<br />

be extremely useful in order to investigate other properties. With CAUCHY’S funda-<br />

mental shift towards arithmetic — rather than algebraic — equality, the numerical con-<br />

vergence <strong>of</strong> partial sums <strong>of</strong> series was given prime importance by using it to define<br />

convergent series. 13 Thus, a property <strong>of</strong> formal series — which could be numerically<br />

convergent or not — was used to define a concept which was subsequently promoted<br />

and investigated. A similar change went on in the theory <strong>of</strong> elliptic functions where<br />

ABEL’S original formal inversion <strong>of</strong> elliptic integrals was replaced by other definitions<br />

<strong>of</strong> elliptic functions. Many <strong>of</strong> the definitions <strong>of</strong> elliptic functions following ABEL’S<br />

original one turned properties — which were results in ABEL’S theory — into defini-<br />

tions. <strong>The</strong> motivations for this change in the status <strong>of</strong> properties <strong>of</strong> elliptic functions<br />

are many; rigor and theoretical applicability figure prominently among them.<br />

21.2.2 Relating concepts<br />

As a result <strong>of</strong> the transition, theorems about concepts and relating concepts came to<br />

dominate mathematics. Two types <strong>of</strong> relations among concepts were <strong>of</strong> principal im-<br />

portance: the representation <strong>of</strong> concepts and the determination <strong>of</strong> the extension <strong>of</strong><br />

concepts.<br />

Representing concepts. Mathematical symbolism and formulae had proved to be<br />

an extremely useful and powerful tool in developing theories in the formula based<br />

paradigm. In order to be able to continue this line <strong>of</strong> research into the concept based<br />

paradigm, representations <strong>of</strong> concepts became quite important. Central instances in-<br />

clude ABEL’S classification <strong>of</strong> explicit algebraic expressions and the multitude <strong>of</strong> rep-<br />

resentations <strong>of</strong> elliptic functions which he developed. A particularly revealing exam-<br />

ple <strong>of</strong> the benefits <strong>of</strong> representations was illustrated in section 18.1 where ABEL’S use<br />

<strong>of</strong> infinite representations in the theory <strong>of</strong> transformation was discussed. <strong>The</strong> study<br />

<strong>of</strong> concepts in their entirety and not the individual objects meant that statements con-<br />

cerning the impossibility <strong>of</strong> certain representations could also be made and proved.<br />

This is particularly true <strong>of</strong> ABEL’S pro<strong>of</strong> <strong>of</strong> the insolubility <strong>of</strong> the quintic (see chapter<br />

6) in which a representation <strong>of</strong> all explicit algebraic expressions was proved not to be<br />

sufficiently powerful to encompass the implicitly defined algebraic expression corre-<br />

sponding to a solution <strong>of</strong> the general fifth degree equation. <strong>The</strong> very same example<br />

also serves to illustrate the problem <strong>of</strong> distinguishing concepts.<br />

Distinguishing concepts. With the focus on concepts, it also became an important<br />

question to determine whether two concepts were identical or differed in their ex-<br />

tensions. One <strong>of</strong> the very best examples is the debate which during the nineteenth<br />

12 (Lakatos, 1976).<br />

13 See section 11.1.


21.3. <strong>The</strong> role <strong>of</strong> counter examples 395<br />

century separated the concepts <strong>of</strong> continuous and differentiable functions by construct-<br />

ing ever more pathological functions belonging to the former concept but not to the<br />

latter one. 14 <strong>The</strong> process <strong>of</strong> investigating concepts can <strong>of</strong>ten be thought <strong>of</strong> as a dialectic<br />

effort alternating between limiting and extending the domain <strong>of</strong> the concept. In point-<br />

ing out the existence <strong>of</strong> objects within a concept and differences between concepts,<br />

examples and counter examples became very important mathematical tools. A num-<br />

ber <strong>of</strong> similar uses <strong>of</strong> examples and counter examples can also be found in ABEL’S<br />

works. <strong>The</strong> most conspicuous example is in the theory <strong>of</strong> equations where ABEL’S<br />

pro<strong>of</strong> on the quintic interpreted as limiting the class <strong>of</strong> solvable equations is precisely<br />

in this line <strong>of</strong> results. <strong>The</strong> use <strong>of</strong> counter examples as limitations on concepts is a quite<br />

modern one which is only meaningful within the concept based paradigm (see section<br />

21.3).<br />

Delineating concepts. One type <strong>of</strong> questions concerning the relation between con-<br />

cepts is so important that it deserves special attention; I have called it delineation <strong>of</strong><br />

concepts. This notion refers to a set <strong>of</strong> questions which concern the precise characteri-<br />

zation <strong>of</strong> the extension <strong>of</strong> a concept by some external and applicable criterion. In other<br />

words, these questions ask for a (feasible) method <strong>of</strong> determining whether a given par-<br />

ticular object falls within the extension <strong>of</strong> a concept or not. In analogy with the steps<br />

<strong>of</strong> limiting and extending the extension <strong>of</strong> concepts (figures 6.1 and 7.3, respectively),<br />

a graphical representation <strong>of</strong> the delineation <strong>of</strong> concepts is produced in figure 21.2.<br />

ABEL’S unfinished research on a general theory <strong>of</strong> algebraic solubility was moti-<br />

vated by precisely this problem <strong>of</strong> determining whether or not a given equation could<br />

be solved algebraically. Similarly, the search for complete criteria <strong>of</strong> convergence also<br />

sought to delineate the extension <strong>of</strong> convergent series once and for all. <strong>The</strong> search for<br />

delineation <strong>of</strong> solvable equations came to a fruitful conclusion when GALOIS’ criterion<br />

was finally accepted as an answer. <strong>The</strong> complete determination <strong>of</strong> the concept <strong>of</strong> con-<br />

vergent series was never so successful; the only complete characterization obtained<br />

was the Cauchy criterion (see page 212) which did not fully meet the demand for being<br />

external and easy to apply.<br />

21.3 <strong>The</strong> role <strong>of</strong> counter examples<br />

It has been described how the problem <strong>of</strong> investigating the extension <strong>of</strong> concepts<br />

led to a particular use <strong>of</strong> examples and counter examples. Inspired by ABEL’S cu-<br />

rious remarks about his “exception” to Cauchy’s <strong>The</strong>orem (see section 12.5), I suggest<br />

that counter examples played fundamentally different roles in the two paradigms dis-<br />

cussed here.<br />

14 See (K. Volkert, 1987; K. Volkert, 1989).


396 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

❅❅❘ ✬<br />

❅❅■<br />

✫<br />

��✒ ��✠<br />

Super-concept<br />

❄<br />

✻<br />

��✠ ✩<br />

��✒ ✲✛ Concept ✲✛<br />

❄<br />

✻<br />

❅❅❘ ✪<br />

❅❅■<br />

Figure 21.2: Delineating the border between a concept and its super-concept.<br />

21.3.1 <strong>The</strong>orems with exceptions<br />

In his binomial paper, ABEL described how he found Cauchy’s <strong>The</strong>orem to “suffer ex-<br />

ceptions” and I find it puzzling to investigate how theorems could possibly admit<br />

exceptions in the 1820s. First, however, the very phrasing <strong>of</strong> ABEL’S statement must<br />

be considered. <strong>The</strong>n, by way <strong>of</strong> recalling arguments carried out “in general”, the con-<br />

nection between exceptions and the formula based paradigm opens up.<br />

<strong>The</strong> authenticity <strong>of</strong> the wording. One may try to explain ABEL’S wording away as<br />

a result <strong>of</strong> his shyness and veneration for CAUCHY. For instance, in his criticism <strong>of</strong><br />

L. OLIVIER, ABEL used the mild phrase “this part does not seem to be true” in the<br />

printed version rather than the more severe judgement “Mr. Olivier is seriously mis-<br />

taken” which we find in ABEL’S notebooks. 15 This would suggest that exceptions were<br />

a milder form <strong>of</strong> criticism than outright counter examples or even paradoxes which<br />

were also terms found in ABEL’S vocabulary. Besides, the problem remains that we<br />

only have A. L. CRELLE’S (1780–1855) translation <strong>of</strong> ABEL’S original manuscript at<br />

our disposal and single words in an edited manuscript can easily be over-interpreted.<br />

Nevertheless, when CAUCHY eventually reacted to ABEL’S exception, he did so explic-<br />

itly stating that he wanted to correct the statement <strong>of</strong> his theorem so that “it no longer<br />

admitted exceptions” 16 (see section 14.1.2). Thus, the word “exceptions” was chosen<br />

in this connection, and I believe that the following interpretation makes it plausible<br />

that ABEL actually meant that Cauchy’s <strong>The</strong>orem suffered an exception — or rather, a<br />

number <strong>of</strong> exceptions.<br />

15 See section 13.2.<br />

16 “Au reste, il est facile de voir comment on doit modifier l’énoncé du théorème, pour qu’il n’y ait<br />

plus lieu à aucune exception.” (A.-L. Cauchy, 1853, 31–32).


21.3. <strong>The</strong> role <strong>of</strong> counter examples 397<br />

Arguments carried out “in general”. In the formula based paradigm, a situation<br />

sometimes arose in which the formula carrying the mathematical argument did not<br />

apply in all (numerical) cases. A number <strong>of</strong> such examples have been described above<br />

starting with EULER’S awareness that peculiar numerical results could emerge if specific<br />

values were inserted in expressions which were formally equal. 17<br />

As J. V. GRABINER describes, 18 J. L. LAGRANGE (1736–1813) held a strong and<br />

lifelong belief in the concept <strong>of</strong> “the general” in mathematics. Not only could formulae<br />

which were valid “in general” be <strong>of</strong> high importance — a general approach and system<br />

in mathematics was also strived for. When LAGRANGE presented his argument that<br />

“all” functions could be expanded in Taylor series, he was also aware that this might<br />

indeed fail to be true for particular functions at particular points. 19 However, these<br />

instances where the general results failed to be true were particular, peculiar, and <strong>of</strong><br />

little interest to mathematicians ascribing to the formula based paradigm.<br />

In connection with ABEL’S Paris memoir, an even more elaborate case was pre-<br />

sented. At a crucial point in his argument to determine the number µ <strong>of</strong> independent<br />

integrals, ABEL employed a generalized degree operator called h. 20 Just as is the case<br />

for the ordinary degree operator <strong>of</strong> polynomials deg P (which ABEL also used), the<br />

degree <strong>of</strong> a sum may fail to be the maximum <strong>of</strong> the two degrees,<br />

deg (P1 + P2) ? = max {deg P1, deg P2} , (21.1)<br />

if deg P1 = deg P2. However, in the Paris memoir, ABEL was not interested in peculiar-<br />

ities and he simply argued that the equality corresponding to (21.1) was true “in gen-<br />

eral”, i.e. with the exception <strong>of</strong> some particular cases <strong>of</strong> little interest (see page 362).<br />

Once the paradigms had shifted, the precise determination <strong>of</strong> the number µ (called<br />

the genus) and the investigation and exposure <strong>of</strong> the necessary assumptions became<br />

a hot topic <strong>of</strong> mathematics.<br />

In a similar situation, ABEL concluded his summary <strong>of</strong> well known properties <strong>of</strong><br />

elliptic functions in the Précis by the statement:<br />

“<strong>The</strong> formulae which have been presented above uphold with certain restrictions<br />

if the modulus c is arbitrary, real or imaginary.” 21<br />

ABEL’S way <strong>of</strong> obtaining the important formulae — <strong>of</strong>ten through tedious manipula-<br />

tions <strong>of</strong> infinite representations — could result in particular cases for which the formu-<br />

lae degenerated or produced false results. However, these cases were few and did not<br />

constitute an obstacle to presenting the formulae.<br />

17 See section 10.1.<br />

18 See (Grabiner, 1981a, 317) or (Grabiner, 1981b, 39).<br />

19 See e.g. (Lagrange, 1813, 29–30).<br />

20 See section 19.3.<br />

21 “Les formules présentées dans ce qui précède ont lieu avec quelques restrictions, si le module c est<br />

quelconque, réel ou imaginaire.” (N. H. <strong>Abel</strong>, 1829d, 245).


398 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

<strong>The</strong> number <strong>of</strong> exceptions. As indicated, in the formula based paradigm, results<br />

which suffered a few exceptions could still be very useful and the existence <strong>of</strong> ex-<br />

ceptions did not immediately lead to the overthrow <strong>of</strong> theorems. This suggests an<br />

interesting way <strong>of</strong> interpreting the last part <strong>of</strong> ABEL’S famous footnote: Besides in-<br />

troducing his exception, ABEL also claimed that many similar functions existed. This<br />

indicates that the number <strong>of</strong> exceptions played a role. A similar remark can also be<br />

found in connection with CAUCHY’S example <strong>of</strong> a non-zero function whose Maclaurin<br />

series is the zero-function, 22<br />

1 −<br />

f (x) = e x2 .<br />

This function represented an exception to the general belief in the expansion in power<br />

series which laid at the heart <strong>of</strong> the Lagrangian approach to analysis. 23 In 1822 and<br />

1829, 24 CAUCHY presented this example and observed how to construct other func-<br />

tions with the same property <strong>of</strong> not being represented by their Maclaurin series except<br />

at a single point.<br />

Both these examples suggest that if theorems in the formula based paradigm con-<br />

tained a quantification as “for all . . . ”, it might be necessary to introduce a statistical<br />

interpretation <strong>of</strong> the for-all quantification as K. VOLKERT has suggested. 25 Exceptions<br />

and their numbers were noticed but no clear distinction between refuted (false) the-<br />

orems and theorems with exceptions can be drawn. <strong>The</strong>orems could be valid even if<br />

they suffered exceptions as long as the known exceptions were not too many or too<br />

important.<br />

Exceptions and the formula based paradigm. Thus, the argument is that exceptions<br />

did have a place in mathematics <strong>of</strong> the formula based paradigm. <strong>The</strong> highly compu-<br />

tational deductions based on long sequences <strong>of</strong> manipulations with finite and infinite<br />

representations occasionally led to results which were (only) true “in general”. In-<br />

stead <strong>of</strong> discarding such results, they were accepted with the knowledge or intuition<br />

that they should not be uncritically applied. However, as this intuition and general un-<br />

derstanding <strong>of</strong> mathematics shifted towards the concept based paradigm, exceptions<br />

became oddities — and counter examples became very powerful tools <strong>of</strong> argument in<br />

this new paradigm.<br />

21.3.2 Counter examples and concepts<br />

In the concept based paradigm, counter examples acquired a position much closer to<br />

their modern usage. As noted, counter examples are very instrumental in pointing<br />

out the differences between concepts and thereby helping to determine the extension<br />

22 Strictly speaking, the function should also be defined at the origin, f (0) = 0. For a good discussion<br />

on this issue, see (Bottazzini, 1990, lxix).<br />

23 See section 10.2.<br />

24 (A.-L. Cauchy, 1822, 277) and (A.-L. Cauchy, 1829, 394–395).<br />

25 (K. T. Volkert, 1986, 144–145).


21.3. <strong>The</strong> role <strong>of</strong> counter examples 399<br />

<strong>of</strong> concepts. Used as tools <strong>of</strong> criticism, a theorem to which a counter example could<br />

be presented was certainly false in the concept based approach. <strong>The</strong>re was no room<br />

for theorems with exceptions. In a sense, the concept based approach adhered to a<br />

viewpoint similar to the Lakatosian one that theorems with counter examples should<br />

either be discarded or modified to range over a smaller domain. <strong>The</strong>re is an abun-<br />

dance <strong>of</strong> such applications <strong>of</strong> counter examples in the 1820s. ABEL presented one very<br />

elaborate example in his refutation <strong>of</strong> OLIVIER when he showed that no criterion <strong>of</strong><br />

the proposed form could ever be constructed having the properties which OLIVIER<br />

had sought. However, the young mathematician who made the most use <strong>of</strong> counter<br />

examples in the 1820s and 1830s was probably DIRICHLET.<br />

In 1829, 26 when DIRICHLET presented his famous result on the convergence <strong>of</strong><br />

Fourier series, he started the paper with a scrutiny <strong>of</strong> an earlier paper by CAUCHY. 27<br />

In particular, DIRICHLET criticized a point in the pro<strong>of</strong> where CAUCHY had used an<br />

implicit assumption which DIRICHLET identified as follows: If the series ∑ an was<br />

convergent, any other series ∑ bn such that lim bn = 1 would also be convergent.<br />

an<br />

Against this argument, DIRICHLET presented the counter example<br />

an = (−1)n<br />

√ and bn =<br />

n (−1)n<br />

�<br />

√ 1 +<br />

n<br />

(−1)n<br />

�<br />

√<br />

n<br />

<strong>of</strong> which the series ∑ an was convergent but the series ∑ bn diverged. DIRICHLET de-<br />

scribed CAUCHY’S conclusion as “not permissible” 28 because it was easy to construct<br />

a counter example.<br />

To DIRICHLET, the existence <strong>of</strong> one single, local counter example thus seems to<br />

have rendered the theorem false; in particular, we find none <strong>of</strong> the above remarks that<br />

“infinitely many similar counter examples may be found or constructed” in DIRICH-<br />

LET’S papers. 29 In some instances, a counter example led DIRICHLET to dismiss the<br />

faulty theorems as false and begin his own deductions from other principles. In other<br />

situations, DIRICHLET drew inspiration from his counter examples to revise existing<br />

pro<strong>of</strong>s in ways which later led to pro<strong>of</strong> analysis.<br />

Later in the nineteenth century, counter examples acquired their modern status as<br />

complete refutations <strong>of</strong> theorems. To a mathematician educated at one <strong>of</strong> the German<br />

universities in the second half <strong>of</strong> the nineteenth century, a theorem could absolutely<br />

not admit exceptions and the precise formulation <strong>of</strong> theorems and pro<strong>of</strong>s had truly<br />

become one <strong>of</strong> the trademarks <strong>of</strong> mathematics.<br />

ABEL’S use <strong>of</strong> counter examples seems to fall in both paradigms. As noted, sense<br />

can be made <strong>of</strong> ABEL’S exception to Cauchy’s <strong>The</strong>orem if it is interpreted in the formula<br />

26 (G. L. Dirichlet, 1829, 120).<br />

27 (A.-L. Cauchy, 1827). Actually, DIRICHLET referred to a paper published in 1823 in the Mémoires de<br />

l’Académie des Sciences; but no paper with these details can be found in CAUCHY’S Œuvres. Thus, it<br />

is here assumed that DIRICHLET actually meant (ibid.).<br />

28 “Mais cette conclusion n’est pas permise” (G. L. Dirichlet, 1829, 158).<br />

29 (G. L. Dirichlet, 1829; G. L. Dirichlet, 1837).


400 Chapter 21. ABEL’s mathematics and the rise <strong>of</strong> concepts<br />

based paradigm. On the other hand, ABEL’S dismissal <strong>of</strong> OLIVIER’S criterion <strong>of</strong> con-<br />

vergence shows signs <strong>of</strong> a concept based refutation. <strong>The</strong>re, a single counter example<br />

was invoked to refute OLIVIER’S claim and an elaborate analysis was employed to<br />

show that the concept <strong>of</strong> tests <strong>of</strong> convergence could not contain a criterion <strong>of</strong> a slightly<br />

generalized form. 30<br />

<strong>The</strong> role <strong>of</strong> mathematical intuition. Importantly, the changing status <strong>of</strong> counter ex-<br />

amples also reflects a change in a mental entity which may be called mathematical in-<br />

tuition. 31 This intuition comprises the expertise, prejudices, and expectations <strong>of</strong> active<br />

mathematicians who have gained an insight into their objects and is thus part <strong>of</strong> T. S.<br />

KUHN’S (1922–1996) disciplinary matrix.<br />

During the eighteenth century, mathematicians built up a high degree <strong>of</strong> insight<br />

into representations <strong>of</strong> functions, in particular into power series or other algebraic<br />

expressions. When this insight was formulated, it <strong>of</strong>ten took the form <strong>of</strong> formulae<br />

relating certain entities by means <strong>of</strong> algebraic notation and the formulae were consid-<br />

ered to have aesthetic properties described as simplicity or degrees <strong>of</strong> symmetry. As<br />

illustrative examples, consider the solution formulae for general equations <strong>of</strong> low de-<br />

gree or the power series expansions <strong>of</strong> elementary transcendental functions. <strong>The</strong> art <strong>of</strong><br />

mathematics also consisted <strong>of</strong> the trained ability to recognize patterns and manipulate<br />

representations to obtain various generalizations.<br />

As a result <strong>of</strong> the change <strong>of</strong> paradigms, the contents and role <strong>of</strong> mathematical in-<br />

tuition also changed. A new kind <strong>of</strong> intuition emerged which helped mathematicians<br />

see differences and similarities between concepts and suggested ways <strong>of</strong> obtaining re-<br />

lations among concepts. As an indication <strong>of</strong> this change in intuition, mathematicians<br />

occasionally brought over intuitions from the old paradigm into the new one. This<br />

could lead them to generalize results into forms in which they were then no longer<br />

permitted. Thus, the changing intuitions are intimately connected with the process <strong>of</strong><br />

concept stretching which LAKATOS has discussed as part <strong>of</strong> interpreting mathematical<br />

development. 32<br />

21.4 Conclusion<br />

<strong>The</strong> analytical scheme <strong>of</strong> a transition from a formula based paradigm to one based on<br />

concepts has shown its applicability in interpreting events in the disciplines <strong>of</strong> algebra<br />

and analysis in the 1820s. In particular, the role <strong>of</strong> new definitions, the coexistence <strong>of</strong><br />

theorems and exceptions, and the new problems <strong>of</strong> delineation have contributed to<br />

throwing ABEL’S mathematical production into perspective.<br />

30 See chapter 13.<br />

31 For a discussion <strong>of</strong> mathematical intuition, see also (K. T. Volkert, 1986).<br />

32 E.g. (Lakatos, 1976).


21.4. Conclusion 401<br />

It is beyond the present scope to analyze and speculate as to the causal reasons for<br />

the purported change <strong>of</strong> paradigms in analysis and algebra. Neither is it the present<br />

purpose to discuss at length the general applicability <strong>of</strong> this frame <strong>of</strong> interpretation.<br />

However, it must be noticed that the interpretation might be limited to the disciplines<br />

described here; in particular, it does not appear to be immediately applicable to geom-<br />

etry. With some right, one could argue that the transition could be interpreted simply<br />

as a maturing <strong>of</strong> the involved theories. Still, I believe that the simultaneous instances<br />

<strong>of</strong> the change <strong>of</strong> style as described above are sufficient to suggest that a general change<br />

in the modes <strong>of</strong> thought was involved.<br />

New questions, new standards, new objects. In the presentation and analysis <strong>of</strong><br />

ABEL’S mathematical production, three local themes were introduced. Based on the<br />

description <strong>of</strong> his works in algebra, I have argued that a new type <strong>of</strong> questions was<br />

being introduced into mathematics. <strong>The</strong>se new questions were indicative <strong>of</strong> the fun-<br />

damental change <strong>of</strong> paradigms. Concerning ABEL’S works in the foundations <strong>of</strong> anal-<br />

ysis, it was illustrated how the change <strong>of</strong> basic definitions and standards <strong>of</strong> pro<strong>of</strong> also<br />

reflected the new focus on concepts. Finally, a cross section <strong>of</strong> ABEL’S works on new<br />

transcendentals illustrated how these transcendental objects were being treated with<br />

the help <strong>of</strong> algebraic methods and also how the introduction <strong>of</strong> new objects led to<br />

important questions <strong>of</strong> representation.<br />

Throughout the description and analysis <strong>of</strong> ABEL’S works, much attention has<br />

been paid to their mathematical contexts. <strong>The</strong> inspirations which ABEL drew from<br />

his predecessors and contemporaries have been described in order to illustrate how<br />

ABEL’S works grew continuously out <strong>of</strong> the mathematical contexts. At the same time,<br />

ABEL’S works were — at a number <strong>of</strong> points — remarkably novel and due attention<br />

has been paid to these aspects. To generalize, ABEL’S methods and the problems<br />

which he attacked were generally well established whereas the questions which he<br />

raised and the approaches which he took in attacking these problems were <strong>of</strong>ten new<br />

and ground-breaking. In connection with the fundamental transition, this manifested<br />

itself in the sense that ABEL had one foot firmly placed in each <strong>of</strong> the two paradigms.


Appendix A<br />

ABEL’s correspondence<br />

In 1881, when ABEL’s collected works were published in their second edition, SYLOW<br />

and LIE included some <strong>of</strong> ABEL’s correspondence. 1 In 1902, a centennial Festschrift on<br />

ABEL’s life, work, and correspondence appeared both in Norwegian and in French<br />

including transcriptions <strong>of</strong> all known letters to and from ABEL. 2 In the twentieth<br />

century, however, additional letters have been found, and for the convenience <strong>of</strong> the<br />

reader, the table ?? provides a list <strong>of</strong> all the letters pertaining to ABEL known to the<br />

author. All the letters appearing in the Norwegian version <strong>of</strong> the Festschrift were also<br />

included in the French version. Missing information for year, month, or date indicate<br />

that that particular information was not available in the letter.<br />

Table A.1: Correspondence sorted by sender<br />

1822/01/18 (<strong>Abel</strong>→Aas, Kristiania, 1822/01/18. In Kragemo, 1929, 49)<br />

1822/01/25 (<strong>Abel</strong>→Aas, Kristiania, 1822/01/25. In ibid., 49–50)<br />

1822/02/06 (<strong>Abel</strong>→Aas, Christiania, 1822/02/06. In ibid., 50)<br />

1826/10/16 (<strong>Abel</strong>→<strong>Abel</strong>, Paris, 1826/10/16. In N. H. <strong>Abel</strong>, 1902a, 41–43)<br />

1827/02/26 (<strong>Abel</strong>→Boeck, Berlin, 1827/02/26. In ibid., 55–56)<br />

1827/01/15 (<strong>Abel</strong>→Boeck, Berlin, 1827/01/15. In ibid., 52–55)<br />

1826/11/01 (<strong>Abel</strong>→Boeck, Paris, 1826/11/01. In ibid., 47–48)<br />

1828/08/18 (<strong>Abel</strong>→Crelle, Christiania, 1828/08/18. In ibid., 67–73)<br />

1827 (<strong>Abel</strong>→Crelle, Christiania, 1827. In ibid., 60–61)<br />

1827/11/15 (<strong>Abel</strong>→Crelle, Christiania, 1827/11/15. In ibid., 61–62)<br />

1826/03/14 (<strong>Abel</strong>→Crelle, Freiberg, 1826/03/14. In N. H. <strong>Abel</strong>, 1881, 266)<br />

1827/11/15 (<strong>Abel</strong>→Crelle, Christiania, 1827/11/15. In ibid., 268)<br />

1828/10/18 (<strong>Abel</strong>→Crelle, Christiania, 1828/10/18. In ibid., 269–270)<br />

1828/10/18 (<strong>Abel</strong>→Crelle, Christiania, 1828/10/18. In Biermann, 1967, 27–29)<br />

1828? (<strong>Abel</strong>→Crelle, 1828?. In Biermann and Brun, 1958, 85)<br />

1826/12/04 (<strong>Abel</strong>→Crelle, Paris, 1826/12/04. In N. H. <strong>Abel</strong>, 1902a, 50–51)<br />

1826/03/14 (<strong>Abel</strong>→Crelle, Freyberg, 1826/03/14. In ibid., 21–22)<br />

1 (N. H. <strong>Abel</strong>, 1881)<br />

2 (N. H. <strong>Abel</strong>, 1902e; N. H. <strong>Abel</strong>, 1902f).<br />

403


404 Appendix A. ABEL’s correspondence<br />

Table A.1: Correspondence sorted by sender (cont.)<br />

1826/08/09 (<strong>Abel</strong>→Crelle, Paris, 1826/08/09. In N. H. <strong>Abel</strong>, 1902a, 38–39)<br />

1826/08/09 (<strong>Abel</strong>→Crelle, Paris, 1826/08/09. In N. H. <strong>Abel</strong>, 1881, 267)<br />

1826/12/04 (<strong>Abel</strong>→Crelle, Paris, 1826/12/04. In ibid., 268)<br />

1824/03/02 (<strong>Abel</strong>→Degen, Christiania, 1824/03/02. In P. Heegaard, 1935, 33–37)<br />

1824/03/02 (<strong>Abel</strong>→Degen, Christiania, 1824/03/02. In P. Heegaard, 1937, 1–5)<br />

1828/09/22 (<strong>Abel</strong>→Fru Hansteen, Christiania, 1828/09/22. In N. H. <strong>Abel</strong>, 1902a, 67)<br />

1828/07/29 (<strong>Abel</strong>→Fru Hansteen, Froland, 1828/07/29. In ibid., 65)<br />

1828/08 (<strong>Abel</strong>→Fru Hansteen, Froland, 1828/08. In ibid., 65–66)<br />

1828/07/21 (<strong>Abel</strong>→Fru Hansteen, Froland, 1828/07/21. In ibid., 63–64)<br />

(<strong>Abel</strong>→Fru Hansteen. In ibid., 62)<br />

1827/03 (<strong>Abel</strong>→Fru Hansteen, Berlin, 1827/03. In ibid., 58–59)<br />

1827/08/18 (<strong>Abel</strong>→Fru Hansteen, Christiania, 1827/08/18. In ibid., 60)<br />

1826/01/16 (<strong>Abel</strong>→Fru Hansteen, Berlin, 1826/01/16. In ibid., 19–20)<br />

1825?/12/08 (<strong>Abel</strong>→Fru Hansteen, Berlin, 1825?/12/08. In ibid., 12–13)<br />

1828/11 (<strong>Abel</strong>→Fru Hansteen, Christiania, 1828/11. In ibid., 75–78)<br />

[1826]/01/30 (<strong>Abel</strong>→Hansteen, Berlin, [1826]/01/30. In ibid., 20–21)<br />

1826/03/29 (<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. In N. H. <strong>Abel</strong>, 1881, 263–265)<br />

1826/05/28 (<strong>Abel</strong>→Hansteen, Grätz, 1826/05/28. In N. H. <strong>Abel</strong>, 1902a, 32)<br />

1826/03/29 (<strong>Abel</strong>→Hansteen, Dresden, 1826/03/29. In ibid., 22–26)<br />

1826/08/12 (<strong>Abel</strong>→Hansteen, Paris, 1826/08/12. In ibid., 39–41)<br />

1825/12/05 (<strong>Abel</strong>→Hansteen, Berlin, 1825/12/05. In ibid., 9–12)<br />

1823/06/15 (<strong>Abel</strong>→Holmboe, Kjøbenhavn, 1823/06/15. In ibid., 3–4)<br />

1823/08/04 (<strong>Abel</strong>→Holmboe, Kjøbenhavn, 1823/08/04. In ibid., 4–8)<br />

1828/06/29 (<strong>Abel</strong>→Holmboe, Froland, 1828/06/29. In ibid., 64–65)<br />

1826/01/16 (<strong>Abel</strong>→Holmboe, 1826/01/16. In ibid., 13–19)<br />

1823/08/03 (<strong>Abel</strong>→Holmboe, Copenhague, 1823/08/03. In N. H. <strong>Abel</strong>, 1881, 254–258)<br />

1826/10/24 (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. In ibid., 259–261)<br />

1826/12 (<strong>Abel</strong>→Holmboe, Paris, 1826/12. In ibid., 261–262)<br />

1827/03/04 (<strong>Abel</strong>→Holmboe, Berlin, 1827/03/04. In ibid., 262)<br />

1827/03/04 (<strong>Abel</strong>→Holmboe, Berlin, 1827/03/04. In N. H. <strong>Abel</strong>, 1902a, 56–58)<br />

1826/12 (<strong>Abel</strong>→Holmboe, Paris, 1826/12. In ibid., 51–52)<br />

1827/01/20 (<strong>Abel</strong>→Holmboe, Berlin, 1827/01/20. In ibid., 55)<br />

1826/04/16 (<strong>Abel</strong>→Holmboe, Wien, 1826/04/16. In ibid., 26–31)<br />

1826/10/24 (<strong>Abel</strong>→Holmboe, Paris, 1826/10/24. In ibid., 43–47)<br />

1826/06/15 (<strong>Abel</strong>→Holmboe, Bolzano, 1826/06/15. In ibid., 33–37)<br />

1825/09/15 (<strong>Abel</strong>→Holmboe, Kjøbenhavn, 1825/09/15. In ibid., 9)<br />

1826/07/05 (<strong>Abel</strong>→Keilhau, Zurich, 1826/07/05. In ibid., 37)<br />

1826/11/01 (<strong>Abel</strong>→Külp, Paris, 1826/11/01. In Hensel, 1903, 237–240)<br />

1828/11/25 (<strong>Abel</strong>→Legendre, Christiania, 1828/11/25. In N. H. <strong>Abel</strong>, 1902a, 78–86)<br />

1828/11/25 (<strong>Abel</strong>→Legendre, Christiania, 1828/11/25. In N. H. <strong>Abel</strong>, 1881, 271–279)<br />

1823 (<strong>Abel</strong>→Olsen, Christiania, 1823. In Brun and Jessen, 1958, 22–23)


Table A.1: Correspondence sorted by sender (cont.)<br />

1828/09/10 (Crelle→<strong>Abel</strong>, 1828/09/10. In N. H. <strong>Abel</strong>, 1902a, 66)<br />

1828/05/18 (Crelle→<strong>Abel</strong>, 1828/05/18. In ibid., 62)<br />

1829/04/08 (Crelle→<strong>Abel</strong>, Berlin, 1829/04/08. In ibid., 89–90)<br />

1826/11/24 (Crelle→<strong>Abel</strong>, Berlin, 1826/11/24. In ibid., 48–50)<br />

1829/05/10 (Crelle→Holmboe, Berlin, 1829/05/10. In N. H. <strong>Abel</strong>, 1902b, 97–98)<br />

1840/05/15 (Crelle→Holmboe, Berlin, 1840/05/15. In ibid., 102)<br />

1821/05/21 (Degen→Hansteen, Kjøbenhavn, 1821/05/21. In ibid., 93–96)<br />

1830/07/24 (Det franske Institut→<strong>Abel</strong>’s efterladte, Paris, 1830/07/24. In N. H. <strong>Abel</strong>, 1902a, 101)<br />

1752 (Euler→Goldbach, 1752. In Euler and Goldbach, 1965)<br />

1829 (Jacobi→Legendre, Potsdam, 1829. In Legendre and Jacobi, 1875)<br />

1830/02/22 (Keilhau→Boeck, Froland, 1830/02/22. In N. H. <strong>Abel</strong>, 1902b, 98–100)<br />

1781 (Lagrange→d’Alembert, Berlin, 1781. In Lagrange, 1867–1892, vol. 13, 368–370)<br />

1828/10/25 (Legendre→<strong>Abel</strong>, Paris, 1828/10/25. In N. H. <strong>Abel</strong>, 1902a, 74–75)<br />

1829/01/16 (Legendre→<strong>Abel</strong>, Paris, 1829/01/16. In ibid., 87–89)<br />

1832/04/11 (Löwenhielm→Hansteen, Paris, 1832/04/11. In N. H. <strong>Abel</strong>, 1902b, 101–102)<br />

1824/08/02 (Schumacher→Hansteen, Altona, 1824/08/02. In ibid., 97)<br />

1829/04/07 (Smith→Holmboe, Froland, 1829/04/07. In ibid., 97)<br />

1881/12/09 (Weierstrass→Lie, Berlin, 1881/12/09. In ibid., 103)<br />

1882/04/10 (Weierstrass→Lie, Berlin, 1882/04/10. In ibid., 103–104)<br />

1873 (Weierstrass→du Bois-Reymond, 1873. In K. Weierstrass, 1923, 199–201)<br />

405


Appendix B<br />

ABEL’s manuscripts<br />

Manuscripts and drafts constitute an important source for historical inquiry, especially<br />

when the historian aims at discussing the genesis <strong>of</strong> certain ideas. In the case <strong>of</strong> ABEL,<br />

the extent sources not published within or immediately after his lifetime fall into two<br />

categories:<br />

1. Manuscripts — to various degrees <strong>of</strong> completion — <strong>of</strong> papers not published in<br />

ABEL’s lifetime.<br />

2. Drafts and notebooks documenting the working mathematician but not intended<br />

for publication.<br />

Items in the first category was to some extent considered and included in the com-<br />

pilations for both editions <strong>of</strong> ABEL’s collected works (N. H. <strong>Abel</strong>, 1839; N. H. <strong>Abel</strong>,<br />

1881). In the second volume <strong>of</strong> the second edition (1881) collected works, SYLOW also<br />

included a general presentation <strong>of</strong> the known items from the second category 1 .<br />

<strong>The</strong> present appendix has as its aim to document the whereabouts <strong>of</strong> archival mate-<br />

rial concerning ABEL. This aim is achieved by reproducing registrants <strong>of</strong> the archives<br />

at the Manuscript Collection, University <strong>of</strong> Oslo (see table B.1) and the Mittag-Leffler<br />

Institute, Djursholm, Sweden (table B.5).<br />

1 (N. H. <strong>Abel</strong>, 1881, vol. 2, 283–289)<br />

2 <strong>The</strong> author gratefully acknowledges the participation <strong>of</strong> KLAUS FROVIN JØRGENSEN in obtaining<br />

the information presented in table B.5.<br />

407


408 Appendix B. ABEL’s manuscripts<br />

MS:592 Manuscripts and letters (table B.2)<br />

MS:434 Fragments from manuscripts (table B.3)<br />

MS:435 Note accompaning the <strong>Abel</strong> manuscripts<br />

(fragmentary, 6 pages)<br />

MS:969-4 Nogle Bemærkninger om Vinkelfunktionerne<br />

(fragmentary, 3 pages)<br />

MS:589 Belonging to the Théorie de la résolution<br />

algébrique des équations (fragmentary, 2<br />

pages)<br />

MS:592 Précis de la théorie des fonctions elliptiques<br />

(168 pages)<br />

MS:920-4 <strong>Niels</strong> <strong>Abel</strong> Berlin-Paris 1825–182? (11<br />

pages)<br />

MS:188-8 Note über die Function . . . (2 pages)<br />

MS:969-4 Paa Froland og ved <strong>Abel</strong>s grav 4. og 5. august<br />

(4 pages)<br />

Multiple ABEL’s notebooks (table B.4)<br />

Table B.1: <strong>Abel</strong> manuscript collections in the Manuscript Collection, University Library,<br />

Oslo.<br />

1. Mémoire sur une classe particulière d’équations résolubles<br />

algébriquement (64 pages)<br />

2. Note sur quelques formules elliptiques (18 pages)<br />

3. Théorèmes sur les fonctions elliptiques (11 pages)<br />

4. Démonstration d’une propriété générale d’une certaine<br />

classe de fonctions transcendantes (4 pages)<br />

5. Matematiske uddrag fra N. H. <strong>Abel</strong>’s breve (13 pages)<br />

Table B.2: <strong>Abel</strong> manuscripts in the Manuscript Collection, University Library, Oslo,<br />

MS:592.


1. Précis d’une théorie des fonctions elliptiques (fragmentary,<br />

8 pages)<br />

2. Om transformationer af elliptiske funktioner, hvorved<br />

de to perioder divideres med hvert sit tal (fragmentary,<br />

9 pages)<br />

3. Divisionsligningers opløsning (fragmentary, 2 pages)<br />

4. Unidentified (fragmentary, 1 page)<br />

5. Belonging to the Précis or the Recherches (fragmentary,<br />

2 pages)<br />

6. Belonging to the Paris mémoire (fragmentary, 2 pages)<br />

7. Belonging to the Précis (fragmentary, 7 pages)<br />

8. Belonging to the Recherches second part (fragmentary,<br />

2 pages)<br />

9. Table <strong>of</strong> contents for a work on elliptic functions (fragmentary,<br />

4 pages)<br />

10. Integration ved hjælp af algebraiske, logaritmiske og<br />

eksponentielle funktioner (fragmentary, 2 pages)<br />

11. Belonging to the <strong>Abel</strong>ian <strong>The</strong>orem for elliptic functions<br />

(fragmentary, 2 pages)<br />

Table B.3: <strong>Abel</strong> manuscripts in the Manuscript Collection, University Library, Oslo,<br />

MS:434.<br />

MS:351:A Notebook A Mémoires de Mathématiques par N.<br />

H. <strong>Abel</strong>. Paris le 9 Août 1826 (202<br />

pages fol.)<br />

MS:436 Notebook B Without title (178 pages fol.)<br />

MS:351:C Notebook C Without title (215 pages fol.)<br />

MS:696 Notebook D Remarques sur divers points de<br />

l’analyse par N. H. <strong>Abel</strong>, 1er Cahier<br />

le 3 Sept. 1827 (136 pages 4:o)<br />

MS:829 Notebook E Mathematiske Udarbeidelser af <strong>Niels</strong><br />

<strong>Henrik</strong> <strong>Abel</strong> (192 pages 4:o)<br />

MS:749 Matematiske Afhandlinger (170<br />

pages 4:o)<br />

Table B.4: <strong>Abel</strong>’s mathematical notebooks in the Manuscript Collection, University<br />

Library, Oslo.<br />

409


410 Appendix B. ABEL’s manuscripts<br />

1. Manuskript af <strong>Abel</strong> (V.Terquem, Bulletin, T. I, p. 56)<br />

2. Manuskript i 4:o: Note sur quelques formules elliptiques<br />

3. Manuskript i 8:o: Théories sur les fonctions elliptiques<br />

4. P.M. af Phragmén<br />

5. Acta korrektur af Recherches sur les fonctions elliptiques<br />

6. Tryckt: Mémoire sur une propriété générale d’une<br />

classe très-étendue de fonctions transcendantes. 1841.<br />

4:o.<br />

7. Mémoire sur les équations algébriques etc. Chr:ia<br />

1824. 4:o.<br />

8. Oplösning af et Par Opgaver ved Hjelp af bestemte<br />

Integraler. 8:o<br />

9. Almindelig methode til at finde Funktioner af een<br />

variable Störrelse. . . 8:o.<br />

Table B.5: <strong>Abel</strong> manuscripts in the Mittag-Leffler Institute, Djursholm, Sweden. 2


Bibliography<br />

<strong>Abel</strong> (MS:351:A). “Mémoires de Mathématiques par N. H. <strong>Abel</strong>. Paris le 9 Août 1826<br />

(Notebook A)”. Håndskriftsamlingen, Oslo MS:351:A (202 pages fol.). Dated 1826.<br />

— (MS:351:C). “Notebook C”. Håndskriftsamlingen, Oslo MS:351:C (215 pages fol.).<br />

Dated 1828.<br />

— (MS:592). “Mémoire sur une classe particulière d’équations résolubles algébriquement”.<br />

Håndskriftsamlingen, Oslo MS:592 (64 pages).<br />

— (MS:696). “Remarques sur divers points de l’analyse par N. H. <strong>Abel</strong>, 1er Cahier<br />

le 3 Sept. 1827 (Notebook D)”. Håndskriftsamlingen, Oslo MS:696 (136 pages 4o).<br />

Dated 1827.<br />

— (MS:829). “Mathematiske Udarbeidelser af <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> (Notebook E)”.<br />

Håndskriftsamlingen, Oslo MS:829 (192 pages 4o). Dated 1820.<br />

<strong>Abel</strong>, N. H. (1823). “Oplösning af et Par Opgaver ved Hjelp af bestemte Integraler”. in:<br />

Magazin for Naturvidenskaberne 1.1–2. Reproduced in (ibid., I, 11–27), 55–68, 205–15.<br />

— (1824a). “Berigtigelse”. in: Magazin for Naturvidenskaberne 2, 143–144.<br />

— (1824b). “Mémoire sur les équations algébriques, on l’on démontre l’impossibilité<br />

de la résolution de l’équation générale du cinquième degré”. in: Oeuvres Complètes<br />

de <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow and S. Lie. 2. vol. 1. 2 vols. First published<br />

Christiania: Grøndahl. Christiania: Grøndahl, 28–33.<br />

— (1824c). “Om Maanens Indflydelse paa Pendelens Bevægelse”. in: Magazin for<br />

Naturvidenskaberne 1, 219–226.<br />

— ([1825] 1839a). “Propriétés remarquables de la fonction y = φx déterminée<br />

par l’équation f y dy − dx � (a − y)(a1 − y)(a2 − y) . . . (am − y) = 0, f y étant une<br />

fonction quelconque de y qui ne devient pas nulle ou infinie lorsque y =<br />

a, a1, a2, . . . am”. in: Oeuvres Complètes de <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow and S.<br />

Lie. 2. vol. 2. 2 vols. Undated manuscript, probably written prior to <strong>Abel</strong>’s European<br />

tour (Holmboe), i.e. before 1825. First published (N. H. <strong>Abel</strong>, 1839, vol. II,<br />

pp. 51–53). Christiania: Grøndahl, 40–42.<br />

— ([1825] 1839b). “Théorie des transcendantes elliptiques”. in: Oeuvres Complètes de<br />

<strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow and S. Lie. 2. vol. 2. 2 vols. Undated manuscript,<br />

probably written prior to <strong>Abel</strong>’s European tour (Holmboe), i.e. before 1825. First<br />

published (ibid., vol. II, pp. 93–184). Christiania: Grøndahl, 87–188.<br />

— ([1826] 1841). “Mémoire sur une propriété générale d’une classe très étendue de<br />

fonctions transcendantes”. in: Oeuvres Complètes de <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow<br />

and S. Lie. 2. vol. 1. 2 vols. First published Mémoires présentés par divers savants,<br />

t. VII, 1841. Delivered to l’Académie des sciences October 26, 1826. Christiania: Grøndahl,<br />

145–211.<br />

— (1826a). “Beweis der Unmöglichkeit, algebraische Gleichungen von höheren<br />

Graden, als dem vierten, allgemein aufzulösen”. in: Journal für die reine und angewandte<br />

Mathematik 1.1. Reproduced in (N. H. <strong>Abel</strong>, 1881, I, 66–87), 65–84.<br />

411


412 Bibliography<br />

<strong>Abel</strong>, N. H. (1826b). “Beweis eines Ausdruckes, von welchem die Binomial-Formel<br />

ein einzelner Fall ist”. in: Journal für die reine und angewandte Mathematik 1.2. Reproduced<br />

in (ibid., I, 102–103), 159–160.<br />

— (1826c). “Démonstration de l’impossibilité de la résolution des équations algébriques<br />

générales d’un degré supérieur au quatrième; par M. <strong>Abel</strong>. (Journ. der<br />

Mathemat., de M. Crelle; t. I, p. 65.)” in: Bulletin des sciences mathématiques, as-<br />

tronomiques, physiques et chimiques 6, 347–354.<br />

— (1826d). “Ueber die Integration der Differential Formel<br />

ρ dx<br />

√ R , wenn R und ρ ganze<br />

Functionen sind”. in: Journal für die reine und angewandte Mathematik 1.3. Reproduced<br />

in (ibid., I, 104–144), 185–221.<br />

— (1826e). “Untersuchung der Functionen zweier unabhängig veränderlicher Größen<br />

x und y, wie f (x, y), welche die Eigenschaft haben, daß f (z, f (x, y)) eine symmetrische<br />

Function von z, x und y ist”. in: Journal für die reine und angewandte Mathematik<br />

1.1. Reproduced in (ibid., I, 61–65), 11–15.<br />

m·(m−1)<br />

x + 2 x2 +<br />

x3 . . . u.s.w”. in: Journal für die reine und angewandte Mathematik<br />

— (1826f). “Untersuchungen über die Reihe 1 + m m·(m−1)(m−2)<br />

2·3<br />

1<br />

1.4. Reproduced in (ibid., I, 219–250), 311–339.<br />

— ([1827] 1881). “Sur les séries”. in: Oeuvres Complètes de <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L.<br />

Sylow and S. Lie. 2. vol. 2. 2 vols. Undated manuscript, probably written in the<br />

second half <strong>of</strong> 1827 (Sylow and Lie). Christiania: Grøndahl, 197–205.<br />

— (1827a). “Recherche de la quantité qui satisfait a la fois a deux équations algébriques<br />

données”. in: Oeuvres Complètes de <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow and<br />

S. Lie. 2. vol. 1. 2 vols. First published Annales de Mathématiques pures et appliquées,<br />

t. XVII, Paris 1827. Christiania: Grøndahl. chap. 13, 212–218.<br />

— (1827b). “Recherches sur les fonctions elliptiques”. in: Journal für die reine und angewandte<br />

Mathematik 2.2. Reproduced in (ibid., I, 263–388), 101–181.<br />

— (1827c). “Ueber die Functionen, welche der Gleichung φx + φy = ψ (x f y + y f x)<br />

genugthun”. in: Journal für die reine und angewandte Mathematik 2.4. Reproduced in<br />

(ibid., I, 389–398), 386–394.<br />

— ([1828] 1839). “Sur la résolution algébrique des équations”. in: Oeuvres Complètes de<br />

<strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow and S. Lie. 2. vol. 2. 2 vols. Undated manuscript,<br />

probably written in the second half <strong>of</strong> 1828 (Sylow and Lie). First published (N. H.<br />

<strong>Abel</strong>, 1839, vol. II, pp. 185–209). Christiania: Grøndahl, 217–243.<br />

— (1828a). “Note sur le mémoire de Mr. L. Olivier No. 4. du second tome de ce journal,<br />

ayant pour titre “Remarques sur les séries infinies et leur convergence””. in:<br />

Journal für die reine und angewandte Mathematik 3.1. Reproduced in (N. H. <strong>Abel</strong>, 1881,<br />

I, 399–402), 79–82.<br />

— (1828b). “Recherches sur les fonctions elliptiques”. in: Journal für die reine und angewandte<br />

Mathematik 3.2. Reproduced in (ibid., I, 263–388), 160–190.<br />

— (1828c). “Remarques sur quelques propriétés générales d’une certaine sorte de<br />

fonctions transcendantes”. in: Journal für die reine und angewandte Mathematik 3.4.<br />

Reproduced in (ibid., I, 444-456), 313–323.<br />

— (1828d). “Solution d’un problème général concernant la transformation des fonctions<br />

elliptiques”. in: Astronomische Nachrichten 6.138. Reproduced in (ibid., I, 403–<br />

428), 365–388.<br />

— (1828e). “Sur le nombre des transformations différentes, qu’on peut faire subir a<br />

une fonction elliptique par la substitution d’une fonction donnée e premier degré”.


Bibliography 413<br />

in: Journal für die reine und angewandte Mathematik 3.4. Reproduced in (ibid., I, 457–<br />

465), 394–401.<br />

— (1829a). “Addition au mémoire sur les fonctions elliptiques, inséré dans le Nr. 138<br />

de ce journal”. in: Astronomische Nachrichten 7.147. Reproduced in (ibid., I, 429–443),<br />

33–44.<br />

— (1829b). “Démonstration d’une propriété générale d’une certaine classe de fonctions<br />

transcendantes”. in: Journal für die reine und angewandte Mathematik 4.2. Reproduced<br />

in (ibid., I, 515–517), 200–201.<br />

— (1829c). “Mémoire sur une classe particulière d’équations résolubles algébriquement”.<br />

in: Journal für die reine und angewandte Mathematik 4.2. Reproduced in (ibid.,<br />

I, 478–507), 131–156.<br />

— (1829d). “Précis d’une théorie des fonctions elliptiques”. in: Journal für die reine und<br />

angewandte Mathematik 4.4. Reproduced in (ibid., I, 518–617), 236–277, 309–348.<br />

— (1839). Oeuvres Complètes de N. H. <strong>Abel</strong>, mathématicien, avec des notes et développements.<br />

ed. by B. M. Holmboe. 1. 2 vols. Christiania: Chr. Gröndahl.<br />

— (1881). Oeuvres Complètes de <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. ed. by L. Sylow and S. Lie. 2. 2 vols.<br />

Christiania: Grøndahl.<br />

— (1902a). “Breve fra og til <strong>Abel</strong>”. in: Festskrift ved Hundredeaarsjubilæet for <strong>Niels</strong> <strong>Henrik</strong><br />

<strong>Abel</strong>s Fødsel. ed. by E. Holst, C. Størmer, and L. Sylow. Kristiania: Jacob Dybwad.<br />

— (1902b). “Breve om <strong>Abel</strong>”. in: Festskrift ved Hundredeaarsjubilæet for <strong>Niels</strong> <strong>Henrik</strong><br />

<strong>Abel</strong>s Fødsel. ed. by E. Holst, C. Størmer, and L. Sylow. Kristiania: Jacob Dybwad.<br />

— (1902c). “Recherches sur les fonctions elliptiques (Second mémoire)”. in: Acta Mathematica<br />

26, 3–42.<br />

<strong>Abel</strong>, N. H. and E. Galois (1889). Abhandlungen über die Algebraische Auflösung der Gleichungen.<br />

ed. by H. Maser and H. Maser. Berlin: Verlag von Julius Springer.<br />

<strong>Abel</strong>, N. H. (1902d). “Dokumenter angaaende <strong>Abel</strong>”. in: Festskrift ved Hundredeaarsjubilæet<br />

for <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>s Fødsel. ed. by E. Holst, C. Størmer, and L. Sylow. Kristiania:<br />

Jacob Dybwad.<br />

— (1902e). Festskrift ved Hundredeaarsjubilæet for <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>s Fødsel. ed. by E.<br />

Holst, C. Størmer, and L. Sylow. Kristiania: Jacob Dybwad.<br />

— (1902f). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. Memorial publie a l’occasion du centenaire de sa naissance.<br />

ed. by E. Holst et al. Kristiania: Jacob Dybwad etc.<br />

Andersen, E. (1975). Heinrich Christian Schumacher. Et mindeskrift. København:<br />

Geodætisk Instituts Forlag.<br />

Andersen, K. (1999). “Wessel’s work on complex numbers and its place in history”. in:<br />

Caspar Wessel. On the Analytic Representation <strong>of</strong> Direction. An Attempt Applied Chiefly<br />

to Solving Plane and Spherical Polygons. ed. by B. Branner and J. Lützen. vol. 46:1.<br />

Matematisk-fysiske Meddelelser, Det Kongelige Danske Videnskabernes Selskab.<br />

Copenhagen: C. A. Reitzel, 65–98.<br />

Anderssen, A. E. G. (1848). “Beleuchtung der wesentlichsten Argumente des<br />

<strong>Abel</strong>’schen Beweises der Unmöglichkeit algebraische Gleichungen von höhern<br />

Graden als dem vierten aufzulösen”. in: Zu der am 13ten, 14ten und 15ten April 1848<br />

stattfindenden Prüfung der Schüler des Königlichen Friedrichs-Gymnasiums. ed. by F.<br />

Wimmer. Breslau: Graß, Barth und Comp., 1–21.<br />

Anonymous (1826). “Beweis der Unmöglichkeit eine vollständige algebraische Gleichung<br />

mit einer unbekannten Grösse, deren Grad den vierten übersteigt, durch<br />

eine geschlossene algebraische Formel aufzulösen”. in: Zeitschrift für Physik und<br />

Mathematik. Mathematische Abteilung 1.2, 253–262.


414 Bibliography<br />

Ayoub, R. G. (1980). “Paolo Ruffini’s Contributions to the Quintic”. in: Archive for History<br />

<strong>of</strong> Exact Sciences 23, 253–277.<br />

Ayoub, R. (1982). “On the nonsolvability <strong>of</strong> the general polynomial”. in: <strong>The</strong> American<br />

Mathematical Monthly 89, 397–401.<br />

— (1984). “<strong>The</strong> Lemniscate and Fagnano’s Contributions to Elliptic Integrals”. in:<br />

Archive for History <strong>of</strong> Exact Sciences 29.2, 131–149.<br />

Begehr, H. G. W. et al., eds. (1998). <strong>Mathematics</strong> in Berlin. Berlin, Basel, Boston:<br />

Birkhäuser Verlag.<br />

Bell, E. T. (1953). “Genius and poverty: <strong>Abel</strong> (1802–29)”. in: Men <strong>of</strong> <strong>Mathematics</strong>. First<br />

published 1937. Melbourne, London, Baltimore: Penguin Books.<br />

Bertrand, J. (1842). “Règles sur la convergence des séries”. in: Journal de Mathématiques<br />

pures et appliquées 7. Serie I, 35–54.<br />

Biermann, K.-R. (1966). “Karl Weierstraß. Ausgewählte Aspekte seiner Biographie”.<br />

in: Journal für die reine und angewandte Mathematik 223, 191–220.<br />

— (1967). “Ein unbekanntes Schreiben von N. H. <strong>Abel</strong> an A. L. Crelle”. in: Nordisk<br />

Matematisk Tidskrift 15, 25–32.<br />

— (1981). “Historische Einführung”. in: Mathematisches Tagebuch 1796–1814. Ostwalds<br />

Klassiker der exakten Wissenschaften 256. Leipzig: Akademische Verlagsgesellschaft,<br />

Geest & Portig K.-G., 7–20.<br />

— (1988). Die Mathematik und ihre Dozenten an der Berliner Universität 1810–1933. Stationen<br />

auf dem Wege eines mathematischen Zentrums von Weltgeltung. Berlin: Akademie-<br />

Verlag Berlin.<br />

Biermann, K.-R. and V. Brun (1958). “Eine Notiz N. H. <strong>Abel</strong>s für A. L. Crelle auf einem<br />

Manuskript Otto Auberts”. in: Nordisk Matematisk Tidskrift 6, 84–86.<br />

Bjerknes, C. A. (1880). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. En skildring af hans liv og videnskabelige virksomhed.<br />

Published as an appendix to Nordisk tidsskrift för vetenskap, konst och industri.<br />

Stockholm: P. A. Norstedt & Söner.<br />

— (1885). <strong>Niels</strong>-<strong>Henrik</strong> <strong>Abel</strong>. Tableau de sa vie et de son action scientifique. Paris: Gauthier-<br />

Villars.<br />

— (1930). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. Eine Schilderung seines Lebens und seiner Arbeit. ed. by V.<br />

Bjerknes. Berlin: Verlag von Julius Springer.<br />

Bois-Reymond, P. du (1871). “Notiz über einen Cauchy’schen Satz, die Stetigkeit von<br />

Summen unendlicher Reihen betreffend”. in: Mathematische Annalen 4, 135–137.<br />

Bolzano, B. (1816). Der binomische Lehrsatz, und als Folgerung aus ihm der polynomische,<br />

und die Reihen, die zur Berechnung der Logarithmen und Exponentialgrößen dienen,<br />

genauer als bisher erwiesen. Prag: C. B. Endersches Buchhandlung.<br />

— (1817). Rein analytischer Beweis des Lehrsatzes, daß zwischen je zwey Werthen, die ein<br />

entgegengesetztes Resultat gewähren, wenigstens eine reelle Wurzel der Gleichung liege.<br />

Prag: Gottlieb Haase.<br />

Bos, H. J. M. (1974). “<strong>The</strong> Lemniscate <strong>of</strong> Bernoulli”. in: For Dirk Struik. Scientific, Historical<br />

and Political Essays in Honor <strong>of</strong> Dirk J. Struik. ed. by R. S. Cohen, J. J. Stachel,<br />

and M. W. Wart<strong>of</strong>sky. Boston Studies in the Philosophy <strong>of</strong> Science 15. Dordrecht:<br />

D. Reidel Publishing Company, 3–14.<br />

Bos, H. J. M. (2001). Redefining Geometrical Exactness. Descartes’ Transformation <strong>of</strong> the<br />

Early Modern Concept <strong>of</strong> Construction. Sources and Studies in the History <strong>of</strong> <strong>Mathematics</strong><br />

and Physical Sciences. New York etc.: Springer.<br />

Bottazzini, U. (1986). <strong>The</strong> Higher Calculus: A History <strong>of</strong> Real and Complex Analysis from<br />

Euler to Weierstrass. New York: Springer-Verlag.


Bibliography 415<br />

— (1990). “Geometrical Rigour and ’Modern Analysis’. An Introduction to Cauchy’s<br />

Cours d’Analyse”. in: Augustin-Louis Cauchy: Cours d’Analyse de l’École Royale Polytechnique.<br />

Première partie. Analyse Algébrique. Instrumenta Rationis. Sources for the<br />

History <strong>of</strong> Logic in the Modern Age VII. Editrice CLUEB Bologna, xi–clxvii.<br />

Brill, A. and M. Noether (1894). “Die Entwicklung der <strong>The</strong>orie der algebraischen<br />

Functionen in ältere und neuere Zeit”. in: Jahresbericht der Deutschen Mathematiker-<br />

Vereinigung 3, 107–566.<br />

Briot and Bouquet (1853a). “Note sur le développement des fonctions en séries convergentes<br />

ordonnées suivant les puissances croissantes de la variable”. in: Comptes<br />

rendus hebdomadaires des séances de l’Académie des Sciences 36. Dated 21 February,<br />

334–335.<br />

— (1853b). “Recherches sur les séries ordonnées suivant les puissances croissantes<br />

d’une variable imaginaire”. in: Comptes rendus hebdomadaires des séances de<br />

l’Académie des Sciences 36. Dated 7 February, 264.<br />

Brun, V. (1949). “Det tapte <strong>Abel</strong>manuskript”. in: Norsk matematisk tidsskrift 31, 56–58.<br />

— (1953). “Det gjenfunne manuskript til <strong>Abel</strong>s Parisavhandling”. in: Nordisk Matematisk<br />

Tidskrift 1, 91–97.<br />

— (1962). Regnekunsten i det gamle Norge fra Arilds tid til <strong>Abel</strong>. Oslo, Bergen: Universitetsforlaget.<br />

Brun, V. and B. Jessen (1958). “Et ungdomsbrev fra <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>”. in: Nordisk<br />

Matematisk Tidskrift 6, 21–24.<br />

Burkhardt, H. (1892). “Die Anfänge der Gruppentheorie und Paolo Ruffini”. in:<br />

Historisch-literarische Abteilung der Zeitschrift für Mathematik und Physik 38, 119–159<br />

suppl.<br />

Cauchy, A.-L. (1815a). “Mémoire sur le nombre des valeurs qu’une fonction peut acquérir,<br />

lorsqu’on y permute de toutes les manières possibles les quantités qu’elle<br />

renferme”. in: Œuvres Complètes d’Augustin Cauchy. vol. II.1. 12+15 vols. second series.<br />

First published Journal de l’École Polytechnique, 10(17), pp. 1–28, 1815. Paris:<br />

Gauthier-Villars, 64–90.<br />

— (1815b). “Mémoire sur les fonctions qui ne peuvent obtenir que deux égales ou<br />

de signes contraires par suite de transpositions opérées entre les variables qu’elles<br />

renferment”. in: Œuvres Complètes d’Augustin Cauchy. vol. II.1. 12+15 vols. 2nd series.<br />

First published Journal de l’École Polytechnique, 10(17), pp. 29–112, 1815. Paris:<br />

Gauthier-Villars, 90–169.<br />

— (1821a). Cours d’Analyse de l’École Royale Polytechnique. Première partie. Analyse Algébrique.<br />

Reproduced in (A.-L. Cauchy, 1990, ). Paris: L’Imprimerie Royale.<br />

— (1823). “Résumé des leçons données a l’École Royale Polytechnique sur le calcul<br />

infinitésimal”. in: Œuvres Complètes d’Augustin Cauchy. vol. II.4. 12+15 vols. 2nd<br />

series. First published Paris: de l’imprimérie royale; 1823. Paris: Gauthier-Villars,<br />

9–261.<br />

— (1826). “Sur un nouveau genre de calcul analogue au calcul infinitésimal”. in: Œuvres<br />

Complètes d’Augustin Cauchy. vol. 6 (2nd series). 12+15 vols. First published<br />

Exercises de mathématiques, 1826. Paris: Gauthier-Villars, 23–37.<br />

— (1828). Lehrbuch der algebraischen Analysis. Königsberg: Bornträger.<br />

— (1990). Augustin-Louis Cauchy: Cours d’Analyse de l’École Royale Polytechnique. Première<br />

partie. Analyse Algébrique. Instrumenta Rationis. Sources for the History <strong>of</strong><br />

Logic in the Modern Age VII. Editrice CLUEB Bologna.<br />

Cauchy, A.-L. (1821b). Cours d’analyse de l’École Royale Polytechnique. Premier partie.<br />

Analyse algébrique. Paris: l’Imprimerie Royale.


416 Bibliography<br />

Cauchy, A.-L. (1822). “Sur le développement des fonctions en séries et sur l’intégration<br />

des équations différentielles ou aux différences partielles”. in: Œuvres Complètes<br />

d’Augustin Cauchy. vol. II.2. 12+15 vols. 2nd series. First published Bulletin de la<br />

Société Philomatique, p. 49–54; 1822. Paris: Gauthier-Villars, 276–282.<br />

— (1825). “Mémoire sur les intégrales définies, prises entre des limites imaginaires”.<br />

in: Œuvres Complètes d’Augustin Cauchy. vol. II.15. 12+15 vols. 2nd series. First published<br />

Paris: Bure, August 1825. Paris: Gauthier-Villars, 41–89.<br />

— (1827). “Mémoire sur les développements des fonctions en séries périodiques”.<br />

in: Œuvres Complètes d’Augustin Cauchy. vol. 2. 12+15 vols. 1st series. Read<br />

to l’Académie royale des Sciences February 27, 1826. First published Mémoires de<br />

l’Académie des Sciences, t. VI, p. 603, 1827. Paris: Gauthier-Villars, 12–19.<br />

— (1829). “Leçons sur le calcul différentielle”. in: Œuvres Complètes d’Augustin Cauchy.<br />

vol. II.4. 12+15 vols. 2nd series. First published Paris: de Bure Frères; 1829. Paris:<br />

Gauthier-Villars, 263–609.<br />

— (1853). “Note sur les séries convergentes dont les divers termes sont des fonctions<br />

continues d’une variable réelle ou imaginaire, entre des limites donné”. in:<br />

Œuvres Complètes d’Augustin Cauchy. vol. I.12. 12+15 vols. 1st series. Read to the<br />

Académie des Sciences on March 14, 1853 and originally published Comptes Rendus<br />

vol. XXXVI p. 454. Paris: Gauthier-Villars, 30–36.<br />

Cayley, A. (1861). “Note on Mr. Jerrard’s researches on the equation <strong>of</strong> the fifth degree”.<br />

in: <strong>The</strong> Collected Mathematical Papers. vol. 5. 13 vols. First published Phil. Mag.,<br />

vol. XXI (1861), pp. 210–214. Cambridge: at the University Press. chap. 310, 50–54.<br />

Chebyshev, P. (1853). “Sur l’intégration des différentielles irrationnelles”. in: Journal de<br />

mathématiques pures et appliquées 18, 87–111.<br />

Cleave, J. P. (1971). “Cauchy, Convergence and Continuity”. in: British Journal for the<br />

Philosophy <strong>of</strong> Science 22, 27–37.<br />

Cockle, J. (1862). “Note on the Remarks <strong>of</strong> Mr. Jerrard”. in: <strong>The</strong> London, Edinburgh, and<br />

Dublin Philosophical Magazine, and Journal <strong>of</strong> Science 23. 4th series, 196–198.<br />

— (1863). “Concluding Remarks on a recent Mathematical Controversy”. in: <strong>The</strong> London,<br />

Edinburgh, and Dublin Philosophical Magazine, and Journal <strong>of</strong> Science 26. 4th series,<br />

223–224.<br />

Cooke, R. (1989). “<strong>Abel</strong>’s <strong>The</strong>orem”. in: <strong>The</strong> History <strong>of</strong> Modern <strong>Mathematics</strong>. ed. by D. E.<br />

Rowe, J. McCleary, and E. Knobloch. vol. 1. Proceedings <strong>of</strong> the Symposium on the<br />

History <strong>of</strong> Modern <strong>Mathematics</strong>, Vassar College, Poughkeepsie, New York, June 20–24,<br />

1989. 3 volumes. Academic Press, 389–421.<br />

Craik, A. D. D. (1999). “Calculus and Analysis in Early 19th-Century Britain: <strong>The</strong> Work<br />

<strong>of</strong> William Wallace”. in: Historia Mathematica 26, 239–267.<br />

Crelle (1827). “Einige Nachrichten von Büchern”. in: Journal für die reine und angewandte<br />

Mathematik 2, 399–400.<br />

— (1828). “Einige Nachrichten von Büchern”. in: Journal für die reine und angewandte<br />

Mathematik 3, 410–412.<br />

Crelle, A. L. (1826). “Vorrede”. in: Journal für die reine und angewandte Mathematik 1.1,<br />

1–4.<br />

— (1827). “Einige Nachrichten von Büchern”. in: Journal für die reine und angewandte<br />

Mathematik 2.4, 399–400.<br />

— (1829a). “Démonstration nouvelle du théorème du binome”. in: Journal für die reine<br />

und angewandte Mathematik 4.3, 305–308.<br />

— (1829b). “Nécrologe”. in: Journal für die reine und angewandte Mathematik 4.4, 402–<br />

404.


Bibliography 417<br />

— (1830). “Mémoire sur la convergence de la série du binome; pour faire suite à la<br />

démonstration du théorème du binôme, donnée tome III. de ce journal, cahier 3.,<br />

page 305”. in: Journal für die reine und angewandte Mathematik 5.2, 187–196.<br />

Dahan, A. (1980). “Les Travaux de Cauchy sur les Substitutions. Étude de son approche<br />

du concept de groupe”. in: Archive for History <strong>of</strong> Exact Sciences 23, 279–319.<br />

Dauben, J. W. (1981). “<strong>Mathematics</strong> in Germany and France in the Early 19th Century:<br />

Transmission and Transformation”. in: Epistemological and Social Problems in the Sciences<br />

in the Early Nineteenth Century. ed. by H. N. Jahnke and M. Otte. Dordrecht:<br />

D. Reidel Publishing Company, 371–399.<br />

Degen, C. F. (1799). “Et Bidrag til Æqvationernes <strong>The</strong>orie”. in: Nye Samling af det Kongelige<br />

Danske Videnskabernes Selskabs Skrifter 5, 572–582.<br />

Degen, C. F. (1817). Canon Pellianus sive Tabula simplicissimam æqvationis celebratissimæ<br />

y = ax + 1 solutionem, pro singulis numeri dati valoribus ab 1 usque ad 1000, in numeris<br />

rationalibus iisdemqve integris exhibens. Hafn.<br />

Del Centina, A. (2002). “<strong>The</strong> Manuscript <strong>of</strong> <strong>Abel</strong>’s Parisian Memoire Found in Its Entirety”.<br />

in: Historia Mathematica 29.1, 65–69.<br />

— (Feb. 2003). “Corrigendum to: “<strong>The</strong> manuscript <strong>of</strong> <strong>Abel</strong>’s Parisian memoir found in<br />

its entirety” [Historia Mathematica 29 (2002) 65–69]”. in: Historia Mathematica 30.1,<br />

94–95.<br />

Delambre, J. B. J. (1810). Rapport historique sur les progres des sciences mathématiques<br />

depuis 1789. Unchanged reprint <strong>of</strong> the original edition, Amsterdam: B. M. Israël,<br />

1966. Paris: l’Imprimerie Impériale.<br />

Descartes, R. (1637). La géométrie. Facsimile in [?]. Leiden: Ian Maire.<br />

Dhombres, J. (1985). “French Mathematical Textbooks from Bézout to Cauchy”. in:<br />

Historia Scientiarum 28, 91–137.<br />

Dhombres, J. and M. Pensivy (1988). “Esprit de rigueur et présentation mathématique<br />

au XVIII ème siècle: le cas d’une démonstration d’Aepinus”. in: Historia Mathematica<br />

15, 9–31.<br />

Dhombres, J. (Dec. 1986). “Mathématisation et communauté scientifique française<br />

(1775–1825)”. in: Archives internationales d’histoire des sciences 36.117, 249–287.<br />

— (1992). “Le rôle des équations fonctionelles dans l’Analyse algébrique de Cauchy”.<br />

in: Revue d’Histoire des Sciences 45.1. Études sur Cauchy (1789–1857), 25–49. ISSN:<br />

0151-4105.<br />

Dickson, L. E. (1959). Algebraic <strong>The</strong>ories. Former editions were entitled Modern Algebraic<br />

<strong>The</strong>ories. New York: Dover Publications.<br />

Dirichlet, G. L. (1829). “Sur la convergence des séries trigonométriques qui servent<br />

à représenter une fonction arbitraire entre des limites données”. in: Journal für die<br />

reine und angewandte Mathematik 4.2. Reproduced in (G. L. Dirichlet, Werke, I, 117–<br />

132), 157–169.<br />

— (1837). “Beweis des Satzes, dass jede unbegrenzte arithmetische Progression, deren<br />

erstes Glied und Differenz ganze Zahlen ohne gemeinschaftlichen Factor sind,<br />

unendlich viele Primzahlen enthält”. in: G. Lejeune Dirichlet’s Werke. vol. 1. First<br />

published Abhandlungen der Königlich Preussischen Akademie der Wissenschaften von<br />

1837, S. 45–81. Berlin: Georg Reimer, 313–342.<br />

Dirichlet, G. L. (1852). “Gedächtnisrede auf Carl Gustav Jacob Jacobi”. in: C. G. J. Jacobi’s<br />

Gesammelte Werke. ed. by C. W. Borchardt, K. T. W. Weierstrass, and A. Clebsch.<br />

vol. 1. 8 vols. Reprinted New York: Chelsea Publishing Company, 1969. Berlin:<br />

G. Reimer, 3–28.


418 Bibliography<br />

Dirichlet, G. L. (1862). “Démonstration d’un théorème d’<strong>Abel</strong>”. in: Journal de mathématiques<br />

pures et appliquées 7. Serie II, 253–255.<br />

— (Werke). G. Lejeune Dirichlet’s Werke. 1889–1897.2 volumes.Edited by L. Kronecker<br />

andL. Fuchs. Berlin: Georg Reimer.<br />

Dugac, P. (1989). “Sur la correspondance de Borel et le théorème de Dirichlet–Heine–<br />

Weierstrass-Borel–Schoenflies–Lebesque”. in: Archives internationales d’histoire des<br />

sciences 39.122, 69–110.<br />

Eccarius, W. (1976). “August Leopold Crelle als Herausgeber wissenschaflicher<br />

Fachzeitschriften”. in: Annals <strong>of</strong> Science 33, 229–261.<br />

Eccarius, W. (1975). “Der Techniker und Mathematiker August Leopold Crelle (1780–<br />

1855) und sein Beitrag zur Förderung und Entwicklung der Mathematik in<br />

Deutschland des 19. Jahrhunderts”. in: NTM-Schriftenr. Gesch. Naturwiss., Technik,<br />

Med. 12.2, 38–49.<br />

Elliot (1876). “Détermination du nombre des intégrales abéliennes de première espèce”.<br />

in: Annales scientifiques de l’École supérieure 2, 399–444.<br />

Eneström, G. (1912–1913). “Kleine Bemerkungen zur letzten Auflage von Cantors<br />

“Vorlesungen über Geschichte der Mathematik””. in: Bibliotheca Mathematica 13.<br />

3rd series, 339–351.<br />

Epple, M. (2000). “Genies, Ideen, Institutionen, mathematische Werkstätten: Formen<br />

der Mathematikgeschichte”. in: Mathematische Semesterberichte 47.2, 131–163.<br />

Euler, L. (1732a). “Specimen de constructione aequationum differentialium sine indeterminatarum<br />

separatione”. in: Leonhardi Euleri opera omnia. ed. by F. Rudio et<br />

al. vol. I:20. ? vols. 1st series. First published Commentarii academiae scientiarum<br />

Petropolitanae 6 (1732/33), 1738, p. 168–174. Leipzig and Bern: B. G. Teubner, 1–7.<br />

Euler, L. (1972). Elements <strong>of</strong> Algebra. trans. by J. Hewlett. New York etc.: Springer-<br />

Verlag.<br />

Euler, L. (1732b). “De formis radicum aequationum cuiusque ordinis coniectatio”. in:<br />

Leonhardi Euleri opera omnia. ed. by F. Rudio et al. vol. I:6. ? vols. 1st series. First<br />

published Commentarii academiae scientiarum Petropolitanae 6 (1732/33), 1738, pp.<br />

216–231. Leipzig and Bern: B. G. Teubner, 1–19.<br />

— (1748). Introductio in analysin infinitorum. vol. 1. Lausanne: Marcum-Michaelem<br />

Bousquet.<br />

— (1754). “Subsidium calculi sinuum”. in: Leonhardi Euleri opera omnia. ed. by F. Rudio<br />

et al. vol. I:14. ? vols. 1st series. First published Novi commentarii academiae scientiarum<br />

Petropolitanae 5 (1754/55), 1760, pp. 164–204. Leipzig and Bern: B. G. Teubner,<br />

542–584.<br />

— (1755). “Institutiones calculi differentialis cum eius usu in analysi finitorum ac doctrina<br />

serierum”. in: Leonhardi Euleri opera omnia. ed. by F. Rudio et al. vol. I:10.<br />

? vols. First series. First published Petropolitanae: Academiae imperialis scien-<br />

tiarum. Leipzig and Bern: B. G. Teubner.<br />

— (1756/57). “De integratione aequationis differentialis m dx<br />

√ (1−x 4 ) =<br />

n dy<br />

√ (1−y 4 ) ”. in:<br />

Leonhardi Euleri opera omnia. ed. by F. Rudio et al. vol. I:20. ? vols. 1st series 251<br />

(Eneström). First published Novi commentarii academiae scientiarum Petropolitanae 6<br />

(1756/7), 1761, p. 37–57. Leipzig and Bern: B. G. Teubner, 58–79.<br />

— (1760). “De seriebus divergentibus”. in: Leonhardi Euleri opera omnia. ed. by F. Rudio<br />

et al. vol. 14. ? vols. First series. First published Novi commentarii academiae scientiarum<br />

Petropolitanae 5 (1764/55), 1760, p. 205–237. Leipzig and Bern: B. G. Teubner,<br />

585–617.


Bibliography 419<br />

— (1768). “Institutionum calculi integralis”. in: Leonhardi Euleri opera omnia. ed. by<br />

F. Rudio et al. vol. I:11–13. ? vols. 1st series. First published Petropoli: Impensis<br />

Academiae Imperialis Scientiarum. Leipzig and Bern: B. G. Teubner.<br />

— (1775). “Demonstratio theorematis Neutoniani de evolutione potestatum binomii<br />

pro casibus quibus exponentes non sunt numeri integri”. in: Leonhardi Euleri opera<br />

omnia. ed. by F. Rudio et al. vol. I:15. ? vols. First series. First published Novi commentarii<br />

academiae scientiarum Petropolitanae 19 (1774), 1775, p. 103–111. Leipzig and<br />

Bern: B. G. Teubner, 207–216.<br />

— (1788–1791). Leonhard Eulers Einleitung in die Analysis des Unendlichen. ed. by J. A. C.<br />

Michelsen. 3 vols. Berlin: Carl Maßdorff.<br />

Euler, L. and C. Goldbach (1965). Briefwechsel 1729–1764. Abhandlungen der<br />

Deutschen Akademie der Wissenschaften zu Berlin 1. Edited by A. P. Juškevič and<br />

E. Winter. Berlin: Akademie-Verlag.<br />

Fourier, J. (1822). “Théorie analytique de la chaleur”. in: Œuvres de Fourier. vol. 1. First<br />

published Paris: Firmin Didot, Père et fils. Paris: Gauthier-Villars et fils.<br />

Francœur, L.-B. (1809). Cours complet de mathématiques pures. 2 vols. Paris: Courcier.<br />

Francœur, L.-B. (1815). Lehrbegriff der reinen Mathematik. Aus d. Franz. übersetzt u. mit<br />

erläuternden Anmerkungen und Zusätze begleitet von Carl Ferd. Degen. Kph.<br />

Fuss, P.-H., ed. (1968). Correspondance mathématique et physique de quelques célèbres<br />

géomètres du XVIIIème siècle. <strong>The</strong> Sources <strong>of</strong> Science 35. Reprint <strong>of</strong> the original St.-<br />

Pétersbourgh 1843 print. New York and London: Johnson Reprint Corporation.<br />

Galois, É. (1830). “Analyse d’un mémoire sur la résolution algébrique des équations”.<br />

in: Écrits et mémoires mathématiques d’Évariste Galois. ed. by R. Bourgne and J.-P.<br />

Azra. First published Bulletin des sciences mathématiques de Ferrusac, XIII, §138, April<br />

1830. Paris: Gauthier-Villars, 163–165.<br />

— (1831a). “Mémoire sur la division d’une fonction elliptique de première classe”. in:<br />

Écrits et mémoires mathématiques d’Évariste Galois. ed. by R. Bourgne and J.-P. Azra.<br />

Paris: Gauthier-Villars, 153–161.<br />

— (1831b). “Note sur <strong>Abel</strong>”. in: Écrits et mémoires mathématiques d’Évariste Galois. ed.<br />

by R. Bourgne and J.-P. Azra. Paris: Gauthier-Villars, 35.<br />

— (Oct. 1831c). “Préface à deux mémoires d’analyse par E. Galois”. in: Écrits et mémoires<br />

mathématiques d’Évariste Galois. ed. by R. Bourgne and J.-P. Azra. Paris:<br />

Gauthier-Villars, 3–11.<br />

Gauss and Bessel (1880). Briefwechsel zwischen Gauss und Bessel. Leipzig: Königliche<br />

Preußische Akademie der Wissenschaften, Wilhelm Engelmann.<br />

Gauss, C. F. (1818). “Determinatio attractionis quam in punctum quodvis positionis<br />

datae exerceret planeta si eius massa per totam orbitam ratione temporis quo<br />

singulae partes describuntur uniformiter esset dispertita”. in: Carl Friedrich Gauss<br />

Werke. vol. 3. 12 vols. First published Commentationes societatis regiae scientiarum<br />

Gottingensis recentiores, Vol. IV, 1818. Göttingen: Königliche Gesellschaft der Wissenschaften,<br />

331–355.<br />

— (1890). Die vier Gauss’schen Beweise für die Zerlegung ganzer algebraischer Functionen<br />

in reelle Factoren ersten oder zweiten Grades. (1799–1849). ed. by E. Netto. Ostwald’s<br />

Klassiker der exakten Wissenschaften 14. Leipzig: Verlag von Wilhelm Engelsmann.<br />

— (1981). Mathematisches Tagebuch 1796–1814. Ostwalds Klassiker der exakten Wissenschaften<br />

256. Leipzig: Akademische Verlagsgesellschaft, Geest & Portig K.-G.<br />

Gauss, C. F. (1799). “Demonstratio nova theorematis omnem functionem algebraicam<br />

rationalem integram unius variabilis in factores reales primi vel secundi gradus re-


420 Bibliography<br />

solvi posse”. in: Carl Friedrich Gauss Werke. vol. 3. 12 vols. First published Helmstad:<br />

Fleckeisen. Translated into German in [?, 3–36]. Göttingen: Königliche Gesellschaft<br />

der Wissenschaften, 1–30.<br />

Gauss, C. F. (1801). “Disquisitiones arithmeticae”. in: Carl Friedrich Gauss Werke. vol. 1.<br />

12 vols. First published Leipzig: Gerh. Fleischer. Göttingen: Königliche Gesellschaft<br />

der Wissenschaften, 3–474.<br />

— (1813). “Disquisitiones generales circa seriem infinitam 1 + αβ<br />

α(α+1)β(β+1)<br />

x + 1·2·γ(γ+1) xx +<br />

1·γ<br />

α(α+1)(α+2)β(β+1)(β+2)<br />

1·2·3·γ(γ+1)(γ+2) x3 + etc.”. in: Carl Friedrich Gauss Werke. vol. 3. 12 vols.<br />

First published Commentationes societatis regiae scientiarum Gottingensis recentiores.<br />

Vol. II. Gottingae MDCCCXIII. Göttingen: Königliche Gesellschaft der Wissenschaften,<br />

125–162.<br />

— (1815). “Demonstratio nova altera theorematis omnem functionem algebraicam rationalem<br />

integram unius variabilis in factores reales primi vel secundi gradus resolvi<br />

posse”. in: Carl Friedrich Gauss Werke. vol. 3. 12 vols. First published Commentationes<br />

societatis regiae scientiarum Gottingensis recentiores, vol. III, Gottingae 1816.<br />

Translated into German in [?, 37–60]. Göttingen: Königliche Gesellschaft der Wissenschaften,<br />

31–56.<br />

— (1863–1933). Carl Friedrich Gauss Werke. 12 vols. Reprinted Hildesheim / New<br />

York: Georg Olms Verlag, 1973–1981. Göttingen: Königliche Gesellschaft der Wissenschaften.<br />

— (1888). Allgemeine Untersuchungen über die unendliche Reihe 1 + αβ<br />

1·γ x +<br />

α(α+1)β(β+1) α(α+1)(α+2)β(β+1)(β+2)<br />

1·2·γ(γ+1) xx + 1·2·3·γ(γ+1)(γ+2) x3 + u.s.w.. ed. by H. Simon. Berlin: Verlag<br />

von Julius Springer.<br />

— (1986). Disquisitiones Arithmeticae. English Edition. Edited by A. A. Clarke and W. C.<br />

Waterhouse. New York, Berlin, Heidelberg, Tokyo: Springer-Verlag.<br />

— (Fa, Kapsel 46a, A1–A13). “Convergenz der Reihen, in welche die periodischen<br />

Functionen einer veränderlichen Größe entwickelt werden”. in: Carl Friedrich Gauss<br />

Werke. vol. 10. 12 vols. Handwritten note found in the Nachlass. Probably not written<br />

prior to 1831. Göttingen: Königliche Gesellschaft der Wissenschaften, 400–406.<br />

Gericke, H. (1970). Geschichte des Zahlbegriffs. B. I. Hochschultaschenbücher 172.<br />

Mannheim/Wien/Zürick: Bibliographisches Institut.<br />

Gillies, D., ed. (1992). Revolutions in <strong>Mathematics</strong>. Oxford: Clarendon Press.<br />

Gillispie, C. C., ed. (1970–80). Dictionary <strong>of</strong> Scientific Biography. vol. I–XVI. New York:<br />

Charles Scribner’s Sons.<br />

Giusti, E. (1984). “Gli ‘errori’ di Cauchy e i fondamenti dell’analisi”. in: Bollettino di<br />

Storia delle Scienze Matematiche 4.2, 24–54.<br />

Glaisher, J. W. L. (1902). “On the relation <strong>of</strong> the <strong>Abel</strong>ian to the Jacobian elliptic functions”.<br />

in: Acta Mathematica 26, 241–248.<br />

Goar, M. (Dec. 1999). “Olivier and <strong>Abel</strong> on Series Convergence: An Episode from Early<br />

19th Century Analysis”. in: <strong>Mathematics</strong> Magazine 72.5, 347–355.<br />

Grabiner, J. V. (1981a). “Changing Attitudes Toward Mathematical Rigor: Lagrange<br />

and Analysis in the Eighteenth and Nineteenth Centuries”. in: Epistemological and<br />

Social Problems in the Sciences in the Early Nineteenth Century. ed. by H. N. Jahnke<br />

and M. Otte. Dordrecht: D. Reidel Publishing Company, 311–330.<br />

— (1981b). <strong>The</strong> Origins <strong>of</strong> Cauchy’s Rigorous Calculus. Cambridge (Mass.): MIT Press.<br />

— (1990). <strong>The</strong> Calculus as Algebra. J.-L. Lagrange, 1736–1813. Harvard Dissertations in<br />

the History <strong>of</strong> Science. New York: Gerland Publishing, Inc.


Bibliography 421<br />

Grattan-Guinness, I. (1971). “Materials for the History <strong>of</strong> <strong>Mathematics</strong> in the Institut<br />

Mittag-Leffler”. in: ISIS 62, 363–374.<br />

Grattan-Guinness, I. (1970a). “Bolzano, Cauchy and the “New Analysis” <strong>of</strong> the Early<br />

Nineteenth Century”. in: Archive for History <strong>of</strong> Exact Sciences 6, 372–400.<br />

— (1970b). <strong>The</strong> development <strong>of</strong> the foundations <strong>of</strong> mathematical analysis from Euler to Riemann.<br />

Cambridge (Mass.): MIT Press.<br />

— (1972). Joseph Fourier 1768–1830. A survey <strong>of</strong> his life and work, based on a critical edition<br />

<strong>of</strong> his monograph on the propagation <strong>of</strong> heat, presented to the Institut de France in 1807.<br />

Cambridge (Mass.): MIT Press.<br />

— (1982). “Paris <strong>Mathematics</strong>, 1795–1830: A Study in Competition”. in: Acta historiae<br />

rerum naturalium nec non technicarum 13, 261–270.<br />

— (1986). “<strong>The</strong> Cauchy-Stokes-Seidel Story on Uniform Convergence Again: Was<br />

there a Fourth Man?” in: Bulletin de la Société Mathématique de Belgique (serie A) 38,<br />

225–235.<br />

— (1990). Convolutions in French <strong>Mathematics</strong> 1800–1840. vol. 2–4. Science Networks —<br />

Historical Studies. Basel, Boston, Berlin: Birkhäuser Verlag.<br />

Gray, J. J. (1984). “A commentary on Gauss’s mathematical diary, 1796–1814, with an<br />

English translation”. in: Expositiones Mathematicae 2, 97–130.<br />

Gray, J. (1992). “Cauchy — elliptic and abelian integrals”. in: Revue d’Histoire des Sciences<br />

45.1. Études sur Cauchy (1789–1857), 69–81. ISSN: 0151-4105.<br />

Gårding, L. (1992). “<strong>Abel</strong> och lösbare ekvationer av primtalsgrad”. in: Nordisk Matematisk<br />

Tidskrift 40.1, 1–13.<br />

Gårding, L. and C. Skau (1994). “<strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> and Solvable Equations”. in:<br />

Archive for History <strong>of</strong> Exact Sciences 48.1, 81–103.<br />

Göpel, A. (1847). Entwurf einer <strong>The</strong>orie der <strong>Abel</strong>’schen Transcendenten erster Ordnung. Ostwald’s<br />

Klassiker der exakten Wissenschaften 67. 1895. First published in Latin in<br />

Crelle’s Journal, Bd. 35, 1847, S. 277–312. Leipzig: Wilhelm Engelmann.<br />

Hamburg, R. R. (1976). “<strong>The</strong> <strong>The</strong>ory <strong>of</strong> Equations in the 18th Century: <strong>The</strong> Work <strong>of</strong><br />

Joseph Lagrange”. in: Archive for History <strong>of</strong> Exact Sciences 16.1, 17–36.<br />

Hamilton, W. R. (1839). “On the Argument <strong>of</strong> <strong>Abel</strong>, respecting the Impossibility <strong>of</strong><br />

expressing a Root <strong>of</strong> any General Equation above the Fourth Degree, by any finite<br />

Combination <strong>of</strong> Radicals and Rational Functions”. in: Transactions <strong>of</strong> the Royal Irish<br />

Academy 18.2. Read 22 May 1837. Reprinted in [?, vol. 3, 517–569], 171–259.<br />

Hamilton, W. R. (1836). “<strong>The</strong>orems respecting Algebraic Elimination, connected with<br />

the Question <strong>of</strong> the Possibility <strong>of</strong> resolving in finite Terms the general Equation <strong>of</strong><br />

the Fifth Degree”. in: <strong>The</strong> mathematical papers <strong>of</strong> Sir William Rowan Hamilton. ed. by<br />

H. Halberstam and R. E. Ingram. vol. 3. 3 vols. First published Phil. Mag. vols. VIII<br />

(1836), pp. 538–43 and IX (1836), pp. 28–32. Cambridge: at the University Press.<br />

chap. XLVII, 471–477.<br />

— (1843). “On Equations <strong>of</strong> the Fifth Degree: and especially on a certain System <strong>of</strong><br />

Expressions connected with these Equations, which Pr<strong>of</strong>essor Badano has lately<br />

proposed”. in: <strong>The</strong> mathematical papers <strong>of</strong> Sir William Rowan Hamilton. ed. by H. Halberstam<br />

and R. E. Ingram. vol. 3. 3 vols. First published Trans. Roy. Irish Acad. vol.<br />

XIX (1843), pp. 329–76. Read 4 August 1842. Cambridge: at the University Press.<br />

chap. XLII, 572–602.<br />

— (1844). “On a Method proposed by Pr<strong>of</strong>essor Badano for the solution <strong>of</strong> Algebraic<br />

Equations”. in: <strong>The</strong> mathematical papers <strong>of</strong> Sir William Rowan Hamilton. ed. by H.<br />

Halberstam and R. E. Ingram. vol. 3. 3 vols. First published Proc. Roy. Irish Acad. vol.


422 Bibliography<br />

II (1844), pp. 275–6. Communicated 4 August 1842. Cambridge: at the University<br />

Press. chap. XLI, 570–571.<br />

Hankins, T. L. (1972). “William Rowan Hamilton (1805–1865)”. in: Dictionary <strong>of</strong> Scientific<br />

Biography. ed. by C. C. Gillispie. vol. 6. New York: Charles Scribner’s Sons,<br />

85–93.<br />

— (1980). Sir William Rowan Hamilton. Baltimore and London: <strong>The</strong> Johns Hopkins University<br />

Press.<br />

Hansteen (1862). “<strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>”. in: Illustreret Nyhedsblad 11.9–10. 2den og 9de<br />

Marts, 37–38, 41–42.<br />

Hauch, M. (June 1997). “Bernard Bolzano. Bolzanos tidlige matematiske arbejder set i<br />

relation til hans matematikfilos<strong>of</strong>i.” Specialeopgave. Institut for de Eksakte Videnskabers<br />

Historie, Aarhus Universitet.<br />

Hawkins, T. (1970). Lebesgue’s theory <strong>of</strong> integration. Its origins and development. Madison<br />

and London: <strong>The</strong> University <strong>of</strong> Wisconsin Press, xv+227.<br />

Heegaard, P. (1937). “Ein Brief von <strong>Abel</strong> an Degen”. in: Acta Mathematica 68, 1–6.<br />

Heegaard, P. (1935). “Et brev fra <strong>Abel</strong> til Degen”. in: Norsk matematisk tidsskrift 17, 33–<br />

38.<br />

“Ein Brief von <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> an Edmund Jacob Külp” (1903). in: Journal für die<br />

reine und angewandte Mathematik 125.4. ed. by Hensel, 237–240.<br />

Hirano, Y. (1984). “Note sur les diffusions de la théorie de Galois. Première clarification<br />

des idées de Galois par Liouville”. in: Historia Scientiarum 27, 27–41.<br />

H<strong>of</strong>mann, J. E. (1949). Die Entwicklungsgeschichte der Leibnizschen Mathematik während<br />

des Aufenthaltes in Paris (1672–1676). München: Leibniz Verlag.<br />

Holmboe, B. (1829). “Necrolog. Kort Fremstilling af <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>s Liv og videnskabelige<br />

Virksomhed”. in: Magazin for Naturvidenskaberne, 334–354.<br />

Holst, E. (1902). “Historisk Indledning”. in: Festskrift ved Hundredeaarsjubilæet for <strong>Niels</strong><br />

<strong>Henrik</strong> <strong>Abel</strong>s Fødsel. ed. by E. Holst, C. Størmer, and L. Sylow. Kristiania: Jacob<br />

Dybwad.<br />

Houzel, C. (1986). “Fonctions elliptiques et intégrales abéliennes”. in: Abrégé d’histoire<br />

des mathématiques. ed. by J. Dieudonné. Paris: Hermann. chap. 7, 293–314.<br />

Jacobi, C. G. J. (1827a). “Demonstratio theorematis ad theoriam functionum ellipticarum<br />

spectantis”. in: C. G. J. Jacobi’s Gesammelte Werke. ed. by C. W. Borchardt,<br />

K. T. W. Weierstrass, and A. Clebsch. vol. 1. 8 vols. First published Schumacher Astronomische<br />

Nachrichten, Bd. 6. Nr. 127. December 1827. Berlin: G. Reimer, 37–48.<br />

— (1827b). “Extraits de deux lettres de M. Jacobi de l’université de Königsberg à M.<br />

Schumacher”. in: C. G. J. Jacobi’s Gesammelte Werke. ed. by C. W. Borchardt, K. T. W.<br />

Weierstrass, and A. Clebsch. vol. 1. 8 vols. First published Schumacher Astronomische<br />

Nachrichten, Band 6. Nr. 123. September 1827. Berlin: G. Reimer, 29–36.<br />

— (1828). “Note sur les fonctions elliptiques”. in: Journal für die reine und angewandte<br />

Mathematik 3.2, 192–195.<br />

— (1832a). “Anzeige von Legendre’s Théorie des fonctions elliptiques, dritte und letzte<br />

Supplement. Nachrichten von Büchern”. in: Journal für die reine und angewandte<br />

Mathematik 8. Reproduced in (JacobiWerke), 413–417.<br />

— (1832b). “Considerationes generales de transcendentibus <strong>Abel</strong>ianis”. in: Journal für<br />

die reine und angewandte Mathematik 9, 394–403.<br />

Jacobi, C. G. J. (1829). “Fundamenta nova theoriae functionum ellipticarum”. in: C.<br />

G. J. Jacobi’s Gesammelte Werke. ed. by C. W. Borchardt, K. T. W. Weierstrass, and<br />

A. Clebsch. vol. 1. 8 vols. First published Regiomonti: Bornträger, 1829. Berlin: G.<br />

Reimer, 49–239.


Bibliography 423<br />

Jahnke, H. N. (1987). “Motive und Probleme der Arithmetisierung der Mathematik in<br />

der ersten Hälfte des 19. Jahrhunderts – Cauchys Analysis in der Sicht des Mathematikers<br />

Martin Ohm”. in: Archive for History <strong>of</strong> Exact Sciences 37, 101–182.<br />

— (1990). Mathematik und Bildung in der Humboldtschen Reform. Studien sur<br />

Wissenschafts-, Social- und Bildungsgeschichte der Mathematik 8. Göttingen: Vandenhoeck<br />

& Ruprecht.<br />

— (1992). “A structuralist view <strong>of</strong> Lagrange’s algebraic analysis and the German combinatorial<br />

school”. in: <strong>The</strong> Space <strong>of</strong> <strong>Mathematics</strong>. Philosophical, Epistemological, and<br />

Historical Explorations. ed. by J. Echeverria, A. Ibarra, and T. Mormann. Berlin/New<br />

York: Walter de Gruyter, 280–295.<br />

— (1993). “Algebraic Analysis in Germany, 1780–1840: Some Mathematical and Philosophical<br />

Issues”. in: Historia Mathematica 20.3, 265–284.<br />

— (1994). “Cultural influences on mathematics teaching: the ambiguous role <strong>of</strong> applications<br />

in nineteenth-century Germany”. in: Didactics <strong>of</strong> mathematics as a scientific<br />

discipline. Dordrecht: Kluwer Acad. Publ., 415–429.<br />

— (1996). “<strong>The</strong> development <strong>of</strong> algebraic analysis from Euler to Klein and its impact<br />

on school mathematics in the 19th century”. in: Vita mathematica (Toronto, ON, 1992;<br />

Quebec City, PQ, 1992). Washington, DC: Math. Assoc. America, 145–151.<br />

— (1999). “Die algebraische Analysis des 18. Jahrhunderts”. in: Geschichte der Analysis.<br />

ed. by H. N. Jahnke. Texte zur Didaktik der Mathematik. Heidelberg and Berlin:<br />

Spektrum Akademischer Verlag. chap. 4, 131–170.<br />

Johnsen, K. (1984). “Zum Beweis von C. F. Gauss für die Irreduzibilität des p-ten Kreisteilungspylonoms”.<br />

in: Historia Mathematica 11, 131–141.<br />

Kiernan, B. M. (1971). “<strong>The</strong> Development <strong>of</strong> Galois <strong>The</strong>ory from Lagrange to Artin”.<br />

in: Archive for History <strong>of</strong> Exact Sciences 8.1–2, 40–154.<br />

Klein, F. (1967). Vorlesungen über die Entwicklung der Mathematik im 19. Jahrhundert. First<br />

published in three volumes 1926–27. New York: Chelsea Publishing Company.<br />

Kline, M. (1990). Mathematical Thought from Ancient to Modern Times. 3 volumes.First<br />

published 1972. Oxford: Oxford University Press.<br />

Knobloch, E. (1998). “<strong>Mathematics</strong> at the Prussian Academy <strong>of</strong> Sciences 1700–1810”.<br />

in: <strong>Mathematics</strong> in Berlin. ed. by H. G. W. Begehr et al. Berlin, Basel, Boston:<br />

Birkhäuser Verlag, 1–8.<br />

Koenigsberger, L. (1879). Zur Geschichte der <strong>The</strong>orie der elliptischen Transcendenten in den<br />

Jahren 1826–29. Leipzig: B. G. Teubner.<br />

Kolmogorov, A. N. and A. P. Yushkevich, eds. (1992–1998). <strong>Mathematics</strong> <strong>of</strong> the 19th century.<br />

3 vols. Translated by R. Cooke. Basel, Boston, Berlin: Birkhäuser Verlag.<br />

Kracht, M. and E. Kreyszig (1990). “E. W. von Tschirnhaus: His Role in Early Calculus<br />

and His Work and Impact on Algebra”. in: Historia Mathematica 17.1, 16–35.<br />

Kragemo, H. B. (1929). “Tre brev fra <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> og hans bror til pastor John<br />

Aas”. in: Norsk matematisk tidsskrift 11, 49–52.<br />

Krazer, A. (1909). “Zur Geschichte des Umkehrproblems der Integrale”. in: Jahresbericht<br />

der deutchen Mathematiker-Vereinigung 18, 44–75.<br />

Kronecker, L. (1853). “Über die algebraisch auflösbaren Gleichungen (I)”. in: Leopold<br />

Kronecker’s Werke. ed. by K. Hensel. vol. 4. First published Monatsberichte der<br />

Königlich Preussischen Akademie der Wissenschaften zu Berlin, 1853, pp. 365–374.<br />

Leipzig and Berlin: B. G. Teubner. chap. 1, 1–11.<br />

— (1856). “Über die algebraisch auflösbaren Gleichungen (II)”. in: Leopold Kronecker’s<br />

Werke. ed. by K. Hensel. vol. 4. First published Monatsberichte der Königlich Preussis-


424 Bibliography<br />

chen Akademie der Wissenschaften zu Berlin, 1856, pp. 203–215. Leipzig and Berlin: B.<br />

G. Teubner. chap. 4, 25–37.<br />

Kronecker, L. (1879). “Einige Entwickelungen aus der <strong>The</strong>orie der algebraischen Gleichungen”.<br />

in: Leopold Kronecker’s Werke. ed. by K. Hensel. vol. 4. First published<br />

Monatsberichte der Königlich Preussischen Akademie der Wissenschaften zu Berlin, 1879,<br />

pp. 205–229. Read on March 3rd 1879. Leipzig and Berlin: B. G. Teubner. chap. 10,<br />

73–96.<br />

Kuhn, T. S. (1962). <strong>The</strong> Structure <strong>of</strong> Scientific Revolutions. Chicago and London: <strong>The</strong><br />

University <strong>of</strong> Chicago Press.<br />

Königsberger (1869). “Berichtigung eines Satzes von <strong>Abel</strong>, die Darstellung der algebraischen<br />

Functionen betreffend”. in: Mathematische Annalen 1.2, 168–169.<br />

Lacroix, S. F. (1797). Traité du Calcul différentiel et du Calcul intégral. vol. 1. Paris: J. B. M.<br />

Duprat.<br />

— (1798). Traité du Calcul différentiel et du Calcul intégral. vol. 2. Paris: J. B. M. Duprat.<br />

— (1800). Traité des différences et des séries, faisant suite au Traité du Calcul différentiel et<br />

du Calcul intégral. Paris: J. B. M. Duprat.<br />

Lagrange, J.-L. (1770–1771). “Réflexions sur la Résolution algébrique des équations”.<br />

in: Œuvres de Lagrange. ed. by J.-A. Serret. vol. 3. 15 vols. First published Nouveaux<br />

Mémoires de l’Académie royale des Sciences et Belles-Lettres de Berlin, 1770–1771. Paris:<br />

Gauthier-Villars, 203–421.<br />

— (1784–1785). “Sur une nouvelle méthode de calcul intégral pour les différentielles<br />

affectées d’un radical carré sous lequel la variable ne passe pas le quatrième degré”.<br />

in: Œuvres de Lagrange. ed. by J.-A. Serret. vol. 2. 15 vols. First published<br />

Mémoires de l’Académie royale des Sciences de Turin, t. II, 1784–1785. Paris: Gauthier-<br />

Villars, 252–312.<br />

— (1813). Théorie des fonctions analytiques. Nouvelle édition. Reprinted in (Lagrange,<br />

1867–1892, vol. IX). Paris: Courcier.<br />

— (1867–1892). Œuvres de Lagrange. ed. by J.-A. Serret. 15 vols. Paris: Gauthier-Villars.<br />

Lakatos, I. (1976). Pro<strong>of</strong>s and Refutations. <strong>The</strong> Logic <strong>of</strong> Mathematical Discovery. Cambridge:<br />

Cambridge University Press.<br />

Lange-<strong>Niels</strong>en, F. (1927). “Zur Geschichte des <strong>Abel</strong>schen <strong>The</strong>orems. Das Schicksal der<br />

Pariserabhandlung”. in: Norsk matematisk tidsskrift 9, 55–73.<br />

— (1929). “<strong>Abel</strong> og “Académie des Sciences” i Paris”. in: Norsk matematisk tidsskrift 11,<br />

13–17.<br />

Larvor, B. (1998). Lakatos: An Introduction. London and New York: Routledge.<br />

Laudal, O. A. and R. Piene, eds. (2004). <strong>The</strong> Legacy <strong>of</strong> <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. <strong>The</strong> <strong>Abel</strong> Bicentennial,<br />

Oslo, 2002. Berlin etc.: Springer.<br />

Laugwitz, D. (1987). “Infinitely Small Quantities in Cauchy’s Textbooks”. in: Historia<br />

Mathematica 14, 258–274.<br />

— (1988–89). “Definite Values <strong>of</strong> Infinite Sums: Aspects <strong>of</strong> the Foundations <strong>of</strong> Infinitesimal<br />

Analysis around 1820”. in: Archive for History <strong>of</strong> Exact Sciences 39, 195–<br />

245.<br />

— (1999). Bernhard Riemann 1826–1866. Turning Points in the Conception <strong>of</strong> <strong>Mathematics</strong>.<br />

Translated by A. Shenitzer from the 1996 German original. Boston, Basel, Berlin:<br />

Birkhäuser.<br />

Legendre and Jacobi (1875). “Correspondance mathématique entre Legendre et Jacobi”.<br />

in: Journal für die reine und angewandte Mathematik 80. Reproduced in (JacobiWerke),<br />

205–279.


Bibliography 425<br />

Legendre, A. M. (1811–1817). Exercises de calcul intégral sur divers ordres de transcendantes<br />

et sur les quadratures. 3 vols (1811, 1817, 1816). Paris: Courcier.<br />

— (1825–1828). Traité des fonctions elliptiques et des intégrales eulériennes. 3 volumes<br />

(1825, 1826, 1828). Paris: Imprimerie de Huzard-Courcier.<br />

Legendre, A.-M. (1793). Mémoire sur les transcendantes elliptiques. Paris: du Pont and<br />

Firmin Didot.<br />

Loria, G. (1902). Spezielle algebraische und transscendente Ebene Kurven: <strong>The</strong>orie und<br />

Geschichte. ed. by F. Schütte. Leipzig: Druch und Verlag von B. G. Teubner.<br />

Lützen, J. (1978). “Funktionsbegrebets udvikling fra Euler til Dirichlet”. in: Nordisk<br />

Matematisk Tidskrift 25/26.1, 5–32.<br />

— (1990). Joseph Liouville 1809–1882: Master <strong>of</strong> Pure and Applied <strong>Mathematics</strong>. Studies in<br />

the History <strong>of</strong> <strong>Mathematics</strong> and Physical Sciences 15. New York: Springer-Verlag.<br />

— (1999). “Grundlagen der Analysis im 19. Jahrhundert”. in: Geschichte der Analysis.<br />

ed. by H. N. Jahnke. Texte zur Didaktik der Mathematik. Heidelberg and Berlin:<br />

Spektrum Akademischer Verlag. chap. 6, 191–244.<br />

Malfatti, G. (1804). “Dubbj proposti al socio Paolo Ruffini sulla sua dimostrazione<br />

dell’impossibilità di risolvere le equazioni superiori al quarto grado”. in: Gianfrencesco<br />

Malfatti Opere. vol. 2. First published Memorie di Matematica e Fisica della<br />

Società Italiana, 11 (1804) pp. 579–607. Bologna: Edizioni Cremonese, 723–751.<br />

Malmsten, C. J. (1847). “In solutionem Aequationum Algebraicarum Disquisitio”. in:<br />

Journal für die reine und angewandte Mathematik 34.1, 46–74.<br />

Martini, L. (1999). “<strong>The</strong> First Lectures in Italy on Galois <strong>The</strong>ory: Bologna, 1886–1887”.<br />

in: Historia Mathematica 26.3, 201–223.<br />

Mehrtens, H. (1981). “Mathematicians in Germany Circa 1800”. in: Epistemological and<br />

Social Problems in the Sciences in the Early Nineteenth Century. ed. by H. N. Jahnke<br />

and M. Otte. Dordrecht: D. Reidel Publishing Company, 401–420.<br />

Mertens, F. (1875). “Ueber die Multiplicationsregel für zwei unendliche Reihen”. in:<br />

Journal für die reine und angewandte Mathematik 79, 182–184.<br />

Mikulinsky, S. R. (1982). “<strong>The</strong> Historical Features <strong>of</strong> the Development <strong>of</strong> Science in the<br />

First Half <strong>of</strong> the 19th Century”. in: Acta historiae rerum naturalium nec non technicarum<br />

13, 31–58.<br />

Mittag-Leffler, G. (1907). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. Paris: La revue du mois.<br />

Natucci, A. (1971). “Giulio Carlo Fagnano dei Toschi (1682–1766)”. in: Dictionary <strong>of</strong><br />

Scientific Biography. ed. by C. C. Gillispie. vol. 4. New York: Charles Scribner’s Sons,<br />

515–516.<br />

Nørgaard, S. (1990). “Elliptiske funktioner og kompleks funktionsteori 1825-1860. Aspekter<br />

af inversionsproblemet for elliptiske integraler set i relation til udviklingen<br />

af en kompleks funktionsteori i Frankrig”. Speciale. Institut for de Eksakte Videnskabers<br />

Historie, Aarhus Universitet.<br />

Olivier, L. (1826a). “Bemerkungen über die Form der Wurzeln algebraischer Gleichungen”.<br />

in: Journal für die reine und angewandte Mathematik 1.2, 97–116.<br />

— (1826b). “Entwickelung einer beliebigen Potenz eines Cosinus durch die Cosinus<br />

der vielfachen Bogen”. in: Journal für die reine und angewandte Mathematik 1.1, 16–36.<br />

— (1826c). “Ueber einige Definitionen in der Geometrie”. in: Journal für die reine und<br />

angewandte Mathematik 1.3, 241–252.<br />

— (1827). “Remarques sur les séries infinies et leur convergence”. in: Journal für die<br />

reine und angewandte Mathematik 2.1, 31–44.<br />

— (1828). “Remarque de Mr. L. Olivier”. in: Journal für die reine und angewandte Mathematik<br />

3.1, 82.


426 Bibliography<br />

Ore, Ø. (Jan. 1950). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. Kurze Mathematiker-Biographien 8. Basel: Verlag<br />

Birkhäuser.<br />

— (1954). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. Et geni og hans samtid. Oslo: Gyldendal Norsk Forlag.<br />

— (1957). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>. Mathematician Extraordinary. Minneapolis: University <strong>of</strong><br />

Minnesota Press.<br />

Otero, M. H. (1997). Joseph-Diez Gergonne (1771–1859). Logique et philosophie des sciences.<br />

vol. 37. Histoire et philosophie des sciences. Nantes: Université de Nantes.<br />

Otte, M. (1982). “<strong>The</strong> Early 19th Century as a Period <strong>of</strong> Transitional Character”. in:<br />

Acta historiae rerum naturalium nec non technicarum 13, 159–161.<br />

Pensivy, M. (1994). “<strong>The</strong> binomial theorem”. in: Companion encyclopedia <strong>of</strong> the history<br />

and philosophy <strong>of</strong> the mathematical sciences. ed. by I. Grattan-Guinness. vol. 1. 2 volumes.<br />

London: Routledge. chap. 4.1, 492–498.<br />

Pesic, P. (2003). <strong>Abel</strong>’s Pro<strong>of</strong>. An Essay on the Sources and Meaning <strong>of</strong> Mathematical Unsolvability.<br />

Cambridge (MA)/London: <strong>The</strong> MIT Press.<br />

Pieper, H., ed. (1998). Korrespondenz Adrien-Marie Legendre – Carl Gustav Jacob Jacobi.<br />

Correspondance mathématique entre Legendre et Jacobi. Teubner-Archiv zur Mathematik<br />

19. Stuttgart, Leipzig: B. G. Teubner.<br />

Pierpont, J. (1896). “On the Ruffini-<strong>Abel</strong>ian theorem”. in: Bulletin <strong>of</strong> the American Mathematical<br />

Society 3, 200–221.<br />

Pierpont, J. (Apr. 1898). “Early history <strong>of</strong> Galois’ theory <strong>of</strong> equations”. in: Bulletin <strong>of</strong><br />

the American Mathematical Society 4, 332–340.<br />

Poggendorff, J. C., ed. (1965). Biographisch-literarisches Handwörterbuch zur Geschichte<br />

der exacten Wissenschaften. Unedited reproduction <strong>of</strong> the Leibzig 1863 edition. Amsterdam:<br />

B. M. Israël n. v.<br />

Poisson (1811). “Note sur le développement des puissances des sinus et des cosinus,<br />

en séries de sinus ou de cosinus d’arcs multiples”. in: Correspondance sur l’Ecole<br />

Impériale Polytechnique 2.3, 212–217.<br />

Pringsheim, A. (1898–1904). “Irrationalzahlen und Konvergenz unendlicher<br />

Prozesse”. in: Encyklopädie der Mathematischen Wissenschaften mit Einschluss<br />

ihrer Anwendungen. ed. by W. F. Meyer. vol. I,1. Arithmetik und Algebra A.<br />

Leipzig: B. G. Teubner. chap. 3, 47–146.<br />

Pyenson, L. (1983). Neohumanism and the Persistence <strong>of</strong> Pure <strong>Mathematics</strong> in Wilhelmian<br />

Germany. Philadelphia: American Philosophical Society.<br />

Radl<strong>of</strong>f, I. (1998). “<strong>Abel</strong>s Unmöglichkeitsbeweis im Spiegel der modernen Galoistheorie”.<br />

in: Mathematische Semesterberichte 45.2, 127–139.<br />

— (2002). “Évariste Galois: Principles and Applications”. in: Historia Mathematica 29,<br />

114–137.<br />

Ramskov, K. (1995). “Matematikeren Harald Bohr”. Danish. Licentiatafhandling. Institut<br />

for de eksakte videnskabers historie, Aarhus Universitet.<br />

Riemann, B. (1857). “<strong>The</strong>orie der <strong>Abel</strong>’schen Functionen”. in: Journal für die reine und<br />

angewandte Mathematik 54, 115–155.<br />

Riemann, B. (1854). “Ueber die Darstellbarkeit einer Function durch eine<br />

trigonometrische Reihe”. in: <strong>The</strong> Collected Works <strong>of</strong> Bernard Riemann (Gesammelte<br />

mathematische Werke und wissenschaftlicher Nachlass). First published Abhandlungen<br />

der Königlichen Gesellschaft der Wissenschaften zu Göttingen, volume 13. New York:<br />

Dover Publications, 227–264.<br />

Rosen, M. I. (1995). “<strong>Niels</strong> Hendrik <strong>Abel</strong> and Equations <strong>of</strong> the Fifth Degree”. in: <strong>The</strong><br />

American Mathematical Monthly 102.6, 495–505.


Bibliography 427<br />

Rosen, M. (1981). “<strong>Abel</strong>’s theorem on the lemniscate”. in: <strong>The</strong> American Mathematical<br />

Monthly 88, 387–395.<br />

Rosenhain, G. (1851). Abhandlung ueber die Functionen zweier Variabler mit vier Perioden.<br />

Ostwald’s Klassiker der exakten Wissenschaften 65. 1895. First published in French<br />

in Mém. des savants, Bd. IX, 1851. Leipzig: Wilhelm Engelmann.<br />

Rothman, T. (1982). “Genius and Biographers: the Fictionalization <strong>of</strong> Evariste Galois”.<br />

in: <strong>The</strong> American Mathematical Monthly 89, 84–106.<br />

Rowe, D. E. (1998). “<strong>Mathematics</strong> in Berlin, 1810–1933”. in: <strong>Mathematics</strong> in Berlin. ed.<br />

by H. G. W. Begehr et al. Berlin, Basel, Boston: Birkhäuser Verlag, 9–26.<br />

Rudio, F. (1895). “Eine Autobiographie von Gotthold Eisenstein. Mit ergänzenden biographischen<br />

Notizen”. in: Gotthold Eisenstein: Mathematische Werke. vol. 2. 2 vols.<br />

First published Zeitschrift für Mathematik und Physik, Supplement zum 40. Jahrgang,<br />

Leipzig (1895), pp. 143–168. New York, N. Y.: Chelsea Publishing Company.<br />

chap. 47, 879–904.<br />

Rudio, F. (1921). “Vorwort”. in: Leonhardi Euleri opera omnia. ed. by F. Rudio et al. vol. 6.<br />

? vols. 1st series. Leipzig and Bern: B. G. Teubner, vii–xxvi.<br />

Ruffini, P. (1799). “Teoria generale delle equazioni, in cui si dimostra impossible la<br />

soluzione algebraica delle equazioni generali di grado superiore al quarto”. in:<br />

Opere Matematiche. ed. by E. Bortolotti. vol. 1. 3 vols. First published in two parts<br />

Bologna: Tommaso d’Aquino. Palermo: Tipografia matematica, 1–324.<br />

— (1805). “Risposta di Paolo Ruffini ai dubbi propostigli dal socio Gian-Francesco<br />

Malfatti sopra la insolubilità algebraica dell’equazioni di grado superiore al<br />

quarto”. in: Opere Matematiche. ed. by E. Bortolotti. vol. 2. 3 vols. First published<br />

Società Italiana delle Scienze, vol. 12. Palermo: Tipografia matematica, 51–90.<br />

— (1813). “Riflessioni intorno alla soluzione delle equazioni algebraiche generali”.<br />

in: Opere Matematiche. ed. by E. Bortolotti. vol. 2. 3 vols. First published Modena:<br />

Presso la società tipografica. Palermo: Tipografia matematica, 155–268.<br />

— (1915–1954). Opere Matematiche. ed. by E. Bortolotti. 3 vols. Palermo: Tipografia<br />

matematica.<br />

[Saigey] (1826). “Développement d’une puissance quelconque d’un cosinus par les<br />

cosinus des arcs multiples; par M. Olivier. (Ibid. p. 16.)” in: Bulletin des sciences<br />

mathématiques, astronomiques, physiques et chimiques 6. Review no. 58, 112.<br />

Schlesinger, L. (1922–1933). “Über Gauss’ Arbeiten zur Funktionentheorie”. in: Carl<br />

Friedrich Gauss Werke. vol. X.2. 12 vols. Edited version <strong>of</strong> the 1912 original. Göttingen:<br />

Königliche Gesellschaft der Wissenschaften.<br />

Schneider, I. (1981). “Herausragende Einzelleistungen im Zusammenhang mit der<br />

Kreisteilungsgleichung, dem Fundamentalsatz der Algebra und der Reihenkonvergenz”.<br />

in: Carl Friedrich Gauß (1777–1855). Sammelband von Beiträgen zum 200.<br />

Geburtstag von C. F. Gauß. ed. by I. Schneider. Wissenschaftsgeschichte. Beiträge<br />

aus dem Forchungsinstitut des Deutschen Museums für die Geschichte der Naturwissenschaften<br />

und der Technik. München: Minerva Publikation, 37–63.<br />

Scholz, E. (1990). “Die Entstehung der Galoistheorie”. in: Geschichte der Algebra: eine<br />

Einführung. ed. by E. Scholz. vol. 16. Lehrbücher und Monographien zur Didaktik<br />

der Mathematik. Mannheim, Wien, Zürich: Bibliographisches Institut, Wissenschaftsverlag.<br />

chap. 14, 365–398.<br />

Scott, J. F. (1976). “Edward Waring (∼1736–1798)”. in: Dictionary <strong>of</strong> Scientific Biography.<br />

ed. by C. C. Gillispie. vol. 14. New York: Charles Scribner’s Sons, 179–181.<br />

Scriba, C. J. (1980). “Viggo Brun in memoriam (1885–1978)”. in: Historia Mathematica<br />

7.1, 1–6.


428 Bibliography<br />

Seidel, P. L. (1847). “Note über eine Eigenschaft der Reihen, welche discontinuirliche<br />

Functionen darstellen”. in: Die Darstellung ganz willkürlicher Functionen durch Sinusund<br />

Cosinusreihen von Lejeune Dirichlet (1837) und Note über eine Eigenschaft der Reihen,<br />

welche discontinuirliche Functionen darstellen von Philipp Ludwig Seidel (1847). ed.<br />

by H. Liebmann. First published Abhandl. der Math. Phys. Klasse der Kgl. Bayerischen<br />

Akademie der Wissenschaften, V (1847), 381–394. Leipzig: Verlag von Wilhelm Engelmann.<br />

Shafarevich, I. R. (1974). Basic Algebraic Geometry. Die Grundlehren der mathematischen<br />

Wissenschaften in Einzeldarstellungen mit besondere Berücksichtigung der<br />

Anwendungsgebiete 213. Translated by K. A. Hirsch. Berlin, Heidelberg, New<br />

York: Springer-Verlag.<br />

Shils, E. (1961). “Centre and Periphery”. in: <strong>The</strong> Logic <strong>of</strong> Personal Knowledge: Essays<br />

Presented to Michael Polanyi. London: Routledge & Kegan Paul. chap. 11, 117–130.<br />

Siegel, C. L. (1959). “Zur Vorgeschichte des Eulerschen Additionstheorems”. in: Sammelband<br />

der zu Ehren des 250. Geburtstages Leonhard Eulers der Deutschen Akademie<br />

der Wissenschaften zu Berlin vorgelegten Abhandlungen. ed. by K. Schröder. Berlin:<br />

Akademie-Verlag, 315–317.<br />

Skau, C. (1990). “Gjensyn med <strong>Abel</strong>s og Ruffinis bevis for umuligheten av å løse den<br />

generelle n’tegradsligningen algebraisk når n ≥ 5”. in: NORMAT 38.2, 53–84.<br />

Smith, D. E. and M. L. Latham, eds. (1954). <strong>The</strong> geometry <strong>of</strong> René Descartes. Facsimile<br />

and English translation. New York: Dover Publications, Inc.<br />

Spalt, D. D. (1981). Vom Mythos der mathematischen Vernunft. Eine Archäologie zum<br />

Grundlagenstreit der Analysis oder Dokumentation einer vergeblichen Suche nach der Einheit<br />

der Mathematischen Vernunft. Darmstadt: Wissenschaftliche Buchgesellschaft.<br />

— (2002). “Cauchys Kontinuum. Eine historiografische Annäherung via Cauchys<br />

Summensatz”. in: Archive for History <strong>of</strong> Exact Sciences 56, 285–338.<br />

Stokes, G. G. (1847). “On the Critical Values <strong>of</strong> the Sums <strong>of</strong> Periodic Series”. in: Mathematical<br />

and Physical Papers. vol. 1. 2 vols. Read to the Cambridge Philosophical<br />

Society on December 6, 1847. Cambridge: at the University Press, 236–313.<br />

Stolz, O. (1904). “Die Bedeutung der <strong>Abel</strong>’schen Abhandlung über die binomische<br />

Reihe für die Functionentheorie”. in: Acta Mathematica 28, 303–306.<br />

Stubhaug, A. (1996). Et foranskutt lyn — <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> og hans tid. Oslo: Aschehoug.<br />

— (2000). <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> and his Times. Called too soon by Flames Afar. trans. by R. H.<br />

Daly. Berlin etc.: Springer.<br />

Sylow (1861). “Om algebraisk Opløsning af Ligninger”. in: Forhandlinger ved de Skandinaviske<br />

Naturforskeres ottende Møde. København, 536–548.<br />

Sylow, L. (1902). “<strong>Abel</strong>s Studier og hans Opdagelser”. in: Festskrift ved Hundredeaarsjubilæet<br />

for <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>s Fødsel. ed. by E. Holst, C. Størmer, and L. Sylow. Kristiania:<br />

Jacob Dybwad.<br />

Sørensen, H. K. (Nov. 1999). “Matematikken i starten af 1800-tallet. <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong> i<br />

en matematisk brydningstid”. Progress-raport. Aarhus: Institut for Videnskabshistorie,<br />

Aarhus Universitet.<br />

— (2002). “<strong>Abel</strong> and his mathematics in contexts”. in: NTM: International Journal <strong>of</strong><br />

History and Ethics <strong>of</strong> Natural Sciences, Technology and Medicine 10, 137–155. DOI: ���<br />

���������������.<br />

— (2005). “Exceptions and counterexamples. Understanding <strong>Abel</strong>’s comment on<br />

Cauchy’s <strong>The</strong>orem”. in: Historia Mathematica 32.4, 453–480. DOI: ��������������<br />

��������������.


Bibliography 429<br />

Toti Rigatelli, L. (1994). “<strong>The</strong> theory <strong>of</strong> equations from Cardano to Galois 1540–1830”.<br />

in: Companion encyclopedia <strong>of</strong> the history and philosophy <strong>of</strong> the mathematical sciences.<br />

ed. by I. Grattan-Guinness. 2 volumes. London: Routledge. chap. 6.1, 713–721.<br />

— (1996). Evariste Galois (1811–1832). Vita Mathematica 11. Basel/Boston/Berlin:<br />

Birkhäuser Verlag.<br />

Vandermonde (1771). “Mémoire sur la résolution des équations”. in: Histoire de<br />

l’académie royale des sciences. Volume not published until 1774, 365–416.<br />

Volkert, K. T. (1986). Die Krise der Anschauung. Eine Studie zu formalen und heuristischen<br />

Verfahren in der Mathematik seit 1850. Studien zur Wissenschafts-, Sozial- und<br />

Bildungsgeschichte der Mathematik 3. Göttingen: Vandenhoeck & Ruprecht.<br />

Volkert, K. (1987). “Die Geschichte der pathologischen Functionen. Ein Beitrag zur<br />

Entstehung der mathematischen Methodologie”. in: Archive for History <strong>of</strong> Exact Sciences<br />

37.3, 193–232.<br />

— (1989). “Zur Differenzierbarkeit stetiger Funktionen. Ampère’s Beweis und seine<br />

Folgen”. in: Archive for History <strong>of</strong> Exact Sciences 40.1, 37–112.<br />

Waerden, B. L. van der (1985). A History <strong>of</strong> Algebra. From al-Khwārizmī to Emmy Noether.<br />

Berlin etc.: Springer-Verlag.<br />

Wantzel (1845). “Démonstration de l’impossibilité de résoundre toutes équations algébriques<br />

avec des radicaux”. in: Nouvelles annales de mathematique 4, 57–65.<br />

Waring, E. (1770). Meditationes Algebraicæ. 2nd. English translation in [?] to which<br />

pages refer. Cantabrigiæ: typis academicis excudebat J. Archdeacon veneunt apud<br />

J. Woodyer.<br />

— (1991). Meditationes Algebraicae: An English Translation. Edited and translated from<br />

Latin by D. Weeks. Providence, R.I.: AMS.<br />

Weierstrass, K. (1923). “Briefe an Paul du Bois-Reymond”. in: Acta Mathematica 39,<br />

199–225.<br />

Weierstrass, K. ([1841] 1894). “Zur <strong>The</strong>orie der Potenzreihen”. in: Mathematische Werke<br />

von Karl Weierstrass. vol. 1. Manuscript <strong>of</strong> 1841. First published 1894. Berlin: Mayer<br />

und Müller, 67–74.<br />

— (1876). “Zur <strong>The</strong>orie der eindeutigen analytischen Functionen”. in: Mathematische<br />

Werke von Karl Weierstrass. vol. 2. First published Abhandlungen der Königl.<br />

Akademie der Wissenschaften, 1876. Berlin: Mayer und Müller, 77–124.<br />

Whiteside, D. T. (1972). “James Gregory (1638–1675)”. in: Dictionary <strong>of</strong> Scientific Biography.<br />

ed. by C. C. Gillispie. vol. 5. New York: Charles Scribner’s Sons, 524–530.<br />

Wussing, H. (1969). Die Genesis des abstrakten Gruppenbegriffes. Ein Beitrag zur Entstehungsgeschichte<br />

der abstrakten Gruppentheorie. Berlin: VEB Deutscher Verlag der Wissenschaften,<br />

258.<br />

— (1975). “Evariste Galois”. in: Biographien bedeutender Mathematiker. Köln: Aulis Verlag<br />

Deubner & Co KG. chap. 6, 389–400.<br />

— (1982). “C. F. Gauss und die Tendenzen der Mathematik während der ersten Hälfte<br />

des 19. Jahrhunderts”. in: Impact <strong>of</strong> Bolzano’s epoch on the development <strong>of</strong> science<br />

(Prague, 1981). Prague: ČSAV, 283–304.<br />

Youschkevitch, A. P. (1976). “<strong>The</strong> Concept <strong>of</strong> Function up to the Middle <strong>of</strong> the 19th<br />

Century”. in: Archive for History <strong>of</strong> Exact Sciences 16.1, 37–85.


Index <strong>of</strong> names<br />

Abbati, Pietro (1768–1842), 88, 125<br />

<strong>Abel</strong>, Hans Mathias (1800–1842), 19<br />

<strong>Abel</strong>, Søren Georg (1772–1820), 18<br />

Andersen, Kirsti (⋆1941), 261<br />

Artin, Emil (1898–1962), 185<br />

Badano, P. Gerolamo, 134<br />

Bader, Hans Peter (1790–1819), 19<br />

Bernoulli, Jakob I (1654–1705), 289<br />

Bernoulli, Johann I (1667–1748), 289<br />

Bertrand, Joseph Louis François (1822–1900),<br />

274<br />

Bjerknes, Carl Anton (1825–1903), 18, 332<br />

Bjørling, 279<br />

Bolyai, Farkas (1775–1856), 34<br />

Bolzano, Bernard (1781–1848), 42, 198, 201–<br />

204, 221<br />

Bombelli, Rafael (1526–1572), 61<br />

Bos, Henk J. M., xix<br />

Bottazzini, Umberto (⋆1947), xix, 13, 207<br />

Brill, 374<br />

Brun, Viggo (1885–1978), 261, 375<br />

Burg, Adam von (1797–1882), 123<br />

Cantor, Georg (1845–1918), 212<br />

Cardano, Girolamo (1501–1576), 57, 59, 60,<br />

102, 103<br />

Cauchy, Augustin-Louis (1789–1857), xi, xii,<br />

3, 7–9, 12, 13, 31, 35, 36, 42, 50, 69,<br />

70, 84, 85, 89–95, 100, 101, 108–110,<br />

112, 119–121, 123, 125, 128, 135, 137,<br />

138, 181, 183, 186, 191–193, 196, 201,<br />

206–219, 221–224, 226–231, 233, 235,<br />

237, 238, 241, 246, 248, 251, 253, 260–<br />

263, 267, 275, 278–282, 306, 307, 329,<br />

339, 349, 350, 372, 373, 375, 392–<br />

394, 396, 398, 399<br />

Cayley, Arthur (1821–1895), 132<br />

Chebyshev, Pafnuty Lvovich (1821–1894), 340<br />

Cockle, James, 132<br />

Crelle, August Leopold (1780–1855), 4, 5,<br />

14, 17, 28–37, 39, 41, 44, 73, 98–101,<br />

122, 123, 126–130, 155, 157, 163, 178,<br />

431<br />

180, 206, 218, 223, 226, 241, 252, 261,<br />

265, 277, 278, 299, 300, 337, 341, 346,<br />

376, 377, 396<br />

d’Alembert, Jean le Rond (1717–1783), 43,<br />

58, 199–201<br />

Dedekind, Julius Wilhelm Richard (1831–<br />

1916), 212<br />

Degen, Carl Ferdinand (1766–1825), 22, 23,<br />

25–28, 54, 97, 98, 261, 297, 299<br />

Delambre, Jean-Baptiste Joseph (1749–1822),<br />

43, 44, 84<br />

Descartes, René du Perron (1596–1650), 53,<br />

57<br />

Dirichlet, Gustav Peter Lejeune (1805–1859),<br />

xiv, 73, 219, 228, 236–239, 247, 248,<br />

280, 391, 399<br />

Dirksen, Enno Heeren (1788–1850), 32<br />

du Bois-Reymond, Paul David Gustav (1831–<br />

1889), 248–251<br />

Elliot, 362<br />

Epple, Moritz, 13<br />

Euclid (∼295 B.C.), 79, 159, 348<br />

Euler, Leonhard (1707–1783), xiv, 5, 6, 8, 10,<br />

21, 22, 24, 26, 27, 29, 32, 49, 53, 58,<br />

62–67, 73, 75, 80, 82, 83, 145, 153,<br />

176, 179, 180, 193–197, 203, 207, 209,<br />

214, 216, 221, 238, 240, 261, 285, 287–<br />

292, 297, 303, 306, 328, 329, 388, 389,<br />

393, 397<br />

Eytelwein, Johann Albert (1764–1849), 32<br />

Fagnano dei Toschi, Giulio Carlo (1682–1766),<br />

153, 287, 289, 290<br />

Fermat, Pierre de (1601–1665), 75<br />

Ferrari, Ludovico (1522–1565), 61<br />

Ferro, Scipione (1465–1526), 59<br />

Ferrusac, baron de (1776–1836), 98, 123, 126<br />

Fourier, Jean Baptiste Joseph (1768–1830),<br />

204, 205, 218, 226, 240<br />

Francoeur, Louis Benjamin (1773–1849), 21,<br />

22


432 Index <strong>of</strong> names<br />

Frobenius, Georg Ferdinand (1849–1917), 280<br />

Galilei, Galileo (1564–1642), 24<br />

Galois, Evariste (1811–1832), 5, 10, 49, 52,<br />

54, 55, 66, 67, 75, 82, 84, 97, 125, 137,<br />

138, 161, 180–187, 390, 395<br />

Garnier, Jean Guillaume (1766–1840), 21, 22<br />

Gauss, Carl Friedrich (1777–1855), 3, 6, 10,<br />

21, 23, 27, 32, 34, 49, 51–54, 58, 72–<br />

80, 82–84, 90, 95–97, 124, 126, 139,<br />

142, 150–154, 157–160, 174, 175, 198–<br />

201, 296, 297, 300, 307, 308, 316–<br />

319, 349, 356, 367, 383<br />

Gergonne, Joseph Diaz (1771–1859), 29, 348,<br />

372, 393<br />

Girard, Albert (1595–1632), 57, 59<br />

Goldbach, Christian (1690–1764), 290<br />

Grabiner, Judith V., 197, 393, 397<br />

Grattan-Guinness, Ivor, 42, 222, 279<br />

Gregory, James (1638–1675), 61, 287<br />

Gruson, Johann Philipp (1768–1857), 32<br />

Gårding, Lars, 171<br />

Göpel, Adolph (1812–1847), 377<br />

Hamilton, William Rowan (1805–1865), 104,<br />

130, 131, 133–136, 138<br />

Hansteen, Cathrine Andrea Borch (1787–1840),<br />

17<br />

Hansteen, Christopher (1784–1873), 22–24,<br />

26–28, 33–36, 40, 97, 125, 224, 225,<br />

246, 250, 297, 331, 339, 348<br />

Heegaard, Poul (1871–1948), 375<br />

Heine, Heinrich Eduard (1821–1881), 280<br />

Hilbert, David (1862–1943), 54<br />

Hindenburg, Carl Friedrich (1741–1808), 218<br />

Hirsche, Meier (1765–1851), 98<br />

Holmboe, Bernt Michael (1795–1850), xviii,<br />

4, 14, 17–22, 31, 34–36, 40, 42, 97,<br />

117, 127, 128, 132–134, 138, 160, 161,<br />

163, 175, 179, 221, 222, 225, 226, 260,<br />

297, 298, 306, 375<br />

Holst, Elling Bolt (1849–1915), 26<br />

Humboldt, Alexander von (1769–1859), 32,<br />

34<br />

Humboldt, Wilhelm von (1767–1835), 32, 41<br />

Huygens, Christiaan (1629–1695), 24<br />

Ivory, James (1765–1842), 42<br />

Jacobi, Carl Gustav Jacob (1804–1851), 6, 17,<br />

36, 42, 44, 45, 49, 55, 97, 155, 157,<br />

290, 294, 298, 300, 302, 307, 309, 310,<br />

331–333, 337, 346, 375, 377, 380, 382,<br />

383<br />

Jahnke, Hans <strong>Niels</strong> (⋆1948), 391<br />

Jerrard, George Birch (1804–1863), 129–132,<br />

134, 138, 179<br />

Johnsen, Karsten, 76<br />

Kemp, Christine (1804–1862), 17<br />

Klein, Christian Felix (1849–1925), 391<br />

Kline, Morris (⋆1908), 95<br />

Kronecker, Leopold (1823–1891), 51, 136–138,<br />

149, 152, 171, 181<br />

Kuhn, Thomas S. (1922–1996), 11, 13, 400<br />

Königsberger, Leo (1837–1921), 104, 131, 133,<br />

134, 138, 332<br />

Külp, Edmund Jacob (⋆1801), 22, 115–117,<br />

128, 129, 134, 138, 166<br />

l’Hospital, Guillaume-François-Antoine de<br />

(1661–1704), 358<br />

Lacroix, Sylvestre François (1765–1843), 21,<br />

186, 198<br />

Lagrange, Joseph Louis (1736–1813), xi, 3,<br />

5, 7, 10, 21, 22, 32, 43, 44, 49–51, 53,<br />

58, 61, 64–73, 75, 81–86, 88, 92, 101,<br />

104, 107, 108, 110, 113, 114, 116, 118,<br />

119, 124, 126, 132, 133, 139, 170, 171,<br />

181, 183, 185, 186, 191, 193, 197, 198,<br />

207, 219, 294, 372, 376, 397<br />

Lakatos, Imre (1922–1974), 11–13, 128, 278,<br />

393, 400<br />

Laplace, Pierre-Simon, marquis de (1749–<br />

1827), 20<br />

Laugwitz, Detlef (1932–2000), 248, 391<br />

Legendre, Adrien-Marie (1752–1833), xiv, 3,<br />

6, 24, 26, 27, 35, 36, 44, 75, 126, 127,<br />

153, 288, 292–297, 299, 307, 309, 310,<br />

321, 325, 327, 331, 333, 336, 337, 339,<br />

375<br />

Lehmus, Daniel Christian Ludolf (1780–1863),<br />

32<br />

Leibniz, Gottfried Wilhelm (1646–1716), 58,<br />

61, 62, 287<br />

Libri, Guglielmo (1803–1869), 348, 361, 375<br />

Lie, Marius Sophus (1842–1899), 244, 245,<br />

252, 325, 328, 329, 375<br />

Liouville, Joseph (1809–1882), 52, 181, 185,<br />

236, 340, 374, 382<br />

Littrow, Karl Ludwig von (1811–1877), 35,<br />

123<br />

Lubbe, Samuel Ferdinand (1786–1846), 32<br />

Lubbock, John William (1803–1865), 130


Index <strong>of</strong> names 433<br />

Lützen, Jesper, xix, 340<br />

MacCullagh, James (1809–1847), 134<br />

Malfatti, Gian Francesco (1731–1807), 88, 125<br />

Malmsten, Carl Johann (1814–1886), 171<br />

Maurice, Jean Frédéric Théodor, 127<br />

Mertens, Franz (1840–1927), 247<br />

Mittag-Leffler, Magnus Gustav (1846–1927),<br />

379<br />

Newton, Isaac (1642–1727), 24, 59, 70–72,<br />

195, 203, 286, 287<br />

Ohm, Georg Simon (1789–1854), 31<br />

Ohm, Martin (1792–1872), 31–33, 206, 218<br />

Olivier, Louis, 126, 206, 265–272, 393, 396,<br />

399, 400<br />

Ore, Øystein (1899–1968), 14, 261<br />

Poisson, Siméon-Denis (1781–1840), 21, 186,<br />

206, 226<br />

Popper, Karl Raimund (1902–1994), 12<br />

Rasmussen, Søren (1768–1850), 20, 23, 24,<br />

26, 27, 36, 40, 339<br />

Riemann, Georg Friedrich Bernhard (1826–<br />

1866), xiv, 219, 236, 247, 250, 369,<br />

373, 388, 391<br />

Rosenhain, Johann Georg (1816–1887), 377<br />

Rudio, Ferdinand (1856–1929), 390, 391<br />

Ruffini, Paolo (1765–1822), 5, 10, 50, 53, 54,<br />

83–92, 97, 100, 101, 107, 108, 110,<br />

112, 119–125, 128, 135, 137, 139<br />

Saigey, Jacques Frédéric (1797–1871), 123<br />

Schlesinger, 296<br />

Schmidten, <strong>Henrik</strong> Gerner v. (1799–1831),<br />

29<br />

Schneider, 200<br />

Schumacher, Heinrich Christian (1784–1873),<br />

24, 28, 297, 331, 346<br />

Seidel, Philipp Ludwig von (1821–1896), 217,<br />

279, 280<br />

Serret, Joseph Alfred (1819–1885), 133<br />

Simonsen, Anne Marie (1781–1846), 18<br />

Skau, Christian, 106, 171<br />

Spalt, Detlef, 280<br />

Stokes, George Gabriel (1819–1903), 217, 279,<br />

280<br />

Stubhaug, Arild (⋆1948), xix, 14, 17, 18, 24<br />

Sylow, Peter Ludvig Mejdell (1832–1918), 14,<br />

26, 32, 42, 122, 155, 167, 169, 171,<br />

175, 176, 180, 243, 245, 252, 253, 306,<br />

307, 351, 355, 357, 358, 361, 362, 365,<br />

366, 373<br />

Tartaglia, Niccolò (1499/1500–1557), 59<br />

Taylor, Brook (1685–1731), 202, 222<br />

Thune, Erasmus Georg Fog (1785–1829), 25<br />

Tralles, Johann Georg (1763–1822), 32<br />

Tschirnhaus, Ehrenfried Walter (1651–1708),<br />

61, 80<br />

Vandermonde, Alexandre-Théophile (1735–<br />

1796), 49, 64, 66, 99<br />

Viète, François (1540–1603), 59, 71<br />

Volkert, Klaus, 398<br />

Wallace, William (1768–1843), 196<br />

Wantzel, Pierre Laurent (1814–1848), 89, 135,<br />

247<br />

Waring, Edward (∼1736–1798), 51, 59, 70–<br />

72, 80, 82, 88, 133, 143<br />

Weber, Heinrich (1842–1913), 185<br />

Weierstrass, Karl <strong>The</strong>odor Wilhelm (1815–<br />

1897), 44, 45, 219, 223, 228, 246, 248,<br />

250, 280, 327, 382, 388<br />

Wessel, Caspar (1745–1818), 261<br />

Wussing, Hans, 64, 186<br />

Young, Thomas (1773–1829), 42


Index<br />

École Normale, 40, 187<br />

École Polytechnique, 3, 31, 40, 207, 218<br />

a priori properties <strong>of</strong> roots, 58<br />

<strong>Abel</strong>’s Vienna review, xix<br />

<strong>Abel</strong>’s exception, 395<br />

<strong>Abel</strong>’s manuscripts, xix<br />

<strong>Abel</strong>ian equations, 49, 51, 52, 141, 142, 145,<br />

149, 150, 152, 163, 165, 171, 172, 176,<br />

178, 180, 186, 314, 316, 334, 392<br />

<strong>Abel</strong>ian groups, 149<br />

<strong>Abel</strong>ian <strong>The</strong>orem, 15, 358, 361, 376<br />

absolutely convergent series, 215<br />

Académie des Sciences, 6, 35, 40, 42, 64, 347,<br />

348, 375, 376<br />

Akademie der Wissenschaften, 136<br />

Algebra<br />

Fundamental <strong>The</strong>orem <strong>of</strong>, 5, 10, 58, 80,<br />

82, 95<br />

algebraic expression, 61<br />

algebraically solvable equations, 392<br />

amplitude, 309<br />

anachronism, 10<br />

Analytical Society, 42<br />

Annales de mathématiques pures et appliquées,<br />

29, 205, 348, 372<br />

anomaly, 11<br />

arithmetic-geometric mean, 296, 297<br />

arithmetical foundation <strong>of</strong> calculus, 9<br />

arithmetical concept <strong>of</strong> equality, 9<br />

Astronomische Nachrichten, 24, 331–333, 337,<br />

338<br />

basic curves, 8<br />

basic functions, 381<br />

Berlin Academy, 32<br />

Berlin Academy, 65<br />

binomial formula, 221<br />

binomial theorem, 221<br />

binomial formula, 238<br />

binomial series, 227, 241, 254, 255, 259, 260,<br />

263<br />

435<br />

binomial theorem, 221, 224, 227, 233, 238,<br />

241, 251, 254, 260, 261, 263<br />

British Association, 129<br />

Bulletin, 123, 126<br />

Bulletin des sciences mathématiques, astronomiques,<br />

physiques et chimiques, 98, 123<br />

Cauchy criterion, 395<br />

Cauchy’s <strong>The</strong>orem, 251, 395, 396, 399<br />

Cauchy-ism, 7<br />

Cauchy-Riemann-equations, 307<br />

Cauchy-Ruffini theorem, 100, 101, 108, 110,<br />

112, 119–121, 128, 135, 137<br />

chains <strong>of</strong> roots, 145, 146<br />

chains <strong>of</strong> roots, 142, 143<br />

change, 11<br />

Christiania cathedral school, 4<br />

circle, 287<br />

circle <strong>of</strong> permutations, 93<br />

circular substitution, 94<br />

Collegium academicum, 339<br />

coming to know objects, 8<br />

commutativity <strong>of</strong> rational dependencies, 147<br />

complex <strong>of</strong> roots, 75<br />

complexum, 75<br />

composite permutations, 85<br />

computational mathematical style, 52<br />

concept<br />

definition, 8<br />

concept based mathematics, xiii, xiv, xvii, 8,<br />

9, 13, 15, 55, 124, 141, 187, 192, 286,<br />

387–395, 398–400<br />

concept stretching, 7, 12, 400<br />

conceptual mathematical style, 52<br />

conjugate roots, 76<br />

constructibility, 74<br />

constructions, 74<br />

convergence<br />

d’Alembert, 200, 201<br />

Gauss, 200<br />

Olivier, 265<br />

convergence <strong>of</strong> Fourier series, 399


436 Index<br />

convergence tests, 265<br />

counter example, 7, 12<br />

crisis, 11<br />

critical attitude, 9<br />

critical revision, xiii, xiv, 9, 191, 192, 281,<br />

392<br />

cumulative nature <strong>of</strong> mathematics, 281<br />

cyclotomic equation, 52, 79<br />

cyclotomic equation, 72, 83, 142, 150–152,<br />

316<br />

cyclotomic equations, 150<br />

Danmarks Tekniske Universitet, 24<br />

degree <strong>of</strong> equivalence, 92<br />

degree <strong>of</strong> substitution, 91<br />

degree <strong>of</strong> equivalence, 86, 87<br />

delineation problem, 400<br />

delineation <strong>of</strong> concepts, xiv, 139, 272, 390,<br />

395<br />

delineation <strong>of</strong> solvable equations, 395<br />

delineation problem, xiv, 341<br />

Den Polytekniske Læreanstalt, 24<br />

descent down string <strong>of</strong> equations, 148<br />

dialectics, 11<br />

differentiable functions, 395<br />

Dirichlet boxing-in principle, 109<br />

disciplinary matrix, 400<br />

divergent series, 265, 267<br />

division <strong>of</strong> the lemniscate, 315<br />

division problem, 155, 157, 313<br />

for circular functions, 150<br />

for the lemniscate, 154<br />

for elliptic functions, 142, 150, 153<br />

for the circle, 6, 74, 151, 154<br />

for the lemniscate, 154<br />

<strong>of</strong> the circle, 150<br />

domain <strong>of</strong> rationality, 159<br />

elementary symmetric relations, 59<br />

epistemic configuration, 13<br />

epistemic object, 13<br />

epistemic technique, 13<br />

equality<br />

arithmetical, 9<br />

formal, 9<br />

Euclid’s division algorithm, 142<br />

Euclid’s division algorithm, 141, 158, 159<br />

Euclidean construction, 74<br />

Euler’s hypothesis, 80<br />

exception, 7, 396<br />

exception barring, 12<br />

exceptions, xiii<br />

false roots, 57<br />

falsification, 12<br />

Fermat primes, 79, 318<br />

fictuous roots, 57<br />

form <strong>of</strong> expression, 67<br />

formal concept <strong>of</strong> equality, 9<br />

formula based mathematics, xiii, xiv, 13, 263,<br />

373, 374, 378, 387–392<br />

French Revolution, 3, 40, 43<br />

functional equations, 261<br />

Fundamental <strong>The</strong>orem <strong>of</strong> Algebra, 5, 10, 58,<br />

80, 82, 95, 159<br />

Galois theory, 99<br />

Grand prix, 375<br />

groupe, 75<br />

habituation, xiii, xiv<br />

hyperbola, 285, 287<br />

hyperelliptic functions, 377<br />

hyperelliptic integrals, 350, 351, 376, 377<br />

hypergeometric series, 198–201, 296<br />

identical substitution, 91<br />

imagined roots, 57<br />

imprimitive groups, 86<br />

indecomposable functions, 158<br />

indeterminate series, 265, 267<br />

index, 87<br />

index <strong>of</strong> a function, 92, 93<br />

indicative divisor, 92<br />

inertia <strong>of</strong> mathematical community, 80<br />

inherited property, 148<br />

Institut, 348<br />

Institut de France, 90, 181, 186, 348<br />

intransitive groups, 86<br />

irreducibility, 157<br />

irreducible equations, 159<br />

irreducible <strong>Abel</strong>ian equations, 176<br />

irreducible <strong>Abel</strong>ian equations, 52<br />

irreducible equations, 52, 141, 142, 146, 158<br />

Journal, 33, 35–37, 39, 122, 126, 128, 129, 155,<br />

223, 244, 261, 265, 269, 278, 299, 300,<br />

303, 337, 339, 341, 376, 377<br />

Journal d’École Polytechnique, 90<br />

Journal de mathématiques pures et appliquées,<br />

181, 236<br />

Journal für die reine und angewandte Mathematik,<br />

4, 5, 29, 73, 98, 126<br />

Königlichen Friedrichs-Gymnasium, 135<br />

l’Hospital’s rule, 322


Index 437<br />

Lagrange interpolation, 349, 355, 356, 360,<br />

372, 373, 381<br />

Lagrange’s <strong>The</strong>orem, 68<br />

Lagrangian program, 207<br />

Mémoires, 65<br />

Mémoires de l’Académie des Sciences, 399<br />

Mémoires présentés par divers savants, 348, 375<br />

Maclaurin series, 398<br />

Magazin, 24<br />

Magazin for Naturvidenskaberne, 23, 132<br />

mathematical intuition, 400<br />

mathematical workshop, 13<br />

multi-valued function, 63<br />

Napoleonic Wars, 4<br />

Napoleonic Wars, 40<br />

neo-humanist movement, 3, 41<br />

Nouvelles annales de mathematique, 135<br />

orbit <strong>of</strong> theta, 145<br />

order <strong>of</strong> substitution, 91<br />

paradigm, 11<br />

period <strong>of</strong> roots, 78<br />

sum, 78<br />

permutations, 85, 91, 92<br />

circle, 93<br />

composite, 85<br />

simple, 85<br />

permutazione, 85<br />

Philosophical Magazine, 130, 132<br />

pigeon hole principle, 109<br />

powers <strong>of</strong> a substitution, 91<br />

primitive groups, 86<br />

primitive roots, 73<br />

principle <strong>of</strong> double periodicity, 310<br />

product <strong>of</strong> substitutions, 91<br />

pro<strong>of</strong> analysis, 12<br />

pro<strong>of</strong> revision, 12, 280<br />

Prussia, 3, 41<br />

pure equations, 80<br />

pure equations, 80<br />

quadrature <strong>of</strong> hyperbola, 287<br />

quadrature <strong>of</strong> hyperbola, 285, 287<br />

quadrature <strong>of</strong> circle, 287<br />

quadrature <strong>of</strong> hyperbola, 295<br />

rationally known quantities, 183<br />

reciprocal roots, 76<br />

rectification <strong>of</strong> hyperbola, 287<br />

Regentsen, 23<br />

residues, 306<br />

resolvent, 62<br />

resolvent equation, 62<br />

revolution, 11<br />

rigorization program, 191<br />

roots<br />

false, 57<br />

imagined, 57<br />

true, 57<br />

Royal Danish Academy <strong>of</strong> Sciences and Letters,<br />

26<br />

Royal Danish Academy <strong>of</strong> Sciences and Letters,<br />

25<br />

Royal Danish Academy <strong>of</strong> Sciences and Letters,<br />

24<br />

Royal Danish Academy <strong>of</strong> Sciences and Letters,<br />

22, 97<br />

Royal Irish Academy, 130<br />

Ruffini’s missed subgroups, 88<br />

ruler-and-compass constructibility, 74<br />

Savants étrangers, 361<br />

semblables fonctions, 66<br />

simple permutations, 85<br />

Società Italiana delle Scienze, 84<br />

St. Petersburg Academy, 21, 62<br />

substituions<br />

order, 91<br />

substitutions, 91<br />

circular, 94<br />

degree, 91<br />

identical, 91<br />

powers, 91<br />

product, 91<br />

sum <strong>of</strong> period <strong>of</strong> roots, 78<br />

symmetric group, 72<br />

systems <strong>of</strong> conjugate substitutions, 85<br />

Taylor series, 312, 397<br />

teleology, 10<br />

tests <strong>of</strong> convergence, 400<br />

<strong>The</strong> Enlightenment, 9<br />

Transactions, 130<br />

Transactions <strong>of</strong> the Royal Danish Academy <strong>of</strong><br />

Science, 23<br />

transformation, 11<br />

true value <strong>of</strong> series, 267<br />

Tuberculosis, 4<br />

Videnskabernes Selskab, 22<br />

Zeitschrift für Physik und Mathematik, 35, 123


<strong>RePoSS</strong> issues<br />

#1 (2008-7) Marco N. Pedersen: “Wasan: Die japanische Mathematik der Tokugawa<br />

Ära (1600-1868)” (Masters thesis)<br />

#2 (2004-8) Simon Olling Rebsdorf: “<strong>The</strong> Father, the Son, and the Stars: Bengt Strömgren<br />

and the History <strong>of</strong> Danish Twentieth Century Astronomy” (PhD dissertation)<br />

#3 (2009-9) <strong>Henrik</strong> Kragh Sørensen: “For hele Norges skyld: Et causeri om <strong>Abel</strong> og<br />

Darwin” (Talk)<br />

#4 (2009-11) Helge Kragh: “Conservation and controversy: Ludvig Colding and the imperishability<br />

<strong>of</strong> “forces”” (Talk)<br />

#5 (2009-11) Helge Kragh: “Subatomic Determinism and Causal Models <strong>of</strong> Radioactive<br />

Decay, 1903–1923” (Book chapter)<br />

#6 (2009-12) Helge Kragh: “Nogle idéhistoriske dimensioner i de fysiske videnskaber”<br />

(Book chapter)<br />

#7 (2010-4) Helge Kragh: “<strong>The</strong> Road to the Anthropic Principle” (Article)<br />

#8 (2010-5) Kristian Peder Moesgaard: “Mean Motions in the Almagest out <strong>of</strong> Eclipses”<br />

(Article)<br />

#9 (2010-7) Helge Kragh: “<strong>The</strong> Early Reception <strong>of</strong> Bohr’s Atomic <strong>The</strong>ory (1913-1915): A<br />

Preliminary Investigation” (Article)<br />

#10 (2010-10) Helge Kragh: “Before Bohr: <strong>The</strong>ories <strong>of</strong> atomic structure 1850-1913” (Article)<br />

<strong>#11</strong> (2010-10) <strong>Henrik</strong> Kragh Sørensen: “<strong>The</strong> <strong>Mathematics</strong> <strong>of</strong> <strong>Niels</strong> <strong>Henrik</strong> <strong>Abel</strong>: <strong>Continuation</strong><br />

and New Approaches in <strong>Mathematics</strong> During the 1820s” (PhD dissertation)<br />

#12 (2009-2) Laura Søvsø Thomasen: “In Midstream: Literary Structures in Nineteenth-<br />

Century Scientific Writings” (Masters thesis)<br />

#13 (2011-1) Kristian Danielsen and Laura Søvsø Thomasen (eds.): “Fra laboratoriet til<br />

det store lærred” (Preprint)<br />

#14 (2011-2) Helge Kragh: “Quantenspringerei: Schrödinger vs. Bohr” (Talk)<br />

#15 (2011-7) Helge Kragh: “Conventions and the Order <strong>of</strong> Nature: Some Historical Perspectives”<br />

(Talk)<br />

#16 (2011-8) Kristian Peder Moesgaard: “Lunar Eclipse Astronomy” (Article)<br />

#17 (2011-12) Helge Kragh: “<strong>The</strong> Periodic Table in a National Context: Denmark, 1880-<br />

1923” (Book chapter)<br />

#18 (2012-1) <strong>Henrik</strong> Kragh Sørensen: “Making philosophy <strong>of</strong> science relevant for science<br />

students” (Talk)


<strong>RePoSS</strong> (Research Publications on Science Studies) is a series <strong>of</strong> electronic<br />

publications originating in research done at the Centre for Science Studies,<br />

University <strong>of</strong> Aarhus.<br />

<strong>The</strong> publications are available from the Centre homepage<br />

(��������������������).<br />

ISSN 1903-413X<br />

Centre for Science Studies<br />

University <strong>of</strong> Aarhus<br />

Ny Munkegade 120, building 1520<br />

DK-8000 Aarhus C<br />

DENMARK<br />

�������������<br />

Faculty <strong>of</strong> Science<br />

Department <strong>of</strong> Science Studies<br />

1

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!