11.09.2019 Views

Historical Painting Techniques, Materials, and Studio Practice

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

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

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>,<br />

<strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


PUBLICATIONS COORDINATION: Dinah Berl<strong>and</strong><br />

EDITING & PRODUCTION COORDINATION: Corinne Lightweaver<br />

EDITORIAL CONSULTATION: Jo Hill<br />

COVER DESIGN: Jackie Gallagher-Lange<br />

PRODUCTION & PRINTING: Allen Press, Inc., Lawrence, Kansas<br />

SYMPOSIUM ORGANIZERS:<br />

Erma Hermens, Art History Institute of the University of Leiden<br />

Marja Peek, Central Research Laboratory for Objects of Art <strong>and</strong> Science,<br />

Amsterdam<br />

© 1995 by The J. Paul Getty Trust<br />

All rights reserved<br />

Printed in the United States of America<br />

ISBN 0-89236-322-3<br />

The Getty Conservation Institute is committed to the preservation of<br />

cultural heritage worldwide. The Institute seeks to advance scientiRc<br />

knowledge <strong>and</strong> professional practice <strong>and</strong> to raise public awareness of<br />

conservation. Through research, training, documentation, exchange of<br />

information, <strong>and</strong> ReId projects, the Institute addresses issues related to<br />

the conservation of museum objects <strong>and</strong> archival collections,<br />

archaeological monuments <strong>and</strong> sites, <strong>and</strong> historic bUildings <strong>and</strong> cities.<br />

The Institute is an operating program of the J. Paul Getty Trust.<br />

COVER ILLUSTRATION<br />

Gherardo Cibo, "Colchico," folio 17r of Herbarium, ca. 1570. Courtesy<br />

of the British Library.<br />

FRONTISPIECE<br />

Detail from Jan Baptiste Collaert, Color Olivi, 1566-1628. After<br />

Johannes Stradanus. Courtesy of the Rijksmuseum-Stichting,<br />

Amsterdam.<br />

Library of Congress Cataloguing-in-Publication Data<br />

<strong>Historical</strong> painting techniques, materials, <strong>and</strong> studio practice :<br />

preprints of a symposium [held at] University of Leiden, the<br />

Netherl<strong>and</strong>s, 26-29 June 1995/ edited by Arie Wallert, Erma<br />

Hermens, <strong>and</strong> Marja Peek.<br />

p. cm.<br />

Includes bibliographical references.<br />

ISBN 0-89236-322-3 (pbk.)<br />

1. <strong>Painting</strong>-<strong>Techniques</strong>-Congresses. 2. Artists' materials-<br />

-Congresses. 3. Polychromy-Congresses. I. Wallert, Arie,<br />

1950- II. Hermens, Erma, 1958- . III. Peek, Marja, 1961-<br />

ND1500.H57 1995<br />

751' .09-dc20 95-9805<br />

CIP<br />

Second printing 1996<br />

iv


Contents<br />

vii<br />

viii<br />

Foreword<br />

Preface<br />

1 Leslie A. Carlyle, Beyond a Collection of Data: What We Can<br />

Learn from Documentary Sources on Artists' <strong>Materials</strong> <strong>and</strong><br />

<strong>Techniques</strong><br />

6 Sylvana Barrett, Dusan C. Stulik, An Integrated Approach for the<br />

Study of <strong>Painting</strong> <strong>Techniques</strong><br />

12 Helen Glanville, Varnish, Grounds, Viewing Distance, <strong>and</strong><br />

Lighting: Some Notes on Seventeenth-Century Italian <strong>Painting</strong><br />

Technique<br />

20 Ann Massing, From Books of Secrets to Encylopedias: <strong>Painting</strong><br />

<strong>Techniques</strong> in France between 1600 <strong>and</strong> 1800<br />

30 Sally A. Woodcock, The Roberson Archive: Content <strong>and</strong><br />

Significance<br />

38 Arie Wallert, Libro Secondo de Diversi Colori e Sise da Mettere a<br />

Oro: A Fifteenth-Century Technical Treatise on ManUSCript<br />

Illumination<br />

48 Erma Hermens, A Seventeenth-Century Italian Treatise on<br />

Miniature <strong>Painting</strong> <strong>and</strong> Its Author(s)<br />

58 Beate Federspiel, Questions about Medieval Gesso Grounds<br />

65 Renate Woudhuysen-Keller, Aspects of <strong>Painting</strong> Technique in the<br />

Use of Verdigris <strong>and</strong> Copper Resinate<br />

70 Josephine A. Darrah, Connections <strong>and</strong> Coincidences: Three<br />

Pigments<br />

78 Kate I. Duffy, Jacki A. Elgar, An Investigation of Palette <strong>and</strong> Color<br />

Notations Used to Create a Set of Tibetan Thangkas<br />

85 Zuzana Skalova, New Evidence for the Medieval Production of<br />

Icons in the Nile Valley<br />

91 Helen C. Howard, <strong>Techniques</strong> of the Romanesque <strong>and</strong> Gothic Wall<br />

<strong>Painting</strong>s in the Holy Sepulchre Chapel, Winchester Cathedral<br />

105 Eddie Sinclair, The Polychromy of Exeter <strong>and</strong> Salisbury<br />

Cathedrals: A Preliminary Comparison<br />

111 Andrea Rothe, Andrea Mantegna's Adoration of the Magi<br />

117 Ulrich Birkmaier, Arie Wallert, Andrea Rothe, Technical<br />

Examinations of Titian's Venus <strong>and</strong> Adonis: A Note on Early<br />

Italian Oil <strong>Painting</strong> Technique<br />

127 Catherine A. Metzger, Barbara H. Berrie, Gerard David's St.<br />

Anne Altarpiece: Evidence for Workshop Participation<br />

v


135 Molly Faries, Christa Steinbuchel, <strong>and</strong> J. R. J. van Asperen de<br />

Boer, Maarten van Heemskerck <strong>and</strong> Jan van Scorel's Haarlem<br />

Workshop<br />

140 E. Melanie Gifford, Style <strong>and</strong> Technique in Dutch L<strong>and</strong>scape<br />

<strong>Painting</strong> in the 1620s<br />

148 J¢rgen Wadum, Johannes Vermeer (1632-1675) <strong>and</strong> His Use of<br />

Perspective<br />

155 Ilze Poriete, Dace Choldere, A Technical Study of the <strong>Materials</strong><br />

<strong>and</strong> Methods Used by the Painters of the Latvian Churches in the<br />

Seventeenth Century<br />

158 Melissa R. Katz, William Holman Hunt <strong>and</strong> the «Pre-Raphaelite<br />

Technique"<br />

166 Jo Kirby, Ashok Roy, Paul Delaroche: A Case Study of Academic<br />

<strong>Painting</strong><br />

176 Joyce H. Townsend, <strong>Painting</strong> <strong>Techniques</strong> <strong>and</strong> <strong>Materials</strong> of Turner<br />

<strong>and</strong> Other British Artists 1775-1875<br />

186 Stephen Hackney, Art for Art's Sake: The <strong>Materials</strong> <strong>and</strong><br />

<strong>Techniques</strong> of James McNeill Whistler (1834-1903)<br />

191 John R. Gayer, <strong>Painting</strong> on a Photographic Substrate: Notes<br />

Regarding <strong>Materials</strong> <strong>and</strong> <strong>Techniques</strong> over the Past 100 Years<br />

196 Ernst van de Wetering, Reflections on the Relation between<br />

Technique <strong>and</strong> Style: The Use of the Palette by the Seventeenth­<br />

Century Painter<br />

204 Index of Contributors<br />

vi


Foreword<br />

One of the first events organized by the Getty Conservation Institute as<br />

it began its activities about ten years ago was a symposium on paintings<br />

conservation. We felt at that time that our particular approach, based on<br />

multiple disciplines looking at a single problem, could contribute significantly<br />

to the field of conservation. Over the course of the past years we<br />

have continued to develop in that direction <strong>and</strong> with that belief, so it is<br />

particularly appropriate to see these important preprints come to light for<br />

the University of Leiden's symposium.<br />

The history of painting techniques is by nature a multidiSCiplinary area<br />

of study, combining research in science, conservation, <strong>and</strong> art history as<br />

well as specific expertise in paintings. Members of each one of these professions<br />

bring to the area their own detailed knowledge in artists' materials,<br />

techniques, or methods, whether it be information specific to pigments,<br />

binding media, signature style, or archival research.<br />

The field of history of painting techniques has been evolving very rapidly<br />

in recent years <strong>and</strong> opening enormous opportunities for further research<br />

<strong>and</strong> connOisseurship, as this important group of papers demonstrates. Not<br />

only scientific methods of examination but also bibliographical <strong>and</strong> archival<br />

research are making significant contributions.<br />

The authors of these preprints bring a wide array of expert knowledge<br />

as well as many fresh points of view that are certain to provoke serious<br />

questions <strong>and</strong> debate. The editors-Arie Wallert, Erma Hermens, <strong>and</strong><br />

Marja Peek-have assembled this volume with a most inSightful approach<br />

<strong>and</strong> focus. So it is with a great deal of satisfaction <strong>and</strong> pleasure that we<br />

present these papers for the study, <strong>and</strong> enjoyment, of a fascinating <strong>and</strong><br />

challenging field.<br />

Miguel Angel Corzo, Director<br />

The Getty Conservation Institute<br />

vii


Preface<br />

This volume of preprints, prepared for an international symposium on<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong> at the University<br />

of Leiden, the Netherl<strong>and</strong>s, 26-29 June 1995, contains the results<br />

of work on historical painting techniques from all parts of the world. The<br />

suggestion to organize such a meeting was raised during the symposium<br />

on the Technology of Art Works from the Central European Region organized<br />

by the Archives of Art Technology in Prague in 1993. The Prague<br />

symposium emphasized Czech painters <strong>and</strong> their techniques. To broaden<br />

the scope of attention, Erma Hermens of the Art History Institute of the<br />

University of Leiden <strong>and</strong> Marja Peek of the Art <strong>Historical</strong> Department of<br />

the Central Research Laboratory for Objects of Art <strong>and</strong> Science in Amsterdam<br />

took on the task of organizing a second meeting in Leiden.<br />

The purpose of the symposium <strong>and</strong> this publication is to promote a<br />

greater underst<strong>and</strong>ing of the changing boundaries <strong>and</strong> interaction between<br />

art historians, conservators, <strong>and</strong> conservation scientists working in the fields<br />

of historical painting techniques-including wall paintings <strong>and</strong> polychrome<br />

sculpture-painting materials, <strong>and</strong> studio practice.<br />

In recent years, there has been an increasing interest in historical painting<br />

techniques. The study of the painting techniques <strong>and</strong> materials used<br />

throughout history <strong>and</strong> in various cultures is by nature an interdisciplinary<br />

exercise. In the past such studies were sometimes conducted with little<br />

interaction between art historians, conservators, materials scientists, <strong>and</strong><br />

historians of science, because each discipline tends to present the results<br />

of its studies to different forums. This volume aims to present different<br />

approaches to the study of historical painting techniques in the hope that<br />

it will encourage cooperation among these various disciplines.<br />

Information about painting techniques can be gained in a variety of<br />

ways, including the chemical or physical analyses of the materials found<br />

in the paintings. Analyses of a large number of paintings attributed to<br />

certain regions, schools, workshops, or individual masters can contribute<br />

to a history of painting techniques. Analytical results can also help art<br />

historians assess attributions <strong>and</strong> can support or reject their hypotheses.<br />

The analysts, however, need the art historians to inform them about the<br />

stylistic idiosyncrasies <strong>and</strong> significance of those schools, workshops, or masters.<br />

It is often rewarding to see what artists have said about their own work<br />

<strong>and</strong> to study their written sources on painting techniques. This kind of<br />

information has come down to us in diaries, such as Neri di Bicci's Ricordanze;<br />

in painting h<strong>and</strong>books such as Cennini's Libra dell'Arte, in anonymous<br />

recipe books; <strong>and</strong> even in model books, such as Stephan Schriber's<br />

Musterbuch. These recipe books tell scientists what substances to look for<br />

in analysis. Both scientific <strong>and</strong> art historical information help us underst<strong>and</strong><br />

the significance of sometimes rather obscure recipes <strong>and</strong> tell us whether<br />

the methods <strong>and</strong> materials described are common or exceptional for the<br />

particular period. Today it is not unusual to find information from historical<br />

sources incorporated in the examination of individual paintings.<br />

viii


The discussions in this volume present historical painting techniques<br />

from a variety of professional perspectives. With its wide range of topics<br />

<strong>and</strong> approaches to the study of historical painting techniques, this publication<br />

is presented in the hope that it will provide an impetus for further<br />

studies that involve material science, art history, conservation, archaeometry,<br />

<strong>and</strong> the history of science. We also hope it will be one in a series<br />

of such interdisciplinary <strong>and</strong> collaborative volumes.<br />

In addition to thanking all the colleagues at our institutions who helped<br />

us with advice <strong>and</strong> expertise, we also would like to acknowledge the<br />

invaluable work <strong>and</strong> support of several individuals in particular. Miguel<br />

Angel Corzo, director of the Getty Conservation Institute, immediately<br />

supported the idea to publish the presentations of the symposium <strong>and</strong><br />

generously provided his time <strong>and</strong> support in the production of the present<br />

volume. Agnes Grafin Ballestrem, director of the Central Research Laboratory<br />

for Objects of Art <strong>and</strong> Science in Amsterdam, <strong>and</strong> A. W. A. Boschloo<br />

of the Art History Institute of the University of Leiden supported<br />

the entire project from its initial stages <strong>and</strong> advised us during the editorial<br />

process. Corinne Lightweaver <strong>and</strong> Jo Hill, independent editorial consultants,<br />

carried out invaluable work on the manuscripts, assisted by Joy Hartnett.<br />

In the preparation of these preprints for publication, we also wish to<br />

thank Dinah Berl<strong>and</strong>, publications coordinator, the Getty Conservation<br />

Institute.<br />

Arie Wallert<br />

The Getty Conservation Institute<br />

Erma Hermens<br />

Art History Institute of the University of Leiden<br />

Marya Peek<br />

Central Research Laboratory for Objects of Art <strong>and</strong> Science, Amsterdam<br />

ix


Plate la, left. Bartolomeo Vivarini, active 145149 1,<br />

Saint Peter, 1490. Tempera on partel. From Polyptych<br />

with Saint James the Greater, The Virgin <strong>and</strong> Child<br />

<strong>and</strong> Other Saints. The]. Paul Getty Museum, Malibu<br />

(71.PB.30).<br />

Plate 1 b, above. Detail from Vivarini's Saint Peter, showing<br />

clearly visible, individual brush strokes.


Plate 2a, above right. Attributed to Raphael, 1483- 1520, Portrait<br />

of a Young Man, 1505. Oil on panel. The J. Paul Getty<br />

Museum, Malibu (78. PB. 364}.<br />

Plate 2b, right. Detail from Portrait of a Young Man, illustrating<br />

the clear, luminous color of the transparent oil techniques.


Plate 3a, above. Peter Paul<br />

Rubens, 1577-1640, The<br />

Virgin as the Woman of<br />

the Apocalypse, 1623.<br />

Oil on wood. The]. Paul<br />

Getty Museum, Malibu<br />

(85.PB. 146).<br />

Plate 3b, right. Detail from<br />

The Virgin as the Woman<br />

of the Apocalypse, illustrating<br />

the use of impasto<br />

white under highlighted<br />

areas.


Plate 4a, above. Jacques­<br />

Louis David, 1748-1825,<br />

The Sisters Zenalde <strong>and</strong><br />

Charlotte Bonaparte,<br />

1821. Oil on canvas. The<br />

J. Paul Getty Museum,<br />

Malibu (86.PA. 740).<br />

Plate 4b, right. Detail Jrom<br />

The Sisters Zenalde <strong>and</strong><br />

Charlotte Bonaparte, illustrating<br />

the imperceptible<br />

shift from dark to light tones<br />

resulting from physically<br />

blending pigments on the<br />

suiface


Plate 5, hft. Colorjading tests, 1837-<br />

1878. Fitzwilliam Museum, UnilJwil)'<br />

of Cambridge.<br />

Plate 6, above. Pigmelll {(llilailicrs: j)"illl<br />

bladder to collapsible Illbe. Fil':'lI'illi,1/1/<br />

Museum, Universil), of Clll//Jri(!c.


Plate 7a, above. Valerio<br />

Mariana da Pesaro, Battle<br />

of San Fabiano, 1618-<br />

1620. Photograph by E.<br />

Buzzegoli, Laboratorio di<br />

Restauro. Courtesy of the<br />

Ujfizi Gallery, Florence.<br />

Plate 7b, right. Detail<br />

from Battle of San Fabiano,<br />

1618-1620, showing<br />

the l<strong>and</strong>scape in the<br />

background built up with<br />

several transparent washes<br />

of color. Photograph by E.<br />

Buzzegoli, Laboratorio di<br />

Restauro. Courtesy of the<br />

Ujfizi Gallery, Florence.


Plate 8. Gherardo Cibo, "Colchico, " from Herbarium (MS ADD 22333), jo lio 17r, ca.<br />

1570. Courtesy oj the British Library.


Plate 9. Marco Palmezzano, Mystic Marriage of Saint Catherine, signed <strong>and</strong> dated<br />

1537. Photograph by Christopher Hurst. Private collection.


Plate 10. Detail from Marco<br />

Palmezzano's Mystic Marriage<br />

of Saint Catherine,<br />

showing green glaze dabbed on<br />

St. Catherine's garment. Area<br />

shown measures approx. 3 X<br />

4 cm. Photograph by Christopher<br />

Hurst. Private collection.<br />

Plate 11. Cross section from PalmezzarLO's Mystic Marriage<br />

of Saint Catherine, showing of the green baldachin<br />

of the throne in normal light (X200). Photograph by Christopher<br />

Hurst. Private collection.<br />

Plate 12. Cross section from Palmezzano's Mystic Marriage<br />

of Saint Catherine, showing the oiling-out layer between<br />

the underpainting <strong>and</strong> the final glaze in UV light<br />

(X220). Photograph by Christopher Hurst. Private collection.


Plate 13, right. Surya<br />

in his Chariot, ca.<br />

1770, Bundi miniature.<br />

Smalt in the blue sky.<br />

Photograph by Paul<br />

Robins (Photo <strong>Studio</strong>),<br />

courtesy oj the Victoria &<br />

Albert Museum (D.379-<br />

1889).<br />

Plate 14, below. Krishna<br />

<strong>and</strong> Girls, 1730-<br />

1735, Basohli mInIature.<br />

Smalt with high arsenic<br />

content in the sky<br />

<strong>and</strong> Krishna's skin. Photograph<br />

by Paul Robins<br />

(Photo <strong>Studio</strong>), courtesy<br />

if the Victoria & Albert<br />

Museum (I.M. 87-1930).


Plate 15. The Buddha Shakyamuni<br />

Preaching at Dhanyakataka.<br />

Glue bound paint on cotton, 80.4 X<br />

3 cm. Denman Waldo Ross Collection,<br />

Museum of Fine Arts, Boston<br />

(06.333).


Plate 16. The Fifth King of<br />

Shambhala. Clue bound paint on<br />

cotton, 80.9 X 43.3 em. Denman<br />

Waldo Ross Collection, Museum of<br />

Fine Arts, Boston (06.324).


Plate 17, top. Icon of Virgin with Child Enthroned<br />

between Archangels, Nine Church<br />

Fathers <strong>and</strong> Nine Coptic Monks in the church<br />

of St. Mercurius Abu's-Saifain. This photograph<br />

documents its state of preservation in 1990. Photograph<br />

by Ayman el-Kranat, courtesy if the Archive<br />

of the church of St. Mercurius Abu's-Saifain.<br />

Plate 18a, b, above lift <strong>and</strong> right. Icon of Six<br />

Equestrian Saints in the church of St. Mercurius<br />

Abu's-Saifa in, lift <strong>and</strong> right sides, showing the<br />

beam's state of preservation in 1990. Photograph<br />

by Ayman el-Kranat, courtesy of the Archive of<br />

the church of St. Mercurius Abu's-Saifain.<br />

Plate 19, right. Detail of the head of Joseph of<br />

Aramathea from Deposition on the upper tier of<br />

the east wall. Photograph courtesy of the Conservation<br />

of Wall <strong>Painting</strong> Department, Courtauld<br />

Institute of Art, London.


Plate 20. Cross section oj Sample 3, taken<br />

Jrom the olive-green background near the leg<br />

oj Nicodemus in the Deposition on the<br />

east wall. The charcoal black inclusions<br />

within the plaster substrate are clearly visible;<br />

over this, a layer oj natural ultramarine<br />

is present. The present green appearance<br />

is due to the layer oj vivianite on the<br />

suiface oj the sample. Originally a blue color,<br />

some oj the particles have altered to a<br />

yellow color, <strong>and</strong> in the center oj some oj<br />

the yellow particles, the original blue color<br />

is just discernible. Photograph courtesy oj<br />

the Conservation oj Wall <strong>Painting</strong> Department,<br />

Courtauld Institute oj Art, London.<br />

Plate 21. Photomicrograph oj sample from<br />

central porch tympanum, Salisbury. Cross<br />

section shows complex structure, representing<br />

several repaintings <strong>and</strong> including vermilion,<br />

red <strong>and</strong> white lead, iron-oxide red<br />

<strong>and</strong> yellow, black, blue, <strong>and</strong> possibly gold.<br />

Plate 22. Photomicrograph oj sample Jrom<br />

Exeter west Jront. Cross section shows pink<br />

primer, white lead, <strong>and</strong> verdigris top layer.


Plate 23, top. Andrea Mantegna, The<br />

Adoration of the Magi. Distemper on<br />

fine linen, 54.6 X 69 .2 cm. The). Paul<br />

Getty Museum, Malibu (85.PA.417).<br />

Plate 24, lift. Detail from The Adoration<br />

of the Magi, upper lift-h<strong>and</strong> comer,<br />

showing feathered-out colors on the edges<br />

of the painting, originally under theframe.<br />

The ). Paul Getty Museum, Malibu<br />

(85.PA.417).<br />

Plate 25, above. Detail from The Adoration<br />

of the Magi, lower edge, showing<br />

feathered-out colors on the edges of the<br />

painting, as in Plate 24.


Plate 26, top. Titian, Venus <strong>and</strong> Adonis. 160 X 196. 5 em. The] . Paul Getty Museum,<br />

Malibu (92.PA.42).<br />

Plate 27, bottom. Titian, Venus <strong>and</strong> Adonis. 107 X 136 em. Courtesy of the National<br />

Gallery of Art, Washington, D.C., Widener Collection (1942.9.84 [PAJ).


Plate 28. Gerard David <strong>and</strong> Workshop, St. Anne Altarpiece, ca. 1506: lift, St. Nicholas, 236 X 76 cm; center, St. Anne with the<br />

Virgin <strong>and</strong> Child, 236 X 96 cm; right, St. Anthony of Padua, 235 X 76 cm. Courtesy of the National Gallery of Art, Washington,<br />

D. C. Widener Collection.


Plate 29. Cross sections of the St. Anne Altarpiece, shown in Plate 28. Cross section (a)<br />

through the underdrawing line below the cherub's foot (photographed at X100) shows fo ur<br />

layers: ground, underdrawing, pale gray, <strong>and</strong> blue-gray. Cross section (b) through a line in<br />

St. Anne's red drapery (photographed at Xl 00) shows multiple layers: ground, type III<br />

underdrawing, two layers of orange red, <strong>and</strong> multiple layers of red glaze (the layering is not<br />

observable in normal light). Cross section (c) through the "correction" layer in the Virgin's<br />

sleeve (photographed at X 100) shows five layers: ground, underdrawing, pale underpaint,<br />

"correction" layer, <strong>and</strong> dark violet glaze. Sampling by B. Miller. Cross section (d) through<br />

the Virgin's blue drapery (photographed at Xl 00) shows fo ur layers: ground, underdrawing,<br />

pale blue, <strong>and</strong> dark blue. Cross section (e) through the dark green part of the cloth of honor<br />

shows seven layers: ground, scattered black particles, blue-green, dark green, dark-green <strong>and</strong><br />

green glazes, <strong>and</strong> discolored varnish. Cross section (f) through St. Anthony's cassock (photographed<br />

at Xl 00) shows three layers: ground, underdrawing, <strong>and</strong> violet-gray. Sampling by<br />

B. Miller.


Plate 30. Cross section from Baptism,<br />

near an edge: layer 1 (at the bottom), intermediate<br />

white layer, here thicker; layer<br />

2, pink layer with madderlike particles;<br />

layer 3 (on top), white with natural ultramarine.<br />

Photograph by]. R. J. van Asperen<br />

de Boer.<br />

Plate 31. Paint cross section, light green<br />

Jo liage, upper part oj central trees. Workshop<br />

oj Jan Bruegel, Noah's Ark. Layers<br />

Jrom bottom: (1) chalk ground; (2) imprimatura:<br />

red <strong>and</strong> black in a translucent medium;<br />

(3) sky: white lead, pale smalt,<br />

black; (4) Joliage underpaint: azurite,<br />

earth, white lead, black. Magnifrcation on<br />

35 mmfilm: X64.<br />

Plate 32. Paint cross section (dark sky at<br />

upper edge) from Esaias van de Velde's<br />

Winter L<strong>and</strong>scape. Layers Jrom bottom:<br />

(1) ground: chalk with white lead, earth,<br />

<strong>and</strong> black; (2) sky: white lead, smalt, <strong>and</strong><br />

earth black; (3) darker sky: smalt <strong>and</strong><br />

white lead; (4) overpaint. Magnifrcation<br />

on 35 mm film: X70.


Plate 33. Paul Delaroehe, The Execution of Lady Jane Grey, 1833. Canvas, 246 X 29 7 em. London, National Callery (1909).


Plate 34a, b, c, d. Paint sampLes Jrom<br />

The Execution of Lady Jane Grey: (a)<br />

cross section, pinkish red oj executioner's<br />

tights, original magnification X540; (b)<br />

cross section, grayish shadow on Lady Jane's<br />

dress, original magnification X750;<br />

(c) cross section, red <strong>and</strong> brown brocade<br />

draped over Lap if seated attendant Left,<br />

originaL magnification X750; (d) unmounted<br />

fragment, bLack oj Sir John Brydges's<br />

gown, photographed from the reverse<br />

at X275.


Plate 35, right. Detail Jrom the lower lifth<strong>and</strong><br />

comer oj J. 1'1'1.. W. Turner's The<br />

Dawn of Christianity shows where Turner<br />

used the canvas to try out colors.<br />

Plate 36, below. Detail oj sheep Jrom<br />

William Holman Hunt's Strayed Sheep<br />

(Our English Coasts), 1852. 432 X<br />

584 mm. Courtesy oj the Tate Gallery,<br />

London (N05665).


Plate 37, lift. James Whistler, Arrangement in<br />

Flesh Color <strong>and</strong> Black: Portrait of Theodore<br />

Duret, 1883-1884, oil on canvas. Wolfe Fund,<br />

Catharine Lorillard Wolfe Collection, 1913<br />

(13.20). © 1984 by The Metropolitan Museum<br />

if Art.<br />

Plate 38, above. Detail of a reflection from Whistler's<br />

Nocturne: Blue <strong>and</strong> Silver, showing that<br />

the paint has been dragged horizontally when both<br />

the white overpaint <strong>and</strong> blue underlayer were still<br />

wet, to produce the iffect if light on the water surface.


Abstract<br />

This paper argues for an interdisciplinary<br />

approach to the study of artists'<br />

painting materials <strong>and</strong> painting<br />

practices. Recent research into British<br />

documentary sources on nineteenth-century<br />

oil painting reveals<br />

information useful in technical examinations<br />

of paintings. Examples<br />

illustrate how important a full underst<strong>and</strong>ing<br />

of the artist's physical<br />

environment <strong>and</strong> contemporary beliefs<br />

can be in accurately interpreting<br />

evidence from a painting.<br />

Beyond a Collection of Data: What We Can Learn<br />

from Documentary Sources on Artists' <strong>Materials</strong><br />

<strong>and</strong> <strong>Techniques</strong><br />

Leslie A. Carlyle<br />

Canadian Conservation Institute<br />

1030 Innes Road<br />

Ottawa, Ontario KiA OC8<br />

Canada<br />

Introduction<br />

In the preface to the catalogue for the National Gallery of London's exhibition,<br />

Art in the Making: Impressionism, the sponsor remarks, "We see not<br />

just painted surfaces, but are given a multi-disciplined information which<br />

brings the paintings themselves to life" (1). The desire to know more about<br />

the whole object, to go beyond the image, the preface continues, to get<br />

"behind the pictures, <strong>and</strong> even through them," is very much a feature of our<br />

era, just as the delight in finding the "real" information hiding behind our<br />

conventional views fu els so many contemporary investigative endeavors in all<br />

disciplines. Our enthusiasm for complete knowledge is something we have<br />

in common with our predecessors of the late eighteenth century.<br />

Fresh from the age of enlightenment, one author of a late eighteenth-century<br />

technical manual on oil painting writes in his preface (2):<br />

Facts judiciously arranged, <strong>and</strong> published from time to time as they accumulate,<br />

are productive of infinite advantage . ... Every branch of science<br />

is much fa cilitated <strong>and</strong> advanced by public communication, which distinctly<br />

points out the present, <strong>and</strong> opens a free channel to future discoveries . ...<br />

Records oj this kind act, therefore, as stimulants to general improvement:<br />

what is already known need not be retraced, <strong>and</strong> what is discovered in<br />

future [sic] may be occasionally added: thus, the needy <strong>and</strong> diffident will<br />

be taught with oeconomy [sic] <strong>and</strong> ease, <strong>and</strong> mystery will be unfolded <strong>and</strong><br />

converted into truth.<br />

Few today would argue for such a completely linear view of knowledge, but<br />

the desire to "know all" has not left us.<br />

Nineteenth-century technical literature on oil painting materials <strong>and</strong> practices<br />

shows that the search for <strong>and</strong> collection of the "facts" was underway in earnest<br />

at that time. By the end of the century, however, we find the optimism<br />

somewhat chastened: all the new chemical knowledge, all the new facts, still<br />

could not provide oil painters with any guarantees, once <strong>and</strong> fo r all, for the<br />

durability of their work.<br />

A similar sort of optimism existed for those in the twentieth century who<br />

examined the role of science in unraveling the mysteries of oil paintings. The<br />

belief was that with enough instruments <strong>and</strong> enough analysis, we could know<br />

the secrets of the old masters, we could know of what a painting is made.<br />

However, as one of our discipline's critics Michael Daly points out, "Such<br />

technical analysis can only ever say what a material is, never what its purpose<br />

was" (3).<br />

Although scientific instrumental analysis is a highly sophisticated branch of<br />

inquiry in itself, results from it alone are not sufficient. It is only in partnership<br />

with other forms of investigation that we can hope to unravel the meaning<br />

behind what we find through analysis. Now, nearing the end of the twentieth<br />

century, having penetrated much of the "mystery" <strong>and</strong> converted it into truth,<br />

we find that the gifts of science are not enough. It is to "multidisciplinary<br />

information" that we turn in order to underst<strong>and</strong> the purpose of the materials<br />

we find.<br />

Carlyle 1


Fortunately, the spirit of the eighteenth century in Britain fired a great enthusiasm<br />

for publications on technical matters, including the materials <strong>and</strong><br />

techniques of oil painting. Thus, we find a rich source of information in the<br />

various treatises, manuals, <strong>and</strong> h<strong>and</strong>books that continued to be published into<br />

the nineteenth century. The following will be a discussion of the kind of<br />

information these documentary sources can provide <strong>and</strong> how this information<br />

can influence our interpretation of cross sections <strong>and</strong> analytical results as well<br />

as fu rther our underst<strong>and</strong>ing of painters' techniques.<br />

The painter's environment<br />

The anonymous eighteenth-century author quoted previously placed great<br />

faith in "facts judiciously arranged, <strong>and</strong> published from time to time as they<br />

accumulate." Yet, however important a discrete piece of information, such as<br />

the date of introduction of a new pigment, may be in the study of painters'<br />

instruction books, it is not always this information that provides insight into<br />

the painter's choice of materials. Sometimes it is the tangential information<br />

about the experience of living at a given time that provides a context fo r<br />

what we observe now.<br />

In cross sections of paint, the build-up of dirt between layers of paint must<br />

be interpreted in relation to past conditions for lighting <strong>and</strong> heating. We<br />

cannot, based on our present-day experience, extrapolate from the thickness<br />

of a dirt layer the length of time between episodes of painting. Here, those<br />

who study the history of technology <strong>and</strong> of domestic life are of great help.<br />

In one source on the history of domestic environments, we find that even as<br />

early as 1700 the use of coal fo r heating in London resulted in a "Tartanous<br />

Smoak" that sullied the environment both indoors <strong>and</strong> out: "All sorts of<br />

Hangings, especially the Tapestry, are in a few Years totally defil'd by it ..."(4).<br />

Because painters who fo llowed the technique of "painting in stages" were<br />

obliged to wait between applications of paint for the underlayers to dry, a<br />

fairly rapid build-up of dirt could be expected under the conditions described,<br />

far more than our late twentieth-century environments would convey.<br />

The level of air pollution in the days of coal heating also caused great concern<br />

among artists <strong>and</strong> their chemist advisors with regard to the role of lead in<br />

paintings. It was thought that the high levels of sulfur in the air caused reactions<br />

with lead-white pigment <strong>and</strong> with lead-treated oil, resulting in an<br />

overall darkening of these materials due to the reaction product, lead sulfide.<br />

Various solutions to this problem, including the application of nonreactive<br />

zinc white over lead-white underlayers, were recommended. This advice to<br />

apply zinc white over lead white was given not only for paint layers, but fo r<br />

grounds as well; it was believed that a lead-white ground preparation could<br />

also be subject to darkening. Cross sections taken from a painting in which<br />

this advice had been followed show layers of two different white paints, the<br />

presence of which would not be immediately obvious without the knowledge<br />

of the remedial steps taken to obviate the so-called lead-sulfide darkening (5).<br />

Aside from the dirt <strong>and</strong> soot from coal heat <strong>and</strong> tallow c<strong>and</strong>les, interior<br />

environments were also substantially colder in the winter months. In the<br />

absence of central heating, painters found that their colors dried significantly<br />

more slowly during the winter months, hence the advice to add materials<br />

that hasten drying at this time of year (see below) .<br />

Beliefs influencing artists' practices<br />

Artists' practices were also influenced by views <strong>and</strong> beliefs that are foreign to<br />

our era. Conservators have discovered empirically that it was not uncommon<br />

for nineteenth-century painters to use similar varnishes in the paint medium<br />

to those used as a final varnish. In the literature, painters were quite explicit;<br />

they believed that using the same resin in the medium as in the final varnish<br />

would, by ensuring homogeneity of materials, reduce the likelihood of cracking<br />

(6). In a pharmaceutical dictionary <strong>and</strong> recipe book published in 1764,<br />

2<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


we find a description of the affinity of like materials that could well be the<br />

antecedent of this nineteenth-century practice (7) .<br />

The widespread addition of varnish to the oil painting medium was itself<br />

fostered by the belief that this was the method used by the old masters to<br />

achieve the particular translucent quality observed in their paint (8).<br />

Seeing the past through present-day lenses<br />

David Lowenthal, in his book The Past Is a Foreign Country, wrote (9):<br />

However fa ithfully we preserve, however authentically we restore, however<br />

deeply we immerse ourselves in bygone times, life back then was based on<br />

ways oj being <strong>and</strong> believing incommensurable with our own. We cannot<br />

help but view <strong>and</strong> celebrate it through present day lenses.<br />

Just as we can enhance our underst<strong>and</strong>ing of artists' practices by learning<br />

more about the implicit assumptions they made based on the beliefs common<br />

to their era, we must be especially careful not to impose our own assumptions<br />

on the past. For example, the identification of fu gitive colors in a nineteenthcentury<br />

painting could lead to the conclusion that the artist knowingly opted<br />

for fugitive colors. Because information regarding which pigments were unstable<br />

was widely available at this time, the presence of these colors in a<br />

painter's work must mean that the painter did not "care" if the colors would<br />

fade. But this is not necessarily true; in fact, the painter may well have conscientiously<br />

purchased stable colors, but may have been unknowingly supplied<br />

with substituted materials by the colormen.<br />

A thorough study of the literature, combined with scientific analyses of paint<br />

samples, revealed that in the nineteenth century the name of a color did not<br />

always offer a reliable indication of composition. Naples yellow is a good case<br />

in point. Traditionally a lead-antimony compound, by the late nineteenth<br />

century Naples yellow was reported to have been substituted with more reliable<br />

coloring agents: lead white <strong>and</strong> cadmium yellow. But the sample labeled<br />

Naples yellow in a Winsor & Newton oil-sample book actually consisted of<br />

lead white, red lake, <strong>and</strong> yellow lake. Since nineteenth-century lake colors<br />

were not particularly stable to light, this particular Naples yellow would be<br />

unlikely to retain its hue indefinitely (10).<br />

Documentary research also indicates that media analysis sampling that is restricted<br />

to one color area cannot be assumed to apply to the whole painting.<br />

Early instruction books, in which the artists were still instructed in grinding<br />

their own paint, indicate that the paint medium would be changed according<br />

to the character of the pigment used. Because of its initial yellow color <strong>and</strong><br />

because it was believed to after-yellow the most, linseed oil was generally<br />

recommended for dark colors, <strong>and</strong> the less-colored poppy <strong>and</strong> nut oils were<br />

reserved for light colors. There were exceptions, however. Lake colors dried<br />

slowly; therefore, linseed oil-the fastest of the three to dry-was recommended<br />

for use with these "light" colors (11). Drying oil in combination<br />

with copal varnish was recommended as the medium for Kings yellow or<br />

orpiment (12). Many other pigments received individual treatment <strong>and</strong> admixtures<br />

with varnish (13).<br />

As we have seen, the choice of oil depended not only on the pigment used,<br />

but also on the season. Oils treated either by boiling alone or in conjunction<br />

with metallic compounds (driers) to hasten their drying time were sometimes<br />

recommended for the winter months only, when damp, cold weather lengthened<br />

drying time. Conversely, painters were warned that such treated oils<br />

would "in summer ... dry so soon as to be troublesome" (14). Since the<br />

essential nature of oil paint had not changed by the early nineteenth century,<br />

it is not surprising to find this kind of advice appearing in print much earlier.<br />

In 1693 Marshall Smith recommended, "If in the hottest weather your greatest<br />

Dryers dry too fast, as White, Umber, &c <strong>and</strong> so grow too stiff to work<br />

with, you may prevent it by mixing a little Sallat Oyle with Colours" (15).<br />

Carlyle 3


Just as we cannot extrapolate to the whole painting from media analysis in<br />

one color area alone, documentary sources indicate that the medium could<br />

change not only from color to color, but also from paint layer to paint layer.<br />

Once again, this was not confined to the nineteenth century. Marshall Smith<br />

instructed that lead white be mixed with nut oil, but noted that linseed oil<br />

could be used in dead-coloring (16). Advice to vary the medium according<br />

to the layer continued to appear in the literature, the faster-drying linseed oil<br />

again being recommended for underlayers such as dead coloring, with poppy<br />

or nut oil in the finishing layers (17).<br />

Interestingly, the medium for the first lay or dead coloring need not have<br />

been oil at all. There were references to the use of watercolor, egg tempera,<br />

<strong>and</strong> a combination of two-thirds starch to one-third oil (18).<br />

Contemporary experience with oil painting materials may also lead to assumptions<br />

that require examination. Today, if we wish to prepare a "traditional"<br />

lead-white ground, the first step would be to size the canvas using a<br />

hide glue such as rabbit-skin or parchment size. Although there are indications<br />

that the use of glue size, including isinglass, was common in the past, this was<br />

not the only material used. Starch was also employed as a size layer <strong>and</strong><br />

appears in recipes throughout the nineteenth century. There were also indications<br />

that the addition of a plasticizer or humectant such as honey, sugar,<br />

or glycerine would not have been unusual. Near the end of the century, we<br />

find a reference to the use of collodion (cellulose nitrate) as a replacement<br />

for the size layer (19).<br />

Our present-day lenses can also result in our underestimating the importance<br />

of materials that in our own day are no longer in use or have become precious<br />

<strong>and</strong> rare. Isinglass, a glue prepared from the swim bladder of the Russian<br />

sturgeon, is not widely available today. In eighteenth- <strong>and</strong> nineteenth-century<br />

Engl<strong>and</strong>, isinglass was commonly used for a variety of purposes: to clarifY<br />

beer, wine, <strong>and</strong> soup, <strong>and</strong> as a sizing agent for fabric, ribbons, <strong>and</strong> paper.<br />

Therefore, the presence of what is now quite rare, but was then a relatively<br />

commonplace glue in the size layer for an oil painting, is not surprising.<br />

Another such material that has dropped out of use entirely is sugar of lead<br />

(lead acetate). A white crystalline powder widely used as a drier fo r oil paint,<br />

it was added directly to the pigment-oil mixture <strong>and</strong> was also present in<br />

medium recipes. Lead acetate could be purchased easily from apothecaries<br />

<strong>and</strong> appears to have been in wide use by painters in the late eighteenth <strong>and</strong><br />

nineteenth centuries. By the twentieth century, however, it was never mentioned<br />

in sources on oil painting materials <strong>and</strong> techniques, although other<br />

traditional lead driers, such as litharge or metallic lead, do receive notice. As<br />

a result, the important role that this material played has never been acknowledged<br />

or studied in the twentieth century.<br />

Conclusion<br />

Although we must accept Lowenthal's observation that life in the past was<br />

"based on ways of being <strong>and</strong> believing incommensurable with our own," we<br />

should not see the exercise of studying past practices <strong>and</strong> materials as fu tile<br />

(20). Rather, we should equip ourselves with the knowledge that we are<br />

h<strong>and</strong>icapped by our late twentieth-century st<strong>and</strong>point. By making use of a<br />

variety of disciplines, by not concentrating our energies too much on only<br />

one avenue of inquiry, we can continue the search to "know all." As much<br />

as possible, we should look outside of our immediate disciplines for researchers<br />

who are also mining the past, as it is this multidisciplinary approach that<br />

will enrich our underst<strong>and</strong>ing <strong>and</strong> interpretation of the "facts."<br />

Notes<br />

1. Forester, Sir Archibald. 1990. Sponsor's preface. In Art in the Making: Impressionism,<br />

D. Bomford,]. Kirby, ]. Leighton, <strong>and</strong> A. Roy. London: The National Gallery<br />

<strong>and</strong> Yale University Press.<br />

4<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


2. A Practical Treatise on <strong>Painting</strong> in Oil Colours. 1795. London, v.<br />

3. Daly, M. 1993. The varnished truth. Art Review. Vol. XLV (November): 57.<br />

4. Davidson, C. 1986. A Woman's Work 15 Never Done. London: Chatto & Windus,<br />

132.<br />

5. Carlyle, L., <strong>and</strong> J. Townsend. 1990. An investigation oflead sulphide darkening<br />

of nineteenth-century painting materials. In Dirt <strong>and</strong> Pictures Separated, London:<br />

United Kingdom Institute fo r Conservation, 40-43.<br />

6. Carlyle, L. 1990. British nineteenth-century oil painting instruction books: a<br />

survey of their recommendations fo r vehicles, varnishes, <strong>and</strong> methods of paint<br />

application. In Cleaning, Retouching, <strong>and</strong> Coatings: Technology of <strong>Practice</strong> fo r Easel<br />

<strong>Painting</strong>s <strong>and</strong> Polychrome Sculpture. London: International Institute for Conservation,<br />

79.<br />

7. "Experiments concur with daily Observation to prove, that different Bodies,<br />

whether Principles or Compounds, have such a mutual Conformity, Relation,<br />

Affinity, or Attraction, as disposes some of them to join <strong>and</strong> unite together,<br />

while they are incapable of contracting any Union with others. Substances having<br />

an affinity together, will unite <strong>and</strong> form one Compound. It may be laid<br />

down as a general Rule, that similar Substances have an Affinity with each other,<br />

as Water with Water, Earths with Earths, &c." James, R. 1764. Pharmacopoeia<br />

Universalis: Or, A New Universal English Dispensatory. London, 11.<br />

8. Carlyle, L. 1991. A critical analysis of artists' h<strong>and</strong>books, manuals <strong>and</strong> treatises<br />

on oil painting published in Britain between 1800-1900: with reference to selected<br />

eighteenth-century sources (Ph.D. diss., Courtauld Institute of Art, University<br />

of London), Vol. I, 27-28.<br />

9. Lowenthal, D. 1985. The Past 15 a Foreign Country. Cambridge: Cambridge University<br />

Press, xvi.<br />

10. Townsend, J., L. Carlyle, S. Woodcock, <strong>and</strong> N. Kh<strong>and</strong>ekar. n.d. Later nineteenth-century<br />

pigments: analysis of reference samples, <strong>and</strong> evidence for adulterants<br />

<strong>and</strong> substitutions. The Conservator. United Kingdom of Conservation.<br />

Forthcoming.<br />

11. Templeton, J. S. 1846. The Guide to Oil <strong>Painting</strong>. London, 28.<br />

12. Ibid., 58.<br />

13. Carlyle 1990, op. cit., 76.<br />

14. Williams, W. 1787. An Essay on the Mechanic of Oil Colours . ... Bath, Engl<strong>and</strong>,<br />

47.<br />

15. Smith, M. 1693. The Art of <strong>Painting</strong> . ... London, 73.<br />

16. Ibid., 71.<br />

17. For example, "The oils are the mixture of oil <strong>and</strong> turpentine; <strong>and</strong> as the portrait<br />

advances towards the finishing sitting, nut or poppy oil may be substituted in<br />

the mixture fo r boiled oil." Cawse, J. 1822. Introduction to the Art of <strong>Painting</strong> in<br />

Oil Colours. London, 15.<br />

18. Carlyle 1991, op. cit., 288.<br />

19. Ibid., 247.<br />

20. Lowenthal, op. cit.<br />

Carlyle 5


Abstract<br />

Methods of art historical research,<br />

painting expertise, <strong>and</strong> scientific research<br />

can contribute to the detailed<br />

study of painting techniques. Conversely,<br />

the knowledge of painting<br />

techniques could support art historical<br />

research. The need for collaboration<br />

between individual disciplines<br />

during all stages of research is<br />

strongly stressed, from the initial formulation<br />

of working questions <strong>and</strong><br />

hypotheses to final conclusions about<br />

the techniques of an artist.<br />

An Integrated Approach for the Study<br />

of <strong>Painting</strong> <strong>Techniques</strong><br />

Sylvana Barrett <strong>and</strong> Dusan C. Stulik*<br />

The Getty Conservation Institute<br />

4503 Glencoe Avenue<br />

Marina del Rey, California 90292<br />

USA<br />

Introduction<br />

The study of artists' techniques in general <strong>and</strong> the study of an individual<br />

artist's techniques in particular are important for several reasons:<br />

(a) Art historians can use detailed knowledge of an artist's technique <strong>and</strong> its<br />

developmental evolution throughout the artist's career in the authentication<br />

process. This information can also aid in the establishment of a proper chronology<br />

for the known works of a given artist.<br />

(b) Artists of various historical periods were able to achieve specific visual<br />

effects by the use of special artists' materials or by methodical application of<br />

proven painting techniques. Because detailed documentation is seldom available<br />

today to help artists learn the steps needed to re-create a given visual<br />

effect, artists must rely on the results of systematic art research to learn old<br />

master techniques.<br />

(c) Museum conservators, in order to ensure a safe working strategy when<br />

planning a conservation or restoration treatment, rely on specific information<br />

about pigments, binding media, <strong>and</strong> materials, including those of earlier restorations,<br />

as well as detailed knowledge of the structural arrangements of these<br />

materials.<br />

The art historian <strong>and</strong> art research<br />

The role of the art historian in the realm of art research is critical. It is the<br />

art historian who must set the foundation, into which the information gathered<br />

by individual researchers of a painting-technique research team is organized,<br />

for final interpretation. In the study of painting techniques, two tools<br />

used by art historians are very important: connoisseurship <strong>and</strong> archival research.<br />

Connoisseurship. Beyond establishment of individual <strong>and</strong> historical chronologies,<br />

connoisseurship itself-the expert knowledge of style <strong>and</strong> technique that<br />

the art historian develops through the course of a career-is of immense help<br />

in the research of artists' techniques. This keen sense of discrimination can<br />

be used to identify idiosyncrasies particular to a given artist's work. When<br />

drawn from works of undisputed provenance, the art historian can use the<br />

idiosyncrasies to establish a signature of style, materials, <strong>and</strong> techniques for<br />

any given artist. This "signature style" is critical for researchers in all the<br />

related disciplines. The signature style for a particular artist establishes the<br />

st<strong>and</strong>ard of measure against which all data can be judged.<br />

Archival research. The study of primary documents-such as municipal, guild,<br />

or financial records <strong>and</strong> chronicles in which artists' names can be directly<br />

located-provides crucial information about artists' lives, training, professional<br />

<strong>and</strong> social st<strong>and</strong>ing, <strong>and</strong> other socio-economic factors that influenced the<br />

development of their working methods <strong>and</strong> personal painting styles. When<br />

an artist's notes, letters, books, diaries, <strong>and</strong> travel journals are available, the<br />

study of these materials often provides important information leading to<br />

* Author to whom correspondence should be addressed.<br />

6<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


knowledge <strong>and</strong> underst<strong>and</strong>ing of the artist's intent, as well as the rationale<br />

behind the use of a particular material <strong>and</strong> painting technique.<br />

The study of primary sources-such as books <strong>and</strong> treatises on painting, old<br />

recipes for the preparation of artist materials, <strong>and</strong> guild practices <strong>and</strong> procedures-can<br />

contribute to our underst<strong>and</strong>ing of the artists' working environment.<br />

The study of such documents can be invaluable, but extreme caution<br />

should be applied when interpreting old manuscripts. A number of misunderst<strong>and</strong>ings<br />

have arisen because of the often confusing nomenclature used<br />

in old texts <strong>and</strong> recipes <strong>and</strong> the distorted facts found in old biographies of<br />

artists. Even so, the study of secondary documents, such as biographies of<br />

painters, provides important information of the creative, social, <strong>and</strong> economical<br />

environment in which the artist worked.<br />

Artists <strong>and</strong> art research<br />

A painting is composed of elements that can be separated for study. As there<br />

are schools of artistic thought, there are schools of painting technique. Each<br />

school of painting technique has a specific procedural approach to painting<br />

construction. Although there are many elements common to all painting<br />

techniques, there are also specific elements unique to each. It is possible to<br />

isolate <strong>and</strong> define these unique qualities for each technical approach <strong>and</strong> to<br />

establish markers for a detailed study. These markers, when encountered during<br />

an examination, provide keys to the likely construction of visual effects<br />

within the piece.<br />

An initial, st<strong>and</strong>ardized visual examination of a painting by an artist specializing<br />

in painting techniques can id in any art historical or subsequent scientific<br />

investigation, not only helping to orient the researcher but also assisting<br />

in the interpretation of the results. If, for example, it was determined through<br />

visual examination that a particular piece was a multilayered, glazed construction<br />

on panel, it might then be assumed, based on knowledge about the<br />

particular technique, that the piece would have an oily imprimatura on a<br />

gesso ground. Knowing this beforeh<strong>and</strong>, a researcher who discovered an oily<br />

component in the ground layer might investigate the possibility of its having<br />

been absorbed from the imprimatura by the lean ground rather than assuming<br />

the discovery of a novel gesso recipe.<br />

Technical approaches used for creating the illusion of volume<br />

A basic goal of all representational artists-to present an illusion of volumeis<br />

accomplished in painting through the juxtaposition of dark <strong>and</strong> light values,<br />

<strong>and</strong> of highlights <strong>and</strong> shadows. This illusion is accomplished traditionally<br />

through one of four basic techniques. A brief analysis of these techniques will<br />

illustrate the possibilities of st<strong>and</strong>ardizing the visual examination of paintings<br />

<strong>and</strong> the usefulness of the visual markers that can be established as a result of<br />

this approach.<br />

Basic technique. In a direct approach, dark <strong>and</strong> light values are placed by single,<br />

individual brush marks onto the surface of the painting. There is no blending<br />

of the pigments themselves. Dark values are used to indicate shadow <strong>and</strong> light<br />

values to indicate highlights, effectively indicating volume. Despite the obvious<br />

simplicity of this approach, it can be highly successful. The tempera<br />

paintings of the Italian Trecento illustrate the effectiveness of this technique.<br />

Visual markers for this technique include a uniform surface of clearly defined,<br />

individual brush strokes that retain their original distinct color <strong>and</strong> do not<br />

physically blend into surrounding pigments (Plate la, b).<br />

Transparent oil technique. This more complex, systematic approach, which was<br />

developed with the advent of transparent oil media, is exemplified by the<br />

Flemish <strong>and</strong> early Netherl<strong>and</strong>ish masters. These artists conceived of the painting<br />

from its inception as a multilayered object with a structural separation of<br />

color <strong>and</strong> fo rm. Volume, developed through highlights <strong>and</strong> shadows in a monochromatic<br />

underpainting, was followed by color embellishments.<br />

Barrett <strong>and</strong> Stulik 7


Well aware of the optical properties of both light <strong>and</strong> color, the artist worked<br />

on a highly reflective, white ground layer. The underpainting could be a<br />

complete gray-toned version of the finished image, painted in a manner such<br />

as that described in the preceding basic technique. It could also be constructed<br />

through a more sophisticated technique, as seen in the unfinished panel of<br />

Santa Barbara by Jan van Eyck. An underdrawing, which establishes contours<br />

<strong>and</strong> darks on the white ground layer, is covered with an imprimatura, a thin,<br />

transparent layer of paint that allows the drawing to show through the ground<br />

while also establishing a middle tone throughout the painting. Highlights<br />

could then be added in white paint where appropriate, thus, with less work,<br />

completing the values <strong>and</strong> creating a finished monochromatic underpainting.<br />

Regardless of the approach taken toward the underpainting, its creation was<br />

essential to the technique itself. Color applied as thin transparent glazes allowed<br />

the fully developed underpainting to define the fo rms while the color<br />

itself remained clean, pure, <strong>and</strong> unadulterated. Highlighted areas could be<br />

achieved with the thinnest possible application of local color, as the white of<br />

the underpainting had merely to be tinted appropriately. Dark tones, however,<br />

posed some problems with the clear transparent pigments: many layers were<br />

required to cover the underdrawing <strong>and</strong> establish the proper local color.<br />

By fo cusing on these highlights <strong>and</strong> shadows, visual identification of the technique<br />

is quite simple. Highlighted areas are very thin <strong>and</strong> fine. Color applied<br />

in thin glazes tends to be clear, luminous, <strong>and</strong> devoid of brush marks. Shadows<br />

<strong>and</strong> dark colors, however, appear as thickly built-up surfaces, creating ridges<br />

clearly visible in raking light where they come into contact with the delicate<br />

light areas (Plate 2a, b).<br />

Highlighting with impasto white. Allowing a freer painting style <strong>and</strong> facilitating<br />

larger formats, this more flexible technique is typical of the Baroque masters.<br />

Any support suffices; the underdrawing is optional. The artist tones the surface<br />

with a middle or darker value, then creates the image with an underpainting<br />

of washes that may be controlled or completely free <strong>and</strong> spontaneous. The<br />

areas of the painting to be highlighted are now created with a heavy impasto<br />

white paint.<br />

This simple procedure accomplishes the same optical effects as the complete<br />

monochromatic underpainting of the previous transparent oil technique, yet<br />

it allows the image to evolve as it is constructed. The continued separation<br />

of value from color still allows for beautiful luminous color. Because the image<br />

originates in the loose, dark washes, contours need not be highly defined <strong>and</strong><br />

extreme chiaroscuro is possible. The resulting work is often quite dramatic in<br />

nature.<br />

This technique provides very specific optical markers. The darks are thin <strong>and</strong><br />

transparent, often revealing the preliminary wash or imprimatura. The highlights<br />

that define the volume appear thick <strong>and</strong> visibly raised from the painted<br />

surface (Plate 3a, b).<br />

Direct suiface blending. In the controlled technique of surface blending, individual<br />

colors <strong>and</strong> values are mixed <strong>and</strong> applied to appropriate locations of<br />

the surface to indicate highlight <strong>and</strong> shadow. Each new application of color<br />

is carefully blended into the surrounding paint, resulting in a smooth, continuous<br />

flow. The underdrawing <strong>and</strong> underpainting serve only as a guide for<br />

the surface painting; they do not actively affect the surface itself.<br />

All traces of brushwork can be blended out if desired; consequently, the technique<br />

lends itself well to smooth, detailed, controlled styles such as found in<br />

the work of the Neoclassicists.<br />

Visual markers fo r the technique include a smooth, continuous surface with<br />

gradual, imperceptible shifts from highlight to shadow. Direct blending of the<br />

pigments creates an opaque quality in contrast to the luminous character of<br />

colors in the multilayered approaches (Plate 4a, b).<br />

8<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Table 1. Scientific methods used Jo r identification oj critical, unaltered parts oj a painting.<br />

M ethod In ormation LimitalJon<br />

observatIOn under UV light I dark areas of repamlJng III - does not worK lor all<br />

comparison wilh light areas<br />

binding media<br />

of heavily oxidized oil paints - possible interference with<br />

(autofluorescence)<br />

some varnishes<br />

UV photography documentation of repainted as above<br />

areas<br />

false color infrared<br />

identification of repainting if only for orientation, should<br />

photography done using pigments of the be confirmed using other<br />

same color but different methods of chemical<br />

chemical composition analysis<br />

X -ray radiography localization of support high contrast only for heavy<br />

alterations, ground <strong>and</strong> paint chemical elements (Pb, Hg,<br />

layer repairs <strong>and</strong> alterations Au)<br />

X-ray f1uorescence - identification of repainted - possible difficulties with<br />

spectrometry (XRF) areas if done with inorganic mixed pigments <strong>and</strong><br />

pigments of different complex multilayer paint<br />

composi tion than original structure analysis<br />

(e.g., zinc or titanium white - no analytical information<br />

instead of lead white) for low atomic weight<br />

- a non contact analysis elements (Z :$ Na)<br />

microchemIcal methods Identification of repainted - microsampling 01 paint<br />

(polarized light microscopy, areas based on pigment material needed<br />

chemical microscopy, identification - uncertain identification of<br />

electron microprobe, X-ray<br />

some organic pigments <strong>and</strong><br />

diffraction)<br />

dyes<br />

The visual traces left by the physical manipulation of painting materials on<br />

paintings can be identified fo r study. Idiosyncrasies of color, brushwork, paint<br />

consistency, form, <strong>and</strong> so on can each be evaluated to reveal specific unique<br />

qualities within given predefined techniques. This information can then be<br />

used for the general analysis of historical paintings. The works of individual<br />

artists can be analyzed in a similar manner, identifYing a signature style <strong>and</strong><br />

painting procedure through specifiC visual clues.<br />

The role of scientific <strong>and</strong> technical examinations<br />

The process of applying scientific methodology to the study of painting techniques<br />

can be divided into the following three major steps: (1) identification<br />

of the critical (unaltered) parts of the painting, (2) authentication of the painting,<br />

<strong>and</strong> (3) study of detailed physical <strong>and</strong> chemical structure of paintings.<br />

Identification oj the critical (unaltered) parts of the painting. The majority of paintings<br />

in museums or private collections have a long history of cleanings, restorations,<br />

<strong>and</strong> alterations. Before embarking on a study of painting techniques,<br />

it is essential to identifY areas of the painting in which the painting technique<br />

of the original artist has not been altered by later treatments. Scientific methods<br />

for such a study are described briefly in Table 1.<br />

Authentication oj the painting. When a series of paintings by a particular artist<br />

is examined with the goal of studying painting techniques, it is crucial that<br />

the authorship of the pieces is established "beyond a reasonable doubt." Scientific<br />

research cannot establish the relationship between the artwork <strong>and</strong> the<br />

artist. What scientific research can do very successfully is to effectively eliminate<br />

paintings that, based on clearly defined scientific facts, could not have<br />

been created during the active life of the artist in question. Several powerful<br />

scientifiC methodologies that can be used to help authenticate paintings are<br />

described in Table 2.<br />

Study of detailed physical <strong>and</strong> chemical structure of paintings. Usually, only the top<br />

paint layer, with its corresponding brushwork <strong>and</strong> surface treatment, is accessible<br />

to visual observation. This is a very severe limitation when studying<br />

painting techniques because before the alia prima technique became widely<br />

used during the second half of the nineteenth century, the majority of paintings<br />

were created in complex multiple-step <strong>and</strong> multilayer processes. Table 3<br />

shows several scientific methodologies that can be used in probing <strong>and</strong> analyzing<br />

such painting structures.<br />

Barrett <strong>and</strong> Stulik 9


Table 2. Scientific techniques Jor the authentication oj paintings.<br />

VISI ble observation <strong>and</strong> X- visible <strong>and</strong> hidden tool - knowledge 01 ancient<br />

I Method Inlonnation Limitation<br />

ray radiography marks; materials (machine technologies required<br />

made canvas, nails, etc.) - new repairs may pose a<br />

problem<br />

pigment chronology identification of the - detailed knowledge of<br />

inconsistency studies systematic use of pigments pigment chemistry <strong>and</strong><br />

(microchemical methods) in the painting which were technology required<br />

not available in times when - problem of inteIpretation<br />

the painting was presumed when only repainted areas<br />

to be painted<br />

analyzed<br />

- the negati ve resul ts cannot<br />

be considered a certain proof<br />

of authenticity<br />

radiocarbon dating absolute dating method of - does not provide good<br />

the actual age of natural results for materials younger<br />

product organic materials in than 300 years<br />

paintings (wood, canvas, - prOblems of impurities<br />

binding media)<br />

dendrochronology or tree absolute dating method quality sample containing a<br />

ring dating actual age of wood sam pIes number of easily measured<br />

tree rings <strong>and</strong> corresponding<br />

calibration data is needed<br />

INTEGRATED APPROACH<br />

FOR THE STUDY OF<br />

PAINTING TECHNIQUES<br />

Figure 1. Diagram of integrated approach for<br />

the study of painting techniques.<br />

Conclusion<br />

Art historians, conservation scientists, <strong>and</strong> artists use different means to study<br />

artists' techniques. Each of their approaches can contribute valuable information<br />

about the painting techniques of an individual artist, a school, or<br />

an art historical period or movement. But each approach leaves something<br />

unexplained, something missing from the whole picture that encompasses<br />

everything from the artist's brushwork to preference for certain materials <strong>and</strong><br />

formulas. To provide a real underst<strong>and</strong>ing of artists' techniques, it is necessary<br />

to establish a close collaboration between all the above-mentioned disciplines,<br />

not only to secure a more complete set of data but, more importantly, to<br />

stimulate interdisciplinary formulation of more holistic answers about artists'<br />

techniques (Fig. 1). A painting should not be studied by individual specialists<br />

from each discipline, but rather by representatives of all disciplines who view<br />

Table 3. Scientific methods used Jo r the analysis oj painting structures.<br />

Method In onnation Llmltallon<br />

mfrared refiectography Identillcation <strong>and</strong> study of I difficult to Idenllly<br />

underdrawing<br />

underdrawing beneath thick<br />

layer of IR-opaque paint<br />

layer<br />

X-ray radiography study of support, lead white - superimposition of several<br />

underpainting, brushwork, layers of painting in one<br />

<strong>and</strong> changes in composition X-ray radiograph<br />

- difficult inteIpretation for<br />

features done in organic<br />

materials or pigments of low<br />

atomic number<br />

X-ray tomography detailed study of individual - methodology under<br />

paint layers<br />

development<br />

- high cost<br />

- problem of interpretation<br />

when individual paint layers<br />

are of uneven thickness<br />

cross section analysis detailed material - cross section sample<br />

(microchemical methods, IR identification (pigments, needed<br />

microscopy) binding media) <strong>and</strong> - high cost of analysis<br />

sequence of individual paint - positive identification of<br />

<strong>and</strong> material layers of the some organic pigments<br />

painting<br />

might need additional<br />

sampling, followed by<br />

organic microchemical<br />

analysis<br />

- additional sampling might<br />

also be needed for detailed<br />

analysis of binding media<br />

using gas chromatography -<br />

mass spectrometry<br />

10<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


the painting together, share background information, <strong>and</strong> actively collaborate<br />

in formulating a working hypothesis, work strategy, <strong>and</strong> research goals. Successful<br />

research calls fo r broad international <strong>and</strong> multidisciplinary collaboration<br />

when art technique findings are used for authentication purposes.<br />

Barrett <strong>and</strong> Stulik 11


Abstract<br />

This paper provides an overview of<br />

technical aspects of the search for<br />

verisimilitude in seventeenth-century<br />

Italian painting. In particular the role<br />

of varnishing will be examined in<br />

relation to technical problems caused<br />

by absorbing grounds. In addition,<br />

some theories on viewing distance<br />

<strong>and</strong> lighting will be discussed.<br />

Varnish, Grounds, Viewing Distance, <strong>and</strong> Lighting:<br />

Some Notes on Seventeenth-Century Italian<br />

<strong>Painting</strong> Technique<br />

Helen Glanville<br />

au Fourquet 471 20<br />

Pardaillan<br />

France<br />

Introduction<br />

In the sixteenth century, the dichotomy between Disegno <strong>and</strong> Colore, between<br />

Titian <strong>and</strong> Raphael, was seen as one between those artists who chose to<br />

imitate nature <strong>and</strong> those who chose the Antique as their model. In the seventeenth<br />

century, this dichotomy could be reduced, simplistically speaking, to<br />

two groups of artists: those who were more strongly influenced by Raphael<br />

<strong>and</strong> the Antique <strong>and</strong> who painted with a smoother, apparently more finished<br />

technique; <strong>and</strong> those who were more influenced by Titian <strong>and</strong> the Venetian<br />

school of painting <strong>and</strong> the more open texture that accompanied this type of<br />

representation of naturalistic effect-what Poussin's friend Du Fresnoy termed<br />

"the great Lights <strong>and</strong> Shadows, the Effect of the whole together" (1). These<br />

two tendencies have been seen to coexist through the end of the nineteenth<br />

century.<br />

Although these two schools differed in their approaches to h<strong>and</strong>ling paint,<br />

they both subscribed to the idea that painting should be the representation<br />

of natural appearances on a flat surface; <strong>and</strong> most importantly, that through<br />

this representation the public should be able to grasp a higher <strong>and</strong> greater<br />

truth. Ideas as to what form this imitation of nature should take varied, but<br />

the essential concept can be found in the writings of theorists as divergent<br />

in other respects as the arch-Venetian Boschini <strong>and</strong> Bellori, the epitome of<br />

Roman Classicism (2, 3).<br />

The first Academy of <strong>Painting</strong>, founded in 1586 by the Carracci in Bologna,<br />

was crucial to the development of painting in seventeenth-century Italy. Painters<br />

such as Domenichino, Reni, Albani, <strong>and</strong> Guercino (as well as Annibale<br />

Carracci) who had come to Rome after training at the Carracci Academy<br />

profoundly influenced their contemporaries in Rome. They brought not only<br />

the teachings of their masters (i.e., that the painter had to emulate nature<br />

accurately on a flat surface, while also illustrating the essence, the "Truth," of<br />

what was depicted, that which was beyond simple appearances). This concept,<br />

essential to painting in its newly reacquired status as a liberal art, was also in<br />

complete accordance with the tenets laid down by the Counter-Reformation.<br />

Two camps emerged concurrently: those who described the thing itself, <strong>and</strong><br />

those who described the impression on the beholder.<br />

St. Philip Neri <strong>and</strong> Paleotti both required that artists, through verisimilitude<br />

or realistic representation, appeal to the hearts <strong>and</strong> minds of the people.<br />

This paper provides an overview of technical aspects of the search fo r verisimilitude<br />

in seventeenth-century Italian painting. In particular the role of<br />

varnishing will be examined in relation to technical problems caused by absorbing<br />

grounds. In addition, some theories on viewing distance <strong>and</strong> lighting<br />

will be discussed.<br />

The use of varnish<br />

Andre Felibien, who moved in the artistic circles of Rome in the 1640s <strong>and</strong><br />

who was a friend of Poussin, Guercino, <strong>and</strong> Cigoli, <strong>and</strong> probably knew Galileo,<br />

wrote, "When a work is painted to the last degree of perfection, it can<br />

be considered from close to [sic]: it has the advantage of appearing stronger<br />

<strong>and</strong> three-dimensional" (4). This same effect, wrote Felibien, can be created<br />

12<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


y distance with "the aid of the air interposed between the eye <strong>and</strong> the object,<br />

using different distances," or by the application of a varnish "which is why<br />

we cover paintings with a varnish that blunts that brightness <strong>and</strong> sharp edge<br />

[qui emousse cette pointe brillante et cette vivacite] which at times appears too<br />

strongly or unevenly in freshly painted works; <strong>and</strong> this varnish gives them<br />

more strength <strong>and</strong> softness [douceur]. ... We use all these different methods<br />

to give painted objects that relief, that roundness that they require in order<br />

to resemble what one is imitating" (5).<br />

Marco Boschini, the Venetian art theorist who in 1660 published his poem<br />

on the art of painting, "La Carta del Navigar Pitoresco," had a low opinion<br />

of the glossy varnishes which he says "foreigners" used: "They make such a<br />

commotion about it, that it would seem that gloss is the only beauty, <strong>and</strong><br />

varnish the apogee of art" (6). Later in the poem, he makes the distinction<br />

that is central to artistic theory <strong>and</strong> to seventeenth-century Italian artists in<br />

Venice as well as Rome. Comparing the aforementioned "foreigners," Boschini<br />

praises the Venetian painter fo r imitating the effects of gloss (in armour<br />

<strong>and</strong> mirrors) in paint, rather than resorting to the use of varnish to produce<br />

this effect physically on the painting's surface, writing, "Ii ha fatti straluser co'i<br />

colori" (he has made them gleam with his paints).<br />

As Boschini wrote, "diligent painting can be copied," but not what he termed<br />

the colpi di dottrina (which rather defies translation). It is the element of skill<br />

that is paramount, rather than the materials.<br />

A high-gloss varnish may have been a requirement of northern seventeenthcentury<br />

artists, at least according to Boschini, but this paper investigates the<br />

role played by varnish in Italian paintings of the same period, both those<br />

requiring "close scrutiny" <strong>and</strong> those "made to be seen from afar," as Horace<br />

discussed, distinguishing these two kinds of fundamentally different paintings,<br />

when pleading for flexibility in the judgment of poetry, saying that it should<br />

be judged like painting which exhibits not only a detailed style that requires<br />

close scrutiny, but also a broad impressionistic style that will not please unless<br />

viewed from a distance (7).<br />

Interesting information on the use of varnish in seventeenth-century Italian<br />

painting can be gained from the highly illuminating lecture entitled "II Lustrato"<br />

delivered on 29 December 1691 by Filippo Baldinucci to his colleagues<br />

at the Accademia della Crusca in Florence. He dealt with the subject of<br />

varnish in contemporary painting as opposed to its use in ancient times <strong>and</strong><br />

in the fourteenth century (8):<br />

VVhat we are told oj the practices oj painters in antiquity leads us to believe<br />

that they did not use oil as their medium; what we are told oj the practice<br />

oj Apelles always remains with me, that is, that he fo und a certain dark<br />

color [9], or maybe a varnish, which no one was able to imitate. This<br />

varnish he applied to his works after he had finished them, <strong>and</strong> with such<br />

a skill that the bright colors did not offe nd the eye <strong>and</strong> appeared from afa r<br />

as through a glass (<strong>and</strong> please take note of this detail). Harsh colors acquired<br />

through it an element of austerity, oj saturation. This is precisely what our<br />

painters in the fo urteenth century did, bifore the discovery of oil as a<br />

medium; that is, they applied a varnish to their panels, which was of a<br />

composition that gave their languished paintings a certain depth <strong>and</strong> added<br />

strength, <strong>and</strong> by dimming the overpowering highlights, brought the whole<br />

closer to natural appearances. And then if you hear it told that modern<br />

painters sometimes also use a varnish on their oil paintings, I reply that<br />

such practice (which is con f ined to a few painters) is not to make up a<br />

deficiency in oil painting-that is, to bring depth to the dark colors, <strong>and</strong><br />

render the highlights softer <strong>and</strong> less garish-things of which oil painting<br />

has no need, but rather to remedy an accidental mishap. This can occur<br />

because oj the imprimatura, the paint mixture which one applies to the<br />

canvases or panels, or because oj a difect of the canvases or panels themselves.<br />

They attract the oil liquid so strongly, almost stealing it from the<br />

Glanville 13


paint, that the paint remains drained [prosciugato], so that it can no<br />

longer be seen evenly on the suiface, as would have been the case when<br />

such an accident would have been prevented. By means oj another unctuous<br />

substance, which is the varnish, applied to those areas where the oil is<br />

missing on the suiface (<strong>and</strong> this is the crux of the matter), the dark colors<br />

are made to reappear. These are the dark pigments which are really present<br />

in the oil painting, not those darks that just appear dark but are not<br />

physically present, as was in fa ct the iff ect created by Apelles's varnish in<br />

some very small areas on his paintings.<br />

Many of Baldinucci's points require <strong>and</strong> deserve investigation; I will confine<br />

myself to examining his assertion that varnish in his time was applied as a<br />

retouching remedy locally rather than over the whole work (as the oil medium<br />

itself provided the saturation required to "bring the whole closer to<br />

natural appearances"). I shall also fo llow up his point linking this localized<br />

"sinking-in" with absorbency of the ground layer.<br />

We are indebted to the Englishman Richard Symonds, an amateur painter<br />

who spent 1646-1647 in Rome, fo r the very detailed <strong>and</strong> painstaking notes<br />

he took while watching Gian Angelo Canini painting in his studio. Canini<br />

was a fr iend of Poussin's, <strong>and</strong> like the latter, trained in Domenichino's studio<br />

in Rome.<br />

Symonds describes on several occasions that portions of Canini's paintings<br />

were "sinking in" (prosciugated), <strong>and</strong> he also described the remedies that<br />

Canini applied. Giving an account of the painting of a portrait of Sir Thomas<br />

Killigrew, he describes how: "the face <strong>and</strong> field were prosciugated ... nitwithst<strong>and</strong>ing<br />

oyle was putt upon the back side of the cloth." This he blamed<br />

on the fact that the "cloth" (<strong>and</strong> he marked "imprimatura" in the margin)<br />

"was not as perfectly dry as it ought. I askt him how he would fetch the<br />

colour of the face, he sayd he would give it a semplice chiara d'uovo [eggwhite]<br />

beaten together . .. or olio di sasso, another varnish" (10).<br />

Elsewhere, Symonds describes Robert Spenser's portrait, painted in one sitting<br />

by Canini (1 1):<br />

The scaife which was crimson he did with lacca / biacca Qake <strong>and</strong> lead<br />

white]. The whole scaife being done, <strong>and</strong> afo re it was dry he putt on gold<br />

colour for the fringe, all which kept his fresh colour <strong>and</strong> needed no varnish.<br />

Not three days after, when one would thinke it was scarce dry he with a<br />

pencill of setola [hog] putt on his varnish over the first field & face &<br />

Armour & h<strong>and</strong>s, but not the scaife or benda, / this kind of varnish he<br />

esteemed above that of Olio di Sasso.<br />

For the preferred varnish, he gives the following recipe on f.20: "2 oz. of<br />

seven times distilled aquavita (spirits of wine) 112 oz. of ground s<strong>and</strong>arac<br />

112 oz. of olio d'Abezzo [Strasbourg Turpentine]." This "final" varnish was<br />

only applied to those areas which he felt required it. He was also quite clear<br />

that certain pigments should not be varnished.<br />

Symonds questioned Canini about the sky in a monumental painting that he<br />

executed on the subject of Anthony <strong>and</strong> Cleopatra: "2 days after the Azzurro<br />

ayre was dry, <strong>and</strong> I askt him what if it should prosciugare [sink in], so much<br />

the better said he, because you never put vernish over azzurro" (12).<br />

The reason for this answer may be in the yellowing characteristics of the<br />

varnish or in a deliberate choice fo r the matte quality of the blue with a<br />

consequent increase in scattered reflection, <strong>and</strong> of the white component in<br />

light. This effect, to be avoided in darks, but desirable for achieving the effects<br />

of aerial perspective, was advised by Leonardo in his treatise on painting. One<br />

should not forget that all the artists in Poussin's circle had a strong interest in<br />

optical matters, <strong>and</strong> studied mathematics <strong>and</strong> optics with the best mathematicians;<br />

in addition, Galileo moved in this same circle.<br />

Canini was also aware that certain pigments, such as blacks, caused problems<br />

by drying matte: "Black, even varnished, sinks in," meaning that unless cor-<br />

14<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


ected, shadows would appear matte <strong>and</strong> therefore cool <strong>and</strong> come forward<br />

rather than giving the intended effects of depth (13).<br />

Armenini is also aware of this problem. In his De' veri precetti della pittura<br />

(1587), speaking of shadows <strong>and</strong> contours he says: "They never remain sufficiently<br />

finished ... but will always sink in, or the darks become cruder so<br />

that it is necessary to come back to them many times in order to make them<br />

fresh, blended, soft, <strong>and</strong> pleasing" (14).<br />

Absorbency of the grounds<br />

The problem of overabsorbent grounds, which is identified by Baldinucci as<br />

the cause for localized applications of varnish, is also a problem that reoccurs<br />

in Symonds's notes <strong>and</strong> obviously concerned Canini <strong>and</strong> presumably the other<br />

artists in this circle.<br />

Almost all of the samples taken from paintings produced in central Italy circa<br />

1600-1800 showed the presence of chalk in varying but significant proportions.<br />

Sometimes chalk was found as a single ground layer, as in the case of<br />

Domenichino's Vision of St. Jerome of 1603 (National Gallery, London); in<br />

Annibale Carracci's Boy Drinking (Christ Church Gallery, Oxford) which<br />

probably dates from the end of the sixteenth century; <strong>and</strong> in the much later<br />

Cavallino, Christ Driving the Traders from the Temple (National Gallery, London).<br />

This type of ground derives from Venetian painting <strong>and</strong> is found, for instance,<br />

in Titian's Madonna <strong>and</strong> Child (ca. 1525), in which what appears to be the<br />

dark ground is in fact the saturated color of the canvas. The refractive index<br />

of chalk is so low that it is almost completely transparent when associated<br />

with oil.<br />

Not enough analyses have been carried out to ascertain whether the chalk<br />

was applied in an aqueous or oil medium. More often the chalk is found<br />

mixed with earth pigments <strong>and</strong> small inclusions of other pigments such as<br />

charcoal black, or with driers such as lead white or an umber, as is the case<br />

with Caravaggio's works in the National Gallery in London (15). In a thin<br />

section taken from Poussin's Finding of Moses (National Gallery, London), one<br />

can see that the absorbency of the ground was recognized <strong>and</strong> an isolating<br />

oil layer applied. The sample was taken from a blanched area, in which both<br />

the ground <strong>and</strong> the paint film were very lean. Interestingly enough, in other<br />

areas containing lead white, the paint film was less lean <strong>and</strong> less absorbent.<br />

Symonds wrote, "It is good to give a mano [h<strong>and</strong>] of gesso first, <strong>and</strong> then oil<br />

on top." Earth from which bricks are made is ground <strong>and</strong> used for the imprimatura,<br />

Symonds wrote, recording the recipe for the ground favored by<br />

Canini: After sizing the canvas, he applied a mixture of "red earth, a little<br />

white lead, a little charcoal black, <strong>and</strong> chalk" (16). Until now the latter seemed<br />

always to have been interpreted as white clay, although brick clay seems perfectly<br />

adequate. This mixture was applied to the canvas in an oil medium.<br />

The absorbency of the ground seems to have been both feared (because of<br />

the problem of sinking in) <strong>and</strong> desired (because it kept the colors fresh) . As<br />

Symonds wrote, "These cloths that have gesso in their imprimatura-the<br />

gesso makes the colour keep fresher <strong>and</strong> does drink up the evil of the oil,<br />

but they crack sooner, <strong>and</strong> that is the worse of the gesso" (17). The sinkingin<br />

of colors, Symonds's "prosciugare," seems to be linked in their minds to<br />

the poor drying of the ground, either because the commercial primer left<br />

out the lead white for economy's sake (18), which seems to reduce the porosity<br />

of the paint film, or because it had not dried sufficiently before the<br />

painting was executed.<br />

Then, as today, the poor drying qualities of earth pigments were recognized<br />

but not fully understood. From analyses we know that when the pigment in<br />

the oil film is an earth color, there is a surprisingly high percentage of scission<br />

Glanville 15


products in the dried oil film (19). These short-chain scission products, by<br />

giving the structure a swollen matrix, confer flexibility to the oil film. Earth<br />

pigments are then particularly suitable as grounds on flexible supports. To<br />

speed the drying process of these grounds, artists advocated the use of prepolymerized<br />

oil (20).<br />

Most of the inorganic components of these colored grounds are hygroscopic<br />

<strong>and</strong> are not easily wetted by organic media. Free fa tty acids in the oil would<br />

act as surface active agents, <strong>and</strong> therefore an unpolymerized oil would be<br />

preferable to ensure wetting of the particles <strong>and</strong> pigment aggregates.<br />

Groen <strong>and</strong> Burnstock have fo und air pockets with earth pigments bound in<br />

unpolymerized linseed oil (21, 22). This problem can only be aggravated by<br />

the use of an oil with a reduced proportion of free fatty acids (23). The<br />

presence of air pockets makes the ground porous, causing the paint layers<br />

above to "sink in" as the paint medium is drawn down into the absorbent<br />

ground.<br />

Speedy drying achieved by applying the ground as an emulsion, as advocated<br />

in Pierre Lebrun's Treatise, would also result in a porous film because of the<br />

voids formed during evaporation of the aqueous phase (24). The absorbency<br />

of the ground would be particularly problematic with artists such as Guercino<br />

who, in his first phase, painted aUa prima, using the dark ground fo r his<br />

shadows <strong>and</strong> often applying only one layer of paint. Malvasia, describing<br />

Guercino's technique before the artist left fo r Rome, says that Guercino had<br />

"an extraordinary speed of execution, in one go laying-in (bozz<strong>and</strong>o) <strong>and</strong><br />

finishing" (25). In the case of Guercino, the complaints came from his clients<br />

who felt that since they paid Guercino by the figure, they wanted to see the<br />

whole figure, not one in which more than half was drowned in shadow.<br />

Lanfranco <strong>and</strong> later Giordano were also known for their speed of execution.<br />

This speed was castigated by most art theorists, because it was seen as pratica,<br />

simple manual facility rather than the fruit of matured intellectual thought<br />

<strong>and</strong> skill.<br />

Exposed or thinly covered porous grounds would pose a problem when it<br />

came to varnishing (for the restorer as well as the artist), but grounds suffering<br />

from what the French call lithargeage present a similar problem in appearance.<br />

The coarsely ground lead white would present a less absorbent ground, but<br />

the rough granular surface scatters light, obscuring detail in the worst cases<br />

<strong>and</strong> flattening the composition. Many seventeenth-century paintings suffer<br />

from this problem (26).<br />

Viewing distance <strong>and</strong> lighting<br />

It is clear that in a painting such as Guido Reni's Abduction of Helen (Louvre),<br />

which does not depend on the warmth of its shadows for the depiction of<br />

depth <strong>and</strong> its illusion of space but rather on what Leonardo termed aerial<br />

perspective <strong>and</strong> the diminution of colors, the impact oflithargeage is minimal,<br />

<strong>and</strong> saturation of the picture surface not essential.<br />

This is the case with paintings of what I have termed the "high finish" school;<br />

that is, those by painters who did not paint alIa prima, nor availed themselves<br />

in the Venetian manner of the artifice of large areas of shadow <strong>and</strong> light, but<br />

rather achieved the illusion of space through "the diminution of the hues, as<br />

much through their quality as through their strength" (27).<br />

The relative importance of saturation in paintings that rely on their dark<br />

ground for illusion of space was especially visible in a recent exhibition of<br />

Neapolitan painting in Bordeaux. Two paintings by Caravaggio were hung<br />

next to each other. One was a privately owned Doubting Thomas, with a high<br />

gloss varnish, <strong>and</strong> the other painting was a Salome with the Head of St. John<br />

the Baptist (National Gallery, London), which had a much more matte appearance<br />

(28).<br />

16<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


The differences in saturation, clearness of detail, <strong>and</strong> illusion of space were<br />

very clear at a normal viewing distance. If one stepped back far enough to<br />

be able to see both paintings, the detail <strong>and</strong> saturation of the two works<br />

appeared to be the same, <strong>and</strong> both surfaces had a matte appearance. The very<br />

effect of the air interposed between the eye <strong>and</strong> the object was alluded to by<br />

Felibien. According to this phenomenon, first objectively described <strong>and</strong> explained<br />

by Leonardo, objects in the distance will appear more blue because<br />

of the interposition of the tiny particles contained in the air.<br />

The warming, darkening effect of varnish, that of light transmitted through<br />

the varnish, will be counteracted by the blue quality of the reflected light.<br />

This means that with paintings executed to be seen at a distance, a glossy<br />

saturated varnish would be unnecessary; at the correct viewing distance, the<br />

painting will appear somewhat matte.<br />

Boschini, characteristically, writes, "Do you think painting is enjoyed more<br />

directly under the eye or from an appropriate distance? Certainly at a distance;<br />

because once you have hung the pictures <strong>and</strong> placed them at the wanted<br />

height, there's no reason for you to go clambering up there" (29).<br />

Another variable that is too infrequently considered when exhibiting paintings<br />

is the lighting. <strong>Painting</strong>s from that period were never painted in uniform<br />

daylight, nor were they to be seen in such conditions. Daylight would have<br />

been directional, <strong>and</strong> artificial light would have been warm c<strong>and</strong>lelight. Restorers<br />

sometimes have the opportunity oflighting the painting with tungsten<br />

lights from the direction in which the painting was executed, <strong>and</strong> the effect<br />

in increased naturalism <strong>and</strong> illusion of space can be quite dramatic (30).<br />

This is not a new observation. In 1628, in a text on the artist Cigoli, the<br />

biographer Commodi wrote about Cigoli's Martyrdom of St. Lawrence: "It<br />

appeared much more admirable in the room in which it was painted which,<br />

receiving little light through a small window, was indeed in proportion both<br />

with the h<strong>and</strong>ling <strong>and</strong> with what was represented in the scene. Having seen<br />

the painting in this room several times, <strong>and</strong> seeing it again in the open air,<br />

he said to himself 'This just is not the same, even though everyone praises it<br />

also out here' " (31).<br />

Viewing distance is also crucial for works executed with a painting method<br />

that defies legibility at close scrutiny. This is what we now may revere as the<br />

"magic" of painting, but to Poussin <strong>and</strong> his circle it came close to pratica,<br />

manual skill not guided by the intellect <strong>and</strong> therefore "inferior" <strong>and</strong> not<br />

worthy of painting as a liberal art.<br />

Rather surprisingly, Bellori appreciated the skill involved in producing effects<br />

to be appreciated from a distance. In his biography ofLuca Giordano, speaking<br />

of a sotto-in-su figure of Christ, he describes how Giordano would constantly<br />

descend from the scaffolding to check on the effect of his painting <strong>and</strong> the<br />

"huge brush strokes" <strong>and</strong> "coarse <strong>and</strong> garish highlights in the blue drapery,"<br />

which the distance of the eye blends <strong>and</strong> harmonizes together so beautifully<br />

<strong>and</strong> makes one underst<strong>and</strong> how great his mastery is" (32).<br />

Boschini's poem, a eulogy of Venetian painting <strong>and</strong> technique, extols the<br />

macchia of which the "high finishers" are so scornful. He carefully distinguishes<br />

this from the rapid, sketchy alla prima technique he terms prontezza (33). He,<br />

too, emphasizes the skill involved, rather than the material aspect. The Venetian<br />

macchia "is what unites the wet paint with the dry paint beneath,<br />

which is so important to the beauty of this way of painting" (34).<br />

Like Armenini <strong>and</strong> Felibien, he emphasizes the h<strong>and</strong>ling of the paint <strong>and</strong> the<br />

color, rather than its nature <strong>and</strong> strongly argues that painting should appeal<br />

to the mind <strong>and</strong> not the eye-that the spectator must see the h<strong>and</strong> inside<br />

the glove, while the eye only sees the glove (35, 36).<br />

This brings us back to the essential message of the Counter-Reformation<br />

<strong>and</strong> of the new stature of painting as a liberal art: to convey essential truths<br />

Glanville 17


epresentationally so as to appeal to the heart <strong>and</strong> mind as well as to the<br />

senses. As Horace said, "<strong>Painting</strong> should instruct as well as delight by the two<br />

ways open to it: close scrutiny or distant viewing."<br />

Notes<br />

1. Du Fresnoy, C. 1750. Observations on the Art of <strong>Painting</strong>. Dryden translation, first<br />

published in 1667.<br />

2. Boschini, M. 1660. La Carta del Navegar Pittoresco. Venice.<br />

3. Bellori, G. P. 1672. Le vite de' Pittori, Scultori e Architetti Moderni, first published<br />

in 1672. Pis a 1821 edition.<br />

4. Felibien, A. 1685. Entretiens sur les vies et les ouvrages des plus excellents peintres<br />

anciens et modernes, 2nd ed. Paris. Vo!' 2, Entretien VII:240.<br />

5. Ibid., 242.<br />

6. This passage in a different translation is quoted by Robert Ruuhrs in his article<br />

"Matte or glossy? Varnish for oil paintings in the seventeenth century," which<br />

appeared in an English translation in the ICA Newsletter, Oberlin, Ohio: November<br />

1985. The article was originally published in Maltechnik-Restauro in July 1983:<br />

"E i forestieri certo in quela parte / Fa tanto cavedal, che la lustrezza / A lori<br />

ghe par l'urUca belezza. / Ghe par che la sigila tuta l'arte."<br />

7. Horace. Ars Poetica. 361. See Lee, R. W 1940. Ut pictura poesis: the humanistic<br />

theory of painting. The Art Bulletin (XXII).<br />

8. Baldinucci, F. 1692. II Lustrato. Paper presented at the Accademia della Crusca<br />

29 December 1691 <strong>and</strong> 5 January 1692. Florence. On page 25: "Trovansi bene<br />

notizie di Pittori antichi, che fanno credere, che tale uso non vi fosse; fra Ie quali<br />

potra sempre appresso di me cio che fu scritto d' Apelle, cioe, ch' egli fu ritrovatore<br />

d'un certo color bruno, 0 vernice che si fosse, la quale rUuno seppe imitare, e<br />

davala all'opere dopo averle finite; e che servivane con tanto giudizio, che i colori<br />

accesi la vista non offendevano, facendoli vedere da lungi come per un vetro (e<br />

notate questa particolarita) e che Ie tinte lascive, mediante quella acquistavano<br />

un certo che d'austero, 0 di scuro, che e tutto quello appunto, che facevano i<br />

nostri pittori del 1300, avanti il ritrovamento della tempera coll'olio, cioe, che<br />

davano sopra Ie tavole una vernice, che era una certa mestura, che alia loro<br />

dilavata pittura un certo che di pili profondo, e di forza maggiore aggiungeva,<br />

ed il soverchio chiaro alquanto smorz<strong>and</strong>o, riduceva a maggior somiglianza del<br />

naturale ...." On page 31: "Se poi sara detto, che i moderni Pittori usano anch'essi<br />

talvolta vernice sopra Ie lor pitture a olio, io rispondo, che tale usanzach'e<br />

di pochi-non e per supplire al mancamento della pittura a olio, cioe, per<br />

render pili profondi gli scuri, e i chiari pili mortificati, e pili carnosi, cose tutte<br />

delle quali la pittura a olio non ha bisohno rna bensi per rimediare ad un'accidental<br />

disgrazia, che occorre taJora a cagione dell'imprimatura, mestica, 0 altro<br />

che dassi sopra Ie tele, 0 tavole, 0 pure proviene dalle medesime tele, 0 tavole,<br />

cioe d'atrarre cosi forte il liquido dell'olio, quasi rub<strong>and</strong>olo al colore, ch'e' venga<br />

in qualche luogo prosciugato per modo, ch'e' non possa farsi vedere in superficie<br />

per tutto ugualmente, com'egli avrebbe fa tto col cessare di tale accidente; con<br />

che per mezzo d'un'altra cosa untuosa, che e la vernice data dove I'olio in<br />

superficie manco, fassi apparire (e questo e il punto stretto e forte) con che fassi<br />

apparire 10 scuro, che gia nella pittura fatta a olio veramente e, non quello che<br />

non v'e, ch'era appunto l'effetto, che in qualche piccolissima parte faceva aile<br />

sue pitture la Vernice d'Apelle."<br />

9. In Italian, the word used is bruno. It should be noted that imbrunire is the word<br />

used for burnishing gold, that is, making it darker <strong>and</strong> increasing its sheen.<br />

10. Beal, M. 1984. A Study of Richard Symonds: His Italian Notebooks <strong>and</strong> Their Relevance<br />

to Seventeenth-Century <strong>Painting</strong> <strong>Techniques</strong>. New York: Garl<strong>and</strong> Publishing,<br />

fo!' 46.<br />

11. Ibid., fo!' 143.<br />

12. Ibid., fo!' 75.<br />

13. Ibid., fo!' 25.<br />

14. Armenini, G. B. 1587. De veri precetti de la pittura. Ravenna.<br />

15. A good example can be found in Caravaggio's Boy with Lizard. National Gallery,<br />

London.<br />

16. Symonds in Beal, op. cit., fo!' 10. "Terra Rossa, biacca da corpo a little. Creta<br />

un tantino, negro Carbone. He complayned the ordinary imprimers putt in no<br />

biacca."<br />

17. Ibid.<br />

18. Ibid.<br />

18<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


19. White, R. Personal communication. National Gallery, London. I am indebted to<br />

Raymond White for much enlightening discussion around the subject.<br />

20. Symonds, op.cit., fol. 10. "To make imprimatura wich shall st<strong>and</strong> abroad in ye<br />

ayre, tis best to use olio Cotto. Also tis good dare una mano di gesso prima /<br />

poi oglio suopra."<br />

21. Groen, K. 1988. Scanning electron microscopy as an aid in the study of blanching.<br />

Hamilton Kerr Institute Bulletin 1 :48-65.<br />

22. Burnstock, A. An examination of blanching in the green paints from a selection<br />

of seventeenth-century French School paintings. Ph.D. diss. Courtauld Institute<br />

of Art, London.<br />

23. Groen, op. cit.<br />

24. Lebrun, P. 1635. Recueuil des essaies des merveilles de la peinture. In Merrifield,<br />

M. p. 1849. Original Treatises on the Arts of <strong>Painting</strong>. Vol. II. London:John Murray,<br />

759-841. On page 821: "To prime a canvas quickly, so that a person may paint<br />

on it the same day that it has been primed, you must grind together some<br />

parchment glue <strong>and</strong> oil priming, <strong>and</strong> immediately prime the canvas with this; it<br />

will harden directly, but the priming is apt to scale off when the canvas is rolled<br />

up."<br />

25. Malvasia. 1646. Vite di Pittori Bolognesi. Unpublished addenda. In Sce/te di curiosita<br />

letterarie, inedite 0 rare dal XIII al XIX, ed. Adriana Arfelli. Speaking of<br />

Guercino's Circumcision in the Chiesa delle Monache di Gesu Maria in Bologna<br />

which was executed overnight: "La sua formidabil velocita nell'operare, e alla<br />

prima bozz<strong>and</strong>o, e fmendo nello stesso tempo."<br />

26. For instance, Caravaggio's Salome with the Head of St. John the Baptist (National<br />

Gallery London) <strong>and</strong> Mattia Pretis's The Beheading of St. John the Baptist (National<br />

Gallery of Irel<strong>and</strong>, Dublin).<br />

27. Felibien, op. cit., Entretien VIII, 401. "[L]a diminution des teintes ... autant par<br />

la qualite que par la force des couleurs."<br />

28. Either an optical effect due to the pronounced lithargeage of the ground or the<br />

inherent matte appearance of the varnish itself.<br />

29. Boschini, op. cit., 295. "Credela che se goda la Pitura / Piu soto l'ochio, 0 in<br />

debita distanza? / Certo in distanza debita: perche / Qu<strong>and</strong>o have messi i quadri,<br />

e situdi / In l'altezza de iluoghi destinadi, / Non e rason, che la ve rampeghe."<br />

30. This was the case during the examination of Gainsborough's The Linley Sisters<br />

in the Dulwich Picture Gallery. Gainsborough had painted the portrait with the<br />

light coming from the left through holes he had drilled in his shutters! See<br />

Glanville, H. 1988. Gainsborough: Artist or Artisan? (original unedited version).<br />

In A Nest of Nightingales, s.l. exhibition catalogue.<br />

31. Biography of Ludovico Cardi, in Cardi, G. B. 1628. II Cigoli. Florence, 19.<br />

"[M]olto piu ammirabil si rendeva in quella stanza dove fu lavorata, la qual per<br />

una picciol finestra ricevendo poco lume era appunto proporzionato et alla maniera<br />

et a quello che nella storia si rappresentava; et avendola ivi veduta piu volte,<br />

e, rivedendola poi all' aria aperta, diceva fra se: questa non e quella, sebbene quivi<br />

ancora lodata da tutti."<br />

32. Bellori, op. cit. Vita di Luca Giordano, 51. "[E]ssendo il tutto condotto con grossi<br />

trattizzi ne' contorni, e cosi anche il Cristo, benche sia alquanto piu terminato,<br />

come figura principale del soggetto, ne! cui panno azzurro si veggono chiari<br />

terribili di biacca imbrattata di carnatura, con tratti di pennello piu grossi, il quale<br />

unisce ed accorda tanto bene con la distanza dell'occhio, che fa comprendere<br />

quanto sia gr<strong>and</strong>e la maestria de! suo pennello."<br />

33. Boschini, op. cit., 339. "[L]a prontezza ... impastar figure in copia, e senza natural.<br />

..."<br />

34. Boschini, op. cit., 339. ''L'e que! unir quel fresco con quel seco / Che in sto<br />

operar si belo importa tanto."<br />

35. Boschini, op. cit., 294. "Perche se se vardasse al color solo, / EI piu vago sarave<br />

anche il piu belo / Ma varv<strong>and</strong>ose el trato del pene!o / Casca sta consequentia<br />

a rompicolo."<br />

36. Boschini, op. cit., 290. "[L]'ochio no'vede altro, che'l guanto."<br />

Glanville 19


Abstract<br />

French publications on painting materials<br />

<strong>and</strong> techniques before 1800<br />

include books of "secrets," treatises<br />

written by artists themselves or<br />

about their practice, articles from<br />

learned journals, dictionaries, <strong>and</strong><br />

encyclopedias. The traditional methods<br />

of painting listed in these publications<br />

include dhrempe (a reference<br />

to a water-based media, which can<br />

be glue, gum, or egg tempera), fresco,<br />

miniature, <strong>and</strong> oil painting,<br />

enamel <strong>and</strong> glass painting, as well as<br />

painting on porcelain. Some of the<br />

more unusual techniques are outlined:<br />

encaustic, eludoric, <strong>and</strong> glass<br />

painting, including glass transfer<br />

techniques, <strong>and</strong> finally a satirical<br />

contribution, "peinture en fromage<br />

ou en ramequm."<br />

From. Books of Secrets to Encyclopedias: <strong>Painting</strong><br />

Te chniques in France between 1600 <strong>and</strong> 1800<br />

Ann Massing<br />

Hamilton Kerr Institute<br />

Whittlesford<br />

Cambridge CB2 4NE<br />

United Kingdom<br />

Introduction<br />

Until the mid-seventeenth century, the business of painting in France was<br />

strictly regulated by guilds. The skills of the painting craft, h<strong>and</strong>ed down from<br />

master to apprentice, were treated as valuable personal possessions <strong>and</strong> protected<br />

by secrecy (1). Few records of the painting techniques themselves were<br />

published before the mid-eighteenth century. The first information contributing<br />

to the history of painting materials <strong>and</strong> techniques in the French language<br />

is found in books of "secrets," collections of recipes on various topics,<br />

often compiled from diverse authors. The "secrets" of the earliest such books<br />

were not contemporary recipes, however, but translations from Italian or Latin,<br />

often with reference to ancient Greek or Roman authors.<br />

Books of secrets<br />

The first of the books of "secrets," Les secrets de reverend Alexis Piemontois (Paris,<br />

1557) was a compilation, translated from the Italian, that included remedies<br />

for illness, wounds, <strong>and</strong> accidents, as well as instructions on how to make<br />

perfumes,jams, dyes, <strong>and</strong> pigments. The book was reprinted in 1573 <strong>and</strong> was<br />

exp<strong>and</strong>ed to include recipes from Dioscorides, Galen, <strong>and</strong> others, as well as<br />

some recipes relating to the history of painting techniques such as gilding,<br />

pigments, inks, <strong>and</strong> the making of varnishes. It was republished several times<br />

during the sixteenth <strong>and</strong> seventeenth centuries, as were other secret books,<br />

such as Les secrets et merveilles de nature (Lyon, 1586) by Jean Jacques Wecker,<br />

a doctor from Colmar. Wecker's recipes fo r magic <strong>and</strong> sorcery, health, cooking,<br />

beauty, <strong>and</strong> the manufacture of pigments, "des secrets des falseurs et vendeurs<br />

de couleurs et peintures," were also taken from Latin sources; he credits seventy-five<br />

authors from Aristotle onward. Fortunately, the book includes a<br />

table of contents, fo r the recipes are not arranged alphabetically but in a<br />

systematic order that reflects Wecker's vision of the universe <strong>and</strong> places painter's<br />

secrets on the same level as recipes fo r cooks, druggists, carpenters, <strong>and</strong><br />

Jomers.<br />

Several books of secrets were published in the seventeenth century, but perhaps<br />

the most popular <strong>and</strong> the most relevant to our topic was Le Sieur<br />

D'Emery's Recueil des curiositez rares et nouvelles des plus admirables iff ets de la<br />

nature ret de l'art] (Paris, 1674) (2). Included were many recipes related to the<br />

history of painting techniques, recipes on how to copy drawings, make pastels,<br />

imitate marble, stain wood, gild, <strong>and</strong> make engravings look like old master<br />

paintings.<br />

Books of secrets continued to be written <strong>and</strong> reprinted throughout the eighteenth<br />

century <strong>and</strong> beyond. In the eighteenth century, the two compilations<br />

most frequently reissued were Secrets des arts et metiers (Brussels, 1755) <strong>and</strong><br />

L' Albert moderne ou nouveau secrets eprouves et licites (Paris, 1768). This "modern"<br />

Albert was a revision of "old" Albertus Magnus, with the superstitions <strong>and</strong><br />

enchantments deleted, keeping only "useful" advice, such as how to cure a<br />

toothache with two live moles. (The method begins by holding a mole in<br />

each h<strong>and</strong> <strong>and</strong> squeezing gently, without letting go, until they die; this process<br />

should take about five hours.) Large sections are devoted to medical <strong>and</strong><br />

agricultural recipes; the third section includes "divers moyens de se faire une<br />

20<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


occupation agreable pour charmer l'ennui," with recipes about liqueurs, flowers,<br />

painting secrets, <strong>and</strong> so on.<br />

Technical treatises<br />

The most widely read of the early painting manuals, judging by the number<br />

of editions, was Claude Boutet's treatise on miniature painting, published in<br />

1672, called by the author an "ABC de La mignature" (3). Following the conventions<br />

of his time, Boutet dedicated his treatise not to the specialist but to<br />

an amateur, a Mlle. Fouquet, who wished to perfect her abilities in the craft.<br />

In the introduction, Boutet claims his "secrets" came from the Italians, <strong>and</strong><br />

states that although he could have profited by keeping them to himself, he is<br />

giving away his knowledge fo r the benefit of others. He provides an extremely<br />

useful summary of French painting practice in the second half of the seventeenth<br />

century. The treatise discusses the various pigments <strong>and</strong> how to use<br />

them, based on a three-step sequence (as in oil painting of the period) : ebaucher,<br />

pointiller, <strong>and</strong> finally finir or rehauser.<br />

A seventeenth-century treatise of great importance for the history of technique<br />

is De La Fontaine's Academie de La peinture (Paris 1679), dedicated to<br />

the tutor of the future king of France (4) . He summarizes much practical<br />

detail, especially on pigment mixtures, mentioning dhrempc, fresco, pastel, grisaille,<br />

<strong>and</strong> miniature as well as oil painting. He refers to the art of perspective,<br />

which he believes a painter should know both theoretically <strong>and</strong> practically.<br />

His dependence on Italian painting is evident, <strong>and</strong> he lists the names of the<br />

Antique painters who have contributed to this knowledge (5). He also discusses<br />

the origin of painting; although several "philosophers" attributed the<br />

origin of painting to a shepherd who began to trace his shadow <strong>and</strong> saw the<br />

resemblance to the human fo rm, De La Fontaine allows that the more common<br />

opinion is that the discovery of painting was made by the Hebrews <strong>and</strong><br />

that they transmitted it to the Greeks <strong>and</strong> to the Romans. Until the late<br />

eighteenth century, it was believed that this knowledge was then "lost," to be<br />

"rediscovered" by "Jean de Bruges" Oan van Eyck) .<br />

Seventeenth-century painters thus considered oil painting a "new" method<br />

of painting, <strong>and</strong>, judging from the source books, the preparation of the medium<br />

was obviously a vexing problem. De La Fontaine suggests nut oil with<br />

lead white thickened by sunlight (6). He follows this with the recommendation<br />

that in order to dry a layer oflead white, or a grisaille layer, one should<br />

mix the white pigment with oil of turpentine, which will cause the layer to<br />

dry as the turpentine evaporates (7). This is followed by advice on how to<br />

prevent layers of drapery from running. He suggests placing the painting flat<br />

on the floor or on a table <strong>and</strong> scattering small pieces of absorbent paper onto<br />

the surface, especially onto the shadows of the drapery. When the paper has<br />

drawn out the oil, but before the paint layer has dried, one should pick up<br />

the painting <strong>and</strong> let it fall gently on one corner, so that the paper comes away<br />

(8) .<br />

Although it is not possible to go into detail about early technical practice<br />

here, I will mention briefly De La Fontaine's method of preparing supports<br />

for painting, methods used well into the eighteenth century. In the case of<br />

canvas, the material was stretched onto a loom, smoothed with a pumice<br />

stone, sized with glue, <strong>and</strong> then given a double ground, the first colored with<br />

raw umber <strong>and</strong> red brown spread on with a knife, the second with lead white<br />

<strong>and</strong> just enough carbon black to make gray. For panels, three layers of a fine<br />

chalk ground ("blanc d'Espagne, comme on Ie vend chez les Ch<strong>and</strong>eliers")<br />

was recommended, with either another layer of glue on top or a gray oil<br />

layer (9).<br />

The Acadernie Royale <strong>and</strong> the theoretical treatises<br />

Before 1800 the history of oil painting in France, <strong>and</strong> consequently the history<br />

of its technique, was dominated by Italian influence. In the sixteenth<br />

Massing 21


century, the royal patronage for the School of Fontainebleau was crucial. Later,<br />

most ambitious French painters traveled to Italy, often spending much of their<br />

professional lives there. The influence of the French Academy, the Academie<br />

Royale de Peinture et de Sculpture founded in 1648, was also of central<br />

importance for any aspect of the history of painting in France. By 1655 it<br />

had become a royal enterprise <strong>and</strong> was soon the most powerful art institution<br />

in Europe. The purpose of the Academy was to convey the principles of art<br />

to its members by means of lectures <strong>and</strong> to instruct students through life<br />

classes. In 1673 the Academy also began organizing exhibitions fo r its members.<br />

These exhibitions were not opened to a wider public until 1791 (10).<br />

With the noble aim of nurturing good craftsmanship, the Academy defined<br />

the principles from which painters were not allowed to deviate. The goal was<br />

to train students in one particular style of drawing. The necessity of copying<br />

from the ancients was stressed, <strong>and</strong> this emphasis continued until the French<br />

Revolution. Nowhere outside the Academy was a life class allowed, even in<br />

an artist's private studio. The Academy did not provide fo r the whole of the<br />

professional education of a young painter, however; the student painter still<br />

learned the basics of his craft in the workshop of his master, in whose house<br />

he often lived, just as in the Middle Ages. Thus, craft "secrets" were still<br />

essential to a painter's success, <strong>and</strong> only gradually were detailed accounts of<br />

methods <strong>and</strong> materials forced into print. The publication of Diderot's encyclopedia<br />

in the mid-eighteenth century <strong>and</strong> the French Revolution a few<br />

years later provide evidence of the culmination of a trend toward the dissemination<br />

of knowledge which had proved impossible to stem.<br />

The personalities of Jean-Baptiste Colbert (1619-1683), the elected "Protector"<br />

of the Academy, <strong>and</strong> Charles LeBrun (1619-1690), "Premier Peintre du<br />

Roi," were crucial in the development of the Academy <strong>and</strong> thus to the history<br />

of French painting. Since it was a royal academy, the king's (i.e., Colbert's)<br />

intentions were imposed, <strong>and</strong> Colbert's tight dictatorship led to a centralization<br />

from which France has still not fully emerged. The approval of the<br />

Royal Academy was a necessity fo r every endeavor within its sphere of influence.<br />

Colbert was also responsible fo r the creation of many other academies<br />

in addition to the Academie de la Peinture, including the Academie des<br />

Sciences in 1666; the importance of this fact to the subject will be indicated<br />

later.<br />

The lectures of the Royal Academy were the basis fo r many of the published<br />

treatises of the seventeenth <strong>and</strong> even the eighteenth centuries. Aesthetics,<br />

beauty <strong>and</strong> proportion, lighting <strong>and</strong> perspective, <strong>and</strong> the expressions of the<br />

figures portrayed were discussed at length by seventeenth-century authors.<br />

Charles LeBrun's famous lectures on human expression (1668) influenced<br />

generations of artists. Most of the academic authors, such as Rol<strong>and</strong> Freart<br />

de Chambray, Andre Felibien, Charles Alphonse Du Fresnoy, <strong>and</strong> Roger De<br />

Piles, were not professional painters, however, <strong>and</strong> were more concerned with<br />

theory than practice. The dispute between the Ancients (Poussinistes) <strong>and</strong> the<br />

Moderns (Rubenistes) dominated the Academy lectures for more than twenty<br />

years. The study of the lives of ancient Greek <strong>and</strong> Roman as well as Renaissance<br />

painters was also a popular subject, as was the history of the origin<br />

<strong>and</strong> "rediscovery" of painting. Although the academic lectures did not usually<br />

discuss techniques, Jean Baptiste Oudry's (1686-1 755) lectures were an important<br />

exception. As president of the Academy, Oudry gave a lecture in 1752<br />

on painting techniques as practiced by the members (1 1).<br />

The spokesman of the Poussinistes, Roger De Piles (1635-1709), an amateur<br />

painter, connoisseur, <strong>and</strong> member of the Academy, was a prolific writer on<br />

the theory of painting. He did, however, compose one book on technique in<br />

1684, Les premiers Clemens de la peinture practique. As edited <strong>and</strong> augmented by<br />

Charles Antoine Jombert in 1776, the text became the most important <strong>and</strong><br />

informative collection of recipes on French painting technique of the seventeenth<br />

<strong>and</strong> early eighteenth centuries (Figs. 1,2). Another important author<br />

fo r the history of technique was Philippe de La Hire of the Academie Royale<br />

22<br />

Histon·cal <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


DE<br />

PEJN'FURE PRATIQUE.<br />

Par<br />

demit::<br />

M. DE.<br />

Royale<br />

PILES,<br />

de Pcinture<br />

de l'Aca­<br />

Sculpture.<br />

&<br />

NOUVELLE EDITION<br />

Entierement refonduc. & augmende con·<br />

lidefablement.<br />

Par CHARL!S-ANTOna J olinItY; ,<br />

-I$<br />

A AMSTERDAM lIT<br />


..<br />

did sink, however, he suggested applying with a small sponge a layer of nut<br />

oil mixed with a siccative before continuing to paint. He added that although<br />

this layer should dry before adding fu rther details, it was possible to paint fine<br />

lines of architectural elements while the layer is still wet (12).<br />

Learned journals <strong>and</strong> periodicals<br />

l'! ENCYCLOP BDIE,<br />

ou<br />

DICTIONNAIRE RAISONNE<br />

DES SCIENCES,<br />

PAR. UNE SOCIETE DE GENS DE LETTRES.<br />

DES ARTS ET DES METIERS.<br />

Nio"'onIr


In spite of the laments of certain people who complain if the taste of the<br />

century fo r dictionaries, this predilection is increasing, a proof if the benifrt<br />

the public is receiving. We are inundated, <strong>and</strong> if this torrent is not stopped,<br />

we will be learning only from dictionaries . ... We want to know everything-or<br />

rather speak if every thing <strong>and</strong> pretend to be ignorant if nothing.<br />

We must submit, therifore, to this "gout du siecle. JJ<br />

Encaustic painting<br />

The history of oil painting was an eighteenth-century fascination. In the mideighteenth<br />

century Diderot still believed, as De La Fontaine previously asserted<br />

in his seventeenth-century treatise, that the art of painting was reborn<br />

when "Jean de Bruges" Gan van Eyck) discovered the "secret" of oil painting.<br />

It was generally accepted that the Ancient painters possessed superior techniques,<br />

<strong>and</strong> that it should be possible to "rediscover" their methods. By the<br />

eighteenth century there was also concern about the state of seventeenthcentury<br />

paintings, which no longer retained the freshness of those more recently<br />

painted (23).<br />

MEMOIRE<br />

U. PIINTURE A L'ENCAUSTIQUE<br />

SU k U. PiINTUl.t A LA CIIE.o<br />

,.. JI. C-::.:..:::.. *r"' ...<br />

f>JI._.nr.e-"Io.F_.M6iI


ived from the Greek words fo r oil <strong>and</strong> water, the two main components of<br />

the method. The method follows (27):<br />

The support, usually of fine cloth primed with poppy oil, is placed under<br />

a thin film


While cooking their eggs, two young Spartans quarreled over a piece if<br />

cheese which had just disappeared in the frying pan. Both young men armed<br />

themselves with a spoon <strong>and</strong> searched the scrambled eggs, but in vain. The<br />

cheese was gone. They came to blows, one claiming the cheese was still<br />

there, the other that it had disappeared, carried off by a genie. The noise<br />

if this battle drew a crowd, <strong>and</strong> the cause if the dispute became known.<br />

It was suggested that the experience be repeated. The ability if egg yolk<br />

to dissolve cheese was recognized.<br />

The author consoled himself that he was not the first to have discovered this<br />

secret, <strong>and</strong> so made it public, congratulating himself on his generosity. Others<br />

might have kept the critical ingredient secret in an attempt to profit from it.<br />

Selling his secret would have been underst<strong>and</strong>able, after all, as reimbursement<br />

fo r the heavy expenses in eggs <strong>and</strong> cheese. Rouquet continues his description<br />

with all the seriousness of a technical treatise:<br />

Mix gruyere cheese cut in f ine strips with two beaten egg yolks over a bainmarie<br />

until the cheese has melted. For the question of a support fo r the<br />

paintinsince cheese does not adhere well to panel or canvas, it is better<br />

to follow its natural association with bread. Therifore, take flour <strong>and</strong> make<br />

dough with a little milk. Finally, it would be bemificial to add a bitter<br />

substance to discourage worms, mice, <strong>and</strong> children from eating the paintings.<br />

However, by leaving it out, poor painters could at least dine on their own<br />

paintings.<br />

Conclusion<br />

As we approach the twenty-first century, our curiosity increases about how<br />

paintings of past generations were created. More <strong>and</strong> more, we look toward<br />

modern methods of scientific analysis to answer our questions about historical<br />

painting materials, but a return to the written sources on painting techniques<br />

is an important first step toward a proper underst<strong>and</strong>ing. By definition, painters<br />

are practically oriented, however, <strong>and</strong> have rarely composed with the pen<br />

as well as the brush; records of their materials <strong>and</strong> techniques must be plucked<br />

from various publications. Undoubtedly during the coming decades, historians<br />

of painting technique will be recovering more information from the<br />

source books about the history of studio practice, painting materials, <strong>and</strong><br />

techniques.<br />

Notes<br />

1. Patent rights existed in Europe in the mid-eighteenth century, but enforcement<br />

came only much later; in order to benefit from a discovery, the inventor still<br />

relied on secrecy. Singer gives an example involving an improvement to the<br />

system of production of liquid bleach made in 1789 fo r which the patent owner<br />

enjoyed the protection of his patent for only four years. Singer, c., et al. 1958.<br />

A history oj technology (4): Oxford, 247.<br />

2. A book of secrets, Essay des merveilles de nature et des plus nobles entifices (Rouen,<br />

1622), proved so popular that by 1657, it was already in the thirteenth edition.<br />

The author was Etienne Binet, the pseudonym of Rene Franyois, Predicateur<br />

du Roy.<br />

3. Boutet, C. 1672. Traite de la mignature, pour apprender aisement a peindre sans martre,<br />

et Ie secret de Ja ire les plus belles couleurs, I'or bruny, et I'or en coquille, Paris. Other<br />

editions were published with slight variations in the title <strong>and</strong> contents. Some<br />

editions appeared anonymously as Escole de la mignature. At least twenty-five<br />

editions were published between 1674 <strong>and</strong> 1800.<br />

4. De La Fontaine. 1679. L'Academie de la peinture [etc.], Paris. De La Fontaine<br />

dedicates his treatise to "Mesire Charles de Sainte Maure, . .. Gouverneur de<br />

Monseigneur Ie Dauphin," for the education of the future king. Two other important<br />

treatises of the seventeenth century are now available in facsimile editions.<br />

These are Bernard Dupuy du Grez's 1699 Traite sur la peinture pour en apprendre<br />

la theorie, & se peifectionner dans la pratique, To ulouse: J & A. Pech; <strong>and</strong> Le Blond<br />

de la Tour's 1669 Lettre a un de ses amis, contenant quelques instructions touchant la<br />

peinture. Bourges et Bordeaux. By the eighteenth century, the number of treatises<br />

Massing 27


increased considerably; the reader is best referred to Massing's 1990 bibliography<br />

(available from the author) .<br />

5. De La Fontaine, op. cit., epistre aiiii: "Je pretend les instruire Des jeunes] au<br />

melange des couleurs, a la force des teintes, & a la union des jours & des ombres<br />

que les Romains nomment clair-obscur."<br />

6. De La Fontaine, op. cit., 15: "Autre huille grasse pour faire seicher les couleurs.<br />

Vous prendres de I'huile de noix dedans une fiolle, vous y mettres de la mine<br />

de plomb, & du blanc de plomb pille ensemble, la mettres au soleil, cela s'engraisse,<br />

& sera fo rt clair: on en mesle avec des couleurs, qu<strong>and</strong> on est presse pour<br />

les faire."<br />

7. De La Fontaine, op. cit. "Beau secret pour faire seicher Ie blanc de plomb sans<br />

changer. Qu<strong>and</strong> vous voudres faire seicher du blanc de plomb, couche entiere,<br />

ou bien avec du gris, en grisaille, vous prendres de l'huile de terebentine meslee<br />

avec vostre blanc, il Ie fera seicher: notes que I'huile de terebentine fait tout<br />

seicher, parcequ'elles'en va en fu mee. " [Author's emphasis.]<br />

8. De La Fontaine, op. cit., 17.<br />

9. De La Fontaine, op. cit., 27-28. For the preparation of a copper plate: ''Lon<br />

prend une plantaine de cuivre bien poly, apres vous prendrez du blanc de plomb<br />

bien broye avec de la terre d'ombre, & noir de charbon mesle ensemble, avec<br />

une brosse vous frotterez par dessus la plataine bien unie, & avec un linge & du<br />

cotton dedans vous frapperez dessus, pour la rendre mieux unie, qu<strong>and</strong> la couleur<br />

sera seiche vous prendrez un cousteau & passerez par dessus pour unir davantage,<br />

apres vous la chargerez encore une fois, & ferez encore de mesme; Ie laissant<br />

seicher, vous pouvez travailler apres pour peindre sur une pierre ou piastre ...."<br />

10. The history of the French Academy has been written by several authors. See<br />

Pevsner, N. 1940. Academies of Art Past <strong>and</strong> Present. Cambridge; <strong>and</strong> more recently,<br />

Tradition <strong>and</strong> Revolution in French Art 1700- 1880. 1993. London: National Gallery.<br />

11. Oudry. 1753. Discours sur la pratique de la peinture et ses procedes principaux: ebaucher,<br />

peindre a fo nd, et retoucher. Paper presented to the French Academy, 2 December<br />

1753, published from the manuscript in the Ecole des Beaux-Arts. In Le Cabinet<br />

de l'amateur, 1861-1862. 1863. Ed. E. Piot. Paris, 107-17. For fu rther discussion<br />

of Oudry's lecture, especially as it relates to his use of varnish, see Swicklik, M.<br />

1993. French painting <strong>and</strong> the use of varnish, 1750-1900. In Conservation Research.<br />

Washington, D.c.: National Gallery of Art, 157-74.<br />

12. La Hire, Philippe de. 1730. Traite de la practique de la peinture. Paris, 710-14.<br />

13. PJ. Macquer published several small brochures printed by the Imprimerie Royale<br />

containing information on pigments used both in the dyeing industry <strong>and</strong> for<br />

painting. In the series Description des arts et metiers fa ites 014 approuvees par Messieurs<br />

de l'Academie Royale des Sciences, many articles were published which were publications<br />

in their own right; subjects include pigments used fo r dyeing, pigments<br />

for painting on porcelain, different types of glues, <strong>and</strong> the making of parchment.<br />

14. For example, a translation of a publication of the Royal Society of London on<br />

minerals by E. Hussey with reference to their use in painting was published by<br />

Dijonval. See Delaval. 1778. Recherches experimentales sur fa cause des changemens de<br />

couleurs dans les corps opaques et naturellement colores. Translated from the English<br />

by Dijonval. Quatremere: Paris de l'imprimerie de Monsieur.<br />

15. Mengs, A.-R. 1786. Oeuvres completes d'Antoine-Raphael Mengs, premier peintre du<br />

roi d'Espagne, &c. Contenant d!ff erens, traites sur la theorie de la peinture. Traduit de<br />

l'italien. Vols. I-II. Paris. Mengs wrote various treatises on the theory of art. See<br />

also Lairesse, Gerard de. 1787. Le gr<strong>and</strong> livre des peintres, 014 f'art de la peinture<br />

considere dans toutes ses parties, et de,nontre par principes; avec des riflexions sur les<br />

ouvrages de quelques bons ma/tres, et sur les difauts qui s'y trouvent. Auquel on a joint<br />

les principes du dessin du meme auteur. Traduit du Holl<strong>and</strong>ois sur la seconde edition.<br />

Vols. I-II: Paris. (Dutch original: Groot Schilderboek. Haarlem. 1740) . Lairesse was<br />

considered among the best painters of his time. His treatise was well known to<br />

eighteenth-century writers <strong>and</strong> artists, <strong>and</strong> his work is representative of practice<br />

in the early eighteenth century. Book IV is entirely devoted to colors <strong>and</strong> coloring;<br />

there are also chapters on drawing, composition, portraits, ceiling painting,<br />

engraving, etc.<br />

16. Observations periodiques sur fa physique, et les beaux-arts continued from 1756 to<br />

1823, although the title changed to Journal de Physique et chemie d'histoire naturelle<br />

et des arts.<br />

17. In the Journal Oeconomique, 1751-1772, in addition to various articles of interest,<br />

the tables of merch<strong>and</strong>ise arriving into the ports of London, Amsterdam, etc.,<br />

were published, with prices <strong>and</strong> exchange rațes for various materials including<br />

28<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


some used for painting (i.e., amber, brazilwood <strong>and</strong> cochineal, gum arabic, indigo,<br />

saffron, s<strong>and</strong>arac, shellac, verdigris, vermilion, etc.).<br />

18. The Royal Academy in London was founded in 1768. By that period, the exchange<br />

of information all over Europe was so rapid that translations of new<br />

methods <strong>and</strong> techniques happened very quickly. See also Carlyle, L. 1991. A<br />

critical analysis of artists' h<strong>and</strong>books, manuals <strong>and</strong> treatises on oil painting published<br />

in Britain between 1800-1900: with reference to selected eighteenthcentury<br />

sources. PhD. diss. Courtauld Institute of Art, University of London.<br />

Also, compare early German source books, including Schiessl, U. 1989. Die<br />

deutschsprachige Literatur zu Werkstoifen und Techniken der Malerei von 1530 bis ca.<br />

1950. Worms.<br />

19. The layout of Diderot's great work is not alphabetical but systematic, after the<br />

division of the sciences following the scheme of Francis Bacon (1561-1626) fo r<br />

an Encyclopaedia oj Nature <strong>and</strong> Art, which was published in 1620 (the encyclopedia<br />

itself was never published). <strong>Painting</strong> is listed under two categories: philosophy<br />

<strong>and</strong> imagination. Alphabetical arrangement of the contents of an encyclopedia<br />

came into general use in the eighteenth century, <strong>and</strong> indexing was not employed<br />

until the 1830s. For fu rther information of the dramatic history of Diderot's<br />

encyclopedia, see Collison, R. 1964. Encyclopaedias: Their History throughout the<br />

Ages. London.<br />

20. Copying from previous authors was commonplace in the seventeenth <strong>and</strong> eighteenth<br />

centuries; everyone did it, <strong>and</strong> it was unusual when a source was cited.<br />

One splendid example is the recipe for making drying oil by boiling nut oil<br />

with litharge, water, <strong>and</strong> one onion. None of the authors really believed in the<br />

efficacy of the onion, but dutifully copied it nonetheless. See Pernety, D. A,-J.<br />

1757. Dictionnaire portatif de peinture, sculpture et gravure. Paris, Ixxxvii-Ixxxviii.<br />

Pernety copied the recipe from La Hire (La Hire, op. cit., 708): "II y en a qui<br />

font cuire avec l'huile un oignon coupe en plusieurs morceaux pour la degraisser<br />

& pour la rendre plus coulante & moins gluante, a ce qu'ils pretendent." De Piles<br />

copied it again with similar doubts <strong>and</strong> suggested using a piece of bread instead<br />

(De Piles. 1776. Elemens de peinture pratique. Charles-Antoine Jombert, 141-42).<br />

21. Lacombe, J. 1752. Dictionnaire portatif des beaux-arts. Paris. See also Macquer,<br />

p. J. 1766-1767. Dictiormaire portatif des arts et metiers. Paris. For information on<br />

the many dictionairies published in the later half of the eighteenth century, see<br />

Massing's 1990 bibliography (available from the author).<br />

22. Pernety, op. cit. (note 20), iii-iv.<br />

23. Several authors refer to the imperfections of the oil medium. De Piles. 1766, op.<br />

cit. (note 20), 97-98. De Piles writes of the deterioration of oil painting, "Cette<br />

espece de peinture est moderne en comparison des autres . ... II n'y a pas de<br />

donte qu'elle seroit la plus parfaite de toutes les manieres de peindre, si les<br />

couleurs ne se ternissoient point par la suite des tems; mais elles brunissent<br />

toujours de plus en plus & tirent sur un jaune brun, ce qui vient de I'huile avec<br />

laquelle to utes les couleurs sont broyees & incorporees."<br />

24. Caylus, M. Ie Comte de, <strong>and</strong> M. Majault. 1755. Memoire sur la peinture a l'encaustique<br />

et sur la peinture a la eire. Geneva <strong>and</strong> Paris.<br />

25. Eighteenth-century authors such as the Count of Caylus believed that the Ancient<br />

Greek painters used an encaustic painting technique. The wax was "fixed"<br />

to the wooden support by placing the completed painting next to an open fire.<br />

Since the only painting materials Pliny mentioned were wax, pigments, fire, <strong>and</strong><br />

brushes, Caylus dismissed the use of solvents such as turpentine to soften wax<br />

<strong>and</strong> make it "paintable." To paint as the ancients did, Caylus considered that the<br />

wax needed to be heated. The watertight metal boxes depicted here could be<br />

filled with boiling water, which kept the colored waxes at a constant temperature.<br />

The box on the left has a roughened glass top used to grind the colors. A similar<br />

box with an opaque top would be a palette.<br />

26. Fratrel,J. 1770. La cire alliee avec l'/wile ou la peinture a /wile-eire. Trouvee a Manheim<br />

par M. Charles Baron de Taubenheim experimerttee decrite & dedie a l'electeur. Manheim.<br />

27. For further discussion of this method of painting, which was taken quite seriously<br />

not only by the inventor himself, but by most contemporary source books, see<br />

Massing, A. 1993. Arnaud Vincent de Montpetit <strong>and</strong> eludoric painting. Zeitschrifi<br />

jur Kunsttechnologie 7 (2):359-68.<br />

28. For a discussion of this popular technique, see Massing A. 1989. From print to<br />

painting. The technique of glass transfer painting. Print Quarterly 6 (4):382-93.<br />

29. Rouquet,J. A. 1755. L'art nouveau de la peinture en jromage, ou en remequin, inventee<br />

pour suivre Ie louable projet de trouver graduellement des jafons de peindre injerieures a<br />

celles qui existent. Marolles.<br />

Massing 29


Abstract<br />

The Roberson Archive, the archive<br />

of the artists' colorman Charles<br />

Roberson & Co. (1819-1985) is<br />

housed at the Hamilton Kerr Institute<br />

in Cambridge. Roberson was<br />

one of the most important colormen<br />

in nineteenth-century London. The<br />

material in the archive gives detailed<br />

information about the internal<br />

workings of the company <strong>and</strong> its relations<br />

to both its customers <strong>and</strong><br />

suppliers. It is the largest artists' colorman<br />

archive in the United Kingdom<br />

<strong>and</strong> covers the period from<br />

1815 to 1944, providing a record of<br />

the materials <strong>and</strong> techniques of<br />

many of the leading nineteenth- <strong>and</strong><br />

twentieth-century British artists. In<br />

addition to catalogues, sample books,<br />

<strong>and</strong> more than three hundred account<br />

ledgers, there is a collection of<br />

pigments <strong>and</strong> objects sold by Roberson<br />

that provides material fo r analysis<br />

<strong>and</strong> study. A three-year research<br />

project is currently being undertaken<br />

to catalogue the archive, compile a<br />

database of the most important documents,<br />

<strong>and</strong> make the archive accessible<br />

for research.<br />

Figure 1. Roberson's shop at 154 Piccadilly,<br />

ca. 189{}-1911. Fitzwilliam Museum, University<br />

of Cambridge.<br />

The Roberson Archive: Content <strong>and</strong> Significance<br />

Sally A. Woodcock<br />

Hamilton Kerr Institute<br />

University of Cambridge<br />

Whittlesford<br />

Cambridge CB2 4NE<br />

United Kingdom<br />

The company's history<br />

Charles Roberson opened his first recorded shop to sell artists' materials at<br />

54 Long Acre, London. Although accounts of the firm's history give the date<br />

as 1810, Henry Matley is listed as "colourman to artists" at this address in a<br />

trade directory of 1817, <strong>and</strong> a note in a recipe book dating from around the<br />

turn of the century states that Roberson did not succeed Matley until 1819<br />

(1). At that time, the shop was in the heart of the artistic area of London,<br />

with the Royal Academy Schools being based nearby in Somerset House on<br />

the Str<strong>and</strong>. In 1828 Roberson <strong>and</strong> his assistant Thomas Miller became partners<br />

<strong>and</strong> moved to 51 Long Acre (2). The partnership dissolved in 1840.<br />

The company continued to trade on Long Acre, moving to the premises they<br />

built at number 99 in 1853. In 1868 the Royal Academy Schools moved to<br />

Burlington House, Piccadilly, <strong>and</strong> Roberson opened a retail branch on Piccadilly<br />

in 1890 (Fig. 1) (3). Over the following years, however, many artisans<br />

<strong>and</strong> manufacturing trades moved away from the center of London to be<br />

replaced by retailers <strong>and</strong> offices; in 1937 Roberson closed both the Piccadilly<br />

<strong>and</strong> Long Acre branches, moving to Parkway, Camden (4). They retained a<br />

West End presence for an unspecified length of time, arranging with the<br />

Medici Galleries in Grafton Street to move a "representative stock of Artists<br />

materials" from their closed Piccadilly branch in January 1940, to be sold<br />

from the galleries on commission (5). Two successive French addresses appeared<br />

in their catalogues after 1870, which they described as their depots in<br />

Paris, but which were in fact the premises of Parisian colormen (6).<br />

The company remained in the Roberson family until sold to a Dutch firm<br />

in the 1970s. G. F. Roberson Park remained a board member (7). In 1985 it<br />

went into liquidation. At liquidation the name was bought by the London<br />

colorman Cornelissen <strong>and</strong> a small range of high-quality materials bearing the<br />

Roberson name is still produced (8).<br />

The arrival of the archive at the Hamilton Kerr Institute<br />

In 1975 Roberson's historic accounts ledgers were moved to the Hamilton<br />

Kerr Institute (9). They were loaned to the Institute for safekeeping <strong>and</strong><br />

research <strong>and</strong>, when the company was dissolved, became part of the Fitzwilliam<br />

Museum's manuscript collection. The archive is now housed in the Institute<br />

<strong>and</strong> will be available for research once cataloguing is complete.<br />

The archive consists of a collection of objects, including a large number of<br />

pigment samples, <strong>and</strong> more than three hundred ledgers dating from 1815 to<br />

1944. There are several continuous series of different types of ledgers, but<br />

there are many gaps, <strong>and</strong> some sets appear to be incomplete (10). Despite<br />

this, it is the most extensive artists' colorman archive in the United Kingdom.<br />

The manuscripts in the Roberson Archive give a detailed picture of the<br />

company's activities. Only the most important categories are discussed here,<br />

as many of the warehouse, order, day, <strong>and</strong> sundries books contain information<br />

duplicated in the main accounts books.<br />

Catalogues <strong>and</strong> sample books<br />

The catalogues preserved in the archive date from 1840-1 853 to 1926-1933<br />

<strong>and</strong> illustrate goods sold by Roberson. They offered a fairly conventional<br />

30<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


ange of artists' materials, books on art, <strong>and</strong> plaster casts of antique statues.<br />

Although they sold some patented equipment (such as a sketching stool that<br />

converted to a walking stick <strong>and</strong> the "Eiffel," a stool made of steel), Roberson<br />

avoided the more innovative artists' equipment patented in the nineteenth<br />

century, such as easels that also served as bicycles or rowing machines (11).<br />

Most products were sold under the Roberson name, although a few materials<br />

such as paper were usually identified by the name of the manufacturer. It is<br />

therefore difficult to establish which materials were made by Roberson <strong>and</strong><br />

which brought in to be made up <strong>and</strong> labeled on the premises. It appears that<br />

the company prepared their own canvas throughout the company's history, a<br />

room used fo r this was still extant in the Parkway workshop until liquidation<br />

(12). They also mixed their own paints, buying pigments <strong>and</strong> other materials<br />

from wholesale suppliers, in common with most other artists' colormen, not<br />

manufacturing the raw pigments themselves. In general they only bought<br />

small quantities of prepared paints from other artists' colormen, sometimes as<br />

little as one tube or cake bought for a specific customer. However, the popularity<br />

of Roberson's Medium, used fo r oil painting, was such that both British<br />

<strong>and</strong> foreign colormen bought wholesale quantities of it from Roberson.<br />

After the First World War, when Roberson's business was beginning a gradual<br />

decline, trading practices began to change <strong>and</strong> by the period following the<br />

Second World War, a reciprocal arrangement with the London colorman<br />

Rowney was made to divide trade into two spheres, Roberson supplying<br />

Rowney with canvas <strong>and</strong> Rowney supplying Roberson with paints (13).<br />

The catalogues indicate the cost of artists' materials <strong>and</strong> demonstrate prices<br />

were stable for much of the nineteenth century. There is a clear differentiation<br />

between luxurious pigments such as ultramarine or carmine <strong>and</strong> other, cheaper<br />

colors; the catalogues show that an ounce (28.35 g) of genuine ultramarine<br />

was twenty-eight times as expensive as its artificial substitute between around<br />

1840 <strong>and</strong> 1911, with prices almost completely stable in that period. The<br />

catalogues also reflect the introduction of new pigments <strong>and</strong> demonstrate that<br />

there could be a considerable time lag between the discovery of a pigment<br />

<strong>and</strong> its commercial application. Roberson & Co. appears to have been comprised<br />

of rather reactionary colormen, who contributed little to technical<br />

innovation, emphasizing the h<strong>and</strong>-prepared nature of the company's products<br />

until long after many of the other colormen had introduced a degree of<br />

mechanization; their reaction to new developments, therefore, may not be<br />

characteristic of the market as a whole.<br />

Three different prices were given in the catalogues until 1920: wholesale,<br />

retail, <strong>and</strong> professional. After 1920, discounts offered by artists' colormen to<br />

professional artists were abolished after complaints of abuse (14).<br />

In addition to catalogues, the archive contains a number of sample books of<br />

canvas, paper, <strong>and</strong> paints, both from Roberson <strong>and</strong> other companies, that are<br />

of great use in providing identified, untreated material fo r analysis. Projects<br />

have been carried out to establish the composition of the ground on the<br />

canvas samples <strong>and</strong> the pigments in the paint samples.<br />

Price books. recipe books. <strong>and</strong> notebooks<br />

Price books indicate both the cost to Roberson of a range of materials <strong>and</strong><br />

the resale price. Many of the cost prices are in an alphabetical code, illustrating<br />

the importance of secrecy to the company; an alphabetical code is also used<br />

in the recipe books.<br />

The recipe books in the archive contain a number of formulae for paints,<br />

media, <strong>and</strong> grounds, as would be expected from an artists' colorman, but also<br />

have recipes for trifle, lemon pickle, wine, boot polish, <strong>and</strong> blacking fo r harnesses.<br />

This illustrates the early connections between colormen, grocers, <strong>and</strong><br />

apothecaries, professions which were separating into distinct trades in the<br />

nineteenth century (15).<br />

Woodcock 31


Figure 2. Charles Roberson's recipe fo r Roberson's Medium, originally in sealed envelope. Fitzwilliam<br />

Museum, University of Cambridge.<br />

The books also demonstrate a degree of experimentation in the formulation<br />

of artists' colors; the composition of mixed colors in particular, such as olive<br />

green or neutral tint, was not absolutely fixed. Notes in the recipe books'<br />

margins indicate that formulae were not always successful, <strong>and</strong> in the later<br />

recipes a degree of substitution was effected (16).<br />

The recipe books show the importance given to secret recipes <strong>and</strong> trade<br />

secrets. Many of the recipes are written in code or shorth<strong>and</strong> or have certain<br />

portions in Greek. The fo rmula of Roberson's Medium, the company's most<br />

successful product, has been struck out in a recipe book (17). It is contained<br />

in a once-sealed envelope from Charles Roberson to his nephew Charles<br />

Park that is marked, "To be opened when I am dead," <strong>and</strong> dated 1868, ten<br />

years before Roberson's death (Fig. 2) (18). Roberson had reason fo r secrecy.<br />

When a Mr. Ellis advertised in the Daily Telegraph of 1900 that he would sell<br />

"the fo rmula fo r making this famous medium fo r imparting permanency to<br />

oil paintings" at a price of seventy guineas, Roberson quickly threatened<br />

prosecution. Ellis's solicitors responded the next day stating that he had "no<br />

intention ... of selling the medium as 'Roberson's Medium,' " had effected<br />

no sale of either the fo rmula or the medium, <strong>and</strong> would discontinue the<br />

advertisement (19).<br />

Personal accounts<br />

The personal account books are among the most important records in the<br />

archive. They are incomplete, but provide detailed information about the<br />

materials <strong>and</strong> techniques of many of the most prominent nineteenth- <strong>and</strong><br />

twentieth-century artists in Britain. They list in most cases what an artist<br />

bought <strong>and</strong> how much, <strong>and</strong> when he or she paid (Fig. 3) . They also demonstrate<br />

the range of services Roberson would perform, such as sending<br />

workers to artists' studios to carry out a variety of tasks, transporting paintings<br />

to <strong>and</strong> from exhibitions <strong>and</strong> loaning equipment to artists. William Holman<br />

Hunt's accounts show that he frequently changed his mind as to the proportions<br />

of his paintings, <strong>and</strong> that Roberson regularly sent workers to add strips<br />

of canvas <strong>and</strong> enlarge stretchers, particularly toward the end of the artist's life.<br />

Account holders were not only confined to artists living in Britain. Jozef<br />

Israels in the Hague, Louis Raemaekers in Brussels, Fritz Voellmy in Basel,<br />

<strong>and</strong> the American artist Wilton Lockwood all bought from Roberson.<br />

The accounts record extensive use oflay figures, usually rented by artists from<br />

Roberson, as their purchase price was very high (Fig. 4) There are a number<br />

32<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 3. Personal account book, 1881-1894, MS 250- 1993, 37-8. Fitzwilliam Museum, University<br />

of Cambridge.<br />

of both full- <strong>and</strong> child-size figures in the archive; their composition-stuffed<br />

knitted textile over a wooden <strong>and</strong> metal frame with papier-mache <strong>and</strong> gesso<br />

heads-helps to explain why they were so frequently sent fo r repair <strong>and</strong> why<br />

Roberson entered into disputes with artists over damage sustained in use (20) .<br />

Despite this, they were probably cheaper <strong>and</strong> more reliable than living models,<br />

of which Roberson kept a register to assist customers (21).<br />

The accounts also show the range of Roberson's customers from wealthy,<br />

well-established painters to professional graphic artists, scene painters, <strong>and</strong><br />

others. Many amateur artists also held accounts with Roberson, including<br />

several members of royalty, both in Engl<strong>and</strong> <strong>and</strong> abroad. Roberson appears<br />

to have been supplying the upper end of the market <strong>and</strong> although it did not<br />

hold the royal warrant fo r Queen Victoria, Roberson supplied the Queen<br />

with sketching materials during her reign <strong>and</strong> was granted the royal warrant<br />

to Queen Alex<strong>and</strong>ra in 1912 <strong>and</strong> to King Vittorio Emanuele III of Italy in<br />

1903. In 1949, Gerald Kelly remembered in a letter to Roberson, "When I<br />

first started painting I could not afford to deal with what I was then told was<br />

the best firm in London, but as soon as I got a little money I got my things<br />

from you, <strong>and</strong> gradually I have bought more <strong>and</strong> more ..." (22).<br />

Figure 4. Lay figure: Child no. 98. Fitzwilliam<br />

Museum, University of Cambridge.<br />

Some artists had their materials specially made to their own recipes or to a<br />

formula supplied in the past which they particularly liked. Leighton asked for<br />

his paints to be made "extra stiff," a preference alluded to in a letter from a<br />

fellow artist in 1888: "I have used your colours fo r several years, <strong>and</strong> I can<br />

with the greatest confidence say that they have given me every satisfaction.<br />

There is one quality which is greatly esteemed-<strong>and</strong> this Sir Frederick Leighton<br />

says is a great desideratum, that is stiffness" (23).<br />

Roberson, however, did not always think the customers' dem<strong>and</strong>s were advisable.<br />

A notebook entry dated 27 November 1882 records, "Sir F. Leighton<br />

P. R. A. in 1881, 82, had canvases prepared with different grounds made to<br />

his order & warranted not to st<strong>and</strong>; plaster of Paris &c. with Lac Varnish over<br />

&c. &c." (24). In June 1888, Millais is also recorded as having "canvases prepared<br />

thickly with one coat of turpentine color to his order, also prepared<br />

his own canvases. Not warranted to st<strong>and</strong>" (25).<br />

Woodcock 33


Bought ledgers<br />

The bought ledgers help to complete the picture of Roberson's trading practices.<br />

They date from 1854 to 1931 <strong>and</strong> indicate the limited extent of Roberson's<br />

manufacturing activities <strong>and</strong> its widespread trade with manufacturers<br />

<strong>and</strong> wholesale suppliers, both in Britain <strong>and</strong> abroad. From this, it is possible<br />

to reconstruct the chain of supply from raw material to artist. The ledgers<br />

also show how often Roberson & Co. printed its catalogues <strong>and</strong> where the<br />

company advertised; the ledgers even give details of the fixtures <strong>and</strong> fittings<br />

of their shops <strong>and</strong> workshops.<br />

One of the primary uses of the bought ledgers is to indicate the dating <strong>and</strong><br />

extent of use of a variety of pigments or pigment mixtures introduced<br />

throughout the period in which Roberson traded. The popularity of a new<br />

pigment <strong>and</strong> the decline of its predecessor can be gauged from Roberson's<br />

purchases from the wholesale supplier, providing a fu ndamental record of the<br />

dissemination of new colors <strong>and</strong> materials.<br />

Letters<br />

The letters preserved in the archive are, by their nature, a disparate group of<br />

materials, but provide detailed <strong>and</strong> wide-ranging information about the company.<br />

Roberson kept much of its more interesting correspondence, particularly<br />

letters expressing approval of its products or restoration treatments. From<br />

1919 to 1940 they also kept h<strong>and</strong>written (<strong>and</strong> later typed) copies of their<br />

answers to correspondence. Some letters were later used in their catalogues<br />

as testimonials; for example, the letter from W P. Frith of January 1897 referring<br />

to his painting in the Tate Gallery, London: " . .. you would like to<br />

know that it is just fifty years since I painted the Derby Day with your colours<br />

<strong>and</strong> medium, <strong>and</strong> that in so long a time there is no change whatsoever in<br />

the materials used" (26) .<br />

The letters also record advice both to <strong>and</strong> from artists concerning technique<br />

<strong>and</strong> the fo rmulation <strong>and</strong> application of materials. This can often be most<br />

informative, such as Thomas Gambier Parry's letter in 1874 from Ely, where<br />

he was working on the cathedral; the letter provides both a recipe <strong>and</strong> instructions<br />

for making up the "usual common encaustic."<br />

The letters also help to confirm information about developments within the<br />

company, such as a letter to the Factories Inspector announcing the installation<br />

of electric color grinding machinery in 1919, contradicting its catalogue<br />

note of around 1926 in which the company states that it is convinced that<br />

paints "which are ground by h<strong>and</strong> under the muller give results superior to<br />

those ground by machinery" <strong>and</strong> that the company, therefore, continued to<br />

use the former system (27). They also give some details of the conditions of<br />

the work force Roberson employed, mentioning industrial accidents <strong>and</strong> the<br />

possibility of redundancy due to ill health.<br />

The accounts books demonstrate how long customers were prepared to delay<br />

payment of their bills <strong>and</strong> many of the letters refer to attempts to recover<br />

bad debts, using an investigation agency if necessary. This demonstrates why<br />

Roberson asked fo r customers to provide references from reputable London<br />

businesses before opening an account. The bad debts were not always the<br />

result of fraud. Before the establishment of the welfare state in Britain, medical<br />

treatment could be ruinously expensive; therefore illness or death often left<br />

dependents at the mercy of creditors.<br />

Throughout the nineteenth century, artists <strong>and</strong> writers expressed unease about<br />

both the quality of artists' materials <strong>and</strong> the soundness of contemporary technique.<br />

This erupted into a public debate in 1880 when William Holman<br />

Hunt, a member of the Pre-Raphaelite Brotherhood <strong>and</strong> a long-st<strong>and</strong>ing<br />

customer of Roberson, gave a paper to the Royal Society of Arts that criticized<br />

the quality of artists' materials <strong>and</strong> accused the color men of either being<br />

deceived by their suppliers or of adulterating the products themselves. Hunt's<br />

34<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


paper resulted in a volume of correspondence <strong>and</strong> a leading article in the<br />

Times. The background to this can be traced in the Roberson Archive, as<br />

the anonymous colorman cited by Hunt as supplying him with adulterated<br />

paint was, in fact, Roberson.<br />

It is significant that Hunt did not transfer his account from Roberson to<br />

another colorman <strong>and</strong> continued to buy from the company until shortly<br />

before his death. Roberson took his accusations seriously <strong>and</strong> painted out<br />

samples of pigment <strong>and</strong> made up a chart of a number of aging tests they had<br />

carried out (Plate 5). The company's concern for the quality of its pigments<br />

is evident in the twentieth century when it used a chemical laboratory to<br />

analyze samples of pigment from the wholesale suppliers. Roberson also sent<br />

materials to artists fo r testing; <strong>and</strong> he corresponded with both Professor<br />

Church, professor of chemistry at the Royal Academy, <strong>and</strong> Helmut Ruhemann<br />

at the National Gallery in developing <strong>and</strong> assessing materials. A project<br />

to analyze pigments in the archive fo r adulterants has been carried out at the<br />

Hamilton Kerr Institute with mixed results: a small number of spectacularly<br />

adulterated pigments (with much use of fillers <strong>and</strong> extenders) were found,<br />

but there was less adulteration than the literature of the time <strong>and</strong> Hunt's<br />

accusations would suggest (28).<br />

Information from objects<br />

The archive contains a number of objects, most of which can be identified<br />

in the company's catalogues. These objects provide a useful set of references<br />

against which to check the documentary information. For example, analysis<br />

of the pigment collection has revealed the unreliability of many of the recipe<br />

books, with adulteration <strong>and</strong> substitution being common. The objects in the<br />

collection also illustrate the development in the nineteenth century of paint<br />

containers, including both a paint bladder <strong>and</strong> the ivory pins used fo r piercing<br />

it, a set of brass paint syringes, the forerunner of the collapsible tube, <strong>and</strong><br />

tubes adopted in the 1840s after they were first patented <strong>and</strong> advertised (Plate<br />

6) .<br />

Other activities of the company<br />

In addition to supplying artists' materials, Roberson was involved in a number<br />

of other activities that can be traced in various sources in the archive. These<br />

included restoration, dealing, <strong>and</strong> publishing. Trade directories show that it<br />

was not uncommon fo r color men to be dealers <strong>and</strong> restorers in the nineteenth<br />

century, but gradually the professors came to be listed separately, although<br />

Roberson's accounts indicate that the company continued all three<br />

pursuits until its records end in the 1940s.<br />

<strong>Historical</strong> context<br />

Many aspects of the history of the Victorian <strong>and</strong> Edwardian period can be<br />

gleaned from the Roberson Archive. The presence of an extensive empire is<br />

felt both in terms of the materials bought from the colonies <strong>and</strong> in the<br />

products supplied, which were designed to withst<strong>and</strong> extremes of climate not<br />

experienced in Britain. Stylistic movements were also reflected in Roberson's<br />

catalogues; materials for illumination, missal painting, <strong>and</strong> heraldic hatchments<br />

became popular during the Gothic revival. Roberson also responded to the<br />

Victorian expansion in public building <strong>and</strong> the popularity of painted interior<br />

decoration by fo rmulating a number of media <strong>and</strong> specially prepared canvases<br />

to imitate fresco. The company was involved in both supplying the materials<br />

<strong>and</strong> erecting the canvases in a number of public buildings in Britain.<br />

Roberson also supplied materials for the great explorations of the period<br />

since, even though photography was used by this time, artists were still often<br />

sent on expeditions to record the results. A tube of yellow ochre taken on<br />

the 1912 Shackleton South Pole expedition was given to Roberson by the<br />

expedition artist George Marston on his return. When the expedition was<br />

Woodcock 35


str<strong>and</strong>ed by pack ice that crushed their ship, Marston mixed his paints with<br />

lamp wick to caulk the seams of the open boat in order to successfully fe tch<br />

help. The Egyptologist Howard Carter was also a customer; he bought artists'<br />

materials, using Roberson as an agent, <strong>and</strong> sent cases of drawings of his finds<br />

to the company fo r distribution.<br />

The effect of both world wars is seen, to some extent, mainly as a series of<br />

problems of supply, raw materials being difficult to obtain <strong>and</strong> materials such<br />

as paper subject to government restrictions. Some pigments were also in short<br />

supply, vermilion increasing in price by 120 percent in 1940 because of shortages<br />

of mercury, which was used fo r shell <strong>and</strong> mine detonators, <strong>and</strong> because<br />

difficulties with chrome colors were being anticipated as they were used fo r<br />

dyes fo r khaki <strong>and</strong> in the manufacture of poison gas (29). Roberson was<br />

fortunate in sustaining only slight damage from an incendiary bomb in 1940<br />

<strong>and</strong> avoiding the more extensive damage experienced by other colormen.<br />

Post-war disruption is also evident; letters from Roberson in 1919 showed<br />

the difficulties of supply <strong>and</strong> delivery facing the colormen. Canvas was in<br />

short supply, materials usually coming from Russia were unobtainable, <strong>and</strong><br />

turpentine was scarce <strong>and</strong> expensive; even tin tubes in which to pack the<br />

colors were difficult to find (30). Similar difficulties were experienced during<br />

World War II. The world wars, the depression of the 1930s, <strong>and</strong> the rise of<br />

new reproduction methods contributed to Roberson's decline in the twentieth<br />

century.<br />

The archive as a research resource<br />

The Roberson Archive is chiefly relevant to the study of the materials <strong>and</strong><br />

techniques of British artists. Its relevance outside the United Kingdom is<br />

confined to the small number of foreign artists who bought from the company<br />

<strong>and</strong> the larger number of suppliers <strong>and</strong> retailers who traded with them.<br />

These contacts include firms in all five continents, but the most numerous<br />

contacts were in France, Germany, <strong>and</strong> the United States.<br />

The archive has been used in the last year to follow up a number of queries<br />

from both the United Kingdom <strong>and</strong> abroad. Although access will be restricted<br />

until the end of 1996 (when the cataloguing project ends), thereafter it will<br />

be available for study. A database of account holders, regarded as the most<br />

informative part of the archive, is being compiled, along with details of their<br />

purchases, <strong>and</strong> it is hoped that a checklist of account holders will be published.<br />

The database will make it possible to search for both particular artists <strong>and</strong><br />

specific materials <strong>and</strong>, when used in conjunction with information from recipe<br />

books <strong>and</strong> catalogues, should provide a clearer picture of what paintings<br />

were made of <strong>and</strong> what artists really used in the nineteenth <strong>and</strong> early twentieth<br />

centuries.<br />

Acknowledgments<br />

The Roberson Archive is published by permission of the Syndics of the Fitzwilliam<br />

Museum, University of Cambridge <strong>and</strong> is generously funded by the Leverhulme<br />

Trust.<br />

Notes<br />

1. Gallagher, M. 1991. The Roberson Archive: inventory oj written <strong>and</strong> printed material.<br />

Unpublished typescript. Cambridge: Hamilton Kerr Institute, 2. See also Kent's<br />

Original London Directory. 1817. Hamilton Kerr Institute (HKI) MS 785-1993,<br />

fol. 54v, 223. See also C. Roberson & Co. Ltd. 1969-1970 Catalogue, ii. Charles<br />

Roberson does not appear in trade directories at this address until Pigot's directory<br />

was published (Pigot & Co. 's Metropolitan new alphabetical directory. 1922-<br />

1923, 68). A letter to the National Gallery in London dated 1923 states that<br />

Charles Roberson founded the company in 1819 (HKI MS 862-1993, 293).<br />

2. Kent's Original London Directory. Op. cit. (note 1).<br />

3. HKI MS 204-1993, 638.<br />

36<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


4. HKI MS 865-1993, 497. The move is attributed to increasing congestion caused<br />

by the growth of Covent Garden Market in a note on the firm's history in a<br />

Roberson catalogue of 1969-1970, ii. A letter of 18 January 1940, however, states<br />

that they were not going to renew their lease on the Piccadilly branch "owing<br />

to conditions caused by the war," HKI MS 866-1993, 47.<br />

5. HKI MS 866-1993, 51-2.<br />

6. The first address is 26 Rue Chaptal, the premises of E. Mary & Fils, later Geo.<br />

Mary, who gave up business in 1920. The second address is 3 Quai Voltaire,<br />

where G. Sennelier & Fils are listed between 1920 <strong>and</strong> 1935.<br />

7. Gallagher, op. cit., 3.<br />

8. Walt, N. 1993. Personal communication. L. Cornelissen & Son, 105 Great Russell<br />

Street, London WC1B 3RY<br />

9. Cob be, A. 1994. Personal communication. A full account of the arrival of the<br />

archive at the Institute is recorded in Gallagher, op. cit., 3-4.<br />

10. The archive is listed in the Cambridge Union Catalogue of Departmental <strong>and</strong><br />

College Libraries, <strong>and</strong> was compiled by Paul Woudhuysen, Keeper of Manuscripts<br />

<strong>and</strong> Printed Books, Fitzwilliam Museum.<br />

11. McGrath, M. 1973. A catalogue of some eighteenth <strong>and</strong> nineteenth century patents<br />

dealing with artists' instruments. Unpublished typescript. London: Courtauld Institute<br />

of Art.<br />

12. Walt, N. 1994. Personal communication.<br />

13. Walt, N. 1993. Personal communication.<br />

14. HKI MS 865-1993, 191.<br />

15. This is demonstrated to even greater effect by H. Hodder, described as "Chemist<br />

& Druggist, Oil <strong>and</strong> Colorman, Tallow Ch<strong>and</strong>ler, Tea Dealer, <strong>and</strong> Tobacconist,"<br />

whose recipe book (1823-1858) covers fo od <strong>and</strong> drink, household cleaning<br />

agents, cures fo r both human <strong>and</strong> animal ailments, <strong>and</strong> recipes fo r varnishes, oils,<br />

<strong>and</strong> paints (Guildhall Library MS 16262, by permission of the Guildhall Library,<br />

Corporation of London) . The preface to a 1875 Post Office Directory clarifies<br />

the separation of the trades: "Groceries in early times consisted chiefly of spices<br />

. . . but subsequently came to comprise confectionery, dyes, drugs, chemicals,<br />

whale oil, &c.; the confectioner, the druggist, <strong>and</strong> the oilman have now branched<br />

off into separate <strong>and</strong> distinct trades" (The 1875 post office directory of the grocery<br />

<strong>and</strong> oil <strong>and</strong> color trades. 1875, iv) .<br />

16. Townsend, J. H., L. Carlyle, S. Woodcock, <strong>and</strong> N. Kh<strong>and</strong>ekar. n.d. Later nineteenth<br />

century pigments: analyses of reference samples, <strong>and</strong> evidence fo r adulterants<br />

<strong>and</strong> substitutions. The Conservator. Forthcoming.<br />

17. HKI MS 788-1993, fo1. 45v.<br />

18. HKI MS 892-1 993. Charles Roberson died <strong>and</strong> was succeeded by his nephew<br />

in 1878, according to the note on the firm's history in a 1969-1970 catalogue,<br />

11.<br />

19. HKI MS 891-1993.<br />

20. HKI MS 864-1993, 142.<br />

21. HKI MS 865-1993, 373.<br />

22. HKI MS 527-1993.<br />

23. HKI MS 500-1993.<br />

24. HKI MS. 785-1993, fo1. 37r.<br />

25. Ibid., fo1. 41r.<br />

26. HKI MS 597-1993.<br />

27. HKI MS 860-1993, 30.<br />

28. Townsend, J. H., et aI., op. cit. See also Woodcock, S., <strong>and</strong> N. Kh<strong>and</strong>ekar, n.d.<br />

The Adulteration of Twentieth Century Pigments in the Roberson Archive. Forthcommg.<br />

29. HKI MS 866-1993, 164, 219.<br />

30. HKI MS 860-1993, 6, 15, 22, 83, 85.<br />

Woodcock 37


Abstract<br />

Two texts from Simone de Monte<br />

Dante's medieval Italian manuscript<br />

are presented here in English translation.<br />

With their many recipes, these<br />

texts shed light on the technique<br />

<strong>and</strong> materials of fifteenth-century<br />

Italian manuscript illumination <strong>and</strong><br />

the historical development of the illuminator's<br />

craft. Studying the relationship<br />

of this manuscript to other<br />

treatises may provide a better underst<strong>and</strong>ing<br />

of the mechanisms involved<br />

in the adaptations <strong>and</strong> transmission<br />

of medieval art technology.<br />

Figure 1. Rome, Biblioteca Casanatense,<br />

MS 1793, cover.<br />

Figure 2. Rome, Biblioteca Casanatense,<br />

MS 1793, jol. 10v.<br />

38<br />

Libro Secondo de Diversi Colori e Sise da Mettere a Oro:<br />

A Fifteenth-Century Technical Treatise on Manuscript<br />

Illumination<br />

Arie Wallert<br />

The Getty Conservation Institute<br />

Museum Services Laboratory<br />

The J. Paul Getty Museum<br />

17985 Pacific Coast Highway<br />

Malibu, California 90265<br />

USA<br />

Introduction<br />

The Biblioteca Casanatense in Rome holds a very interesting fifteenth-century<br />

technical art manuscript. It has the shelf mark "MS 1793" <strong>and</strong> shows on<br />

folio 2r ("r" for recto) the author's name <strong>and</strong> the date of the manuscript:<br />

"Questo libro et ne di Simone de Monte Dante dela Zazera . .. 10 chopiai<br />

el mese di novembre 1422" (Fig. 1).<br />

The manuscript contains two closely related treatises on folios 10v-13v ("v"<br />

for verso) <strong>and</strong> 15v-20v. Written in different h<strong>and</strong>s, these texts contain recipes<br />

devoted to the practice of manuscript illumination. Various aspects of the art<br />

are described, such as ink production, parchment preparation, gold leaf application,<br />

<strong>and</strong> the preparation <strong>and</strong> application of pigments <strong>and</strong> various transparent<br />

organic colorants.<br />

The recipes in the Simone manuscript are not related to the well-known<br />

Mappae Clavicula tradition, the texts from Theophilus' Schedula de Diversarum<br />

artium, the Eraclius texts, or the texts of the Liber de Coloribus. What makes<br />

this manuscript so important is the fact that it appears to be related to a<br />

particular group of technical art treatises comprising a number of littleknown,<br />

mostly unpublished recipe texts that are specifically devoted to the<br />

miniaturist's craft.<br />

The relationships of parts in Simone's treatise with those manuscripts in the<br />

Siena Biblioteca Comunale, the Modena Biblioteca Estense, the Bologna Biblioteca<br />

Universitaria, the Oxford Bodleian Library, <strong>and</strong> the Florence Biblioteca<br />

Nazionale are quite intriguing. The recipes of both treatises in Simone's<br />

manuscript are given here in translation, with notes indicating the relationships<br />

to other recipe texts, as well as references to relevant technical literature<br />

(1) .<br />

Relationships to recipes in the well-known Bolognese manuscript are referred<br />

to according to the numbering <strong>and</strong> pagination of Merrifield's edition (2).<br />

Relationships with two treatises in the Siena Biblioteca Comunale follow the<br />

numbering in Wallert's edition (3). To facilitate the underst<strong>and</strong>ing of these<br />

relationships, Simone's recipes-originally not numbered-are given numbers<br />

in brackets in this translation.<br />

The Simone manuscript is clearly not written by a person with a firsth<strong>and</strong><br />

knowledge about the products <strong>and</strong> techniques he describes. Occasionally,<br />

Simone prescribes the wrong ingredients, gives only part of a recipe, or<br />

gives some recipes in which the final outcome clearly will not be the result<br />

promised by the heading; obviously, he was not a craftsman. Sometimes his<br />

descriptions are too succinct to be fully understood by the uninitiated or<br />

modern reader. In those cases I have added to the Simone text, placing these<br />

additions in brackets to distinguish them from the original text <strong>and</strong> to make<br />

the recipes more intelligible.<br />

With folio 10v, Simone begins a book on various colors <strong>and</strong> grounds to lay<br />

gold (Fig. 2). Following is the text of folios 10v-13v (page changes in the<br />

original will be noted here with the symbol {}):<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


[1] How to make a gold ground. Take fine, slaked gypsum, as much as the size<br />

of a nut, <strong>and</strong> grind it with a little bit of clear water. Then take Armenian<br />

bole, as much as a chickpea or a bean, <strong>and</strong> grind this separately, <strong>and</strong> also with<br />

water. Having done this, mix them together, take some dissolved glue <strong>and</strong><br />

add it. Next have a bit of c<strong>and</strong>y or sugar <strong>and</strong> some earwax <strong>and</strong> grind it gently<br />

together. The glue must be so strong that the whole matter sticks a little bit<br />

to the porphyry stone when you grind it. Then you keep it warm above<br />

some hot coals. When you work with it you can add some glue so that it<br />

will be a bit fluid. Let the dissolved glue st<strong>and</strong> a few days so that it comes<br />

from the fire with better strength. Now you can work with it. When the<br />

work is dry, scrape it until it is smooth. Then wet it with clear water <strong>and</strong> lay<br />

the gold on it. Press on it when it is dry. If the gold ground would be too<br />

weak for you, distemper it: you can put a bit of glair in the water (4).<br />

[2] To make lines of burnished gold. Take slaked gypsum, grind it with glair <strong>and</strong><br />

a bit of honey, some earwax, <strong>and</strong> a bit of glue. Let it dry [when applied], then<br />

huff on it <strong>and</strong> lay the gold. Burnish it (5).<br />

[3] To make lines of gold in another way. Take [gum] ammoniac <strong>and</strong> grind it<br />

without water or anything else. Then take garlic juice <strong>and</strong> grind it with it.<br />

Next add a little bit of Armenian bole to it. Work with it. When this is done;<br />

huff on it <strong>and</strong> lay the gold on it (6).<br />

[4] To make the blue clothlets. Take a plant called turnsole, of which you pick<br />

one-by-one the flowers <strong>and</strong> clustering {llr} berries. Then you put these in<br />

a mortar, according to the quantity you want to make. Crush them <strong>and</strong> press<br />

the juice out of the aforesaid berries. You put this juice in a glazed bowl <strong>and</strong><br />

then you take pieces of cloth that are white <strong>and</strong> coarse. Then you put these<br />

into the aforementioned bowl [with the juice] <strong>and</strong> let them soak well in it.<br />

Do this three times, <strong>and</strong> every time that you take it out of the mentioned<br />

bowl, you let it dry well on a wooden bench. Before you reach this stage,<br />

make sure you find a remote quiet hole in the ground, <strong>and</strong> urinate much in<br />

it, six days before you come to make the said turnsole. Take straw <strong>and</strong> make<br />

a screen over the urine <strong>and</strong> place the cloths above the humidity of this urine.<br />

Thus it will become darker <strong>and</strong> more beautiful, <strong>and</strong> let it st<strong>and</strong> there for at<br />

least twenty days above the aforesaid urine <strong>and</strong> then the turnsole will have<br />

become beautiful.<br />

You prepare it as follows: Take glair according to the quantity you think this<br />

turnsole needs to have to become deeply colored. When you want to work<br />

with it, take this glair in a shell or some glazed small pot <strong>and</strong> then take the<br />

turns ole <strong>and</strong> put some of it in according to the right amount <strong>and</strong> let it st<strong>and</strong><br />

[in the glair] for an hour <strong>and</strong> then press it out with your finger <strong>and</strong> the juice<br />

that comes out of the clothlet is the turnsole. With this you can work, using<br />

a little bit at a time (7).<br />

[5] To make a fine ink. If you want to make a fine ink, take one pound of<br />

galls. Break it into smaller pieces <strong>and</strong> put it in ten pounds of rainwater or<br />

water from the pot. Let it boil in a pot that has not previously been used,<br />

until the volume of this water is reduced to half. Next sieve it <strong>and</strong> throw out<br />

the substance of the galls <strong>and</strong> put the water back in after you have washed<br />

the jar. {11 v} Having this done, put in five pounds of the softest white wine<br />

you can find, i.e. replacing the amount of water that has evaporated. Let it<br />

cook <strong>and</strong> when it starts to boil, put in half a pound of perfectly pulverized<br />

gum Arabic, adding this bit by bit while stirring with an iron cane until the<br />

gum is dissolved. When that is done, take the aforesaid vase from the fire <strong>and</strong><br />

then put in half a pound of perfectly pulverized Roman vitriol. Then cover<br />

it, so that no dust can get into it <strong>and</strong> let it st<strong>and</strong> for three or fo ur days on a<br />

quiet place, because when you leave it alone it becomes more beautiful <strong>and</strong><br />

blacker. Then without forcing it, sieve it through a cloth into a glazed pot.<br />

And if you want this to be right, you sieve it after 15 days three times. Know<br />

that the vitriol loses its quality through smoke, that is why one does not heat<br />

it on the fire. If you cannot find white wine, you may take red wine, which<br />

Wallert 39


should be fresh <strong>and</strong> clear <strong>and</strong> of little color. Know that the quality of the<br />

right gum can be recognized in that it is difficult to break. The gum that<br />

comes in small grains is much better. The good quality Roman vitriol can<br />

be recognized by its blue color <strong>and</strong> [its texture], in that it is hard <strong>and</strong> grainy,<br />

like coarse salt. It is sufficiently good when the rain water or river water is<br />

soft, but better is that which comes from the fo untain. Good galls are recognized<br />

by their being small <strong>and</strong> wrinkled, <strong>and</strong> hard inside (8).<br />

[6] To make a good yellow. Take an egg <strong>and</strong> take out the white of it <strong>and</strong> put<br />

it in a glazed pot. Put in as much saffron as seems right to you. Take a wooden<br />

mixer <strong>and</strong> beat it so much that the egg white breaks <strong>and</strong> becomes nicely<br />

yellow. Then take a fine clothlet <strong>and</strong> sieve [the glair mixture] through it in a<br />

glazed dish <strong>and</strong> leave it {12r} in the sun to dry. Then take it out, keep it dry,<br />

<strong>and</strong> when you want to work with it, grind it with a little bit of water <strong>and</strong> a<br />

few drops of glair (9).<br />

[7] To make a fine lac. Take a man's urine, as much as you need, put it in a<br />

bowl <strong>and</strong> let it st<strong>and</strong> for eight days. Then put it in a pot <strong>and</strong> let it boil so<br />

long that there is no foam anymore. Then percolate it through strong ashes,<br />

so that the liquid that comes through is like a lye. When you have sieved it;<br />

take coarse gum lac <strong>and</strong> crush it until gets the appearance of bread crumbs,<br />

put it in a new pot, <strong>and</strong> put in the said ... (10).<br />

[8] To make indigo. Take flower of woad <strong>and</strong> flour of grain. Make a dough<br />

of it with urine <strong>and</strong> with vinegar. Make a cake of it <strong>and</strong> dry it in the sun.<br />

And if it is too light, take more flower of woad <strong>and</strong> mix it again until it takes<br />

on the color you want (11).<br />

[9] How to make vermilion. Take one part of mercury <strong>and</strong> one of white sulfur,<br />

as much of one as of the other. Put it in a glass bottle, thoroughly clad with<br />

clay. Put it on a moderate fire <strong>and</strong> cover the mouth of the bottle with a tile.<br />

Close it when you see yellow smoke coming out of the bottle, until you see<br />

the red <strong>and</strong> almost vermilion-colored smoke. Then take it from the fire <strong>and</strong><br />

the vermilion will be ready (12).<br />

[10] To make burnished gold lines. Take slaked gypsum, ground with glair <strong>and</strong><br />

a bit of honey, <strong>and</strong> some earwax subtly ground together with a little bit of<br />

bone glue. Let it dry when you have applied it. Then huff on it, lay on the<br />

gold, <strong>and</strong> polish it (13).<br />

[11] A way to test whether the ultramarine blue is good. Take an iron knife <strong>and</strong><br />

hold it in the fire. Then take a little bit of the blue <strong>and</strong> put it on this<br />

knife, <strong>and</strong> if it is good {12v} it will become more beautiful <strong>and</strong> of a tender<br />

color. If it is not good, it will become black like ink (14).<br />

[12] To make blue. Take a glass pot <strong>and</strong> put in it fresh sulfur <strong>and</strong> mercury,<br />

thoroughly mixed. And take two parts of sulfur to three parts of mercury.<br />

Clad the pot with white potters clay <strong>and</strong> make sure that this is thoroughly<br />

mixed with horse dung. Put this to heat well on the fire, until the smoke<br />

comes out of it. Then take it off the fire <strong>and</strong> it will be a perfect blue (15).<br />

[13] To make a blue water. Take from the elder tree the berries with the<br />

kernels, at the time when they are in the stage between green <strong>and</strong> ripe. Cook<br />

them in a kettle <strong>and</strong> when they are boiled enough, take a cloth <strong>and</strong> squeeze<br />

the juice that comes out of them in a glass or glazed pot. Then temper it<br />

with urine <strong>and</strong> use it as a painter's color. When it is dry you can lightly paint<br />

with the juice oflilies with a brush on it, <strong>and</strong> it will be a blue color. Similarly,<br />

you can dye cloths with it or paper to write on with letters in gold.<br />

[14] How to make red lead. Heat litharge together with rasped lead in the fire.<br />

This makes the red lead.<br />

[15] How to make the dark "verzino. " If you wish to make the dark verzino,<br />

take about a "quarto" of rasped brazilwood. Put it in a beaker. Take white of<br />

egg, thoroughly beaten fine with a sponge <strong>and</strong> put it fo r eight days in a bottle<br />

with some realgar. Mix it together after those days. Then sieve this verzino<br />

40<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


into a glass, <strong>and</strong> put in a bit of well ground alum, of about the size of a<br />

chickpea. Then take a piece of white linen <strong>and</strong> sieve the verzino <strong>and</strong> put it<br />

{ 13r} in a glazed bowl <strong>and</strong> put it to dry in a place where there is no sun,<br />

until it is dry. With the verzino you can make red. When you work with it,<br />

put a little bit of it in the shell <strong>and</strong> add a little bit of clear water (16).<br />

[16] How to make green. Take arzica <strong>and</strong> grind it <strong>and</strong> then you add pale azurite<br />

blue, so much that it takes the color you want. Know that the greens that are<br />

made with arzica need only a little bit of medium.<br />

[17] To make "p urpurina. " Take similar amounts of mercury <strong>and</strong> tin <strong>and</strong> have<br />

them amalgamated together <strong>and</strong> when they thus are thoroughly combined,<br />

let it cool down <strong>and</strong> grind it together. But first take "caschina" [?], which is<br />

separately ground very well (17). Then take an oven with red hot coals <strong>and</strong><br />

put [the ingredients] in a bottle or pot on the lit coals, <strong>and</strong> let it st<strong>and</strong> until<br />

smoke comes off, <strong>and</strong> make sure that mouth of the bottle stays open.<br />

[18] To take letters off from the parchment. Take one pound of saltpeter, one<br />

pound of vitriol, <strong>and</strong> distill it. Then take a sponge, dip it into this water, rub<br />

it on the parchment <strong>and</strong> the letters will disappear (18).<br />

[19] To make "origno" (19). Take lead <strong>and</strong> sulfur <strong>and</strong> nitric salt <strong>and</strong> mercury<br />

in the same quantities together <strong>and</strong> put it in a bottle of which the mouth is<br />

wide (20, 21). Put it on a moderate fire <strong>and</strong> when it starts to look glittery<br />

like gold on the mouth of the bottle, take it off the fire <strong>and</strong> it will be done<br />

(22).<br />

[20] To make red. Take very well ground eggshells, treat them on the stone<br />

<strong>and</strong> put them in a new, hollow brick. And take verzino <strong>and</strong> boil it with a<br />

little bit of roche alum <strong>and</strong> pour it on the eggshells. Mix it <strong>and</strong> of this you<br />

add until is becomes as light as you wish (23). {13v}<br />

[21] To make the yellow "arzicha. " Take very well ground eggshells <strong>and</strong> put<br />

them in the hollow of a new brick. Take the weld herb of the textile dyers<br />

<strong>and</strong> let it boil in water with a little bit of alum. Pour it on the eggshells <strong>and</strong><br />

thus make it as light or as dark as you wish (24).<br />

[22] To make green. Take arzicha <strong>and</strong> azurite blue <strong>and</strong> mix it together <strong>and</strong> if<br />

you want to make a darker green, you must add more blue. Distemper it<br />

with gum water.<br />

[23] To make shades on the arzicha. Take saffron, tempered with white of egg.<br />

[24] To make gold ground. Take lead white <strong>and</strong> Armenian bole, some earwax<br />

<strong>and</strong> saffron. Temper it with white of egg a little bit of saffron <strong>and</strong> a bit of<br />

"b<strong>and</strong>oli"(?). Grind it with water.<br />

[25] To make letters in gold with the pen. Take lead white, <strong>and</strong> make eight parts<br />

of it, <strong>and</strong> take an amount of honey of one of these eight parts <strong>and</strong> grind it<br />

with white of egg <strong>and</strong> a pinch of saffron.<br />

[26] To lay bole on gesso. You want to mix it with glair <strong>and</strong> then grind it with<br />

clear water.<br />

[27] To make gold letters. Take a bit of bole <strong>and</strong> some burned chalk, about<br />

half the amount of the bole <strong>and</strong> take three or four times as much slaked<br />

gypsum <strong>and</strong> grind it with glair.<br />

[28] To make lines in gold. Take slaked gypsum, ground with glair in a bit of<br />

honey, a little bit of earwax, <strong>and</strong> a bit of glue. Let it dry. Then huff on it, lay<br />

the gold leaf on it <strong>and</strong> burnish it (25).<br />

[29] To make gold you can apply with the pen. If you want to prepare gold that<br />

you can work with the pen, take beaten gold leaves-Praise God.<br />

End. {15v}<br />

The following text is from folios 15v-20v, one page of which is shown in<br />

Figure 3:<br />

Wallert 41


[30] To make gum water for paints. Take of the white gum Arabic, the quantity<br />

you need <strong>and</strong> put it in a beaker. Next take clear water <strong>and</strong> put it in this<br />

beaker, so much that it covers the gum about two fingers wide. Place it in<br />

the sun. If there is no sun, put it near the fire <strong>and</strong> let it st<strong>and</strong> there for three<br />

or fo ur hours, until the gum dissolves. Then take it from the fire, <strong>and</strong> let it<br />

st<strong>and</strong> for two days, but {16r} make sure that you stir it every fo ur hours.<br />

Next sieve it through a cloth <strong>and</strong> cover it until you will work with it. This<br />

IS proven.<br />

[31] To make egg medium to temper the colors. Take as much white of egg as<br />

you want <strong>and</strong> put it in a very clean pot. Then take fo ur sponges bound<br />

together at the top <strong>and</strong> knead with both h<strong>and</strong>s so that the white of eggs<br />

completely turns into such a froth that it will not even fall out if you turn<br />

the pot upside down. Let it st<strong>and</strong> from the morning until the evening until<br />

it settles <strong>and</strong> then put it in a bottle. And put some camphor in this bottle so<br />

that, not even after a year, will it go bad.<br />

Figure 3. Rome, Biblioteca Casanatense,<br />

MS 1793, fol. 17r.<br />

[32] To make a watercolor like gold. Take this previously described glair <strong>and</strong> put<br />

a little bit of saffron in it; <strong>and</strong> if you put in more yellow it will all be like<br />

gold.<br />

[33] To grind <strong>and</strong> clean vermilion for writing <strong>and</strong> miniature painting. Take as much<br />

vermilion as you think you need, grind it very fine with water on the porphyry<br />

stone. Then let it dry on the stone. Then grind it very fine fo r the<br />

second time. Next put it in a glass horn <strong>and</strong> temper it with the aforesaid<br />

glair <strong>and</strong> let it st<strong>and</strong> fo r three or fo ur hours until the vermilion has settled<br />

on the bottom. Then carefully decant the supernatant liquid <strong>and</strong> replace it<br />

fo r new glair. And keep on doing this for two or three times. Thereafter write<br />

with this on your paper. When you think it is not shiny enough, make the<br />

letters you want, large or small, let them dry <strong>and</strong> then lightly lay with the<br />

brush some of the previously described yellow on it <strong>and</strong> it will be very shiny<br />

(26).<br />

[34] To make green for writing <strong>and</strong> miniature painting. Take purple lilies <strong>and</strong> pick<br />

only the petals of these. Pound them <strong>and</strong> take of the juice. You should know<br />

that this juice is of a purple color. When you add a bit of roche alum, that<br />

is, to a beaker of juice about a chestnut of well pulverized alum, it will turn<br />

blue. Then take pieces of linen cloth <strong>and</strong> put them to soak very well in this<br />

juice. Then lay it [the cloth] on the grid [to dry], <strong>and</strong> then it must be soaked<br />

another time so that it will better take on the color. And in between you let<br />

it dry on the grid.<br />

[35] To make blue clothlet colors. Take certain flowers, that grow between the<br />

corn, which have five petals <strong>and</strong> have the color of a purple lily. Take the<br />

petals <strong>and</strong> rub these leaves in a piece of old linen. Let it dry <strong>and</strong> if you feel<br />

that the linen does not take enough color, repeat the treatment fo r a second<br />

time. You can also take the flowers that are called "fior di lebio" or "ibio,"<br />

which also grow between the corn. Pick, when you begin, the flowers that<br />

have a sky blue color <strong>and</strong> have lots of petals. Pick the petals of these flowers<br />

<strong>and</strong> squeeze them bit-by-bit with your fingers. Next press them well together,<br />

take out the juice <strong>and</strong> then soak the linen in it <strong>and</strong> let it dry on the grid.<br />

{17v} Also: take the petals of the "fior di lino" <strong>and</strong> crush them <strong>and</strong> put the<br />

pieces of linen in this juice (27). Let them dry in the same way as described<br />

before <strong>and</strong> if the pieces of linen would not take up enough color the first<br />

time, dip them in again (28).<br />

[36] To make red clothlets fo r writing <strong>and</strong> miniature painting. Take the leaves of<br />

"bietoli," also called "gelosia," <strong>and</strong> let them dry a bit <strong>and</strong> then rub these leaves<br />

in a piece of linen cloth <strong>and</strong> if it does not have enough color for you, rub<br />

in more until you think it is right (29, 30). You should know that the leaves<br />

need to be dried because they have a certain humidity of water in them.<br />

Therefore, one lets them dry, which consumes this humidity, so that only the<br />

color remains.<br />

42<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


[37] To make crude verzino for writing <strong>and</strong> miniature painting. Take the brazilwood<br />

<strong>and</strong> finely scrape of the quantity you think you need, with a piece of glass.<br />

Put these scrapings in a new glass. Add so much vinegar to it, that it covers<br />

the brazilwood half a fmger's width, <strong>and</strong> not more. The vinegar must be<br />

strong <strong>and</strong> clear. Add to the substance in this glass the quantity of about a<br />

bean of gum Arabic <strong>and</strong> the same amount of well-pulverized roche alum.<br />

Place it in the sun <strong>and</strong> make sure that the glass is well covered. Keep it this<br />

way for three or four days. If there is no sun, then keep it for four days so<br />

close to the fire that it gets lukewarm. When it is warmed this way, put it in<br />

shells as shown here, <strong>and</strong> let it st<strong>and</strong> in these shells, until it seems done to<br />

you. Know that the longer it remains, the better it is. Next put it in a bottle.<br />

{17v} Write with this on your work. If you want to make red clothlet colors<br />

of it, dip in to soak the pieces of cloth <strong>and</strong> let them dry in the previously<br />

described way, except that in the making of the clothIets [neither] the alum<br />

nor the gum Arabic should be used.<br />

[38] To make a purple clothlet for writing <strong>and</strong> miniature painting. Take the petals<br />

of the "papatello," which is a sort of wild poppy. Crush them <strong>and</strong> take out<br />

the juice. To one glass of juice it needs to have about the size of a bean of<br />

well-pulverized roche alum. Then dip in the piece of cloth, once or twice,<br />

until you think that it has taken enough of the color.<br />

[39] To make a beautiful blue at little cost. Take quicklime <strong>and</strong> green <strong>and</strong> ground<br />

verdigris <strong>and</strong> sal armoniac, as much of one as of the other. Grind these all<br />

together with urine <strong>and</strong> you will see a beautiful blue. Temper it with the<br />

previously described glair, when you want to work with it (31).<br />

[40] To make invisible letters, which you cannot see unless near the fire. Take sal<br />

armoniac, the amount of it is about the size of a chestnut to half a glass of<br />

water. Write with this water when it is dissolved, <strong>and</strong> if you want to read it<br />

you heat the parchment near the fire <strong>and</strong> the letters will appear as if they<br />

were written in ink.<br />

[41] To make black letters with every water you write. Take galls <strong>and</strong> vitriol, as<br />

much of one as of the other. Grind it <strong>and</strong> sieve it <strong>and</strong> put it on the parchment<br />

like a varnish coating. Next write on this with any water you want <strong>and</strong> while<br />

you are writing the letters will appear black.<br />

[42] To erase letters from goatskin parchment without a scraping iron. Take the juice<br />

of a lemon or of a strong orange <strong>and</strong> immerse {18r} a sponge in it. Rub<br />

with this sponge on the letters <strong>and</strong> they will be taken off as if they never had<br />

been written.<br />

[43] To make a ground fo r gold. Take slaked gypsum, as much as you need, <strong>and</strong><br />

Armenian bole about one third of the gypsum, <strong>and</strong> of aloe one third of the<br />

bole. Then grind everything together very fine with water on the porphyry<br />

stone. Let it dry on the porphyry stone <strong>and</strong> then grind it again very fine.<br />

Add a little bit of c<strong>and</strong>y sugar <strong>and</strong> when it is ground enough, gather it in a<br />

glazed horn <strong>and</strong> temper it with equal quantities of the aforesaid glair <strong>and</strong><br />

water of animal glue, that is hide glue. Make this gold ground so liquid that<br />

it flows from the pen <strong>and</strong> write whatever you want. And if it appears not to<br />

have enough volume, write over it a second time, like before. Scrape it carefully<br />

with a scraping iron when it is dry, so that the letter will be plane <strong>and</strong><br />

smooth. When you want to lay the gold, huff on the letter <strong>and</strong> immediately<br />

lay the leaf of gold or silver. Press on it with the calf's tooth. Then clean it<br />

with some cotton wool so that what sticks out from the letter will be wiped<br />

off. If you want to make a head of a frame with this gold ground with the<br />

brush like described, paint it over the first application <strong>and</strong> then scrape it <strong>and</strong><br />

lay the gold.<br />

[44] To make French glue for many purposes. Take animal glue, i.e. bone glue,<br />

<strong>and</strong> put it to soak in so much water that it covers the glue. Put it to soak in<br />

the evening <strong>and</strong> put it on the fire in the morning. When you soak it, it does<br />

not matter {18v} if there is not enough water to cover it completely, because<br />

Wallert 43


when the lather covers it about the width of a finger it is good enough.<br />

When you have had it cooking fo r a while, take it from the fire <strong>and</strong> immediately<br />

add four or five hard <strong>and</strong> fresh pieces of chicken dung. Then you<br />

sieve it through a straining cloth, without wringing it. Then put as much as<br />

you need of it back in the water <strong>and</strong> use it as it pleases you.<br />

[45] To color <strong>and</strong> lay gold on oranges <strong>and</strong> other fruits. Take previously described<br />

glue <strong>and</strong> douse your h<strong>and</strong>s in it. Rub with these washed h<strong>and</strong>s the orange<br />

or any another apple very well. Let it dry. Then wet your h<strong>and</strong>s again <strong>and</strong><br />

rub this fruit again until it has taken on the glue well. Then give it the color<br />

you want it to take. But if you wish to be it in gold, be it as letters or in<br />

gold leaves, do as follows: Take armoniac, very well ground with urine on<br />

the porphyry stone. Distemper it with urine, but not so much that it flows<br />

from the pen, because the letters or the leaves want to be done with the<br />

brush over the colors with this armoniac. Let it dry. Next take the gold leaf<br />

<strong>and</strong> put it on the letters. But you should know that the letters first need to<br />

be warmed up with the knife. Then, when the gold is laid, immediately press<br />

lightly on it with cottonwool. Then rub with the cottonwool over the letters<br />

so that the gold that does not adhere to the letters is taken off. You can also<br />

give it beautiful, nice colors.<br />

[46] To make a tragacanth gum paste oj any color you want, to make reliifs. Take<br />

as much tragacanth gum as you want <strong>and</strong> put it in a glass or a dish, <strong>and</strong> add<br />

so much vinegar that it covers the tragacanth. But the vinegar you put in<br />

should be strong. Let it st<strong>and</strong> to soak fo r two days <strong>and</strong> two nights. Then take<br />

slaked gypsum which is dry. Next take lead white <strong>and</strong> triturate it very well<br />

with the tragacanth in a mortar. But put on the bottom of the mortar {19r}<br />

a bit of the gypsum so that the lead white <strong>and</strong> the tragacanth will not stick<br />

to the mortar. To prevent it from sticking to the pestle, put also some gypsum<br />

on that. Pound it well <strong>and</strong> it will become a paste as white as cottonwool. If<br />

you want the paste to become black, do as follows: Take a lit c<strong>and</strong>le <strong>and</strong> over<br />

the flame you hold a piece of tinned iron so close that it touches the flame,<br />

whose smoke will deposit in a black spot. Take this deposit <strong>and</strong> rub it together<br />

with the paste. And the more you put into it the better it will be. Grind it<br />

with indigo if you want it to be blue, <strong>and</strong> if you put in little, it will be light<br />

blue, <strong>and</strong> it will be dark blue if you put much more indigo in it. If you want<br />

it to be red, put in vermilion which has been finely ground on the porphyry<br />

stone. The more you put in, the redder it will be. If you add little of it, it<br />

will be of a flesh tone. Put in a bit of well pulverized verdigris if you want<br />

.it to be green. If you want it to be yellow, put in some well ground orpiment.<br />

If you want it to be a purple color, put in a bit of lac. Know that the more<br />

of these colors you put in, the stronger colored it will be. The paste you thus<br />

make should be a little bit harder <strong>and</strong> you make it in such a way that you<br />

grind it again many times with the previously mentioned things. To make a<br />

good paste <strong>and</strong> to prevent it from sticking, either to the mortar, or to the<br />

pestle, knead a piece of it in your h<strong>and</strong>. When you want make it in a mold,<br />

take some of the soft paste that fits on the blade of a knife, or even less. But<br />

if you want to let it harden in a lead or clay mold, clad the form with some<br />

cotton so that the paste will not get stuck to the mold. Apply some warm<br />

bone glue in the areas where you want to do some modeling or make foliage.<br />

That will do it.<br />

[47] Papier mache to make reliifs. Take scraps of the paper cuttings from books<br />

<strong>and</strong> put them to boil in a kettle. When it has boiled, take the scraps out <strong>and</strong><br />

pound them fine in a bronze {19v} mortar. Then let it boil a second time.<br />

Take the kettle from the fire once it has boiled <strong>and</strong> let it cool down. When<br />

it is cold enough to have your h<strong>and</strong>s in it, take this paper pulp out. Squeeze<br />

the water out, make balls of it <strong>and</strong> let them dry. When you want to make<br />

reliefs with it, put these balls to soak in hot water during one or two days.<br />

When the water is cooled down enough to have your h<strong>and</strong>s in it, break the<br />

balls <strong>and</strong> stir. It will become like a dough. This paste spreads smoothly. Apply<br />

44<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


it on what you are working, with a scuilie, so that it spreads. Press everywhere<br />

on it with a sponge, as you work with the paper pulp without glue.<br />

[48] To make cork board fo r making reliif. Take good writing paper, which is<br />

smooth <strong>and</strong> with a little glue. Place them between washcloths. Always fold<br />

the cloths in four quarters, placing the paper in between. Put it in boiling<br />

water, making sure that it does not become brownish or stained by lye. Then<br />

take it out <strong>and</strong> let it dry on the grid. When it is a bit dry, you can work<br />

with it. It is best to work with water, so warm that you can just st<strong>and</strong> to have<br />

your h<strong>and</strong> in it.<br />

[49] To make letters, big or small, in gold. Take the milky juice of a fig tree<br />

<strong>and</strong> put as much as you think you need in a dish. Next take good black ink,<br />

which is rich in gum, <strong>and</strong> put so much of it in that the juice looks black.<br />

Write with it. When the letters are written, you huff on it <strong>and</strong> immediately<br />

lay the gold on it. Burnish with the tooth, <strong>and</strong> then rub the letters with<br />

cottonwool <strong>and</strong> where you have huffed on the ground, the gold will stick<br />

<strong>and</strong> the rest will be taken off (32).<br />

[50] Another way. Take Armenian bole <strong>and</strong> grind it with clear water very fine<br />

on the porphyry stone. Let it dry on the stone. Then grind it again with<br />

water <strong>and</strong> let it again dry on the stone. Then grind it again with the aforesaid<br />

milky juice. When ground, {20r} gather what you have in a horn of glass,<br />

or of glazed earthenware. Temper this so much that it flows from the pen.<br />

[51] To make secret writings. Take well pounded <strong>and</strong> sieved galls <strong>and</strong> put it to<br />

soak in water. Let it st<strong>and</strong> in the water until it settles to the bottom. Four<br />

glasses of this powder go to one glass of water. Write with this water whatever<br />

you want. Take, when you want to read it, vitriol <strong>and</strong> put it in water, with<br />

which you wet these letters. The letters, which you will be able to read, will<br />

now appear.<br />

[52] To make the blue clothlet. Take a plant which is called turnsole whose<br />

flowers <strong>and</strong> clustering berries you pick one-by-one. Then you put these in<br />

a mortar, according to the quantity you want to make of it. Crush them <strong>and</strong><br />

take the juice out. Put this juice in a glazed bowl <strong>and</strong> then you take pieces<br />

of cloth that are white <strong>and</strong> coarse. Then you immerse them in the bowl [with<br />

the juice] <strong>and</strong> let them soak well in it.<br />

Do this three times, <strong>and</strong> every time that you that you take it out of this bowl,<br />

you let it dry well on a wooden bench. Before you reach this stage, you have<br />

to find a remote quiet place in the ground, <strong>and</strong> urinate much in it, six days<br />

before you come to make the said turnsole. Then, to place it above the<br />

humidity of this urine, take straw <strong>and</strong> make a fabric over the urine, <strong>and</strong> place<br />

the cloths on it. Let it st<strong>and</strong> fo r at least twenty days above the aforesaid urine<br />

to make it darker <strong>and</strong> more beautiful.<br />

The best turnsole is made in this way: First, take glair according to the quantity<br />

you want to make. You may wish to use amounts small enough that the<br />

said turns ole will be {20v} strongly colored. When you want to work with<br />

it, take the glair <strong>and</strong> put it in a shell or in a small glazed jar. Then take the<br />

turnsole <strong>and</strong> put some of it in, according to the right proportion <strong>and</strong> let it<br />

st<strong>and</strong> [in the glair] fo r an hour. Then press it out with your finger <strong>and</strong> the<br />

juice that comes out of the clothlet is the turnsole. Always prepare a little bit<br />

at a time because that is better.<br />

Notes<br />

1. Constraints of space in the present volume made it impossible to publish a true<br />

bilingual text edition (with a transcript of the original printed next to the translation).<br />

A copy of my transcript, however, is available upon request.<br />

2. Merrifield, M. P. 1849. Original Treatises, Dating from the XIIth to the XVIIlth<br />

Centuries on the Arts of <strong>Painting</strong> in Oil, Miniature, Mosaic, <strong>and</strong> on Glass; of Gilding,<br />

Dyeing, <strong>and</strong> the Preparation of Colours <strong>and</strong> Artificial Gems, Vol. II. London: John<br />

Murray, 325-547. References to the so-called Bolognese manuscript (Segreti per<br />

Wallert 45


Colori, Biblioteca Universitaria MS 1536) are given as Bo1. MS, followed by<br />

recipe number <strong>and</strong> page number in Merrifield's publication.<br />

3. Wallert, A. 1991. Kookboeken en Koorboeken, Techniek en productie van boekverluchting<br />

in koorboeken van het klooster Monte Oliveto Maggiore. Groningen: Groningse Universiteits<br />

Drukkerij. References to Ambruogio da Sienas Ricepte daffare pili colori,<br />

Siena Biblioteca Comunale MS 1.JI.19, fols. 99r-106r, are given as Siena<br />

MS I, followed by recipe number <strong>and</strong> page number in Wallert's edition. The<br />

indication Siena MS II with recipe <strong>and</strong> page number refers to the edition in the<br />

same publication of Bartolomeo da Sienas Experientie, Siena Biblioteca Comunale,<br />

MS L.XI.41, fols. 34v-39v.<br />

4. This recipe is an abbrieviated version of Siena MS II, XVIII, p. 45. A version in<br />

Latin of the same recipe is in Bo1. MS, 160, p. 467.<br />

5. Alex<strong>and</strong>er, S. M. 1965. Medieval recipes describing the use of metals in manuscripts.<br />

Marsyas (XII) :34-51.<br />

6. An identical recipe is in Siena MS I, recipe XXIV, p. 35. An abbreviated version<br />

of this recipe is in Oxford (Bodleian Library, MS Conon. Misc. 128, fo1. 2v): "A<br />

mettere oro in carta."<br />

7. Wallert, A. 1990. Chrozophora tinctoria Juss.: Problems in identifying an illumination<br />

colorant. Restaurator (1 1):141-55.<br />

8. Identical recipe in Siena MS I, XVIII, p. 31. Similar descriptions of ingredients<br />

are given in Siena MS II, III, p. 37; Bo1. MS 389; <strong>and</strong> Bo1. MS 390, p. 597. Also<br />

see Zerdoun Bat-Jehouda, M. 1980. La fabrication des encres noires d'apres les<br />

textes, Codicologica 5; Les materiaux du livre manuscrit, Litterae textuales Leiden.<br />

9. Kuhn, H., 1961, Safran und dessen Nachweis durch Infrarotspektrographie in<br />

Malerei und Kunsth<strong>and</strong>werk, Leitz-Mitteilungen fUr Wissenschaft und Technik,<br />

Bd 11, (1):24-28. See also Straub, R. E. 1965. Der Traktat "De Clarea" in der<br />

Burgerbibliothek Bern, Schweizerisches Institut ju r Kunstwissenschajt, jahresbericht<br />

(12):89-1 14.<br />

10. The same but more complete recipe can be found in Bo1. MS 129, p. 447.<br />

11. Identical recipe Siena MS I, XXIII, p. 35. This recipe recurs in a slightly more<br />

elaborate form in Oxford, Bodleian Library, MS Conon. Mise. 128, fo1. 21r <strong>and</strong><br />

also in Bo1. MS, 77, p. 415, <strong>and</strong> Bo1. MS, 80, p. 417.<br />

12. Brachert, E <strong>and</strong> T. 1980. Zinnober. Maltechnik / Restauro (86):145-59.<br />

13. Siena MS II, XlV, p. 49.; Siena MS I, Xv, p. 29.<br />

14. Siena MS II, XII, p. 41. Also see Kurella, A., <strong>and</strong> I. Strauss. 1983. Lapis lazuli und<br />

naturliches Ultramarin. Maltechnik / Restauro (89):34-55.<br />

15. This recipe will not result in the promised blue pigment, but rather in the<br />

formation of the red mercuric sulphide. On the "azure-vermilion tangle," see<br />

Thompson, D. V 1956. The <strong>Materials</strong> <strong>and</strong> <strong>Techniques</strong> oj Medieval <strong>Painting</strong>. New<br />

York: Dover Reprints, 155.<br />

16. An identical recipe can be fo und in Siena MS I, XII, p. 29. Also similarities with<br />

Modena, Biblioteca Estense, MS ex.T.7.3., fo1. lOv. Also see Wallert, A. 1986.<br />

Verzino <strong>and</strong> Roseta Colours in 15th Century Italian Manuscripts. Maltechnik /<br />

Restauro (92):52-70.<br />

17. The recipe requires at this point the use of sulphur to be practically applicable.<br />

It is not clear what is meant by "cashina."<br />

18. Similar in Siena MS I, XXI, p. 32, <strong>and</strong> Siena MS II, XXI, p. 46.<br />

19. Probably a corruption of "porporino," a golden glistening tin (II) sulphide.<br />

20. The recipe calls for lead. Any useful pigment, however, can only be obtained if<br />

this lead were replaced with tin.<br />

21. The nitric salt, "sal nitro," should have been "sal armonic," i.e., al11110nium chloride.<br />

22. This recipe is slightly more intelligible than Simone's recipe 17. See Schiessl, U.<br />

1981. Musivgold, eine pigmentgeschichtliche Skizze. Maltechnik / Restauro (87):<br />

219-30.<br />

23. Florence, Biblioteca Nazionale, MS Magliabecchiana CL XV 8bis, fo1. 14v. More<br />

elaborate variations are in Siena MS I, III, p. 23, <strong>and</strong> in Florence, Biblioteca<br />

Laurenziana, MS Ashburnhamia 349, fols. 80v-81r.<br />

24. The weld herb of the dyers is Reseda luteola L. See Stoll, A. M. 1981. Gelbe<br />

Pflanzenlacke, Schuttgelb und Safran. Maltechnik / Restauro (87):73-1 11.<br />

25. An identical recipe can be found in Florence, Biblioteca Nazionale, MS Palatina<br />

916, fo1. 110r, "Affare profili d'oro brunito." A related recipe occurs in Bo1. MS,<br />

162, p. 469.<br />

26. This recipe occurs in a slighly more elaborate version in Siena MS II, VI, p. 39.<br />

27. Fior di lino is the cornflower, Centaurea cyanus L.<br />

46<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


28. On these "clothlet colors," see Wallert, A. 1993. Natural organic colorants on<br />

medieval parchment: anthocyanins. ICOM Committee fo r Conservation Preprints,<br />

516-23; <strong>and</strong> Willhauk, N. 1981. Farbprobleme spatmittelalterlicher Buchmalerei.<br />

Restaurator (1-2):103-34.<br />

29. Bietola is a beet or garden beet; bietolina = weld (Reseda luteola) (?).<br />

30. A plant with the same name, gilosia (Amaranthus tricolor?), is mentioned in Bol.<br />

MS, 117, p. 439, for making a purple clothlet color.<br />

31. Gettens R. J., <strong>and</strong> E. West Fitzhugh. 1966. Azurite <strong>and</strong> blue verditer. Studies in<br />

Conservation (2): 54-61.<br />

32. Quite similar is Siena MS I, IV, p. 22. In Simone's manuscript, however, the<br />

vermilion prescribed by Ambruogio is replaced for a black ink. The recipe recurs<br />

in a rather abbreviated version in Florence, Biblioteca Laurenziana, MS Ashburnhamia<br />

349, fol. 84v.<br />

Wallert 47


Abstract<br />

Primary sources can be extremely<br />

valuable for the study of historical<br />

painting techniques, materials, <strong>and</strong><br />

studio practice. Most authors of<br />

these treatises <strong>and</strong> manuals are either<br />

anonymous or little known. Quite<br />

often, therefore, no context for these<br />

books can be traced. In some cases,<br />

however, research on the author(s), if<br />

known, can reveal information that<br />

helps to interpret these sources,<br />

shedding light on the contents <strong>and</strong><br />

their fu nction in studio practice. An<br />

example can be fo und in a seventeenth-century<br />

Italian treatise on<br />

miniature painting which, according<br />

to its title, was written by the miniaturist<br />

Valerio Mariani da Pesaro in<br />

1620. Research on this treatise has<br />

revealed some unexpected aspects of<br />

its authorship.<br />

A Seventeenth-Century Italian Treatise on Miniature<br />

<strong>Painting</strong> <strong>and</strong> Its Author(s)<br />

Erma Hermens<br />

Kunsthistorisch Instituut<br />

Rijksuniversiteit Leiden<br />

Doelensteeg 16<br />

Postbus 9515<br />

2300 RA Leiden<br />

The Netherl<strong>and</strong>s<br />

The manuscripts<br />

The Leiden University Library owns a seventeenth-century Italian manuscript<br />

titled Della Miniatura, del Signor Valerio Mariani da Pesaro, Miniatore del serenissimo<br />

Signor Duca d' Urbino, del Signore Capitano Giorgio Maynwaringe, inglese,<br />

l'anno del Signore 1620, in Padova (sign. Voss. Ger. Gall. 5q). The manuscript,<br />

consisting of three sections, contains a treatise on miniature painting technique.<br />

The first section covers technical recipes for the preparation of pigments,<br />

dyestuffs <strong>and</strong> inks, binding media, <strong>and</strong> utensils such as pencils <strong>and</strong><br />

brushes. The second section gives instructions on how to paint l<strong>and</strong>scapes,<br />

<strong>and</strong> describes the mixtures of colors to use <strong>and</strong> the build-up in layers fo r the<br />

execution of different types of trees, weather conditions, water, sea, rocks,<br />

plants, mountains, villages, close-up <strong>and</strong> distant views, <strong>and</strong> so fo rth. The third<br />

<strong>and</strong> smallest section describes the mixtures of colors <strong>and</strong> the layering of colors<br />

used to depict skin <strong>and</strong> cloth in portrait painting.<br />

The recipes in the first section seem to be quite accurate, a result in part of<br />

the continuation of traditional methods, <strong>and</strong> in part of personal experiments<br />

by the writer. In general, one can say that the technique described is a watercolor<br />

technique fo r paper <strong>and</strong> parchment, in which the final effect is<br />

reached by using a mixture of pigments <strong>and</strong> dyestuffs or, as is more often <strong>and</strong><br />

more typical fo r this treatise, by several layers of washes of colors.<br />

During the research on the treatise, two more manuscripts were found to be<br />

connected with the Leiden manuscript. One manuscript, owned by the Beinecke<br />

Library, Yale University (sign. MS 372), is entitled Della miniatura del<br />

Signore Valerio Mariani da Pesaro, miniatore del Duca d'Urbino con aggiunte d'altre<br />

cose per l'istessa professione dal Signor D. Antonello Bertozzi scrittore e miniatore in<br />

Padova, per me Francesco Manlio Romano, l'anno MDCXX. The second manuscript,<br />

entitled Ricordi di belli colon:, is part of the Urbinati Latini collection in<br />

the Biblioteca Apostolica Vaticana, Citta del Vaticano, Italy (sign. Urb. Lat.<br />

1280). The three manuscripts will be referred to here as the Leiden, Yale, <strong>and</strong><br />

Rome manuscripts.<br />

The Yale manuscript, despite the differences in its title, is very similar to the<br />

Leiden manuscript, with only some minor variations in language. It contains,<br />

however, some extra pages at the end with additional recipes fo r pigments,<br />

paper dyeing, etching materials, as well as recipes fo r gunpowder, payments<br />

for books, a list of the quarters of Venice <strong>and</strong> other notes, some of them<br />

dated between 1614 <strong>and</strong> 1628. The extra recipes could be the "aggiunte<br />

d' altre cose per l'istessa professione dal Signor D. Antonello Bertozzi scrittore<br />

e miniatore in Padova."<br />

The Leiden manuscript contains a price list of pigments <strong>and</strong> dyestuffs <strong>and</strong><br />

also a list of all materials needed by the miniatore. Both are lacking in the<br />

Yale manuscript (see Appendix doc. 1).<br />

The Rome manuscript covers only the second part of the Leiden <strong>and</strong> Yale<br />

manuscripts, that is, the part on l<strong>and</strong>scape painting. We find here many differences<br />

in language <strong>and</strong> also in the sequence <strong>and</strong> length of the chapters. The<br />

L<strong>and</strong>scape section in the Leiden <strong>and</strong> Yale manuscripts is more extensive, as<br />

48<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


it also includes instructions fo r composition, the build-up of a miniature from<br />

a pencil drawing up to the final execution in paint, <strong>and</strong> so on. Some chapters<br />

are a combination of two or three short chapters from the Rome manuscript,<br />

covering more or less the same subjects. It seems that the writer of the treatise<br />

rearranged <strong>and</strong> exp<strong>and</strong>ed the text from the Rome manuscript to improve the<br />

contents (1).<br />

Valerio Mariani da Pesaro<br />

The miniaturist Valerio Mariani da Pesaro-the writer of the treatise, according<br />

to its title-is mentioned by Thieme-Becker as being active from<br />

1560 to 1600, <strong>and</strong> by Zani as living from 1565 to 1611, but further information<br />

on the treatise <strong>and</strong> its author seemed difficult to fmd (2). Archival<br />

research in Italy however proved otherwise, <strong>and</strong> the treatise can now be provided<br />

with an ample context.<br />

Mariani is described in Lancellotti's L'Hoggidi (1636) as a pupil of Giovanni<br />

Maria Boduino, a miniaturist who worked in Friuli (3). In a document from<br />

1582, Mariani is mentioned as the twelve-year-old pupil of Boduino, establishing<br />

Mariani's date of birth as 1570 (4). About Mariani's working environment<br />

after his training in Friuli, little information has been fo und so far. He<br />

worked, as he mentions in the treatise (Chapter XVI: Giallo Santo), for the<br />

Duke of Savoy: "Giallo santo is a colour which is extracted from the flower<br />

of the broom, as a painter in Borgo in Brescia taught me, while I stayed in<br />

that place in the service of the honourable memory of Duke Emanuele Filiberto,<br />

Duke of Savoy ..." (5). And, as can be read in the title of the treatise,<br />

he was also employed by the Duke of Urbino. According to the date of the<br />

treatise, this must have been Francesco Maria II della Rovere. The last known<br />

dates in Mariani's chronology are 1618, when he is mentioned as heir in his<br />

brother's will, <strong>and</strong> 1625, when he is mentioned as a debitore in the duke's<br />

bookkeeping (6).<br />

So far no documents on his work for the Savoy court or other employers<br />

have been fo und; therefore, we concentrate on Mariani's activities fo r the<br />

Urbino court, which provides us, as will become clear, with the context in<br />

which the treatise was written.<br />

Francesco Maria II della Rovere began his reign in 1574, after his father<br />

Guidobaldo II della Rovere died, leaving him with a bankrupt state. The<br />

young duke, who had a strict <strong>and</strong> sober upbringing at the court of Philip II,<br />

started a period of severe economizing, including trimming the court's cultural<br />

expenses. It was 1580 before the duke regained interest <strong>and</strong> renewed<br />

his financial backing fo r cultural <strong>and</strong> artistic activities. Many documents, as<br />

well as artworks he commissioned, survive the period of his reign, creating a<br />

clear image of the cultural policy <strong>and</strong> interests of the duke (7).<br />

As a contemporary chronicle states, "When Francesco Maria, the last Duke,<br />

had paid the large debts of his father Guidobaldo by stopping the many arts<br />

<strong>and</strong> famous crafts that were executed here, he changed his mind after seeing<br />

the severe consequences <strong>and</strong> ordered the fo undation of several workshops at<br />

his court where very excellent masters in every art <strong>and</strong> profession were put<br />

to work" (8). One such master was the miniaturist Valerio Mariani da Pesaro.<br />

These workshops fo rmed a well-organized business, providing artworks "to<br />

order" for the duke. As not only masters but apprentices were present, the<br />

workshops seem to provide the appropriate milieu for writing a technical<br />

manual. A sketch of the workshops will, therefore, be given.<br />

The workshops<br />

The duke made personal notes during the years 1580-1620 of all his "artistic<br />

expenses," including monthly payments to the botteghini (workshops) (9). He<br />

employed painters, sculptors, watchmakers, ebony workers, bookbinders, goldsmiths,<br />

<strong>and</strong> miniaturists in a lively <strong>and</strong> businesslike organization (10).<br />

Hermens<br />

49


The actual working areas of the botteghini were situated on the ground floor<br />

in the left wing of the ducal palace in Pesaro. There seems to have been an<br />

internal entrance from the palace (11). The botteghini also opened onto the<br />

street like small shops; indeed, the artists were permitted by the duke to accept<br />

orders from other clients. Some artists had their living quarters above the<br />

workshops; others who had families received money to rent housing elsewhere<br />

in Pesaro. They also received provisions, c<strong>and</strong>les, <strong>and</strong> firewood, as can<br />

be read in a ledger (12).<br />

A supervisor was appointed with responsibility for the daily organization,<br />

provisions, <strong>and</strong> the supply of materials. Bills with lists of ordered materials,<br />

including pigments <strong>and</strong> brushes, are among the many surviving documents<br />

concerning the workshops (see Appendix doc. 2). The duke's commissions<br />

were not given by the duke personally but passed on by intermediaries, mostly<br />

noblemen employed as his secretaries or persons suitable because of their<br />

profession. These intermediaries also acted as the duke's representatives in the<br />

search for artists <strong>and</strong> artisans for the workshops. For example, the duke's ambassadors<br />

were ordered to search in Rome, Venice, <strong>and</strong> Florence fo r artists of<br />

the highest quality. The proposed artists had to send proof of their capacities<br />

which, according to the documents, was quite often rejected by the duke.<br />

In 1581 the duke asked his ambassador in Rome, Baldo Falcucci, to find him<br />

a miniaturist. From 1582 on, a "miniatore" is mentioned in the lists of workshop<br />

employees (13). In many instances, the name of the artist is not given<br />

but a "maestro Valerio miniatore" appears from 1603 to 1605. Although employee<br />

lists from 1605 on are missing, Mariani's activities at the court of<br />

Urbino most likely continued.<br />

Figure 1. Detail from the Battle of San Fabiano,<br />

illustrated in Plate 7a, showing the<br />

center scene where the sketchy underdrawing is<br />

clearly visible. Photograph by E. Buzzegoli,<br />

LAboratorio di Restauro. Courtesy of the UjJ<br />

izi Gallery, Florence.<br />

Mariani's activities can be traced not only in documents but also in many<br />

miniatures. The Biblioteca Apostolica Vaticana has three illustrated manuscripts<br />

attributed to Mariani, in addition to a part of the Purgatorio <strong>and</strong> all<br />

the miniatures of the famous Dante Urbinate's Paradiso (14). The Galleria<br />

Palatina in Florence owns two series of miniatures, framed in black ebony<br />

frames, with scenes from the lives of Christ <strong>and</strong> Mary, miracles, <strong>and</strong> martyred<br />

saints. The Uffizi owns the only miniature with Mariani's signature, a battle<br />

scene from the life of Federico da Montefeltro (15) (Plate 7a). This miniature<br />

is of great interest as we can compare the techniques used with those as<br />

described in the treatise, thus comparing practice with theory. The miniature<br />

is kept in a dark cabinet; the colors are, therefore, in extremely good condition.<br />

A study of the miniature using microscopy made clear that the st<strong>and</strong>ard<br />

technique of washes was used. Specific comparisons (i.e., the build-up of the<br />

trees, background with the village, bluish mountains <strong>and</strong> sky, etc.) with chapters<br />

in the l<strong>and</strong>scape section of the treatise, show that Mariani is following<br />

the instructions given there. The miniature has a very sketchy underdrawing-which<br />

can be seen with the naked eye but is even more clearly apparent<br />

with an infrared camera-that has been traced in some places with a brownish<br />

ink (Fig. 1) (16). A blue-black ink is also used to emphasize contours <strong>and</strong><br />

shadows. On the underdrawing, the paint is most often added in transparent<br />

layers using the techniques described in the treatise.<br />

Mariani was known for his technical abilities; Lancellotti described him as a<br />

pupil of Boduino who "surpassed his master in patience, <strong>and</strong> his miniatures<br />

were owned by the most important princes in the world." They both "kept<br />

a secret of how to grind gold in the Persian way for miniature painting <strong>and</strong><br />

writing, that stayed stable as the antique [gold] " (17).<br />

Although the preceding paragraphs comprise a short resume of the research<br />

results so far, it is clear that we are considering a treatise that can be placed<br />

in an interesting context, offering a unique possibility of comparing the theory<br />

of the treatise with the practice as laid down in Mariani's work. The<br />

context of the workshops, <strong>and</strong> Mariani's obvious technical skills, make it plausible<br />

that the treatise was written as a manual to be used in the workshop.<br />

50<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Ricordi di belli colori<br />

Initially it seemed reasonable to assume that the Rome manuscript was writter.<br />

Sy Mariani as a draft version of the l<strong>and</strong>scape section. The manuscript<br />

wa, then reworked <strong>and</strong> exp<strong>and</strong>ed for the larger treatise of which, presumably,<br />

both the Leiden <strong>and</strong> Yale manuscripts are copies. However, there are some<br />

remarkable features that made us continue the search fo r the origin of the<br />

Rome manuscript, leading to interesting results.<br />

The manuscript's text covers most of the chapters on specific l<strong>and</strong>scape elements<br />

that are present in the Leiden <strong>and</strong> Yale manuscripts, but lacks the<br />

chapters on composition <strong>and</strong> underdrawing. There are also considerable differences<br />

in the order <strong>and</strong> composition of the chapters <strong>and</strong> in the language,<br />

as well as some repetitions <strong>and</strong> crossing out of text. First, a comparison with<br />

other Italian treatises on painting technique from the sixteenth <strong>and</strong> early<br />

seventeenth century shows that no specific attention is given to l<strong>and</strong>scape<br />

painting, especially not in such great detail as we see in the Rome manuscript.<br />

L<strong>and</strong>scape painting of that period was simply a Flemish specialty. It reached<br />

its peak after Pieter Breughel the Elder introduced the new spatial concept<br />

of several planes leading to a distant vista in which the religious theme became<br />

a detail instead of the main, foreground scene. The l<strong>and</strong>scapes produced by<br />

the many Flemish artists working in Italy were enormously popular, <strong>and</strong><br />

l<strong>and</strong>scape print series circulated widely (18). The Rome treatise seems to<br />

reflect a Flemish sense of detail concerning l<strong>and</strong>scape elements, an idea supported<br />

by the presence of Flemish painter Michiel Gast's name in a chapter<br />

on how to paint villages in the background. Gast's methods are found to be<br />

exemplary. We know little about Michiel Gast today except that he came to<br />

Rome as a pupil of Lorenzo of Rotterdam <strong>and</strong> was known for his paintings<br />

of ruins (19). The citation of Gast's methods <strong>and</strong> the detailed attention to<br />

l<strong>and</strong>scape elements suggests a strong Flemish influence on the author of Ricordi<br />

di belli colori.<br />

Another remarkable feature of the Rome manuscript are the chapters on<br />

flowers <strong>and</strong> plants. Although the plants described are quite common in Italy,<br />

the names used are often botanical ones, probably most familiar to a person<br />

with a special botanical interest. In these chapters some personal remarks are<br />

present. For example, the writer says, "I have made a yellow flower in my<br />

book with herbs, the biggest one without l<strong>and</strong>scapes" <strong>and</strong> "to represent well<br />

the fruit of the Jusaina or otherwise nocella as that is how it is called in the<br />

village of Rocca Contrada." It is striking that all such personal remarks made<br />

by the author are omitted in the l<strong>and</strong>scape section in the Leiden <strong>and</strong> Yale<br />

manuscripts, while personal remarks are retained in the sections of those<br />

manuscripts containing recipes for pigments <strong>and</strong> dyestuffs .<br />

A reading of the manuscript suggests that the writer lived in Rocca Contrada,<br />

painted l<strong>and</strong>scapes, came from a Flemish background (or was strongly influenced<br />

by Flemish l<strong>and</strong>scape painting), had botanical interests, <strong>and</strong> probably<br />

illustrated a herbarium. Knowing this, should we still consider Valerio Mariani<br />

da Pesaro as the writer of the Rome manuscript or should we look elsewhere?<br />

It was in fact the citation of Michie! Gast's name which helped us to answer<br />

this question.<br />

Gherardo Cibo<br />

One of the very few known works by Gast can be found in the collection<br />

of Gherardo Cibo, an Italian nobleman-artist-botanist (1512-1600). Cibo<br />

lived for most of his life in the small village of Rocca Contrada, now called<br />

Arcevio, in the Marche, the region of the Duchy of Urbino. Cibo started an<br />

ecclesiastic career in Rome but he did not pursue it, settling in 1539 in Rocca<br />

Contrada where he occupied himself with artistic <strong>and</strong> botanical activities.<br />

Although in obscurity until the 1980s, research has revealed that Cibo was<br />

responsible fo r many l<strong>and</strong>scape drawings in major collections that had been<br />

anonymous or wrongly attributed, often to Flemish artists (Fig. 2). A 1989<br />

Hermens 51


Figure 2. Gherardo Cibo, L<strong>and</strong>scape. Courtesy of the Printrooltl if the University of Leiden, the<br />

Netherl<strong>and</strong>s (il1v. nr. A W 441).<br />

exhibition <strong>and</strong> catalogue dedicated to Gherardo Cibo shed much light on<br />

the artist's career <strong>and</strong> personality (20). A large group of l<strong>and</strong>scape drawings,<br />

several herbaria, <strong>and</strong> illustrated editions of Mattioli's translated version of<br />

Dioscorides can now be attributed to Cibo (Plate 7b, Fig. 3). According to<br />

contemporary testimonies, Cibo must have been an interesting personality<br />

who came from a wealthy family. Praised by the local population for his<br />

virtues <strong>and</strong> charity work, Cibo occupied himself with botanical expeditions<br />

<strong>and</strong> painting. A local seventeenth-century historian alludes to Cibo's high<br />

artistic qualities <strong>and</strong> his method of preparing colors by extracting the dyestuffs<br />

from herbs, fruits, <strong>and</strong> seeds (21).<br />

Figure 3. Gherardo Cibo, "Hell/ionite, "<br />

from Herbarium (MS ADD 22332),<br />

fol. 143r, ca. 1570. Courtesy of the British<br />

Library.<br />

Could Gherardo Cibo be the writer of Ricordi di belli colori? Proof was to<br />

be found by comparing Cibo's works with the Rome manuscript. In the<br />

manuscript we not only find instructions on the mixtures of colors <strong>and</strong> the<br />

build-up of transparent washes of paint to achieve certain effects but also<br />

instructions specifically directed toward working on pen drawings or prints<br />

(i.e., Chapter XXIX: "Beautiful green to use on prints <strong>and</strong> on plants drawn<br />

with pen"; Chapter XXXVI: "To colour herbs, printed or drawn with pen").<br />

Cibo's works show that he did both. Many of his l<strong>and</strong>scape drawings were<br />

executed in pen <strong>and</strong> ink <strong>and</strong> then colored with transparently applied colors.<br />

He also colored prints in several editions of Mattioli's translation of Dioscorides<br />

<strong>and</strong> added l<strong>and</strong>scape backgrounds as a personal touch.<br />

Additional proof can be found in the text. The personal remarks cited above<br />

concerning the chapters on plants point to the herbaria illustrated by Cibo.<br />

Two of these drawings are kept in the British Library (22). The largest herbarium<br />

illustrated contains only plants, drawn in ink <strong>and</strong> colored with watercolor,<br />

without l<strong>and</strong>scapes in the background: " . .. my largest book on herbs<br />

without l<strong>and</strong>scapes." The other smaller herbarium does have l<strong>and</strong>scapes; on<br />

folio 183v, one reads, "Fusaina salva[ti]ca, nocella qui chiamata a Roccha<br />

C[ontrada]," which corresponds fully with the first lines of Chapter XLV of<br />

the manuscript, in which the instructions on how to paint this flower are<br />

given. The author describes several color mixtures for the Jusaina or nocella<br />

flowers <strong>and</strong> folio 184v in the herbarium can be fo und many color samples<br />

<strong>and</strong> some fusaina flowers where these mixtures are apparently tested (Figs. 4,<br />

5).<br />

Finally, examination reveals that the Rome manuscript is written in two<br />

h<strong>and</strong>s. One author wrote only two pages; the script on these pages can be<br />

52<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 4. Gherardo Cibo, "Fusaina [. .. J<br />

nocella qui chiamata a Roccha C [ontrada J, "<br />

Herbarium, (MS ADD 22332), Jo l.<br />

183v. Courtesy of the British Library.<br />

Figure 5. Gherardo Cibo, several proofs oj<br />

colors <strong>and</strong> "Fusaina" flowers, Jrom Herbarium<br />

(MS ADD 22332), 101. 184v. Courtesy<br />

of the British Library.<br />

identified as Cibo's h<strong>and</strong>writing. The conclusion must be that Cibo-not<br />

Mariani-was the author of the Rome manuscript.<br />

Gherardo Cibo <strong>and</strong> Valerio Mariani da Pesaro<br />

Although Gherardo Cibo's part in the origin of the treatise is established, we<br />

may still assume that Valerio Mariani composed the treatise, using Cibo's<br />

specific knowledge <strong>and</strong> thus incorporating the section on l<strong>and</strong>scape painting.<br />

There is no reason, thus far, not to assume that Mariani wrote the first section<br />

with technical recipes, especially as some personal notes present cannot be<br />

attributed to Cibo, pointing to a professional miniaturist who was working<br />

on a commission basis.<br />

Did Cibo <strong>and</strong> Mariani meet? Cibo died in 1600 when Mariani was thirty<br />

years old. His name appears in the employee lists for the first time in 1603,<br />

which does not indicate he was not working in Pesaro before that date, as<br />

most of the time the workshop employees are only indicated by their function<br />

(i.e., miniatore) <strong>and</strong> the duke had started his search for capable miniaturists<br />

in 1581. If they did meet, Cibo must have been in his late seventies or early<br />

eighties but still active, as a letter he wrote to the Duke Francesco Maria II<br />

della Rovere indicates. In the letter, dated 1580, Cibo tells the duke he is<br />

very honored by his request to illustrate an edition of Mattioli's Dioscorides;<br />

Cibo finished the work at nearly seventy years of age (23). This letter, however,<br />

also indicates a clear contact between Cibo <strong>and</strong> the duke, but there were<br />

apparently more contacts with the Urbino court.<br />

In the Biblioteca Comunale in Jesi (Marche, Italy), an album is kept with<br />

l<strong>and</strong>scape drawings, mainly by Cibo. The text on the cover of the album is<br />

in Cibo's h<strong>and</strong>writing <strong>and</strong> says that the album contains "a little l<strong>and</strong>scape on<br />

paper from the h<strong>and</strong> of the Flemish painter who serves our illustrious Duke<br />

of Urbino, which Sir Cavaliere Ardoino sent me, in April I think. 1591. And<br />

he names himself M[aestro] Giovanne. There are here two drawings of<br />

M[aestro] giovanne fiame[n]go from l<strong>and</strong>scapes on coloured paper ...." It<br />

also mentions drawings of "the Painter from ForB" (24).<br />

Cavaliere (knight) Ardoino can be identified as Girolamo Ardovino (also Ardoino<br />

or Arduini), the duke's architect (25). In many documents, he figures<br />

Hermens 53


as an intermediary between the duke <strong>and</strong> the workshops. Maestro Giovanne<br />

Fiamengo <strong>and</strong> the Pittore da ForIi both appear in the many documents concerning<br />

the workshops as being employed by the duke. Cibo was obviously<br />

acquainted with Ardovino <strong>and</strong>, therefore, must have been familiar with the<br />

workshops <strong>and</strong> the artists working there. He might have known the young<br />

Mariani if the latter did arrive in Pesaro before 1600. In any case, he probably<br />

knew the miniaturists' workshop.<br />

The connection<br />

As mentioned earlier, the Rome manuscript is written in two h<strong>and</strong>s. The<br />

two pages in Cibo's h<strong>and</strong>writing comprise two chapters, one chapter on the<br />

mixture for meadows <strong>and</strong> fields (with two color samples in the margin), <strong>and</strong><br />

one chapter with more general instructions on mixtures of colors (see Appendix<br />

doc. 3). At the end of the latter chapter, the following is written: " . ..<br />

such that if you Sir, will try to be a little less lazy <strong>and</strong> exercise more often,<br />

the exercises will work out very well." This last sentence makes clear that the<br />

manuscript was intended as a manual for someone who wanted to learn, or<br />

was ordered to learn, to paint <strong>and</strong> draw l<strong>and</strong>scapes like Cibo did, that is, in<br />

the Flemish way. Although the manuscript is part of the Urbinati Latini collection<br />

of the Vatican Library, it seems unlikely that it was written fo r the<br />

duke himself, as the final remark is not very suitable fo r addressing a duke. It<br />

is reasonable to assume, however, that the duke, obviously impressed by Cibo's<br />

artistic capabilities, asked him to write down his techniques for the execution<br />

of l<strong>and</strong>scapes in the Flemish manner in a manual that could be used by the<br />

duke's miniaturists' workshop. The different h<strong>and</strong>writing, corrections, <strong>and</strong><br />

repetitions may indicate that Cibo dictated the text to someone, except for<br />

the chapter in which he addresses the person for whom the manuscript was<br />

meant. This person might have been the young Mariani. If not, then Mariani<br />

found the little book on l<strong>and</strong>scape painting, reworked it, <strong>and</strong> included it in<br />

his treatise. The pigments <strong>and</strong> dyestuffs that appear in the l<strong>and</strong>scape sectionsome<br />

of them rather uncommon, such as giallo de' vasari <strong>and</strong> bruno d'Inghilterra--are<br />

all described in Mariani's recipes (in the first section of the Leiden<br />

<strong>and</strong> Yale manuscripts) . Cibo really did experiment with extracting colors from<br />

plants <strong>and</strong> fruits, as well as with mixtures of pigments, as similar proofs of the<br />

color samples described previously can be found throughout all Cibo's works,<br />

even in the margin of the text of the Rome manuscript. These ricordi (reminders)<br />

may have inspired Mariani to describe exactly those pigments <strong>and</strong><br />

dyestuffs Cibo used. Interesting is the use of the name giallo de' vasari (potter's<br />

yellow) for lead-antimone yellow. In the Marche, the most important <strong>and</strong><br />

famous majolica industry of Italy was flourishing in the sixteenth century.<br />

Cibo presumably obtained this pigment directly from the potters. In the recipe<br />

for bruno de Inghilterra, Mariani says the color was known to spetiali et<br />

pittori (pharmacists <strong>and</strong> painters) but that so far he had not been able to<br />

discover the exact composition. Most of the other recipes are traditional,<br />

except for those using plants <strong>and</strong> fruits to extract dyestuffs; the latter recipes<br />

may also come from Cibo. The extra chapters on composition <strong>and</strong> underdrawing,<br />

although not present in Ricordi di belli colori, seem to reflect Cibo's<br />

techniques when compared with his work.<br />

Although no definite proof can be given, a personal exchange of information<br />

between Mariani <strong>and</strong> Cibo cannot be excluded.<br />

A comparison between practice <strong>and</strong> theory is possible, as ample material can<br />

be researched. Not only Mariani's miniatures, especially the signed one, but<br />

also Cibo's many l<strong>and</strong>scape drawings <strong>and</strong> the herbaria <strong>and</strong> illustrated Dioscorides<br />

editions, provide this opportunity. It is clear that Mariani used the<br />

same method of building up transparent layers in the background l<strong>and</strong>scape<br />

of his signed miniature as is described in Cibo's instructions (Plate 8). Mariani's<br />

technique, although clearly guided by a personal interpretation, is very<br />

similar to that used in Cibo's l<strong>and</strong>scape drawings executed in ink <strong>and</strong> colored<br />

with transparent watercolors.<br />

54<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


The treatise is typical for the late sixteenth <strong>and</strong> early seventeenth century in<br />

describing the techniques for the execution of two extremely important genres<br />

of painting in that period: l<strong>and</strong>scapes <strong>and</strong> miniature portraits. Both authors<br />

provide us with valuable information on the materials used, the techniques,<br />

<strong>and</strong> the studio practice, as well as the setting <strong>and</strong> the possible cooperation<br />

that provided the inspiration to write the treatise.<br />

Appendix<br />

Document 1. List of materials needed by the miniaturist, most described in<br />

the recipes in the first section of the treatise (Leiden, University Library, Voss.<br />

Ger. Gall. Sq. c.2v.):<br />

Biacca, Cinabro, Minio, Biadetto chiaro e scuro, Azurro oltramarino,<br />

Azurro grosso, Giallolino f ino e sottile, Giallo santo, Terra negra, Terra<br />

rossa, Terra gialla, Terra d'ombra, Acqua verde, Tomasole, Terra gialla<br />

brugnata, Lacca di grana, Lacca di verzino, Verde di giglio paonazzo,<br />

Fuligine, Anaraccio, Negro Ju me, Negro d'osso di persico, Terra di campana,<br />

Orpimento, Terra verde 0 verde porro, Verd'azzurro e verdetto,<br />

Indico, ZaJfarano, Verde rame orientale, Bol armenio 0 ocria lemnia, S<strong>and</strong>racca,<br />

Bruno d'Inghilterra 0 color di sale, Sangue di drago, Acqua gialla<br />

di spino cavino, Marchesita arsa, Giallo de' vasari, Ocria, Ocria abrugiata,<br />

Acqua di verzino, Carbone, Acqua e aceto gommato e calcinato, Inchiostro,<br />

Liscia, Lapis rosso e negro, Piombino per disegnare.<br />

Document 2. List of materials ordered by the supervisor Felice Antonio di<br />

Letiero for the workshops (Archivio di Stato, Florence, Fondo Urbino, Appendice,<br />

Filza 54, c.n.n.):<br />

A dl 15 9bre 1607. [ ... ] In per ordine di Maestro Felice Antonio di<br />

Letiero a di 24 detto per Ii botteghini di S[ua] A{ltezza]: Libra 2 112<br />

Azurro di Spagna fino [ ... ]; Biadetto libre 4 [ ... ]; Verdetto 3 112 [ ... ]<br />

Giallolino di Fi<strong>and</strong>ra libre 4 1/2 [ ... ]; Verdetto libre 6 112 [ ... ]; Azurro<br />

di Spagna chiaro libre 4 oncie 3 [ . .. ]; Lacca di grana fina libre 2 [ ... ];<br />

Smaltino di Fi<strong>and</strong>ra libre 10 1/2 [ ... ]; Sma/tine diverse con cor po libra 1<br />

oncie 9 [ ... ]; Penelli di sedoli con asta [ ... ]; Penelli grossi di vano [ ... ];<br />

Penelli mezani [ ... ]; Penelli piccoli [ ... ]; Cocciole di madre perla no. l00<br />

[ ... ]; Jacchino per portare detti robbe [ ... ]. A di 14 di Xbre 1607. 10 Felice<br />

Antonio di Lettiero ho ricevuto Ii sopradetti robbe per servitio di botteghini<br />

di Sua Altezza. Umilmente Felice Antonio.<br />

Document 3. Rome, Biblioteca Apostolica Vaticana, Urb. Lat. 1280, c.18v-<br />

19r.<br />

Avertimenti sopra Ie mestiche.<br />

Bisogna di avertire che tutte Ie sorte de mach ie, tanto per arbori quanta<br />

per sassi e per greppi et prati, che e necessario di Ja rle piu oscure, 0 piu<br />

chiare secondo la qualita del paese, anchora che tirino piu al verde, 0 piu<br />

al giallo, 0 piu all' azzurro, 0 piu al rossigno, 0 piu al negro secondo che<br />

piu piacera, 0 secondo che ricercara il disegno. La prima coperta non vol<br />

mai essere troppo chiara per potere dar poi i lumi secondo il bisogno, pero<br />

sempre sara bene mesticare con la prima machia quei colori che haveranno<br />

da s f oleggiare, poi illuminarle secondo si conviene e che satiffacci all'occhio<br />

et tanto si havera da Ja re per prati, campi et sassi quanto per Ii arbori.<br />

Non si po Jar questo senza un po' de Ja tica et perdimento de' colori. Pero<br />

chi vorra Ja re che Ie cose passino secondo il dovere, bisognara che componendo<br />

Ie mestiche vadi prov<strong>and</strong>o 0 sopra carte, 0 sopra tele, secondo dove<br />

si havera da des igna re, et haver tanto di pazienza che si asciughi et parendole<br />

poi 0 troppo chiara 0 troppo scura, 0 troppo negra, 0 troppo azurra,<br />

o troppo gialla, 0 troppo rossegiante, potra sempre rimediare con giongere<br />

piu di quel colore che sara necessario. Et sempre e bene Jar tanta mesticha<br />

che piu presto possi avanzare che manchare, perche non sola la persona se<br />

ne potra servire ad ogni hora per quel disegno, ma perche manc<strong>and</strong>o havera<br />

gr<strong>and</strong>issima Ja tica de poterla rifare come Jece la prima volta. Questo e piu<br />

Hermens 55


necessario di fa re che di haver 10 esempio delle machie perche si potra fa re<br />

qui una machia che volendola usar simile 0 sara gran fatica farla dello<br />

istesso colore, avera sara 0 troppo chiara 0 troppo scura 0 de altra qualita<br />

che non fara a proposito per il disegno. Si che Vostro Signore se risolvi de'<br />

spigrirsi un poco et de esercitarse pitA spesso et cosi I esercitio passara benissimo.<br />

Acknowledgments<br />

This research has been carried out with the financial support of The Netherl<strong>and</strong>s<br />

Organization fo r Scientific research (NWO). I especially want to thank Tim Bedford<br />

for a critical reading <strong>and</strong> correction of the manuscript <strong>and</strong> fo r his encouragement.<br />

Notes<br />

1. From the results of linguistic research of the Leiden <strong>and</strong> Yale manuscripts, we<br />

may assume that both manuscripts are copies from the same, presumably original,<br />

manuscript. The h<strong>and</strong>writing as well as the paper quality <strong>and</strong> watermarks of the<br />

Leiden manuscript seem to point at an early date, possibly early or mid-seventeenth<br />

century. The Yale manuscript contains extra notes, some of them dated<br />

between 1614 <strong>and</strong> 1628, suggesting the early 1620s as the date of the copy. A<br />

complete edition of the treatise, with technical <strong>and</strong> art historical comments by<br />

the present author, is forthcoming.<br />

2. Thieme, U, <strong>and</strong> F. Becker. 1930. Algemeines Lexicon der hildende Kunstler von der<br />

Antike his der Gegenwart, XXXI, Leipzig, 64-65. See also Zani, P. 1823. Enciclopedia<br />

metodico-critico-ragionata delle helle arti, part I, Vo!' XIII. Parma, 44, 457, note 28.<br />

3. Lancellotti, S. 1636. L'Hoggid£. Venezia, 310.<br />

4. Zuccolo Padrono, G. M. 1969. Miniature manieristiche nelle comn"lissioni dogali<br />

del 110 cinquecento presso il Museo Correr. Bolletino dei Musei Civici Veneziani<br />

72 (XIV):6, 16, note 6.<br />

5. "II giallo santo e un colore che si cava dal fiore della ginestrella, come m'insegno<br />

un pittore a borgo in Brescia, mentre stava in qual luogo nel servitio del serenissima<br />

memoria del Duca Emanuele Filiberto di Savoia ..." Leiden: University<br />

Library (Voss. Germ. Gall. 5q, fo!' 16r)<br />

6. For the will, see Pesaro, Biblioteca Oliveriana (MS 455), Spogli Almerici, fo!'<br />

338v. For Mariani's debt, see Florence, Archivio di Stato, Fondo Urbino, Classe<br />

III, Filza XXIV, fo!' 242v.<br />

7. Most documents are kept in the Archivio di Stato, Florence <strong>and</strong> the Biblioteca<br />

Oliveriana in Pesaro. See also Gronau, G. 1935. Documenti Artistici Urhinati. Firenze.<br />

8. Biblioteca Oliveriana, Pesaro (MS Olivo 1009) Cavaliere Domenico Bonamini,<br />

Biografie degli Uomini Illustri Pesaresi. fo!' 290: ". .. l'ultimo Duca Francesco Maria<br />

che avendo prima risecate la grossa spesa del Duca Guidobaldo suo padre col<br />

far cessar tanti arti e celebri manifatte che qui si professavano, che vedendone in<br />

seguito il gravissimo derivo, si ricredette e com<strong>and</strong>o che si esigessero sotto la di<br />

lui corte, varie ollicine, aile quali deputo eccellentissimi maestri in ogni professione<br />

ed arti."<br />

9. These lists can be fo und in the Archivio di Stato, Florence, Fondo Urbino, Classe<br />

III, Filza XXIII (1581-1610 <strong>and</strong> 1613-1620), Classe I, Div. G., Filza CVI (161 1-<br />

1612).<br />

10. Archivio di Stato, Florence, Fondo Urbino, Classe III, Filza XXIII, fo!' 474r:<br />

"Lista della famiglia e salariati di S.A.S. e provisioni del'anno 1603: Maestro<br />

Propertio Grilli orologgiero, bocche 1; Maestro Cesare Maggiero pittore, bocche<br />

2; Maestro Giovanni Fiamengo, bocche 2; Maestro Simuntio miniatore, bocche<br />

2; Maestro Valerio miniatore, bocche 2; Maestro Nicolo argentiero, bocche 2;<br />

Maestro Boniforte orefici, bocche 2; Maestro Jacopo Cossa lavora d'ebbano, bocche<br />

2; Maestro Gervasio libraro ...." The name of the artist <strong>and</strong> the number of<br />

people working in the workshop are mentioned: the artist, plus an assistant or<br />

apprentice, equals hocche 2 (two mouths).<br />

11. Archivio di Stato, Florence, Fondo Urbino, Appendice, filza 48, not numbered.<br />

In this document the decoration of some rooms on the first floor of the palace<br />

is described: "Dalla porta che entra nella galleria nuova alia porta che va aile<br />

botteghe ...."<br />

12. Archivio di Stato, Florence, Fondo Urbino, Appendice, Filza 32: fo!' 13r a.f.<br />

13. The letters from Falcucci to the duke concerning the search fo r a miniaturist<br />

can be fo und in the Archivio di Stato, Florence, Fondo Urbino, Classe I, Div. G,<br />

56<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Filza CXLIII, fols. 1122r, 1125v, 1129v, 1133r; <strong>and</strong> Biblioteca Oliveriana, Pesaro<br />

(MS 375), vol. XXIV, fols. 227v, 212r, 207r, 179r. On the miniaturists, see also<br />

Meloni Trkulja, S. 1981. I miniatori di Francesco Maria II della Rovere. In<br />

Omaggio ai Della Rovere, ed. P Dal Poggetto. Pesaro.<br />

14. Michelini-Tocci, L. 1965. II Dante Urbinate della Biblioteca Vaticana. Facsimile.<br />

Milano.<br />

15. Meloni Trkulja, op. cit. (note 13). From the same author: 1980. Firenze e la<br />

Toscana dei Medici nell'europa del Cinquecento, vol. Palazzo Vecchio: commitenza e<br />

collezionismo medicei, Florence: 200-205. See fo r more information on Mariani<br />

<strong>and</strong> his miniatures: Hermens, E. 1993. Valerio Mariani da Pesaro: a seventeenthcentury<br />

Italian miniaturist <strong>and</strong> his treatise. Miniatura (3/4):80-93.<br />

16. See Hermens, E. 1993. Valerio Mariani da Pesaro <strong>and</strong> his treatise: theory <strong>and</strong><br />

practice. Technologia Artis (3):109-12. The infrared research was executed by E.<br />

Buzzegoli <strong>and</strong> D. Kunzelman, Laboratorio di Restauro, Uffizi, Florence.<br />

17. Lancellotti, op. cit., 310. "Valerio Mariani di Urbino fu suo discepolo e avanzo<br />

il maestro di patienza e delle sue miniature hebbero i maggiori Prencipi del<br />

mondo .... Havevano un degreto di macinar l' oro alia Persiana per miniare e<br />

per iscrivere, che stava saldo come I'antico."<br />

18. Tongiorgi Tomasi, L. 1989. Gherardo Cibo: Visions of l<strong>and</strong>scape <strong>and</strong> the botanical<br />

sciences in a sixteenth-century artist. Journal oj Carden History 9 (4) :208.<br />

19. Thieme <strong>and</strong> Becker, op. cit., XIII, 240. See also Bertolotti, A. 1880. Artisti Belgi<br />

ed Ol<strong>and</strong>esi, Roma, 41-4. See also Hoogewerff, G. 1. 1912. Nederl<strong>and</strong>sche schilders<br />

in Italie in de XIVe eeuw, Utrecht, 142-43.<br />

20. Nesselrath, A., 1989. Gherardo Cibo: "qui non e cognito." In cherardo Cibo alias<br />

"Ulisse Severino da Cingoli, " ed. A. Nesselrath. Comune di San Severino Marche,<br />

5-36.<br />

21. Celani, E. 1902. Sopra un erbario di Gherardo Cibo conservato nella R. Biblioteca<br />

Angelica di Roma. Malphigia XVI:I0.<br />

22. British Library, London, (MS Add 22333 <strong>and</strong> 22332). See Tongiorgi, op. cit.<br />

23. Cibo's letter can be fo und in the Archivio di Stato, Florence, Fondo Urbino,<br />

Classe I, div.G, Filza CCLX, fol. 6r. For the edition of Diose or ides, see Nesselrath,<br />

op. cit., catalogue no. 20, 100. Roma, Biblioteca Aless<strong>and</strong>rina (inv. Rari 278).<br />

24. Idem, catalogue no.40: 123-27. Jesi, Biblioteca Communale, Album A: "Qui dentro<br />

sta un paesetto in carta reale, de mano del pittor[e] Fiamengo ch[e] serve il<br />

n[ostro] Duca Sereniss[im]o di Urbino: che mi mano il S[ignor]e Cavalier Ardoino.<br />

Di Aprile mi pare. 1591. E chiamase M Giovanne." He also mentions<br />

drawings "quelli de mano d[e]l pittor[e] da ForE ...."<br />

25. Thieme <strong>and</strong> Becker, op. cit., II, 78, under the name Girolamo Arduini. See also<br />

Zani op.cit. (note 3), II, 181, 320, note 138. Arduini is said to have written a<br />

treatise titled Trattato del modo di piantare e Jortificare una cittCt.<br />

Hermens 57


Abstract<br />

The problems of interpretation of<br />

written sources on painting technique<br />

are well known. Through loss<br />

of the technical tradition, within<br />

which details of information were<br />

well understood at the time of writing,<br />

technical information is obscured<br />

fo r later generations. In<br />

courses on historical techniques of<br />

painting at the School of Conservation<br />

in Copenhagen, attempts at reconstructing<br />

the kind of gesso<br />

ground used in early Italian painting<br />

have prompted investigation into the<br />

actual meaning of the "giesso uolteriano"<br />

mentioned by Cennino Cennini<br />

in his treatise. This paper examines<br />

the problem from three angles:<br />

(1) the possible meanings of Cennini's<br />

text on this point; (2) the preparation<br />

of gesso grounds from the<br />

possible fo rms of gesso resulting<br />

from the first point (dihydrate, hemihydrate,<br />

anhydrite); <strong>and</strong> (3) technical<br />

evaluation of the reconstructions <strong>and</strong><br />

comparison with the results of scientific<br />

examination of grounds in early<br />

Italian painting.<br />

Questions about Medieval Gesso Grounds<br />

Beate Federspiel<br />

Konservators Skolen<br />

Danske Kunst Akademie<br />

Esplanaden 34<br />

1263 Kopenhagen-K<br />

Denmark<br />

Introduction<br />

The problems in interpreting written sources on painting techniques are well<br />

known. Besides the paintings themselves, the written sources are the only<br />

testimony of materials <strong>and</strong> techniques used in former times. Advanced methods<br />

of scientific analysis employed in the examination of paintings do not<br />

always answer questions about materials <strong>and</strong> techniques. And the written<br />

sources do not always provide easy access to painting techniques. Time has<br />

obscured the comprehension of the texts. The pure linguistic translation of a<br />

written source is often far from sufficient, but may be greatly aided by reconstructing<br />

the technical details described <strong>and</strong> comparing the results of the<br />

scientific analyses.<br />

Underst<strong>and</strong>ing Cennini's text<br />

During the courses in historical techniques of painting at the School of Conservation<br />

in Copenhagen, it was increasingly dissatisfying for the author <strong>and</strong><br />

others to reconstruct the gesso grounds as described by Cennino Cennini,<br />

<strong>and</strong> it was eventually necessary to scrutinize his text concerning grounds.<br />

Merely reading Cennini's instructions in the English translation by D. V<br />

Thompson in 1933 was not sufficient; it was obvious that things were not as<br />

easy as they may have seemed (1). Not that Cennini is imprecise in his instructions<br />

on this point; he is more thorough in his instructions on ground<br />

than in his description of paint application. But how can certain important<br />

passages be interpreted 600 years later?<br />

The original manuscript by Cennini being lost, the question of which surviving<br />

copy to use as a source remains, of course, a central one. That aspect<br />

will not be addressed in this article; the source used here is Lindberg's Swedish<br />

version of Cennini's Codex Laurentianus. Lindberg's translation shows semantic<br />

details, absent in previous translations, that are important for the underst<strong>and</strong>ing<br />

of decisive technical details (2).<br />

In Chapter CXV of his treatise, for example, Cennini describes the preparation<br />

of the ground for painting on panel. <strong>Painting</strong> in the Middle Ages<br />

included gilding. Gilding was the main reason for the great efforts invested<br />

in creating a perfect ground. Gypsum was the material used in the preparations<br />

of grounds for painting <strong>and</strong> gilding throughout the whole Mediterranean<br />

area as far back as the first millennium B.C.E. (3). The first written<br />

evidence of a ground for painting made of gypsum appears in the ninthcentury<br />

Lucca manuscript, which mentions a ground consisting of gypsum<br />

<strong>and</strong> glue for gilding on wood (4) . Cennini clearly distinguishes between gesso<br />

grosso <strong>and</strong> gesso sotille; that is, a double-structured ground consisting of several<br />

layers of a coarse ground on top of which are applied several layers of a finer<br />

ground. In both structures, the medium is animal glue (5). Such grounds, with<br />

local variations, were found in fourteenth- <strong>and</strong> fifteenth-century paintings<br />

from Florence <strong>and</strong> Siena (6) .<br />

Lindberg's translation, here translated from Swedish into English by the author<br />

(7), differs in several crucial passages from the 1933 English translation by D.<br />

V Thompson. The passage concerned is the following: The Italian text says,<br />

"poi abbi giesso grosso cioe uolteriano che e purghato ede tamigiato amodo<br />

di farina, ..." which Thompson translates as, "then take some gesso grosso,<br />

58<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


that is, plaster of Paris, which has been purified <strong>and</strong> sifted like flour" (8, 9).<br />

The Swedish translation says: "Tag sedan grovgips, det vill saga Volterra-gips,<br />

som er renad och siktad som mj ol. ..." [Then take coarse gypsum, that is to<br />

say Volterra gypsum that has been purified <strong>and</strong> sifted like flour. ... J<br />

The question is, apart from the obvious fact that the gypsum mentioned was<br />

mined in the quarries at Volterra, what was then understood by this statement?<br />

Initial attempts by the author at reconstructions were based on<br />

Thompson's translation. The use of his "plaster of Paris," a gypsum burned<br />

to the hemihydrate fo rm, resulted in immediate setting in water (10).<br />

The next chapter, which describes the preparation of the gesso sottile, reveals<br />

details about the gesso volteriano not mentioned in Chapter CXV Chapter<br />

CXVI says, "Ora si uuole chettu abbi dun giesso elquale sichiama giesso<br />

sottile elquale e diques to medesimo giesso mae purghato perbene unmese<br />

tenuto in molle innun mastello rinuoua ogni di laqua chesquasi siinarsiscie<br />

edesciene fuori ogni fo chor di fuocho e uiene morbido chome seta ...." (11).<br />

In Lindberg's translation: "Nu will man att du skall ha en gips som man kallar<br />

fingips, vilken besrar av denna samma gips, men den ar renad i gott och val<br />

en manad, lagd i blot i en balja. Byt varje dag vattnet tills den ar nastan torr,<br />

och varje glad av eld gar ut den, och den blir mj uk som silke ...." [Now<br />

you should have a gesso, which is called fine gesso, which is this same gesso,<br />

but it has been purified for a good month, soaked in a basin. Change the<br />

water every day until it is almost dry, <strong>and</strong> every glow of fire leaves it, <strong>and</strong> it<br />

will be soft as silk ....<br />

J<br />

There are two things to be noticed here: Lindberg argues that the remark<br />

about the glow of fire leaving the gypsum-the same gypsum that was used<br />

for the gesso grosso-can only be understood in the sense that the gypsum<br />

was indeed burned. The question is, what form resulted from the process?<br />

The remark about the soaking in water constitutes another important point:<br />

What kind of procedure is meant? The existing translations are not very clear<br />

about this point. It seems quite conceivable that a double purpose-a washing<br />

process <strong>and</strong> a process of changing the morphology <strong>and</strong> chemical composition<br />

of the material-was served by this treatment.<br />

The preparation of gesso grounds<br />

The naturally occurring gypsum, calcium sulfate dihydrate CaS04'2H20, can<br />

be burned at various temperatures. Burning at 128 °C produces CaS04'H20,<br />

the hemihydrate form. Burning at 130-160 °c creates an anhydrite <strong>and</strong> hemihydrate<br />

mixture (12). This is the so-called plaster of Paris or stucco plaster,<br />

which sets quickly with water <strong>and</strong> thus returns to the dihydrate form. Between<br />

163 °C <strong>and</strong> 300 DC, soluble anhydrite, CaS04, is formed, which also<br />

reacts quickly with water. According to Mora, et aI., the dihydrate form will,<br />

at temperatures above 250 °C, turn into insoluble anhydrite that is no longer<br />

able to set with water (13). Experiments at the School of Conservation have<br />

shown, however, that anhydrite burned at 300 °C, 400 DC, <strong>and</strong> even 500 °C<br />

is still able to react with water, forming dihydrate again (Figs. 1 , 2) (14). The<br />

Merck Index even gives 650 °C as the limit above which the insoluble anhydrite<br />

is formed (15). Above 900 DC, the so-called Estrich gypsum (a combination<br />

of anhydrite <strong>and</strong> calcium oxide) is fo rmed; this material sets very<br />

slowly with water <strong>and</strong> becomes extremely hard. Gypsum can also occur in<br />

nature as insoluble anhydrite, CaS04.<br />

In Cennini's time, the mined gypsum was burned in rather primitive kilns.<br />

Gettens describes a very ancient kiln with little temperature control, in which<br />

blocks of gypsum are stacked <strong>and</strong> a fire lighted at the base of the kiln (16).<br />

The result of this process must have been a mixture of anhydrite, hemihydrate,<br />

<strong>and</strong> even dihydrate forms. Overburned or dead-burned insoluble anhydrite<br />

lumps must have been prevalent at the bottom of the stack near the source<br />

of heat. In the upper tiers, the lumps burned at lower temperatures, hemihydrate/<br />

anhydrite forms must have been present in larger relative amounts.<br />

Federspiel 59


70<br />

60<br />

50<br />

40<br />

a 3500 3000<br />

Transmittance I Wavenumber (cm-1)<br />

2500 2000 1500 1000 500<br />

80<br />

b 3500 3000<br />

Transmittance I Wavenumber (em-1)<br />

2500 2000 1500 1000 500<br />

60<br />

40<br />

2<br />

c 3500<br />

Transmittance I<br />

Wavenumber<br />

II<br />

3000<br />

(em-1)<br />

2500 2000 1500 1000 500<br />

Figure 1. Fourier transform infrared spectra of calcium sulfa te Lilith d!fJerent amounts if crystal water:<br />

(a) anhydrite; (b) hemihydrite, dihydrate baked at 135 °Cfor two hOLm; (c) dihydrite.<br />

It is even possible, as suggested by Gettens <strong>and</strong> Mrose, that if the lumps were<br />

fairly large, that the outside would have been burned to anhydrite while the<br />

inside would be only partially dehydrated (17). Some of the inside of the<br />

gypsum lumps in the primitive kiln stack could have been heated below 100<br />

°C so that the gypsum remained hydrated. It is interesting to note here that<br />

Theophilus gives the following instructions: " ... take some gypsum burned<br />

in the fashion of lime ...." which must be understood to have been burned<br />

at very high temperatures, that is, above 900 °C (18). The resulting substance<br />

will be quicklime (calcium oxide), which is slaked in water <strong>and</strong> transformed<br />

into lime (calcium hydroxide).<br />

Gettens describes another kind of kiln that worked in the same way as ancient<br />

bread-baking ovens; after a fire had preheated the inside of the brick oven<br />

until the interior was red hot, the fire was extinguished, the coals removed,<br />

<strong>and</strong> the bread-or, in this case, lumps of gypsum-were placed in the oven<br />

overnight. This kind of operation allowed for a more uniform heating. This<br />

60<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


100<br />

80<br />

60<br />

40<br />

a 3500 30.00 2500 2000 1500 1000 500<br />

Transmittance J Wavenumber (cm-1)<br />

20<br />

b 3500 3000 2500 2000 1500 1000 500<br />

Transmittance I Wavenumber (cm-1)<br />

70<br />

60<br />

50<br />

20<br />

10<br />

c 3500<br />

Transmittance I<br />

Wavenumber<br />

3000 2500 2000 1500 1000 500<br />

(cm-1)<br />

Figure 2. Fourier traniform irifrared spectra of gypsum bumed at 500 ° C: (a) anhydrite sample before<br />

treatment with water; (b) slurred in water for two days, the sample still consists of anhydrite; (c) slurred<br />

with water for ten days, the sample now consists of dihydrite.<br />

kind of oven had been in use in the Mediterranean civilizations since early<br />

Roman times (19). The painter's manual of Mount Athos actually describes<br />

the baking of gypsum in a similar type of kiln (20). This manual suggests that<br />

the product from the burning process would be soluble anhydrite, which,<br />

soaked in water, would form the dihydrate product again.<br />

In Cennini's directions for the preparation of the gesso sottile, the burnt<br />

gypsum should be soaked in water for about a month. Lindberg states that<br />

Thompson's translation of the word rinuoua as "stir up," referring to stirring<br />

the water every day, is not correct. Thompson may have mistaken the "n"<br />

for "m" (21). Rinuoua literally means renew or change. Lindberg further<br />

argues that siinarsiscie means "to dry," dismissing Thompson's translation of<br />

"rots away" as incorrect. Lindberg offers the following interpretation: the<br />

water should be poured off the soaked gesso every day until the gesso is<br />

almost dry, which simply means that as much water as possible should be<br />

Federspiel 61


poured off each time; such a process is actually a washing process. Watersoluble<br />

impurities, such as salts, possibly present in the gypsum would be<br />

washed away in this process. Such impurities might cause discoloration of the<br />

ground or efflorescence of salts from it. Apart from this, soaking the burned<br />

gypsum in water obviously also had the function of changing the texture <strong>and</strong><br />

the chemical composition of the material, supposing the point of departure<br />

is soluble anhydrite.<br />

Technical evaluation <strong>and</strong> comparison of results<br />

The recent examination of grounds in Italian paintings by the Laboratoire<br />

de Recherches des Musees de France elaborates on the double structure of<br />

the Italian gesso grounds (22). This double structure was also shown in the<br />

examinations by the National Gallery's laboratories in London (23). Unfortunately,<br />

this point was not addressed in the otherwise excellent 1954 examination<br />

by Gettens <strong>and</strong> Mrose, which makes their results of somewhat<br />

limited value in this context.<br />

Concluding from the results of examinations of grounds, the gesso grosso<br />

consists of mainly anhydrite, sometimes with dihydrate present. I am here<br />

referring to the French examination, which even states the ratio of anhydrite<br />

to dihydrate (24). The numerous Florentine <strong>and</strong> Sienese examples show the<br />

following compositions for the gesso grosso: 100% anhydrite or 75:25 anhydrite:dihydrate<br />

or 50:50 anhydrite:dihydrate. For the gesso sottile, two ratios:<br />

100% dihydrate or dihydrate containing 25% anhydrite.<br />

Considering the chemical changes of soluble anhydrite in contact with water<br />

(the binding medium of the ground is animal glue, which always contains a<br />

certain amount of water), it seems puzzling that the gypsum in the layers of<br />

gesso grosso kept its anhydrous fo rm. A possible explanation could be that<br />

the conversion process became slower as the burning temperature increased<br />

(25). As the anhydrite did not change into dihydrate, the evaporation time of<br />

the water in the binding medium must have been shorter than the hydration<br />

time of the anhydrite. It would be extremely improbable that the gesso grosso<br />

material would be a stable anhydrite, which is not able to react with water,<br />

or a dead-burned anhydrite, which is neither form, because the examples<br />

concerned show the double-structured grounds of gesso grosso <strong>and</strong> gesso<br />

sottile. It would be very unlikely that two different kinds of gypsum would<br />

be employed in the process. The analyses of the sottile layers show dihydrate,<br />

which is the soluble anhydrite soaked in water. It must be assumed that the<br />

point of departure in both the grosso <strong>and</strong> the sottile is the same compound:<br />

soluble anhydrite. The presence in the gesso grosso layers of a mixture of<br />

anhydrite <strong>and</strong> dihydrate would be explained by the sometimes poorly controlled<br />

burning process.<br />

In the samples from Umbria, Latium, the Marches, Venice, <strong>and</strong> Ferrara showing<br />

single-structured grounds, the dihydrate present (100%) is claimed to be<br />

the natural unburned dihydrate (26). The raw gypsum (calcium sulfate dihydrate)<br />

does not differ either in chemical composition or crystal structure<br />

from a dihydrate that has gone through the process applied to a gesso sottile<br />

of burning <strong>and</strong> soaking in water (27).<br />

The question remains, is it possible that the dihydrate present in the examples<br />

mentioned above could be the processed dihydrate? The author's reconstructions<br />

indicate that the raw calcium sulfate dihydrate can only with great<br />

difficulty be triturated to a degree that will make it usable as a ground. Even<br />

then, the resulting surface will not facilitate a satisfactory base fo r gilding. It<br />

must be admitted, however, that raw gypsum can vary considerably with<br />

regard to texture (28). Considering this, we can exclude the possibility that<br />

raw gypsum could have been pulverized to yield a satisfactory product. It<br />

seems likely that in most cases the process would be greatly facilitated by first<br />

burning the calcium sulfate to an anhydrite, followed by grinding <strong>and</strong> processing<br />

with a final soak in water to produce the dihydrate form. This is the<br />

62<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


process described by Cennini for gesso sottile as furnishing the best ground<br />

for gilding (29). In fact, Cennini describes in Chapter CXVIII how certain<br />

panels can be grounded only with gesso sottile.<br />

Conclusion<br />

Finally, returning to the meaning of the giesso volteriano in Cennini's text,<br />

we presume that the material must have been calcium sulfate dihydrate<br />

burned at 300-650 °c, thus forming soluble anhydrite, which was used fo r<br />

the gesso grosso. The preparation of the gesso sottile would be a process of<br />

changing the chemical composition, altering the texture, <strong>and</strong> washing.<br />

The author does not agree with the point of view put forward by E. Martin,<br />

et al. that Cennini did not know whether the giesso volteriano he discussed<br />

was unburned gypsum directly from the quarries of Volterra or if it was<br />

processed by being burned (30). Cennini must have had a good underst<strong>and</strong>ing<br />

of these materials. That we do not underst<strong>and</strong> the exact meaning of his words<br />

600 years later is another matter. In his own time <strong>and</strong> geographical sphere, it<br />

was most likely very well understood, at least by other craftspersons.<br />

Acknowledgments<br />

It was extremely difficult to obtain raw calcium sulfate dihydrate in Denmark. I am<br />

very much indebted to the following persons who made the practical experiments<br />

possible: V Meyer (Building <strong>Materials</strong>, Kalkbr;enderihavnsgade 20, 2100 Copenhagen),<br />

who readily took the trouble to import unburned gypsum for my research<br />

purposes, <strong>and</strong> to Niels Erik Jensen (CTO, Aiborg Portl<strong>and</strong>, 9220 Aiborg 0st), who<br />

demonstrated a keen interest <strong>and</strong> readiness to help with this project by providing me<br />

with raw gypsum in various forms. I wish to thank my colleague, chemist Mads Chr.<br />

Christensen (School of Conservation, Copenhagen) fo r his help in carrying out the<br />

analyses, <strong>and</strong> for fruitful discussions. My colleague, geologist Nicoline Kalsbeek,<br />

(School of Conservation), has kindly furnished me with information from the Cambridge<br />

Structural Database. Finally, I wish to thank Professor Bo Ossian Lindberg<br />

(Institutionen fOr Konstvetenskap, University of Lund, Sweden) for reading <strong>and</strong> commenting<br />

on the text.<br />

Notes<br />

1. Cennini, C. 1960. The Craftsman's H<strong>and</strong>book: II libro de/l'arte. Translated by D. V<br />

Thompson, Jr.. New Yo rk, Dover Reprint (Yale University Press, 1933).<br />

2. Cennini, C. Boken om nullarkonsten. Oversiittning och notkommentar av Ossian Lindberg.<br />

Unpublished manuscript provided by Professor Lindberg. Part of the text<br />

has been published in Lindberg, B. 0. 1991. Antologi om milleriteknik. Lund: Institutionen<br />

fo r Konstvetenskap, Lunds Universitet. As shown by Lindberg, the<br />

Codex Laurentianus (Florence, Biblioteca Laurentiana, MS 78 P 23) must be considered<br />

closest to the lost original manuscript. The Codex Riccardianus (Florence,<br />

Biblioteca Riccardiana, MS 2190), however, is our only source for several passages<br />

which are missing in the Laurentianus text. Therefore, Lindberg based his edition<br />

on both the Laurentianus <strong>and</strong> the Riccardianus codices, as Thompson also did.<br />

For the discussion of the surviving transcripts of Cennini's text <strong>and</strong> the various<br />

editions <strong>and</strong> translations, see Lindberg, B. 0. 1991. Cennino Cennini: den obnde<br />

masteren. In Miliningens anatomi. Exhibition catalogue. Kuituren, Lund, 39-48.<br />

3. Oddy, W A. 1981. Gilding through the ages: An outline history of the process<br />

in the Old World. Gold Bulletin <strong>and</strong> Gold Patent Digest (14): 75-79, in particular<br />

page 77. See also, Gilded wood: conservation <strong>and</strong> history. 1991. Gilding Conservation<br />

Symposium, Philadelphia Museum oj Art. Madison, Sound View Press, 34.<br />

4. Berger, E. 1973. Quellen und Technik der Fresko-, Oel-und Temperamalerei des Mittelalters.<br />

Liechtenstein S<strong>and</strong>ig reprint (Munich 1912), 13.<br />

5. The medieval sources mentioning the preparation of grounds also mention (animal)<br />

glue as the medium. The only exception, to my knowledge, is the fifteenthcentury<br />

Bolognese manuscript, which mentions a ground that apparently has no<br />

medium but depends on the setting with water of the hernihydrite. See Merrifield,<br />

M. P 1849. Segreti per colori, Original treatises on the arts oj painting,<br />

Vol. II. London, John Murray, 595.<br />

6. Martin, E., N. Sonoda, <strong>and</strong> A. R. Duval. 1992. Contribution a l'etude des preparations<br />

blanches des tableaux italiens sur bois. Studies in Conservation (37): 82-<br />

Federspiel 63


92. The fifty paintings examined by the authors reveal a clear trend. <strong>Painting</strong>s<br />

belonging to the Florentine <strong>and</strong> Sienese schools showed double-structured<br />

grounds of gesso grosso <strong>and</strong> gesso sottile, identified respectively as anhydrite <strong>and</strong><br />

dihydrate. Venetian examples showed a tendency towards single-structured<br />

grounds consisting only of dihydrate, as also observed in Gettens, R. E. <strong>and</strong> M.<br />

E. Mrose. 1954. Calcium sulphate minerals in the grounds of Italian paintings.<br />

Studies in Conservation (1):174-90. See in particular pages 180-83. The recent<br />

French examination finds that the examined paintings representing schools outside<br />

Tuscany are not numerous enough in their material to conclude anything<br />

general about the treatment of grounds in the rest of Italy.<br />

7. See note 2.<br />

8. The wording of the Codex Laurentianus is quoted from Lindberg, B. 0. 1990.<br />

Feta och magra limmer enligt Cennino Cennini. In Meddelelser om Konservering,<br />

165-87. Lindberg notes that the Codex Riccardianus in this passage has a slightly<br />

different, but not clearer wording than the Laurentianus. Lindberg points to the<br />

fact that the manuscript lacks interior punctuation.<br />

9. Cennini, 1960. Op. cit., 70.<br />

10. Unfortunately, this suggestion by Thompson, which in the author's opinion is<br />

wrong, is apparently reflected in numerous texts on medieval grounds. See, for<br />

instance, Straub, R. E. 1984. Tafel-und Tiichleinmalerei des Mittelalters. In Reclams<br />

H<strong>and</strong>buch der kiinstlerischen Techniken, Vol. I. Stuttgart Philipp Reclam, 156-<br />

57.<br />

11. See note 8.<br />

12. The various sources on gypsum are obviously not identical in their information<br />

on the relationship between the chemical composition of the gypsum <strong>and</strong> burning<br />

temperatures. Rompps Chemielexicon. 1979. Stuttgart, 575-76. See also, Mora,<br />

L., P. Mora, <strong>and</strong> P. Philippot. 1984. The Conservation of Wallpaintings. London,<br />

Butterworth, 39-47. Also see The Merck Index, 9th ed. 1976, 1707.<br />

13. Mora, et aI., op. cit. (note 12), 42.<br />

14. Lindberg, B. 0. , <strong>and</strong> B. Skans, in their experiments carried out at Institutionen<br />

fo r Konstvetenskap i Lund, have also observed that gypsum burned at 300 °C<br />

fo r two hours is still able to react with water. Lindberg, B. 0. , <strong>and</strong> B. Skans. 1990.<br />

Feta och magra limmar enligt Cennino. In Meddelelser om konservering, 184.<br />

15. See note 12.<br />

16. Gettens, R. J. 1954. A visit to an ancient gypsum quarry in Tuscany. Studies in<br />

Conservation (1):4, 190-92.<br />

17. Gettens <strong>and</strong> Mrose, 1954. Op. cit., (note 6), 185.<br />

18. Theophilus on Divers Arts. 1979. Translated by G. Hawthorne <strong>and</strong> C. S. Smith.<br />

New York, Dover Reprints, 27.<br />

19. A History of Technology, Vol. II. 1957. Oxford, The Clarendon Press, 118.<br />

20. The Painter's Manual of Dionysius of Fourna. 1978. Translated by P. Hetherington.<br />

London, 6. It appears from the directions in the Athos book, that right from the<br />

mining, it is important to choose the right kind of gypsum: " . .. see that you<br />

use only what is white <strong>and</strong> glistening."<br />

21. Lindberg, 1990. Op. cit. (note 8), 184.<br />

22. See note 6.<br />

23. Bomford, D., J. Dunkerton, D. Gordon, A. Roy, <strong>and</strong> J. Kirby. 1990. Art in the<br />

Making. London, National Gallery. See in particular pages 17-19.<br />

24. Martin, Sonoda, <strong>and</strong> Duval, 1992. Op. cit., 86-89, Table 2.<br />

25. It must, of course, be below 650 °C, which is the limit beyond which the anhydrite<br />

is no longer able to react with water.<br />

26. Martin, Sonoda, <strong>and</strong> Duval, 1992. Op. cit., 90. See also Dunkerton, J., <strong>and</strong> A.<br />

Roy. 1986. The Technique <strong>and</strong> Restoration on Cima's The Incredulity if S. Thomas.<br />

National Gallery Technical Bulletin (10) :4-27, especially page 5. See also Dunkerton,<br />

J., A. Roy, <strong>and</strong> A. Smith, A. 1987. The unmasking of Tura's "allegorical<br />

figure": A painting <strong>and</strong> its concealed image. National Gallery Technical Bulletin<br />

(11):5-35, especially 19.<br />

27. If a different crystal structure had existed fo r the dihydrate it would most likely<br />

have been included in the CSD (Cambridge Structural Database). This is not<br />

the case.<br />

28. It was quite obvious that not only was the burned gypsum (the anhydrite) much<br />

easier to grind than the raw gypsum, but the anhydrite burned at 700 °C was<br />

much easier to grind than the anhydrite burned at 400 0c.<br />

29. Cennini, 1960. Op. cit., 73.<br />

30. Martin, Sonoda, <strong>and</strong> Duval, 1992. Op. cit., 90<br />

64<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

The frequent use of verdigris in<br />

paintings from the early Gothic<br />

period to the eighteenth century is<br />

reflected by the citation of the pigment<br />

in treatises. The many warnings<br />

against the use of verdigris <strong>and</strong><br />

precautions recommended to prevent<br />

it from discoloring are set against the<br />

fact that verdigris <strong>and</strong> copper resinate<br />

have survived in many cases as an<br />

intensely green color. It is suggested<br />

that the method of preparation, dissolving<br />

verdigris in warm oil, <strong>and</strong> its<br />

application with a piece of canvas<br />

contribute to the stability of the<br />

green glaze. Colored underpainting<br />

<strong>and</strong> yellow glazes on top of the<br />

green layer were used to modify the<br />

green tone.<br />

Aspects of <strong>Painting</strong> Technique in the Use of Verdigris<br />

<strong>and</strong> Copper Resinate<br />

Renate Woudhuysen-Keller<br />

Hamilton Kerr Institute<br />

Fitzwilliam Museum<br />

University of Cambridge, Whittlesford<br />

Cambridge CB2 4NE<br />

United Kingdom<br />

Introduction<br />

Verdigris <strong>and</strong> copper resinate have been fo und in paintings dating from medieval<br />

times to well into the eighteenth century. Among the green substances<br />

available for painting, verdigris had the most intense color but little covering<br />

power compared with malachite. Verdigris was made by exposing strips of<br />

copper to vinegar vapor as described by Theophilus at the beginning of the<br />

twelfth century (1). Many later treatises <strong>and</strong> artists' h<strong>and</strong>books contain recipes<br />

on how to make verdigris, how to distill it by dissolving the material in<br />

vinegar <strong>and</strong> allowing it to recrystallize, <strong>and</strong> how to turn it into the transparent<br />

green color that today is called copper resinate.<br />

The use of verdigris<br />

Verdigris was used ground in oil, in cherry gum, <strong>and</strong> also in egg, but it had<br />

a bad reputation for turning brown or black. Cennino Cennini, at the end<br />

of the fourteenth century, wrote that "it is nice to the eye but it does not<br />

last," <strong>and</strong> in the fifteenth-century Strasburg Manuscript, a note states that<br />

verdigris does not agree with orpiment (2, 3) . Leonardo da Vinci wrote at<br />

the end of the fifteenth century that the beauty of verdigris "vanishes into<br />

thin air if it is not varnished immediately" (4) . The warnings become even<br />

clearer in the Brussels Manuscript, written by Pierre Lebrun in 1635: "Verdigris<br />

is added to charcoal black, or lamp black, to make these colours dry,<br />

but it is used only with the shadows, for it is a poison in painting, <strong>and</strong> it kills<br />

all the colours with which it is mixed" (5). At the same time, between 1620<br />

<strong>and</strong> 1640, De Mayerne in London wrote in his notebook (6):<br />

Le verd de gris} which is only used Jo r glazing} is an enemy oj all colours}<br />

so much so} that it kills them all} especially azurite. Even if you work<br />

with a brush that has been cleaned with oil that has come into contact with<br />

verdigris before} as clean as it may seem} or if you put the colours on a<br />

palette on which there has been verdigris before} it spoils everything. In this<br />

way} he who wishes to work with verdigris has to keep brushes} palette<br />

<strong>and</strong> oil Jor cleaning separate.<br />

In 1757 Pernety wrote disdainfully, "Ver de gris: this is poisonous for all<br />

animals as well as for colours; if one wants to make use of it in oil painting,<br />

one has to use it on its own or at most mixed with black. It ruins all the<br />

colours, <strong>and</strong> even if there was only a little bit in the grain of the canvas it<br />

would spoil all the colours that were laid on it in the process of painting"<br />

(7).<br />

Chemical analysis, however, has shown that the beautiful greens <strong>and</strong> green<br />

glazes that have survived in paintings are indeed very often verdigris <strong>and</strong><br />

copper resinate. Numerous reports published in the National Gallery Technical<br />

Bulletin <strong>and</strong> in the Bulletin de I } Institut Royal du Patrimoine Artistique bear witness<br />

to this fact. In his investigation of verdigris <strong>and</strong> copper resinate, Hermann<br />

Kuhn came to the conclusion that "experiments with paint specimens <strong>and</strong><br />

observations on paintings ... show that the properties of verdigris are not<br />

nearly as harmful as suggested by the literature" (8).<br />

Woudhuysen-Keller 65


Why have some green paint layers of verdigris <strong>and</strong> copper resinate turned<br />

brown while others remained intact? After a search through the treatises for<br />

instructions on how to use this dangerous green pigment <strong>and</strong> a comparison<br />

of the written instructions with the actual methods of applying the pigment<br />

on the paintings, results seemed to suggest that three interesting factors determine<br />

appearance: (1) verdigris has to be thoroughly embedded in oil or<br />

an oil-resin varnish to be protected from air <strong>and</strong> humidity; (2) colored underlayers<br />

contribute to the beauty of the green glaze; <strong>and</strong> (3) admixtures of<br />

yellow lakes or yellow glazes on top of the transparent green layer were<br />

applied to soften the sharpness of the bluish-green tone of verdigris <strong>and</strong><br />

copper resinate.<br />

The instructions for the use of verdigris emphasize that it should be thoroughly<br />

incorporated into the medium <strong>and</strong> covered with varnish as soon as<br />

possible. The early use of verdigris for pictura transludda on gold leaf <strong>and</strong> tinfoil<br />

described by Theophilus, Cennini, <strong>and</strong> in the Tegernsee Manuscript (ca.<br />

1500) points to this fact (9, 10, 11). The Strasburg Manuscript simply states<br />

that all colors should be ground in oil. It proceeds to explain how to mix<br />

colors <strong>and</strong> how to achieve good results by painting in several layers, the<br />

painting technique generally found in fifteenth-century paintings (12).<br />

In On the true precepts if the art of painting (1587), Armenini gives very detailed<br />

instructions on painting technique, particularly on how to paint green drapery<br />

(13):<br />

if the drape is to be green, one does as follows: After the sketch is made<br />

using somewhat coarse green, black <strong>and</strong> white, it is lightly painted with a<br />

mixture if verdigris, a little common varnish, <strong>and</strong> some giallo santo. With<br />

a coarse brush of miniver, one veils the sketch uniformly; next one pats it<br />

either with the palm if the h<strong>and</strong> or with a little wad if cotton wool covered<br />

with linen, unta the given colour is uniform <strong>and</strong> no brush strokes can be<br />

detected. And if the result is not to one's satisfaction, after the veiling is<br />

dry one repaints with the same mixture <strong>and</strong> then pats in the prescribed<br />

way.<br />

He also explains that before veiling, the thoroughly dried sketch has to be<br />

oiled out very thinly to stop the glaze from being repelled by the underlayer.<br />

A painting by Palmezzano depicting the Mystic Marriage if Saint Catherine,<br />

signed <strong>and</strong> dated 1537, shows exactly this technique in the green drapery of<br />

the throne <strong>and</strong> St. Catherine's green garment (Plate 9). The pattern of the<br />

textile used for dabbing on the glaze is clearly visible (Plate 10). The buildup<br />

shown in the cross section also corresponds to Armenini's instructions<br />

(Plate 11); even the oiling-out layer is visible in ultraviolet light (Plate 12).<br />

The glaze was apparently too viscous to be spread out evenly with a brush,<br />

therefore the glaze was spread by dabbing it with a rag. Traces of textile<br />

pattern are also visible in a Flemish altarpiece, painted in Antwerp around<br />

1520, now at Oxburgh Hall in Norfolk. In the process of dabbing on the<br />

green glaze the artist could not always keep within the outline of his green<br />

drapery, so he had to retouch the background in some places. Minor overlaps<br />

were simply left; one can see the weave pattern of the rag, the glaze is partly<br />

discolored <strong>and</strong> some green particles are still visible.<br />

Experimentation<br />

As an experiment, some neutral recrystallized verdigris was ground in linseed<br />

oil <strong>and</strong> the mixture was heated very gently to approximately 50 °C, until the<br />

copper acetate had dissolved <strong>and</strong> the pigment grain had disappeared. The<br />

mixture was intensely green <strong>and</strong> quite viscous. It could be spread with a<br />

brush while warm, but congealed very quickly, making the brush strokes very<br />

coarse <strong>and</strong> imprecise. However, it was quite easy to spread the glaze by dabbing<br />

it on with a piece of canvas. After a few hours, the glaze dried. Within<br />

a few days the thick brush strokes of green glaze showed drying wrinkles,<br />

whereas the dabbed area looked the same as when it was applied, demon-<br />

66<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


strating that spreading this glaze thinly in several layers is not merely a matter<br />

of style but also a technical necessity.<br />

The green tablecloth in Jan Davidsz de Heem's Still Life with Fruit (Cambridge,<br />

Fitzwilliam Museum), painted in 1650, consists of a green glaze over<br />

a dark brown underlayer with yellow-green highlights. Clearly this green<br />

glaze has been applied with a brush, as the fairly clumsy brush marks are<br />

visible. Along the contours of the vine leaves <strong>and</strong> on either side of the thin<br />

stalks of the cherries, the green glaze leaves a gap showing the dark brown<br />

underlayer. It is quite clear that this glaze must have been a rather viscous<br />

liquid <strong>and</strong> therefore difficult to paint out with any precision.<br />

The recipes for making copper resinate, collected by the doctor Theodore<br />

Turquet de Mayerne in London between 1620 <strong>and</strong> 1640, are generally considered<br />

to be the earliest. They call for verdigris to be heated with Venetian<br />

turpentine <strong>and</strong> oil of turpentine as follows: "Beautiful green: take 2 ounces<br />

of Venetian turpentine, 1 112 ounce of oil of turpentine, mix <strong>and</strong> add 2<br />

ounces of verdigris in little pieces. Set it on hot ashes <strong>and</strong> let it boil gently.<br />

Try it on some glass to see if you like the colour; strain it through a cloth"<br />

(14) .<br />

Trying out the recipe, it was found that the verdigris did not dissolve in the<br />

mixture of Venetian turpentine <strong>and</strong> oil of turpentine because there was not<br />

enough resin present; also, the presence of oil of turpentine hampered the<br />

reaction of the copper acetate with the Venetian turpentine. When more resin<br />

in the fo rm of rosin was added, a dark green resinous substance resulted,<br />

which was liquid while hot, but hard <strong>and</strong> glassy as it cooled. This green glassy<br />

substance can be ground in oil like a pigment. When ground in oil, however,<br />

the color is no longer very intense.<br />

Another recipe in the De Mayerne manuscript asks for verdigris, ground in<br />

oil, to which hot common varnish is added: "Painters, i.e. those who paint<br />

as well as those who paint furniture <strong>and</strong> blinds, grind verdigris with linseed<br />

oil <strong>and</strong> then add common varnish, stirring it well. They allow the impurities<br />

to sink down <strong>and</strong> only use the clear liquid, which they apply warm" (15).<br />

With the term "common varnish" a solution of resins in oil is meant. This<br />

recipe was the base for the experiment described above, grinding verdigris<br />

in oil <strong>and</strong> heating it.<br />

The preparation <strong>and</strong> application of the green glaze is not the only secret of<br />

de Heem's tablecloth: the dark reddish-brown undermodeling gives the green<br />

glaze its velvety depth. This observation proved to be very useful during<br />

retouching. The only way to match this intense dark green color was to<br />

reconstruct the build-up of layers exactly. The verdigris in the green glaze<br />

was substituted with the transparent green pigment viridian (because of its<br />

stability) <strong>and</strong> a little synthetic Indian yellow to match the required tone.<br />

The reason a green glaze over a reddish-brown underlayer appears so very<br />

dark lies in the absorption of the waves of the spectrum: green absorbs all<br />

red waves, red absorbs all green waves. The two layers superimposed absorb<br />

practically the entire spectrum of visible light, so that the resulting color is<br />

almost black.<br />

The painter's use of green glazes <strong>and</strong> grisaille<br />

The green curtain in Titian's Tarquin <strong>and</strong> Lucretia (Fitzwilliam Museum)<br />

shows green glazes over an undermodeling in red with broad white highlights.<br />

The idea is ingenious. To start with, Titian underpainted the curtain in gray<br />

with some azurite, then he laid in the modeling with a brownish red containing<br />

some red lake <strong>and</strong> some very generous white highlights. As he applied<br />

the green glaze, the shadows in the folds appeared very dark green, the middle<br />

tones were light green because of the green glaze over white brush strokes,<br />

<strong>and</strong> the flickering highlights remained white from the undermodeling, partly<br />

emphasized with an extra brush stroke. In some places, Titian allowed the<br />

red to shine through, giving the material a wonderful shot-silk effect.<br />

Woudhuysen-Keller 67


In late Gothic painting, grisaille underpainting was fairly common for green<br />

garments or l<strong>and</strong>scape elements. This was found in paintings by van Eyck as<br />

well as by Uccello, north of the Alps as well as in Italy (16, 17). The method<br />

is described in artists' treatises. Palomino, in the first quarter of the eighteenth<br />

century, describes the method of underpainting green cloths with grisaille<br />

(18). The same instructions are found in Croeker's treatise, published in 1743<br />

(19). This documents a long tradition. Grisaille undermodeling was not the<br />

only method recommended. In his Portrait if a Dominican Monk (Upton House<br />

near Banbury), painted around 1525, Lorenzo Lotto used undermodeling in<br />

azurite <strong>and</strong> lead white with the darkest shadows painted in an olive green<br />

mixture of black <strong>and</strong> lead-tin yellow. The next layer contains verdigris with<br />

very little lead-tin yellow <strong>and</strong> some black in the shadows, fo llowed by a green<br />

glaze. This green glaze is a substantial, absolutely homogenous layer, without<br />

any visible brushwork or textile pattern.<br />

In addition to colored undermodeling to modifY the green glaze, there is the<br />

possibility of adjusting the green color by means of yellow glazes. Cennini<br />

recommended saffron for this purpose (20). Leonardo suggested aloe dissolved<br />

in warm spirit of wine: "If you have finished a painting with this simple<br />

green [verdigris in oil] <strong>and</strong> if then you were to glaze it lightly with this aloe<br />

dissolved in spirit of wine, then it would be of a most beautiful colour. Also<br />

this aloe can be ground in oil, either on its own or together with the copper<br />

green <strong>and</strong> with any other colour you like" (21).<br />

De Mayerne noted a recipe fo r a green copper glaze that called fo r some<br />

terra merita, which is curcuma, to be added to improve the color (22). Goetghebeur<br />

<strong>and</strong> Kockaert drew attention to the existence of yellow <strong>and</strong> brown<br />

glazes on top of layers of verdigris or copper resinate (23). In cleaning green<br />

areas in paintings, therefore, it is not sufficient to test a brown layer for copper<br />

to find out whether the paint is discolored copper resinate. Even if it is<br />

not, it can still be an original yellow lake, now discolored. As these organic<br />

lakes are very difficult to identifY, their presence is often suspected without<br />

reassuring proof either way, a most disconcerting situation fo r the restorer. In<br />

addition to research by means of scientific analysis, it might be helpful to<br />

approach the problem via the written sources <strong>and</strong> by reconstructing some of<br />

the recipes.<br />

Notes<br />

1. Hawthorne, J. G. <strong>and</strong> C. S. Smith. 1979. Theophilus on Divers Arts. New York:<br />

Dover Publications, 41. See also Scholtka, A. 1992. Theophilus Presbyter: Die maltechnischen<br />

Anweisungen und ihre Gegenuberstellung wit naturwissenscha J tlichen Untersuchungsbifunden.<br />

Zeitschrift jur Kunsttechnologie und Konservierung 6 (1):1-54.<br />

2. Cennini, C. 1960. The Crcifisman's H<strong>and</strong>book. Translated by D. V. Thompson. New<br />

York: Dover Publications, 33.<br />

3. Borradaile, V. <strong>and</strong> R. Borradaile. n.d. The Strasburg manuscript: a medieval painters'<br />

h<strong>and</strong>book. Munchen: Callwey, 55-7. Passage translated by the author.<br />

4. Ludwig, H. 1882. Lionardo da Vinci. Das such von der Malerei. Nach den Codes<br />

Vaticanus (Urbinas) 1270. Wien: Braumiiller (Quellenschriften fLir Kunstgeschichte<br />

und Kunsttechnik des Mittelalters und der Renaissance), 15:244. Passage<br />

translated by the author.<br />

5. Merrifield, M. P. 1967. Original Treatises on the Arts if <strong>Painting</strong>, Vol. 2. New York:<br />

Dover Publications (2):822.<br />

6. van de Graaf,J. A. 1958. Het De Mayeme Manuscript als bron voor de schildertechniek<br />

van de barak. British Museum, Sloane 2052. Proefschrift ... Mijdrecht: Drukkerij<br />

Verweij, 142. Passage translated by the author.<br />

7. Pernety, A. J. 1972. Dictionnaire portatif de peinture, sculpture et gravure. Geneva:<br />

Minkoff Reprint, 547. Passage translated by the author.<br />

8. Kuhn, H. 1993. Verdigris <strong>and</strong> copper resinate. In Artists' Pigments. Ed. A. Roy.<br />

Washington: National Gallery of Art, 136. For preparation <strong>and</strong> identification of<br />

green copper pigments, see Ellwanger-Eckel, E 1979. Herstellung und Verwendung<br />

kunstlicher gruner und blauer Kupjerpigmente in der Malerei. Stuttgart: Institut fu r<br />

Museumskunde an der Staatlichen Akademie der Bildenden Kunste.<br />

9. Hawthorne <strong>and</strong> Smith, op. cit., 33.<br />

68<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


10. Thompson, op. cit., 61.<br />

11. Berger, E. 1975. Quellen und Technik der Fresco-, Oel-und Tempera-Malerei des Mittelalters<br />

. . . Walluf-Nendeln: S<strong>and</strong>ig Reprint, 195.<br />

12. Borradaile, op. cit., 55-7.<br />

13. Armenini, G. B. 1977. On the true precepts of the art of painting. Edited <strong>and</strong> translated<br />

by E. J. Olszewski. Burt Franklin, 194.<br />

14. van de Graaf, op. cit., 174. Passage translated by the author.<br />

15. van de Graaf, op. cit., 164. Passage translated by the author.<br />

16. Brinkmann, P W F., et al. 1984. Het Lam Godsretabel van van Eyck . .. Bulletin<br />

de l'Institut Royal du Patrimoine Artistique (20):164.<br />

17. Massing, A. <strong>and</strong> N. Christie. 1988. The Hunt in the Forest by Paolo Uccello. Bulletin<br />

of the Hamilton Kerr Institute (1):36.<br />

18. Veliz, Z. 1986. Artists' <strong>Techniques</strong> in Colden Age of Spain. Cambridge: Cambridge<br />

University Press, 168-69.<br />

19. Croeker, J. M. 1982. Der wohl anfuhrende Mahler. Rottenburg: Kremer-Reprint,<br />

110.<br />

20. Thompson, op. cit., 29-30.<br />

21. Ludwig, op. cit., 244-7. Passage translated by the author. See also Chastel, A. n.d.<br />

Leonardo da Vinci. Samtliche Cemalde und die Schriften zur Malerei. [Darmstadt] :<br />

Wissenschaftliche Buchgesellschaft, 351-52.<br />

22. van de Graaf, op. cit., 174.<br />

23. Goetghebeur, N., <strong>and</strong> L. Kockaert. 1980. Le gr<strong>and</strong> Calvaire d'Albert Bouts au<br />

M usee des Beaux-Arts de Bruxelles. Bulletin de l'Institut Royal du Patrimoine Artistique<br />

(18):5-20.<br />

Woudhuysen-Keller 69


Abstract<br />

The discovery of three "new" pigments<br />

is described. Their history is<br />

traced through the literature of pigments,<br />

ceramics, <strong>and</strong> glass. Tin white,<br />

previously undiscovered, is followed<br />

from Iraqi ceramics to its occurrence<br />

as a pigment on Jain miniatures in<br />

India. Burnt green earth, mentioned<br />

by early Italian <strong>and</strong> nineteenth-century<br />

English writers, is identified on<br />

nineteenth-century oil paintings.<br />

Cobalt oxide colorant is followed<br />

from Persia to Europe <strong>and</strong> China<br />

<strong>and</strong> finally, in a new fo rm, to Venetian<br />

enamels <strong>and</strong> as smalt on Hindu<br />

miniatures.<br />

Connections <strong>and</strong> Coincidences: Three Pigments<br />

Josephine A. Darrah<br />

Science Group<br />

Conservation Department<br />

Victoria & Albert Museum<br />

London SW7 2RL<br />

United Kingdom<br />

Introduction<br />

There are certain advantages in being a generalist rather than a specialist. In<br />

this paper, the author hopes to show some of the connections <strong>and</strong> coincidences<br />

that occur when a wide range of materials <strong>and</strong> objects are analyzed<br />

in the same laboratory.<br />

Tin white<br />

The pigment known as tin white is elusive. It is mentioned occasionally in<br />

medieval <strong>and</strong> later European literature, but has never been identified on paintmgs.<br />

Tin oxide was used first as an opacifier in ceramic glazes to reproduce Chinese<br />

porcelains in ninth-century Iraq. From there it spread throughout the<br />

Near East (by the tenth century) <strong>and</strong> to Spain (by the thirteenth century) via<br />

North Africa (1). By the twelfth century it was being used to opacifY glass.<br />

It has been identified in twelfth-century Byzantine colored mosaic tesserae<br />

in Tchernigov, Ukraine (2). In 1612, in the first book devoted to glassmaking,<br />

Neri describes the preparation of enamel by adding calcined tin to calcined<br />

lead (3). The earliest dated European tin white glass is late fifteenth-century<br />

Venetian (4). White glass <strong>and</strong> enamel of the succeeding centuries is almost<br />

always opacified with tin (5).<br />

Writing in the late thirteenth century, Eraclius gives two very similar recipes<br />

for white glaze used for earthenware (6) . White glass was ground very finely,<br />

mixed with sulfur, painted onto the ceramic <strong>and</strong> fired. Tin is not mentioned<br />

but must have been the opacifier. The purpose of the sulfur is not known. It<br />

would not have survived the firing.<br />

The Paduan manuscript written in Venice in the seventeenth century (but<br />

copying sixteenth-century material), has another similar recipe, but here it is<br />

for a pigment: "Un bianco bellissmo-Si piglia cristallo di Venetia ..." [take<br />

some powdered Venice glass, add to it a third part of powdered sulfur] (7).<br />

The mixture was heated to red heat in a pipkin, cooled, <strong>and</strong> ground. Merrifield<br />

suggests that it was used for painting miniatures (8).<br />

Harley quotes an English source (ca. 1500): "For to make Ceruse. Take plates<br />

of tinne <strong>and</strong> beate them as thinne as thowe maist ..." (9). The tin was hung<br />

in a sealed barrel with vinegar for several weeks. The method is exactly the<br />

same as that used to produce lead or flake white. Indeed, the name ceruse<br />

was applied in Engl<strong>and</strong> to both tin <strong>and</strong> lead white, <strong>and</strong> perhaps more correctly<br />

to a mixture of lead white <strong>and</strong> chalk.<br />

Harley says that documentary sources indicate that tin white was used in<br />

manuscripts. Her suggestion that it became obsolete because manuscript illumination<br />

declined cannot be substantiated, as the pigment has not been<br />

identified in any English or European miniatures. In the seventeenth century,<br />

Van Dyck tried it in oil <strong>and</strong> reported that it had insufficient body <strong>and</strong> was<br />

only useful for manuscript illumination (10). My tens fo und that it blackened<br />

in sunlight, spoiled white lead if the two were mixed, was useless in oil, <strong>and</strong><br />

also in distemper if exposed to air (11). In the late eighteenth century, France<br />

Guyton de Morveau experimented with pigments <strong>and</strong> reported that tin white<br />

was unsuitable as it tended to yellow or blue (12).<br />

70<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Tingry, in 1830, gives a recipe for "Another Cremnitz White," which he<br />

describes as a beautiful pearly white, too expensive for house painting, but<br />

"it would, no doubt, be attended with great advantage in painting pictures"<br />

(13). This white was a mixture of tin white, one-fourth part zinc white <strong>and</strong><br />

one-eighth part white clay separated from Briancon white.<br />

Linton describes tin white (oxide of tin) as, "too feeble in body ... to be of<br />

any service to the oil painter . .. " (14). He does, however, add that it may be<br />

unaffected by "injurious gases." Field also thinks the pigment to be poor,<br />

writing that it "dries badly <strong>and</strong> has almost no body in oil or in water, it is<br />

the basis of the best white in enamel painting" (15). In 1951, Mayer merely<br />

states that it is not a paint pigment at all (16).<br />

Figure 1. Tin white pigment, details on a<br />

sixteenth-century Jain miniature. Photograph<br />

by Paul Robins (Photo <strong>Studio</strong>), courtesy of<br />

the Victoria & Albert Museum (I. S.84-<br />

1963 1 15).<br />

Tin white, therefore, has two main uses: as an opacifier in glass, enamel, <strong>and</strong><br />

ceramic glazes from the ninth to twentieth centuries; <strong>and</strong> as a possible pigment<br />

in manuscripts <strong>and</strong> miniatures from the fifteenth to seventeenth centuries.<br />

Even in the short review given here, three quite different recipes for<br />

its preparation are given from the following eras: (1) circa 1500, Engl<strong>and</strong>, tin<br />

oxide; (2) circa 1580, Italy, white glass powder opacified with tin oxide; <strong>and</strong><br />

(3) circa 1800, Engl<strong>and</strong>, a mixture of tin oxide, zinc oxide, <strong>and</strong> white clay.<br />

During a research program in the laboratory of the Victoria & Albert Museum,<br />

methods for nondestructive identification of the pigments on Indian<br />

miniatures were studied (17). Incident light microscopy, energy-dispersive x­<br />

ray fluorescence (EDXRF) spectroscopy, <strong>and</strong> ultraviolet <strong>and</strong> infrared color<br />

reversal photography were used.<br />

Several Jain miniatures were examined; these small, jewel-like, utterly distinctive<br />

paintings from Western India were of the fifteenth <strong>and</strong> sixteenth centuries.<br />

The areas of white paint were frequently restricted to details of textiles<br />

<strong>and</strong> jewelry (Figs. 1 , 2). Two of the miniatures had tin present as the major<br />

constituent in areas of white pigment. Because red lake <strong>and</strong>, in one case, gold<br />

leaf were beneath the white, this was regarded as interesting but not conclusive<br />

of tin white being present.<br />

Tin was then fo und on a third Jain miniature, possibly dating from the fifteenth<br />

century, with carbon black <strong>and</strong> verdigris in the same area. Three more<br />

paintings were chosen from the same manuscript as one of those examined<br />

earlier, as it was possible to focus on areas with no other pigment or only<br />

gold leaf. In all three, tin was the major constituent (Fig. 3) . It is possible,<br />

Figure 2. Tin white pigment, details on a<br />

sixteenth-century Jain miniature. Photograph<br />

by Paul Robins (Photo <strong>Studio</strong>), courtesy of<br />

the Victoria & Albert Museum (I. S. 46-<br />

19591 26).<br />

500<br />

450<br />

400<br />

350<br />

Sn<br />

300<br />

250<br />

200<br />

Sn<br />

Sn<br />

Au<br />

Au<br />

150<br />

Sn<br />

100<br />

50<br />

Au<br />

5<br />

10<br />

15<br />

20<br />

KeV<br />

25<br />

30<br />

35<br />

40<br />

Figure 3. EDXRF spectrum of tin white pigment on a sixteenth-century Jain miniature. Spectrum<br />

prepared by David Ford, Science Group, Victoria & Albert Museum (1. 5.46-19591 49).<br />

Darrah 71


therefore, to say that the pigment is a tin white (Table 1). A mixed tin-lead<br />

white was identified on a seventeenth-century Hindu miniature; much overpainting<br />

with zinc white confused the picture.<br />

It has not been possible to determine the constitution of the Indian tin white;<br />

the size <strong>and</strong> delicacy of the miniatures prohibits the removal of a sample. The<br />

pigment is a clean, brilliant white <strong>and</strong> appears reasonably opaque. It is probably<br />

tin oxide.<br />

Although the date of these miniatures falls into the same period in which<br />

the pigment is said to have been used in Europe, it seems unlikely that the<br />

Jain school was using the Western pigment. It is probable, however, that it<br />

developed from the use of tin oxide in ceramics, ultimately deriving, as it did<br />

in Europe, from the Near East or neighboring Persia.<br />

Burnt green earth<br />

Table 1. White pigments on jifleenth- to<br />

eighteenth-century Jain miniatures.<br />

Accession No. Da" Pigments<br />

IS 2-1972 c. 1460 Kaolin, mica<br />

Isacco 1<br />

15C Calcium white<br />

Isacco 2 15C Tin while<br />

IS 46-1959 f26 16C Tin white<br />

IS 46-1959 f45 16C Tin white<br />

IS 46- 1959 f47 16C Tin white<br />

IS 46-1959 f49 16C Tin white<br />

IS 84-1963 fl5 16C Tin white<br />

IS 82·1963 16C Hindu Calcium white, mica<br />

Private owner 17C Hindu Tin - lead white<br />

IS 2-1984 18C Lead white<br />

Green earth in its various fo rms has been used as a pigment through much<br />

of Eurasia fo r 2,000 years. Vitruvius wrote of it in the first century B.C.E.<br />

Burnt green earth seems to be mentioned first in the sixteenth or seventeenth<br />

centuries. The Paduan manuscript lists it as a color for miniature painting <strong>and</strong><br />

also records that "the shadows of the flesh are made with terra 'ombra, terra<br />

verde burnt, <strong>and</strong> asphaltum" (18). The Volpato manuscript gives the method<br />

of preparation (19). Merrifield quotes Lomazzo, who directs that shadows on<br />

flesh should be made with burnt terra verde <strong>and</strong> nero di campana or umber<br />

(20). She says that "modern writers do not mention this colour, but the use<br />

of it has been revived by an eminent English artist, under the name of 'Verona<br />

Brown' "(21).<br />

Linton mentions it briefly with terre verte: "When calcined, it forms another<br />

beautiful pigment called Verona Brown" (22).<br />

To ch says Verona brown is a "fancy name" fo r a mixture of burnt umber <strong>and</strong><br />

burnt or raw sienna (23). At the time of the Constable <strong>and</strong> Turner research<br />

projects in Engl<strong>and</strong>, the author examined two paint boxes of the relevant<br />

period in the Victoria & Albert Museum collections. One was said to have<br />

Figure 4. Paint box, said to have belonged to William Turner, with bladders of oil paint, including<br />

Gebr. Terra di Verte <strong>and</strong> Gebr. Griine Erde. W. 65- 1920. Courtesy of the Victoria & Albert MuseU/t!.<br />

72<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


elonged to William Mallord Turner (Fig. 4) . It contained seventeen bladders<br />

of solidified oil paint, all of which had been opened <strong>and</strong> sealed with a tack.<br />

Each was labeled in German or English. Two were named as Gebr. Terra di<br />

Verte <strong>and</strong> Gebr. Grune Erde, the term "Gebr." coming from the German<br />

gebrannte for burnt or roasted (24).<br />

EDXRF analysis of the paint on the surface of the bladders showed iron to<br />

be the major constituent with traces of manganese <strong>and</strong> titanium. The Terra<br />

di Verte contained a trace of calcium, the Griine Erde considerably more<br />

with traces of potassium, rubidium, <strong>and</strong> strontium (frequently present in calcium<br />

deposits). Green earth is a complex silicate colored by iron with a<br />

structure similar to mica (25).<br />

Dispersions of the two paints were made. The Gebr. Terra di Verte was a<br />

rich orange-brown comparable to burnt sienna but more translucent. The<br />

Gebr. Griine Erde was similar to raw sienna, but of a greener tone. This must<br />

have been burned at a lower temperature, as a few green particles remained.<br />

Under the microscope, the pigments were identical to green earth, with the<br />

overlapping plates of the crystals visible on the larger particles. Calcite was<br />

seen in both, but the Gebr. Griine Erde had, as indicated by EDXRF, a higher<br />

proportion of it. The labels suggest that both bladders were prepared in Germany,<br />

perhaps from two different sources of green earth.<br />

The appearance of the samples brought to mind the unidentified brown pigment<br />

seen by the author in several nineteenth-century l<strong>and</strong>scape paintings,<br />

usually mixed with Prussian blue, ochres, <strong>and</strong> so forth, to produce greens <strong>and</strong><br />

browns. It was suggested that it was the brown seen in the cross sections from<br />

a painting by Constable then being examined. SEM-EDX analysis of the latter<br />

at the National Gallery laboratory produced a spectrum identical to that for<br />

green earth. It was later identified in several Constable paintings dating from<br />

1811 to 1829 mixed in greens <strong>and</strong> browns (26, 27). Could Merrifield's "eminent<br />

English artist" be John Constable?<br />

The author has tentatively identified burnt green earth in paintings by Peter<br />

de Wint <strong>and</strong> J. F. Millet (28).<br />

Green earth was rarely used in Engl<strong>and</strong>. The author has seen it only in<br />

seventeenth- <strong>and</strong> eighteenth-century wall paintings (oil) <strong>and</strong> cartoons by Verrio,<br />

Laguerre, Thornhill, <strong>and</strong> Robert Adam, all of whom trained in Italy or<br />

France. Verona brown seems to have been adopted in the early nineteenth<br />

century as a translucent addition to the earth <strong>and</strong> organic browns then available.<br />

Verona was a source of one of the better green earths, but was ab<strong>and</strong>oned<br />

earlier this century (29).<br />

By coincidence, Constable Project researcher Sarah Cove visited Brussels <strong>and</strong><br />

brought the author a bottle of pigment, Griine Erde Gebr., from an artists'<br />

suppliers. This modern sample is a darker, duller brown; perhaps burned at a<br />

higher temperature than the earlier examples, it too contains calcite.<br />

Brown pigments tend to be neglected, partly, no doubt, because of the difficulty<br />

in distinguishing the multitude of ochres, organic earths, <strong>and</strong> lakes.<br />

Burnt terra verte has a quite distinctive appearance, is easily identifiable by<br />

EDX, <strong>and</strong> may be more common than previously thought.<br />

Smalt<br />

The earliest blue glass colored with cobalt is from Eridu, Mesopotamia, circa<br />

2000 B.C.E. Recipes survive from Ashurbanipal's library in Nineveh, circa<br />

650 B.C.E. The Indians adopted Sumerian technology <strong>and</strong> were making cobalt<br />

blue glass by the sixth century B.C.E. (30). The Egyptians were using<br />

cobalt by circa 1400 B.C.E. The Romans were familiar with it; it was common<br />

in Western Europe in the seventh century <strong>and</strong> occurs in the Sassanian <strong>and</strong><br />

Islamic periods (31).<br />

The first appearance of smalt is in a wall painting (ca. 1000-1200 C.E.) in<br />

Khara Khoto, Central Asia, <strong>and</strong> in the Church of Our Saviour of the Mon-<br />

Darrah<br />

73


Table 2. Unusual occurrences <strong>and</strong> constituents of smalt <strong>and</strong> blue enamel.<br />

Source Accession Blue pigment Date<br />

Number<br />

or enamel<br />

Minor<br />

EDXRF Analysis<br />

Trace<br />

BasohJi miniature 1.5.50-1953 smalt 1660-70<br />

1.5.51-1953 small 1660-70<br />

I.M.87-1930 small 1730-35<br />

I.M.88-1930 smalt 1730-35<br />

Bundi miniature 0.379-1889 small c.I770<br />

Venetian jug 273-1874 enamel 1472-1525<br />

tazza 5496-1859 enamel late 15th C.<br />

bowl C.170-1936 enamel 1521-23<br />

bowl C. I60-1936 enamel early 16th C.<br />

bowl 5489-1859 enamel 1500-1600<br />

S. German oil painting C.I.A. No 3021 small 15th - 16th C.<br />

Pb'<br />

As Ph·<br />

As<br />

As<br />

Pb'<br />

As<br />

As<br />

As<br />

As<br />

As<br />

KCa Fe Co Ni Cu As<br />

KCa Mn Fe Co Ni Cu Zn Bi<br />

KCa Fe Co Ni Ph Bi<br />

KCa Fe Co Ni Zn Pb Bi<br />

KCa Ti Fe Co Ni Cu As<br />

KCa Mn Fe Co Ni Cu Zn Pb Bi Sr Sn<br />

KCa Mn Fe Co Ni (Au) Ph Bi Sr<br />

KCa Ti Mn Fe Co Ni Cu Ph Bi Sr<br />

KCa Mn Fe Co Ni Cu Zn Ph As Bi Sr (So Sb)t<br />

KCa Mn Fe Co Ni Cu (Au) Ph Bi Sr Sn Sb<br />

Si K Fe Co Ni Pb. Bi<br />

• lead from lead white pigment<br />

t tin <strong>and</strong> antimony from yellow enamel applied over blue<br />

* Courtauld Institute of Art, SEM-EDX analysis<br />

astery of Chora (Kariye Camii), Constantinople (1325-1453 C.E.) (32, 33). It<br />

is likely that the smalt derived from Middle Eastern <strong>and</strong> Near Eastern blue<br />

glass.<br />

The earliest known source of cobalt ore is Khashan, Persia; the Mesopotamians<br />

<strong>and</strong> Egyptians probably obtained it from there. This suggests that the<br />

ore's coloring properties were already known to the Persians. Medieval European<br />

glass was tinted with Damascus pigment or zaffre, the Arabic name<br />

for cobalt oxide, again suggesting a Near Eastern source (34).<br />

Cobalt oxide, known as sulimani, was imported into China from Persia in the<br />

Tang period; by the fourteenth century it was transported by sea from the<br />

Persian Gulf via Sumatra. Muslim merchants residing in China influenced<br />

ceramic designs; much of the blue <strong>and</strong> white ware was produced fo r the<br />

Islamic market (35).<br />

Cobalt blue glass <strong>and</strong> glaze was known from China to Western Europe <strong>and</strong><br />

yet throughout this period (3000 B.C.E. to ca. 700 C.E.) the most important<br />

blue pigment was Egyptian blue, a fr it colored with copper, at its best rivaling<br />

azurite but often appearing in paler turquoise shades. The terminal date is<br />

circa 850 C.E. on a fresco in the church of San Clemente in Rome (36). It<br />

seems unlikely that the secret of making Egyptian blue was lost to the<br />

Romans during the turmoil of the Teutonic invasions, as glass <strong>and</strong> enamel<br />

colored with copper continued to be made both in Italy <strong>and</strong> many other<br />

countries.<br />

In Europe there is an inexplicable hiatus in the use of blue pigments deriving<br />

from glass from about 850 to 1490 C.E. The Venetians were making cobalt<br />

glass by the mid-fifteenth century; the earliest references to smalt are by Leonardo<br />

da Vinci <strong>and</strong> Perugino in the 1490s (37). It has been identified on<br />

an altarpiece (1493) by Michael Pacher (38). Cobalt was discovered in Saxony<br />

in the mid-fifteenth century <strong>and</strong> fully exploited by around 1520. This may<br />

explain the greater utilization of smalt in the sixteenth century (39, 40).<br />

Cobalt ores combine iron with nickel <strong>and</strong>/or arsenic, the latter volatilized in<br />

the smelting process. Cobalt was also obtained from the residue in the separation<br />

of bismuth (41). The metal oxides were fused with s<strong>and</strong> <strong>and</strong> potash<br />

to produce zaffre (the Arabic name still being used) <strong>and</strong> sold to glassmakers.<br />

During research on Indian miniatures, the author found smalt on one Bundi<br />

example (ca. 1770 C.E., central India, Hindu). The pigment, not previously<br />

identified on Indian paintings, contained cobalt, iron, nickel, <strong>and</strong> a little arsenic<br />

(Table 2) (Plate 13).<br />

In a later project examining miniatures from the northern Hindu states, smalt<br />

was fo und on four miniatures from Basohli, a tiny state in the Himalayan<br />

foothills north of Lahore (42, 43). In one example (1660-1670 C.E.), smalt<br />

was used to paint areas of the sky <strong>and</strong> Krishna's skin; the pigment contained<br />

74<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


500<br />

450<br />

As<br />

Ph<br />

As<br />

400<br />

350<br />

Fe<br />

Co<br />

300<br />

Ni<br />

250<br />

200<br />

150<br />

Bi<br />

Ph<br />

Bi<br />

100<br />

50<br />

0<br />

0 2 4<br />

6<br />

8<br />

10<br />

12<br />

14 16 18 20<br />

KeV<br />

Figure 5. EDXRF spectrum of smalt with a high arsenic content <strong>and</strong> traces if bismuth <strong>and</strong> lead.<br />

Krishna <strong>and</strong> Girls, 1730-1735, Basohli. Spectrum prepared by David Ford, Science Group, Victoria<br />

& Albert Museum (I.M.87-1930).<br />

cobalt, iron, nickel, <strong>and</strong> a trace of arsenic. On another of the same date <strong>and</strong><br />

on two later miniatures (1730-1735 C.E.), a very different type of smalt was<br />

fo und in the sky <strong>and</strong> Krishna's skin (Plate 14). EDXRF spectra showed a<br />

high arsenic content with iron, cobalt, nickel, bismuth, <strong>and</strong> lead (Fig. 5). No<br />

orpiment (arsenic) was present on these paintings.<br />

By coincidence, the author had been analyzing a number of enameled Venetian<br />

glass vessels (44). Blue enamel, opacified with tin oxide, occurs on<br />

many of these objects. In a group of five, all late-fifteenth to early-sixteenth<br />

century, the blue enamel appeared to be opacified with arsenic; bismuth <strong>and</strong><br />

lead were also present (Fig. 6) . Lead oxyarsenate, 3Pb3(As04)2'PbO, is said to<br />

have first been used as an opacifier in European heavy lead glass in the eighteenth<br />

century (45). Neither the Venetian enamel nor the Basohli smalt seem<br />

500<br />

450<br />

400<br />

Fe<br />

As & Ph<br />

350 Co<br />

300<br />

250 Bi<br />

Bi<br />

200<br />

150<br />

Ca<br />

Sr<br />

100<br />

50<br />

2 4 6 8 10 12 14 16 18 20<br />

Figure 6. EDXRF spectrum of cobalt blue enamel with a high arsenic content <strong>and</strong> traces of bismuth<br />

<strong>and</strong> lead. Venetian eflamelled glass bowl, 1521-1523. Spectrum prepared by David Ford, Science<br />

Group, Victoria & Albert Museum (C. 170-1936).<br />

KeV<br />

Darrah 75


to contain sufficient lead to fulfill the requirements of the above formula or<br />

for a heavy lead glass.<br />

In Venice, there is blue enamel opacified with arsenic 250 years before it is<br />

known to have been used to opacifY glass. In Basohli two centuries later, a<br />

pigment previously unknown in India was used <strong>and</strong> in three cases; it is very<br />

similar to the Venetian enamel. The enamels are of exactly the same period<br />

as when smalt first appears in European paintings, but in no example yet<br />

identified does it contain arsenic <strong>and</strong> bismuth. Do we have two completely<br />

independent discoveries, or were both importing enamel fr it from some Middle<br />

Eastern source that was using arsenic as an opacifier earlier than the<br />

Venetians? The author has been unable to find any analyses which may hold<br />

the answer to the sources of Persian or Middle Eastern enamels of the relevant<br />

period.<br />

As a footnote, smalt contammg a trace of bismuth was found on a south<br />

German oil painting (ca. 1400-1500 C.E.) in the collection of the Courtauld<br />

Institute of Art (46). In this case, it seems likely that the cobalt oxide colorant<br />

was obtained from the residue of the smelting of bismuth (as described previously),<br />

<strong>and</strong> that it is an accidental constituent.<br />

Conclusion<br />

Trade, industry, <strong>and</strong> art are ancient, ubiquitous, <strong>and</strong> international. This paper<br />

has endeavored to show that "new" pigments can still be found in unexpected<br />

places. The connections <strong>and</strong> coincidences that become apparent to the scientist<br />

through laboratory testing indicate historical trade <strong>and</strong> industrial developments,<br />

<strong>and</strong> may sometimes ultimately depend on the craftsperson who<br />

experiments with new materials. Our ability to underst<strong>and</strong> these connections<br />

<strong>and</strong> coincidences requires knowledge of metalworking, ceramics, glass, <strong>and</strong><br />

enamels to explain these connections <strong>and</strong> coincidences both in time <strong>and</strong><br />

place.<br />

Notes<br />

1. Watson, 0. 1994. Personal communication. Ceramics <strong>and</strong> Glass Collection, Victoria<br />

& Albert Museum, London SW7 2RL.<br />

2. Turner, W E. S., <strong>and</strong> H. P. Rooksby, 1959. A study of the opalising agents in<br />

ancient opal glasses throughout three thous<strong>and</strong> four hundred years. Glastechnische<br />

Berichte 32K (VIII): 17-28.<br />

3. Neri, A. 1612. L'Arte Vetraria. Florence.<br />

4. Turner, op. cit.<br />

5. Darrah, J. A. 1992. Venetian <strong>and</strong> Facon de Venise glass. Unpublished analyses.<br />

London: Victoria & Albert Museum.<br />

6. Merrifield, M. P. 1849. Original treatises Jrom the XIIth to the XVIIlth centuries on<br />

the arts oj painting. John Murray, London, 200-5.<br />

7. Merrifield, op. cit., 704-5.<br />

8. Merrifield, op. cit., cliii.<br />

9. Harley, R. 1982. Artist's pigments c. 160(}-1835. 2nd ed., Butterworth Scientific,<br />

172-73.<br />

10. Harley, op. cit.<br />

11. Ibid.<br />

12. Ibid.<br />

13. Tingry, 1830. The Painter's <strong>and</strong> Colourman's Complete Guide. 3rd ed. London:<br />

Sherwood, Gilbert <strong>and</strong> Piper, 45.<br />

14. Linton, W 1852. Ancient <strong>and</strong> Modem Colours Jrom the Earliest Periods to the Present<br />

Time. Longman, Green, Brown <strong>and</strong> Longman. London, 55.<br />

15. Field, G. 1890. The Rudiments oj Colours <strong>and</strong> Colouring. Strahan <strong>and</strong> Co. London,<br />

63.<br />

16. Mayer, R. 1951. The Artists H<strong>and</strong>book oj <strong>Materials</strong> <strong>and</strong> <strong>Techniques</strong>. London: Faber,<br />

69.<br />

17. Isacco, E., <strong>and</strong> J. A. Darrah. 1993. The ultraviolet <strong>and</strong> infrared method of analysis.<br />

A scientific approach to the study of Indian miniatures. Artibus Asiae, LIII (3/<br />

4):470-91.<br />

18. Merrifield, op. cit., 650-5 1, 700.<br />

76<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


19. Merrifield, op. cit., 745-6.<br />

20. Merrifield, op. cit., ccxxi-ccxxiii.<br />

21. Merrifield, op. cit.<br />

22. Linton, op cit., 64.<br />

23. Toch, M. 1911. <strong>Materials</strong> fo r Permanent <strong>Painting</strong>. Constable & Co., 169.<br />

24. The paint box (W65-1920) also contained Krapplack, Pariser blau, verdigris,<br />

Neapelgelb, Prussian blue, beirnschwarz, lichter ocker, smalt blau, kernschwarz,<br />

Cremserweiss, chrome yellow, Indian red, <strong>and</strong> two red ochres <strong>and</strong> a yellow ochre<br />

(with no labels).<br />

25. Grissom, C. A. 1986. Green earth. In Artists' Pigments. Ed. R. L. Feller. Washington:<br />

National Gallery of Art, 141-67.<br />

26. Cove, S. 1991. Constable's oil painting materials <strong>and</strong> techniques. In Constable. L.<br />

Parris <strong>and</strong> 1. Fleming-Williams. London: Tate Gallery, 493-5 18.<br />

27. I am grateful to Sarah Cove (Constable Research Project) for drawing my attention<br />

to four fu rther references to burnt green earth <strong>and</strong> Verona brown in the<br />

latter half of the nineteenth century, unfortunately too late to include them in<br />

this paper.<br />

28. Darrah,]. A., Victoria & Albert Museum Conservation Department Reports:].<br />

F. Millet, The Wood Sawyers, ca. 1850 (CAl 47);]. F. Millet, The Well at Cruchy,<br />

1854 (CAl 49);]. F. Millet, L<strong>and</strong>scape with Terminal Figure, ca. 1864 (CAl 172);<br />

<strong>and</strong> P. de Wint, A Woody L<strong>and</strong>scape, 1812-1816 (261-1872).<br />

29. Merrifield, op. cit., ccxxi-ccxxiii.<br />

30. Singh, R. N. Ancient Indian Class. Delhi: Parimal Publications, 53, 169-71.<br />

31. Vose, R. H. 1980. Class. Collins Archaeology, 30.<br />

32. Muhlethaler, B., <strong>and</strong> ]. Thissen. 1969. Smalt. Studies in Conservation (14) :47-61.<br />

33. Gettens, R.]., <strong>and</strong> G. L. Stout. 1958. A Monument of Byzantine wall painting:<br />

The method of construction. Studies in Conservation (3):107-18.<br />

34. Frank, S. 1982. Class <strong>and</strong> Archaeology. Academic Press, 23.<br />

35. Medley, M. 1976. The Chinese Potter: A Practical History of Chinese Ceramics. Phaidon<br />

Press, 176-78.<br />

36. Lazzarini, L. 1982. The discovery of Egyptian blue in a Roman fresco of the<br />

medieval period (A.D. ninth century) . Studies in Conservation (27):84-86.<br />

37. Merrifield, op. cit., ccvii.<br />

38. Grissom, op. cit.<br />

39. Gettens, R.]., <strong>and</strong> G. L. Stout. 1966. <strong>Painting</strong> <strong>Materials</strong>. New York: Dover, 157-<br />

59.<br />

40. Kuhn, H. 1973. Terminal dates for paintings derived from pigment analysis. In<br />

Application of Science in the Examination of Works of Art. Seminar proceedings.<br />

Boston, 199-205.<br />

41. Frank, op. cit.<br />

42. I should like to thank V. Sharma, visiting Nehru Fellow of the Shirnla Museum,<br />

India, for bringing these paintings to my attention.<br />

43. Archer, W G. 1973. Indian <strong>Painting</strong>s of the Punjab Hills. Vol. 1. London, 16, 21.<br />

44. Darrah, 1992, op. cit.<br />

45. Turner, op. cit.<br />

46. I am grateful to A. Burnstock, Conservation <strong>and</strong> Technology Department, Courtauld<br />

Institute of Art, London University, for allowing me to use the results of<br />

her analysis of St. John of Patmos (CIA 302).<br />

Darrah 77


Abstract<br />

A comprehensive examination was<br />

conducted on a set of twenty-three<br />

seventeenth-century Tibetan thangkas<br />

owned by the Museum of Fine Arts,<br />

Boston (BMFA). The examination<br />

was undertaken because of the<br />

dearth of technical information on<br />

Tibetan thangkas in Western literature.<br />

Infrared reflectography was used<br />

to document color notations drawn<br />

on the ground layer by the artist(s)<br />

as a guide fo r the artist(s) <strong>and</strong> apprentices.<br />

Samples were taken from<br />

areas displaying such notations <strong>and</strong><br />

multiple analytical techniques were<br />

utilized to identify the pigments.<br />

The existence of more than one color-code<br />

system became evident<br />

when the color notations <strong>and</strong> the<br />

identified pigments from the BMFA<br />

set were compared to those found<br />

on other Tibetan paintings. Nine<br />

comparative paintings were examined,<br />

fo ur from the BMFA <strong>and</strong> five<br />

from the Los Angeles County Museum<br />

of Art.<br />

Figure 1. Detail IR-riiflectogram from The<br />

Buddha Shakyamuni Preaching at Dhanyakataka,<br />

showing color codes throughout an<br />

off ering bowl if jewels. Denman Waldo Ross<br />

Collection, Museum if Fine Arts, Boston<br />

(06.333).<br />

An Investigation of Palette <strong>and</strong> Color Notations<br />

U sed to Create a Set of Tibetan Thangkas<br />

Kate I. Duffy*<br />

Conservation Department<br />

Los Angeles County Museum of Art<br />

5905 Wilshire Boulevard<br />

Los Angeles, California 90036<br />

USA<br />

Jacki A. Elgar<br />

Asiatic Department<br />

Museum of Fine Arts<br />

465 Huntington Avenue<br />

Boston, Massachusetts 021 15<br />

USA<br />

Introduction<br />

Tibetan thangkas are scroll paintings that incorporate Buddhist iconography.<br />

The painting is done on cloth that is stitched into a framework of silk borders.<br />

Along the top edge is a wooden stave from which the painting is hung. Along<br />

the bottom edge is a wooden dowel around which the thangka is easily rolled<br />

for storage <strong>and</strong> transport. In 1906 the Museum of Fine Arts, Boston acquired<br />

a set of twenty-three thangkas (1). The paintings in this set had lost their<br />

original thangka format <strong>and</strong> were mounted on panels upon their arrival.<br />

Originally thirty-three or thirty-four paintings comprised the set, depicting<br />

the thirty-two Kings of Shambhala, the Buddha Shakyamuni, <strong>and</strong> possibly a<br />

Kalachakra M<strong>and</strong>ala. The existing set, referred to as the "Shambhala paintings"<br />

throughout this paper, contains only the Shakyamuni <strong>and</strong> twenty-two<br />

images of the Kings (Plates 15, 16). Over the years the set has been assigned<br />

various dates <strong>and</strong> places of origin. Today it is generally accepted to originate<br />

from late seventeenth-century Tibet.<br />

Initially, each Shambhala painting was surveyed by infrared reflectography<br />

(IRR), a nondestructive technique that enabled color codes on the ground<br />

layer to be viewed (2). The color notations are h<strong>and</strong>written in Tibetan<br />

dbu. med script. In areas with complex juxta positioning of numerous colors<br />

(such as offering bowls containing multicolored jewels, or garments with intricate<br />

folds), notations were observed in abundance (Fig. 1). Once the infrared<br />

data were compiled, X-ray fluorescence (XRF) analyses of twelve to<br />

fifteen areas were carried out. After examining these results, approximately<br />

ten areas were sampled. Multiple analytical techniques were utilized, including<br />

polarized light microscopy (PLM), Fourier transform infrared spectroscopy<br />

(FTIR), electronprobe microanalysis (EPMA), X-ray diffraction (XRD), highperformance<br />

liquid chromatography (HPLC), ultraviolet/visible absorption<br />

spectrometry (UV /vis), <strong>and</strong> fluorescence spectrophotometry (FS). Black pigments,<br />

presumably carbon-based, were not studied. Decorative gold was also<br />

not examined except for initial XRF analyses.<br />

Preparing the support<br />

The Shambhala thangkas are all painted on cotton cloth supports. According<br />

to Jackson <strong>and</strong> Jackson, who documented the practices of living thangka<br />

painters, the painting support is typically first made taut by stitching it to four<br />

pliable sticks (3) . It is then laced into a larger wooden frame, leaving a space<br />

of several inches between outer <strong>and</strong> inner frames. This space is crucial for<br />

adjusting tension. The support is stiffened by sizing both sides with a gelatin<br />

solution. A mixture of finely ground white pigment <strong>and</strong> size solution is then<br />

* Author to whom correspondence should be addressed.<br />

78<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Table 1. Tibetan color names, translations, <strong>and</strong> identified pigments.<br />

COLOR<br />

white<br />

blue<br />

Iliaht blue<br />

OOTA1lQII ENGUSH SPEWNG TIBETAN COLOR NAME ENGLISH TRANSLATION<br />

ka dkar 0 white<br />

Iha mthina azurite<br />

n oo Isnoo Inoht blue<br />

PIGMENTS IDENTIFIED<br />

kaolin 246<br />

azurite (1 2 4,6\<br />

azurite kaolin 2 3 4 6<br />

Iliehter blue<br />

blue/Qreen<br />

blue/oreen 1<br />

Iliaht blue/eraen t=.<br />

dark blue/ reen<br />

ellow<br />

ellow<br />

ellow "<br />

briaht red<br />

dark red<br />

ink<br />

ink<br />

oranae '"<br />

n kva snno skva IJinht blue + white<br />

loa IsoanoDO malachite<br />

azurite kaolin 2 3 4 6<br />

malachite brochantite (2.4)<br />

'= \ 'ana sb ar ra nn "comnounded nreen"<br />

ornimentlreal ar indiao kaolin 1 2 3 4 6<br />

to. , , 'ann sbvar Fana "comnnunded nreen" ornimentlreal ar indiQo kaolin (1 2 3 4 6)<br />

'2.. 'ana kva<br />

fana skva "comoounded reen" + white samole not taken<br />

'= \ IXQ") 'ana naa<br />

fann na "comnnunded teen" + black sam Ie not taken<br />

;:,<br />

.:) <br />

,<br />

'"<br />

fAA.<br />

"-<br />

0( "


Figure 2. Detail IR-refiectogram from The<br />

Twenty-ninth King of Shambhala, showing<br />

the color code ngo kya on the leg of a<br />

goat. Denman Waldo Ross Collection, Museum<br />

if Fine Arts, Boston (06. 335).<br />

The notation ngo (sngo), found on twenty-two Shambhala paintings, means<br />

light blue or sky blue (9). Ngo was used for sashes, linings of garments, jewels,<br />

nimbuses, rock crags, <strong>and</strong> lotus petals. Samples of this color taken from four<br />

paintings were identified as a mixture of azurite <strong>and</strong> kaolin. Further whitening<br />

of this pigment is indicated by the notation ngo kya (sngo skya) <strong>and</strong> is used<br />

to depict the color for water, a goat, the sky, <strong>and</strong> the skin color of a king (10,<br />

11, 12). This notation is shown in Figure 2. The additional character kya<br />

indicates that more than 50 percent of the color mixture is white. An actual<br />

quantitative measurement is difficult to carry out; however, one XRD pattern<br />

of a ngo kya sample indicated a higher kaolin:azurite ratio than an XRD<br />

pattern taken from a ngo sample.<br />

Greens. Basic green is denoted by pa (spang), an abbreviation fo r the Tibetan<br />

word meaning malachite (13). Pa, observed on all twenty-three Shambhala<br />

paintings, is used for rock crags, l<strong>and</strong> masses, garments, jewels, <strong>and</strong> foliage.<br />

The green pigment was identified by FTIR <strong>and</strong> XRD in four samples as a<br />

mixture of malachite <strong>and</strong> brochantite. Since brochantite is associated with<br />

malachite deposits, the combination is probably a natural one (14). The mined<br />

source for malachite in Tibet was probably the same as previously mentioned<br />

for azurite (15).<br />

A mixed green was found on twenty-one Shambhala paintings. Jackson <strong>and</strong><br />

Jackson describe a "compounded green" (sbyar ljang) derived from a mixture<br />

of orpiment <strong>and</strong> indigo (16). The notation uncovered is jang, occasionally<br />

written 'jang, <strong>and</strong> is used for nimbuses, mountains, lotus centers, <strong>and</strong> leaves.<br />

The color varies from deep blue to aqua to yellowish green. Five samples<br />

taken from five paintings were examined by FTIR, XRD, EPMA, <strong>and</strong> PLM.<br />

Examination of two of the five dispersed pigment slides revealed a mixture<br />

of realgar <strong>and</strong> clay. Two other slides contained a mixture of realgar, orpiment,<br />

<strong>and</strong> clay. The fifth contained a mixture of only orpiment <strong>and</strong> clay. Indigo,<br />

tentatively identified in three dispersed pigment slides, was positively identified<br />

by FTIR in one of the samples. UV Ivis spectrophotometry analysis of<br />

these samples is planned fo r the fu ture in hopes of definitively identifYing<br />

indigo. The chromatic differences found in the samples may be due to the<br />

fugitive nature of the indigo or to the discoloration of either realgar or orpimento<br />

Indigo may fade when applied thinly, especially when exposed to<br />

sunlight (17). Although each sample appears to be slightly different, it is<br />

thought that the original ingredients were the same: namely, clay, orpimentl<br />

realgar, <strong>and</strong> indigo. The notation jang kya mang skya), meaning light green,<br />

was revealed on two Shambhala paintings for decorative elements of a king's<br />

throne, clouds, <strong>and</strong> l<strong>and</strong> masses (18). The color notation jang nag jang nag),<br />

meaning dark green, is documented on four Shambhala paintings <strong>and</strong> is used<br />

for leaves <strong>and</strong> l<strong>and</strong> masses (19). The paint covering this notation as well as<br />

the jang kya has yet to be sampled.<br />

Yellow. Color notations fo r yellow (ser po) were documented on nineteen<br />

Shambhala paintings. The notations are actually written as one of the following<br />

three: se, ser, or sare. Se is the most abbreviated <strong>and</strong> sare is a misspelling,<br />

perhaps to facilitate writing. The pigment mixture for these notations was<br />

consistently fo und to contain orpiment mixed with kaolin <strong>and</strong> a small amount<br />

of red lead. Five samples from four paintings were taken from areas with sare.<br />

FTIR analyses indicate a kaolin-type clay, similar to the spectrum described<br />

for ka. Examination of four samples by EPMA identified silicon, aluminum,<br />

magnesium, <strong>and</strong> lead, as well as minor amounts of calcium, iron, <strong>and</strong> arsenic.<br />

PLM revealed clay particles mixed with spherical aggregates of red lead <strong>and</strong><br />

fine to medium particles of orpiment. On one slide several large particles of<br />

realgar were identified. Realgar (arsenic disulfide) is often fo und in natural<br />

deposits with orpiment (arsenic trisulfide). Large deposits of orpiment exist<br />

near Chamdo in eastern Tibet <strong>and</strong> in the Yunnan Province of China (20,<br />

21).<br />

One of the samples taken from the area marked se was nearly colorless. Surface<br />

elemental analysis by XRF identified arsenic as a major element, yet neither<br />

80<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


J<br />

Figure 3. Three-dimensional spectrum of na sample from The Twentieth King of Shambhala.<br />

Denman Waldo Ross Collection, Museum of Fine Arts, Boston (06.334).<br />

orpiment nor realgar was fo und by PLM. The FTIR spectrum showed kaolin.<br />

The remainder of the sample was analyzed by XRD. The majority of the<br />

peaks correlate with those of metahalloysite. The remaining peaks were assigned<br />

to kaolinite <strong>and</strong> arsenolite (As203), a white mineral. It is unlikely that<br />

arsenolite is original to the painting. The yellow orpiment probably faded<br />

<strong>and</strong> became a whitish arsenic trioxide. The impermanence of arsenic trisulfide<br />

has been documented elsewhere (22).<br />

Reds. Two shades of red are observed on the Shambhala paintings. One is a<br />

bright red, the other is muted. The bright red uses the notation ga (rgya mtshal)<br />

for Chinese or Indian vermilion (23). Ga is found on fourteen paintings <strong>and</strong><br />

depicts a color used for garments, lotus petals, <strong>and</strong> nimbuses. PLM, XRD, <strong>and</strong><br />

FTIR identified the color as a mixture of vermilion <strong>and</strong> kaolin. Under the<br />

microscope, the vermilion particles appear as finely ground spherical particles<br />

with a deep orange-red color. Several historical references are given by Jackson<br />

<strong>and</strong> Jackson which state that Tibetan painters had access to both synthetic<br />

mercuric sulfide <strong>and</strong> to the natural mineral cinnabar from China, India, <strong>and</strong><br />

Tibet (24).<br />

The darker red is denoted by ma (dmar po), meaning red color (25). Ma is<br />

found on fourteen paintings <strong>and</strong> describes a color used for garments <strong>and</strong><br />

sashes. The paint consists of a dual layer with a dark red organic coating over<br />

a red pigment layer consisting of vermilion <strong>and</strong> kaolin. Further analyses are<br />

pending <strong>and</strong> no conclusive identification has been made.<br />

Pink. The color pink is used mainly fo r clouds but also for jewels, buckles,<br />

makaras, <strong>and</strong> garudas (26). Pink is denoted by na or, occasionally, na kar, <strong>and</strong><br />

represents lac dye (na rosy mixed with a white pigment (27, 28). Na is fo und<br />

on all twenty-three Shambhala paintings. Areas were sampled on six of the<br />

paintings <strong>and</strong> examination of dispersed pigment slides revealed a clay base<br />

mixed with a red dyestuff. FTIR analyses identified the clay as kaolin, as<br />

described earlier. Identification of the dyestuff was achieved through absorption<br />

spectrometry. The identification was fu rther confirmed by fluorescence<br />

.spectrophotometry, utilizing a method described elsewhere (29). The pink<br />

was identified as a lac dye. Figure 3 shows a three-dimensional plot of the<br />

result, which is characteristic of lac. Jackson <strong>and</strong> Jackson state that much of<br />

the dyestuff was traditionally gathered <strong>and</strong> prepared in Tibet (30). The dye<br />

is extracted from sticks encrusted with a resinous secretion produced by the<br />

lac insect, Kerria laeea Kerr.; the resin is still found today in the eastern Himalayas<br />

where the warmer climate is more conducive to its fo rmation.<br />

Orange. The color orange is represented by la, an abbreviation for the Tibetan<br />

word fo r minium i khri) (31). The notation was uncovered on garments,<br />

nimbuses, finials, crowns, flames, belts, jewelry, roof tiles, wheels, <strong>and</strong> vases. La<br />

was found on fifteen Shambhala paintings. Three sampled areas were examined<br />

by XRD, EPMA, FTIR, <strong>and</strong> PLM. The results identified a mixture of<br />

Duffy <strong>and</strong> Elgar 81


ed lead <strong>and</strong> calcium carbonate. <strong>Historical</strong> references given by Jackson <strong>and</strong><br />

Jackson state that Tibetan painters did not use the natural mineral minium<br />

(32). Instead, the synthetic lead tetroxide was imported from China, Nepal,<br />

<strong>and</strong> India. Calcium carbonate was available in Tibet, particularly in Rinpung,<br />

an area north of Lhasa <strong>and</strong> the seat of government in the sixteenth century<br />

(33). In certain parts of Tibet, the cost of calcium carbonate was prohibitive<br />

<strong>and</strong> less expensive white pigments were often used. It is unknown why calcium<br />

carbonate was used to lighten the red lead. Other pigments were mixed<br />

with kaolin. The choice may be due to the purity <strong>and</strong> color of the calcium<br />

based pigment.<br />

Summary of palette <strong>and</strong> color notations for the Shambhala set<br />

Pigment identification determined that the notations on the ground layer<br />

directly correlate to the pigments used, supporting the belief that these notations<br />

were a guide for the artist(s) <strong>and</strong> apprentices. Pigments used for the<br />

respective color notations are similar throughout, which is not surprising given<br />

that the Shambhala paintings belong to one set.<br />

The artist's palette for the Shambhala paintings from the Museum of Fine<br />

Arts, Boston (BMFA) consists of pigments derived from minerals or synthetic<br />

mineral analogues with the exception of two organic dyes, indigo <strong>and</strong> lac<br />

(Table 1). The pigments identified in this investigation deviate little from the<br />

modern-day painter's palette, as documented by Jackson <strong>and</strong> Jackson. The<br />

pure blue <strong>and</strong> green colors were painted with coarsely ground, unadulterated<br />

mineral pigments. All other colors were mixed with a white pigment either<br />

to lighten the color or to achieve translucency. The orange color is unique<br />

in that it was created by mixing calcium carbonate with red lead. In all other<br />

cases kaolin clay was used, perhaps due to its availability <strong>and</strong> low cost.<br />

All color notations are h<strong>and</strong>written in Tibetan dbu. med script on the ground<br />

layer. Some derivative colors are indicated by additional dbu. med characters<br />

to the root notation. Figure 2 is an infrared reflectogram depicting such a<br />

color notation. In this case, kya is added to the root notation ngo. The entire<br />

notation, ngo kya (sngo skya), indicates a light blue color. The second character,<br />

kya, is used to indicate whitening of the existing color, ngo. Another notation<br />

found indicates the darkening of an existing color. The notation has the<br />

additional characters na <strong>and</strong> ga, pronounced nag. These additions represent<br />

the Tibetan word for black (nag po). In the case of jang nag jang nag), nag<br />

indicates darkening of the existing green color, jang.<br />

Comparative paintings<br />

Comparative paintings from the Museum of Fine Arts, Boston <strong>and</strong> the Los<br />

Angeles County Museum of Art (LACMA) were examined using infrared<br />

reflectography (34, 35). Only a few pigment samples, however, were taken.<br />

The LACMA nineteenth-century commemoration thangka for the bhimaratha<br />

rite from Tashi Lhunpo monastery in central Tibet did not reveal any color<br />

codes.<br />

Figure 4. Detail IR-riiflectogram from A<br />

Mahasiddha <strong>and</strong> Taklungpa Lamas, showing<br />

the color code pkya on a l<strong>and</strong>mass. Los<br />

Angeles County Museum of Art, Los Angeles<br />

(M. 81.206. 12).<br />

Another LACMA painting, A Mahasiddha <strong>and</strong> Taklungpa Lamas (ca. late 1700s)<br />

from Taklung monastery in central Tibet, revealed color codes written in<br />

Tibetan dbu. med script. Many of the notations were similar to those documented<br />

on the BMFA Shambhala paintings. One distinct difference, however,<br />

is in the manner of differentiating derivative colors. For a light green color,<br />

fo r example, the notation used was pkya (spang skya). A subjoined kya is added<br />

to the root character pa to indicate the addition of white (Fig. 4). Three other<br />

paintings examined using IRR revealed color codes in dbu.med script. These<br />

were Shakyamuni with Disciples <strong>and</strong> Dharmatala, two paintings from a set<br />

of five at the BMFA accepted as sixteenth-century eastern Tibetan; <strong>and</strong><br />

LACMA's Portrait oj the Fifth Karmapa, accepted as originating in eighteenthcentury<br />

Kham, a region of eastern Tibet.<br />

82<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 5. Detail IR-reflectogram from Shakyamuni<br />

<strong>and</strong> the Eighteen Arhats, showing<br />

a color code written in Chinese on an Arhat's<br />

back. Los Angeles County Museum of Art,<br />

Los Angeles (M. 83. 105. 18}.<br />

Shakyamuni <strong>and</strong> the Eighteen Arhats, owned by LACMA <strong>and</strong> thought to be<br />

an eighteenth-century work from the Kham region, has color codes on the<br />

ground layer written in Chinese. Unlike the brilliant white magnesite ground<br />

found in the Shambhala paintings, this ground is buff-colored <strong>and</strong> composed<br />

of hydrocerussite <strong>and</strong> kaolinite. Two color-code systems are employed. The<br />

first is numerical: numbers represent specific colors. The second uses Chinese<br />

color names or idioms. Under the dark blue pigment of a begging bowl, IRR<br />

revealed the Chinese character for the number seven. The pigment was identified<br />

by PLM <strong>and</strong> FTIR as azurite. On an Arhat's back, IRR revealed the<br />

Chinese character for hulled rice (Fig. 5). Examination of a dispersed pigment<br />

slide revealed orpiment mixed with a small amount of red lead. A character<br />

not yet translated was uncovered on several deep red colored areas. Examination<br />

of a dispersed pigment slide revealed vermilion mixed with a small<br />

amount of red lead.<br />

Another LACMA painting with Chinese color codes is Palden Remati <strong>and</strong><br />

Her Retinue, accepted as originating from the Gelukpa monastery in Central<br />

Tibet, 1800-1850 C.E. This painting employs the two color-code systems<br />

described earlier with similar notations. The Chinese numbers for three <strong>and</strong><br />

seven were used to denote a light blue color. This color has yet to be sampled.<br />

Presumably "seven" represents azurite <strong>and</strong> "three" represents the second addition<br />

of a white pigment.<br />

Two BMFA paintings displaying Chinese color codes come from a set of five<br />

entitled Stories from the Life of Buddha. The set is thought to be eighteenthcentury<br />

Tibetan. Again, evidence of the two color-code systems described<br />

earlier as well as notations fo r derivative colors were found on these paintings.<br />

The Chinese numbers for two <strong>and</strong> six are used to denote a light green. This<br />

color has yet to be sampled. Presumably "six" represents malachite <strong>and</strong> "two"<br />

the first addition of a white pigment.<br />

Conclusion<br />

Since thangka painting is a tradition passed on from master to apprentice,<br />

determining palette <strong>and</strong> deciphering color-code systems may prove to be<br />

helpful in the identification of specific workshops or painting lineages. The<br />

color-code system used on the BMFA Shambhala paintings has several distinctive<br />

traits. These include an additional character kya (skya) for whitish<br />

tints (Fig. 3) <strong>and</strong> the additional characters na ga (nag) for darker tints. The<br />

artist of the painting A Mahasiddha <strong>and</strong> Taklungpa Lamas indicated whitish<br />

tints with a kya subjoined to the root notation (Fig. 4). This system for<br />

distinguishing derivative colors is similar to one still practiced by some modern<br />

Tibetan thangka painters. The paintings with color codes in Chinese<br />

script pose many questions which are beyond the scope of this paper. In terms<br />

of the ethnic background of the creators of these works, one can only speculate.<br />

Perhaps the painters were Chinese, since a native Tibetan speaker would<br />

be unlikely to write such "private" communications in a fo reign script. On<br />

the other h<strong>and</strong>, a bilingual Tibetan might use Chinese characters to communicate<br />

with Chinese apprentices or coworkers. Likewise, one can only<br />

speculate on whether such paintings were produced in Tibet or China. In<br />

order to answer these questions, more infrared data <strong>and</strong> pigment analyses need<br />

to be compiled from thangkas of known Tibetan as well as Chinese origin.<br />

Acknowledgments<br />

The authors are indebted to the following people at the Museum of Fine Arts,<br />

Boston: Rhona MacBeth, John Robbe, Darielle Mason, Anne Morse, Wu Tung, Arthur<br />

Beale, <strong>and</strong> John Elwood. The authors wish to thank the staff at the Los Angeles<br />

County Museum of Art for their generous support, particularly Victoria Blyth-Hill<br />

<strong>and</strong> John Twilley. For his analytical contributions, the authors are grateful to Arie<br />

Wallert of the Getty Conservation Institute. The authors wish to thank the following<br />

Lamas <strong>and</strong> thangka painters for their help in deciphering color codes: Archung Lama<br />

<strong>and</strong> Gega Lama of Kathm<strong>and</strong>u, Nepal; Megmar, Alex<strong>and</strong>er Kocharov, <strong>and</strong> Chating<br />

Jamyang Lama of Dharamsala, India. For his enthusiasm, support <strong>and</strong> analytical con-<br />

Duffy <strong>and</strong> Elgar 83


tributions, the authors would like to acknowledge Richard Newman at the Museum<br />

of Fine Arts, Boston.<br />

Notes<br />

1. Ross, D. W 1904. The purchase of twenty-six Tibetan paintings from Mme.<br />

Tangeveil, Paris. Unpublished typed letter to M. S. Prichard. Boston: Museum of<br />

Fine Arts.<br />

2. An infrared TV camera equipped with a Hamamatsu No. 2606-06 vidicon was<br />

used at the BMFA.<br />

3. Jackson, D., <strong>and</strong> J. Jackson. 1984. Tibetan Thangka <strong>Painting</strong>: Methods & <strong>Materials</strong>.<br />

London: Serindia Publications, 15-23.<br />

4. Palache, c., H. Berman, <strong>and</strong> C. Frondel. 1951. Dana's System of Mineralogy. Vol.<br />

II. New York: John Wiley <strong>and</strong> Sons, Inc., 165.<br />

5. Shaftel, A. 1986. Notes on the technique of Tibetan thangkas. Joumal of the<br />

American Institute for Conservation 25 (2):98.<br />

6. Jackson, op. cit., 174.<br />

7. Ibid., 175.<br />

8. Ibid., 75.<br />

9. Kocharov, A., <strong>and</strong> ChatingJamyang Lama. 1993. Personal communication. Dharamsala,<br />

India: Library of Tibetan Works & Archives.<br />

10. Ibid.<br />

11. Archumg Lama <strong>and</strong> Gega Lama. 1993. Personal conmumication. Kathm<strong>and</strong>u,<br />

Nepal.<br />

12. Jackson, op. cit., 177.<br />

13. Ibid.<br />

14. Palache, op. cit., 543.<br />

15. Jackson, op. cit., 78.<br />

16. Ibid., 176.<br />

17. Gettens, R. J., <strong>and</strong> G. L. Stout. 1966. <strong>Painting</strong> <strong>Materials</strong>: A Short Encyclopaedia.<br />

New York: Dover Publications, Inc., 120.<br />

18. Kocharov, op. cit.<br />

19. Megmar. 1993. Personal communication. Dharamsala, India: Library of Tibetan<br />

Works & Archives.<br />

20. Jackson, op. cit., 82.<br />

21. Palache et aI., op. cit., 268.<br />

22. FitzHugh, E. W n.d. Orpiment <strong>and</strong> realgar. In Artist's Pigments: A H<strong>and</strong>book of<br />

their History <strong>and</strong> Characteristics, Vol. III. Forthcoming.<br />

23. Jackson, op. cit., 176.<br />

24. Ibid., 80.<br />

25. Ibid., 174<br />

26. Makaras are mythological sea creatures. Garudas are guardians of the sky (part<br />

bird <strong>and</strong> part human).<br />

27. Jackson, op. cit., 92.<br />

28. Kocharov, op. cit.<br />

29. Wallert, A. 1986. Fluorescent assay of quinone, lichen, <strong>and</strong> redwood dyestuffs.<br />

Studies in Conservation 31 (4):145-55.<br />

30. Jackson, op. cit., 113.<br />

31. Ibid., 93.<br />

32. Ibid., 81.<br />

33. Ibid., 82.<br />

34. <strong>Painting</strong>s at LACMA were surveyed with an infrared TV camera, Quantex QVC<br />

2500, equipped with a b<strong>and</strong>pass interference filter (1.6 micron wavelength) to<br />

eliminate chromatic aberrations of the lenses.<br />

35. Kossolapov, A. J. 1993. An improved vidicon TV camera for IR-reflectography.<br />

ICOM Committee fo r Conservation Preprints, 25-31.<br />

84<br />

Histon:cal <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

Icons preserved in the Nile valley<br />

are the least known of Egypt's antiquities.<br />

In the gap between the early<br />

Coptic icons from the fourth to<br />

sixth centuries <strong>and</strong> those of the second<br />

half of the eighteenth century, a<br />

recently discovered group of medieval<br />

icons that are difficult to date<br />

<strong>and</strong> attribute has assumed prominence.<br />

Two beam icons, painted on<br />

similarly made panels of sycamore<br />

timber, were selected to describe in<br />

detail. Even preliminary research on<br />

their technology has resulted in new<br />

material for the study of icon painting<br />

<strong>and</strong> the continuity of historical<br />

Egyptian techniques, materials, <strong>and</strong><br />

mythology into Christian times.<br />

Figure 1. Detail of the reverse side if beam<br />

B. Photograph by Z. Skalova.<br />

New Evidence for the Medieval Production of Icons<br />

in the Nile Valley<br />

Zuzana Skalova<br />

Foundation for the Preservation of Icons in the Middle East<br />

Conservation of Coptic Art Project<br />

12 Tolombat Street, Apt. 17<br />

Garden City, Cairo<br />

Egypt<br />

Introduction<br />

Even the earliest preserved Coptic icons demonstrate that they were created<br />

for monastic circles. They repeatedly portray local saints (often monks) or<br />

they express the doctrines of Coptic religious thought. This monastic patronage<br />

remained constant.<br />

Many Coptic icons were until recently in a poor state of preservation. Today,<br />

only a small number of icons in the Nile valley are restored to their full<br />

advantage (1). Consequently, few art historians have looked for old icons or<br />

recognized them (2) . However, important medieval icons hanging in the<br />

church of St. Mercurius Abu's-Saifain in Old Cairo were admired <strong>and</strong> quite<br />

correctly assessed by Alfred Butler during the 1870s. Although they were<br />

"dim with age <strong>and</strong> indistinguishable," he concluded that "the icons generally<br />

speaking are ancient <strong>and</strong> well executed." Butler's comments on the disfIgurement<br />

of the pictures caused by remarkably careless technology, when compared<br />

with Italian panels, can serve as the point of departure fo r this study<br />

(3) .<br />

During the last five years, the author has fo und, studied, <strong>and</strong> in some cases<br />

restored, some twenty medieval icons, which can be preliminarily dated between<br />

the thirteenth <strong>and</strong> the fifteenth centuries. They are nearly all large<br />

icons, clearly made for public veneration in Eastern churches. Some of these<br />

reflect distinct Coptic patronage through their iconography (4).<br />

In assembling <strong>and</strong> attributing the group, the author relied primarily on technical<br />

aspects emphasizing the icons' kinship, which clearly shows that they<br />

emerged from the same local tradition of workmanship. Common features<br />

include the use of indigenous wood (usually sycamore), the awkward construction<br />

of the panels, a ground layer containing anhydrite, a limited range<br />

of pigments, the use oflow-quality azurite blue, an unburnished golden background,<br />

<strong>and</strong> a thin gray layer of varnish (5).<br />

Stylistic <strong>and</strong> iconographic aspects are more hybrid, but also reflect a rather<br />

peripheral <strong>and</strong> culturally mixed background. This suggests that local craftsmen<br />

<strong>and</strong> foreign painters were working together in Egypt. Thus, though these<br />

paintings are varied in style, the panels have such consistent parallels in the<br />

choice of wood <strong>and</strong> peculiar carpentry, that they can be classified together.<br />

They are manufactured from narrow, roughly assembled planks held together<br />

with huge traverses <strong>and</strong> narrow boards, <strong>and</strong> nailed with big iron nails driven<br />

in from the front. The rifts between the planks are filled on both sides with<br />

plaster <strong>and</strong> covered with palm bark fiber, <strong>and</strong> sometimes textiles, to smoothen<br />

the surface (Fig.1). When compared with panels attributed to the Greek icon<br />

workshops, they appear clumsy <strong>and</strong> extremely heavy to transport, a fact that<br />

would additionally testifY to local provenance. The choice of omnipresent<br />

sycamore wood might have been dictated by scarcity <strong>and</strong> the costs of more<br />

suitable material in the Nile valley (6) .<br />

The finish of the back sides of these paintings is also characteristic for the<br />

group. They were invariably covered with a thick layer of plaster, decorated<br />

with alternating lines of wavy pink-brown <strong>and</strong> gray-blue brush strokes. This<br />

is a rare feature in icon painting, <strong>and</strong> it may suggest the formula of a work-<br />

Skalova 85


shop. Dissemination of such formulae may even have caused decoration of<br />

this type to occur on the back sides of icons in the collection of St. Catherine's<br />

Monastery at Sinai (7). However, these Sinaitic panels are skillfully<br />

made from imported wood; <strong>and</strong> they are analogous to many Byzantine icons<br />

preserved in Europe, rather than to the Coptic icons (8).<br />

The icon supports in the Nile valley must have been made according to a<br />

different studio tradition. According to their technology, they clearly fit into<br />

the context of native (presumably Coptic) practice. Furthermore, the majority<br />

of the previously mentioned technical characteristics can be traced back to<br />

the industries <strong>and</strong> materials of Graeco-Roman <strong>and</strong> Pharaonic Egypt (9).<br />

Two medieval beam icons<br />

Figure 2a, b. One of two cuts, top, showing<br />

medalliollS of all arChat1gel alld an Old Testament<br />

prophet; m'ld Blessing Christ, depicted<br />

ill the central medal/ioll, below. Photographs<br />

by Z. Skalova.<br />

Two unique beam icons in the church of St. Mercurius Abu's-Saifain may be<br />

singled out for discussion. This church was the seat of the Coptic patriarchy<br />

in the Middle Ages <strong>and</strong> it is likely that both sacred pictures were made for<br />

it (10).<br />

The first icon, The Virgin with Child Enthroned between Archangels, Nine Church<br />

Fathers <strong>and</strong> Nine Coptic Monks, measures 44.5 X 246.5 cm, <strong>and</strong> is shown in<br />

Plate 17 <strong>and</strong> Figures 3 (left) <strong>and</strong> 4a, b. The second icon, Six Equestrian Saints<br />

(originally ten saints), measures 45 X 207.5 cm <strong>and</strong> is shown in Plates 18a,<br />

b <strong>and</strong> in Figures 2a, b; 3 (right); <strong>and</strong> 4c-e. The two icons will be referred to<br />

here as beam A <strong>and</strong> beam B, respectively.<br />

Thanks to Butler's description <strong>and</strong> drawing, the conservation history of these<br />

two icons can be traced back more than a century (Fig. 3). Both icons were<br />

repeatedly restored. Beam A is preserved in its original form, which has<br />

helped in reconstructing the structurally altered beam B. Butler counted only<br />

four horsemen in beam B (Fig. 3, right). Today, remarkably, beam B consists<br />

of six horsemen. Some time in the past, this longitudinal icon was cut into<br />

pieces. Meanwhile, two more horsemen from the original beam were added<br />

to Butler's fragment (Figs. 2a, 4c).<br />

The equestrian saints are depicted in sculptured <strong>and</strong> gilded arches that carry<br />

eleven alternating medallions with Old Testament prophets, archangels, <strong>and</strong><br />

the blessing Christ. Each horseman is identified by inscriptions in Coptic <strong>and</strong><br />

Arabic.<br />

Additionally, the symmetry of beam B is marred by the fact that the six holy<br />

horsemen are riding in conflicting directions. It is obvious that their initial<br />

order <strong>and</strong> number have been changed. Two cuts, visible in the restored panel,<br />

confirm this observation (Fig. 4c). Such an intervention must have been the<br />

work of a person who failed to underst<strong>and</strong> the symbolism of this picture. It<br />

seems unlikely that these saints were arranged to gallop away from the Savior<br />

instead of toward him, as is proper in a hieratic composition. Christ, depicted<br />

en buste in one of the medallions, can be seen on Butler's fragment (Fig. 2b).<br />

If the placement of Christ's medallion is accepted as being in the middle of<br />

the beam icon, the number of horsemen comes to ten. When reconstructed,<br />

beam B should be about 350 cm long (Fig. 4e). Clearly, both beams were<br />

conceived together to form part of broader didactic program.<br />

A peculiar aspect of both paintings is their carpentry. They are assembled<br />

from irregularly cut horizontal pieces, three across the width <strong>and</strong> nine across<br />

the length (Figures 4b <strong>and</strong> 4d). Additional sculptured arches are nailed on the<br />

front. On the reverse, vertically placed traverses hold the planks together. Both<br />

\lnnooonnl llnoOOII<br />

FiJ!urc 3. Bea1l1 A (lift); bealll B (riJ!ht). Drawillg by A!fred BIIIler, 1884.<br />

86<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


I )--------------------------1 I i<br />

(b)<br />

------ ---<br />

------- ------<br />

----<br />

--------------------<br />

-------<br />

----.--<br />

---<br />

---<br />

-----<br />

(c)<br />

t<br />

t !<br />

1- - -<br />

---<br />

---<br />

1---<br />

- - - ---<br />

(d)<br />

018<br />

(e)<br />

i<br />

FIgure 4a. Beam A, front. Drawing by Sameh Adly.<br />

Figure 4b. Beam A, reverse. Drawing by Sameh Ad/y.<br />

Figure 4c. Beam B, front. Drawing by Sameh Adly.<br />

Figure 4d. Beam B, reverse as preserved today. Drawing by Sameh Ad/y.<br />

Figure 4e. Beam B, recol1structed. Drawil1g by Sameh Ad/y.<br />

beam icons in the church of St. Mercurius Abu's-Saifain are not only constructed<br />

from planks of the same irregular dimensions, but the wood looks<br />

so similar that it could come from one tree. The wood was identified as<br />

sycamore (Ficus sycomorus L.): the traverses are made from cypress (Cupressus<br />

sempervirens var. horizontalis Gord.) (1 1).<br />

The Egyptian sycamore tree<br />

The peculiar shape <strong>and</strong> dimensions of the support planks of the beam icons<br />

A <strong>and</strong> B might have been dictated by the diameter of the tree trunk that<br />

provided the timber. It does not seem probable that the patron of such significant<br />

icons would economize on the timber. The choice of painter <strong>and</strong><br />

quality of materials usually testifY to the resources <strong>and</strong> intentions of the client.<br />

For the art historian, the interest of this research lies not only in the date <strong>and</strong><br />

provenance of the wood (provide by dendrochronological tests), but mainly<br />

in the reason why sycamore timber was preferred (12). The use of sycamore<br />

for these icons might have derived from the holy attributes of the tree, which<br />

would have been of great importance to the patron.<br />

Skalova 87


Figure 5. One of many scenes from Pharaonic tombs: an image of the tree goddess pouring blessings<br />

011 the deceased. Under the sycamore's shadow are the souls of these deceased ill the form of Ba birds.<br />

After N. de Garis Davis's Seven Private To mbs at Kurnah. Mond Excavations at Thebes [I.<br />

1948. Ed. A. H. Gardiner. London, plate XXXIV.<br />

The symbolic importance of the sycamore tree for Copts originates in Luke<br />

19:4, where the sycamore tree was said to have been climbed by Zachaeus<br />

in his eagerness to see Christ. In Coptic folklore, the sycamore apparently<br />

symbolizes the Coptic people (13). This suggests strong continuity with the<br />

traditional worship of the sycamore.<br />

In ancient Egypt, the sycamore was so common that one of the names of<br />

Egypt was "L<strong>and</strong> of the Sycamore" (14). It was considered the most holy<br />

tree, thanks to the deep shadow its protective crown offered in this sunny<br />

country. Hathor, Nut, <strong>and</strong> Isis-the three ancient goddesses-were believed<br />

to dwell in the sycamore <strong>and</strong> were often depicted nestling in its crown, mostly<br />

as a personification of the tree itself (Fig. 5) (15).<br />

The Pharaonic sycamore cult seems to have survived into local Christian<br />

mythology (16). Even today, people do not like to cut old sycamore. The tree<br />

grows in village cemeteries to provide protective shadow. Associations with<br />

the Virgin Mary resting in Egypt under the sycamore also remain alive to<br />

this day. At the well at Matariyya in Heliopolis, today a suburb of Cairo, a<br />

centuries-old sycamore still grows that is believed to have been visited by the<br />

Virgin with the infant Jesus. This holy place is abundantly described by many<br />

pilgrims as having contained, through the Middle Ages, an enchanting orchard<br />

of balsam <strong>and</strong> other exotic trees, such as cypresses (17).<br />

Christian Ethiopia, which is closely connected with Coptic Egypt, still believes<br />

that in each sycamore one Maria lives; they call the sycamore Marianet.<br />

The link between the ancient Egyptian goddess Hathor, Lady of the Sycamore,<br />

<strong>and</strong> Christian Virgin Mary can be surmised (18).<br />

Conclusion<br />

In Egypt, image veneration <strong>and</strong> sycamore tree veneration have been practiced<br />

since antiquity, <strong>and</strong> great importance was attached to the use of special wood<br />

for sacred images. This tradition was so strong that its survival into Coptic<br />

times is not surprising. Thus the Coptic icons are the repository of an earlier<br />

heritage.<br />

88<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


The underst<strong>and</strong>ing of an icon <strong>and</strong> its appropriate restoration according to its<br />

role <strong>and</strong> spiritual meaning in the church requires not only modern conservation<br />

training <strong>and</strong> scientific apparatus, but also a certain degree of theological<br />

sensitivity <strong>and</strong> awareness. The symbolism with respect to the actual process<br />

of icon painting used to have a fu ndamental significance. What was common<br />

knowledge to contemporaries must be reinterpreted after centuries of oblivion<br />

(19, 20). The icon cannot be understood apart from the wider cultural<br />

<strong>and</strong> theological context to which it belongs, as local traditions differ.<br />

With these typically Egyptian traditions in mind, the author wonders but<br />

cannot scientifically prove, if the twin beam icons might have been made<br />

from wood of an ancient sycamore from some sacred Coptic place, such as<br />

Matariyya. This would perhaps explain why the quality of timber was deemed<br />

unimportant.<br />

The construction of panels <strong>and</strong> the use of sycamore, as well as the distinctive<br />

Coptic iconography, all strengthen the attribution of these sacred pictures to<br />

the Nile valley. Future comparative study of the ancient Egyptian beliefs <strong>and</strong><br />

technologies, <strong>and</strong> their survival through icon painting in the Nile valley, may<br />

enrich our knowledge about one of the oldest <strong>and</strong> most conservative pictorial<br />

traditions in the world.<br />

Acknowledgments<br />

The training program, "Conservation of Coptic Icons," in Egypt was made possible<br />

by the generous support of the Netherl<strong>and</strong>s Directorate General of the Ministry of<br />

Foreign Affairs <strong>and</strong> the Egyptian Antiquities Organisation. The author, an art historian<br />

specializing in the conservation <strong>and</strong> restoration of icons, has been working<br />

since 1989 as the Field Director of the Coptic Museum in Cairo.<br />

Notes<br />

1. Skalova, Z. 1990. Conservation problems in Egypt: icons, preliminary classification<br />

<strong>and</strong> some case studies. In ICOM Committee fo r COl/servation Preprints, 777-<br />

82.<br />

2. Skalova, Z. 1991. A little noticed thirteenth-century Byzantine icon in the<br />

church of St. Barbara in Old Cairo, "The Virgin with Child Enthroned." Bulletin<br />

de la Societe d'Archeologie Copte, xxx. Cairo, 93-103.<br />

3. Butler, A. 1884. The Anciel/t Coptic Churches of Egypt, Vol I. Reprint 1970. Oxford,<br />

104.<br />

4. These <strong>and</strong> other medieval icons in Egypt will be studied <strong>and</strong> included in my<br />

Ph.D. dissertation, Mediaellal lcol/s ill Egypt, at the University of Leiden.<br />

5. I anL very indebted to Paolo <strong>and</strong> Laura Mora, who, when in Egypt as directors<br />

of the Conservation team of the Nefertari Wall <strong>Painting</strong> Conservation Project<br />

1986-1992, kindly shared their experience <strong>and</strong> advised me in this matter.<br />

6. Rutschowscaya, M-H. 1986. Catalogue des bois de l'Egypte copte. Paris. See also<br />

Rutschowscaya, M-H. 1991. Coptic Woodwork. Coptic EI/cyclopedia, Vol. 7,<br />

2325-47. To my knowledge, the research on Coptic wood <strong>and</strong> carpentry is<br />

limited <strong>and</strong> until recently there has been little scientific investigation on medieval<br />

icon panels extant in the Nile valley.<br />

7. Mouriki, D. 1990. Icons from the twelfth to the fifteenth century. In Sinai: Treasures<br />

of the MOl/astery of St. Catherine. Ed. K. A. Manafis. Athens, 385, note 25.<br />

8. This conclusion is based on my observations of the beam icons hanging in the<br />

church of the Transfiguration in St. Catherine Monastery at Sinai <strong>and</strong> various<br />

medieval icons in the European collections.<br />

9. Lucas, A., <strong>and</strong> J. R. Harris. 1962. Allciel/t Egyptiar/ <strong>Materials</strong> <strong>and</strong> Illdustries. 4th ed.<br />

London.<br />

10. Sawirus Ibn al-Mukaffa. 1970. History of Patriarchs of the Egyptian Church. Translated<br />

by A. Kather <strong>and</strong> 0. H. E. KHS-Burmester. Cairo. Vol. III, Part II.<br />

11. [ am very grateful to Prof. dr. P Baas, Onderzoekinstituut Rijksherbariuml<br />

Hortus Botanicus, Rijksuniversiteit Leiden, fo r identifying these species.<br />

12. Rutschowscaya, M-H. 1990. Le bois dans l'Egypte chrt:tienne. In Artistes, artisal/s<br />

et productiol/ artistiqlle au Moyel/ Age, [I. Colloque international. 2-6 Mai 1983.<br />

Ed. Xavier Barral I Altet. Paris: Universite de Rennes, 214-1 5. See the discussion<br />

following her article about difficulties encountered when analysing wooden objects<br />

from Egypt at the Louvre by the carbon-14 method <strong>and</strong> dendrochronology.<br />

Skalova 89


13. MacCoull, L. S. B. 1993. The Apa Apollos Monastery ofPharaon <strong>and</strong> its papyrus<br />

archive. In Museon. Tome 106, note 11. MacCoull cites Les Sycamores, a book<br />

about the accomplishments of 1930s Coptic cultural figures that was published<br />

in 1978 in Cairo.<br />

14. Baum, N. 1988. Arbres et Arbrusles d'Egypte ancienne. Orientalia Lovainiensis Analecta.<br />

Leuven, 17-75.<br />

15. Moftah, R. 1959. Die Heiligen Baume im Altem Agypten. Ph.D. diss. Gottingen:<br />

Georg-August-University.<br />

16. My thanks to Egyptologists Ramses Moftah <strong>and</strong> Michael Jones in Cairo fo r<br />

encouraging this line of interdisciplinary inquiry. OlafE. Kaper provided suitable<br />

illustration.<br />

17. Zanetti, U. 1993. Matarieh, La sainte famille et les baumiers. In Analecta Boll<strong>and</strong>iana<br />

(1 11):21-68.<br />

18. Moftah, op. cit., 23.<br />

19. These are concepts that belong to religious tradition <strong>and</strong> are transm.itted in a<br />

way that cannot always be scientifically investigated, which should not dim.inish<br />

their sign.ificance. I would like to thank the priests of the church of St. Mercurius<br />

Abu's-Saifain for their kind support in the project, as well as my senior trainees<br />

Magdi Mansour Badawy, Hanan Nairouz, <strong>and</strong> Mona Hussein, for their assistance.<br />

20. Abu-el-Makarim, a Coptic priest from the church Harat Zuweilah in Cairo, left<br />

a manuscript dated 1206 that describes the monasteries <strong>and</strong> churches of the Nile<br />

Delta. He mentions that in the church of St. George at Netel near Port Said,<br />

one icon of St. George is painted on a piece of the very wood used to torture<br />

the martyr saint himself. See Father Samuel. 1990. Icons et iconographie en<br />

Egypte au XIIe siecle d'apres Ie manuscrit d'Abu-el-Makarim, publie en Arabe<br />

en Caire en 1984. Le Monde Copte, No. 18. Limoges.<br />

90<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

As part of a five-year systematic survey<br />

of the techniques of English<br />

medieval wall paintings, the highly<br />

important late Romanesque <strong>and</strong> early<br />

Gothic paintings in the Holy<br />

Sepulchre Chapel in Winchester are<br />

being examined. The use of red lake<br />

in the Romanesque scheme is the<br />

earliest recorded use of this pigment<br />

in English wall painting; the identification<br />

of vivianite, now partially altered<br />

to a yellow form, is the first in<br />

any English medieval wall painting.<br />

All aspects of the technique of both<br />

schemes, including their complex<br />

laying out, pentimenti, pigments,<br />

media, <strong>and</strong> gilding are discussed in<br />

the context of contemporary European<br />

painting.<br />

Te chniques of the Romanesque <strong>and</strong> Gothic<br />

Wall <strong>Painting</strong>s in the Holy Sepulchre Chapel,<br />

Winchester Cathedral<br />

Helen C. Howard<br />

Leverhulme Research Fellow<br />

Conservation of Wall <strong>Painting</strong> Department<br />

Courtauld Institute of Art, Somerset House<br />

Str<strong>and</strong>, London WC2R ORN<br />

United Kingdom<br />

Introduction<br />

The exquisite Romanesque <strong>and</strong> Gothic paintings in the Holy Sepulchre<br />

Chapel constitute what is arguably the finest medieval painted interior in<br />

Engl<strong>and</strong>. Dating from circa 1175 <strong>and</strong> circa 1220, respectively, their iconography<br />

<strong>and</strong> style have been discussed exhaustively by Park (1). From the technical<br />

point of view, they provide fascinating insight into the execution of two<br />

schemes of exceptionally high quality, separated in date by scarcely forty years,<br />

but painted in very different techniques. Although the technique of the earlier<br />

decoration is rooted in the tradition of painting a fresco, an additional proteinaceous<br />

binding medium <strong>and</strong> also lead pigments have been identified,<br />

which, together with the sweeping compositional changes made at an advanced<br />

stage of the painting process, suggest that significant portions were<br />

completed a secco. The Gothic painting is perhaps more typical of English<br />

medieval wall painting; though carbonation of lime is still the principal mechanism<br />

of binding, the use of large plaster patches, inclusion of additional<br />

organic binding media, <strong>and</strong> incorporation of pigments unsuitable for application<br />

in alkaline conditions preclude describing the technique as fresco.<br />

Investigation of the painting technique was undertaken as part of a comprehensive<br />

five-year study, funded by the Leverhulme Trust, of English medieval<br />

wall painting techniques <strong>and</strong> in conjunction with a conservation campaign<br />

in the chapel undertaken by the Courtauld Institute of Art <strong>and</strong> sponsored by<br />

the Skaggs Foundation (2) .<br />

The paintings<br />

The most striking area of late twelfth-century painting is on the east wall of<br />

the Chapel, comprising a Deposition in the upper register <strong>and</strong>, below the<br />

dividing geometric border, an Entombment with the Maries at the Sepulchre <strong>and</strong><br />

the Harrowing of Hell. Uncovered when the overlying thirteenth-century<br />

scheme was detached in the 1960s, the painting has escaped the ravages of<br />

multifarious conservation treatments to which it would almost certainly have<br />

been subjected had it been exposed earlier. The surface was severely keyed<br />

in preparation for the thirteenth-century plaster, but despite these numerous<br />

damages the superb quality of the painting is evident (Fig. 1 <strong>and</strong> Plate 19).<br />

Traces of the Romanesque scheme survive elsewhere, as on the south wall<br />

above the western recess where a Resurrection of the Dead with an angel blowing<br />

the last trumpet is positioned on either side. Originally situated in the<br />

corresponding position above the eastern recess, but now transferred to the<br />

north wall of the chapel, is a striking mitered head <strong>and</strong> a sinopia for three<br />

scenes set under architectural canopies. To date, this sinopia is a unique find<br />

in the context of English medieval wall painting, <strong>and</strong> provides important<br />

evidence fo r the technique of the Romanesque scheme.<br />

Extensive remodeling of the chapel in the early thirteenth century, including<br />

the insertion of a rib vault, destroyed large portions of the Romanesque<br />

scheme <strong>and</strong> necessitated a complete redecoration. The resulting scheme is<br />

also of exceptionally high quality, but has been marred by a series of invasive<br />

conservation treatments. The painting originally decorating the east wall (<strong>and</strong><br />

Howard 91


Figure ,' , Holy Sepulchre Chapel, Winchester Cathedrar General view of the twe!Jih-celltllry paillfill.R Oil the east wan Photograph courtesy if the<br />

COllservation if Wall Paintillg DepartlUCllt, Courtauld Institflle of Art, LOHdon,<br />

92 <strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


later transferred to an artificial support at the west end of the hapel) mirrored<br />

the iconography of the Romanesque painting beneath with a few minor<br />

adjustments. The Magdalene in the Deposition was replaced by a centurion<br />

holding a scroll, <strong>and</strong> the Harrowing of Hell was moved to the recess of the<br />

adjacent south wall <strong>and</strong> paired with the Noli me Tangere. Above this recess is<br />

the Entry into Jerusalem <strong>and</strong> a newly identified scene of the Washing of Christ's<br />

Feet (3). Within the western recess of the south wall are scenes of the Martyrdom<br />

of St. Catherine of Alex<strong>and</strong>ria. A striking image of Christ as Pantocreator,<br />

surrounded by evangelist symbols, fills the eastern segment of the vault,<br />

while the remaining segments are painted with Infancy scenes (Annunciation,<br />

Nativity, <strong>and</strong> Annunciation to the Shepherds) , foliage, <strong>and</strong> busts in roundels.<br />

Other thirteenth-century painting survives on the northern arches of the<br />

chapel.<br />

Conservation history<br />

Until the conservation program in the 1960s, very little of the twelfth-century<br />

painting was visible, as it lay beneath the later scheme. The thirteenth-century<br />

painting was itself in extremely poor condition, due to the combination of<br />

humidity <strong>and</strong> the wax coating <strong>and</strong> hair nets applied by Professor Tristram in<br />

the 1920s to consolidate the surface (4) . The conservation campaign in the<br />

1960s involved the following: transfer of the thirteenth-century painting on<br />

the east wall to an artificial support at the west end; detachment <strong>and</strong> replacement<br />

in situ of other areas of thirteenth-century painting; transfer of the<br />

mitered head <strong>and</strong> sinopia from the twelfth-century scheme onto the north<br />

wall; partial removal of the wax coating; <strong>and</strong> consolidation with skimmed<br />

milk <strong>and</strong> lime water. This complex conservation history has far-reaching implications<br />

for the conclusions that can be drawn for the original technique<br />

of the thirteenth-century paintings.<br />

Technique of the twelfth-century painting<br />

Plaster. Information on the construction mortar, depth, <strong>and</strong> stratigraphy of<br />

the Romanesque plaster is largely concealed by edging repairs <strong>and</strong> fills. There<br />

are, however, two areas which supply valuable evidence for the original technique.<br />

The ashlar support is visible at dado level where the plaster has been<br />

lost, <strong>and</strong> it is clear that a single plaster layer approximately 0.5 cm in thickness<br />

was applied to the stone. Elsewhere there are two distinct plaster layers: the<br />

uppermost layer, the intonaco, is approximately 0.5 cm in thickness, while<br />

below that a slightly coarser <strong>and</strong> more yellow plaster appears to be about 1.0<br />

cm thick. It seems reasonable to assume that a layer of plaster was applied to<br />

level the masonry where necessary, <strong>and</strong> that the intonaco was then either<br />

applied over this, or directly on to the stone.<br />

Unfortunately, it was not possible to take a sufficiently large sample for full<br />

analysis of the original plaster, but from visual examination of the polished<br />

cross sections it was estimated that a fmely graded inert aggregate, principally<br />

subangular quartz, was combined with lime in a ratio of approximately 2: 1<br />

to create the intonaco. The conspicuous addition of chopped straw <strong>and</strong> fibers<br />

to the plaster suggests that they were included to increase the mechanical<br />

strength of the render. This organic material may have had the additional<br />

function of acting as a mechanical buffer by absorbing moisture during the<br />

setting of the plaster (5). In some samples, it is also clear that charcoal particles<br />

have been incorporated, with the greatest concentration beneath areas that<br />

are predominantly blue (Plate 20, samples 3 <strong>and</strong> 5). It seems likely that the<br />

charcoal was incorporated to reduce the light scatter from the white substrate,<br />

thereby increasing the covering power of the blue mineral pigments.<br />

Although the surface texture itself is fairly smooth, the overall topography of<br />

the painted surface is remarkably uneven . The application of the plaster is<br />

rather crude, with deep undulations readily visible in raking light (6) .<br />

There are two principal horizontal zones of plaster corresponding to the<br />

narrative registers, <strong>and</strong> another for the ornamental border that divides these<br />

Howard 93


egisters (Fig. 2). Additionally, within each of the principal plaster zones, there<br />

are further divisions that roughly conform to the various individual figures<br />

or groups of figures. These secondary plaster joins, many of which are rather<br />

indistinct (particularly in the lower register) make an underst<strong>and</strong>ing of the<br />

plastering sequence problematic. It is evident however, that the narrow central<br />

plaster patch for the geometric border was applied before the plastering of<br />

either narrative register. In the upper register, the roughly executed plaster<br />

joins, applied wet-over-dry, clearly indicate that the patch for the central cross<br />

was applied first, followed by that fo r the main figure group, <strong>and</strong> finally those<br />

for the flanking figures. By contrast, the plaster joins in the lower register are<br />

indistinct, <strong>and</strong> the application appears to have been wet-on-wet, allowing the<br />

edges to merge. Careful examination of the surface in raking light does, however,<br />

indicate that the central portion may have been applied first.<br />

The plastering of the east wall of the chapel differs significantly from that<br />

typically found elsewhere in Engl<strong>and</strong>, where accumulated evidence at Kempley,<br />

Witley, <strong>and</strong> Canterbury suggests that application in broad horizontal<br />

b<strong>and</strong>s is characteristic of the Romanesque period (7). The plaster patches at<br />

Winchester do, however, find parallels both in Engl<strong>and</strong> <strong>and</strong> on the Continent:<br />

in the scheme (ca. 1100) in the parish church at Hardham (Sussex), where<br />

separate plaster patches were applied for each scene as well as for the borders;<br />

<strong>and</strong> in the Romanesque paintings of Vicq (France), in which the plaster was<br />

applied in a grid of large rectangular patches corresponding to the narrative<br />

<strong>and</strong> ornamental divisions (8, 9).<br />

Preparatory techniques. The detachment of a portion of the thirteenth-century<br />

scheme from the south wall in the 1960s provided particularly significant<br />

information concerning the laying out of the Romanesque painting. A small<br />

but exquisite mitered head was uncovered here; underneath was found a<br />

sinopia for three scenes set beneath architectural canopies.<br />

The existence of a sinopia on the east wall, traces of which are visible where<br />

tiny losses in the into naco exist, was noted by Park (10). Evidence from the<br />

stratigraphy of the plaster <strong>and</strong> from the examination of cross sections, such as<br />

that of Sample 29 showing a trace of red pigment (red earth with some<br />

cinnabar) beneath approximately 0.4 cm of plaster, suggests that the sinopia<br />

was applied either directly on the ashlar support or leveling plaster.<br />

Examination of the surface of the Romanesque plaster in raking light reveals<br />

the use of snapped lines marking the position of the central ornamental border.<br />

The geometric elements within the border were set out by incision into<br />

the wet plaster with the aid of a compass. When this border was moved to<br />

a slightly lower position, new incisions were made, but in this case a slightly<br />

different quality of line is evident since the plaster was clearly no longer as<br />

fresh. Incision into the wet plaster is evident in other distinct areas, as in<br />

Christ's arm <strong>and</strong> the flagon of holy oil in the Entombment. In addition, a<br />

preparatory drawing in yellow iron oxide, clearly visible wherever the paint<br />

layers have been lost, was used to place the main pictorial elements within<br />

the visual field.<br />

Pentimenti. The Romanesque paintings are particularly fascinating from the<br />

point of view of the changes made by the painter at an advanced stage of<br />

the painting process. This is particularly evident in the Entombment, in which<br />

the remains of another head <strong>and</strong> a broad-brimmed hat can just be seen to<br />

the left of the Virgin's head, indicating the original position of the figure<br />

anointing Christ's body with oil (in the final version placed to the right of<br />

the Virgin) .<br />

Many other alterations are visible. For instance, the figure of St. John in the<br />

Deposition was finally painted with his h<strong>and</strong> gesturing toward Christ, but the<br />

yellow preliminary drawing indicates that he was originally conceived with<br />

his h<strong>and</strong> held to his face in grief.<br />

Palette. The present research has established that the palette of the Romanesque<br />

paintings included: gold leaf, Au; natural ultramarine,<br />

94<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Winchester Cathedral<br />

Holy Sepulchre Chapel<br />

o 20 60 80 100 em<br />

East Wall<br />

,<br />

§<br />

i<br />

/<br />

.f<br />

J<br />

.j<br />

...... . . ......... ....... ....... ............. .. -. ... _, . ....... , .... , .. ... -........... --.... _ . .,_ .... .<br />

\,<br />

\<br />

t. , "<br />

Incisions<br />

D Snapped Lines<br />

Plaster Joins<br />

F(rzure 2. Diagral1l of the east lIlall scholle, showillg the positioll of plaster joills, illcisioll, allli S/lapped cord Illarks. Diagrmll collrtesy of the COllserPatioll<br />

if Wall Paimillg Department, Courtauld Illstitute of Art, Londoll.<br />

Howard 95


3Na20·3Al203·6Si02·2Na2S; vivianite Fe3 + 2(P04)2·SH20; cinnab:l.r, HgS;<br />

red lake; red lead, Pb304; hematite, Fe203; green earth,<br />

K[(Al,Fe"')·(Fe",Mg)]· (AlSi3,Si4)O,o (OH)2; yellow iron oxide, Fe203·H20;<br />

<strong>and</strong> lime white, CaC03.<br />

One of the most interesting findings is that red lake was used for the unusual<br />

pink color of Nicodemus's robe. This is the earliest identification of this<br />

pigment in English wall painting, though at Miistair (Switzerl<strong>and</strong>) a red lake<br />

pigment thought to be madder has been identified in the Carolinian scheme<br />

of circa SOO (11).<br />

The inclusion of vivianite in the palette is particularly surprising, not only<br />

since this pigment has not previously been identified in English medieval wall<br />

painting, but also because it was clearly selected fo r its distinctive coloristic<br />

qualities rather than as an economic alternative to other mineral blues (12).<br />

The characteristic deep indigo blue of vivianite, set against a pale blue of<br />

natural ultramarine combined with lime white, was employed for the central<br />

details on the vair (bluish gray <strong>and</strong> white squirrel fur) lining of Nicodemus's<br />

cloak in the Deposition. Although the iron phosphate mineral has now altered<br />

<strong>and</strong> appears green, an examination of the Morgan leaf from the Winchester<br />

Bible (ca. 11 70-1 1 SO) shows that vair linings were represented by pale blue<br />

elements with a dark blue center, <strong>and</strong> it is clear that this was also the intention<br />

in the wall paintings.<br />

Binding media. Analysis undertaken initially by microchemical tests on thin<br />

sections <strong>and</strong> cross sections was followed in some cases by Fourier transform<br />

infrared microspectroscopy. Results confirmed that although the carbonation<br />

of lime is the principal binding mechanism, a proteinaceous component is<br />

also present in some samples. A proteinaceous component was identified, for<br />

example, in a paint layer consisting of red lead, lead white, <strong>and</strong> calcium carbonate<br />

in Sample 27, taken from the impasto decoration of the geometric<br />

border. Likewise, a proteinaceous component was identified in a layer of calcium<br />

carbonate applied beneath a resinous mordant for gold leaf in Sample<br />

30, from the halo of the angel in the Entombment.<br />

Instrumental analysis of the media of comparable wall paintings is rare, but it<br />

is significant that, where available, it indicates similar findings. For instance,<br />

in the paintings (ca. 1130) of Idensen (Lower Saxony), the presence of a<br />

proteinaceous binding medium is associated with blue <strong>and</strong> green pigments,<br />

while in the scheme (ca. 1130) in St. Gabriel's Chapel, Canterbury Cathedral,<br />

both linseed oil <strong>and</strong> protein have recently been identified as part of the<br />

original technique (13, 14).<br />

Application. The presence of a sinopia, <strong>and</strong> the application of the into naco in<br />

overlapping patches, indicate that the primary binding mechanism for the<br />

pigments was the carbonation of calcium hydroxide from the lime plaster;<br />

that is, at least the preliminary drawing <strong>and</strong> initial pigment layers were applied<br />

a fresco. However, the stratigraphy of the final painting is remarkably complex<br />

<strong>and</strong> varies substantially across the pictorial surface, from thin single layers<br />

applied directly on the lime plaster substrate to paint applied in considerable<br />

impasto, often over colored grounds (Table 1). In Sample 24, a single layer,<br />

just 30 f.Lm thick, of yellow iron oxide combined with umber <strong>and</strong> a few<br />

charcoal black particles has been applied directly to the white plaster substrate.<br />

Layers such as this are likely to have been applied a fresco. For areas of flesh<br />

painting, multiple layers have been applied to produce complex effects of<br />

modeling. In Sample 7, the mid-dark flesh tone has been produced by the<br />

application of five different layers of earth pigments-green, red, <strong>and</strong> yellowapplied<br />

either singly or in combination, often with the addition of lime white<br />

<strong>and</strong> charcoal black. Here, carbonation of the lime white pigment provides<br />

additional binding capacity within the complex layer structure. By contrast,<br />

red lead combined with an additional proteinaceous binding medium was<br />

applied in a layer some 200 f.Lm thick for the decorative motifs on the central<br />

border, over a thin layer of carbon black applied directly to the lime plaster.<br />

96<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Gilding. Gold leaf was applied for a small number of particular features such<br />

as halos <strong>and</strong> the decorative borders of drapery. Two samples were taken to<br />

establish the gilding technique; in both cases, analysis by FTIR indicated that<br />

a resinous mordant was used to adhere the gold leaf. The supporting layer<br />

consists of calcium carbonate combined with a protein, probably glue. At 5<br />

jJ.m, the gold leaf is exceptionally thick, more than twice that found at Idensen<br />

(1-2 jJ.m) where it was applied over a "bole" of lead white <strong>and</strong> carbon<br />

black bound with prepolymerized linseed oil (15).<br />

Fading oj the lake pigment. Although the susceptibility of red lakes to fa ding<br />

was known in the Middle Ages, they were nonetheless often used in wall<br />

paintings (16). The key factor in the fading of lake pigments is exposure to<br />

ultraviolet radiation, <strong>and</strong>, indeed, a darker pink is apparent where fresh losses<br />

have occurred in linear details overlaying Nicodemus's drapery. Lake pigment<br />

mixed with white, applied over a white ground <strong>and</strong> unprotected by a glaze<br />

(i.e., the Holy Sepulchre Chapel paintings), is particularly vulnerable to internal<br />

reflection (17).<br />

Alteration oj vivianite. Initial identification of the pigment by polarized light<br />

microscopy, in which the pigment is typified by blue-yellow pleochroism<br />

every 90°, was confirmed by X-ray diffraction (18, 19). In recent years, vivianite<br />

has been identified in medieval painting in Germany, as on the Romanesque<br />

lectern at Freudenstadter (ca. 1150) where it was applied over a<br />

gray ground of lead white combined with carbon black (20). In the context<br />

of English medieval polychromy, a preliminary identification of the mineral<br />

has been made on an Anglo-Saxon stone sculpture from York (21).<br />

Vivianite occurs naturally in two discrete environments. It is fo und in the<br />

oxidized upper layers of some metalliferous ore deposits, as at St. Agnes in<br />

Cornwall, where it generally appears as dark indigo, blue-black, or green<br />

crystals (22). It is also fo und in organic, phosphate-rich environments, <strong>and</strong> is<br />

frequently associated with bones, decaying wood, <strong>and</strong> other organic remains.<br />

Vivianite is generally stable <strong>and</strong> dark blue or green in color, though the<br />

mineral may be colorless when initially exposed (23).<br />

It seems likely that in the medieval period vivianite was used only where the<br />

mineral was locally available; thus, mineral deposits are well known in Germany.<br />

Good crystalline deposits of the mineral have been found at Whale<br />

Chine on the Isle of Wight, <strong>and</strong> in its earthy form at Fordingbridge in Hampshire,<br />

both close to Winchester (24). Current work on samples from the<br />

chapel includes an analysis of trace elements <strong>and</strong> examination of the crystalline<br />

structure to determine whether a mined or peaty alluvial deposit was<br />

the source of the mineral.<br />

Vivianite is known to be generally stable in its blue fo rm but at Winchester<br />

some of the particles have altered to a yellow color, giving an overall green<br />

effect. The mineral's color change from colorless to blue on initial exposure<br />

is due to increased ferric ion concentrations, <strong>and</strong> it has been established that<br />

mechanical grinding of the colorless crystals, heating in air, storage in a vacuum,<br />

or chemical treatment of samples can produce a more rapid conversion<br />

from ferrous to ferric ions, <strong>and</strong> so to a blue color (25, 26). It is therefore<br />

interesting to speculate whether grinding of the blue mineral to produce a<br />

particle size suitable fo r use as a pigment may have contributed to an additional<br />

increase in ferric ion concentration <strong>and</strong> ultimately to a further color<br />

alteration from blue to yellow.<br />

Technique of the thirteenth-century paintings<br />

Invasive conservation interventions-including waxing, facing with glue for<br />

detachment, thinning of the original plaster support, consolidation, <strong>and</strong> cleaning-have<br />

compromised the results of the technical examination of these<br />

paintings. Nevertheless, certain conclusions can be drawn.<br />

Plaster. Following the keying of the twelfth-century painting, a single layer<br />

of plaster approximately 5-8 mm thick was applied to the surface. This<br />

Howard 97


Fi,RlIre 3. This thirteellth-cel1tury pailltil1,R, ori,Ril1aily 011 the east wall, has beel1 transferred to al/. artificial slIpport at the west el1d of the chapel.<br />

Photo,Rraph collrtesy of the COllServatioll of Wall PaintilJ,R Department, Courtauld [llStitllte of Art, Londoll.<br />

98 <strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Winchester Cathedral<br />

Holy Sepulchre Chapel<br />

Jo<br />

tOO em<br />

West Wall<br />

I2J P<br />

laster Joins<br />

Figure 4. Diagralll of the thirteenth-century east schenle (now tral1sferred to the west el/d), showil/g the positiol/ of plaster joil/s. Diagralll col/rtesy<br />

of the COIIServatiol1 of Wall <strong>Painting</strong> Departlllent, Courtauld Institl/te


plaster consists of lime combined with an aggregate of subangular quartz in<br />

a ratio of 1:3. It is generally more yellow <strong>and</strong> less well prepared than the<br />

plaster for the earlier decoration, with large lumps of unmixed calcium carbonate<br />

clearly visible.<br />

It is now difficult to establish the sequence of application of the plaster, since<br />

surface features have been completely flattened by the detachment process<br />

(Figs. 3, 4) . Examination under raking light of the surface of the east wall<br />

paintings (now transferred to the west end) does, however, indicate that plaster<br />

joins may originally have run along the upper edge of the central border <strong>and</strong><br />

the lower horizontal of the cross in the upper register. This indicates that<br />

generally the plaster may have been applied in broad horizontal b<strong>and</strong>s, roughly<br />

corresponding to scaffold lifts. However in the east recess of the south wall<br />

there appears to be a vertical join between the Harrowing of Hell <strong>and</strong> Noli me<br />

Tangere, suggesting that in this area at least a fu rther division was considered<br />

necessary.<br />

Preparatory techniques. A preparatory sketch in red iron earth was used to set<br />

out the main features of the composition. Unlike the twelfth-century painting<br />

wherein both a sinopia <strong>and</strong> then fu rther preparatory drawing on the final<br />

plaster layer exist, in the thirteenth-century scheme the preparatory drawing<br />

is confined to the final plaster layer.<br />

Direct incisions into the plaster are also clearly visible, such as the fine, sharp<br />

incisions that outline the curls in the hair of the figure at far right in the<br />

Entombment <strong>and</strong> the hair of St. John in the Deposition.<br />

Pentimenti. An interesting alteration discovered during the detachment process-<strong>and</strong><br />

presumably visible on the reverse of the thinned plaster layer-was<br />

that above the horizontal arm of the cross in the Deposition was a sketch fo r<br />

the sun <strong>and</strong> moon, which are symbols more usually associated with the Crucifixion<br />

(27).<br />

Pigments. The present research has established that the original palette included:<br />

natural ultramarine (3Na20·3Al203·6Si02·2Na2S); vivianite (Fe3 +<br />

2[P04L'8H20); copper chloride green; cinnabar (HgS); hematite (Fe203); yellow<br />

iron oxide (Fe203'H20); lead white (2PbC03·Pb[OHl2); lime white<br />

(CaC03); charcoal black (C).<br />

Binding media. The inclusion of lead pigments in the palette makes highly<br />

likely the incorporation of organic binding media as part of the original<br />

technique but, owing to the previous conservation interventions, it was not<br />

possible to confirm their presence.<br />

Application. Preliminary drawing in red ochre was followed by blocking in of<br />

the basic background colors. Oakeshott considered that the powerful black<br />

outlines, which are such a major feature of the scheme, were also painted at<br />

this stage <strong>and</strong> were thus more firmly bound by the carbonation of the plaster<br />

(28). The stratigraphy is remarkably simple, with one or at most two paint<br />

layers applied directly to the plaster (Tables 2, 3, 4, 5). In some cases the final<br />

paint layer was applied in thick impasto, <strong>and</strong> it seems likely that here an<br />

additional binding medium was used.<br />

Pigment alteration. Although the background colors of the scheme are now<br />

red <strong>and</strong> green, it seemed much more likely on the basis of comparisons with<br />

other paintings of the period that red <strong>and</strong> blue would originally have been<br />

used. Analysis indicated that this was indeed the case: the present green color<br />

is due to the alteration of some of the original blue pigment (vivianite) to<br />

form a yellow alteration product.<br />

The alteration of a lead pigment is evident in the fo rm of black spots on the<br />

paint surface throughout the scheme, <strong>and</strong> analysis confirmed the presence of<br />

a dark red/brown lead-based material. In wall paintings elsewhere where this<br />

type of alteration occurs, the alteration product has been found to be plattnerite<br />

(lead dioxide) (29). In the Holy Sepulchre Chapel however, analysis by<br />

100<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


scanning electron microscopy I energy dispersive spectroscopy (SEM/EDX)<br />

analysis indicated a low oxygen peak, which may indicate that a mixture of<br />

lead dioxide <strong>and</strong> lead sulfide is present here. In addition, cinnabar has altered<br />

to its dark form, metacinnabar, as shown in Sample 37, taken from the darkened<br />

cheek patch of a prophet on the vault.<br />

Acknowledgments<br />

I am grateful to the Leverhulme Trust for their generosity in funding the five-year<br />

study, <strong>and</strong> to the Skaggs Foundation for their support of the conservation program<br />

at Winchester. Particular thanks are also due to G. Cressey <strong>and</strong> P. T<strong>and</strong>y of the Natural<br />

History Museum, <strong>and</strong> to David Park, Sharon Cather, <strong>and</strong> Robyn Pender of the<br />

Conservation of Wall <strong>Painting</strong> Department, Courtauld Institute of Art.<br />

Table 1.<br />

Samples Jrom the east wall oj the Holy Sepulchre Chapel, Winchester Cathedral<br />

Sample No:<br />

3<br />

Location <strong>and</strong> Description<br />

E. wall, upper tier, S. side, olive green<br />

background to left of Nicodemus' foot.<br />

Original Polychromy<br />

• vivianite 20llm<br />

• natural ultramarine 30llm<br />

Indications of binding medialPigment<br />

alterations <strong>and</strong> other additions<br />

• vivianite has partially altered from blue to yellow<br />

________ <br />

-- <br />

-- <br />

-- <br />

-- <br />

--. I ;m e p I t er w ; il i c ha r o o al b la ck ;n cl us ;o n ; s =. 40 O m <br />

<br />

<br />

-- <br />

----------__________<br />

- E. wall, upper tier, S. side, red background<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

21<br />

22<br />

adjacent to cross, red over blue.<br />

E. wall, upper tier, N. side, uppermost edge of<br />

Joseph of Aramathea's hat, grey surface over red<br />

pigment on blue.<br />

E. wall, upper tier, N. side, Joseph of<br />

Aramathea's red drapery, translucent dark red<br />

shadow just below beard.<br />

E. wall, upper tier, Christ's right fOOl, mid-dark<br />

flesh tone.<br />

E. wall, upper tier, Nicodemus' pink drapery, at<br />

hip level, pale pink with highlight.<br />

E. wall, lower tier, N. side, Joseph of<br />

Aramathea's red robe, darkest red tone at uppermost<br />

edge.<br />

E. wall, lower tier, N. side, Christ's proper h<strong>and</strong>,<br />

flesh painting with pink drop of blood on<br />

surface.<br />

E. wall, lower tier, Harrowing a/Hell, deep<br />

purple shadow in drapery.<br />

E. wall, lower tier, Harrowing of Hell, green<br />

outer drapery (query modem retouching).<br />

• cinnabar in red lead matrix with a f e w particles of natural ultramarine • calcium carbonate crust on surface<br />

220J,lm<br />

• natural ultramarine 30llm<br />

- cinnabar IOllm<br />

• natural ultramarine <strong>and</strong> charcoal black in a lime white matrix 10llm<br />

• lime plaster with charcoal black inclusions 250m<br />

• haematite 20m<br />

• yellow iron oxide 2m<br />

• trace calcium carbonate<br />

• cinnabar 551lm<br />

• lime plaster! OOm<br />

• red, yeUow <strong>and</strong> green iron oxides <strong>and</strong> charcoal black 10m<br />

• green iron oxide 45 m<br />

- lime white combined with yellow iron oxide IOOllm<br />

• yellow iron oxide 25).101<br />

• red iron earth 50).1m<br />

• trace I ime plaster<br />

• red lake with a trace of ultramarine in a calcium carbonate matrix<br />

100).1m. The trace of calcium sulphate present in the layer may<br />

represent the substrate of the lake pigment, now faded.<br />

• plaster substrate with charcoal black 80).1m<br />

• cinnabar 30l1m<br />

• yellow iron oxide IOl1m<br />

• plaster substrate with a few charcoal black inclusions 200l1m<br />

• cinnabar 20l1m<br />

• trace yellow iron oxide<br />

• green iron oxide 4511m<br />

• lime white with yellow iron oxide particles 100l1m<br />

• lime white, yellow iron oxide <strong>and</strong> charcoal black 60l1m<br />

• natural ultramarine, charcoal black <strong>and</strong> lime white 90l1m<br />

• lime white 40l1m<br />

• plaster substrate 200Jlm<br />

• red lake pigment has faded<br />

• cobalt green (CoO'nZnO)<br />

23<br />

24<br />

25<br />

26<br />

E. wall, lower tier, lower edge of sarcophagus,<br />

deep purple over yellow.<br />

E. wall, lower tier, Christ's loin cloth, yellow<br />

shadow over deep yellow.<br />

E. wall, upper tier, Magdalene's drapery, hem at<br />

calf level, yellow & white bole over red/purple<br />

<strong>and</strong> black paint layers.<br />

E. wall, upper tier, Nicodemus' cloak, blue <strong>and</strong><br />

white vair lining, green area over blue.<br />

27 E. wall, central decorative border, 4th circular<br />

motif from north side, impasto motif in orange<br />

28<br />

29<br />

30<br />

31<br />

32<br />

<strong>and</strong> white.<br />

E. wall, upper tier, Nicodemus' right proper leg ,<br />

brilliant rich red highlight over red lead <strong>and</strong><br />

yellow earth.<br />

E. wall, upper tier, wavy ground between<br />

Nicodemus' legs, deep red brown over pink.<br />

E. wall, lower tier, censing angel, halo. trace<br />

gilding on bole over brown paint layer <strong>and</strong> red<br />

lead.<br />

E. wall, lowertier, N. side, angel's drapery,<br />

green over yellow near hem.<br />

E. wall, upper tier, N. side, grey/green<br />

background between legs of Joseph of<br />

Aramathea.<br />

• charcoal black, redlbrown iron oxides with a few particles of cinnabar<br />

80/1m<br />

• yeJlow iron oxide combined with lime white 90l1m<br />

• plaster substrate l20/1m<br />

• yellow iron oxide <strong>and</strong> umber with a few charcoal black particles 30/1m<br />

• plaster substrate 100)1m<br />

• resinous layer 50)1m<br />

• white layer (calcium carbonate + protein) 30)1m<br />

• resinous layer lOO).lm<br />

• white layer (calcium carbonate + protein) 20/1m<br />

• resinous layer 100llm<br />

- vivianite in calcium carbonate matrix 90l1m<br />

• natural ultramarine in lime matrix 100J.lm<br />

• plaster substrate 300).lm<br />

• red lead combined with lead white <strong>and</strong> calcium carbonate with a trace<br />

of proteinaceous binding medium present 200).lm<br />

• carbon black 10/1m<br />

• cinnabar combined with red lead 50).lm<br />

• red lead IOO/1m<br />

• red iron oxide 15J.lm<br />

• cinnabar 10).lm<br />

• plaster substrate/ground 400/1m<br />

• underdrawing, red iron oxide <strong>and</strong> cinnabar IOOllm<br />

• gold leaf5/1m<br />

• resinous mordant 20J.lm<br />

• white layer (calcium carbonate + protein) 10/1m<br />

• dark resinous layer 15/1m<br />

• white layer (calcium carbonate + protein) 20m<br />

• dark resinous layer 25).lm<br />

• green iron oxide combined with some charcoal black particles 50/1m<br />

• yellow iron oxide 75m<br />

• yellow iron oxide combined with lime white 180/1m<br />

• plaster substrate with charcoal black particles 120/1m<br />

• charcoal black with a few red <strong>and</strong> yellow iron oxide particles 80).lm<br />

• plaster substrate 200llm<br />

• a proteinaceous binding medium is present in the white layers<br />

• vivianite has partially altered from blue to yellow<br />

• translucent crust of calcium carbonate combined with calcium sulphate on<br />

surface<br />

• a proteinaceous binding medium is present in the red paint layer<br />

• white crust on surface<br />

• white crust <strong>and</strong> trace pigment on surface<br />

• a proteinaceous binding medium is present in the white layers<br />

• trace white crust combined with a few charcoal black particles<br />

The table gives the location where each sample was taken, the sample number. <strong>and</strong> the stratigraphy of the various layers from the top down,<br />

with the thickness of each layer in microns (pm).<br />

Howard 101


Table 2.<br />

Samples from the vest wall if the Holy Sepulchre Chapel, Winchester Cathedral<br />

Sample No:<br />

34<br />

35<br />

36<br />

Location <strong>and</strong> Description<br />

W. wall, upper tier, N. side, green background<br />

colour below white scroll.<br />

W. wall. upper tieT, St. John, drapery below<br />

knee, black over pink.<br />

W. wall, upper tier, S. side, (left as viewed) leg<br />

of Joseph of Aramalhea, dark green linear detail<br />

on back of calf.<br />

Original Polychromy<br />

• viviani!e IOOllm<br />

• charcoal black 80llm<br />

• lime plaster substrate 60llm<br />

• darkened lead pigment l5IJm<br />

• lime plaster substrate 250 11m<br />

• vivianite 150JJm<br />

• charcoal black <strong>and</strong> lime white 40llm<br />

• I ime plaster substrate IOOJJm<br />

• trace of red pigment from 12th century scheme 20JJm<br />

Coatings! Indications of binding media!<br />

Pigment alterations<br />

• beeswax coating containing calcium carbonate <strong>and</strong> calcium sulphate 20llm<br />

• proteinaceous coaling 30llm<br />

• vivianite has partially altered from blue to yellow<br />

• beeswax (trace)<br />

• proteinaceous coating<br />

• lead pigment has altered to fonn a dark product which appears to be a mixtur<br />

of lead dioxide (plattnerite) <strong>and</strong> lead sulphide.<br />

• beeswax (trace)<br />

• proteinaceous coating<br />

• vivianite has partially altered from blue to yellow<br />

Table 3.<br />

Samples from the north wall of the Holy Sepulchre Chapel, Winchester Cathedral<br />

Sample No:<br />

11<br />

12<br />

13<br />

14<br />

Location <strong>and</strong> Description<br />

N. wall, E arch, soffit, E. side, upper figure, left<br />

proper arm, pale pink drapery.<br />

N. wall, E arch, chamfer on S. side, E end, 6th<br />

ashlar block from apex, yellow over green on<br />

plaster.<br />

wall, E arch , S face, E side, C13th vine<br />

scroll in red on plaster.<br />

N. wall, E arch, soffit, E side, lower figure,<br />

thick beige coating over green paint of drapery.<br />

Original Polychromy<br />

• white, lead-containing pigment with charcoal black, calcium carbonate,<br />

calcium sulphate & red pigment particles<br />

• yellow, lead-containing layer which appears white towards perimeter. A<br />

few yellow iron oxide particles are present<br />

• lime ground - calcium carbonate with calcium sulphate <strong>and</strong> a trace of<br />

wax present<br />

• yellow earth pigment in lead white matrix<br />

• vivianite in lead white matrix<br />

• lime white ground<br />

• red earth<br />

• lime ground<br />

• vivianite in lead white matrix<br />

Coatings! Indications of binding media!<br />

Pigment alterations<br />

• proteinaceous coating<br />

• proteinaceous material is indicated in both paint layers<br />

• vivianite has partially altered from blue to yellow<br />

• beeswax<br />

• proteinaceous coating<br />

• beeswax<br />

• proteinaceous coating<br />

• thick lead white layer<br />

• vivianite has partially altered from blue to yellow<br />

The tables give the location where each sample was taken, the sample number, <strong>and</strong> the stratigraphy of the various layers from the top down,<br />

with the thickness of each layer in microns (Jim).<br />

Table 4.<br />

Samples from the south wall of the Holy Sepulchre Chapel, Winchester Cathedral<br />

Sample No:<br />

16<br />

17<br />

18<br />

19<br />

20<br />

41<br />

Location <strong>and</strong> Description<br />

S. wall, E. bay, above recess, blue background<br />

behind donkey's left ear, thick dark impasto<br />

over blue background.<br />

S. wall, W bay, above recess, E. side, old<br />

cleaning test area, coating from in key mark<br />

S. wall, W. bay, above recess, E. side, old<br />

cleaning test area, coating over red! brown paint<br />

layer.<br />

S. wall, W. bay, above recess, W. end, CI2th<br />

angel, area fluoresces in uv, coating over black<br />

<strong>and</strong> orange paint.<br />

S. wall. W. bay, recess, lower zone, central area<br />

of dark red 'drapery', coating over redlblack<br />

paint.<br />

S. wall, E. bay, above recess, Washing offeer,<br />

bright green from 'border' above Christ's head.<br />

Original Polychromy<br />

• darkened lead pigment 100llm<br />

• natural ultramarine (2Sllm)<br />

• organic analysis only<br />

• carbon black pigment 5J1m<br />

• red earth in lime matrix 10JJm<br />

• lime ground 40).lm<br />

• lime plaster 200llm<br />

• black <strong>and</strong> red particles in a lead white matiix 351lm<br />

• lime ground l40llm<br />

- lime plaster substrate 400JJm<br />

• red iron oxide 2SIlm<br />

• charcoal black 40llm<br />

• blue/green copper chloride green layer 50llm<br />

• paler blue/green copper chloride layer 50llm<br />

• calcium carbonate with a little calcium sulphate present 20llm<br />

Coatings! Indications of binding media!<br />

Pigment alterations<br />

• proteinaceous material on surface<br />

• lead pigment has altered to fonn a dark product which appears to be a mixture<br />

of lead dioxide (plattnerite) <strong>and</strong> lead sulphide<br />

• layers of beeswax<br />

• beeswax with a small amount of proteinaceous material<br />

• extremely thin coating present which is almost completely combined with the<br />

upper pigment layer. As yet unidentified<br />

• beeswax<br />

• trace coating on surface<br />

• copper chloride green may be the product of an alteration prOcess<br />

The table gives the location where each sample was taken, the sample number, <strong>and</strong> the stratigraphy of the various layers from the top down,<br />

with the thickness of each layer in microns (/-un).<br />

Table 5.<br />

Samples from the vault if the Holy Sepulchre Chapel, WiNchester Cathedral<br />

Sample No:<br />

15<br />

37<br />

38<br />

39<br />

40<br />

42<br />

43<br />

44<br />

45<br />

46<br />

Location <strong>and</strong> Description<br />

Vault, E. bay, E. segment, under left proper<br />

h<strong>and</strong> of Christ, 'pea' green background of<br />

m<strong>and</strong>orla.<br />

Vault, E. bay, S. segment, W. side, prophet in<br />

roundel, darkened cheek patch.<br />

Vault, E. bay, S. segment, W. side, prophet in<br />

roundel, yellow flesh painting of h<strong>and</strong>.<br />

Vault, E. bay, S. segment, W. side, prophet in<br />

roundel, h<strong>and</strong>, dark flesh area.<br />

Vault, E. bay, W. side, S. segment, prophet in<br />

roundel, left side, dark b<strong>and</strong> around edge of<br />

roundel.<br />

Vault, E. bay, S. segment, W. side, prophet in<br />

roundel, red linear detail of robe over now black<br />

layer.<br />

Vault, E. bay, E. segment, m<strong>and</strong>orla, S. side,<br />

green border with black on surface.<br />

Vault, E. bay. E. segment, S. side, green<br />

background below Christ's h<strong>and</strong>, white<br />

decorative spot over green.<br />

Vault, E. bay, E. segment, m<strong>and</strong>orla, S. side,<br />

green over red.<br />

Vault, E. bay, E. segment, m<strong>and</strong>orla, N. side,<br />

brilliant green of border.<br />

Original Polychromy<br />

• vivianite 75JJm<br />

• black (darkened cinnabar) particles in lead white matrix ISllm<br />

• plaster substrate 140llm<br />

• yellow iron oxide 15JJm<br />

• lime white ground 50l1m<br />

• plaster substrate 200llrn<br />

• charcoal black <strong>and</strong> red iron oxide 40l1m<br />

• lime white ground 100llm<br />

• plaster substrate<br />

• dark red/brown lead-based layer 651lm<br />

• lime white ground 751lm<br />

• red earth 1251lm<br />

• dark redlbrown lead-based material 551lm<br />

• white ground 751lm<br />

• plaster substrate 400llm<br />

• dark red/brown lead-based material 75J.1m<br />

• lime white ground 180llm<br />

• lead white 90l1m<br />

• vivianile 55J.1m<br />

• white layer (calcium carbonate combined with calcium sulphate) 50llm<br />

• trace red pigment<br />

• pale blue/green copper-based layer 351lm<br />

• white ground 751101<br />

• plaster substrate 300llm<br />

Coatings! Indications of binding media!<br />

Pigment alterations<br />

• beeswax<br />

• proteinaceous coating<br />

• vivianite has partially altered from blue to yellow<br />

• proteinaceous coating with additional components of calcium carbonate,<br />

calcium sulphate <strong>and</strong> a trace of wax 151lm<br />

• cinnabar has altered to its black meta fonn<br />

• proteinaceous coating 10llm<br />

• proteinaceous coating with additional components of calcium carbonate <strong>and</strong><br />

calcium sulphate 30JJm<br />

• proteinaceous material has penetrated paint layers<br />

• proteinaceous coating 20llm<br />

• lead pigment has altered to fonn a dark material which appears to be a mixture<br />

of lead dioxide (plattnerite) <strong>and</strong> lead sulphide<br />

• proteinaceous coating 2SI1m<br />

• lead pigment has altered to fonn a dark material which appears to be a mixture<br />

oflead dioxide (plattnerite) <strong>and</strong> lead sulphide<br />

• proteinaceous coating IOllm<br />

• lead pigment has altered to fonn a dark material which appears to be a mixtur<br />

of lead dioxide (plattnerite) <strong>and</strong> lead sulphide<br />

• proteinaceous coating 51lm<br />

• beeswax coating<br />

• proteinaceous material is incorporated in wax coating 351lm<br />

• proteinaceous coating 51lm<br />

The table gives the location where each sample was taken, the sample number, <strong>and</strong> the stratigraphy of the various layers from the top down,<br />

with the thickness of each layer in microns (Jim).<br />

102 <strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Notes<br />

1. Park, D. 1983. The wall paintings of the Holy Sepulchre Chapel. In Mediellal Art<br />

<strong>and</strong> Architecture at Winchester Cathedral. British Archaeological Association Conference<br />

Transactions 1980, VI, 38-62.<br />

2. I am indebted to Stephen Rickerby <strong>and</strong> the students of the Conservation of<br />

Wall <strong>Painting</strong> Course, Courtauld Institute, for their observations during the work<br />

in the chapel, <strong>and</strong> for providing the drawings used in Figures 4 <strong>and</strong> 6. Previous<br />

investigations of aspects of the paintings include: Hluvko, S. 1991. Red piglTlents<br />

in English medieval wall painting; <strong>and</strong> Howard, H. 1988. Blue pigments in English<br />

medieval wall painting. Diploma dissertations. London: Conservation of<br />

Wall <strong>Painting</strong> Department, Courtauld Institute of Art, University of London.<br />

3. This scene, previously described as the Raising of Lazarus, was identified by<br />

Christoph Tinzl.<br />

4. Lee, M. 1975. <strong>Painting</strong>s in the Holy Sepulchre Chapel, Winchester. Unpublished<br />

M.A. report, University of London.<br />

5. For the role of organic additives in plaster, see Sickels, L. 1981. Organics v synthetics:<br />

their use as additives in mortars. In Mortars, Cements <strong>and</strong> Grouts Used in<br />

the Conservation of Historic Buildings . Rome: International Centre fo r the Study<br />

of the Preservation <strong>and</strong> Restoration of Cultural Property, 25-49.<br />

6. Deeply undulating surfaces occur in other Romanesque paintings, such as the<br />

scheme of ca. 1130 in St. Gabriel's Chapel, Canterbury Cathedral. See Cather,<br />

S., <strong>and</strong> H. Howard. 1994. Romanesque wall paintings in the apse of St. Gabriel's<br />

Chapel, Canterbury Cathedral: their technique, condition <strong>and</strong> environment reassessed.<br />

Arbeitslufte zur Denkmalpfiege in Niedersachsen (1 1):143.<br />

7. See Rickerby, S. 1990. Kempley: A technical examination of the Romanesque<br />

wall paintings. In Early Medieval Wall <strong>Painting</strong> <strong>and</strong> Painted Sculpture in Engl<strong>and</strong>.<br />

Ed. S. Cather, et al. Oxford: BAR British Series 219, 249-261; Howard, H. 1992.<br />

All Saints' Church, Witley, Surrey: scientific examination of the Romanesque<br />

wall paintings. Unpublished report. London: Conservation of Wall <strong>Painting</strong> Department,<br />

Courtauld Institute of Art, University of London; Cather <strong>and</strong> Howard,<br />

op. cit. (note 6).<br />

8. These patches were mapped in the 1980s by the Canterbury Cathedral Wallpainting<br />

Workshop.<br />

9. Kupfer, M. 1986. Les fresques romanes de Vicq: E tude technique. Bulletin Monumental<br />

144:98-132.<br />

10. Park, op. cit., 40.<br />

11. Mairinger, F., <strong>and</strong> M. Schreiner. 1986. Deterioration <strong>and</strong> preservation of Carolingian<br />

<strong>and</strong> medieval wall paintings in the Mlistair Convent. Part II: <strong>Materials</strong><br />

<strong>and</strong> rendering of the Carolingian wall paintings. In Case Studies in the Conserl!ation<br />

of Stone <strong>and</strong> Wall <strong>Painting</strong>s. Ed. N. S. Bronmlelle, et al. London: International<br />

Institute fo r the Conservation of Historic <strong>and</strong> Artistic Works, 195-96.<br />

12. Ultramarine was the mineral blue normally employed in Romanesque wall<br />

paintings, though azurite has been identified at Kempley (Rickerby, op. cit., 256)<br />

<strong>and</strong> Marslet, Denmark (Graebe, H., K. Trampedach, <strong>and</strong> M. Jensen. 1986. Kalkmalerierne<br />

i Marslet Kirke. National museete Arbejdsmark 23:164-82).<br />

13. Matteini, M., <strong>and</strong> A. Moles. 1994. Mural paintings in Wunstorf-Idensen: Chemical<br />

investigations on paintings, materials <strong>and</strong> techniques. Arbeitschifte zur Denkmalpfiege<br />

in Niedersachsen 11:85-86.<br />

14. Cather <strong>and</strong> Howard, op. cit., 144-46.<br />

15. Matteini <strong>and</strong> Moles, op. cit., 85.<br />

16. Cennino Cennini. II Libro Dell'Arte: The Craftsman's H<strong>and</strong>book. 1933. Translated<br />

by D. V. Thompson. New Haven, 26.<br />

17. The acceleration of the rate of fading over time, due to an increased proportion<br />

of white, has been shown experimentally fo r a lake mixed with white, <strong>and</strong> for<br />

a lake used as a glaze over a white ground. See Johnston-Feller, R., <strong>and</strong> C. W<br />

Bailie. 1982. An analysis of the optics of paint glazes: fa ding. In Science <strong>and</strong> Technology<br />

in the Service of Conserl/ation. Ed. N. S. Brommelle, et al. London: International<br />

Institute for Conservation of Historic <strong>and</strong> Artistic Works, 180-85.<br />

18. Faye, G. H., P. G. Manning, <strong>and</strong> E. H. Nickel. 1968. The polarized optical absorption<br />

spectra of tourmaline, cordierite, chloritoid <strong>and</strong> vivianite: ferrous-ferric<br />

electronic interaction as a source of pleochroism. The American Mineralogist (53):<br />

1174-201.<br />

19. I am grateful to Dr. G. Cressey of the Natural History Museum for carrying out<br />

analysis by X-ray diffraction using a Debye-Scherrer camera. The possible ad-<br />

Howard 103


ditional presence of metavivianite (a polymorph of vivianite) was indicated by<br />

this analysis.<br />

20. Richter, E. L. 1988. Seltene pigmente im Mittelalter. Zietschrift Jur Kunsttechnologie<br />

und Konservierung 2 (1):171-77.<br />

21. J. Darrah. 1987. Personal communication. Victoria & Albert Museum, London.<br />

22. Mineral deposits are fo und in Engl<strong>and</strong>, France, Germany, Italy, Portugal, Serbia,<br />

<strong>and</strong> the Ukraine, as well as in the United States <strong>and</strong> elsewhere. I am grateful to<br />

Mr. Peter T<strong>and</strong>y of the Natural History Museum, London for providing this<br />

information <strong>and</strong> for guiding me through the extensive range of samples in the<br />

collection. I am also indebted to Dr. Brian Young of the British Archaeological<br />

Survey for his enthusiastic interest <strong>and</strong> informed opinion on the provenance of<br />

the mineral.<br />

23. Watson, T. L. 1918. The color change in vivianite <strong>and</strong> its effect on the optical<br />

properties. The American Mineralogist 3 (8): 159-61.<br />

24. P. T<strong>and</strong>y. 1994. Personal corrununication. Natural History Museum, London.<br />

25. Following initial exposure of the colourless mineral, oxidation of one of the<br />

paired ferrous ions may occur at the surface <strong>and</strong> along crystal cleavages. The<br />

resulting intervalency charge transfer between the paired ions (now Fe2+ <strong>and</strong><br />

Fe3+) produces a blue color. Cressey, G. 1994. Personal communication. Natural<br />

History Museum, London.<br />

26. Hanzel, D., W Meisel, 0. Hanzel, <strong>and</strong> P. Glitlich. 1990. Mossbauer effect study<br />

of the oxidation of vivianite. Solid State Communications 76 (3):307-10.<br />

27. This information was kindly provided by Mr. David Perry <strong>and</strong> published by<br />

Park, op. cit.<br />

28. Oakeshott, W. 1981. The paintings of the Holy Sepulchre Chapel. Winchester<br />

Cathedral Record 50, plate 98a.<br />

29. Welford, P. 1991. Investigation of the phenomenon of dark flesh areas in English<br />

medieval wall painting. Diploma diss. London: Conservation of Wall <strong>Painting</strong><br />

Department, Courtauld Institute of Art, University of London.<br />

104<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

It is rare for ancient external paint<br />

to survive the English climate, particularly<br />

in the original location of<br />

the paint. It is vital that evidence of<br />

color be documented, <strong>and</strong> studied<br />

current recognition of this need has<br />

created opportunities to survey<br />

the fabric of important historic<br />

buildings for evidence of paint.<br />

Analysis of paint fragments fo und on<br />

the richly sculpted west front of Exeter<br />

Cathedral has revealed evidence<br />

of a magnificent polychromy on<br />

both architecture <strong>and</strong> sculpture, giving<br />

vital information on fourteenth<strong>and</strong><br />

fifteenth-century materials <strong>and</strong><br />

techniques. A similar investigation on<br />

the west front of Salisbury Cathedral<br />

has just commenced. An initial look<br />

at surviving evidence of polychromy<br />

enables some comparisons to be<br />

made, in addition to other relevant<br />

examples.<br />

The Polychromy of Exeter <strong>and</strong> Salisbury Cathedrals:<br />

A Preliminary Comparison<br />

Eddie Sinclair<br />

10 Park Street<br />

Crediton<br />

Devon EX17 3EQ<br />

United Kingdom<br />

Introduction<br />

There has been in recent years an increasing awareness of the historic importance<br />

<strong>and</strong> vital role of color in medieval architecture, as there is at last<br />

recognition that our ancient buildings, <strong>and</strong> not just the artifacts within them,<br />

were painted as an integral part of their overall design.<br />

The Reformation in Engl<strong>and</strong> in the sixteenth century resulted in the destruction<br />

or obliteration of much polychromy. Where any evidence of color still<br />

exists it is frequently only fragmentary, but those fragments retain much valuable<br />

information.<br />

Conservation work on the west-front image screen of Exeter Cathedral carried<br />

out from 1979 to 1984 revealed much evidence of a rich polychromy<br />

in the form of paint fragments surviving in the most sheltered corners of<br />

both architecture <strong>and</strong> sculpture. Although the study of this polychromy is<br />

discussed in detail elsewhere, the establishment of a large archive of paint<br />

samples has created an invaluable resource that can be used with other emerging<br />

fragmentary evidence (1, 2). As well as providing useful reference material,<br />

these samples contain much information not yet explored that future similar<br />

proj ects may yet discover (3) .<br />

The work on the Exeter polychromy, while not the first English cathedral to<br />

receive such attention, was on an unprecedented scale (4). As conservation<br />

work is carried out on more cathedrals <strong>and</strong> other important buildings in<br />

Engl<strong>and</strong>, investigations into the color become a vital element of the work<br />

undertaken. Similar investigations in other European countries have provided<br />

a wealth of detailed information, summarized by Rossi-Manaresi <strong>and</strong> more<br />

recently by Brodrick (5, 6). Each cathedral adds its own invaluable evidence<br />

to the complex picture of materials, techniques, <strong>and</strong> workshop practice of<br />

medieval times.<br />

With conservation work due to commence on the west front of Salisbury<br />

Cathedral in winter 1994, an inspection fo r polychromy was requested. A<br />

preliminary investigation with access to only part of the facade has shown<br />

evidence of color. Some analysis has been carried out to date, with further<br />

work anticipated in 1995.<br />

Exeter Cathedral west front<br />

Figure 1. Exeter Cathedral, west front,<br />

showing image screen. Photograph courtesy of<br />

the Cathedral of St. Peter in Exeter.<br />

The west-front image screen of Exeter Cathedral dates from the fourteenth<strong>and</strong><br />

fifteenth-centuries, although much of the crenellated parapet <strong>and</strong> some<br />

portions of the architectural elements, along with six sculptures <strong>and</strong> the heads<br />

of four others, have been replaced over the years (Fig. 1). Most of the fabric<br />

is built from Beer stone, a local compact, close-grained limestone.<br />

Much vital information on the materials <strong>and</strong> techniques of the medieval<br />

period are provided in the Exeter Cathedral fabric accounts for the period<br />

1279-1353 (7, 8). Although the fabric accounts are missing for most of the<br />

period during which the west front was being constructed <strong>and</strong> decorated,<br />

they provide a wealth of information relating to the polychromed bosses of<br />

the high vault which, combined with an examination of the fabric discussed,<br />

Sinclair 105


enhances our picture of the working process (9). This, too, can be of immense<br />

value for comparative purposes with the west front.<br />

Much of the west front of Salisbury Cathedral (Fig. 2) was restored in the<br />

nineteenth century, when most of the empty niches were filled with new<br />

sculpture. Only portions of eight sculptures derive from the original scheme<br />

of 1245-1260, which may never have been completed. The majority of architectural<br />

<strong>and</strong> all of the ornamental <strong>and</strong> sculptural stonework on Salisbury's<br />

west front is constructed of the local Chilmark stone, a s<strong>and</strong>y limestone. With<br />

much of the stone weathered <strong>and</strong> covered in lichen, it is not always immediately<br />

apparent whether the fabric is original or replacement. Unfortunately,<br />

while there is documentation for much of the restoration work, there appears<br />

to be none recording the original fabric.<br />

Figure 2. Salisbury Cathedral, west front.<br />

Photograph courtesy rif the Courtauld Institute<br />

rif Art.<br />

A preliminary inspection for paint on the facade revealed evidence only on<br />

the moldings above the lower register figures on the north face of the north<br />

turret, the most protected part of the west front (Fig. 3). More evidence of<br />

paint may become apparent as cleaning commences on the west front in 1995.<br />

The inspection for paint extended into the sheltered central porch where,<br />

again, much of the fabric is from the nineteenth century. On the tympanum,<br />

however, the medieval surfaces that have remained are in areas still thickly<br />

painted, with ample evidence of a magnificent polychromy (Fig. 4) . The only<br />

original carving is that of four heads at the apex of the tympanum; they are<br />

particularly well protected <strong>and</strong> retain paint on their hair, beards, <strong>and</strong> eyes.<br />

An examination of the roof bosses in the west walk of Salisbury cloister,<br />

which date from 1263-1270, also reveals evidence of color though here the<br />

paint is extensive, with some bosses retaining almost all their color. While<br />

their polychromy cannot be seen as part of the same scheme as that of the<br />

west front <strong>and</strong> central porch it serves as valuable additional reference material,<br />

as well as an important surviving fact in its own right.<br />

Figure 3. Salisbury Cathedral, north face rif<br />

north turret, lower register. Evidence rif paint<br />

was found 011 the moldings above the figures.<br />

Photograph courtesy of Dean <strong>and</strong> Chapter,<br />

Salisbury Cathedral.<br />

Figure 4. Salisbury Cathedral central porch,<br />

ly,upauul11. Photograph courtesy rif Dean <strong>and</strong><br />

Chapter, Salisbury Cathedral.<br />

Paint samples were taken from all three locations at Salisbury to analyze the<br />

type of pigments <strong>and</strong> binding media used by the medieval painters. These<br />

analyses are still in progress <strong>and</strong> much remains to be done. Further evidence<br />

may yet come to light that will alter current perception of the preliminary<br />

results.<br />

Salisbury polychrorny<br />

Although the scarcity of evidence of paint on the west-front facade makes it<br />

rather premature to talk of differences or similarities, a preliminary examination<br />

of the paint from all three locations immediately reveals a difference<br />

in technique (Fig. 5).<br />

Those west-front samples examined thus far show the existence of a thick<br />

white ground with a single, colored layer on top (Fig. Sa), while those from<br />

the cloister bosses show a thin, translucent white ground with one or two<br />

colored layers on top <strong>and</strong> extensive use of gilding, which may also bear painted<br />

decoration (Fig. Sb). By comparison those from the central porch display<br />

a more complex structure (Fig. Sc), with some samples exhibiting up to fifteen<br />

layers <strong>and</strong> several repaintings, often resulting in a change of color, with the<br />

use of a red earth primer, liberal applications of white lead <strong>and</strong> a broad range<br />

of colors (Plate 21). Presumably the paint was reapplied whenever it started<br />

to look shabby; there are three layers of gilding on the beard of one of the<br />

carved heads. Samples examined thus far from the cloister bosses show no<br />

evidence of repainting.<br />

Although more paint may yet be discovered on Salisbury west front, current<br />

results indicate several possibilities. If the sculptural scheme was left unfinished,<br />

perhaps only isolated, completed areas were painted. Alternatively, the scheme<br />

may have been completed, but left largely unpainted due to lack of funds.<br />

Weathering <strong>and</strong> human intervention may have caused loss of most of the<br />

106<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Sa. Location : Salisbury west front,<br />

north face of north turret.<br />

Sb. Location : Salisbury, cloister boss.<br />

2.Smm<br />

I. White ground<br />

2. Iron-oxide red<br />

,, <br />

I<br />

lmm<br />

1. White ground<br />

2. Yellow ochre, gold size?<br />

3. Gold leaf<br />

4. Black decoration<br />

Sc. Location : Salisbury central porch,<br />

tym panum .<br />

Sd. Location : Exeter west front,<br />

drapery of king.<br />

2.Smm<br />

. <br />

I<br />

l.Smm<br />

Multi-layered sample, with<br />

1. Red earth primer,<br />

2. White lead undercoat,<br />

3. Vermilion top layer <strong>and</strong> several<br />

repaintings of primer, white lead<br />

<strong>and</strong> vermilion.<br />

1. Pi nk primer<br />

2. White lead<br />

3. Vermilion<br />

4. Black cement layer<br />

Figure 5. Paint cross sections: a prelimillary view.<br />

paint, but protected corners usually retain odd fragments, as was fo und on<br />

the north return.<br />

At this point in the investigation, the discovery of the existence of paint in<br />

just one area on Salisbury west front could be the result of a poor technique.<br />

The careful preparation of the stone with appropriate sealant, primer, <strong>and</strong><br />

ground played a major role in the durability of the paint layers above. The<br />

initial examination of the north turret samples at Salisbury suggests that the<br />

white ground is of a chalk <strong>and</strong> gesso type, <strong>and</strong> not very tough, though it is<br />

thick <strong>and</strong> would have provided a smooth surface for the paint. Its softness<br />

suggests a loss of medium, though the lack of evidence so far for paint elsewhere<br />

may be due to a poor choice of medium.<br />

Visual analysis, largely through the study of cross sections at this stage, reveals<br />

a typical medieval palette, with a liberal use of costly exotic pigments, particularly<br />

in the central porch. Pigments here include vermilion, verdigris <strong>and</strong><br />

copper resinate greens, black (probably lamp black), red <strong>and</strong> white lead, a dark<br />

blue that appears to be indigo, red <strong>and</strong> yellow ochre, <strong>and</strong> gold leaf. The<br />

cloister pigments include red <strong>and</strong> yellow iron oxide, black, gold leaf, a cool<br />

blue-green <strong>and</strong> a calcium carbonate white. No lead pigments have yet been<br />

identified here. Only two pigments, red <strong>and</strong> yellow ochre (with a chalky<br />

white layer below), have been identified on the west-front facade.<br />

Exeter polychromy<br />

That any paint survives at Exeter, in spite of several major cleaning programs,<br />

must be due to a careful selection of superior quality materials (Fig. Sd) with<br />

abundant use of durable red <strong>and</strong> white lead. Most samples are well bound,<br />

though migrating salts <strong>and</strong> some loss of medium cause some paint to delaminate.<br />

A typical sample from Exeter west front has-on top of an invisible sealanta<br />

red earth primer followed by a pale pink primer consisting of iron-oxide<br />

red, chalk, white <strong>and</strong> red lead, all in varying proportions (Plate 22). On top<br />

Sinclair 107


of this, a lead white undercoat provides a tough resistant layer, often also used<br />

to enhance the colored layer above, which may in turn be covered with a<br />

glaze color.<br />

<strong>Techniques</strong>: Salisbury <strong>and</strong> Exeter<br />

Although the structure of the Exeter west-front samples in no way resembles<br />

that of the samples from Salisbury west front, there is a greater similarity to<br />

those samples from the central porch, where a similar range of pigments are<br />

combined with a complex layer structure. The presence of white lead in the<br />

porch at Salisbury will have played some part in the survival of the paint<br />

there.<br />

The use of three different techniques, apparent even at this early stage of the<br />

Salisbury investigation, with the indication of two different techniques<br />

operating on the west front, is in keeping with evidence found on other<br />

European cathedrals of a considerable variation in technique existing on<br />

large-scale schemes (10, 11).<br />

The richest colors <strong>and</strong> effects at Salisbury may have been reserved for the<br />

main doorway, the gr<strong>and</strong> ceremonial entry into the cathedral (12). If less<br />

expensive preparations were used elsewhere, as is suggested by the north turret<br />

evidence, this could explain their disappearance. At Exeter, the evidence of a<br />

rich repainting around the main entry also suggests the important role played<br />

by color in the liturgy of a great church, where the main entry was not<br />

considered complete without its polychromy <strong>and</strong> was renewed more frequently<br />

than elsewhere (13).<br />

The medium of weathered exterior paint has naturally deteriorated, making<br />

reliable analysis extremely problematic. Both linseed oil <strong>and</strong> egg tempera appear<br />

to be present in the samples from Exeter west front, either as an emulsion<br />

or to temper different layers. While no media analysis has yet been carried<br />

out on the Salisbury west-front samples, those from the porch appear to be<br />

in an egg tempera medium <strong>and</strong> those from the cloister in linseed oil. At<br />

Salisbury, however, the usual problems of media analysis are fu rther complicated<br />

by the presence of lichens, algae, bacteria, <strong>and</strong> also the identification of<br />

wax. A protective coating applied in a previous restoration has left a brownyellow<br />

wrinkled skin over much of the stonework, fu rther complicating anal­<br />

YSlS.<br />

With most surviving paint at Exeter found in deep crevices, much of the<br />

detail of the finely carved Beer stone must have been lost, as layers of paint<br />

fill corners of crockets, nostrils, or finely carved hair. The skill of the medieval<br />

painter was to use his craft not merely to paint the carved form, but to<br />

embellish <strong>and</strong> further enhance it. At times this must have entailed repainting<br />

obscured detail by redefining highlights <strong>and</strong> shadows to carved hair.<br />

Similarly, the moldings around the tympanum <strong>and</strong> the carved heads at the<br />

apex in the central porch at Salisbury must have needed picking out in fresh<br />

color, as the corners lost their clarity over the years, with a build-up of paint<br />

particularly thick where different planes meet <strong>and</strong> overlapping of color occurs.<br />

Where the stone is sound on the west aisle cloister bosses, the carving remains<br />

crisp, as there is no build-up of paint layers to clog up the detail.<br />

The painter did not rely on the play of light alone to give form to the<br />

sculptures, but added depth to hollows with shading or highlights to high<br />

points where paint rarely survives on an exterior. Traces of black in the inner<br />

moldings of the orders around the tympanum at Salisbury show how hollows<br />

were further emphasized by the choice of dark colors. Carvings are thrown<br />

into relief by the background color from which they emerge. Thus, in the<br />

porches at Exeter, the foliage around the doorways, now mostly bare stone,<br />

is defined by the thick traces of indigo still surviving in the corners of the<br />

background.<br />

The painter would add details to carving-even if they would never be seen<br />

by the ordinary viewer-such as the delicate red outline defining the most<br />

108<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


sheltered eye on the southernmost carved head in the porch at Salisbury,<br />

remarkably similar in style to that fo und on one of the polychromed heads<br />

on the cloister bosses.<br />

No metal leaf was found on the facade at Exeter, although the fabric accounts<br />

list the purchase at Christmas 1341 of " gold, silver, <strong>and</strong> various colors for<br />

painting the image of the blessed Peter" high up in the gable of the west<br />

front (14). Instead there were several occurrences of orpiment where gold<br />

might have been expected, such as on crowns <strong>and</strong> hats; presumably this was<br />

used elsewhere for reasons of economy.<br />

A reference to "the painting of the bishop in the gable," makes it clear that<br />

the final costly colors <strong>and</strong> foils were applied in situ (15). As for primings <strong>and</strong><br />

earlier preparations, these may have been carried out in the shelter of the<br />

workshop, as was the case with the interior bosses in the crossing, where red<br />

lead priming can be seen disappearing into the masonry joints under the<br />

medieval mortar (16). However, recent developments have shown that the<br />

quire bosses must have been totally painted in situ (17). Both practices, therefore,<br />

were possible <strong>and</strong> were operating at Exeter. Workshop practice at Salisbury<br />

is yet to be explored, though the presence of dirt beneath the lowest<br />

paint layer on the cloister samples suggests that they were left unpainted for<br />

some time.<br />

Conclusion<br />

While it is too soon to draw conclusions about surviving external color at<br />

Salisbury, the discovery of its very existence is important, placed in the context<br />

of a tradition that was soon to disappear from English ecclesiastical buildings,<br />

although it lingered a little longer in a secular context (18). The search for<br />

further evidence of polychromy will continue at Salisbury, along with a more<br />

detailed study of identified paint traces.<br />

It is hoped that evidence of polychromy from Exeter, <strong>and</strong> now Salisbury,<br />

however fragmentary, can be seen as part of a European Gothic tradition. It<br />

would be interesting to see if a more detailed study reveals any differences in<br />

style to distinguish English architectural polychromy from that of its European<br />

counterparts.<br />

Acknowledgments<br />

I would like to thank the deans <strong>and</strong> chapters of Salisbury <strong>and</strong> Exeter Cathedrals for<br />

supporting this research, <strong>and</strong> all individuals nam.ed in the notes. Recent media analysis<br />

for Salisbury Cathedral was carried out by N. Kh<strong>and</strong>ekar of the Hamilton Kerr<br />

Institute. The research is being carried out with the technical support of the Earth<br />

Resources Centre, Exeter University.<br />

Notes<br />

1. Sinclair, E. 1991. The west front polychromy. In Medieval Art <strong>and</strong> Architecture at<br />

Exeter Cathedral. Ed. F. Kelly. British Archaeological Association, 116-33.<br />

2. Sinclair, E. 1992. Exeter Cathedral: Exterior polychromy. In The Conservator as<br />

Art Historian. Ed. A. Hulbert, et al. London: United Kingdom Institute for Conservation,<br />

7-14.<br />

3. The author, a practicing conservator, carried out this research, with invaluable<br />

help <strong>and</strong> tuition from Drs. Ashok Roy <strong>and</strong> Raymond White of The National<br />

Gallery, Josephine Darrah of the Victoria & Albert Museum, Peter Mactaggart,<br />

<strong>and</strong> Anna Hulbert. However, access to sophisticated analytical equipment <strong>and</strong><br />

techniques was limited, <strong>and</strong> most results are based on the study of cross sections<br />

<strong>and</strong> dispersions.<br />

4. Conservation work on Wells Cathedral west front, carried out from 1974-1 986,<br />

uncovered extensive evidence of polychromy that will be reported on in Sampson,<br />

J. Wells Cathedral: West front archaeology <strong>and</strong> conservation. Forthcoming.<br />

5. Rossi-Manaresi, R. 1987. Considerazioni tecniche sulla scultura monumentale<br />

policromata, romanica e gotica. Bollettino d' Arte (41): 173-86.<br />

6. Brodrick, A. 1993. <strong>Painting</strong> techniques of early medieval sculpture. In Romanesque<br />

Stone Sculpture from Medieval Engl<strong>and</strong>. 18-27.<br />

Sinclair 109


7. The Accounts of the Fabric of Exeter Cathedral, 1279- 1353. 1981. Part I, 1279-<br />

1326. Ed. A. Erskine. Devon <strong>and</strong> Cornwall Record Society, 24.<br />

8. The Accounts of the Fabric of Exeter Cathedral, 1279- 1353. 1983. Part II, 1328-<br />

1353. Ed. A. Erskine. Devon <strong>and</strong> Cornwall Record Society, 26.<br />

9. Hulbert, A. 1991. An examination of the polychromy of Exeter Cathedral roof<br />

bosses, <strong>and</strong> its documentation. In Medieval Art <strong>and</strong> Architecture at Exeter Cathedral.<br />

Ed. F. Kelly. British Archaeological Association, 188-98.<br />

10. Rossi-Manaresi, op. cit.<br />

11. Sampson, J. Personal communication. At Wells, three different techniques were<br />

identified by Dr. Roy (National Gallery, London) with the use of either white<br />

lead or gypsum (?) as a ground. The niche in which The Coronation if the Virgin<br />

is situated, above the main west door, retains evidence of the more exotic pigments<br />

on a lead white ground.<br />

12. In a search fo r polychromy during recent work on the west front of Bath Abbey,<br />

Jerry Sampson found that, apart from one isolated fragment elsewhere, the only<br />

paint was in the west door. This may be the result of a poor technique resulting<br />

in almost total disappearance of color, but it suggests again the importance of<br />

the main doorway. See Sampson, J. 1992. Bath Abbey West Front: The History of<br />

the Restorations. Private report. Bath City Council <strong>and</strong> English Heritage.<br />

13. Rossi-Manaresi, R., <strong>and</strong> A. Tucci. 1984. The polychromy of the portals of the<br />

gothic cathedral of Bourges. [COM Committee fo r COl1seYilatiol'l preprints, 7th Triennial<br />

Meeting, Copenhagen. During construction of the west front, the south<br />

portal was used as the main entry. Here, twelfth-century carvings, unpainted until<br />

incorporated into the south portal in 1225, were then polychromed in situ,<br />

forming a temporary main entrance. Bourges Cathedral is contemporary with<br />

Salisbury Cathedral.<br />

14. Erskine, 1983. Op. cit., 269.<br />

15. Ibid., 270.<br />

16. Hulbert, op. cit., 192.<br />

17. Hulbert, A. 1994. Personal cOl11l11unication.<br />

18. Sinclair, E. 1990. Appendix: Sampling <strong>and</strong> analysis of paint traces from the exterior<br />

elevations of the front block. In Exeter Guildhall. S. R. Blaylock. Exeter:<br />

Devon Archaeological Society Proceedings (48): 174-6. Research on the exterior<br />

of Exeter Guildhall revealed a rich polychromy that included azurite, vernlilion,<br />

<strong>and</strong> gilding, with the earliest layers dating from 1593 to 1594.<br />

110<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

Extensive examination of Andrea<br />

Mantegna's Adoration of the Magi at<br />

the J. Paul Getty Museum reveals<br />

Mantegna's unusual technique of<br />

painting. Although this painting was<br />

generally described as painted in<br />

tempera or oil, analysis has revealed<br />

the medium to be distemper (animal<br />

glue). The rationale for this technique<br />

is explained here in terms of<br />

aesthetic <strong>and</strong> environmental constraints<br />

of the fifteenth century as<br />

well as the art-historical context of<br />

the period.<br />

Andrea Mantegna's Adoration of the Magi<br />

Andrea Rothe<br />

Department of <strong>Painting</strong>s Conservation<br />

The J. Paul Getty Museum.<br />

17985 Pacific Coast Highway<br />

Malibu, California<br />

USA<br />

Introduction<br />

In 1985 the J. Paul Getty Museum acquired the Adoration oj the Magi by<br />

Andrea Mantegna (1431-1 506) (Plate 23) . Nothing was known about the<br />

painting until the early nineteenth century, when it was thought to have been<br />

brought to Engl<strong>and</strong> by Alex<strong>and</strong>er Baring, first Lord Ashburton (1). It was<br />

first shown publicly in London at the Royal Academy in 1871, <strong>and</strong> then at<br />

New Gallery in 1894; but with few other exceptions, the painting was not<br />

readily accessible. This might explain why it was sometimes confused with a<br />

nineteenth-century copy, now in the Johnson collection in Philadelphia (2) .<br />

The Adoration was shown again in London in 1981 in the exhibition Splendours<br />

of the Gonzaga, <strong>and</strong> has since been widely accepted as a late work by<br />

Mantegna, dated circa 1495-1505. The composition of this popular subject<br />

must have been much admired, for at least seven copies by other artists survive<br />

(3).<br />

The horizontal composition with five half-length figures placed tightly<br />

around the Christ Child bears resemblance to other works by Mantegna, such<br />

as The Virgin <strong>and</strong> Child with Saints in the Galleria Sabauda in Turin <strong>and</strong> The<br />

Presentation in the Temple in the Staatliche Museen zu Berlin. Mantegna seems<br />

to have been the initiator of this type of composition, which inspired Giovanni<br />

Bellini, Cima da Conegliano, <strong>and</strong> Vincenzo Catena, among others. Remarkable<br />

are the descriptive features <strong>and</strong> the distinctive skin color of each<br />

individual figure. Noteworthy is Mantegna's foreshortening of the Christ<br />

Child's right leg, which is reminiscent of the Lamentation in the Brera in<br />

Milan. Most of the execution adheres to Mantegna's original composition,<br />

with the exception of the faces of the Virgin <strong>and</strong> the Christ Child, in which<br />

the X-radiograph shows Mantegna's original reworking or pentimenti<br />

(Fig. 1).<br />

Analysis of painting technique used in the Adoration<br />

In order to properly restore the Adoration, extensive research on the unusual<br />

technique of this painting was necessary. Although the Getty Adoration was<br />

generally described as painted in tempera or oil, recent analysis carried out<br />

by the Getty Conservation Institute has identified the medium as distemper<br />

(animal glue) (4) . Pioneering research done by John Mills of the National<br />

Gallery in London <strong>and</strong> other experts in the 1970s drew attention to the fact<br />

that this medium was widely used during the Gothic <strong>and</strong> Renaissance periods<br />

(5). Within the next decade more accurate methods of analysis were developed<br />

(6). The principal reason fo r the lack of more information has been the<br />

difficulty in analytically distinguishing egg from glue, as both are complex<br />

proteins. The analytical results can also be unreliable because infusions of glues<br />

from later relinings <strong>and</strong> consolidants make it difficult to determine whether<br />

they are part of the original medium.<br />

Throughout the history of painting, the use of glue as a medium has been<br />

quite common, yet very often paintings in this medium are classified as tempera<br />

paintings, such as the mummy portraits of the late Fayum period (third<br />

century C.E.), which were no longer painted in encaustic, but probably with<br />

a glue medium (7). In Germany <strong>and</strong> particularly in the Netherl<strong>and</strong>s of the<br />

fifteenth <strong>and</strong> sixteenth century, thous<strong>and</strong>s of distemper paintings were pro-<br />

Rothe 111


Figure 1. X-radiograph of Andrea Mantegna's The Adoration of the Magi. The]. Paul Getty<br />

Museum, Malibu.<br />

FIgure 2. Dierie Bouts, The Annunciation.<br />

Distemper on linen, 90 X 74.5 em. The].<br />

Paul Getty Museum, Malibu (85.PA.24).<br />

duced, including examples by great artists such as Albrecht Durer <strong>and</strong> Pieter<br />

Brueghel (8) . Some of these that have survived relatively untouched show us<br />

the beauty <strong>and</strong> brilliance of this medium. Well-preserved examples of this<br />

technique include three paintings by Dieric Bouts: (a) the Annunciation in the<br />

collection of the J. Paul Getty Museum, Malibu (Fig. 2); (b) the Resurrection<br />

at the Norton Simon Museum, Pasadena; <strong>and</strong> (c) the Deposition in the National<br />

Gallery, London (9). Many of these paintings, however, have either been<br />

destroyed or severely damaged <strong>and</strong> altered. The glue renders the paint film<br />

brittle <strong>and</strong> readily damaged by water; because of the glue's hygroscopic nature,<br />

it attracts dust <strong>and</strong> c<strong>and</strong>le soot. Consequently, most of these paintings suffered<br />

the fate of being varnished in order to "liven up" the colors, thus becoming<br />

oil or varnish paintings by absorption. Most of them were destroyed because<br />

they were so fragile, <strong>and</strong> less than 100 of the above-mentioned Netherl<strong>and</strong>ish<br />

paintings have survived (10).<br />

In Italy, however, Andrea Mantegna was the only artist to make extensive use<br />

of this technique. Approximately thirty of his distemper paintings have survived.<br />

Mantegna, with some of his fellow painters north of the Alps, must have been<br />

intrigued with the effect of light, or the lack of it, on his paintings. In order<br />

to visualize the effect of different media in the same light, we need only look<br />

at an illuminated manuscript in the subdued light of a modern showcase to<br />

marvel at the glowing colors (11). By contrast, it would be practically impossible<br />

to view an oil painting under the same light conditions because of<br />

the deep saturation of the pigments <strong>and</strong> the different refractive index. Mantegna<br />

probably faced the same dilemma of wanting to create paintings with<br />

incredible detail <strong>and</strong> luminosity that would not lose their visual power due<br />

to lack of light <strong>and</strong> the distracting reflections caused by a varnished surface.<br />

Because we live today in a technically advanced age in which we can regulate<br />

the light we need, it is difficult to imagine relying on c<strong>and</strong>les <strong>and</strong> torches.<br />

Yet, until the wide distribution of electricity less than one hundred years ago,<br />

light in a closed environment was scarce. Until the Renaissance, many palaces<br />

<strong>and</strong> buildings in Italy had only a few small windows; many of Mantegna's<br />

paintings were probably destined to be hung in a private bedroom or chapel<br />

in the fortresslike Ducal Palace in Mantua, thus a technique that would not<br />

112<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


create surface gloss <strong>and</strong> could be viewed in subdued light was ideal. Of all<br />

the techniques, only fresco <strong>and</strong> distemper have these characteristics <strong>and</strong> only<br />

distemper, which is more brilliant, gives the artist a limitless time to work<br />

(12). Distemper paintings are also light <strong>and</strong> portable. A letter by Mantegna<br />

to Ludovico Gonzaga, Marquis of Mantua, suggested painting portraits on<br />

fme linen in this technique, as they can be rolled up on a dowel <strong>and</strong> easily<br />

shipped (13). These advantages <strong>and</strong> characteristics might explain why Mantegna,<br />

the absolute perfectionist, had such a predilection for distemper.<br />

The animal glue for distemper painting was made from either skins or parchment,<br />

preferably sheep or goat (14). Glue made out of fish bones was also<br />

used; it has been identified in the Cult of Cybele by Mantegna at the National<br />

Gallery in London (15). The glues were prepared <strong>and</strong> dried into tablets in<br />

the winter to keep them from growing mold. When the glue was needed, it<br />

was first soaked in water <strong>and</strong> then dissolved in a hot water bath.<br />

<strong>Historical</strong> methods <strong>and</strong> recipes<br />

A rather succinct description of the practice of this distemper technique is<br />

given in the so-called Eraclius manuscript. In recipe XXVI of De Coloribus<br />

et Artibus Romanorum, the text states (16):<br />

If you wish to paint on linen cloth, <strong>and</strong> lay gold upon it, prepare it thus:<br />

Take parchment, or clippings if parchment, <strong>and</strong> put them in a jar with<br />

water, which must be placed over the f ire <strong>and</strong> made to boil as before directed;<br />

then dip a cloth into it, take it out immediately, <strong>and</strong> stretch it out on a<br />

wet panel <strong>and</strong> let it dry. Then burnish or polish it all over with a glass<br />

muller, <strong>and</strong> stretch it out, fastening it on to a wooden frame with the<br />

thread. You may then paint upon it with colours distempered with size,<br />

or egg, or gum.<br />

The recipe describes how the cloth was treated to make it less absorbent, so<br />

that the colors would not be soaked into the cloth <strong>and</strong> spread out. Only such<br />

a pretreated cloth could be painted on with some detail. It is this same property<br />

that is alluded to in Jehan Ie Begue's 1431 recipe compilation Experimenta<br />

de Coloribus (17):<br />

In Engl<strong>and</strong> the painters work with these waters upon closely woven cloths<br />

[sindone?], wetted with gum water, made with gum arabic <strong>and</strong> then dried,<br />

<strong>and</strong> afterwards stretched out on the floor if the soler ... <strong>and</strong> the painters<br />

... paint upon them figures stories <strong>and</strong> other things. And because these<br />

cloths lie stretched out on a flat suiface, the coloured waters do not flow or<br />

spread in painting upon them . . . the touches of the paint brush made<br />

with these waters do not spread, because the gum with which, as already<br />

mentioned, the cloth is wetted, prevents their spreading . ...<br />

Glue gels at room temperature, however, making it difficult to use. The difficulties<br />

of keeping the medium in a workable consistency are described in<br />

the De Coloribus Diversis Modis Tractatur in Sequentibus, written in 1398 by<br />

Johannes Alcherius. In a recipe given to Alcherius by the Flemish painter<br />

Jacob Coene, instructions are provided for painting <strong>and</strong> "laying gold on<br />

parchment, paper, linen cloth, sindone [very fme linen], <strong>and</strong> on primed wooden<br />

panels. The recipe describes the making of glue out of parchment or cutting<br />

of fine leather: "Lastly, let the size, or sized water be warm; I say warm, lest<br />

it may be conglutinated, ... when it is cold it will be congealed like jelly ..."<br />

(18).<br />

Alcherius continues, "Moreover when using a paintbrush, the colour may be<br />

held in the h<strong>and</strong>, which by its warmth or heat will not allow it to congeal<br />

... And in painting with a pen, as well as with a paint brush, it is a good<br />

thing to keep the colour over a slow fire of charcoal at such a warmth, that<br />

it may not congeal, but may remain liquid" (19).<br />

The recipe stresses the fact that-in contrast to painting on panel-when<br />

painting on cloth or Tiichlein, the artist should apply paint in several layers<br />

(20):<br />

Rothe 113


On cloth <strong>and</strong> sindone it is more necessary that this colour should be layd<br />

on twice, wile tempered with size, bi f ore it is put on Jo r the last coat<br />

tempered with white oj egg. And this is because sindone <strong>and</strong> cloth, owing<br />

to their porosity, are too absorbent, flowing, flexible, <strong>and</strong> unstable, <strong>and</strong><br />

theri f ore soak up the colour, so that there does not remain a good <strong>and</strong> firm<br />

substance oj colour upon the cloth or sindone, unless, as usi f ul experience<br />

tells us, it is laid on several times.<br />

After the first layer was applied, the surface of the canvas was burnished to<br />

receive the second layer <strong>and</strong> finally a last layer <strong>and</strong> gilding (21):<br />

Let those things dry which you have drawn <strong>and</strong> painted, <strong>and</strong> when they<br />

are dry burnish them, that is, polish or smooth them gently with a tooth<br />

oj a horse or a boar, or with a polished hard stone fitted Jo r this purpose,<br />

in order that all the roughness may be softened down, ... <strong>and</strong> again paint<br />

over <strong>and</strong> draw upon those same places, with this colour, as bi f ore, <strong>and</strong><br />

afterwards let it dry, <strong>and</strong> then polish <strong>and</strong> burnish it as bi f ore. Afterwards<br />

go over <strong>and</strong> repaint those places which you did bifore, with the same<br />

mordant or colour, but let this third <strong>and</strong> last coat oj colour be tempered<br />

with white oj egg ....<br />

Sometimes painters tried to prepare the medium in such a way that it retained<br />

a workable consistency without being heated. This is why some recipes, including<br />

the following two recipes, survive for keeping glue fluid during preparation<br />

without heating it: "If you have the time, allow the mordant to get<br />

stale, for several days or weeks, for it will be better putrid than fresh" <strong>and</strong> "if<br />

you want to keep it liquid, put in more plain water, <strong>and</strong> let it st<strong>and</strong>; <strong>and</strong> after<br />

a few days it will stay liquid without heating. It may smell bad, but it will be<br />

very good" (22, 23).<br />

Others describe adding more water <strong>and</strong> vinegar with honey to the dissolved<br />

glue (24). The next day the vinegar is poured off, <strong>and</strong> clear water is added<br />

<strong>and</strong> heated. When it cools, the glue is stored <strong>and</strong> eventually becomes liquid.<br />

The small amount of vinegar soaked up by the glue also seems to help preserve<br />

it. The De Arte Illumin<strong>and</strong>i says, "And know that it is a very good plan<br />

to soften parchment or stag's horn glue with the best vinegar; <strong>and</strong> when it<br />

is softened, <strong>and</strong> the vinegar poured off, add plain water <strong>and</strong> melt it, <strong>and</strong><br />

proceed as has been said" (25). Pigments were mixed with this medium in<br />

much the same way as with egg tempera, presumably by first grinding them<br />

in water <strong>and</strong> then adding the medium.<br />

Painted on fine linen, the Adoration measures 54.6 em X 69.2 em. Judging<br />

by the age <strong>and</strong> type of strainer, it is reasonable to presume that the painting<br />

was removed from its original support, which could have been a panel or<br />

another strainer as described in the Alcherius manuscript. This would presumably<br />

have been carried out in the early nineteenth century for the purpose<br />

of mending two horizontal tears. Presumably to repair this damage, a<br />

lining canvas was applied to the back with an animal glue or starch paste. In<br />

what is preserved of the original canvas, no tack holes or pronounced stretch<br />

marks (called "scalloping") are visible. This might indicate that the canvas was<br />

glued to its support, either a panel or a strainer (26). Two paintings by Mantegna<br />

with their original supports still exist. One is the Ecce Homo at the<br />

Musee Jacquemart Andre in Paris, which is glued to a panel, <strong>and</strong> the other<br />

is the Presentation at the Staatliche Museen in Berlin, which has its original<br />

strainer (Fig. 3) (27). A slightly raised edge on the top might indicate that<br />

the Adoration had an engaged frame, such as the Berlin Presentation, which<br />

has a series of splinters adjacent to the painted frame with the color continuing<br />

up along these splinters. An example of an engaged frame survives<br />

(Metropolitan Museum of Art, Lehman Collection) that has a remnant of the<br />

original canvas by an unknown artist s<strong>and</strong>wiched between the strainer <strong>and</strong><br />

the frame (28).<br />

114<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 3. Andrea Mantegl1a, The Presentation in the Te mple. Origil1al strailler with panel inserts<br />

seen from front, 68. 9 X 86.3 CII. Staatliche Museen zu Berlin.<br />

Gold highlighting<br />

The intricate details <strong>and</strong> extensive gold highlighting of the Getty Adoration<br />

are laid out over preparatory base colors such as those visible on the borders<br />

of the Ecce Homo. In the latter, Mantegna probably finished the painting after<br />

it was placed into a frame (which no longer exists), as suggested by the fine<br />

details painted up to a one-half-inch border all around. The feathered-out<br />

colors on two edges of the Getty Adoration give an indication of such preparatory<br />

colors (Plates 24, 25). The figures themselves are placed against a<br />

uniform black background. The halos of St. Joseph, the Virgin, <strong>and</strong> the Christ<br />

Child, painted in shell gold (fmely ground gold in gum arabic), are executed<br />

with fine, concentric brush marks that gradually become longer <strong>and</strong> thicker<br />

toward the outer rim (Plate 24). The simulated Cufic letters on the Virgin's<br />

dress, the coins in Balthazar's cup, the censer <strong>and</strong> red cap of Melchior, <strong>and</strong><br />

the earrings <strong>and</strong> brooch of Balthazar are all highlighted with fm e gold lines.<br />

This all was done in a technique that is consistent with Alcherius' descriptions.<br />

Concluding observations<br />

The recent Mantegna exhibit at the Royal Academy in London <strong>and</strong> the<br />

Metropolitan Museum in New York provided us with a better underst<strong>and</strong>ing<br />

of Mantegna's use of distemper <strong>and</strong> gave us an opportunity to study the<br />

varnished <strong>and</strong> unvarnished paintings next to each other (29).<br />

Many distemper paintings that have been varnished have been considered<br />

irreversibly damaged. The recent restoration of the Adoration was a successful<br />

attempt to improve the aesthetic quality of the painting <strong>and</strong> reverse the glossy<br />

surface to a matte one, a surface that is more consistent with the original<br />

technique.<br />

Notes<br />

1. Important Old Master Pictures (Part 1). 1985. London: Christie, Manson, <strong>and</strong> Woods,<br />

Ltd. 18 April 1985:39-44.<br />

Rothe 115


2. Elam, C. 1982. The Splendours oj the Gonzaga. Catalog. London: Victoria & Albert<br />

Museum, 122, entry no. 32.<br />

3. Lightbown, R. W 1986. Mantegna. Oxford, Berkeley, <strong>and</strong> Los Angeles, 446.<br />

4. I am indebted to Dusan Stulik, Michele Derrick, Michael Schilling, <strong>and</strong> Eric<br />

Doehne from the Getty Conservation Institute fo r their extensive media analysis.<br />

5. Mills,]., <strong>and</strong> R. White. 1978. National Gallery Technical Bulletin (2):71-76.<br />

6. White, R. 1984. The characterization of proteinaceous binders in art objects.<br />

National Gallery Technical Bulletin (8):5-14.<br />

7. Thompson, D. L. 1982. Mummy Portraits in the J. Paul Getty Museum. Malibu,<br />

California: The ]. Paul Getty Museum, 7.<br />

8. Wolf thai, D. 1989. The Beginnings oj Netherl<strong>and</strong>ish Canvas <strong>Painting</strong>: 1400-1530.<br />

Cambridge: Cambridge University Press.<br />

9. Leonard, M., F. Preusser, A. Rothe, <strong>and</strong> M. Schilling. 1988. Dieric Bouts's Annunciation.<br />

The Burlington Magazine Guly): 517-522.<br />

10. Wolf thaI, op. cit., 6. See also 38-87 in catalogue.<br />

11. Illuminated manuscript artists used media, such as gum arabic, that are similar to<br />

watercolors <strong>and</strong> have very similar optical properties to distemper paint.<br />

12. Mantegna was known to have been an exasperatingly slow painter, as revealed<br />

in the correspondence with his benefactor, Ludovico Gonzaga, the Marquis of<br />

Mantua.<br />

13. Kristeller, P 1902. Andrea Mantegna. London, 534, doc. 69.<br />

14. Cennini, C. 1960. The Crciftsman's H<strong>and</strong>book. Translated by D. V Thompson. New<br />

York: Dover Publications, 87.<br />

15. White, op. cit., 5-14.<br />

16. Merrifield, M. P 1849. Original Treatises on the Arts oj <strong>Painting</strong>. London: John<br />

Murray, 232.<br />

17. Ibid., 89-90.<br />

18. Ibid., 255.<br />

19. Ibid., 262.<br />

20. Ibid.<br />

21. Ibid., 262-64.<br />

22. Ibid., 282.<br />

23. De Arte Iliumin<strong>and</strong>i. 1933. Translated from the Latin (Naples MS XII.E.27) by<br />

D. V Thompson <strong>and</strong> G. H. Hamilton. New Haven: Yale University Press, 12-<br />

13.<br />

24. Eastlake, C. L. 1960. Methods <strong>and</strong> <strong>Materials</strong> oj <strong>Painting</strong> oj the Great Schools <strong>and</strong><br />

Masters. Vol. 1. New York: Dover Publications, 106-7.<br />

25. De Arte Iliumin<strong>and</strong>i, op. cit., 13.<br />

26. Strainers differ from stretchers in that they have no keys. Stretchers became<br />

widely used after the eighteenth century.<br />

27. The Ecce Homo has the original panel, confirmed by carbon-14 dating, on which<br />

the original canvas is glued only by its edges to the back of the panel. The<br />

Presentation still has its original strainer with two original backing boards inserted<br />

into the strainer, as if it were on panel. The Lamentation in the Brera, which no<br />

longer has its original support, quite probably had one similar to the Presentation.<br />

28. Newberry, T., G. Bisacca, <strong>and</strong> L. Kanter. 1990. Italian Renaissance Frames. New<br />

York: The Metropolitan Museum of Art, no. 63 (diagram erroneously under no.<br />

64).<br />

29. So far, not a single painting of Mantegna's has been identified as painted in<br />

oil, even though there are records of his having ordered walnut oil.<br />

116<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

In September 1992 the J Paul Getty<br />

Museum acquired Titian's VertUS <strong>and</strong><br />

Adonis, which is considered to be<br />

one of the few of Titian's autograph<br />

paintings of the subject. The inorganic<br />

<strong>and</strong> organic pigments, oil<br />

medium of the paint layers, <strong>and</strong> proteinaceous<br />

medium of the ground<br />

were examined using various analytical<br />

methods. In addition, the authors<br />

studied Titian's other, much later<br />

version of Venus <strong>and</strong> Adonis located<br />

in the National Gallery in Washington,<br />

D.c. The comparison revealed<br />

distinct differences in style <strong>and</strong> technique.<br />

This multidisciplinary study<br />

has shown to be very useful in gaining<br />

a better underst<strong>and</strong>ing of Titian's<br />

working methods. The examination<br />

also revealed useful information<br />

about various aspects of early Italian<br />

oil painting methods in general <strong>and</strong><br />

helps place the technique of this<br />

painting by Titian in context with<br />

other paintings of the period.<br />

Te chnical Examinations of Titian's Venus <strong>and</strong> Adonis:<br />

A Note on Early Italian Oil <strong>Painting</strong> Technique<br />

Ulrich Birkmaier<br />

Conservation Department<br />

National Gallery of Art<br />

4th Street <strong>and</strong> Constitution Ave. NW<br />

Washington, D.c. 20565<br />

USA<br />

Arie Wallert*<br />

The Getty Conservation Institute<br />

Museum Services Laboratory<br />

The J Paul Getty Museum<br />

17985 Pacific Coast Highway<br />

Malibu, California 90265<br />

USA<br />

Andrea Rothe<br />

Department of <strong>Painting</strong>s Conservation<br />

The J Paul Getty Museum<br />

17985 Pacific Coast Highway<br />

Malibu, California 90265<br />

USA<br />

Introduction<br />

The story of Venus <strong>and</strong> Adonis, from Book X of Ovid's Metamorphoses, ends<br />

with the name of a fragile flower: anemone. As the tale is told, Adonis was a<br />

young man of unequaled beauty, who became the lover of Venus (1). Venus<br />

was so much in love with him "that she even stayed away from heaven,<br />

preferring Adonis to the sky. She used to hold him in her arms, <strong>and</strong> became<br />

his constant companion." She warned her lover not to hunt dangerous animals:<br />

"Your youth <strong>and</strong> beauty, <strong>and</strong> the charms which make Venus love you,<br />

have no effect upon lions or bristling boars, or the eyes <strong>and</strong> minds of other<br />

wild beasts. The fierce boar deals a blow with his fangs, as swift as a lightning<br />

flash ..."<br />

Of course, Adonis did go out hunting. His dogs fo und a fresh trail, followed<br />

it, <strong>and</strong> roused a wild boar. Adonis tried to kill it, but with the help of its<br />

crooked snout the boar dislodged the spear. The boar pursued Adonis <strong>and</strong><br />

"it sank its teeth deep in his groin, bringing him down, mortally wounded<br />

on the yellow s<strong>and</strong>." Venus, on her way to Cyprus, driving through the air<br />

in her chariot, heard her lover's groans. She went down, <strong>and</strong> with the dying<br />

Adonis in her arms, said that an everlasting token of her grief would remain<br />

there. His blood dripping to the ground would change into a flower, the<br />

anemone. But, just like their love, the enjoyment of this flower is brief "for<br />

it is so fragile, its petals so lightly attached, that it quickly falls, shaken from<br />

its stem by those same winds that give it its name, anemone."<br />

This story is represented in a series of poesie paintings of Classical subjects, by<br />

Titian. Such paintings were very popular among wealthy patrons, as their<br />

mythological subject matter usually provided an excuse for the depiction of<br />

overt sensuality. King Philip II of Spain commissioned a series of poesies from<br />

Titian. The Venus <strong>and</strong> Adonis now located in the Prado Museum in Madrid<br />

is considered the first of the series to reach its patron in 1554. The last one<br />

was The Rape oj Europa of 1562, now in the Isabella Steward Gardener Museum<br />

in Boston.<br />

Another version of Venus <strong>and</strong> Adonis by Titian recently came into the collection<br />

of the J. Paul Getty Museum. This work, showing Venus attempting<br />

* Author to whom correspondence should be addressed.<br />

Birkmaier, Wallert, <strong>and</strong> Rothe 117


to prevent Adonis from leaving for the hunt, is considered one of the few of<br />

Titian's autograph paintings on the subject (Plate 26). In addition to this work<br />

<strong>and</strong> the Venus <strong>and</strong> Adonis in the Prado, another of Titian's paintings by the<br />

same name is located at the National Gallery of Art in Washington, D.c.<br />

(Plate 27).<br />

Conservation work on the Getty Museum's recent acquisition of Venus <strong>and</strong><br />

Adonis gave the authors the opportunity to perform a technical examination<br />

of the painting. It was examined with the use of the following: microscopic<br />

methods, including polarized light microscopy (PLM) <strong>and</strong> scanning electron<br />

microscopy (SEM); spectrometric methods, including ultraviolet visible spectroscopy<br />

(UV Ivis), Fourier transform infrared (FTIR) spectroscopy, fluorescence<br />

(FS) spectroscopy, <strong>and</strong> energy dispersive X-ray fluorescence (EDXRF)<br />

spectroscopy; <strong>and</strong> chromatographic methods, including gas chromatography<br />

(GC) <strong>and</strong> thin-layer chromatography (TLC). X-ray diffraction (XRD) <strong>and</strong> a<br />

number of staining tests <strong>and</strong> microchemical tests were also used (2). The<br />

cooperation between conservators <strong>and</strong> scientists proved instrumental in gaining<br />

a better underst<strong>and</strong>ing of Titian's painting technique. This underst<strong>and</strong>ing<br />

was greatly improved through the cooperation of David Bull of the Washington<br />

National Gallery who gave us the opportunity to examine a later<br />

version of the Venus <strong>and</strong> Adonis in the gallery's collection using a stereomicroscope.<br />

This examination made it possible to compare Titian's execution<br />

of the same theme at different stages in his career. Our study was greatly<br />

facilitated by Joyce Plesters's pioneering publications on the examination of<br />

Titian pieces in the London National Gallery. Her work served as an example<br />

for our own research <strong>and</strong> as a continuing source for comparison, <strong>and</strong> helped<br />

in the interpretation of the analytical results. It also helped the authors place<br />

the technique of this particular painting in relation to that of other paintings<br />

of the same period, as well as those of Titian's later <strong>and</strong> earlier periods.<br />

Support<br />

The painting measures 160 X 196.5 cm. The original canvas is a plain weave<br />

linen, having 16 threads per centimeter in the warp direction <strong>and</strong> 18 threads<br />

per centimeter in the weft direction. It is made of two strips of canvas, joined<br />

by a vertical seam. The right strip measures 101 cm; the left-h<strong>and</strong> strip is<br />

slightly smaller <strong>and</strong> measures 95.5 cm. The painting has been cut on all sides;<br />

it is possible that the left side was cut by 5-9 cm. PI esters has found that the<br />

loom width of sixteenth-century Venetian canvases tends to range between<br />

1.06 <strong>and</strong> 1.10 cm (3).<br />

The painting was first documented in the collection of the Queen Christina<br />

of Sweden <strong>and</strong> then in that of the Duc d'Orleans. The French eighteenthcentury<br />

engraver Delignon made prints of objects in the duke's collection<br />

(4). One of these prints shows the Getty Venus <strong>and</strong> Adonis, in which it can<br />

be seen that the original composition may have extended out a bit more on<br />

the left side <strong>and</strong> that it was not originally cut off as close to Venus's fo ot as<br />

it is now. Since the scalloping of the original canvas on the left side is clearly<br />

still visible, the losses cannot have been too extensive.<br />

In the Prado version, which measures 186 X 207 cm (as compared to the<br />

National Gallery version, which measures 107 X 136 cm), the painting's<br />

surface extends a bit further to the left. Comparison of the loss in the Getty<br />

painting with corresponding areas in the print <strong>and</strong> in the Prado version leads<br />

to the conclusion that the width of the loss may be accounted for by the<br />

difference between the present 101 cm strip <strong>and</strong> the 106-1 1 0 cm width of<br />

the average sixteenth-century Venetian loom. At the right of the Getty painting,<br />

a relatively large section of the dog's tail, still complete in the Delignon<br />

print, is cut off. The amount of the loss is difficult to estimate, but judging<br />

from comparison with the print (the Prado version has suffered an even<br />

greater loss), it is probably not more than 4-5 cm.<br />

The loss at the upper side, again based on comparison with the Delignon<br />

print, <strong>and</strong> the London <strong>and</strong> Prado versions, is more extensive. The shaft of the<br />

118<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


arrow in Adonis's right h<strong>and</strong> ends at the edge of the painting. The loss may<br />

be estimated at 5-9 cm. On the basis of a comparison with the Delignon<br />

print, the original size of the painting may have been 172 X 213 cm.<br />

Ground<br />

The painting has a conventional gesso ground. The ground was chemically<br />

identified as calcium sulfate. XRD has shown that it consisted of a dihydrate<br />

gypsum. This composition is common fo r Venetian painting at that time.<br />

Gypsum of the dihydrate form occurs when it is used in its natural state. The<br />

use of unburned gesso-as opposed to burned gesso, where crystallization<br />

water is driven off to an anhydrite form-seems to occur primarily on Venetian<br />

paintings. PLM examination, however, has shown that a considerable<br />

amount of anhydrite particles were also present. <strong>Painting</strong>s from Florence <strong>and</strong><br />

Sienna more often show the presence of anhydrite or hemihydrite grounds<br />

(5). Gesso grounds of this type were the same as those used for ground layers<br />

on panel paintings, <strong>and</strong> its use stems from this tradition. In panel paintings,<br />

however, several layers of gesso-gesso grosso <strong>and</strong> gesso sottile--were applied.<br />

The brittleness of the gypsum layers, which form a good first ground on<br />

panels, would cause it to crack off too easily from a canvas. Therefore, artists<br />

began using thinner gesso layers on canvas; eventually, in the eighteenth century,<br />

an oil priming was used instead.<br />

The ground on the Getty Titian is rather thinly applied so that it only fills<br />

the spaces between the warps <strong>and</strong> wefts of the canvas. Cross sections show<br />

that it is an unpigmented ground. In some areas, an oily layer containing a<br />

few charcoal-black particles was found between the white ground <strong>and</strong> the<br />

first paint layer. It was only later, in the paintings by Tintoretto, for instance,<br />

that a colored ground became more common. A passage in the Volpato<br />

manuscript refers to Titian's use of white gesso grounds as opposed to the<br />

increasing use of colored grounds by other, more modern Venetian painters.<br />

Staining of microscopic cross sections with specific reagents gave strong indication<br />

that the binding medium of the ground contains proteinaceous material;<br />

this was confirmed by the presence of specific FTIR absorption b<strong>and</strong>s.<br />

Infrared mapping allowed us to locate the presence of specific b<strong>and</strong>s in the<br />

cross sections.<br />

Underdrawing<br />

Infrared reflectography revealed some evidence of underdrawing. The interpretation<br />

of the infrared reflectograms is difficult, however, as some features<br />

that showed up as broad, dark lines in the reflectogram could be carbon black<br />

used in the paint, rather than the underdrawing. It is not immediately obvious,<br />

for instance, whether the dog's curled tail in the original underdrawing was<br />

intended to be straight. The broad, dark form in the reflectogram may represent<br />

a dark, carbon-black pigment in the painted tail, which was actually<br />

intended to be curled in the underdrawing. The reflectogram shows sketchy<br />

lines in the trees in the background. The sleeping Cupid seems also to have<br />

been rapidly sketched before painting. In all, no significant deviations from<br />

the preparatory drawing appear to have been made.<br />

Paint layers: the medium<br />

Staining of cross sections indicated that the actual painting was executed in<br />

an oil medium. This was confirmed by the presence of characteristic absorption<br />

b<strong>and</strong>s in the FTIR spectrum (Fig. 1). Three samples were selected for<br />

examination with gas chromatography. One sample consisted Inainly of paint<br />

fo r the golden vase in the painting's lower left corner. The other samples<br />

contained blue particles for the sky. Contemporary sources often indicate the<br />

use of walnut oil, which was generally considered to yellow less with age, for<br />

the making of blue paints. Earlier studies of Titian's paintings have revealed<br />

that the artist used both types of oil on different occasions. Chromatography<br />

showed an azelaic:palmitic acid ratio in all three different samples, indicating<br />

Birkmaier, Wallert, <strong>and</strong> Rothe 119


60<br />

55<br />

"! ,... '"<br />

.. ..; ,.:<br />

50<br />

N 0<br />

<br />

'"<br />

:;!; <br />

, , , , ,<br />

IT '000 3600 3200 2800<br />

'00<br />

Wavenumber (cm-1)<br />

2000 raoo 1600 "00 1200 1000 800 600<br />

Figure 1. FTIR spectrum of paint sample (no. 17) takenfrol11 blue if the sky area. Hydrocarbon<br />

stretch b<strong>and</strong>s at 2923 CI11-1 <strong>and</strong> 2851 cm-I• as well as the hydrocarbon bend at 1404 cm-I, indicate<br />

the presence of an oi/ medium. The b<strong>and</strong> at 2342 cm -I is characteristic for natural ultramarine.<br />

the use of a drying oil. The palmitic:stearic acids ratios of the samples suggest<br />

that the sky was executed in walnut oil or in a mixture of walnut <strong>and</strong> linseed<br />

oils in contrast to other areas of the painting, in which the faster drying, but<br />

more yellowing, linseed oil was used.<br />

<strong>Painting</strong><br />

The thinness <strong>and</strong> relatively simple structure of the paint layers are in accordance<br />

with what seems to have been Titian's practice around the 1550s.<br />

Examination of the cross sections shows that most of the paint was fairly<br />

directly <strong>and</strong> thinly applied, as opposed to the technique he used in his later<br />

paintings. In the foreword to Marco Boschini's Ricche Minere della Pittura Veneziana<br />

(1664), an authoritative <strong>and</strong> contemporary description is given by<br />

Palma Giovane of the manner in which Titian gave form to the paintings he<br />

made after the 1550s (6):<br />

He used to sketch in his pictures with a great mass if colours, which served<br />

as a base fo r the compositions he then had to construct. [The compositions<br />

wereJformed with bold strokes made with brushes laden with colours, sometimes<br />

with a pure red earth, which he used for a middle tone, <strong>and</strong> at other<br />

times of white lead; <strong>and</strong> with the same brush tinted with red, black <strong>and</strong><br />

yellow he formed an accent; <strong>and</strong> thus he made the promise if a figure<br />

appear in fo ur strokes . ... Having constructed these precious fo undations<br />

he used to turn his pictures to the wall <strong>and</strong> leave them without looking at<br />

them, sometimes for several months. When he wanted to apply his brush<br />

again . .. he would treat his picture like a good surgeon would his patient,<br />

reducing if necessary some swelling or excess of flesh, straightening an arm<br />

if the bone structure was not exactly right . ... After he had done this,<br />

while the picture was drying, he would turn his h<strong>and</strong> to another <strong>and</strong> work<br />

on it the same way. Thus he gradually covered those quintessential forms<br />

with living flesh, bringing them by many stages to a state in which they<br />

lacked only the breath of life . ... In the last stages he painted more with<br />

his fingers than with his brushes.<br />

This process could involve not just two or three weeks but, with several steps<br />

<strong>and</strong> with long interruptions, could continue for months, even years, resulting<br />

in a painting with several paint layers.<br />

The process, as described by Palma Giovane, is distinctly different from the<br />

earlier manner in which Titian painted. This difference was already noticed<br />

by Vasari in 1566:<br />

120<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


For the early works are executed with a certain f inesse <strong>and</strong> an incredible<br />

diligence, so that they can be seen from close as well as from a distance<br />

while these last pictures are executed with broad <strong>and</strong> bold strokes <strong>and</strong><br />

smudges, so that from nearby nothing can be seen whereas from a distance<br />

they seem peifect.<br />

In the case of the Venus <strong>and</strong> Adonis at the Getty, the earlier technique was<br />

used. The paint layers are few <strong>and</strong> relatively thinly applied. Where visible in<br />

infrared reflectography, the underdrawing seems to have been closely followed.<br />

No significant changes from the carefully planned composition have been<br />

made. Some later additions seem to have been made, however, in areas such<br />

as Adonis's arms <strong>and</strong> Venus's legs <strong>and</strong> back. Some areas in the l<strong>and</strong>scape also<br />

appear to be built up with several paint layers. The pigments seem to be<br />

thoroughly ground directly into the binding medium. This contrasts strongly<br />

with the pigments in the National Gallery's Venus <strong>and</strong> Adonis. Observed<br />

under a stereo microscope, many lumps of pigment are visible, suggesting that<br />

in making the paint, the pigments were not really ground into, but simply<br />

stirred into the oil. Some differences can be noted between the version in<br />

the Getty <strong>and</strong> the one in the Prado. Judging from the development as described<br />

by the two contemporary comments, the Getty painting is more<br />

loosely executed than the Prado version, <strong>and</strong> must have been made at a later<br />

date, perhaps around 1560. The National Gallery version can definitely be<br />

considered to have been painted after the Getty <strong>and</strong> Prado versions.<br />

Paint layers: the pigments<br />

Yellow pigments. The highlights of Adonis's belt were executed in a bright<br />

yellow with occasional small orange-red dots placed on the highlights. When<br />

examined with PLM, samples taken from the area showed that the yellow<br />

consisted of highly birefringent mica-like particles (n>1 .66). The yellow pigment<br />

had all the optical characteristics of orpiment, including a laminated<br />

form <strong>and</strong> a waxy luster of the fairly large crystalline particles. The identification<br />

of orpiment was confirmed by finding larger amounts of arsenic in<br />

the sample. A small microchemical test showed that the sample also contained<br />

sulfides.<br />

The red dots on Adonis's belt <strong>and</strong> the orange-red of Adonis's shoe were also<br />

executed in an arsenic-containing pigment realgar. These pigments were<br />

identified by the usual means of microscopic <strong>and</strong> microchemical analysis. In<br />

addition, the findings were confirmed by XRD <strong>and</strong> XRF (Fig. 2).<br />

The use of orpiment <strong>and</strong> realgar in European easel paintings is comparatively<br />

rare, but it is not unusual to find them in Venetian painting; the use of<br />

orpiment occurred more often in manuscript illuminations. This lack of popularity<br />

may be due to the fact that both pigments are highly poisonous, <strong>and</strong><br />

good alternatives, such as lead-tin yellow or yellow lakes fo r orpiment or<br />

vermilion for realgar, were readily available. Because of their sulfidic nature,<br />

both orpiment <strong>and</strong> realgar are not very compatible with many other pigments.<br />

The Theophilus manuscript even states that orpiment cannot be mixed with<br />

any other color because it would destroy them. While realgar was sometimes<br />

used to prevent putrefaction of binding media by bacteria, fungi, <strong>and</strong> microorganisms,<br />

historical recipes describing the use of realgar as an actual pigment<br />

are extremely rare (7).<br />

Considering their drawbacks, it is surprising to see how frequently these<br />

pigments were used in sixteenth-century Venetian painting. They have been<br />

identified in several paintings by Titian <strong>and</strong> Giorgione <strong>and</strong> in many paintings<br />

by Tintoretto <strong>and</strong> Bassano. The reasons for this preference are not immediately<br />

evident. Both mineral substances could not be fo und in the immediate<br />

vicinity <strong>and</strong> had to be imported. They are conspicuously absent in paintings<br />

made in the area around Naples where the minerals actually do occur in the<br />

fumaroles near Mount Vesuvius. Their frequent use in sixteenth-century Venice<br />

may relate to the increasing use of oil as the paint medium. The pigment<br />

Birkmaier, Wallert, <strong>and</strong> Rothe 121


!:;<br />

;Dr -ID'LdDtDw<br />

at' ______________________________-' __ r-__ ,, __ '-'-,<br />

Titian YlIOIl low .pot on<br />

baIt Adon i.<br />

80/51" tcrg8t.<br />

240 50kV. 3. 3m".<br />

acac. 11. I. 1993, AW<br />

A.<br />

t-<br />

..<br />

7L-____________________________________________________________ <br />

KloV 10 11 12 13<br />

Figure 2. XRF spectrum of yellow area on belt of Adonis. High peaks fo r arsenic <strong>and</strong> sulphur suggest<br />

the presence qf orpil11ent.<br />

particles, being enclosed in the oily medium, would no longer be incompatible<br />

with other pigments, as they were in the previously used tempera techniques.<br />

The radiant brilliance of their color in this new medium may have<br />

contributed to their popularity.<br />

Ochres were used to paint the l<strong>and</strong>scape, the brownish color of the dogs, <strong>and</strong><br />

the golden vase near Venus's seat. Yellow ochres (hydrous iron oxides) were<br />

identified under the microscope by their optical properties <strong>and</strong> by the presence<br />

of high peaks fo r iron in the XRF spectrum.<br />

Blue pigments. The major blue pigment used on this painting was natural<br />

ultramarine (Fig. 1). Samples taken from the deep blue of the mountain range<br />

appear under the microscope as pale blue, splintery particles with a low refractive<br />

index (n < 1.66). In the cross section, one can see a single layer of<br />

densely packed blue particles.<br />

XRF of several blue areas showed high peaks for cobalt, potassium, <strong>and</strong> arsenic,<br />

strongly suggesting the presence of smalt, an artificial pigment made<br />

from potassium-rich glass deeply pigmented with cobalt oxide <strong>and</strong> ground<br />

to a powder. Gettens <strong>and</strong> Stout suggest that the earliest occurrence of cobaltcolored<br />

glass in Europe may have been in the early fifteenth-century Venetian<br />

glass industry (8) . As the colorant for the glass, a substance called zafran is<br />

mentioned. A recipe in a fifteenth-century treatise already mentions the preparation<br />

of smalt as a smalto cilestro (9).<br />

The source of the smalt in the Getty painting may have been Saxony or<br />

Bohemia, where sixteenth-century glassmakers used the locally mined cobaltite<br />

(CoAsS) <strong>and</strong> smaltite (CoAs2) minerals, which contain large amounts<br />

of arsenic, to make smalt. The high peaks of arsenic measured by XRF, in<br />

combination with those of cobalt, seem to indicate that cobaltite or smaltite<br />

minerals were the source for the blue pigment. It is unclear whether the<br />

zafran, or zaffer of Italian descriptions, has the same composition as the northern<br />

European cobaltite. In other areas of the painting's sky, the original ultramarine<br />

was scumbled over with a pale, milky blue. Examination of cross<br />

sections of those areas shows that the pale blue consists of a layer of almost<br />

completely discolored glassy particles.<br />

Red pigments. Vermilion appears in Adonis's red sleeve. Its presence was established<br />

by discovering high mercury peaks with XRF <strong>and</strong> confirmed by<br />

122<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


J<br />

Figure 3. Fluorescence spectrul1'l of red lake pigment on shirt of Adonis. Quantum yields at A em 445<br />

nl1'l to A ex 397 nm <strong>and</strong> A em 433 nl1'l to A ex 355 rim indicate preSe/1ce of a cochineal dyestuff.<br />

microchemical tests <strong>and</strong> PLM of a sample taken from the area. The other<br />

reds in Adonis's tunic <strong>and</strong> the clothing draped over Venus's seat were of an<br />

organic nature. In a sample taken from the seat, an organic colorant was fo und<br />

that also showed the presence of a few textile fibers. This led us to believe<br />

that the red lake may very well have been a so-called lacca de cimatura de grana.<br />

These cimatura lakes were made by an early recycling process. The dyestuffs<br />

were precipitated into a pigment lake extracted from red textile clippings or<br />

shearings (cimature) from a tailor's workshop. The making of such lakes was a<br />

fairly common practice, <strong>and</strong> its description can be fo und in numerous fifteenth-<br />

<strong>and</strong> fourteenth-century recipe texts (10). The procedure could be<br />

performed with most of the available red textiles. The dyestuffs most likely<br />

to be found in these lakes are anthraquinone-type dye, such as kermes, cochineal,<br />

lac dye, <strong>and</strong> madder. The analyses by FS <strong>and</strong> TLC showed the presence<br />

of more than one organic colorant; the red appeared to be a mixture. A sample<br />

taken from Adonis's garment showed a good match with a library scan of the<br />

cochineal st<strong>and</strong>ard (Fig. 3). The results were confirmed by chromatography,<br />

which also indicated the presence of purpurin, a colorant present in madder<br />

type dyestuffs. No spot for alizarin, the major coloring component of common<br />

madder (Rubia tinctorum), could be found. Several possible conclusions<br />

are indicated: (a) wild madder (Rubia peregrina) was used, (b) a specific technique<br />

was used for the dyeing of the original textile, or (c) the alizarin was<br />

not fully extracted in the cimatura process. This finding supports the suggestion<br />

that the lake was a cimatura de grana-lake, <strong>and</strong> the painters who extracted<br />

the dyestuffs from textile shearings had no control over the actual dyestuff<br />

composition. Since they are similar in color, <strong>and</strong> not readily distinguishable,<br />

any madder-dyed textiles could easily have been included in a larger batch<br />

cochineal-dyed clippings.<br />

Lake pigments are known to dry poorly. It was common practice, therefore,<br />

to add ground glass to lakes. Sixteenth-century glass was typically very rich<br />

in lead; the lead in the glass would act as a siccative, thus promoting the<br />

drying of the paint. An interesting passage in the Paduan manuscript describes<br />

the process: "To make lake indigo <strong>and</strong> lamp-black, dry quickly. Grind them<br />

with oil, then take glass ground to a very fine powder, <strong>and</strong> incorporate with<br />

the colors by grinding them together again; <strong>and</strong> thus, in the space of 24<br />

hours, they will dry" (11).<br />

Examination of a microsample of the red lake of Adonis's shirt did, indeed,<br />

show the presence of small glass particles. The purplish red of Venus's seat,<br />

on the other h<strong>and</strong>, was made of a mixture of an organic red lake <strong>and</strong> smalt<br />

(Fig. 4). In this case, the smalt had a double function. It gave the carmine lake<br />

a more purple color <strong>and</strong> at the same time acted as a siccative.<br />

White pigments. The white pigment found on the painting was lead white.<br />

XRF revealed the presence of lead in the sky <strong>and</strong> in the flesh tints. Polarized<br />

Birkmaier, Wallert, <strong>and</strong> Rothe 123


Titian r&d 91 aZI!;! on aQot<br />

12.1.199:3. 80/5,.. target. 240 SOkV, ::I. grru'<br />

eee. AW<br />

F.<br />

Pb<br />

A.<br />

Pb<br />

Figure 4. XRF spectrum if purple area on seat of Venus. High peaks fo r cobalt <strong>and</strong> arsenic suggest<br />

the presence if smalt.<br />

light microscopy of pigment samples taken from those areas showed them to<br />

contain birefringent particles with a high refractive index (n> 1.66) that<br />

matched lead-white laboratory st<strong>and</strong>ard. This conclusion was confirmed by<br />

staining cross sections with Lugol stain. X-radiographs of the painting show<br />

how Titian used the lead white, particularly to enhance the contrasts in his<br />

composition. It helped to make Adonis's silhouette st<strong>and</strong> out against lighter<br />

areas in the sky, <strong>and</strong> emphasized the reddish blush on the face of Venus against<br />

the bright crimson shirt of Adonis (Fig. 5).<br />

Figure 5. X-radiograph if Venus <strong>and</strong><br />

Adonis, showing Titian's use of lead-whitebased<br />

highlights.<br />

A sample taken from the sky area showed that there was not very much<br />

pigment present. The sky appeared to be executed in thin glazes over an<br />

anhydrite gesso ground, with the white of the ground showing through. The<br />

finding of the lathlike anhydrite particles came as a surprise since anhydrite<br />

is generally considered to have been used in more southern areas of Italy. In<br />

the Venice region, the dihydrate fo rm was usual. We do not yet have a satisfactory<br />

explanation for this. In samples taken from the sky area, a relatively<br />

large amount of splintery, isotropic particles with low refractive index<br />

(n


This property was deliberately used in the making of copper resinates. These<br />

are complex compositions of copper salts with various resinous acids, such as<br />

abietic <strong>and</strong> succinic acids, achieved by heating verdigris with Venetian turpentine<br />

or pine resins. The resulting substance shows no particulate matter,<br />

<strong>and</strong> a strong, bright green color. Copper resinate became the pigment of<br />

choice whenever bright green glazes were desired. An early account fo r the<br />

deliberate making of such a compound can be found in a fifteenth-century<br />

manuscript in the Biblioteca Casanatense (13). Unfortunately, the copper resinates<br />

show the same tendency to discolor, giving the meadow <strong>and</strong> the trees<br />

in the background of the Getty Titian their present brown color in place of<br />

the originally deep green color. In various areas on the painting, however,<br />

the copper resinate has retained some of its original color. In particular, the<br />

green leaves of grass near Adonis's foot st<strong>and</strong> out because of their bright,<br />

strong color. Examination of a cross section taken from that area showed that<br />

the copper resinate was not laid over the underpaint but rather mixed with<br />

lead white. This mixture may possibly have helped in preserving much of its<br />

original color. The green of the trees in the background was produced with<br />

copper resinate.<br />

The green that was used to paint the meadow near the vase at the lower-left<br />

corner consisted of a mixture of copper resinate <strong>and</strong> yellow ochre. The cross<br />

section showed that there were only two different paint layers on the gesso<br />

ground: one layer of yellow ochre, possibly applied in two coatings, <strong>and</strong> an<br />

upper layer of copper resinate. The leaves of the plants appear to be highlighted<br />

with white, <strong>and</strong> glazed with copper resinate.<br />

No other green pigments could be identified. This finding tallies with the<br />

use of greens in Titian's later Tarquin <strong>and</strong> Lucrezia (14). We were surprised<br />

that malachite, fo und in Titian's almost contemporary Bacchus <strong>and</strong> Ariadne or<br />

in Tintoretto's paintings in the National Gallery, could not be detected in the<br />

samples we took from the Getty painting (15, 16).<br />

Conclusion<br />

Our examinations show that the execution of the J. Paul Getty Museum's<br />

Venus <strong>and</strong> Adonis represents a stage between the Prado version <strong>and</strong> the National<br />

Gallery version. It st<strong>and</strong>s technically in the middle between two extremes<br />

in Titian's stylistic development. The earliest stage is the Prado version<br />

in which an already conceived image is carefully designed <strong>and</strong> then executed<br />

<strong>and</strong> filled in accordingly. In Titian's later style, however, the painting grows<br />

out of an interaction between matter <strong>and</strong> concept. As every touch of the<br />

brush has its impact on all previous touches, there is a shift in the appearance<br />

of the final painting. Rather than resulting from a fixed plan, this way of<br />

creating a painting with the total problem of the picture in mind, is apt to<br />

be a continually developing <strong>and</strong> self-revising one.<br />

The Getty piece already represents a concept of painting in which form does<br />

not merely follow function, but rather grows out of a continuous interaction<br />

between the dem<strong>and</strong>s of the material <strong>and</strong> the artistic idea: "Obwohl das Werk<br />

erst im Vollzug des Schaffens wirklich wird und so in seiner Wirklichkeit<br />

von diesem abhangt, wird das Wesen des Schaffens vom Wesen des Werkes<br />

bestimmt" (17).<br />

Notes<br />

1. Ovid. 1986. Metamorphoses. Translated <strong>and</strong> introduced by M. M. 1. Middlesex,<br />

Penguin Books, Ltd., 239, 244-45.<br />

2. van Asperen de Boer. J. R. J. 1975. An introduction to the scientific examination<br />

of paintings. Netherl<strong>and</strong>s Yearbook Jo r History if Art (26): 1-40.<br />

3. Plesters, J. 1980. Tintoretto's <strong>Painting</strong>s in the National Gallery, National Callery<br />

Technical Bulletin (4):37.<br />

4. Wethey, H. E. 1975. The <strong>Painting</strong>s oj Titian, Vol. III: The Mythological <strong>and</strong> <strong>Historical</strong><br />

<strong>Painting</strong>s. London: Phaidon, plate 189.<br />

Birkmaier, Wallert, <strong>and</strong> Rothe 125


5. Gettens, R.]., <strong>and</strong> M. E. Mrose. 1954. Calcium sulphate minerals in the grounds<br />

of Italian paintings. Studies in Conservation (4):182-83.<br />

6. Valcanover, F. 1990. An introduction to Titian. In Titian, Prince oj Painters. Venice,<br />

3-28.<br />

7. The only documentary evidence for it can be fo und in the fifteenth-century<br />

manuscript in Modena Biblioteca Estense (MS a.T.7.3), fol. 3r. Also see Wallert,<br />

A. 1984. Orpiment <strong>and</strong> realgar, some pigment characteristics. MaltechniklRestauro<br />

(90):45-58.<br />

8. Gettens, R.]., <strong>and</strong> G. L. Stout. 1966. <strong>Painting</strong> <strong>Materials</strong>: A Short Encyclopedia. New<br />

York: Dover Publications.<br />

9. Rome, Biblioteca Casanatense MS 2265, fol. 91r, describes the making of the<br />

blue glass. On fol. 88r, the manuscript describes the actual processing of the glass<br />

into a pigment for painting: "A macinare ismalto azuro che poi pingere in suo<br />

colori. Pilia il smalto et rompelo sotile poi macinelo sopra il porfido cun sorro<br />

de ovo cun uno pocho di melle. Masinalo sutile como vopi et se <strong>and</strong>asse secondo<br />

azonze uno pocho de aqua. Et qu<strong>and</strong>o e masinato lava fora il mele et 10 ovo<br />

cun l'aqua e remanera cossa bellisima."<br />

10. Wallert, A. 1991. "Cimatura de grana"; Identification of natural organic colorants<br />

<strong>and</strong> binding media in mediaeval manuscript illumination. Zeitschrijt ju r Kunsttechnologie<br />

und Konservierung (5):74-83.<br />

11. Merrifield, M. P. 1849. Original Treatises on the Arts if <strong>Painting</strong>, Vol. II. London:<br />

John Murray, 667.<br />

12. Kiihn, H. 1964. Griinspan und seine Verwendung in der Malerei. Farbe und Lack,<br />

(70):703-1 1.<br />

13. Rome, Biblioteca Casanatense MS 2265, fol. 99r : "A fa r stagnolo verde. Piglia<br />

vernice liquida e olio de linosa equal misura e meteli denmtro verderame bene<br />

pulverizato. Poi fa bolire tanto che habia il colore a tuo modo. Poi 10 adopeera."<br />

The result of this recipe is a complex copper resinate/oleate compound. This<br />

recipe shows that copper resinates were not always made of Venice turpentine<br />

as is generally assumed on the basis of a description in De Mayerne's treatise.<br />

14. Jaffe, M., <strong>and</strong> K. Groen. 1987. Titian's Tarquin <strong>and</strong> Lucrezia in the Fitzwilliam.<br />

The Burlington Magazine, (CXXIX):162-71.<br />

15. Lucas, A., <strong>and</strong> J. Plesters. 1978. Titian's Bacchus <strong>and</strong> Ariadne. National Callery<br />

Technical Bulletin (2) :25-48, especially 4 1.<br />

16. Plesters, op. cit., 32-48.<br />

17. Heidegger, M. 1970. Der Ursprung des Kunstwerkes. Stuttgart: Reclam Verlag, 66.<br />

126<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

The St. Anne Altarpiece (National<br />

Gallery of Art, Washington, Widener<br />

Collection) is the largest work attributed<br />

to the Bruges artist Gerard<br />

David (ca. 1460-1523). Several panels<br />

of the altarpiece were examined<br />

to identify the materials <strong>and</strong> methods<br />

used. The results show that the<br />

center panel was painted using a different<br />

technique from that for the<br />

side <strong>and</strong> predella panels. There are<br />

distinct groupings of underdrawings<br />

among the panels. The main panel<br />

was laid out more specifically at the<br />

underdrawing stage, with three separate<br />

drawing campaigns <strong>and</strong> a correction<br />

on top of a paint layer. In<br />

the center panel only, all the major<br />

fo rms were laid out in the first paint<br />

stage with a lighter value of the final<br />

paint color, with multiple layers of<br />

progressively more saturated paint<br />

built up to create the final appearance.<br />

In contrast, the predella <strong>and</strong><br />

side panels were painted directly, using<br />

fewer layers of paint. The program<br />

of painting would have allowed<br />

for workshop participation, especially<br />

in the main panel.<br />

Gerard David's St. Anne Altarpiece: Evidence for<br />

Workshop Participation<br />

Catherine A. Metzger <strong>and</strong> Barbara H. Berrie*<br />

National Gallery of Art<br />

Washington, D.c. 20565<br />

USA<br />

Introduction<br />

Gerard David (ca. 1460-1523) is known to have been a member of the painters'<br />

guild in Bruges by January 1484. It is generally accepted that he operated<br />

a workshop, evidenced by the use of pricked drawings <strong>and</strong> his documented<br />

dispute with Ambrosius Benson over the alleged theft of such drawings (1 ,<br />

2). The recent restoration of the St. Anne Altarpiece provided an opportunity<br />

fo r its reexamination <strong>and</strong> technical analysis (3). This paper addresses the evidence<br />

for workshop participation in the altarpiece. The full results of the<br />

technical investigation will be published in the future.<br />

The altarpiece<br />

The St. Anne Altarpiece depicts St. Anne with the Virgin <strong>and</strong> Child flanked<br />

by St. Nicholas (left) <strong>and</strong> St. Anthony of Padua (right) (Plate 28). John H<strong>and</strong><br />

has described the altarpiece (4). Six pre della panels are accepted as part of the<br />

original altarpiece; the three panels at the National Gallery of Scotl<strong>and</strong> illustrate<br />

three miracles from the life of St. Nicholas, while the three panels at the<br />

Toledo Museum of Art show three miracles from the life of St. Anthony.<br />

Another panel, Lamentation of Christ at the Foot of the Cross (The Art Institute<br />

of Chicago, Mr. <strong>and</strong> Mrs. Martin A. Ryerson Collection), is possibly part of<br />

this altarpiece. According to dendrochronological data, David painted the altarpiece<br />

around 1506, but the stylistic evidence for dating is less precise (5).<br />

The stiffness <strong>and</strong> lack of expression of the figures in certain altarpiece panels,<br />

especially the central panel <strong>and</strong> the St. Anthony pre della panels, have led to<br />

doubts regarding the attribution to David. Workshop participation in the St.<br />

Anne Altarpiece was proposed as early as 1905 (6) . Marlier suggested the participation<br />

of Ambrosius Benson in the altarpiece (7). Scillia has attributed the<br />

central panel <strong>and</strong> the St. Anthony predella panels to an assistant (8). Recently,<br />

it has been suggested that the entire altarpiece is by a follower of David (9).<br />

The support<br />

Figure 1. Secondary eiectroll image of the<br />

chalk ground of the center panel of Gerard<br />

David <strong>and</strong> Workshop's St. Anne Altarpiece,<br />

ca. 1506. The coccolith is typical of<br />

those found in northern European deposits.<br />

The two side panels of the altarpiece, which have been reduced in height,<br />

now measure 214 X 76 cm. If the semicircular upper profile is recreated<br />

based on the truncated arcs still visible at the upper extremity of each panel,<br />

the original height of these panels can be estimated at 236 cm. By comparison<br />

to contemporary Italianate altarpieces, the central panel can be assumed to<br />

have been even greater in height. The predella panels are 56 cm high. The<br />

altarpiece is the largest work in David's oeuvre <strong>and</strong> one of the largest surviving<br />

from the period (10). The scale alone suggests that David would have<br />

had assistance in this production. The patron fo r such a sumptuous <strong>and</strong> extravagant<br />

commission is not yet known.<br />

The central panel is assembled from fo ur boards, each side panel from three.<br />

From the relative widths of the central <strong>and</strong> side panels, it is clear that this<br />

was a fixed, not a closing, altarpiece. The oak panels were prepared in the<br />

traditional manner with coatings of chalk in glue. Coccoliths in the chalk<br />

ground prove a natural, <strong>and</strong> Northern, source (Fig. 1) (11).<br />

* Author to whom correspondence should be addressed.<br />

Metzger <strong>and</strong> Berrie 127


Figure 2. Infrared rriflectogram assembly if the cherub st<strong>and</strong>ing 01'/ the throlle (lift side, area if rriflectogram:<br />

23 X 10 C/'I'/). Type I drawing is visible, characterized by an even, relatively narrow width.<br />

Rriflectograph by M. Faries.<br />

The underdrawing<br />

Infrared reflectography reveals distinct campaigns of underdrawing that can<br />

be described in four groups (12). Type I is characterized by a narrow, even<br />

line that skips over the surface of the ground. It is the first drawing stage in<br />

all three of the main panels <strong>and</strong> the St. Nicholas predella panels. There is only<br />

a small area of type I drawing in one of the St. Anthony predella panels. The<br />

drawing made using this material is summary. An example of this drawing<br />

type in the cherub on the throne is illustrated in Figure 2. This underdrawing<br />

128<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 3. Infrared reJlectogram assembly of the Virgin's head (area of reJlectogram: 23 X 17 em).<br />

Type II underdrawing characterized by variable density <strong>and</strong> line width. ReJlectograph by M. Faries.<br />

can be seen in the cross section illustrated in Plate 29a, obtained from the<br />

location marked in Figure 3. The underdrawing layer is only a few microns<br />

thick.<br />

Type II underdrawing is characterized by variable line width <strong>and</strong> intensity.<br />

These properties, together with the distinctly dropletlike forms visible in some<br />

places, suggest type II lines were produced using a liquid. Type II lines are<br />

found in the figural group of the center panel, where they describe the forms<br />

more firmly. Figure 3 shows the infrared reflectogram of St. Anne's head.<br />

Here sketchy type I lines have been overdrawn with the type II underdrawing.<br />

Nearly all the underdrawing in the St. Anthony predella panels is executed<br />

with a liquid type II drawing material that lays out the composition concisely.<br />

In the central figural group of the main panel, however, the line is often rigid<br />

<strong>and</strong> repetitive as, for example, in the shading to the right of the Christ Child.<br />

In addition to these kinds of underdrawing there is a third form, type III,<br />

that can be seen with the naked eye in St. Anne's red robe (Fig. 4). These<br />

lines are characterized by their bold appearance with h<strong>and</strong>ling qualities more<br />

similar to that of a paint than of an ink. A cross section shows type III was<br />

laid down thickly, <strong>and</strong> then covered with a translucent layer (Plate 29b). Type<br />

III has only been found in St. Anne's red drapery <strong>and</strong> no other place in either<br />

the side or predella panels. A fo urth type of marking of fo rm is evident in<br />

the reflectogram of the Virgin's sleeve (Fig. 5). Here, a dark irregular line lies<br />

over the first layers of paint (Plate 29c). The subsequent layer of paint follows<br />

the line denoted by this "correction" layer. Optical microscopy showed the<br />

pigment here is charcoal (13). Two panels, The Miracle if the Purse <strong>and</strong> The<br />

Metzger <strong>and</strong> Berne 129


Figure 4. Infrared niffectogram assembly of part of St. Anne's red drapery (area of riflectogram: 24 X<br />

21 em). Type III underdrawing, which is characterized by its painterly quality, is illustrated. Riflectograph<br />

by M. Faries.<br />

Miracle of the Dead Child, have w<strong>and</strong>ering marks that recollect the appearance<br />

of the correction in the Virgin's sleeve. The dotted quality of these lines has<br />

led researchers to interpret them as having been produced by pouncing. However,<br />

microscopical examination of the paint surface showed that in these<br />

locations, the black paint protrudes through the uppermost white layers, giving<br />

an impression of a series of dots. These spotty "correction" lines in the<br />

predella panels were not followed in the final paint layers, as they were in the<br />

center panel.<br />

The looseness <strong>and</strong> lack of formality of the underdrawing in the two side<br />

panels <strong>and</strong> the predella panels, coupled with the fact that there are many<br />

repositionings from the underdrawn to the paint stage, suggests that someone<br />

with the authority to interpret the sketch <strong>and</strong> deviate from it, most likely the<br />

master himself, was working on the side panels <strong>and</strong> the predella panels. The<br />

careful additional delineation of outlines in the St. Anne suggests that someone<br />

who was less intuitive was to work on this panel. The "correction" mark<br />

in the Virgin's sleeve, which is followed in the final paint layer, also suggests<br />

a second h<strong>and</strong> in this process. The " correction" marks in the predellas, which<br />

are ignored in the final paint stage, suggest the finishing touches were completed<br />

by someone who could <strong>and</strong> did act independently. Participation of<br />

workshop members at the underdrawing stage has been noted. For example,<br />

Fairies found at least three h<strong>and</strong>s in the underdrawing in the Crucifixion triptych<br />

in the Rijksmuseum Het Catherijnenconvent in Utrecht (14). The master,Jan<br />

van Scorel, drew only in the exterior wings; the drawing in the interior<br />

wings <strong>and</strong> the central panel was executed by two anonymous assistants.<br />

130<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 5. Irifrared niflectogram assembly of the Virgin 's sleeve, showing the "correction" layer which<br />

appears as dotted lines (area of refiectogral1l: 24 X 21 an). Refiectograph by M. Faries.<br />

The paint layers<br />

The preparatory layers. Staining the cross sections shows that the chalk ground<br />

in the main panels was not coated with a layer of glue before application of<br />

the underdrawing. In some of the cross sections, however, there is a dusting<br />

of red <strong>and</strong> black particles in an oleaginous layer; these particles are especially<br />

visible under the green cloth of honor behind St. Anne <strong>and</strong> the Virgin.<br />

Scattered particles such as this are rarely noted in painting cross sections;<br />

however, Coremans, et al. <strong>and</strong> later Brinkman, et al. found a similar layer in<br />

the Ghent altarpiece (15, 16). There is a clearer indication of the use of a<br />

sealing layer over the underdrawing in St. Anne's red drapery. Here the hatching<br />

lines, which are visible to the naked eye, have a translucent layer over<br />

them, separating the red paint from the drawing lines.<br />

The X-radiographs of all six predella panels have clear indications for an<br />

overall, r<strong>and</strong>omly brushed underlayer which must contain lead white to register<br />

in the X-radiograph. The few samples from the pre della panels show<br />

Metzger <strong>and</strong> Berrie 131


there is a lead-white underpaint or imprimatura. X-radiographs of the central<br />

panel show a similar vigorous, striated, <strong>and</strong> r<strong>and</strong>omly applied layer in some<br />

areas.<br />

The relationship between the underdrawing <strong>and</strong> the final image. In the main <strong>and</strong><br />

predella panels, the layers of paint follow the outlines of the underdrawing in<br />

a general fashion. There are many major <strong>and</strong> minor departures from the<br />

underdrawn outlines, however, <strong>and</strong> additional elements have been added to<br />

the composition in the paint stage without any redrawing. There are numerous<br />

examples in the predel1a panels. For example, in St. Nicholas Slips a<br />

Purse through the Window if an Impoverished Nobleman, the h<strong>and</strong> <strong>and</strong> purse<br />

were painted slightly lower than they are drawn, <strong>and</strong> a shoe <strong>and</strong> stocking that<br />

were not included in the underdrawing appear. The praying h<strong>and</strong>s of the<br />

foremost boy in St. Nicholas Restores Three Dismembered Ch£ldren to Life were<br />

lowered in the paint stage <strong>and</strong> St. Nicholas's crosier was raised in the same<br />

scene.<br />

In the side panels, St. Anthony's foreground fo ot is drawn in a lower position<br />

than it was painted. The elaborate gold crosier that St. Nicholas grasps is<br />

changed from the drawn to the painted version. These changes were made<br />

at the initial stage of painting without any redrawing.<br />

In the central panel, there are fewer changes from the underdrawing to the<br />

final image, however, the Virgin's head was painted higher than it was drawn.<br />

The X-radiograph shows this change was made after the first layer of St.<br />

Anne's wimple had been laid down, after which the head was painted in its<br />

new position, without any detectable redrawing.<br />

The pigments <strong>and</strong> layer structure. The painting technique of the center panel of<br />

the altarpiece is similar to that found in the Ghent altarpiece (17). There is<br />

evidence fo r an oil-based sealing layer toned with red <strong>and</strong> black pigments<br />

over the underdrawing <strong>and</strong> under the paint. The paint structure is multilayered<br />

<strong>and</strong> the colors are worked from light to dark. The cross sections show<br />

that each large area of color was blocked in using a light tone of the final<br />

color. For example, the Virgin's mantle is underpainted in pale blue (Plate<br />

29d) ; a blue-green layer blocked in the whole area of the cloth of honor<br />

(Plate 2ge); <strong>and</strong> the throne, on which statues of putti st<strong>and</strong>, is underpainted<br />

with a lighter gray (Plate 29a) (18). Type III underdrawing provides the midtone<br />

in St. Anne's drapery, which was painted using two layers of vermilion<br />

<strong>and</strong> red lake covered by multiple layers of a transparent red glaze (Plate 29b).<br />

The paint layer structure in the side panels is not the same as in the center;<br />

the build-up of color <strong>and</strong> form is quite different. St. Anthony's robe has only<br />

a single violet-gray layer (Plate 29f). St. Nicholas's chasuble was painted by<br />

laying in a yellow ochre base, on which shadows were created with a thick<br />

dark layer of black pigment mixed with vermilion. Whereas in the center<br />

panel the layers of paint are always tonally related (e.g., red covering red) , in<br />

St. Nicholas yellow paint is used as a base for red <strong>and</strong> green in the saint's<br />

vestments. In the red embroidered chasuble, the cut velvet was painted with<br />

a layer of vermilion mixed with red lake. Red glazes were applied to depict<br />

the texture.<br />

The technique of Lamentation at the Foot if the Cross, as described by Butler,<br />

is similar to the technique employed on the predel1a panels in the use of an<br />

imprimatura containing lead white <strong>and</strong> in the relatively simple layer structure<br />

(19). Certain similarities to David's The Virgin <strong>and</strong> Child with Saints <strong>and</strong> a<br />

Donor are seen in the paint structure (20).<br />

Summary<br />

The central figural group of the center panel was produced using a different<br />

technique from the rest of the altarpiece. Its structure is similar to those found<br />

in earlier fifteenth-century paintings, such as in the Ghent altarpiece (21).<br />

This method of working-sealing the underdrawing, laying in light tones,<br />

132<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


followed by modeling layers <strong>and</strong> glazes, with the addition of new drawing as<br />

guidelines-would have lent itself to the participation of the workshop. The<br />

repositioning of the Virgin's head during the painting process could indicate<br />

the continued involvement of the master. <strong>Painting</strong> with modifications <strong>and</strong><br />

corrections fronl the underdrawing, <strong>and</strong> modeling of tone using very few<br />

layers of paint, as is seen in the side panels, requires self-assurance <strong>and</strong> authority;<br />

it would be harder to use assistants in the production of a painting<br />

when this was the method of working. Workshop participation in the center<br />

panel rather than the side <strong>and</strong> predella panels is not usually expected, although<br />

there is precedent for it.<br />

Acknowledgments<br />

The authors thank Molly Faries, Maryan Ainsworth, <strong>and</strong> John O. H<strong>and</strong> for many<br />

helpful discussions <strong>and</strong> the staff of the Department of Visual Services, National Gallery<br />

for their assistance.<br />

Notes<br />

1. Taubert,]. 1975. Pauspunkte in Tafelbildern des 15. und 16.Jahrhunderts. Bulletin<br />

de l'Institute Royale du Patrimoine Artistique (XV) :390.<br />

2. Bruges, State Archives, Aanwinsten 3804: Register van de burgerlijke vonissen<br />

van de Schepenkamer te Brugge 1518-1519, fols. 82v-83r. See also Parmentier,<br />

R. A. 1937. Bescheiden omtrent Brugsche Schilders in de 16e eeuw: 1. Ambrosius<br />

Benson. H<strong>and</strong>elingen van het genootschap Societe d'Emulation te Brugge LXXX, 92-<br />

3. See also Marlier, G. 1957. Ambrosius Benson et la peinture a Bruges au temps de<br />

Charles-Quint, Damme, 42-43.<br />

3. The St. Anne Altarpiece was the subject of the Curatorial Colloquy III of the<br />

Center fo r Advanced Study in the Visual Arts, 20-24 May 1991.<br />

4. H<strong>and</strong>, ]. O. 1986. In Early Netherl<strong>and</strong>ish <strong>Painting</strong>. Eds. ]. 0. H<strong>and</strong> <strong>and</strong> M. Wolff.<br />

Washington: National Gallery of Art, 63-74. See also H<strong>and</strong>,]. O. 1992. The Saint<br />

Anne Altarpiece. Exhibition brochure. Washington: National Gallery of Art.<br />

5. Klein, P 1986. Dendrochronological Report on The St. Anne Altarpiece. 23 April 1986.<br />

Klein determined that there are two groups of boards. Three of the boards, two<br />

from St. Anthony <strong>and</strong> one from St. Nicholas, come from the same tree. The<br />

felling date of this tree was estimated as 1496 (using the assumption of fifteen<br />

sapwood growth rings) . The other group of boards have rings that suggest a<br />

felling date of 1402. Assuming the wood was stored for ten years, Klein estimates<br />

the altarpiece could have been manufactured from 1506 onward.<br />

6. von Bodenhausen, E. E 1905. Gerard David und seine Schule. Munich: E Bruckman.<br />

7. Marlier, op. cit. (note 2), 46-50.<br />

8. Scillia, D. G. 1975. Gerard David <strong>and</strong> Manuscript Illumination in the Lowl<strong>and</strong>s.<br />

Ph.D. diss. Case Western Reserve University, Ohio.<br />

9. van Miegroet, H.]. 1989. Gerard David. Antwerp: Mercatorfonds, 313-14.<br />

10. Verougstraete-Marcq, H., <strong>and</strong> R. Van Schoute. 1989. Cadres et supports dans la<br />

peinture fiam<strong>and</strong>e aux 15e et 16e siecles. Heure-Ie-Romain, 9, 76-77.<br />

11. Gettens, R.]., E. W FitzHugh, <strong>and</strong> R. L. Feller. 1993. Calcium Carbonate Whites.<br />

In Artists' Pigments: A H<strong>and</strong>book of their History <strong>and</strong> Characteristics, Vo!' 2. Washington:<br />

National Gallery of Art, 203-26.<br />

12. Infrared reflectography of the main panels was undertaken (1981-1982) by Molly<br />

Faries <strong>and</strong> in 1990 by Maryan Ainsworth. Reflectography of the Toledo panels<br />

was completed in 1984 by Faries, <strong>and</strong> of the Edinburgh panels in 1990 by the<br />

National Gallery of Art. Faries used a high resolution Grundig FA 70H video<br />

camera with a TV Macromar 1:2.8/36 mm lens <strong>and</strong> a Kodak 87 A Wratten filter<br />

on a Hammamatsu N214 vidicon detector. The assemblies were prepared by<br />

photographing the image as it appeared on the monitor <strong>and</strong> pasting the photographs<br />

together.<br />

13. A Leitz Orthoplan microscope was used with apochromat lenses. The samples<br />

were mounted in Cargille Meltmount <strong>and</strong> examined at xl00-630.<br />

14. Faries, M. 1986. Division of labour in the production of panel paintings in the<br />

workshop of Jan van Score!' In Kunst voor de beeldenstorm. Exhibition catalogue.<br />

's-gravenhage: 113.<br />

15. Coremans, P, R. ]. Rutherford, <strong>and</strong> ]. Thissen. 1952. La technique des Primitifs<br />

Flam<strong>and</strong>s. Etude scientifique des materiaux, de la structure et de la technique<br />

picturale. Studies in Conservation (1):1-29.<br />

Metzger <strong>and</strong> Berrie 133


16. Brinkman, P W. E, L. Kockaert, L. Maes, L. Masschelein-Kleiner, E Robaszynski,<br />

<strong>and</strong> E. Thielen. 1984. Het lam Godsretabel van van Eyck. Een heronderzoek<br />

naar de materialen en schildermethoden. 1. De plamuur, de isolatie!aag, de tekening<br />

en de grondtonen. Bulletin Institut Royale Patrimoine Artistic-KIK (20):137-<br />

66.<br />

17. Brinkman, P W. E, L. Kockaert, L. Maes, E. M. M. Thielen, <strong>and</strong> J. Wouters. 1988-<br />

1989. Het lam Godsretabe! van van Eyck een Heronderzoek naar de materialen<br />

en schildermethoden. 2. De hoofdkleuren blauw, groen, gee! en rood. Bulletin<br />

Institut Royale Patrimoine Artistic-KIK (20):26-49. See also Coremans, et al., op.<br />

cit. <strong>and</strong> Brinkman, et a]., 1984, op. cit.<br />

18. A lEOL 6300 scanning electron microscope <strong>and</strong> a Link energy dispersive spectrometer<br />

with the Super ATW Si(Li) detector were used. Analysis of a green<br />

particle in the bottom layer of paint in the cloth of honor showed it was composed<br />

of copper, chlorine, <strong>and</strong> oxygen.<br />

19. Butler, M. H. 1976. A technical investigation of the materials <strong>and</strong> technique used<br />

in two Flemish paintings. Museum Studies (The Art Institute of Chicago) (8):59-<br />

71.<br />

20. Wyld, M., A. Roy, <strong>and</strong> A. Smith. 1979. Gerard David's "The Virgin <strong>and</strong> Child<br />

with Saints <strong>and</strong> a Donor." National Callery Technical Bulletin, Vol. 3. London:<br />

National Gallery, 5-65.<br />

21. Brinkman, 1988-1989, op. cit. See also Coremans, et al., op. cit. <strong>and</strong> Brinkman,<br />

et al., 1984, op. cit.<br />

134<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

Based on extensive information from<br />

cross sections <strong>and</strong> infrared reflectography,<br />

this paper presents some aspects<br />

of painting technique that were<br />

held in common by Jan van Scorel,<br />

the head of a productive sixteenthcentury<br />

North Netherl<strong>and</strong>ish workshop,<br />

<strong>and</strong> Maarten van Heemskerck,<br />

his best-known assistant. Some of the<br />

idiosyncrasies in Heemskerck's painting<br />

technique differ from Scorel's<br />

studio routine <strong>and</strong> are more apparent<br />

in this artist's early works.<br />

Maarten van Heemskerck <strong>and</strong> Jan van Scorel's Haarlem<br />

Workshop<br />

Molly Faries*<br />

Henry Radford Hope School of Art<br />

Indiana University<br />

Bloomington, Indiana 47405<br />

USA<br />

Christa Steinbuchel<br />

Wallraf-Richartz-Museum<br />

Bischofsgarten 1<br />

W-5000 Kaln<br />

Germany<br />

J. R. J. van Asperen de Boer<br />

Groningen University<br />

Department for the History of Art <strong>and</strong> Architechture<br />

Oude Boteringestraat<br />

p. 0. Box 716<br />

9700 AS Groningen<br />

The Netherl<strong>and</strong>s<br />

Introduction<br />

From the records of the Mariakerk in Utrecht where Jan van Scorel held<br />

clerical office, the precise dates of his stay in Haarlem are known: 29 April<br />

1527 to 28 September 1530. Karel van M<strong>and</strong>er's Schilder-Boeck, the primary<br />

source on the workshop Scorel established in Haarlem during that time, reports<br />

that Scorel rented a house in order to take on students. In the biographies<br />

of other sixteenth-century North Netherl<strong>and</strong>ish painters, including<br />

Jan Swart van Groningen or Jan Vermeyen, M<strong>and</strong>er implies that these painters<br />

were either Scorel's students or were somehow in contact with his shop.<br />

Modern scholars have added other likely (<strong>and</strong> not so likely) names to M<strong>and</strong>er's<br />

list, including Comelis Buys, Herman Postma, <strong>and</strong> Jan Stephan van Calcar,<br />

but the assistant whom M<strong>and</strong>er described at some length was Maarten<br />

van Heemskerck. According to M<strong>and</strong>er, Heemskerck applied himself diligently<br />

in Scorel's Haarlem studio, eventually producing works that were indistinguishable<br />

from those of his master. In M<strong>and</strong>er's rather dramatic account,<br />

Heemskerck finally surpassed the de Const (artistry) of his master, only to be<br />

dismissed from the shop, ostensibly because of Scorel's jealousy (1). However<br />

anecdotal this story might seem, art historians have indeed had difficulties<br />

distinguishing early works by Heemskerck from Scorel's work. For most of<br />

the twentieth century, early Heemskerck artworks were attributed to Scorel;<br />

it was not until the 1980s that several key attributions were changed, primarily<br />

in the 1986 Art Before Iconoclasm exhibition (2). Jeff Harrison's recently<br />

published dissertation on Heemskerck is the first to outline a new <strong>and</strong> plausible<br />

chronology for his early works (3).<br />

Jan van Scorel's technique<br />

Full technical studies have been carried out by Molly Faries <strong>and</strong> J. R. J. van<br />

Asperen de Boer on a number of paintings by Jan van Scorel from this period,<br />

including examination by binocular microscope, infrared reflectography, X­<br />

radiography, dendrochronology, <strong>and</strong> sampling. Until the new shifts in attribution,<br />

few early works of Heemskerck had been studied as thoroughly. This<br />

situation changed with the recent cleaning of Lamentation (Wallraf-Richartz­<br />

Museum, Cologne), a painting attributed variously to Scorel, Heemskerck, or<br />

an anonymous artist from the same period. The research in conjunction with<br />

this restoration, carried out under the direction of Christa Steinbuchel, has<br />

* Author to whom correspondence should be addressed.<br />

Faries, Steinbuchel, <strong>and</strong> van Asperen de Boer 135


Figure 1. Jan van Scoref, Baptism of Christ, ca. 1527-1530. Photograph by M. Faries (after<br />

Dingjan), courtesy of Frans Hafsmuseum, Haarlem.<br />

provided critical comparative evidence (4). The results help to define the art<br />

historical attributions <strong>and</strong> suggest some changes to Harrison's chronology.<br />

This material not only brought greater clarity regarding the typical painting<br />

procedures used in Scorel's shop, but also signaled the steps Heemskerck took<br />

in his evolution away from Scorel's shop.<br />

During the Haarlem years, Scorel st<strong>and</strong>ardized his painting technique in response<br />

to the needs of an active workshop (5). The most overt clue to the<br />

st<strong>and</strong>ardization is revealed in the layout of underdrawings from this period.<br />

The Baptism if Christ in the Frans Halsmuseum, Haarlem, must have been<br />

one of the most prestigious commissions the artist received 1527-1530 (Fig.<br />

1). The layout of this painting is fully worked out, from the assured contours<br />

<strong>and</strong> loosely marked shaded zones in the figures of the main scene to the light<br />

<strong>and</strong> dark b<strong>and</strong>s in the l<strong>and</strong>scape <strong>and</strong> other background detail. Areas where a<br />

known motif was to be placed were no longer just left as a blank space in<br />

the underdrawing, but marked with indicative shapes referring to the motif.<br />

The underdrawing was laid out as a more recognizable <strong>and</strong> emphatic guide.<br />

The Baptism was also painted following what had evolved into Scorel's usual<br />

practice. This practice comprised the following: (a) an application of lead<br />

white as a continuous intermediate layer between the ground <strong>and</strong> paint, (b)<br />

an underdrawing in black chalk on this layer, <strong>and</strong> (c) a preference fo r certain<br />

paint-layer structures <strong>and</strong> color combinations. A white intermediate layer certainly<br />

occurs in artistic groups other than that of Scorel's workshop. It has<br />

been found in some early German panels, <strong>and</strong> sporadically in the Hans Memling<br />

<strong>and</strong> Gerard David groups, for instance. Scorel, however, could not have<br />

learned the use of such a layer from his master in Amsterdam. It begins to<br />

show up in some of Scorel's early works produced during his journey to Italy,<br />

but only appears consistently after his return north. It was the three-fold<br />

fu nction of this layer that was peculiar to the efficacy of Scorel's painting<br />

technique: it isolated the ground, added to the painting's luminosity, <strong>and</strong> provided<br />

"tooth" for the underdrawing, a fu nction almost unique to the Scorel<br />

group (6). In many infrared documents, the underdrawing can be seen to skip<br />

or crumble on top of this ridged surface. One cross section from Baptism<br />

shows this typical paint layer structure: a first layer consisting of lead white<br />

at a maximum thickness of 12 fL, a second layer composed of clumps of black<br />

136<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


o . (Vv..-.<<br />

1 2.<br />

underdrawing, <strong>and</strong> a third layer consisting of a dark green mixture of green,<br />

blue, <strong>and</strong> white; no ground is used (Fig. 2). In addition, Scorel often paints<br />

wet in wet with the usual run of pigments, except for his use of natural<br />

ultramarine <strong>and</strong> a blue-over-rose structure. The latter superimposition of colors,<br />

with ultramarine on the surface (as seen again in a cross section from<br />

Baptism), has been linked to Scorers observation of Italian painting technique<br />

(Plate 30) (7).<br />

Figure 2. Cross section from Baptism with<br />

measurements in microns: layer 1, traces of<br />

the intermediate white layer; layer 2, particles<br />

of black chalk underdrawing; layer 3, mixture<br />

of light green, blue, <strong>and</strong> white. Photograph<br />

by M. Faries (after J. R. J. van Asperen de<br />

Boer).<br />

Marteen van Heemskerck's technique compared<br />

Before the conservation of the Cologne Lamentation <strong>and</strong> the recent revisions<br />

in art historical opinion, what was known of Heemskerck's painting technique<br />

showed little connection to Scorel (Fig. 3). The current acceptance of Lamentation,<br />

long attributed to Scorel, as a Heemskerck by scholars Faries, Harrison,<br />

<strong>and</strong> W Th. Kloek is by no means unchallenged. A study of the painting's<br />

technique shows why a separation of "h<strong>and</strong>s" has been so difficult.<br />

Lamentation was begun as a Scorel studio piece <strong>and</strong> finished as a Heemskerck,<br />

as seen in both the underdrawing <strong>and</strong> the painting technique. Infrared reflectography<br />

examination, undertaken by Faries in 1991, disclosed exceptionally<br />

complicated compositional change in this work (8). The painting was<br />

taken through as many as six stages while the composition was changed from<br />

a profile Scorelesque Entombment scene to a Heemskerck frontal presentation<br />

of the Lamentation.<br />

The numerous samples required because of the piece's complex restoration<br />

history provided more than enough evidence about Lamentation's painting<br />

procedure. The initial preparation of the ground <strong>and</strong> position of the underdrawing<br />

in the paint layer structure match the st<strong>and</strong>ard practice of Scorers<br />

Haarlem shop. A thin layer of lead white covers the entire surface of the<br />

ground; it appears consistently in sections that show the entire paint layer<br />

structure. The black chalk underdrawing is found on top of the intermediate<br />

layer (9). Although some of the lines of the different compositional stages<br />

must have crossed, no noticeable overlapping or disjunctures in any samples<br />

that include the underdrawing layer can be seen. The presence of the lead<br />

white intermediate layer coupled with the black chalk underdrawing can be<br />

Figure 3. Maarten van Heemskerck, Lamentation of Christ, ca. 1530. Photograph by C. Steinbachel,<br />

courtesy of Wallraj-Richartz-Museum, Cologne.<br />

Faries, Steinbiichel, <strong>and</strong> van Asperen de Boer 137


considered a kind of technical signature, although not of an individual h<strong>and</strong>:<br />

It is the signature of Scorel's shop. Most of the pigments used in completing<br />

the image are, of course, also common to those found in Scorel's paintings.<br />

No ultramarine has been found, however, <strong>and</strong> the blue used throughout the<br />

painting is azurite. This no doubt proves not only that Jan van Scorel is the<br />

only north Netherl<strong>and</strong>ish artist documented to have used natural ultramarine<br />

in his works, but that he also used it selectively, saving it for his most important<br />

commissions (10). It was, therefore, not available to his assistant, or at<br />

least not for this painting. Not unexpectedly, the blue-over-rose paint layer<br />

structure is also lacking in Lamentation.<br />

Figure 4. Infrared reflectogram detail if<br />

Lamentation, showing the dark undermodelil1g<br />

strokes if Nicodemus's turbal1 in the upper<br />

background. I'!frared reflectography by M.<br />

Faries.<br />

;;<br />

I<br />

I<br />

s) (.10<br />

'1) 10-'3£<br />

3) q -2-1<br />

Z) g-I&<br />

Figure 5. Cross sectiol1 from Maarten van.<br />

Heemskerck's St. Luke with measuremen.ts in<br />

microns: layer 1, possible trace of ground;<br />

layer 2, gray intermediate layer; layer 3,<br />

compact layer if black with black particles;<br />

layer 4, admixture if white, black, <strong>and</strong> red;<br />

layer 5, white layer with one large azurite<br />

crystal; layer 6, varnish. Photograph by M.<br />

Faries (after J. R. J. van Asperen de Boer).<br />

Other ways in which the painting practice in Lamentation deviates from ScoreI's<br />

st<strong>and</strong>ard require a different explanation, since these differences form critical<br />

links with other paintings that can now be considered early Heemskerck.<br />

Dark modeling strokes executed with a brush have been detected in the<br />

finishing stages of Lamentation. Some strokes that underlie the surface colors<br />

can be disclosed by infrared (Fig. 4). Other black-colored modeling is applied<br />

in the final paint layer, especially in reds. The cross section from John the<br />

Evangelist's drapery clearly shows the black particles in the vermilion <strong>and</strong> red<br />

lake admixture (11).<br />

Although Scorel's Baptism <strong>and</strong> Heemskerck's Lamentation are the focus of this<br />

study, their techniques may be compared to a substantial amount of related<br />

technical material. Ample evidence proves that the paint-layer structure described<br />

above typifies the Scorel group paintings during Scorel's time in Haarlem.<br />

In addition, since nine paintings now attributed by some scholars to the<br />

early Heemskerck period have also been examined by various technical<br />

means, the evolution of Heemskerck's painting technique can be surveyed<br />

(12). Of the nine paintings, only four have been studied with the binocular<br />

microscope <strong>and</strong> sampled so far. At least three paintings use the blue-overrose<br />

structure typical of the Scorel group although, in each case, the blue<br />

pigment is azurite. Other aspects of the paint layer structure can be documented<br />

by X-ray or infrared reflectography. The infrared vidicon distinguishes<br />

pure colors that look black to our eyes but are transparent to the eye of the<br />

vidicon from "true" blacks, which remain opaque in reflectograms. This instrumental<br />

technique can therefore locate black modeling or under modeling<br />

in the same way it locates underdrawings. Dark modeling strokes used to<br />

block in forms can also be found in at least three other works in the hypothetical<br />

early Heemskerck group. The technique is unknown in the Scorel<br />

group. The presence of a thin layer of lead white over the ground can be<br />

verified in cross sections, as well as by X-radiography where it appears as<br />

broadly-brushed streaking. This practice is evident in a majority of works in<br />

the early Heemskerck group. It is the coupling of underdrawing with this<br />

layer, however, which undergoes a change; <strong>and</strong> finally, the character of the<br />

layer itself changes. In Lamentation, the black chalk underdrawing is supplemented<br />

by dark modeling <strong>and</strong> undermodeling in paint. Several other paintings<br />

have a fo rm of underdrawing, but it is executed in a dark, paintlike<br />

substance rather than chalk. In four works, the underdrawing is undetectable.<br />

No underdrawing could be detected in the St. Luke <strong>Painting</strong> a Portrait of the<br />

Virgin <strong>and</strong> Child (1532), an indisputable early Heemskerck, <strong>and</strong> no underdrawing<br />

appeared in samples. In this work Heemskerck used a gray intermediate<br />

layer, a very different basis upon which to build up the fo rms of his<br />

painting (Fig. 5) (13).<br />

Conclusion<br />

Heemskerck must have worked in Scorel's studio for all or part of the period<br />

from 1527 to 1530, <strong>and</strong> after Scorel left Haarlem in September 1530, Heemskerck<br />

continued to paint in the city as an independent master until he himself<br />

left for Rome in 1532. The artist moved from a graphic black-on-white<br />

layout to more subtle forms of shading <strong>and</strong> undermodeling; based on changes<br />

in painting technique, an evolution is proposed for Heemskerck in this period.<br />

138<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


The key position of Lamentation in this proposed development is obvious, for<br />

it encapsulates both painting methods. Generally, those paintings that still rely<br />

on underdrawing (on the intermediate white layer) as part of the basic layout<br />

have an early date. Whether or not the paintings were produced in Scorers<br />

shop probably cannot be proven, but they were painted following Scorers<br />

shop routine. The paintings without underdrawing but with black-colored<br />

modeling appear to be later works, closer to Heemskerck's dated works of<br />

1531 <strong>and</strong> 1532. A clearer art historical picture of the early Heemskerck is<br />

now possible, but surely the blending of Scorers painting practices into the<br />

build-up of some of the earliest works explains the difficulties art historians<br />

have had in the past in attributing these paintings.<br />

Notes<br />

1. van M<strong>and</strong>er, K. 1604. Het Schilder-Boeck. Haarlem. Davaco ed. 1969, fol. 245.<br />

2. Kunst voor de beeldenstorm. 1986. Exhibition catalogue. Eds. J. P. Filedt Kok, W.<br />

Halsema-Kubes, W. Th. Kloek. Amsterdam: Rijksmuseum. Especially catalogue<br />

number 69, Rest on the Flight into Egypt (National Gallery of Art, Washington,<br />

D.c.) <strong>and</strong> catalogue number 72, Portrait oj a Young Scholar (Boymans van Beuningen<br />

Museum, Rotterdam), which were definitively assigned to Heemskerck.<br />

Around this same time, Adam <strong>and</strong> Eve (Hatfield House, Marquess of Salisbury)<br />

was being shifted from Scorel to Heemskerck by the research of Faries <strong>and</strong><br />

Harrison.<br />

3. Harrison, J. C. 1992. The <strong>Painting</strong>s oj Maerten van Heemskerck. (Ph.D. diss., 1987,<br />

University of Virginia.) Ann Arbor, Michigan, esp. 11-34.<br />

4. For the restoration, see C. Steinbuchel, "The investigation <strong>and</strong> restoration of<br />

Maarten van Heemskerck's Lamentation oj Christ. " In Le dessin sous-jacent dans la<br />

peinture, Colloque X: Le dessin sous-jacent dans le processus de creation, eds. R. Van<br />

Schoute <strong>and</strong> H. Verougstraete-Marcq. Universite Catholique de Louvain, Institut<br />

superieur d'archeologie et d'histoire de l'art. Forthcoming.<br />

5. Faries, M. 1987. Some results of the recent Score! research: Jan van Scorel's<br />

definition of l<strong>and</strong>scape in design <strong>and</strong> color. In Color <strong>and</strong> Technique in Renaissance<br />

<strong>Painting</strong>, Italy <strong>and</strong> the North. Ed. M. B. Hall. New York, 89-103, esp. 96.<br />

6. van Asperen de Boer, J. R. J., M. Faries, <strong>and</strong> J. P. Filedt Kok. 1986. <strong>Painting</strong><br />

technique <strong>and</strong> workshop practice in northern Netherl<strong>and</strong>ish art of the sixteenth<br />

century. In Kunst voor de beelderlstorm. Exhibition catalogue. Eds. J. P. Fi!edt Kok,<br />

W. Halsema-Kubes, W. Th. Kloek. Amsterdam: Rijksmuseum, esp. 108.<br />

7. Faries, op. cit., 92.<br />

8. For these infrared reflectography results, see M. Faries, "Attributing the layers of<br />

Heemskerck's Cologne Lamentation oj Christ." In Le dessin sous-jacent dans la peinture,<br />

Colloque X: Le dessin sous-jacent dans le processus de creation. Eds. R. Van<br />

Schoute <strong>and</strong> H. Verougstraete-Marcq. Universite Catholique de Louvain, Institut<br />

superieur d'archeologie et d'histoire de l'art. Forthcoming.<br />

9. Sample numbers 1,2, <strong>and</strong> 27.1 show this structure. Communication from Christa<br />

Steinbuchel to Molly Faries, 7 October 1994.<br />

10. van Asperen de Boer, et aI., op. cit., 109.<br />

11. Sample number 18. Communication from Christa Steinbuchel to Molly Faries,<br />

7 October 1994.<br />

12. The works studied, in presumed chronological order, are as follows: ca. 1527-<br />

1530: (1) Lamentation oj Christ, private collection, New York; (2) Crucifixion,<br />

Detroit Institute of Arts; (3) Rest on the Flight into Egypt, National Gallery of Art,<br />

Washington, D.c.; (4) Lamentation oj Christ, Wallraf-Richartz-Museum, Cologne;<br />

(5) 1531: Portrait oj a Scholar, Museum Boymans-van Beuningen, Rotterdam; (6)<br />

1531?: Baptism oj Christ, Staatliche Museen Preuf3ischer Kulturbesitz, Berlin; (7)<br />

1532: Christ as the Man oj SorroUls, Museum voor Schone Kunsten, Ghent; (8)<br />

1532: St. Luke <strong>Painting</strong> a Portrait oj the Virgin <strong>and</strong> Cltild, Frans Halsmuseum,<br />

Haarlem; (9) ca. 1530-1532: Family Portrait, Staatliche Kunstsanmilungen, Kassel.<br />

For the lack of underdrawing revealed in the 1529 so-called Bicker Portraits in<br />

Amsterdam, Rijksmuseum, see Kunst voor de beeldenstorm, op. cit., catalogue number<br />

71.1-2.<br />

13. van Asperen de Boer, J. R. J. 1987. A technical study of some paintings by<br />

Maarten van Heemskerck. In Color <strong>and</strong> Technique in Renaissance <strong>Painting</strong>, Italy <strong>and</strong><br />

the North. Ed. M. B. Hall. New York, 105-1 14.<br />

Faries, Steinbiichel, <strong>and</strong> van Asperen de Boer 139


Abstract<br />

In the 1620s, a generation of Dutch<br />

l<strong>and</strong>scape artists began to work in a<br />

naturalistic mode very different from<br />

that of the earlier generation of<br />

Flemish mannerist l<strong>and</strong>scape artists, a<br />

number of whom had recently emigrated<br />

to the Northern Netherl<strong>and</strong>s.<br />

The change from fantastic l<strong>and</strong>scape<br />

to representations of Dutch scenes<br />

reflected political <strong>and</strong> economic<br />

changes as the Northern Netherl<strong>and</strong>s<br />

established independence from Spanish<br />

domination. This stylistic change<br />

is reflected in changes in the painting<br />

materials <strong>and</strong> practices of the realist<br />

painters. In the 1620s, Dutch<br />

painters of naturalistic l<strong>and</strong>scape<br />

adapted the efficient working practices<br />

of the Flemish l<strong>and</strong>scape<br />

painters. They replaced the refined<br />

h<strong>and</strong>ling of paint <strong>and</strong> bright colors<br />

of the mannerist painters with limited<br />

tonalities <strong>and</strong> an abbreviated h<strong>and</strong>ling<br />

of paint to create convincing<br />

views of the Dutch l<strong>and</strong>scape.<br />

Style <strong>and</strong> Technique in Dutch L<strong>and</strong>scape <strong>Painting</strong><br />

in the 1620s<br />

E. Melanie Gifford<br />

Scientific Research Department<br />

National Gallery of Art<br />

Washington, D.c. 20565<br />

USA<br />

Introduction<br />

The early years of the seventeenth century saw a striking change of style in<br />

l<strong>and</strong>scape painting in the Netherl<strong>and</strong>s. At this time, Flemish l<strong>and</strong>scape specialists<br />

produced paintings characterized more by fantasy than by close observation<br />

of nature. In the 1580s <strong>and</strong> 1590s, large numbers of Flemish artists,<br />

some of them l<strong>and</strong>scape specialists, had emigrated to Holl<strong>and</strong> to escape the<br />

political <strong>and</strong> economic hardships of the rebellion against Spanish rule (1).<br />

Around 1620, a very different, naturalistic style of l<strong>and</strong>scape painting developed<br />

in the Northern Netherl<strong>and</strong>s, particularly around Haarlem. These paintings,<br />

often based on drawings after nature, created distinctive images of the<br />

Dutch l<strong>and</strong>scape (2, 3, 4).<br />

Such a dramatic change of style raises fascinating questions about how style<br />

develops. Was the new, naturalistic l<strong>and</strong>scape style sparked by the arrival of<br />

the Flemish immigrants, or does it reflect an indigenous artistic sensibility? Is<br />

artistic style dependent on painting practice learned from other artists, or do<br />

artists modifY their practice to meet the dem<strong>and</strong> for new styles?<br />

In an ongoing technical study, the author has been looking for material evidence<br />

for l<strong>and</strong>scape artists' artistic concerns (5). This study seeks to characterize<br />

the differences of technique between Flemish mannerist l<strong>and</strong>scapes of<br />

the turn of the sixteenth <strong>and</strong> seventeenth centuries, l<strong>and</strong>scapes by Flemish<br />

immigrants to the Northern Netherl<strong>and</strong>s during the same period, <strong>and</strong> the<br />

first naturalistic Dutch l<strong>and</strong>scapes, which date from the second <strong>and</strong> third<br />

decade of the seventeenth century (6). In defining the artists' choices of<br />

working methods <strong>and</strong> painting materials, the study seeks to exp<strong>and</strong> our<br />

underst<strong>and</strong>ing of the motivations, both fo rmal <strong>and</strong> practical, behind the<br />

development of the naturalistic Dutch l<strong>and</strong>scape in the 1620s. The present<br />

contribution centers on developments in the h<strong>and</strong>ling of paint in these l<strong>and</strong>scapes.<br />

From the paintings studied to date, I have chosen five paintings as<br />

illustration: a Flemish mannerist l<strong>and</strong>scape, a mannerist l<strong>and</strong>scape by a Flemish<br />

immigrant, two naturalistic l<strong>and</strong>scapes by Dutch artists, <strong>and</strong> a somewhat later<br />

Dutch "tonal" l<strong>and</strong>scape. A later, more complete publication will present full<br />

analyses of the paintings' materials.<br />

Flemish mannerist l<strong>and</strong>scape<br />

In the Flemish l<strong>and</strong>scape painting tradition, space is defined by formal conventions<br />

that convey an illusion of recession. Dark <strong>and</strong> light passages alternate,<br />

often with shadowed foreground repoussoirs sharply outlined against a brightly<br />

lit area beyond. Optical phenomena such as atmospheric perspective are represented<br />

not illusionistically, but almost symbolically, by a space organized into<br />

three zones defined by distinctly different tonalities. The darkened foreground<br />

zone is typically a rich brown, the middle ground is green, <strong>and</strong> the distance<br />

is a clear blue.<br />

Study of the technique of Flemish mannerist l<strong>and</strong>scapes reveals refined h<strong>and</strong>ling<br />

combined with efficient working methods, which contributed to the<br />

high output of busy workshops. The compositions were usually laid out with<br />

a fairly complete underdrawing, most often including the figures, which were<br />

planned as integral parts of the composition. With the underdrawing as guide,<br />

140<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 1. Workshop of Jan Bruegel, Noah's Ark. Panel, 65. 4 X 94.5 em. The Walters Art Gallery, Baltimore (3 7. 1998).<br />

space for the figures was left in reserve as the l<strong>and</strong>scape was painted. Some<br />

l<strong>and</strong>scape specialists collaborated with figure painters, in which case the figures<br />

would be painted over a fully completed l<strong>and</strong>scape. Other workshops<br />

divided up production in a sort of assembly line, with junior associates filling<br />

in minor details.<br />

The painting materials in the Flemish mannerist group, while not unusual,<br />

were used in ways that enhanced the clear colors. Typically, a light-colored,<br />

opaque ground maintained the luminous tonalities of the paint, but rarely is<br />

the ground itself visible in the final image. Instead, the three dominant zones<br />

of the composition were laid out with broad areas of underpaint: warm brown<br />

in the foreground, soft green in the middle zone, <strong>and</strong> light blue in the distance.<br />

The final image was worked up over these base colors. Working from the<br />

back of the composition to the foreground, each zone was painted in turn<br />

in a limited range of colors harmonizing with that of the underpaint.<br />

A painting of Noah 's Ark at the Walters Art Gallery was produced by the<br />

workshop of Jan Bruegel (Fig. 1) (7, 8). In this painting, the techniques described<br />

above were adapted for team production in a workshop setting. As<br />

shown in a paint cross section from high in the central group of trees, a thick<br />

white chalk ground was sealed <strong>and</strong> lightly toned by a translucent imprimatura<br />

(layers 1 <strong>and</strong> 2) (Plate 31) (9). Over the preparation <strong>and</strong> underdrawing, the<br />

sky was laid in first, followed by a base color loosely brushed for each particular<br />

area of the composition; the upper two layers of the cross section show<br />

the sky (layer 3), which extends under the foliage base tone in the upper part<br />

of the trees, <strong>and</strong> the clear green underpaint (layer 4) of the lightest passage<br />

of the central group of trees. A skilled painter painted the main foliage elements<br />

<strong>and</strong> the figures <strong>and</strong> animals, which are the primary subjects of the<br />

painting, onto the broad areas of underpaint. Only after these major elements<br />

GijJord 141


Figure 2. Roel<strong>and</strong>t Savery, L<strong>and</strong>scape with Animals <strong>and</strong> Figures, signed <strong>and</strong> dated 1624. Panel, 54.6 X 91.4 cm. National Gallery if Art,<br />

Washington, D. C. (1989.22. 1). Gift of Robert H. <strong>and</strong> Clarice Smith in honor if the fiftieth anniversary if the National Gallery of Art.<br />

were completed did a workshop assistant fill in minor foliage details. The<br />

assistant's work has a mechanical, repetitive quality, <strong>and</strong> shows a cautious respect<br />

for the work of the master; the monkey on a low branch at the left is<br />

surrounded by a halo of the base color, where the assistant scrupulously avoided<br />

overlapping the master's work with his blades of grass.<br />

L<strong>and</strong>scapes by Flemish immigrants<br />

L<strong>and</strong>scapes by Flemish immigrants such as Gillis van Coninxloo, who arrived<br />

in Amsterdam from Frankenthal in 1595, <strong>and</strong> Roel<strong>and</strong>t Savery, who arrived<br />

with his family in 1591, are clearly in the Flemish tradition, both compositionally<br />

<strong>and</strong> technically (10). There are, however, variations of emphasis in<br />

the compositions that have consequences for the technique. The compositions<br />

retain the conventions of Flemish l<strong>and</strong>scape. Dark <strong>and</strong> light passages are<br />

strongly juxtaposed <strong>and</strong> the three zones in brown, green, <strong>and</strong> blue organize<br />

the recession into space, but there is a much greater emphasis on the nearer<br />

zones. In Coninxloo's forest l<strong>and</strong>scapes, such as the L<strong>and</strong>scape with Hunters of<br />

1605 in Speyer, the second <strong>and</strong> third zones are reduced to glimpses caught<br />

through the dense growth of monumental trees that fills the foreground zone<br />

(11). The result is a newly limited tonal range, one that emphasizes the browns<br />

<strong>and</strong> deeper greens of the middle <strong>and</strong> foreground. This self-imposed limitation<br />

is in itself a first step toward naturalism.<br />

The techniques of these immigrant artists are consistent with contemporaries<br />

still working in Fl<strong>and</strong>ers. The compositions were planned in underdrawings<br />

that made provision for all but the minor figures. Over a light-colored ground<br />

the artists established a base color for each area with a brushy underpaint;<br />

each area was then worked up with loosely painted, final details in a color<br />

that harmonized with the underpaint showing through from below. To avoid<br />

the danger of monotony in their emphasis on the brown foreground <strong>and</strong><br />

green middle zones, the artists introduced a range of subtle variations within<br />

the harmonies of browns <strong>and</strong> greens.<br />

142<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 3. Camelis Vroom, L<strong>and</strong>scape with a River by a Wood, signed <strong>and</strong> dated 1626. Panel, 31.3 X 44.2 em. National Gallery, London (3475).<br />

In Roel<strong>and</strong>t Savery's L<strong>and</strong>scape with Animals <strong>and</strong> Figures (1624), an area for<br />

the foreground group of figure <strong>and</strong> animals was held in reserve while the<br />

l<strong>and</strong>scape was painted; only the small figures to the rear are painted over the<br />

l<strong>and</strong>scape (Fig. 2). The foreground of this work is painted with final details<br />

in a typical transparent brown over a warm brown underpaint. In the lighter<br />

middle ground of this painting, the same transparent brown is used for the<br />

details over passages of tan underpaint; in a passage underpainted in gray, such<br />

as the stream at the left, the details are worked up in rapidly sketched strokes<br />

of darker gray paint.<br />

L<strong>and</strong>scape in Haarlem in the 1620s<br />

By the 1620s, a number of Dutch painters had taken up l<strong>and</strong>scape as a specialty.<br />

In contrast to the fantasy l<strong>and</strong>scapes of the Mannerists, these paintings<br />

represented recognizable scenes of the local countryside. The colors are limited,<br />

the compositions rely on subtle atmospheric effects to create almost<br />

continuous recession into space, <strong>and</strong> the figures no longer play a significant<br />

role. The techniques of l<strong>and</strong>scape painting in this early period are varied, but<br />

there is evidence that those Haarlem artists who were most adventuresome<br />

compositionally also incorporated new ways of h<strong>and</strong>ling paint, methods<br />

which became integral to the effects of the "tonal l<strong>and</strong>scape" painters in the<br />

1630s.<br />

Comelis Vroom painted the L<strong>and</strong>scape with a River by a Wood, a work now<br />

in London, in 1626 (Fig. 3) . This Haarlem l<strong>and</strong>scape artist did not participate<br />

in the tonal style in the next decade, but in this early l<strong>and</strong>scape, Vroom<br />

modified Flemish practice to create a more subtle spatial recession with his<br />

characteristic color range dominated by browns, yellows, <strong>and</strong> grayish greens.<br />

The color of each area is established using an underpaint, following the Flem-<br />

Giff ord 143


Figure 4. Esaias van de Velde, Winter L<strong>and</strong>scape, signed <strong>and</strong> dated 1623. Panel, 25. 9 X 30.4 em. The National Gallery, London (6269).<br />

ish practice, but this underpaint is more varied, not restricted to sharply<br />

defined b<strong>and</strong>s of color receding into the composition. Throughout the foreground<br />

<strong>and</strong> middle ground the underpaint is a web of grayish green <strong>and</strong><br />

yellow passages, worked wet-into-wet.<br />

The l<strong>and</strong>scape paintings of Esaias van de Velde, produced during his years in<br />

Haarlem (1609-1618) <strong>and</strong> in the Hague until his death in 1630, are central<br />

to the development of the naturalistic l<strong>and</strong>scape. Though Esaias was trained<br />

by the Flemish inunigrant Gillis van Coninxloo, he developed most of the<br />

elements of technique that were to define the work of the naturalistic tonal<br />

l<strong>and</strong>scapists, who included his pupils Jan van Goyen <strong>and</strong> Pieter de Neyn. This<br />

is particularly apparent in his winter scenes, where the bare l<strong>and</strong>scape encourages<br />

a particularly limited tonality. The Winter L<strong>and</strong>scape of 1623, now<br />

in London, is prepared with a thinly rubbed ground that barely fills the grain<br />

of the oak panel (Fig. 4). A cross section from the upper edge of the sky<br />

shows this extremely thin layer, textured on the underside by the grain of<br />

the wood (layer 1) (Plate 32). This preparation is barely perceptible on the<br />

painting; the warm pinkish tone that dominates passages, such as the distance<br />

at the right <strong>and</strong> parts of the fo reground, is created by the wood showing<br />

through the slightly tinted ground (12). In a loose <strong>and</strong> suggestive underdrawing,<br />

the artist situated the main elements of the l<strong>and</strong>scape <strong>and</strong> indicated the<br />

foliage along the horizon with a few looped strokes but made no provision<br />

for the figures (Fig. 5). Following the guide of the underdrawing, he toned<br />

144<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 5. Infrared photograph of Esaias van de Velde's Winter L<strong>and</strong>scape. National Gallery, London (6269).<br />

broad areas of the cOmpOSltlOn, leaving the panel's thinly applied ground<br />

visible in places. At this stage, the sky was painted with varying concentrations<br />

of pale smalt (layers 2 <strong>and</strong> 3), <strong>and</strong> the details of the horizon were painted<br />

wet-into-wet; fo rms in the foreground were sparsely indicated with tan <strong>and</strong><br />

gray-green underpaint. Over this underpainting the artist sketched out the<br />

details of the image in a monochrome paint ranging from gray to black, then<br />

completed the painting with a few touches of creamy mid-tones <strong>and</strong> final<br />

highlights painted wet-into-wet. When the entire l<strong>and</strong>scape was complete,<br />

the artist superimposed the small figures over it, the human presence almost<br />

incidental to the depiction of a raw winter's day.<br />

The tonal l<strong>and</strong>scape painters<br />

At the end of the 1620s, Jan van Goyen <strong>and</strong> two other Haarlem artists, Pieter<br />

Molijn <strong>and</strong> Salomon van Ruysdael, initiated a l<strong>and</strong>scape style in which a<br />

completely convincing representation of the Dutch countryside was rendered<br />

in a deliberately limited, almost monochromatic palette. A work as late as van<br />

Goyen's View oj Dordrecht Jrom the Dordtse Kil (1644) reveals the economy of<br />

h<strong>and</strong>ling that typifies these works: a wet-into-wet application of the sky <strong>and</strong><br />

horizon, followed by a sketchy monochrome design only occasionally amplified<br />

by rapid touches of lightly colored paint (Fig. 6) . The tone of the<br />

wood <strong>and</strong> the grain pattern can be seen through the thin ground; both play<br />

a role in the fmal image. The results of an earlier technical study established<br />

Gifford 145


Figure 6. Jan van Coym, View of Dordrecht from the Dordtse Kil, sigrled <strong>and</strong> dated 1644. Panel, 64. 7 X 95. 9 eln. National Callery of Art,<br />

Washil1gton, D. C. (1978. 1.11). Alicia Mellon Bruce FUl1d.<br />

that both the visible wood grain <strong>and</strong> the monochromatic palette, especially<br />

the use of almost colorless smalt in the grayish skies, resulted from conscious<br />

artistic choices rather than accident (13). The rapidity with which these works<br />

must have been painted may have had an economic motive as well, lowering<br />

the cost of production <strong>and</strong> hence enlarging the market for such works (14).<br />

Conclusion<br />

The transition from Flemish mannerist l<strong>and</strong>scape to a new Dutch form of<br />

naturalistic l<strong>and</strong>scape in the first decades of the seventeenth century represents<br />

a dramatic transition in style. The mode of mythological or biblical episodes<br />

set before brightly colored, imaginary vistas, which Flemish immigrants had<br />

brought north, was replaced in Holl<strong>and</strong> in just a few years by subdued <strong>and</strong><br />

sympathetic renderings of the Dutch countryside.<br />

The preliminary results of this study of the painting techniques of the period<br />

suggest that the change in style was accomplished by deliberate changes in<br />

painting technique. Dutch artists modified the conventionalized Flemish l<strong>and</strong>scape<br />

technique of light grounds, complete underdrawings, <strong>and</strong> colored underpaints,<br />

which established three schematic zones of space. Instead, working<br />

from drawings made from nature, they developed a sketchy shorth<strong>and</strong> style<br />

in which the properties of their painting materials were exploited to evocative<br />

effect. Thinly applied grounds <strong>and</strong> spare applications of underpaint allowed<br />

the panels themselves to play a role in the image. Underdrawings became<br />

quick notations upon which a more complete painted "sketch" was developed,<br />

<strong>and</strong> this monochrome image, only partially colored <strong>and</strong> worked up in<br />

the final stage of painting, was a dominant part of the finished work.<br />

These developments are a fascinating illustration of the ways in which political<br />

<strong>and</strong> social changes fo ster new artistic markets. As members of Dutch society<br />

146<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


at the beginning of the seventeenth century, naturalistic l<strong>and</strong>scape painters<br />

must have shared with their fellow citizens a newly awakened appreciation<br />

for the beauties of the local scene. As business people, they supplied <strong>and</strong><br />

developed a new market governed by this taste. As practitioners of painting,<br />

they developed new working methods, which were both economical <strong>and</strong><br />

superbly suited to expressing their new aesthetic.<br />

Notes<br />

1. Briels,J. G. C. A. 1976. De zuidnederl<strong>and</strong>se inmugrantie in Amsterdam en Haarlem<br />

omstreeks 1572-1630. PhD. diss. Rijksuniversiteit Utrecht.<br />

2. Bengtsson, A. 1952. Studies on the Rise of Realistic L<strong>and</strong>scape <strong>Painting</strong> in Holl<strong>and</strong><br />

1610-1625. Figure 3.<br />

3. Brown, C. 1986. Introduction. In Dutch L<strong>and</strong>scape, The Early Years: Haarlem <strong>and</strong><br />

Amsterdam 1590- 1 650. Exhibition catalogue. London: The National Gallery, 11-<br />

34.<br />

4. Freedberg, D. 1980. Dutch L<strong>and</strong>scape Prints of the Seventeenth Century. London.<br />

5. I am grateful fo r the opportunity to have studied paintings in the collections of<br />

the Frans Halsmuseum, Haarlem; the National Gallery, London; the National<br />

Gallery of Art, Washington; <strong>and</strong> the Walters Art Gallery, Baltimore. I would<br />

particularly like to thank Ashok Roy of the National Gallery <strong>and</strong> Ella Hendriks<br />

of the Frans Halsmuseum for their assistance. I appreciate very much the many<br />

discussions of this material that I have had with Arthur K. Wheelock,Jr. <strong>Painting</strong>s<br />

included in the study were exanuned with a stereonLicroscope, <strong>and</strong> with X­<br />

radiography <strong>and</strong> infrared reflectography, where available. Paint analysis was carried<br />

out on a linuted number of dispersed pigment samples <strong>and</strong> paint cross sections<br />

by light nLicroscopy <strong>and</strong> scanning electron nucroscopy with energy dispersive X­<br />

ray spectroscopy (SEM/EDS). The compositions of paint mediums were estimated<br />

only using biological stains on the cross sections. Additional paintings were<br />

exanLined visually to study the h<strong>and</strong>ling of paint.<br />

6. A survey of several Dutch l<strong>and</strong>scapes from the 1620s <strong>and</strong> the "tonal" period can<br />

be found in Bomford, D. Te chniques of the early Dutch l<strong>and</strong>scape painters. In<br />

Brown, op. cit., 45-56.<br />

7. This painting was sampled by Feliciry Campbell during conservation treatment<br />

in 1988.<br />

8. This painting closely replicates the primary version now at the J. Paul Getry<br />

Museum, Malibu (92.PB82).<br />

9. From the samples, it is not clear whether the underdrawing lies under or over<br />

the imprimatura.<br />

10. Coninxioo died in 1607. Savery left Amsterdam for Prague, where he served the<br />

Emperor Rudolf II until about 1613, when he returned to Amsterdam. In 1619<br />

he moved to Utrecht, where he worked until his death in 1639.<br />

11. Historisches Museum der Pfalz, Speyer (H. M.1957/122). Illustrated in Sutton,<br />

p. C. 1987. Masters of Seventeenth-Century Dutch L<strong>and</strong>scape <strong>Painting</strong>. Exhibition<br />

catalogue. Amsterdam: RijksmuseuIT1; Boston: Museum of Fine Arts; Philadelphia<br />

Museum of Fine Arts. Plate 1.<br />

12. A ground of black, toned with a little yellow earth <strong>and</strong> rubbed into the poplar<br />

support, was observed on the original section of Farms Flanking a Frozen Canal,<br />

1614, North Carolina Museum of Art, Raleigh (52.9.61). See Goist, D. C. 1990.<br />

Case study of an early Dutch l<strong>and</strong>scape. lCOM Committee fo r Conservation preprints,<br />

9th Triennial Meeting, 648-52.<br />

13. Gifford, E. M. 1983. A technical investigation of some Dutch seventeenth-century<br />

tonal l<strong>and</strong>scapes. AlC preprints, 39-49.<br />

14. Montias,J. M. 1987. Cost <strong>and</strong> value in seventeenth-century Dutch art. Art History<br />

10:455-66.<br />

Gifford 147


Abstract<br />

Through extensive study of Vermeer's<br />

paintings, the author demonstrates<br />

that the artist must have used<br />

a chalk line attached to a pin at the<br />

vanishing point in the painting to<br />

create the central perspective in his<br />

pictures. By studying the changes in<br />

the design of the central perspective<br />

throughout his oeuvre, a certain<br />

chronology appears. This conclusion<br />

contradicts the previously accepted<br />

beliefs that Vermeer's interiors were<br />

faithful portraits of actual rooms or<br />

that the use of a camera obscura explained<br />

the realism of his interiors.<br />

Johannes Vermeer (1632-1675) <strong>and</strong> His Use<br />

of Perspective<br />

J orgen Wadum<br />

Mauritshuis<br />

Korte Vijverberg 8<br />

p. O. Box 536, 2501 eM Den Haag<br />

The Netherl<strong>and</strong>s<br />

Introduction<br />

After visiting Vermeer on 21 June 1669, the art collector Pieter Teding van<br />

Berckhout noted in his diary that, among the examples of Vermeer's art he<br />

had seen, the most extraordinary <strong>and</strong> curious were those showing perspective<br />

(1). Three-dimensional interiors, depicted on two-dimensional canvases in<br />

such a way that the eye is deceived into believing the spatial illusion, were<br />

greatly admired by litjhebbers (connoisseurs) in the seventeenth century. The<br />

appreciation of perspective was underlined by the fact that these paintings<br />

had to be executed by artists who were sufficiently technically competent to<br />

be able to create these effects convincingly (2). Architectural pictures or "perspectives"<br />

were therefore often much more expensive than other genres.<br />

Montias documented that around 1650, the price fo r a perspective was fairly<br />

high, at an average of25.9 guilders apiece, while l<strong>and</strong>scapes sold for an average<br />

of 5.6 guilders each (3). The Delft architecture painter Hendrik van Vliet (ca.<br />

161 1-1675) could have observed that one of his perspectives, in the estate of<br />

the art dealer Johannes de Renialme in Amsterdam at his death in 1657, was<br />

valued at 190 guilders (4).<br />

Montias states that despite the fact that interior scenes could also rightfully<br />

be called perspectives, he never came across a description of one in the many<br />

inventories he examined (5). All the more interesting then is the comment<br />

van Berckhout made after visiting Vermeer's atelier.<br />

When the inventory of Vermeer's estate was carried out after his premature<br />

death in 1675, a number of books in folio were found in his back room<br />

together with twenty-five other books of various kinds. Among the easels<br />

<strong>and</strong> canvases in his atelier, three bundles of all sorts of prints were found (6).<br />

It might be interesting to speculate what these books <strong>and</strong> prints were about,<br />

but we shall never know. It is conceivable, however, that some of them were<br />

guides to perspective drawing, works either by Hans Vredeman de Vries<br />

(1526 or 1527-1606) or those published by S. Marolois (ca. 1572-1627),<br />

Hendrik Hondius (1573-1649), <strong>and</strong> F. Desargues (1593-1662) (7, 8, 9, 10). It<br />

can be seen in the perspective design extrapolated from his paintings that<br />

Vermeer was certainly familiar with the principles laid out in these manuals<br />

on perspective.<br />

Unfortunately, nothing is known about Vermeer's apprenticeship. Therefore<br />

one must turn to an extensive examination of his paintings in order to gain<br />

an impression of the development of his method of rendering space.<br />

State of research<br />

Over the years many studies have been made of Vermeer's use of perspective<br />

<strong>and</strong> his spatial constructions, only a few of which shall be referred to here.<br />

Early this century, Eisler made an extensive study of Vermeer's use of space,<br />

<strong>and</strong> he describes the complicated use of triangles, circles, squares, <strong>and</strong> diagonals<br />

that may have fo rmed the basis fo r Vermeer's pictures (11).<br />

Probably inspired by Wilenski, who in 1928 wrote about some special effects<br />

in photography, " . .. perhaps one of the ironies of art history [is] that with<br />

a Kodak any child might now produce by accident a composition that a great<br />

148<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


artist like Vermeer had to use all his ingenuity ... to achieve ... ," Swill ens<br />

published his first detailed study of the paintings in 1929. Swillens concluded<br />

that Vermeer had painted his oeuvre in five different rooms, all thoroughly<br />

recorded (12, 13). In another publication on Vermeer in 1950, he further<br />

elaborated his views on Vermeer's use of spatial illusion <strong>and</strong> his realistic recording<br />

of space (14). Swillens illustrated how Vermeer depicted his interiors<br />

with great accuracy. The position or eye level of the artist was established<br />

<strong>and</strong> thus the height of Vermeer himself <strong>and</strong> of the chair on which he almost<br />

always sat when painting. The work of Swillens, <strong>and</strong> belief that Vermeer<br />

rendered what he actually saw in front of him, has had a major influence on<br />

the scholarly research on Vermeer.<br />

Even in the 1940s, Hyatt Mayor records that highlights in the foreground in<br />

some of Vermeer's paintings "break up into dots like globules of halation<br />

swimming on a ground glass," <strong>and</strong> a few years later Gowing reached the<br />

conclusion that Vermeer had used a camera obscura (15, 16).<br />

In his studies, Seymour continues with this thought, which, apart from halation<br />

around highlights, he based on specific phenomena in the paintings,<br />

such as the diffusion of the contours (17). He also found that the perspective<br />

in certain paintings resembled the distortion obtained when using a wideangle<br />

lens.<br />

Prompted by Seymour's article, Schwarz fu rther suggested that Vermeer may<br />

have used the camera obscura as a technical aid in his painting process (18).<br />

Bearing testimony to Vermeer's use of technical devices for rendering his<br />

images, wrote Schwarz, is the fact that the mathematician <strong>and</strong> physicist Balthasar<br />

de Monconys (1611-1665) attempted to visit Vermeer during his stay<br />

in Delft in 1663, <strong>and</strong> that a friendship probably existed between Vermeer <strong>and</strong><br />

fellow townsman Anthony van Leeuwenhoek (1632-1723), a specialist in<br />

microscopes <strong>and</strong> lenses.<br />

There is a tendency to consider mathematicians otherwise uninvolved in the<br />

creation of visual arts as responsible for developing an intellectual interest in<br />

perspective. However, this view was probably not shared by the artists, who,<br />

fo r their part, were using the simplest <strong>and</strong> at the same time most convincing<br />

methods to create their spatial illusions. We are therefore entitled to believe<br />

that, as de Monconys was also an art collector, it would be much more likely<br />

that he wanted to pay Vermeer a visit in order to see his renowned paintings<br />

or perspectives (19). Vermeer was also the Headman of the Guild of St. Luke<br />

at the time, <strong>and</strong> he would naturally be the person for an art collector to see<br />

when visiting the town.<br />

In 1968 Mocquot suggested that Vermeer might have used double mirrors<br />

to create his perspectives, both in his Allegory of <strong>Painting</strong> <strong>and</strong> in The Concert<br />

(20, 21, 22).<br />

Finck claimed in 1971 to be able to prove that twenty-seven paintings by<br />

Vermeer must have been made with the aid of a camera obscura (23). The<br />

arguments presented here will make it clear that this highly ambitious thesis<br />

has no basis in reality.<br />

Wheelock undertook the most detailed study of the optics <strong>and</strong> perspectives<br />

used by Delft painters around 1650 (24). It is argued that some of Vermeer's<br />

pictures (although far fewer than is asserted by Finck) do indeed have many<br />

effects similar to that which can be achieved using lenses or a camera obscura,<br />

<strong>and</strong> therefore the use of these devices seems highly probable. Wheelock does<br />

make clear, however, that the use of a camera obscura would be extremely<br />

difficult indoors because the light levels were generally insufficient to obtain<br />

an image. In more recent publications, Wheelock increasingly argues that<br />

Vermeer probably did not trace images seen through a camera obscura, but<br />

that he must have been aware of the device <strong>and</strong> used certain special effects<br />

seen through it for his paintings.<br />

Wadum 149


Based on intensive studies of various means used to create " church portraits,"<br />

de Boer concluded in 1988 that optical devices were not generally used by<br />

artists in the Netherl<strong>and</strong>s around 1650 (25). Interestingly, he notes that the<br />

reason for this would probably be the difficulty of combining the use of a<br />

camera obscura (for tracing an image) with actually painting a painting.<br />

Recently, Arasse made a general comparison of the position of the horizon<br />

<strong>and</strong> the viewpoint in Vermeer paintings (26). He notes that the viewpoint<br />

gradually lowered between 1656 <strong>and</strong> 1661. According to Arasse, Vermeer<br />

tended to combine a low viewpoint with a high horizon. Arasse considers<br />

the often very low viewpoint in relation to the depicted figures to be a special<br />

effect that Vermeer deliberately wanted to create in order to draw the viewer<br />

into the scenes. This statement shows that Arasse considers Vermeer's intention<br />

to be the creation of an illusionistic spatial setting as an imaginative<br />

process rather than the rendering of a known space, an opinion this author<br />

shares.<br />

Present research<br />

In the following paragraphs, results from the author's latest research on this<br />

aspect of Vermeer's painting technique are presented. Through a thorough<br />

study of the actual paintings, mostly out of their frames <strong>and</strong> placed under a<br />

stereomicroscope, certain surface phenomena in the paint layer have been<br />

observed. Together with X-radiographs <strong>and</strong> other photoanalytical means such<br />

as ultraviolet <strong>and</strong> infrared photography, a compilation of information has been<br />

possible, leading to the conclusion that Vermeer did not paint "naar het leven"<br />

(after life), as suggested by the majority of scholars mentioned above, but that<br />

as a craftsman he created a spatial illusion with the masterly h<strong>and</strong> of an<br />

outst<strong>and</strong>ing artist.<br />

In 1949 Hulten was the first to actually record a discernible vanishing point<br />

in one of Vermeer's paintings. He observed that just below the left knob of<br />

the map hanging on the rear wall in The Allegory of <strong>Painting</strong> there was a small<br />

irregularity in the paint layer which coincided with the central vanishing<br />

point of the composition (27).<br />

Indeed, fo r Vermeer the central perspective was the main guideline for his<br />

interiors. Current examinations reveal that the vanishing point can still be<br />

fo und in most of his interior scenes (28). It can be seen (with the naked eye<br />

or more easily with a stereomicroscope) that Vermeer must have attached a<br />

pin at the vanishing point in the painting, resulting in the loss of minuscule<br />

amounts of paint <strong>and</strong> ground. X-radiographs can be used to find the black<br />

spot where the ground containing lead white is missing between the threads<br />

of the canvas. Having inserted the pin at the vanishing point, Vermeer would<br />

have used a string to reach any area of his canvas to create perfect orthogonals<br />

for the perspective.<br />

Vermeer's method is far from unique; among the architectural painters of his<br />

time it was well known. Gererd Houckgeest (1600-1661) <strong>and</strong> Emanuel de<br />

Witte (1617-1 692) practiced this method, which Pieter Saenredam (1597-<br />

1665) had already brought to perfection (29, 30). Vermeer's slightly older<br />

colleague Pieter de Hooch (1629-1683) also used a single vanishing point.<br />

Similarly, in paintings by Gerard Dou (1613-1675), Gabriel Metsu (ca. 1629-<br />

1667), <strong>and</strong> others, we again find irregularities in the paint where a pin was<br />

placed at the vanishing point.<br />

The method of using a chalk line to indicate lines is still used by painters<br />

<strong>and</strong> other artists when planning illusionistic interiors (e.g., with marbling, a<br />

specialty developed during the Baroque period) . That this kind of illusionistic<br />

painting was known to Vermeer is clear from the virginals in the two London<br />

paintings, both of which have been "marbled."<br />

In order to transfer the line indicated by the string, chalk is applied to the<br />

string. Holding the string taut from the pin inserted at the vanishing point,<br />

150<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 1. Perspective drawing of Vermeer's Glass of Wine, 1658-1660.<br />

the painter draws the string up a little from the surface, using the free h<strong>and</strong>,<br />

<strong>and</strong> lets it snap back onto the canvas. The powdery chalk is thus applied to<br />

the surface of the painting; the line can be traced with a pencil or brush. The<br />

remaining dust can be gently wiped or blown away, leaving little or no trace<br />

of the method except the pin point.<br />

The distance points, positioned on either side of the vanishing point on the<br />

horizon, provide the basis fo r the diagonals, which in turn form the basis of<br />

the tiled floors. Distortion occurs at the corners if the horizon is placed<br />

relatively high <strong>and</strong> the distance points are close to the vanishing point. Examples<br />

of this are observed in the Glass of Wine (1658-1660), with its viewing<br />

angle of approximately 43° (Fig. 1), <strong>and</strong> the Girl with The Wine Glass (1659-<br />

1660), which has an even smaller angle of approximately 35° (31, 32). The<br />

Music Lesson (1662-1665), in which Vermeer again returns to a large angle,<br />

(approximately 44°), is the last picture in his oeuvre to show a certain distortion<br />

of the floor tiles due to the short interval from the distance points<br />

(33). In The Concert (1665-1666), the angle again returns to about 34°; in The<br />

Allegory of <strong>Painting</strong> (1666-1667), the viewing angle has decreased to around<br />

30° (Fig. 2). In The Love Letter (1667-1670) the angle declines to about 28°,<br />

<strong>and</strong> in the last painting executed by his h<strong>and</strong>, A Lady Seated at the Virginals<br />

(1673-1675), Vermeer reduces the viewing angle to only 22° (Fig. 3) (34,<br />

35).<br />

A growing tendency can clearly be observed over the years to let the distance<br />

points move further away from the scene. By doing this, Vermeer eliminated<br />

the distortion of the floor tiles in the foreground corners, particularly as he<br />

moved his vanishing point toward the edge of the painting at the same time.<br />

As the vanishing point can still be identified in many of Vermeer's pictures,<br />

his method of using threads attached to a pin inserted at the central point is<br />

evident. The distance points, however, could constitute a problem. Would he<br />

be able to determine the position of the diagonals on the edge of his canvas<br />

when space recedes towards the back walls in his interiors? This is hardly<br />

Figure 2. Perspective drawing of Vermeer's Allegory of <strong>Painting</strong>, 1666-1 667.<br />

Wadum 151


Figure 3. Perspective drawing of Vermeer's A Lady Seated at the Virginals, 1673- 1675.<br />

likely because it would imply doing unnecessary calculations, <strong>and</strong> indeed no<br />

trace of marks on the edges of his paintings has so far surfaced.<br />

If there were a simple method of creating perfect central-point perspective,<br />

painters would surely have used it. By placing the canvas against a board (most<br />

of his paintings are small) or a wall, between two nails on either side of the<br />

painting, the painter would be able to use strings for the diagonals as well.<br />

Indications of the use of such a simple method may be deduced from books<br />

on perspective that might have been known to Vermeer. Desargues writes in<br />

his introduction that a painter who wants to know more about the Meetkonst<br />

(the art of measurement) should consult the L<strong>and</strong>meeter (the cartographer)<br />

in order to make use of his expertise (36) . This, he writes, would lead<br />

to a better underst<strong>and</strong>ing of perspective or Doorsicht-kunde. Desargues further<br />

suggests that the painter should look around him in other guilds to take<br />

advantage of the knowledge of carpenters, bricklayers, <strong>and</strong> cabinetmakers.<br />

Furthermore, it appears that constructors of perspective in the seventeenth<br />

century were using drawing tables almost as sophisticated as the ones we use<br />

today. With strings attached to the upper corners of the drawing table, the<br />

draftsman could create any orthogonals he wanted on his paper. The horizon<br />

could be plotted using a sliding ruler at a fixed 90° angle to the horizontal<br />

bottom edge of the table. A horizon would be chosen at the desired level on<br />

this ruler, <strong>and</strong> by sliding the ruler across the paper, a line could be drawn<br />

(37).<br />

Vermeer also worked in this way, as is proven by the presence of the clearly<br />

distinguishable needle point found in the paint in paintings throughout his<br />

whole oeuvre (38).<br />

Conclusion<br />

The extraordinary <strong>and</strong> curious perspectives, so much admired by van Berckhout<br />

in 1669, therefore appear to have been carefully constructed. This leaves<br />

the impression that Vermeer should be regarded first <strong>and</strong> foremost as a practical<br />

<strong>and</strong> skilled master in creating space just the way he wanted it. This<br />

approach departs from the previous conception of the artist as reproducing<br />

the scenes he saw in front of him, either by careful copying using drawing<br />

frames or a camera obscura. The author believes that Vermeer was completely<br />

aware of the spatial illusion he wanted to create, which he produced by<br />

combining his skill in constructing space with his artistic talent fo r composition,<br />

color, technique, <strong>and</strong> iconography (39). He thereby created his images<br />

in such a way that viewers are deceived into believing that the scenes were<br />

real. This was the highest level of artistic ambition to which a seventeenthcentury<br />

painter could aspire, a level Vermeer surely attained.<br />

Acknowledgments<br />

Most sincere thanks are due to the institutes that have been more than generous in<br />

supplying information about their Vermeer paintings <strong>and</strong> letting the author examine<br />

most of them in the conservation studios. The amount of information is overwhelming<br />

<strong>and</strong> goes well beyond the scope of this article; the reader is referred to the<br />

fo rthcoming exhibition catalogue on Johannes Vermeer (Washington, 1995; The<br />

152<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Hague, 1996). The author is also grateful to Rob Ruurs of the University of Amsterdam<br />

for a fruitful <strong>and</strong> stimulating discussion.<br />

Notes<br />

1. Van Berckhout may have been able to see the two town views, The Little Street<br />

<strong>and</strong> The View of Delft, <strong>and</strong> certainly interior scenes such as The Music Lesson, A<br />

Woman Holding a Balance, The Concert, <strong>and</strong> The Allegory of <strong>Painting</strong>. He might<br />

have also seen The Love Letter, which is dated 1667 (Blankert) or 1669-1670<br />

(Wheelock).<br />

2. Ruurs, R. 1991. Functies van architectuurschildering, in het bijzonder het kerkinterieur.<br />

Perspectieven: Saenredam en de architectuurschilders van de 17e eeuw: Boymans-van<br />

Beuningen Rotterdam, 50.<br />

3. Montias, J. M. 1991. "Perspectieven" in zeventiende-eeuwse boedelbeschrijvingen.<br />

Perspectiven: Saenredam en de architectuurschilders van de 17e eeuw: Boymansvan<br />

Beuningen Rotterdam, 28.<br />

4. Bredius, A. 1915-1922. Kunstler-Inventare, Urkunden zur Cesichte der Holl<strong>and</strong>ischen<br />

Kunst des XVIten, XVIIten und XVIIItenjahrhunderts 1-7. Den Haag.Vol. 1:238.<br />

5. Montias, op. cit., 24.<br />

6. Montias, J. M. 1989. Vermeer <strong>and</strong> his Milieu: A Web of Social History. Princeton,<br />

339 (doc. 364).<br />

7. Vredeman de Vries, H. 1604. Perspective. Dat is de hoochgheroemde cOl'lste . . . . Den<br />

Haag.<br />

8. Marolois, S. 1628. Perspective contenant la tMorie, pratique et instructionfondamentale<br />

d'icelle. Amsterdam.<br />

9. Hondius, H. 1647. Crondige onderrichten in de optica of de perspective konste. Amsterdam.<br />

10. Bosse, A. 1664. Franfois Desargues. Algemene manier tot de practyck der perspective<br />

gelyck tot die der meet-kunde met de kleyn voet-maat [ ... J by-een-gevoeght door. Amsterdam.<br />

11. Eisler, M. 1916. Der Raum bei Jan Vermeer. jahrbuch der Kunsthistorischen Sammlungen<br />

der Allerhachsten Kaiserhauses. Wi en-Leipzig: B<strong>and</strong> XXXIII (4):213-91.<br />

12. Wilenski, R. H. 1928. An introduction to Dutch Art. New York, 284-86.<br />

13. Swillens, P. T. A. 1929. Een perspectivische studie over de schilderijen van Johannes<br />

Vermeer van Delft. Oude Kunst (7): 129-61.<br />

14. Swillens, P. T. A. 1950.joharmes Vermeer, Painter of Delft 1632-1675. Utrecht­<br />

Brussels.<br />

15. Hyatt Mayor, A. 1946. The photographic eye. The Metropolitan Museum of Art<br />

Bulletin (5):15-26.<br />

16. Gowing, L. 1953. Vermeer.<br />

17. Seymour Jr., C. 1964. Dark chamber <strong>and</strong> light-filled room: Vermeer <strong>and</strong> the<br />

camera obscura. The Art Bulletin XLVI (3):323-31.<br />

18. Schwarz, H. 1966. Vermeer <strong>and</strong> the camera obscura. Pantheon (XXIV): 170-82.<br />

19. De Monconys visited numerous artists during his travels all over Europe to meet<br />

with fellow scientists.<br />

20. Mocquot, M. 1968. Vermeer et Ie portrait en double mirror. Le club franfaise de<br />

la medaille 18 (1):58-69.<br />

21. Kunsthistorisches Museum, Vienna.<br />

22. Isabella Steward Gardner Museum, Boston.<br />

23. Finck, D. 1971. Vermeer's use of the camera obscura: a comparative study. The<br />

Art Bulletin. (LIII):493-505.<br />

24. Wheelock Jr., A. K. 1977. Perspective, Optics, <strong>and</strong> Delft Artists Around 1650. New<br />

York.<br />

25. de Boer, P. G. 1988. Enkele Delftse zeventiende eeuwse kerkportretten opnieuw bekekert.<br />

Dissertatie: Delft.<br />

26. Arasse, D. 1994. The Art of <strong>Painting</strong>. Paris.<br />

27. Hulten, K. G. 1949. Zu Vermeer's Atelierbild. Konsthistorisk Tidskrift (XVIII):90-<br />

98.<br />

28. The first pin hole we discovered in the vanishing point of a Vermeer painting<br />

was in the Lady Writing a Letter with her Maid from the National Gallery of<br />

Irel<strong>and</strong>. We are grateful to Andrew O'Connor, chief conservator of the gallery,<br />

fo r letting us examine the painting just after its recent rediscovery.<br />

29. Ruurs, R. 1987. Saertredam: The Art of Perspective. Amsterdam, Philadelphia, Gronmgen.<br />

30. Giltaij, J. 1991. Perspectiven, Saenredam en de architectuurschilders van de 17e<br />

eeuw. Perspectiven: Saemedam en de architectuurschilders van de 17e eeuw: Boymansvan<br />

Beuningen Rotterdam, 16.<br />

Wadum 153


31. Herzog Anton Ulrick-Museum, Braunschweig.<br />

32. Staatliche Museen zu Berlin, Preul3ischer Kulturbesitz, Gemaldegalerie, Berlin-<br />

Dahlem.<br />

33. H. M. Queen Elizabeth II, Royal Collection, London.<br />

34. Rijksmuseum, Amsterdam.<br />

35. The National Gallery, London.<br />

36. Bosse, op. cit.<br />

37. Ibid.<br />

38. From Offi cer <strong>and</strong> Laughing Girl, ca. 1658 (The Frick Collection, New York) , to<br />

A Young Woman St<strong>and</strong>ing at the Virginal, ca. 1673-1675 (The National Gallery,<br />

London).<br />

39. See fo rthcoming exhibition catalogue on Johannes Vermeer (Washington, 1995;<br />

The Hague, 1996), in which various aspects of Vermeer's paintings are examined.<br />

154<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

The paper presents some results of<br />

research on the painting materials<br />

<strong>and</strong> methods used in Latvian<br />

churches in the seventeenth century.<br />

The technical research of historical<br />

painting techniques in Latvia is rather<br />

preliminary. The authors concentrate<br />

on the polychromy of the interior<br />

decorations. Results of the<br />

analysis of materials <strong>and</strong> techniques<br />

used in the decorations are given,<br />

stressing the varieties as a result of<br />

the various interpretations of the Baroque<br />

style <strong>and</strong> technique in different<br />

regions of Latvia.<br />

A Technical Study of the <strong>Materials</strong> <strong>and</strong> Methods<br />

Used by the Painters of the Latvian Churches<br />

in the Seventeenth Century<br />

Ilze Poriete*<br />

Restoration Centre History<br />

Museum of Latvia<br />

Vecpilsetas 7<br />

Riga, LV 1050<br />

Latvia<br />

Dace Choldere<br />

State Culture Monuments Department of Latvia<br />

Klostera 5<br />

Riga, LV 1050<br />

Latvia<br />

Introduction<br />

In the second half of the seventeenth century, the territory of Latvia was<br />

ruled by several different monarchs <strong>and</strong> was therefore divided into areas of<br />

various influences. From the middle of the sixteenth century, the western part<br />

was ruled by the Duke ofKurzeme an Zemgale; after the Polish-Swedish war<br />

the southeastern part (Latgale) was subject to the Polish king <strong>and</strong> the northeastern<br />

part (Vidzeme) to the Swedish king. These conditions influenced the<br />

production of Latvian art. Created by artists coming from various parts of<br />

Europe, general styles <strong>and</strong> tendencies were interpreted in various ways (1).<br />

Therefore, the objects of this survey-the churches <strong>and</strong> their pulpits in Nurmuizha<br />

<strong>and</strong> Burtnieki, both decorated in the 1680s <strong>and</strong> representative of the<br />

Baroque-illustrate two different styles using very different techniques <strong>and</strong><br />

materials.<br />

Figure 1. The pulpit <strong>and</strong> the altar of the<br />

Nurmuizha church bifore restoration. Photograph<br />

by M. B. Vanaga <strong>and</strong> R. Kanins.<br />

Figure 2. The pulpit of the Burtnieki church<br />

bifore restoration. Photograph by M. B.<br />

Vanaga <strong>and</strong> R. Kanins.<br />

The church of Nurmuizha is located in the territory of Kurzeme <strong>and</strong> was<br />

founded in 1594; however, the present interior dates from the 1680s. Twisted<br />

columns with rich decorative carvings <strong>and</strong> many sculptures portraying disciples<br />

of Christ decorate the pulpit <strong>and</strong> the church interior. A sculpture of<br />

Moses functions as support for the pulpit. The artist who executed the wood<br />

carving is unknown but his style indicates an eastern Prussian origin. Although<br />

written contemporary sources do not shed any light on when the<br />

decorations were painted, one must assume that it was carried out in the<br />

1680s. Later repairs are relatively extensive, including additional wood carvings<br />

<strong>and</strong> paintings executed by different craftsmen, such as gilders <strong>and</strong> interior<br />

painters (Fig. 1).<br />

The church of Burtnieki is located in the northeastern part of Latvia which<br />

was under the rule of the Swedish king. Here, the basic construction of the<br />

pulpit is very similar to construction of the pulpit in Nurmuizha, but the<br />

decoration is more restricted, rational, <strong>and</strong> even classical. Baroque influences<br />

arrived here by many different routes, thus resulting in a large variety of<br />

interpretations. In both churches, pulpits are situated under the arch of triumph.<br />

In Burtnieki, however, the pulpit's construction is polygonal with<br />

straight stairs, <strong>and</strong> the columns are not twisted but straight <strong>and</strong> narrow. The<br />

sides of the stairs are decorated with panels separated by pilasters instead of<br />

columns. Instead of sculptures, there are paintings on the panels between the<br />

columns, all representing scenes from the New Testament. In contemporary<br />

documents (1691), it is noted that the altar <strong>and</strong> the pulpit were painted <strong>and</strong><br />

outlined with silver. During the present restoration, a rather neutral overpainting<br />

was removed, revealing the original decorative painting, which shows<br />

* Author to whom correspondence should be addressed.<br />

Poriete <strong>and</strong> Choldere 155


marbled columns, multicolored paintings on the panels, <strong>and</strong> blue <strong>and</strong> silver<br />

paintings on the podium. The identities of the artists are unknown (Fig. 2).<br />

Methods<br />

Analysis of the pigments was carried out combining optical microscopy, microchemical<br />

tests, <strong>and</strong> emission spectroanalysis (2, 3). Media were determined<br />

by the use of thin-layer chromatography, infrared spectroscopy, <strong>and</strong> microchemical<br />

tests (4, 5, 6) .<br />

<strong>Materials</strong> <strong>and</strong> results<br />

Technical analysis from paint samples was executed to reveal the original layer<br />

of the pulpit of the Nurmuizha church. Samples were taken from the background<br />

<strong>and</strong> profiles of the pulpit's twisted columns (now black) <strong>and</strong> from the<br />

decorative wood carvings of grapes, masks, <strong>and</strong> reliefs (now gold) . The results<br />

indicate that the ground layer consists of an unpigmented calcium carbonate<br />

bound with an animal glue, as was shown by staining tests. In the cross section,<br />

the ground layer is shown as a white layer with some small brown glue<br />

particles. On a thin transparent layer in which protein has been fo und, there<br />

is a black layer with occasional particles of a blue pigment. The few blue<br />

particles present are transparent; a positive identification of the pigment was<br />

impossible. Later overpaintings are executed in black, containing charcoal <strong>and</strong><br />

oil. The decorative vines are gilded. The underlayer is composed of calcium<br />

carbonate <strong>and</strong> glue. In the cross section, a layer of brown hematite is visible<br />

<strong>and</strong> the presence of glue particles was determined with staining tests. On top<br />

of the gilding there is a layer of mordant gilding: oil pigmented with ochre<br />

<strong>and</strong> minium, the latter probably acting as a drying agent. The bronze layer<br />

on the pigmented oil layer was applied much later.<br />

Figure 3. Cross section if paint sample from<br />

part if the flesh color on a sculpture in the<br />

Nurmuizha church. The layers are as follows:<br />

(1) thin pink ground, (2) main paint layer<br />

containing lead white <strong>and</strong> ochre, (3) transparent<br />

layer containing oil, (4-8) overpaints.<br />

Photograph by M. B. Vanaga <strong>and</strong> R. Kanins.<br />

Samples were taken from the clothes, hair, <strong>and</strong> flesh color of one sculpture<br />

on the pulpit (Fig. 3) . The analyses showed that the gilded wrap was executed<br />

in a water-gilding technique, while the dress itself was originally blue. The<br />

blue layer consists oflead white <strong>and</strong> smalt with tempera as a binding medium<br />

on a chalk-glue ground layer (7). Analysis of the samples taken suggests that<br />

the pulpit had water-gilded, wood-carving details on a blue background. The<br />

same blue color is also fo und in other details of the church's decoration (the<br />

altar <strong>and</strong> the decorative ledge). Later overpaintings, however, have penetrated<br />

the original, damaged layer, changing it <strong>and</strong> making a correct analysis difficult.<br />

During the technical investigation of the pulpit of Burtnieki church, samples<br />

were taken from parts that were well conserved: the background of the pulpit's<br />

construction, the podium's decorative ledge, the decorative grapes, the roof,<br />

<strong>and</strong> the pelican. The results of the analysis show that the pink background<br />

contains hematite <strong>and</strong> occasional particles of calcium carbonate (CaC03) in<br />

oil (Fig. 4). Colors from the original marbling are revealed in the background<br />

(smalt, indigo, hematite, charcoal black, <strong>and</strong> copper resinate). Decorative elements<br />

are silvered, such as the silver leaves, which were glazed with copper<br />

resinate over the pink underlayer (hematite, oil).<br />

Figure 4. Cross section if paint sample from<br />

part if the podium in the Burtnieki church.<br />

The layers are as follows: (1) thin pink<br />

ground, (2) layer of silver leaves, (3) green<br />

layer containing cupric resinate, (4-8) overpaints.<br />

Photograph by M. B. Vanaga <strong>and</strong><br />

R. Kanins.<br />

The decorative bunch of grapes is painted in a dark blue layer containing<br />

indigo, smalt, <strong>and</strong> oil, <strong>and</strong> the green leaves are executed in a green glaze<br />

(copper resinate, oil, gum) over silver leaf applied in an oil-based mordant<br />

technique. The decorative details of the pulpit's baldachin have the same<br />

preparatory ground layer as the podium. The marbling is carried out as described<br />

above. The silvering is done using an oil mordant <strong>and</strong> subsequently<br />

glazed. In some places the silver is covered with colorful glazes (i.e., the drops<br />

of blood on the pelican are executed in an organic red glaze on silver). In<br />

summary, the decorative painting of the Burtnieki church shows a rich polychromy<br />

<strong>and</strong> is executed in an oil medium.<br />

156<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Summary<br />

We can conclude after this preliminary investigation that many differences<br />

between the two pulpits are present, although they both date from the same<br />

period. Two types of grounds were found; in the Nurmuizha pulpit, chalkglue<br />

ground is used, while in the Burtnieki pulpit a ground of ochre in oil<br />

is present. The pigments are similar in that they are probably the most commonly<br />

used pigments in Latvia during that period. During previous investigations<br />

of blue colors in the polychromy of seventeenth-century Latvian<br />

churches, two pigments were always present: indigo <strong>and</strong> smalt. Natural ultramarine<br />

has been found in very few cases.<br />

Details in the decoration of the Nurmuizha church are executed using a<br />

water-gilding technique. It is impossible to determine if a varnish or glaze is<br />

present, as the next gilding layer is applied in an oil-based mordant technique<br />

that fully blends with the first gilded layer.<br />

The decorative wood carvings of the Burtnieki church are silvered <strong>and</strong> in<br />

some places covered with colored glazes.<br />

Different h<strong>and</strong>s are present in both churches, as during the seventeenth century<br />

in Latvia there was a continuous migration of craftspersons. A gilderpainter<br />

could have been called from any workshop, especially in rural areas,<br />

introducing that workshop's typical techniques <strong>and</strong> materials. It is not suggested<br />

that the described methods were the most common in Latvia. Any<br />

concrete conclusions can only be made after fu rther investigations in Latvia,<br />

<strong>and</strong> after a comparison with methods used in other parts in Europe.<br />

Acknowledgments<br />

The authors are grateful to polychrome wooden sculpture restorers S. Rugevica <strong>and</strong><br />

1. Pukse fo r the samples <strong>and</strong> for their help with the work. Thanks are due to M.<br />

Harris for editing the manuscript.<br />

Notes<br />

1. Grosmane, E. 1981. Ventspils koktelnieki 17.gs otra puse- 18.gs.sakums. Riga, Zinatne,<br />

46-61.<br />

2. Microscope MBI-15 (USSR), magnification x160.<br />

3. Emission spectroscopy was carried out by A. Deme, an assistant of the spectroscopy<br />

laboratory of Latvia University.<br />

4. Striegel, M. E, <strong>and</strong> D. Stulik. 1993. High performance thin-layer chromatography<br />

fo r the identification of binding media: techniques <strong>and</strong> applications. In AIC Preprints.<br />

5. Infrared spectroscopy was carried out by 1. Gudele, an assistant at Riga Technical<br />

University.<br />

6. Martin, E. 1977. Some improvements in techniques of analysis of paint media.<br />

Studies in Conservation 22:63-67.<br />

7. The definition is in process.<br />

Poriete <strong>and</strong> Choldere 157


Abstract<br />

The shift in the nineteenth century<br />

from a tradition of artist-prepared<br />

materials to an industry of mass-produced<br />

commercial products greatly<br />

endangered the artistic community<br />

through the widespread distribution<br />

of products of inferior quality <strong>and</strong><br />

unstable properties. For over fifty<br />

years, the British Pre-Raphaelite<br />

painter William Holman Hunt<br />

waged a campaign fo r the reform of<br />

the manufacture of artists' materials<br />

<strong>and</strong> the rights of the artist as a consumer<br />

to expect materials of consistent<br />

quality <strong>and</strong> uniformity. The reevaluation<br />

of the Pre-Raphaelite<br />

technique, in conjunction with an<br />

exploration of Hunt's advocacy on<br />

behalf of artist-consumers, places in<br />

perspective his focus on artistic<br />

traditions at a crucial transition time<br />

in the history of materials <strong>and</strong> tech­<br />

TUques.<br />

William Holman Hunt <strong>and</strong> the<br />

"Pre-Raphaelite Technique"<br />

Melissa R. Katz<br />

Davis Museum <strong>and</strong> Cultural Center<br />

Wellesley ColJege<br />

Wellesley, Massachusetts 02181<br />

USA<br />

Introduction<br />

The late-eighteenth <strong>and</strong> early nineteenth centuries in British painting constitute<br />

a period noted for the rise of a major school of national painting,<br />

dominated by masters such as Sir Joshua Reynolds, Thomas Gainsborough,<br />

<strong>and</strong> J. M. W. Turner, who engaged in technical experiments of dubious value.<br />

Their reliance on gelled mediums, bituminous paints, <strong>and</strong> fugitive pigments,<br />

respectively, has left a body of work disfigured by sunken patches, wide craquelure,<br />

<strong>and</strong> faded color. In search of shortcuts to achieve the luminous glow<br />

of the old masters, they produced, instead, paintings whose technical inadequacies<br />

were known <strong>and</strong> seen by the generation that followed.<br />

This next generation of painters included the Pre-Raphaelite Brotherhood,<br />

formed in 1848 in radical opposition to the training <strong>and</strong> taste imposed on<br />

British art by the Royal Academy. Hunt was a fo unding brother, <strong>and</strong> the only<br />

one of the group to maintain a lifelong adherence to their principles of<br />

fidelity to nature minutely observed, boldness in color <strong>and</strong> lighting, emulation<br />

of early Italian painting, <strong>and</strong> depiction of contemporary or literary subject<br />

matter. Yet the rebellion was short-lived, <strong>and</strong> the Pre-Raphaelites rapidly<br />

became the leading painters of their day, with Hunt as one of the most<br />

popular.<br />

By the 1870s Hunt's position was assured as the celebrated painter of such<br />

Victorian icons as The Light oj the World, The Awakening Conscience, <strong>and</strong> The<br />

Finding if the Saviour in the Temple (Figs. 1,2, 3). Periods spent in the Middle<br />

East seeking Biblical authenticity alternated with spells in pleasant, wellequipped<br />

London studios where, liberated from h<strong>and</strong>-to-mouth struggle, he<br />

was free to contemplate other aspects of art as a career (Figs. 4, 5). The stability<br />

<strong>and</strong> longevity of his paintings were of primary concern to Hunt, who observed<br />

not only the technical inadequacies of the preceding generation, but<br />

also the poor aging qualities of the artworks of his contemporaries.<br />

Nineteenth-century commercial artists' materials<br />

Hunt was among the first artists to note the increasingly poor quality of the<br />

artists' materials fo r sale in the mid-nineteenth century, <strong>and</strong> the most vociferous<br />

in calling attention to their faults <strong>and</strong> advocating their improvement.<br />

From the platform of leading painter of his day, he set out on a crusade to<br />

reform the manufacture of artists' materials, to impose st<strong>and</strong>ards of quality<br />

<strong>and</strong> workmanship, <strong>and</strong> to ensure access to consistently reliable products from<br />

colormen informed of <strong>and</strong> interested in the durability of the goods they were<br />

selling.<br />

Figure 1. William Holman Hunt in his studio,<br />

painting the St. Paul's versioll q{The<br />

Light of the World, ca. 1900.<br />

The market fo r artists' materials in the early <strong>and</strong> mid-nineteenth century had<br />

changed overwhelmingly with the advent of industrial production <strong>and</strong> mass<br />

marketing, <strong>and</strong> the rapid introduction <strong>and</strong> adaptation of newly developed<br />

materials whose aging properties were unknown. Prior to the nineteenth<br />

century, artists had used materials prepared fo r them in their own studios or<br />

by local artisans who followed the exacting st<strong>and</strong>ards of their clients, allowing<br />

the artist to determine the materials used <strong>and</strong> methods of preparation. With<br />

the rise of the commercial colorman in an age of burgeoning capitalism, a<br />

middleman was introduced between the manufacturer <strong>and</strong> the consumer of<br />

158<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


the product, a salesman motivated by the need for profit <strong>and</strong> immediate shortterm<br />

customer satisfaction rather than long-term stability.<br />

The evolution of the art nurket also brought with it an expansion of materials<br />

available fo r sale, aided by advances in scientific fields that greatly exp<strong>and</strong>ed<br />

the range of available pigments <strong>and</strong> dyestuffs. The artist's palette roughly doubled<br />

in the nineteenth century, each decade bringing new colors, starting<br />

with cobalt blue <strong>and</strong> lemon yellow. In the early part of the century, iodine<br />

scarlet, chrome yellow, <strong>and</strong> emerald green were introduced; <strong>and</strong>, in the 1820s,<br />

synthetic ultramarine became available. Zinc oxide (Chinese white) <strong>and</strong> viridian<br />

came in the 1830s, <strong>and</strong> cadmium yellow <strong>and</strong> orange followed in the<br />

1840s. The 1850s brought the aniline dyes, brilliantly colored <strong>and</strong> wildly<br />

impermanent coal-tar derivatives. The 1860s saw the dawn of aureolin yellow,<br />

chromium oxide green, synthetic alizarin, manganese violet, cerulean blue,<br />

<strong>and</strong> so on (1). The colors produced by these new pigments were often dazzling,<br />

but also alarmingly unstable.<br />

Figure 2. William Holman Hunt sculpting a<br />

model of Christ's head to aid him in the<br />

painting


with colormen <strong>and</strong> hired scientists to analyze paint samples from various<br />

suppliers. He badgered the Royal Academy into appointing their first professor<br />

of chemistry to research <strong>and</strong> teach materials science. He fo rmed an<br />

artists' cooperative to secure h<strong>and</strong>-ground pigments <strong>and</strong> pure materials. He<br />

monitored the condition of his own paintings, attentive to the conditions of<br />

their display <strong>and</strong> h<strong>and</strong>ling. He made test panels <strong>and</strong> stored them in his studio<br />

ten, fifteen, <strong>and</strong> twenty years to observe the effects of aging. With ever increasing<br />

obsession, he investigated material permanence, compatibility, <strong>and</strong><br />

composition. Like a true conservator, he fo rsook the artwork for the artmaking,<br />

producing fewer paintings over longer intervals in a painstakingly slow<br />

technique.<br />

Figure 5. William Holman Hunt, Self-Portrait,<br />

1875. Courtesy of the Ujfi zi Gallery,<br />

Florence, Italy. Note the length of the brushes<br />

on the table. The tunic is the same striped<br />

garment used in The Finding of the Saviour<br />

in the Temple.<br />

Figure 6. Nineteenth-century containers for<br />

oil paints (animal bladder with ivory tacks,<br />

piston tuhe, <strong>and</strong> collapsible metal tube).<br />

Courtesy of the Forbes Collection, Harvard<br />

University Art Museums.<br />

Hunt's passions flared in the mid-1870s with the realization that Roberson's<br />

orange vermilion, a favorite commercial tube paint, was being adulterated<br />

with red lead, <strong>and</strong> thus blackening on the canvas. In frustration, Hunt wrote<br />

to his friend <strong>and</strong> fellow artist, John Lucas Tupper (5):<br />

It seems as tho [sic] I were struggling against Fate. Every day sometimes<br />

including Sundays I have been toiling every hour, <strong>and</strong> just as I have got<br />

my task nearly completed the whole thing has fallen into disorder again for<br />

at least five or six times <strong>and</strong> I have had to begin again. At last I have<br />

fo und out what has been the cause oj this: Roberson's tube oj Orange<br />

Vermilion, which I used without suspicion because 25 years ago they sold<br />

this color absolutely pure, is adulterated with 10 percent oj villainy, the<br />

greater part lead, which has blackened so rapidly that when it had got dry<br />

enough fo r the f<br />

inal glazings the flesh had got to such a color [sic] that I<br />

nearly went crazy ... I have had the color analysed <strong>and</strong> at the same time<br />

have taken the opportunity to have others investigated <strong>and</strong> find that the<br />

fraudulent habit is exercised in many other cases. What is to me more<br />

discouraging than this is that many artists I have spoken to about [it] are<br />

quite satisfied to go on dealing in these spurious colors saying "Oh, they<br />

will last my time, " <strong>and</strong> "I never fo und my pictures change" <strong>and</strong> with<br />

base humility "they, " the colors, "are good enough for my work. " Leighton,<br />

when I proposed a little co-operative society for importing <strong>and</strong> grinding<br />

pure colors said, "And what's poor Roberson to do?"<br />

The culmination came on Friday, 23 April 1880, when Hunt addressed the<br />

members of the Royal Society of Arts on the subject, "The Present System<br />

of Obtaining <strong>Materials</strong> in Use by Artist Painters, as Compared with that of<br />

the Old Masters." At a conservative estimate, the talk lasted at least two-<strong>and</strong>a-half<br />

hours. Concerned as much with the decline in knowledge of artists'<br />

techniques as well as materials, Hunt observed, "In the old days the secrets<br />

were the artist's; now he is the first to be kept in ignorance of what he IS<br />

using" (6). Eloquently he informed his audience (7):<br />

I feel called upon to avow that I regard the artists' colourmen oj London<br />

as gentlemen of intelligence, of character, <strong>and</strong> great enterprise, to which<br />

qualities we are much indebted for the comparatively safe positions we enjoy;<br />

for indeed, at the worst, it must be recognised that we might have gone<br />

fu rther astray. It is needful, however, that we should be not only in good<br />

h<strong>and</strong>s, but we should give strong prooJ that we can distinguish between<br />

that which is fa ulty <strong>and</strong> that which is perfect; <strong>and</strong> it is the want of discriminating<br />

power in the painter which produces all the indiff erence on the<br />

part oj the preparer to the permanent character of the materials he supplies.<br />

The painter has really not the power to trace the causes of difects. The<br />

colourman naturally judges of the character of the materials he vends by the<br />

condition they are in while under his own eye. To him, the evils revealing<br />

themselves in the work which has passed through his shop do not exist if<br />

he never sees them; <strong>and</strong> if he hears of them only, as evils untraceable in<br />

their cause which have occurred to one of his customers (who may, sometimes,<br />

have obtained materials elsewhere), his sense oj responsibility is quieted,<br />

when he has received the assurance of his men in the workshop that<br />

the usual rules, which have hitherto resulted in work of a kind not eliciting<br />

160<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


complaints, have been strictly adhered to. The workmen, too, in these shops<br />

are not permanent, <strong>and</strong> there is virtually no responsibility fo r any one<br />

preparation. In most cases the complaint is never made, for the evil may<br />

be a very serious one, <strong>and</strong> yet it may not manifest itself bifore the death<br />

of the artist.<br />

In conclusion, he stated, "The cure that we have to seek is one it is possible<br />

to defme compactly. It is to establish a means of transmitting the practical<br />

wisdom of one generation of painters to another" (8).<br />

With striking prescience, he addressed the need "to fo und a society for looking<br />

after the material interests of painting . .. composed of important members<br />

of the profession of painting ... [joined by] gentlemen of reputation in<br />

chemical science." This society would form "a library ... of all works of<br />

literature which exist on the subject of artistic practices . .. ,establish a workshop<br />

for the preparation of materials . .. ,arrange for the importing of colours<br />

from abroad for collecting specimens of experiments. An important aspect of<br />

the society will be to cultivate the opportunities of obtaining further samples<br />

of every variety of colours existing in the far East, of proving these, <strong>and</strong><br />

putting them on record in our museum fo r all generations to see." A technical<br />

school would be established, <strong>and</strong> artists trained, "that thus we should be the<br />

inheritors, not only of our immediate predecessors, but the heirs of all the<br />

ages, <strong>and</strong> that, though our pretensions would not be ostentatious in our humble<br />

way, we might be proud that we should be repeating the chosen tasks of<br />

the gods, the directing of inert matter to a spiritual end" (9).<br />

The Pre-Raphaelite technique<br />

Fully aware of the frailty of the materials of painting, Hunt became a master<br />

technician. Most of his paintings have remained well preserved, retaining the<br />

exceptional brilliancy of color which he strove so hard to achieve via successive<br />

layering of minute strokes of transparent <strong>and</strong> semi-transparent paint.<br />

He chose quality linen canvas stretched on paneled stretcher supports as "protection<br />

against accidental injuries, such as a push from a corner of a picture<br />

frame in the confusion which precedes <strong>and</strong> follows exhibitions, a kind of<br />

injury which, if not visible at the time, may show years afterwards in starred<br />

cracks in the hard paint" (10).<br />

With some notable exceptions (such as the Liverpool Triumph of the Innocents,<br />

painted on h<strong>and</strong>kerchief linen from an Arab bazaar because the artist was too<br />

impatient to await a delayed shipment of supplies from Engl<strong>and</strong>), many of<br />

Hunt's paintings remain unlined today <strong>and</strong> on their original stretchers, in the<br />

same gilt frames he designed for them. On other occasions, he had the paintings<br />

lined as a prophylactic measure, either during execution or shortly after<br />

completion (11).<br />

Hunt chose his medium <strong>and</strong> surface coating with the same view to permanency<br />

as his pigments <strong>and</strong> support. Instead of the popular megilp, a gelled<br />

preparation of linseed oil <strong>and</strong> mastic varnish much used <strong>and</strong> abused by nineteenth-century<br />

British painters, Hunt used amber colors, smooth-flowing<br />

tube paints with pigments bound in a drying oil <strong>and</strong> copal resin. For once,<br />

Hunt diverged from sound technique, failing to foresee the embrittling effect<br />

to the paint layer over time from the addition of copal to the medium, along<br />

with the eventual yellowing of his colors due to the oxidation of the varnish<br />

component. With the best of intentions, he defended his choice of medium,<br />

noting that "amber varnish ... protects the colours very perfectly, but has<br />

two slight disadvantages, as it lessens the brilliance of the white by the richness<br />

of the yellow tone in the varnish, <strong>and</strong> permits each touch to spread, though<br />

very slightly. Both these difficulties, however, occur immediately <strong>and</strong> may be<br />

calculated for" (12).<br />

In spite of his profound interest in stability of materials <strong>and</strong> techniques, if<br />

Hunt is remembered at all today as a technical innovator, it is as the author<br />

of the Pre-Raphaelite technique, a technique actually attempted by very few<br />

Katz 161


Pre-Raphaelite painters <strong>and</strong> never to any great extent. The myth of the Pre­<br />

Raphaelite technique arose from a single paragraph in Hunt's 1,000-page,<br />

two-volume autobiography. Citing as an example the painting Valentine Rescuing<br />

Sylvia from Proteus (Fig. 7), Hunt stated that he would (13):<br />

Figure 7. William Holman Hunt, Valentine<br />

Rescuing Sylvia from Proteus (Two<br />

Gentlemen of Verona), 1851. Courtesy if<br />

the Birmingham Museums <strong>and</strong> Art Callery,<br />

Birmingham, Engl<strong>and</strong>. This is the painting<br />

Hunt was working on when he described the<br />

wet-ground technique ill his autobiography (I,<br />

276): "The heads of Valentine alld if Proteus,<br />

the h<strong>and</strong>s if these figures, <strong>and</strong> the brighter<br />

costumes in the same painting had been<br />

executed in this way . ... In the country we<br />

had used it, so fa r, mainly for blossoms of<br />

flowers, for which it was singularly valuable. "<br />

Paint daubs if van'ous mixtures of red are<br />

visible on the sp<strong>and</strong>rels.<br />

Select a prepared ground, originally Jor its brightness, <strong>and</strong> renovate it, if<br />

necessary, with Jresh white when first it comes into the studio, white to be<br />

mixed with a very little amber or copal varnish. Let this last coat become<br />

oj a thoroughly stone-like hardness. Upon this suiface complete with exactness<br />

the outline oj the part in h<strong>and</strong>. On the morning Jo r the painting,<br />

with Jresh white (from which all supeifluous oil has been extracted by means<br />

oj absorbent paper, <strong>and</strong> to which a small drop oj varnish has been added)<br />

spread a Ju rther coat very evenly with a palette knife over the part Jor the<br />

day's work, oj such consistency that the drawing should Ja intly show<br />

through. In some cases the thickened white may be applied to the Jo rms<br />

needing brilliancy with a brush, by the aid oj rectified spirits. Over this<br />

wet ground, the colour (transparent <strong>and</strong> semi-transparent) should be laid<br />

with light sable brushes, <strong>and</strong> the touches must be made so tenderly that<br />

the ground below shall not be worked up, yet so Jar enticed to blend with<br />

the superimposed tints as to correct the qualities oj thinness <strong>and</strong> staininess<br />

[sic], which over a dry ground transparent colours used would inevitably<br />

exhibit. <strong>Painting</strong> oj this kind cannot be retouched except with an entire<br />

loss oj luminosity.<br />

In spite of the attention given to this quotation, however, it is rarely pointed<br />

out that Hunt describes this as a technique with which he experimented at<br />

one point in his career, used fo r specific design areas rather than entire canvases<br />

<strong>and</strong> not as a wholesale working method, as has been interpreted. Indeed,<br />

until the appearance of the autobiography in 1905, Hunt appears to repudiate<br />

his early experimental technique. It remains unmentioned in his first version<br />

of his memoirs, nor does it crop up in the frequent h<strong>and</strong>written inscriptions<br />

in which he recorded details of technique on bare sections of canvas or support<br />

(14) (Fig. 8). Hunt dismisses its relevance to his career in an article on<br />

his painting technique that appeared in the magazine Porifolio in 1875, stating<br />

that he used wet grounds only from 1850 to 1854 to capture the effects of<br />

sunlight (15). While these five early years are viewed today as Hunt's period<br />

of greatest productivity <strong>and</strong> success, to his contemporaries his highest achievements<br />

came later in a career that spanned seven decades (16).<br />

The sudden prominence accorded to the technique in 1905, may be due not<br />

to Hunt himself, but to his wife, Edith, seeking to enhance his reputation<br />

through the implication of technical innovation. Suffering from glaucoma,<br />

Hunt dictated much of his memoir to Edith who, according to their gr<strong>and</strong>daughter,<br />

among others, took liberties with the text, "deleting passages ...<br />

she considered unsuitable for posterity" (author's emphasis) (17). Edith's urge<br />

to improve went so far as taking advantage of her husb<strong>and</strong>'s blindness to have<br />

a studio assistant repaint Hunt's portrait of her, secretly slimming her waist<br />

<strong>and</strong> reddening her lips (18). It may well have been at her suggestion that the<br />

wet-ground technique, after fifty years of oblivion, abruptly became of importance<br />

in her husb<strong>and</strong>'s career, serving as one more opportunity to assert<br />

his innovation, skill, <strong>and</strong> pivotal role in the movement he had helped to shape.<br />

Along with the disproportionate attention given to the Pre-Raphaelite technique<br />

comes a misunderst<strong>and</strong>ing of the technique itself, due to a fu ndamental<br />

misreading of the word "ground." In Hunt's writings, the term often refers<br />

to an imprimatura, or underpaint, layer, rather than the intermediary priming<br />

that prepares a solid or fabric support to receive paint. In his diary, fo r example,<br />

Hunt refers to "lay[ing] a new ground for the left shoulder, which I<br />

do of white, cobalt green, <strong>and</strong> cadmium" (19). Further investigation confirms<br />

that Hunt's "wet ground" was actually a layer of paint, the "fresh white" of<br />

his writings referring to flake white oil paint, as he indicated in a letter of<br />

1878 (20):<br />

162<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


F(gure 8. Illfrared rriflectograph of the hal1dwritten notatioll under the paint layer il1 the upper right<br />

corner of The Miracle of the Sacred Fire, Church of the Holy Sepulchre at Jerusalem, at the<br />

HalVard University Art Museums. Inscriptions such as this one are frequel1tly fo und in paintillgs by<br />

HUllt, either on the reverse of the supports, 011 bare sp<strong>and</strong>rels, or ul1derneath the pail1t layers. Image<br />

provided by the author.<br />

I am obliged to wait long Jo r the drying oj the paint I put on to Jo rm a<br />

Jresh ground, <strong>and</strong> while patience is being exercised as a necessity it seems<br />

but a little virtuous to summon hope to help me over the completing oj the<br />

parts where the ground has finally become workably [sic] even <strong>and</strong> dry.<br />

Indeed, Hunt complains here that the underpaint layer is wet, rather than<br />

seeking to exploit its freshness as part of his technique.<br />

This reevaluation of the Pre-Raphaelite technique is consistent with the results<br />

of analyses carried out on cross sections taken from various paintings by<br />

Hunt, which indicate the blending of paint layers wet-into-wet, a common<br />

technique throughout the history of painting, rather than the swirling of paint<br />

layers into wet primings (21). This clarification not only places Hunt's<br />

achievements in a far more plausible technical frame, but also emphasizes, as<br />

he did throughout his career, his connection with <strong>and</strong> indebtedness to the<br />

Katz 163


paintings techniques of the past masters, rather than the faddish disregard for<br />

tradition prevalent among his contemporaries.<br />

Conclusion<br />

In seeking reform, Hunt sought not to stop the wheels of progress but to<br />

channel them. His advocacy for the rights of the consumers in a time of<br />

rampant laissez-faire capitalism had ramifications throughout his century <strong>and</strong><br />

our own, contributing directly to the enactment of regulations governing the<br />

safety of commercial goods <strong>and</strong> legislation regarding truth in advertising. His<br />

sophisticated underst<strong>and</strong>ing of the science of painting, in spite of limited<br />

formal schooling, <strong>and</strong> the impact of environment in the preservation of art<br />

place him as a key figure in the development in the professions of art conservation<br />

<strong>and</strong> museology. His underst<strong>and</strong>ing of historic painting techniques,<br />

well advanced fo r his day as well as our own, generated a revival of the craft<br />

of painting after a generation notorious for technical inadequacies. Too often<br />

he has been dismissed by art historians as a minor painter of deeply tasteless<br />

religious scenes. His contribution to nineteenth-century British painting <strong>and</strong><br />

to the current stability of commercial artists' materials must not be underestimated.<br />

Acknowledgments<br />

The author wishes to thank Kate Olivier (Fogg Art Museum, Harvard University)<br />

for having encouraged the undertaking of this research; Jacqueline Ridge (National<br />

Museums <strong>and</strong> Galleries on Merseyside, Liverpool), Zahira Veliz (private practice, London),<br />

<strong>and</strong> David Bomford (National Gallery, London) fo r facilitating research in Britain;Judith<br />

Bronkhurst (Witt Library, Courtauld Institute, London) <strong>and</strong> Leslie Carlyle<br />

(Canadian Conservation Institute, Ottawa) for their valuable insights into nineteenthcentury<br />

British art <strong>and</strong> techniques; <strong>and</strong> the staff of the following institutions fo r<br />

generous access to curatorial <strong>and</strong> conservation files pertaining to William Holman<br />

Hunt: Fogg Art Museum, Cambridge, Massachusetts; Wadsworth Atheneum, Hartfo<br />

rd, Connecticut; Tate Gallery, London; Courtauld Institute of Art, London; Guildhall<br />

Art Gallery, London; Walker Art Gallery, Liverpool; Lady Lever Art Gallery, Port<br />

Sunlight; Manchester City Art Galleries, Manchester; Birmingham Museums <strong>and</strong> Art<br />

Gallery, Birmingham; <strong>and</strong> the Roberson Archives, Hamilton Kerr Institute, Cambridge<br />

University, Cambridge.<br />

Notes<br />

1. Harley, R. D. 1970. Artists' Pigments c. 1600-1835: A Study in English Documentary<br />

Sources. London: Butterworth Scientific. Also, Gettens, R., <strong>and</strong> G. Stout, 1966.<br />

<strong>Painting</strong> <strong>Materials</strong>: A Short Encyclopaedia. New York: Dover Publications, Inc.<br />

2. Bomford, D., et al. 1990. Art in the Making: Impressionism. London: The National<br />

Gallery.<br />

3. St<strong>and</strong>age, H. C.,1886. The Artists' Manual of Pigments showing their Composition:<br />

Conditions of Permanency, Non-Permanency, <strong>and</strong> Adulterations; Effects in Combination<br />

with Each Other <strong>and</strong> with Vehicles; <strong>and</strong> the Most Reliable Tests of Purity . .. . London:<br />

Crosby, Lockwood & Co.<br />

4. St<strong>and</strong>age, op. cit., 53-54.<br />

5. A Pre-Raphaelite Friendship: The Correspondence of William Holman Hunt <strong>and</strong> John<br />

Lucas Tupper, no. 114 (11 August 1875). 1986. Eds. J. H. Coombs, et al. Ann<br />

Arbor, Michigan: UMI Research Press, 20l.<br />

6. Hunt, W H. 1880. The Present System of Obtaining <strong>Materials</strong> In Use By Artist<br />

Painters, As Compared With That of the Old Masters. Journal of the Society of<br />

Arts 28 (23 April 1880), 492.<br />

7. Hunt, op. cit, 492.<br />

8. Ibid., 493.<br />

9. Ibid.<br />

10. Hamerton, P. G. 1875. Technical Notes-W Holman Hunt, an interview with<br />

the artist. The Porifolio: An Artistic Periodical VI (1):45.<br />

11. Original paint extends over onto the tacking margins of the lining canvas <strong>and</strong><br />

the gummed tape edging the lining of the Birmingham Finding of the Saviour in<br />

the Temple. Conservation files, Birmingham Museums <strong>and</strong> Art Gallery.<br />

12. Hamerton, op. cit., 45.<br />

164<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


13. Hunt, W. H. 1905-1 906. Pre-Raphaelitism <strong>and</strong> the Pre-Raphaelite Brotherhood. London:<br />

Macmillan & Co., (1):276.<br />

14. Hunt, W H. 1886. The Pre-Raphaelite Brotherhood: A fight fo r art. The Contemporary<br />

Review 49 (April):471-88, (May) :737-50, Oune):820-33.<br />

15. Hamerton, op. cit., 46.<br />

16. Of course, the Pre-Raphaelite Brotherhood in its strictest sense only endured<br />

from 1848 to 1854, so a purist might claim it as a Pre-Raphaelite technique,<br />

although all the Brothers quickly returned to traditional painting techniques, <strong>and</strong><br />

only Hunt <strong>and</strong> John Everett Millais had ever developed it to any great extent.<br />

17. Holman-Hunt, D. 1969. My Gr<strong>and</strong>father, His Wives <strong>and</strong> Loves. London: Hamish<br />

Hamilton, 18.<br />

18. AITlOr, A. C. 1989. William Holman Hunt: The True Pre-Raphaelite. London: Constable,<br />

266.<br />

19. Hunt, W. H. 1872. Personal diary (May 25, 1872). Original manuscript in the<br />

John Ryl<strong>and</strong>s University Library, Manchester, Engl<strong>and</strong>.<br />

20. A Pre-Raphaelite Friendship, no. 136 (16 October 1878), 258.<br />

21. Consistent results have been achieved in cross-section analysis carried out by the<br />

author, <strong>and</strong> by conservators at the Tate Gallery <strong>and</strong> University College, London,<br />

among others.<br />

Katz 165


Abstract<br />

The production of a painting in early<br />

nineteenth-century France followed<br />

a clearly defined sequence of<br />

steps. After drawings had been made,<br />

the composition was outlined on the<br />

prepared canvas <strong>and</strong> the modeling<br />

was indicated, often with a reddishbrown<br />

"sauce." Local color, light, <strong>and</strong><br />

shade were laid in; this sketch was<br />

elaborated using a full range of tones<br />

laid out individually on the ebauche<br />

(palette). The final stage of painting<br />

refined this process further. This is<br />

demonstrated in Paul Delaroche's<br />

The Execution of Lady Jane Grey<br />

(1833). A pupil of Watelet <strong>and</strong> Gros,<br />

Delaroche received much popular<br />

acclaim during the 1830s, <strong>and</strong> a<br />

number of eminent painters passed<br />

through his studio. Thus he occupies<br />

a central position in the history of<br />

academic painting.<br />

Paul Delaroche: A Case Study of Academic <strong>Painting</strong><br />

Jo Kirby* <strong>and</strong> Ashok Roy<br />

Scientific Department, National Gallery<br />

Trafalgar Square<br />

London WC2N 5DN<br />

United Kingdom<br />

Introduction<br />

The July Monarchy of Louis Philippe, who came to the throne in 1830, was<br />

a period of technological advance <strong>and</strong> increasing industrialization, marked by<br />

the rise of a wealthy <strong>and</strong> influential middle class. It was distinguished by its<br />

adherence to the philosophy of eclecticism, not only in politics, but also in<br />

the realm of the arts. Official art followed a middle course between the two<br />

dominant trends, Classicism <strong>and</strong> Romanticism, showing the careful composition,<br />

drawing, <strong>and</strong> modeling of the former <strong>and</strong> an interest in the subject<br />

matter <strong>and</strong> emotional content of the latter. The painting of the juste milieu<br />

could justly be described as the art of the bourgeoisie. Paul Delaroche, who<br />

rose to prominence at this time, was one of its most popular <strong>and</strong> successful<br />

exponents (1, 2).<br />

Born in 1797, the son of an art dealer, Delaroche (christened Hippolyte)<br />

entered the studio of Antoine-Jean Gros, a disciple of David, after early training<br />

with the l<strong>and</strong>scape painter Louis-Etienne Watelet <strong>and</strong> with Constant<br />

Desbordes. The recipient of many honors, royal patronage, <strong>and</strong> several official<br />

commissions, Delaroche achieved early Salon success. In 1833, he inherited<br />

Gros's studio <strong>and</strong> became a professor at the Ecole des Beaux-Arts (3, 4, 5, 6).<br />

His atelier was perhaps the busiest <strong>and</strong> most effective of the period; his pupils<br />

included Gerome, Daubigny, Millet, Monticelli, <strong>and</strong> Thomas Couture, himself<br />

the master of Edouard Manet (7).<br />

The Execution of Lady Jane Grey<br />

Delaroche's The Execution if Lady Jane Grey, finished in 1833, achieved considerable<br />

success at the Salon exhibition of 1834. The subject, drawn from<br />

English Tudor history <strong>and</strong> depicted with ostensible accuracy, appealed to popular<br />

taste. The scene depicted-the moment immediately before the beheading-was<br />

that of the greatest dramatic tension; it also touched the sensibilities<br />

of the public without disgusting them. As Etienne-Jean Delecluze wrote,<br />

"The spectator can contemplate the axe ... without horror" (8) (Plate 33).<br />

The blindfolded Lady Jane fu mbles for the block; a figure who is probably<br />

intended to be Sir John Brydges, the Lieutenant of the Tower of London,<br />

gently guides her h<strong>and</strong>. On the left, a despairing lady-in-waiting turns her<br />

face towards the massive column; the other lady-in-waiting, her mistress's<br />

discarded dress across her lap, has fainted. The impassive executioner st<strong>and</strong>s<br />

on the right. Delaroche's historical sources for the painting included the Martyrologue<br />

des Protestans of 1588, quoted in the Salon catalogue (9) . There were<br />

a number of other publications, as well as the works of other painters, upon<br />

which he could have drawn, including Hans Holbein the Younger's painting<br />

of Anne of Cleves, which was at the Louvre (10, 11). Delaroche undertook<br />

exhaustive research before any painting project (12).<br />

Adored by the crowds, the painting was praised <strong>and</strong> condemned in almost<br />

equal measure by the critics. The criticism that Delaroche's treatment was<br />

theatrical rather than dramatic, voiced by Gustave Planche among others, is<br />

interesting as it may reflect an aspect of the artist's practice: Delaroche used<br />

small model rooms, within which he arranged wax figures to assist in the<br />

composition of his paintings (13, 14). According to Edward Armitage, a for-<br />

* Author to whom correspondence should be addressed.<br />

166<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 1. The Execution of Lady Jane Grey, after cleanirlg a/1d before restoration.<br />

mer student, the use of similar "boxes" to study the fall of light upon small<br />

draped nlOdels was not uncommon (15). It is fair comment that the scene<br />

depicted takes place uncomfortably close to the picture plane within a stagelike<br />

space.<br />

The 1834 Salon catalogue states that the painting was owned by Prince Anatole<br />

Demidoff, who paid 8,000 francs for it. It remained in private ownership<br />

until it was bequeathed to the National Gallery by the second Lord Cheylesmore<br />

in 1902 (16). In 1928 it suffered flood damage while at the Tate<br />

Gallery <strong>and</strong> was later described as destroyed (17). In 1973 it returned to the<br />

National Gallery, where examination showed that the condition of the painting<br />

was not as bad as had been feared, consisting principally of tears in the<br />

canvas (the largest tear through the executioner's feet) <strong>and</strong> associated paint<br />

losses. Figure 1 shows the picture before restoration after it had been repaired,<br />

double lined with wax resin, <strong>and</strong> cleaned (1974-1975). Evidently its construction<br />

must have been relatively sound, <strong>and</strong> this was confirmed by its recent<br />

technical examination (18).<br />

Method <strong>and</strong> materials of painting<br />

The highly fmished Salon painting or Academic picture was the result of a<br />

well-established procedure; the artist would make preliminary studies, an esquisse<br />

(a sketch of the intended composition), <strong>and</strong> detailed drawings before<br />

transferring the design to canvas <strong>and</strong> beginning to paint (19, 20, 21,22). These<br />

Kirby <strong>and</strong> Roy 167


Figure 2. Paul Delaroche, Study for The Execution of Lady Jane Grey, ca. 1833. Watercolor<br />

<strong>and</strong> body color over pencil, with varnish, 18.4 X 14.3 em. University of Manchester, f;IIhitworth Art<br />

Gallery.<br />

steps may be observed in The Execution if Lady Jane Grey <strong>and</strong> other works<br />

by Paul Delaroche. It must be remembered that the artist may have been<br />

assisted by students at certain stages in the production of a painting of this<br />

size, even though there is no obvious indication of this here (23).<br />

Compositional sketches for several of Delaroche's paintings survive; Joan of<br />

Arc in Prison (London, Wallace Collection), for example, was painted as the<br />

sketch fo r Joan if Arc . . . Interrogated in Prison by the Cardinal of Winchester<br />

(1824, Rouen, Musee des Beaux-Arts) (24, 25). Most are smaller than the<br />

fmal versions <strong>and</strong> all are more freely painted. Delaroche felt strongly that a<br />

preliminary sketch embodied the artist's imaginative process <strong>and</strong> inspiration<br />

(26). The only known compositional study for The Execution if Lady Jane<br />

Grey is a small watercolor in the Whitworth Art Gallery, Manchester (Fig. 2).<br />

Here, Delaroche sets the figures within a Romanesque interior <strong>and</strong> only the<br />

executioner, who st<strong>and</strong>s in profile holding a sword, is significantly different<br />

from the large final version (27). Many minor alterations can be attributed to<br />

scaling up <strong>and</strong> improving the composition: modifYing the background arcade<br />

<strong>and</strong> staircase, for example, <strong>and</strong> simplifYing the costumes. The striking transformation<br />

is in the color, which is so resonant <strong>and</strong> warm in the finished<br />

version in contrast to the cooler <strong>and</strong> less coherent tonalities of the watercolor.<br />

In the sketch, the executioner is dressed in dull green <strong>and</strong> red; the attendant<br />

facing the column is portrayed in deep blue, rather than rich dark purple. In<br />

the sketch, light plays evenly across the room; in the painting, it is more<br />

concentrated on the figures, although the pattern of light fall is similar. The<br />

squat, oddly appealing figures of the tiny watercolor have been transformed<br />

into an elegant, theatrical "tableau."<br />

The next step was to make drawings for the composition's elements. Much<br />

importance was traditionally attached to drawing, <strong>and</strong> Delaroche produced<br />

hundreds of drawings during his career (28). Several must have been made<br />

for The Execution if Lady Jane Grey. Two studies for it on paper certainly<br />

survive, one in the Musee du Louvre in Paris, the other in the British Museum<br />

in London (Fig. 3) (29). The Louvre sheet shows the figure of the<br />

executioner on the left, squared up fo r transfer: It is very close to that in the<br />

168<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 3. The Execution of Lady Jane Grey. Pencil drawing for the central group, 18 X 16.5<br />

cm. London, British Museum {fo rmerly National Gallery, London}.<br />

finished painting. The two other figures on the sheet are studies for Lady<br />

Jane's ladies-in-waiting; neither has been squared up. That for the st<strong>and</strong>ing<br />

figure has been used with little modification. The other figure in the Louvre<br />

drawing appears to kneel <strong>and</strong> pray facing the central group; this composition<br />

was not used in the final painting, in which the slumped position of the<br />

attendant is very much as it is in the watercolor study-to greater dramatic<br />

effect. The British Museum drawing shows Lady Jane <strong>and</strong> Sir John Brydges;<br />

it has been used with little modification. There are small differences in details<br />

of costume between the drawings used <strong>and</strong> the finished painting.<br />

Apart from his use of draped figures in room sets, discussed above, the artist<br />

was in the habit of making wax or plaster models of elements of the composition<br />

when he felt there was a problem to be solved; a plaster model was<br />

made of the two princes, for example, in The Children if Edward IV (1830,<br />

Musee du Louvre) (30). This appears not to have been necessary for the Lady<br />

Jane Grey.<br />

The chosen drawn versions of the figures would then have been transferred<br />

to the prepared canvas. The canvas for a painting of this size would have been<br />

stretched to order, apparently from a single piece of medium-weight linen;<br />

the largest stock canvas was a toile de 120 (about 1.3 X 1.9 m) (31). The roll<br />

of canvas used was already primed with a ground of lead white in linseed oil.<br />

Over this first priming, Delaroche then had a second ground-also of lead<br />

white, but this time bound in walnut oil-applied to the stretched canvas.<br />

Delaroche seems to have transferred designs for the single figures or groups<br />

individually; infrared reflectography has revealed "squaring-up" lines under<br />

the paint of Lady Jane's dress, her seated attendant, <strong>and</strong> in the area of the<br />

Kirby <strong>and</strong> Roy 169


Figure 4. Il1frared reftectogram from Lady jane Grey's skirt in The Execution of Lady Jane Grey.<br />

block, but the scale of these various grids suggests that there is no single<br />

system of squaring across the entire picture surface. The infrared images show<br />

many changes in the underdrawing; lines drawn across Lady Jane's wrists<br />

suggest that her dress was to be long-sleeved, as in the watercolor sketch (Fig.<br />

4). At the drawing stage, Delaroche also indicated such fe atures as shadows<br />

in drapery folds, as seen in the st<strong>and</strong>ing woman's dress.<br />

In many unfinished French paintings of this period it can be seen that the<br />

underdrawing has been strengthened with a translucent brownish wash, sometimes<br />

known as "sauce" (32, 33). In this case, no such material could be<br />

detected with certainty in any of the cross sections examined, <strong>and</strong> only the<br />

charcoal drawing was fo und; possibly a grayish wash, for example, was applied<br />

in the shadows of folds.<br />

Subsequently the figures <strong>and</strong> background were laid in using brownish or<br />

grayish shades of paint, composed of lead white combined with a variety of<br />

tinting pigments (Cassel earth, ochres, <strong>and</strong> other natural earth pigments, as<br />

well as small quantities of cobalt blue, a red lake pigment <strong>and</strong> black) . The<br />

tonality of this underpaint bears some relation to the color of the paint that<br />

was to be applied on top, thus the background <strong>and</strong> black garments are underpainted<br />

in shades of gray, while the underpaint of the flesh varies from a<br />

warm beige fo r the executioner to a grayish white for Lady Jane. The underpaint<br />

also indicates light <strong>and</strong> shade, by varying the proportion of ochres,<br />

black, <strong>and</strong> cobalt blue in the mixture, the color of the underpaint of Lady<br />

Jane's dress is changed from a pale beige in areas of highlight to a dark brown<br />

in the deepest shadow. Occasionally, the underpaint is similar in color to the<br />

intended local color; for the cushion it is a translucent green consisting of<br />

black, verdigris, Pruss ian blue, yellow ochre, <strong>and</strong> perhaps a yellow lake.<br />

At this point the painting would have had an appearance not unlike a grisaille<br />

version of the final composition, an element of Delaroche's practice mentioned<br />

by Delaborde (34). Infrared reflectography has revealed many pentimenti,<br />

however, bearing out comm.ents by students that Delaroche frequently<br />

reworked passages during painting (35). The executioner appears originally<br />

to have had a sash around his waist, for example, <strong>and</strong> the position of the pike<br />

behind the balustrade <strong>and</strong> the angle of the banisters have been altered. The<br />

underpaint for the executioner's tights is mauve-gray (composed of lead<br />

170<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 5. Detail of Lady Jal1e Grey from The ExeclItion of Lady Jane Grey.<br />

white, red lake, <strong>and</strong> black), similar in color to that used in the watercolor<br />

sketch; the presence of a thin layer of black paint above it, darkening or<br />

obliterating it, suggests a change of mind. Above it is a much warmer version<br />

of the underpaint layer, containing more red lake pigment, followed by the<br />

desired red paint (Plate 34a) .<br />

This grisaille probably represents the ebauche stage of the painting (36).<br />

When it was dry, the local color was applied. Bouvier describes the use of a<br />

series of premixed, graduated middle tones applied side by side <strong>and</strong> then<br />

blended to give an smooth passage from light to dark across each part of the<br />

painting, for both the ebauche <strong>and</strong> the final paint layers (37). The sim.plicity<br />

<strong>and</strong> precision of the layer structure <strong>and</strong> the generally close relationship between<br />

the depth of tone in the grisaille <strong>and</strong> the chiaroscuro of the finished<br />

painting, as demonstrated in the painting of Lady Jane's dress, suggests that<br />

something approaching this may have been done (Fig. 5). Only one or two<br />

quite thin, even layers of paint may be present over the underpaint, with<br />

perhaps an additional glaze or highlight. Observation of the paint surface,<br />

however, <strong>and</strong> the occasional presence of extrem.ely thin scumbles or glazes of<br />

paint (or sim.ply a greater concentration of paint medium) at the top of a<br />

paint layer reveal the careful blending of tones. One paint layer frequently<br />

appears to have been applied over another while the layer below was still<br />

fresh, enabling one mid-tone to be merged into the next. Plate 34b shows a<br />

cross section of paint from the shadow of a fold in Lady Jane's dress. The<br />

grayish paint of the shadow, consisting of lead white with a little cobalt blue<br />

<strong>and</strong> Cassel earth, merges with the creamier paint of the layer below so com-<br />

Kirby <strong>and</strong> Roy 171


pletely that it is difficult to distinguish between them. The lowest paint layer<br />

(above the grayish wash of the drawing <strong>and</strong> the white ground) is the dark<br />

undermodeling of the ebauche, containing Cassel earth, charcoal, <strong>and</strong> cobalt<br />

blue. This is probably sufficiently dark in color to contribute to the observed<br />

chiaroscuro of the dress (a similar mixture is used as a final glaze fo r the<br />

deepest shadows) . The creamy mid-tone of the dress consists principally of<br />

lead white, with traces of yellow ochre, Cassel earth, <strong>and</strong> cobalt blue; the<br />

proportion of the tinting pigments is altered as one tone blends into the next,<br />

however. The blending process is also reflected in the presence of a number<br />

of paint layers in subtly different shades of cream in different parts of the<br />

dress. Only in the lightest highlights is lead white used almost pure, ground<br />

in walnut oil; even here, the presence of a trace of cobalt blue gives it coldness.<br />

This general pattern of paint construction is repeated elsewhere in the painting.<br />

Passages of paint containing a pigment used more or less unmixed are<br />

very rare indeed, but in the red glazes supplying the purplish red of the<br />

brocade dress, on the lap of the seated lady-in-waiting in Figure 6c, <strong>and</strong> the<br />

shadows on the executioner's tights, a crimson lake pigment was used in this<br />

way. In both cases, the dyestuff present was that extracted from a cochineal<br />

insect, Dactylopius coccus Costa, on a substrate consisting largely of hydrated<br />

alumina. The brown paint of the brocade, containing yellow ochre with other<br />

iron oxides <strong>and</strong> black, contained ordinary linseed oil; the red glaze, however,<br />

contained heat pre-polymerized linseed oil <strong>and</strong> a little mastic resin (Plate 34c).<br />

This indicates the use of a varnish of the type recommended as a painting<br />

medium, perhaps the jellylike vernis des Anglais described by Merimee as being<br />

particularly suitable fo r glazes because it could be brushed on so easily (38).<br />

The presence of a resin-containing medium is also suggested by the whitish<br />

fluorescence exhibited by the glaze layer in ultraviolet illumination under the<br />

mIcroscope.<br />

The paint used for the red of the executioner's tights is perhaps surprisingly<br />

complicated, as it contains two red lake pigments mixed with vermilion <strong>and</strong><br />

lead white (Plate 34a). Examination under the microscope suggests that one<br />

is the cochineal lake used in the glaze; the other, less crimson in color, was<br />

not present in sufficient quantity fo r analysis, but its pronounced orange-pink<br />

fluorescence in ultraviolet illumination suggests that the dyestuff may have<br />

been extracted from madder root, the use of which was being developed in<br />

France at the time (39). Curiously, the same lake (mixed with black <strong>and</strong> cobalt<br />

blue) is used rather than the more crimson cochineal lake for the attendant's<br />

purple dress. Even the velvety black of Sir John Brydges' garment is a mixed<br />

color: it contains a subtle combination of black, Prussian blue, red lake, <strong>and</strong><br />

a translucent yellow pigment (Plate 34d). This combination is similar to<br />

Edouard Manet's tinted darks in Music in the Tuileries Gardens, painted thirty<br />

years later (40). Quite marked brushwork is visible in the black garments of<br />

Sir John <strong>and</strong> the executioner; analysis shows that Delaroche used a varnishtype<br />

painting medium of similar composition to that used for the red glaze,<br />

which retained the texture of the brush strokes.<br />

To summarize, Paul Delaroche painted a grisaille of his composition <strong>and</strong> then<br />

colored it in; the labor necessary to produce the finished painting was, however,<br />

considerable barely revealed on its bl<strong>and</strong> surface. The painter's craftsmanship<br />

<strong>and</strong> underst<strong>and</strong>ing of materials cannot be denied; his reliance on<br />

lead white <strong>and</strong> ochres in particular <strong>and</strong> the absence of bitumen have resulted<br />

in a paint film in remarkably good condition, considering the recent history<br />

of the painting. The use of a varnish-type paint medium has often proved to<br />

be a recipe fo r disaster; in this case the paint shows few of the defects often<br />

caused by its use. Delaroche appears to have added only cobalt blue <strong>and</strong><br />

synthetic ultramarine (<strong>and</strong> possibly an improved madder lake) to what could<br />

be described as a conventional eighteenth-century palette.<br />

The production of the Salon painting was not the end of the story. Public<br />

awareness of successful Salon paintings was increased by means of reproductions<br />

in the press; several of Delaroche's paintings, including Children of Edward<br />

172<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 6. Paul Delaroche, The Execution of Lady Jane Grey, ca. 1834. Canvas, 45. 7 X 53.3<br />

e/11. Corporation of London, Guildhall Art Gallery.<br />

IV, were reproduced by lithography. Artists were also able to keep their works<br />

in the public eye by means of high-quality engravings <strong>and</strong> Delaroche had an<br />

arrangement with the print publisher Adolphe Goupil to produce <strong>and</strong> publish<br />

engravings, after his paintings (41). For large paintings, including Lady Jane<br />

Grey, The Children of Edward IV, <strong>and</strong> Joan of Arc, he painted a reduced-scale<br />

copy from which the engraving could be made (42) . In the case of Lady Jane<br />

Grey, it is possible that the painting now in the collection of the Guildhall<br />

Art Gallery, Corporation of London, is the reduced-scale copy (Fig. 6) (43).<br />

The engraving, by Mercurij, one of several engravers who engraved the painter's<br />

works fo r Goupil, was begun in 1834, but was only completed in 1857,<br />

the painter himself having died the previous year (44).<br />

Acknowledgments<br />

The authors would like to thank John Leighton, curator of nineteenth-century paintings<br />

at the National Gallery, fo r generously sharing the information he has gained<br />

during his research with the authors; Rachel Billinge, Leverhulme Research Fellow<br />

at the National Gallery, for infrared reflectography carried out on the painting; Raymond<br />

White fo r examination of the paint medium by gas chromatography mass<br />

spectrometry; Sara Hattrick <strong>and</strong> the Photographic Department for photography.<br />

Notes<br />

1. Boime, A. 1971. The Academy <strong>and</strong> French <strong>Painting</strong> in the Nineteenth Century. London:<br />

Phaidon, 9-21.<br />

2. Boime, A. 1980. Thomas Couture <strong>and</strong> the Eclectic Vision. New Haven <strong>and</strong> London:<br />

Yale University Press, 3-35.<br />

3. Delaborde, H., <strong>and</strong> J. Godde. 1858. Oeuvre de Paul Delaroche, reproduit en photographie<br />

par Bingham, accompagne d'une notice sur la vie et les ouvrages de Paul Delaroche<br />

par Henri Delaborde, et du catalogue raisonne de l'oeuvre par Jules GoddlE. Paris: Goupil.<br />

References to the essay are cited as Delaborde, 1858; those to the catalogue as<br />

Godde, 1858.<br />

4. Ziff, N. D. 1977. Paul Delaroche: a study in nineteenth-century French history painting.<br />

New York <strong>and</strong> London: Garl<strong>and</strong> Publishing, Inc.<br />

5. Boime, 1980. Op. cit., 40-41, 52-60, 70-75.<br />

Kirby <strong>and</strong> Roy 173


6. Foissy-Aufrere, M.-P 1983. La jeartlle d'Arc de Paul Delaroche, Salon de 1824:<br />

dossier d'une oeuvre. Rouen: Musee des Beaux-Arts, 82-87.<br />

7. Hauptman, W 1985. Delaroche's <strong>and</strong> Gleyre's teaching ateliers <strong>and</strong> their group<br />

portraits. Studies in the History of Art (18) :79-82, 89-97.<br />

8. Delecluze, E.-J. 1834. Salon de 1834 (2nd article). joumal des debats, politiques et<br />

litteraires 5 March 1834.<br />

9. Paris Salon de 1834. Explication des ouvrages de peinture, sculpture, architecture, gravure<br />

et lithographie des artistes vivalls exposes au Musee Royal Ie 1" mars, 1834. Paris:<br />

Vinchon. Facsim.ile edition, 1977. New York <strong>and</strong> London: Garl<strong>and</strong> Publishing<br />

Inc., 52 (no. 503).<br />

10. Ziff, op cit., 126-35.<br />

11. Ganz, P 1950. The <strong>Painting</strong>s of Ha1/5 Holbeill. London: Phaidon, 251 (no. 107)<br />

<strong>and</strong> plate 148.<br />

12. Mirecourt, E. de, [C-J.-B.Jacquot]. 1871. Histoire co/I.temporaine: portraits et silhouettes<br />

au XIX" siMe. 119. Delaroche, Decamps. 3rd ed. Paris: Librairie des contemporains,<br />

16-17.<br />

13. Planche, G. 1855. Salon de 1834. In Etudes sur I 'ecole franfaise (1831-1852), peinture<br />

et sculpture. Paris: Michel Levy Freres, Vol. 1, 241.<br />

14. Boime, 1980. Op. cit., 56.<br />

15. Armitage, E. 1883. Lectures on painting, delivered to the students of the Royal Academy.<br />

London: Tri.ibner & Co., 254-56.<br />

16. Gould, C 1975. Delaroche <strong>and</strong> Gautier: Gautier's views on the ((Execution of Lady<br />

jane Grey" <strong>and</strong> on other compositions by Delaroche. Leaflet accompanying the exhibition<br />

of the restored painting. London: Trustees of the National Gallery, Publications<br />

Department.<br />

17. Alley, R. 1959. Tate Gallery catalogues. The fo reign paintings, drawings <strong>and</strong> sculpture.<br />

London: Trustees of the Tate Gallery, 298.<br />

18. Examination of the paint cross sections <strong>and</strong> pigments was carried out by Ashok<br />

Roy using optical mjcroscopy, scanning electron microscopy coupled with energy-dispersive<br />

X-ray analysis, <strong>and</strong> X-ray diffraction. Identification of the paint<br />

media by gas chromatography mass spectrometry was performed by Raymond<br />

White. Identification of lake pigment dyestuffs by high performance liquid chromatography<br />

was carried out by Jo Kirby. The conservation treatment is described<br />

in the National Gallery conservation records.<br />

19. Boime, 1971. Op. cit., 22-47.<br />

20. Bouvier, P-L. 1832. Manuel des jeunes artistes et amateurs en peinture. 2nd ed. Paris:<br />

F. G. Levrault.<br />

21. Paillot de Montabert,J.-N. 1829. Traite complet de la peinture. Paris: Bossange pere,<br />

Vol. 9, 36-65.<br />

22. Arsenne, L.-C 1833. Manuel du peintre et du sculpteur [etc.]. Paris: Librairie encyclopedique<br />

de Roret, Vol. 2, 195-413.<br />

23. For example, Godde states that the ebauche of Straffo rd (1835, location unknown)<br />

was painted by Henri Delaborde from Delaroche's watercolor sketch <strong>and</strong> a wax<br />

model. Godde, 1858. Op. cit. (note 3), plate 19 <strong>and</strong> commentary.<br />

24. Ingamells,J. 1986. The Wallace Collection. Catalogue of Pictures, II, French Nineteenth<br />

Century. London: Trustees of the Wallace Collection, 114 (no. P604).<br />

25. Foissy-Aufrere, op. cit., 21-23.<br />

26. Boime, 1971. Op. cit., 45, 102-3.<br />

27. Godde, 1858. Op. cit. (note 3), plate 16 <strong>and</strong> commentary.<br />

28. Ziff, N. D. 1975. Dessins de Paul Delaroche au cabinet des Dessins du musee du<br />

Louvre. La Revue du Louvre et des Musees de France, XXV: 163-68. A further<br />

donation of drawings was made to the Louvre in 1982.<br />

29. Etude pour la 'Jane Gray" [sic], 1727RF. Pencil <strong>and</strong> red chalk, 23. 1 X 20 cm.<br />

See Guiffrey, J., <strong>and</strong> P Marcel. 1910. Inventaire general des dessins du Musee du<br />

Louvre et du Musee de Versailles. Ecole franfaise. Paris: Librairie centrale d'art et<br />

d'architecture. Vol. 5, 2-3 (no. 3535).<br />

30. Delaborde <strong>and</strong> Godde, op. cit., 15, plate 14 <strong>and</strong> conU11entary.<br />

31. Paillot de Montabert, op. cit., vol. 9, 144-47.<br />

32. Bouvier, op. cit., 236-37.<br />

33. Couture, T. 1879. Conversations on Art Methods lMetlLOde et entretiens d'atelierJ.<br />

(French edition, 1867). New York: G. P Putnam's Sons, 7.<br />

34. Delaborde, op. cit., 18.<br />

35. Armitage, op. cit., 79.<br />

36. Boime, 197 1. Op. cit., 36-41.<br />

37. Bouvier, op. cit., 207-74.<br />

174<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


38. Merimee,J.-F.-L. 1830. De la peinture a I'lwile [etc.]. Paris: Mme. Huzard (facsimjle<br />

reprint, Puteaux: Erec, 1981), 70-71.<br />

39. Paillot de Montabert, op. cit., 258-71; see also Merimee, op. cit., 144-65.<br />

40. Bomford, D., J. Kirby, J. Leighton, <strong>and</strong> A. Roy. 1990. Art in the Making: Impressionism,<br />

London: National Gallery <strong>and</strong> New Haven: Yale University Press, 117,<br />

119.<br />

41. Mirecourt, op. cit., 24.<br />

42. Ingamells, op. cit., 99-100 (no. P276), 103-104 (no. P300).<br />

43. Knight, V 1986. The Works of Art of the Corporation of London: A Catalogue of<br />

<strong>Painting</strong>s, Watercolours, Drawillgs, Prints alld Sculpture. Cambridge: Woodhead­<br />

Faulkner, 72 (no. 1052).<br />

44. Godde, op. cit. (note 3) , plate 16 <strong>and</strong> conunentary. The engraving measured 29<br />

X 36 cm.<br />

Kirby <strong>and</strong> Roy 175


Abstract<br />

Turner's use of sketches on paper, his<br />

development of successful oil sketches<br />

into finished paintings, his preference<br />

for absorbent primings, <strong>and</strong> his<br />

modified oil media are described.<br />

His oil painting techniques circa<br />

1800-1850 are illustrated by studies<br />

of several works. His use of megilps<br />

(varnish-modified oil media) is outlined,<br />

along with his use of newly<br />

available manufactured pigments, <strong>and</strong><br />

is compared with analyses from other<br />

works painted by British artists<br />

1775-1875, ranging from Reynolds<br />

to Whistler.<br />

<strong>Painting</strong> <strong>Techniques</strong> <strong>and</strong> <strong>Materials</strong> of Turner <strong>and</strong> Other<br />

British Artists 1775-1875<br />

Joyce H. Townsend<br />

Senior Conservation Scientist<br />

Tate Gallery, Millbank<br />

London SW 1P 4RG<br />

United Kingdom<br />

Introduction<br />

This paper presents important aspects of J. M. W Turner's technique, discussed<br />

before by the author in greater detail, by examining several paintings not<br />

previously described in this context (1, 2). Turner painted in oil fo r over fifty<br />

years (ca. 1798-1850), <strong>and</strong> it is interesting to compare his materials to those<br />

used in British paintings of the preceding <strong>and</strong> fo llowing twenty-five years (in<br />

the Tate Gallery collection, unless otherwise stated). Thus, this paper presents<br />

a comparison of the use of modified paint media <strong>and</strong> the adoption by various<br />

artists of new pigments produced between 1775 <strong>and</strong> 1875.<br />

Turner's oil painting techniques, compared to others' techniques<br />

Turner spent at least ten years as a watercolorist before he used oil as a paint<br />

medium, developing a range of techniques that he would utilize ever after in<br />

oil. In the earlier watercolors, transparent washes overlie the white paper<br />

except where Turner reserved highlights. The greens were made by mixing,<br />

overlaying, or physically mixing with his fingers, washes of brown <strong>and</strong> blue.<br />

There is very little underdrawing, <strong>and</strong> generally it is free rather than detailed.<br />

The lTlental image was transferred directly to the support. As Farington wrote,<br />

"Turner has no settled process but drives the colour about till he has expressed<br />

the ideas in his mind" (3). Turner produced over 20,000 pencil sketches <strong>and</strong><br />

watercolors, now at the Tate Gallery, but few have a direct counterpart in oil.<br />

He hardly ever produced a detailed oil sketch, even for a commission, <strong>and</strong><br />

when he sketched in oil he developed the best sketches into completed <strong>and</strong><br />

exhibited works, rather than repainting them on a new canvas.<br />

Turner seems to have tried nearly all materials <strong>and</strong> methods once. The following<br />

descriptions apply to many of his paintings, if not all. Once he began<br />

to work in both media (always independently) , he utilized new application<br />

techniques <strong>and</strong> pigment in both media at about the same time. His early<br />

work in watercolor gave him an underst<strong>and</strong>ing of <strong>and</strong> liking fo r light-toned,<br />

absorbent surfaces, <strong>and</strong> honed his skills in the application of optical greens<br />

<strong>and</strong> blacks. Some eighteenth-century artists such as Wright of Derby (4) used<br />

white grounds to lend luminosity to their oil paintings, but many of Turner's<br />

immediate contemporaries were using thicker paint <strong>and</strong> warm-toned grounds<br />

up to 1820. Constable produced oil sketches upon mid-toned buff, red, or<br />

blue grounds, though his exhibited works do not show quite such a variety<br />

of grounds (5). After 1820, more British artists tended to use white grounds.<br />

Many commercially primed canvases are white during this time <strong>and</strong> up to<br />

the 1890s, when Sargent <strong>and</strong> Whistler used gray ones of varying tones in<br />

Engl<strong>and</strong> (6) . Turner used white primings for a fair proportion of his oils, both<br />

exhibited <strong>and</strong> unfinished, in the first decade of the nineteenth century; in<br />

later decades most of his supports had white grounds. In a fair num.ber of the<br />

paintings examined, the primings consist of lead white in whole egg medium<br />

(7), most of the others being lead white in oil, not sized on the surface or<br />

between applications of priming, as had previously been usual. Absorbent<br />

surfaces gave the impetuous Turner a very rapid indication of the final color<br />

of the paint, <strong>and</strong> allowed him to develop the composition rapidly over fastdrying<br />

paint if it looked promising.<br />

176<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


F(gure 1. J. M. W. Turner, Goring Mill <strong>and</strong> Church, ca. 1806-1807. Oil sketch on canvas,<br />

857 X 1162 111111. Courtesy of the Tate Gallery, London (N02704).<br />

Some works were ab<strong>and</strong>oned at this stage, for example Goring Mill <strong>and</strong> Church<br />

(ca. 1806-1807), in which the buildings were lightly drawn in pencil before<br />

washes were applied of thinned linseed oil paint in green <strong>and</strong> brown for grass,<br />

buildings, <strong>and</strong> cattle, <strong>and</strong> in highly thinned, pale blue paint to suggest the<br />

clouds (Fig. 1). Other ab<strong>and</strong>oned oil paintings from this date <strong>and</strong> later have<br />

little or no pencil underdrawing. Turner used it mainly to outline buildings<br />

or ships, where accuracy mattered. By contrast, l<strong>and</strong>scapes <strong>and</strong> trees were<br />

usually freely painted. In later years, Turner used increasingly brightly colored,<br />

thinned paint as a first lay-in, generally brown for the l<strong>and</strong>scape <strong>and</strong> blue for<br />

the sky, leaving white priming in areas that would later be depicted as yellow<br />

sunlight. He glazed down the colors as he worked. Cross sections from most<br />

oils show thin transparent washes, overlaid by thin paint layers in the same<br />

colors, lightened with lead white. Some ab<strong>and</strong>oned works have a patch of<br />

bright red or blue in the foreground, which Turner would have developed<br />

into something appropriate as the image evolved, perhaps a red buoy in the<br />

sea, or a brightly dressed figure in a l<strong>and</strong>scape.<br />

Turner's earliest oils look thickly <strong>and</strong> conventionally painted at first glance or<br />

when viewed through accumulated yellow varnish, but this is deceptive. Dolbadern<br />

Castle, North Wales is a good example (Fig. 2). Turner presented it to<br />

the Royal Academy, London, when he obtained full membership at the age<br />

of twenty-eight. Recent cleaning revealed thin glazes of Mars orange (i.e.,<br />

strongly colored synthetic iron oxide) <strong>and</strong> localized scumbles of black, <strong>and</strong><br />

Mars orange or red, with white in the l<strong>and</strong>scape. The surface was, in fact,<br />

vulnerable <strong>and</strong> potentially sensitive to solvents. Clouds were applied with a<br />

thicker, creamy-looking paint that retained brush marks, small areas were then<br />

scumbled over with quite bright yellows or pinks, applied rather lean. Naples<br />

yellow, reddish brown ochres, a purplish ochre, <strong>and</strong> ivory black provided these<br />

highlights, while the sky itself was painted in ultramarine, lightened with lead<br />

white. Turner continued to use ultramarine for finishing skies throughout his<br />

life, having often done the initial lay-in with smalt, though not in this case.<br />

He used Prussian blue glazes <strong>and</strong> washes over thicker layers of brown ochres<br />

<strong>and</strong> umbers for the greener parts of the l<strong>and</strong>scape <strong>and</strong> the stream. The fastflowing<br />

water is indicated by small flecks of textured white paint. These may<br />

be lean strokes of paint applied in a pure oil medium. In other paintings, such<br />

areas were found to contain linseed oil but very rarely walnut oil, despite the<br />

fact that the latter was known to yellow less. It is much less likely that the<br />

Townsend 177


Figure 2. J. M. W. Tumer, Dolbadern Castle, North Wales, 1802. Oil 011 canvas, 1190 X 889 /11111. Courtesy of the<br />

Royal AcadelllY of Arts, LOlldo/l.<br />

textured clouds were painted in pure oil medium; they more resemble a<br />

megilp, a medium discussed in detail later.<br />

Many of Turner's oils are disfigured by wide contraction cracks long associated<br />

with natural bitumen or asphaltum. The paint of Dolbadern Castle is free<br />

of such defects, possibly because the subject did not require the extreme<br />

contrasts between sunny <strong>and</strong> shaded l<strong>and</strong>scape that Turner achieved so effec-<br />

178<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figl/re 3. Detail if lower left, s/lOwillg wlltractioll cracks ill vitulllell-rich paillt, jrolll J. M. W. TurlIer's<br />

The Hero of a Hundred Fights, ca. 1800, rel/,orked 1847. Oil 01/ cal1vas, 908 X 1213.<br />

COl/rtesy oj the Tate Gallery, LOlldoll (N00551).<br />

tively elsewhere through the use of bitumen, megilp, <strong>and</strong> resinous glazes.<br />

Bitumen has been detected in his Italian l<strong>and</strong>scapes of 1820-1840 in particular,<br />

<strong>and</strong> appears at its worst when used as an underlayer for the foreground.<br />

In other works, Turner added to bitumen's poor drying characteristics by<br />

painting it directly over old damaged paint or by exposing canvas to extreme<br />

dampness <strong>and</strong> dripping water, so the support stressed the paint <strong>and</strong> caused<br />

cracking. It is not certain whether he ever used synthetic bitumen, a material<br />

severely criticized by his contemporaries, <strong>and</strong> reputed never to dry at all (8) .<br />

Figure 3 shows a detail from Turner's The Hero of a Hundred Fights that has<br />

bitumen-rich paint applied over oil paint a few decades old, showing severe<br />

cracking typical of this artist's oeuvre. Figure 4 shows even more disfiguring<br />

cracking, attributed to synthetic bitumen, in a study of Editha's head fo r<br />

Hilton's Editha <strong>and</strong> the Monks Searching fo r the Body of Harold.<br />

The full complexity of Turner's later paintings can be represented by The<br />

Dawn oj Christianity, The Flight into Egypt, exhibited in 1841. Turner used a<br />

square canvas with a white, absorbent priming of lead white <strong>and</strong> oil, <strong>and</strong><br />

planned the work fo r a non-square frame. He sketched in octagonal cut-offs<br />

in pencil, drew a circle with pencil <strong>and</strong> compasses, <strong>and</strong> then painted a more<br />

or less circular image. He prudently made use of the corners of the canvas<br />

to try out colors, <strong>and</strong> quite large brush loads of paint survive there, uncontaminated<br />

by later layers. Figure 5 shows the framed painting, <strong>and</strong> Plate 35<br />

shows a detail of the lower right corner. The most textured paint, nearly free<br />

of pigment <strong>and</strong> by now honey colored, proved to be a megilp. The canvas is<br />

Townsend 179


Figure 4. Detail (showing cracks) from William Hilton the Younger's study of Editha's head for Editha<br />

<strong>and</strong> the Monks Searching for the Body of Harold, ca. 1834. Oil on canvas, 335 X 244<br />

mm. Courtesy rf the Tate Gallery, London (N00333).<br />

unlined. The paint dribbling down the tacking margins at an angle suggests<br />

Turner used a sloping easel, probably the tripod type depicted in his watercolors.<br />

There are very few descriptions of Turner painting, but observation<br />

of the paintings <strong>and</strong> cross sections makes it clear that he thinned paint excessively,<br />

until it contracted into isl<strong>and</strong>s as it dried (visible in the foreground);<br />

at other times, he mixed paint in drying oil on the palette so rapidly as to<br />

leave recognizable blobs of the added oil. Almost certainly, Turner completed<br />

The Dawn of Christianity at the Royal Academy in the three days required of<br />

other artists of his era fo r retouching sunken areas or appling varnish. The<br />

sky paint was applied rather thickly with a palette knife, as was white impasto<br />

in the lower right. Both paints consisted of good quality lead white with few<br />

impurities. Turner sometimes modified the sky with opaque scumbles, but<br />

rarely glazed it, emphasizing the contrast between the sky <strong>and</strong> the highly<br />

glazed l<strong>and</strong>scape in the foreground. The middle ground was painted rapidly<br />

180<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 5. J. M. W. Tumer's The Dawn of Christianity, the Flight into Egypt, exhibited 1841. Oil on cal1vas, 787 X 787 111m.<br />

Courtesy of Ulster Museum, Belfast, Northern Irel<strong>and</strong>.<br />

with megilped paint, <strong>and</strong> hence has high impasto that did not slump as it<br />

dried. Turner may even have mixed the megilp into the oil paint on the<br />

canvas, rather than on the palette. Beeswax <strong>and</strong> spermaceti wax, added to oil<br />

<strong>and</strong> both found in this painting (9), yielded a more flowing medium used to<br />

good effect in the more distant trees, though the prominent tree on the right<br />

has a more strongly textured trunk painted with megilp. The very thin, later<br />

glazes that ran off the edges <strong>and</strong> soaked into the corners of the absorbent<br />

canvas would have soaked less into medium-rich paint. Turner used emerald<br />

green here (<strong>and</strong> quite frequently by this date) to provide a stronger contrast<br />

with chrome yellows <strong>and</strong> orange than that provided by an optical green.<br />

Curiously, green mixtures of opaque blues <strong>and</strong> yellows are virtually unknown<br />

in his paintings. The original varnish may have survived on this painting in<br />

the hollows of textured paint, where it has entrapped hogs' hairs, as in other<br />

Turner oils in which the varnish may be original. Little can be said conclu-<br />

Townsend 181


sively about the varnishes Turner used. He may have varnished later works<br />

only when they were sold.<br />

Megilps compared to oil medium<br />

There are numerous references in artists' manuals <strong>and</strong> critics' reviews to megilps,<br />

combinations of mastic varnish <strong>and</strong> drying oil that gelled on mixing,<br />

<strong>and</strong> could then be mixed into pure oil paint on the palette (10). Megilps<br />

(thixotropic medium modifiers) had excellent h<strong>and</strong>ling characteristics both<br />

for impasto <strong>and</strong> glazing, but a severe tendency to darken, <strong>and</strong> to cause cracking<br />

whenever varnish was applied. Such materials had been used <strong>and</strong> criticized<br />

at least since Reynolds' time. Recent studies have shown that megilps<br />

can be made successfully from linseed oil cooked with lead acetate or litharge,<br />

cooled, <strong>and</strong> later mixed with mastic spirit varnish (11). These megilps subsequently<br />

show different behavior on aging, <strong>and</strong> Turner's paint is somewhat<br />

closer to lead acetate megilp. Pure megilp samples from the corners of The<br />

Dawn of Christianity certainly contain lead <strong>and</strong> behave when heated like artificially<br />

aged lead acetate megilps made from nineteenth-century recipes.<br />

Megilped white paint from this painting <strong>and</strong> the considerably earlier Dolbadern<br />

Castle has rather a different chemical composition from pure drying oil<br />

(12), being more hydrolyzed <strong>and</strong> less oxidized, as though a drier were present<br />

from the beginning. Documentary evidence indicates that Turner used lead<br />

acetate in copious amounts.<br />

Megilps are two-component materials whose properties vary significantly as<br />

the oil:resin proportion changes from 1:3 to 3:1. Film-forming capability, the<br />

tendency of the megilp to segregate afterwards, its tackiness, <strong>and</strong> its ability to<br />

absorb more or less dust than oil paint all depend on the exact proportions.<br />

Turner, who worked fast <strong>and</strong> fu riously, <strong>and</strong> never even paused to grind his<br />

pigments finely, must have used a variety of fo rmulations <strong>and</strong> proportions of<br />

megilp, albeit unconsciously. This has led to variations in degree of yellowing<br />

<strong>and</strong> solubility in Turner's paintings, <strong>and</strong> made them very sensitive to cleaning.<br />

Megilps, as well as paints with a startling variety of melting <strong>and</strong> softening<br />

points, have been fo und in numerous samples from Reynolds's later paintings<br />

of the 1780s. Reynolds's The Death if Dido, one of the "fancy pictures"<br />

wherein he is said to have used paint media that he never allowed his pupils<br />

to use, is particularly rich in megilplike layers <strong>and</strong> has some very striking<br />

surface defects. They arise when a modified oil layer is applied to fairly pure<br />

oil, whereupon microwrinkles form in the previously stable film, <strong>and</strong> the<br />

surface later exhibits a rough texture with drying cracks cutting in deeply.<br />

Figure 6. Detail if Anna's veil, painted over clouds <strong>and</strong> sky, from Joshua Reynolds's The Death of<br />

Dido, ca. 1781. Oil on cal1vas, 1473 X 2407. Courtesy of The Royal Collection, Her Majesty the<br />

Queen.<br />

182<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


TabLe 1. Dates of manufa cture of inorganic pigments introduced 1775- 1875, with the earliest<br />

instance of their use fo und to date in oil paintings at the Tate Gallery.<br />

Pigment Firat manufacture Earlieat dete Artlat Source<br />

Prussian blue early18th c. by mid-18th c. varioos<br />

Marred mid or later 18th c.? 1755 1760 Reynolds a<br />

Mars yellow mid or later 16th c.? 1781 Reynolds a<br />

Mars brown mid or later 18th c.? 1781 Reynolds<br />

patent yellow published ca.1 ns 1781 1 Reynolds a<br />

Scheele's green used 1775 ca. 1606-1807 Turner a<br />

published 1778<br />

b<br />

cobalt green pub 1780 not found yet a<br />

made mid-19th c.? or 1825<br />

C<br />

barium sulphate used 1782 1840s Turner d<br />

Indian yellow mentioned 1766 1781 Reynolds a<br />

Brunswick green "nawn in 1795 not found yet e<br />

opaque oxide of chromium scld by Reid 1815 not found yet a<br />

cobalt blue 1802 in France ca.1801>-1807 Turner f<br />

chrome yellow 1814 1815 exh. 1814 Turner a<br />

pale lemon chrome posl 1814-1815? ca.1822 1823 Turner<br />

chrome orange post 1814-18151 ca.1822 1823 Turner<br />

emerald green disc. 1814 ca.1828 Turner I<br />

synthetic ultramarine 1826-1827 c8.1851 Turner a<br />

1857 Campbell j<br />

ultramarine green prepared 1828, made 1854-6 not found yet 9<br />

Chinese white 1834 in Engl<strong>and</strong> 1835-1840 Turner a<br />

1852 Hunt<br />

viridian 1830s in France exh. 1842 Turner f<br />

cobalt yellow 1831 in Germany 18731 Whistler a<br />

made 1851. 1860 Eng<br />

d<br />

barium chromate French pat 1840s exh. 1843 Turner h<br />

strontium chromate sold by Reid 1835? Mulready 8<br />

1857 Campbell j<br />

orange vermilion ''new in 1835 exh. 1843 Turner i<br />

chrome scarlet 1840 W&N ca.1851 Turner e<br />

cadmium yellow 1843 1855 Millais j, i<br />

antimony orange patented in Engl<strong>and</strong> in 1847 not found yet c<br />

zinc chromate 1850 not found yet 9<br />

cobalt violet dark mentioned 1859 not found yet e<br />

cadmium orange 1860 Rowney 1872 Whistler e<br />

unramarine red dev 1870 1880 not found yet 9<br />

Sources: (a) Harley, R. D. 1982. Artists' Pigments c11835. A Study of English Documentary Sources.<br />

London: Heinemann-Butterworth ; (b) Reid, G. 1835. Chromatography. London: Tilt; (e) Friedstein, H. 1981.<br />

Journal of Chemical Education (58):291-95; (d) Artists' Pigments. A H<strong>and</strong>lx>ok of their History <strong>and</strong> Characteristics.<br />

1986. Ed., A. L Feller. Washington: National Gallery of Art; (8) l. Carlyle, personal communication; (1) Bomford, D.,<br />

J. Kirby, J. Leighton, <strong>and</strong> A. Roy. 1990. Art in the Ma king:Impressionism. London: National Gallery; (g) Gattens, R.<br />

J., <strong>and</strong> G. l. Stout. 1966. <strong>Painting</strong> <strong>Materials</strong>: A Short Encyclopedia. New York: Dover Publications; (h) D.<br />

Saunders <strong>and</strong> J. Kirby, personal communication ; (i) Harley, A. D. 1987. Some new watercolours in the nineteenth<br />

century. The Conservator (11):46-50; m E. Sheldon, personal communication. <strong>Painting</strong>s are privately owned.<br />

Figure 6 shows a detail of this painting, which was lined early in its history.<br />

Turner's techniques were similar, in that he would apparently apply any paint<br />

medium over any other to gain a beautiful but short-lived visual effect.<br />

Some of Turner's imitators achieved effects that today appear similar to his,<br />

but without the shrinkage <strong>and</strong> disruption of the paint. Turner's paintings may<br />

have changed greatly with time, as the critic John Ruskin thought. Several<br />

of Callcott's paintings have numerous thin glaze layers <strong>and</strong> less wet working<br />

than Turner used. Callcott, Etty, <strong>and</strong> other contemporaries of Turner did not<br />

have the patience to wait until previous paint had dried before they added<br />

another layer, but they did use varnish interlayers so that later paint could be<br />

applied safely. The Pre-Raphaelites favored a disciplined approach, too, as is<br />

well known. Hunt's Strayed Sheep (Our English Coasts) has very detailed, localized<br />

layers applied to already dried paint throughout the foreground, <strong>and</strong>,<br />

like works by Mulready <strong>and</strong> Collins, has a very well-preserved surface today<br />

(Plate 36).<br />

The other durable method is to use such slow-drying oil paint that wet<br />

working is possible, or to thin the paint so it fo rms a single layer. This was<br />

Whistler's method in the early 1870s, when he was producing nocturnes <strong>and</strong><br />

harmonies (13). Instead of allowing the paint layer to grow thick, he scraped<br />

it back vigorously <strong>and</strong> began again, sometimes using paint so wet that the<br />

canvas had to be laid flat until it dried.<br />

British artists' adoption of new pigments, 1775-1875<br />

Table 1 summarizes dates of invention, first publication, <strong>and</strong> so on, of pigments<br />

that became available between the eighteenth century <strong>and</strong> 1875, <strong>and</strong><br />

notes the earliest occurrence yet found in oil paintings at the Tate Gallery.<br />

Townsend 183


Reynolds was known to experiment with paint media, <strong>and</strong> he might be<br />

supposed to have tried out new pigments too. Five of his works have been<br />

examined, including three at the Tate. Not surprisingly, all five include traditional<br />

pigments such as lead white, ivory black, asphaltum, vermilion, <strong>and</strong><br />

Naples yellow, while one or more included red lead, orpiment, blue verditer,<br />

smalt, ultramarine, <strong>and</strong> green earth (14). Prussian blue, invented early in the<br />

eighteenth century, was used frequently by Reynolds. Some of this Prussian<br />

blue has a different microscopical appearance from the modern variety; similar<br />

material, previously illustrated in color by Welsh (15), was identified in Turner's<br />

paintings up to about 1840, when Turner began using the "modern"<br />

variety as well. George Jones used the modern variety in 1832. Reynolds also<br />

used Mars red from 1755-1760, <strong>and</strong> Mars orange, red, <strong>and</strong> brown in 1781.<br />

He used Indian yellow in a work dated to 1788, <strong>and</strong> wrote about a material<br />

that may have been Indian yellow in 1784, two years before it has been noted<br />

in the literature (16). Patent yellow has also been tentatively identified in a<br />

painting of 1781. Table 1 shows that all these instances demonstrate early uses<br />

of these pigments. Reynolds' organic pigments (red, blue, <strong>and</strong> a green made<br />

from yellow <strong>and</strong> blue dyes) are still being investigated.<br />

Turner's early use (1800-1850) of new pigments in oil, summarized in Table<br />

1, has been described elsewhere (17). While it is true that more paintings by<br />

Turner than other individual artists have been analyzed, the inference that he<br />

was more innovative than his contemporaries is inescapable. Turner used Mars<br />

colors frequently, as did many of his fellow British artists, including Constable<br />

from circa 1810 (18). In contrast, Arnald, Farington, Hilton, <strong>and</strong> Callcott used<br />

only well-established pigments such as ultramarine, Prussian blue, Naples yellow,<br />

<strong>and</strong> vermilion. Artists who used new pigments quite soon after their<br />

introduction include the following: Constable-cobalt blue in 1817-1818,<br />

chrome yellow in 1816, <strong>and</strong> opaque oxide of chromium in 1837 or earlier;<br />

Briggs-chrome yellow <strong>and</strong> orange in 1826; <strong>and</strong> Mulready-emerald green<br />

in 1842 (19). Barium chromate <strong>and</strong> a pigment tentatively identified as strontium<br />

yellow were found in a Mulready of 1835. Cadmium yellow has been<br />

fo und in a Millais of 1855, <strong>and</strong> strontium yellow in a Campbell of 1857 (20).<br />

The latter included synthetic ultramarine, rarely used before 1850 except by<br />

Turner, because it had a poor reputation (21). Whistler used two shades of<br />

cadmium yellow regularly from 1864 (the earliest of his works at the Tate),<br />

<strong>and</strong> strontium yellow in two works circa 1864-1871 <strong>and</strong> in 1872, respectively<br />

(22) . The earliest tentative identification of cobalt yellow (mixed with barium<br />

chromate) is in one of Whistler's oils from the fo llowing year. No examples<br />

have yet been found of the cobalt violet shades that were available by the<br />

end of this period or by 1900.<br />

White pigments <strong>and</strong> their fillers are also interesting. Artists from the time of<br />

Reynolds <strong>and</strong> Romney tended to use lead white with pipe clay or china clay<br />

extenders, both for priming <strong>and</strong> paint. Gypsum has been found in many paint<br />

samples from Reynolds <strong>and</strong> Wright of Derby in the later eighteenth century<br />

(23). Zinc white was very rarely found in Turner's oils, <strong>and</strong> not yet in his<br />

fellow Royal Academicians works painted before 1847. But Hunt used it in<br />

1852, with lead white for a local imprimatura under the sky of Strayed Sheep<br />

(Our English Coasts) <strong>and</strong> with Prussian blue, presumably supplied as a tube of<br />

paler "Antwerp blue." Barium, attributable to barium sulfate, has been fo und<br />

as an extender in Turner's paintings of the 1840s, in the Hunt piece, <strong>and</strong> very<br />

frequently in Whistler's white paint from 1864-1 875, <strong>and</strong> beyond.<br />

A discussion of pigments that fell out of use during this period would be out<br />

of place here, but work is continuing in this area (24). Analyses of the paint<br />

media used by the artists mentioned here also continue, but as yet there are<br />

insufficient results for comparisons between Turner <strong>and</strong> many of his contemporanes.<br />

Acknowledgments<br />

The Turner research was fu nded by the Leverhulme Trust <strong>and</strong> supported by the Tate<br />

Gallery, through Stephen Hackney. My thanks to the following people: Brian<br />

184<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Kennedy of the Ulster Museum, Viola Pemberton-Piggott of the Royal Collection<br />

<strong>and</strong> Helen Valentine of the Royal Academy, who arranged the examination of the<br />

Turners in their collections, Anna Southall <strong>and</strong> Gillian Osmond, who pointed out<br />

interesting nineteenth-century pigment samples, <strong>and</strong> Stephen Hackney, Tom Learner,<br />

Rica Jones, <strong>and</strong> Anna Southall for editing.<br />

Notes<br />

1. Townsend, ]. 1993. Turner's <strong>Painting</strong> <strong>Techniques</strong>. London: Tate Gallery. Information<br />

on Turner is published here, except where stated otherwise.<br />

2. Townsend,]. H. n.d. Turner's use of materials, <strong>and</strong> implications for conservation.<br />

In Turner's <strong>Painting</strong>s <strong>Techniques</strong> in Context. Ed. ]. H. Townsend, et al. London:<br />

UKIC. Forthcoming.<br />

3. The Diary ofjoseph Farington. 1978. Eds. K. Garlick <strong>and</strong> A. Maclntyre. London:<br />

Yale University Press, Vol. I (1799), 273.<br />

4. Jones, R. 1991. Notes for conservators on Wright of Derby's painting techniques.<br />

The Conservator (15):13-32.<br />

5. Cove, S. 1991. Constable's oil painting materials <strong>and</strong> techniques. In Constable.<br />

Eds. L. Parris <strong>and</strong> 1. Fleming-Williams. London: Tate Gallery, 493-529.<br />

6. These Whistler canvases are in the Birnie Philip Gift, Hunterian Art Gallery,<br />

University of Glasgow.<br />

7. Townsend,]. H. 1994. The materials of]. M. W. Turner, 2: primings <strong>and</strong> supports.<br />

Studies in Conservation (39):145-53.<br />

8. Carlyle, L. A., <strong>and</strong> A. Southall. 1993. No short mechanic road to fame: the<br />

implications of certain artists' materials for the durability of British painting:<br />

1770-1840. In Robert Vernon's Gift: British Artfor the Nation 1847. London: Tate<br />

Gallery, 21-26.<br />

9. Townsend, n.d., op. cit.<br />

10. Carlyle <strong>and</strong> Southall, op. cit.<br />

11. Townsend, n.d., op. cit.<br />

12. Rainford, D., personal communication.<br />

13. Hackney, S. 1994. Colour <strong>and</strong> tone in Whistler's nocturnes <strong>and</strong> harmonies 1871-<br />

1872. The Burlington Magazine CXXXVI (10):695-99.<br />

14. Some identifications were made by Rica Jones, Tate Gallery.<br />

15. Welsh, F. S. 1988. Particle characteristics of Prussian blue in an historical oil paint.<br />

journal of the American Institute fo r Conservation (27):55-63.<br />

16. Dubois, H. 1992. Aspects if Sir joshua Reynolds' <strong>Painting</strong> Technique. Unpublished<br />

typescript. Cambridge: Hamilton Kerr Institute.<br />

17. Townsend, n.d. , op. cit.<br />

18. Cove, op. cit.<br />

19. Ibid.<br />

20. Sheldon, E., personal communication. <strong>Painting</strong>s are privately owned.<br />

21. Townsend,]. H. 1993. The materials of]. M. W Turner, 1: Pigments. Studies in<br />

Conservation (38) :231-53.<br />

22. See note 6.<br />

23. Jones, op. cit.<br />

24. Townsend, ]. H., L. Carlyle, S. Woodcock, <strong>and</strong> N. Kh<strong>and</strong>ekar. Unpublished research.<br />

Townsend 185


Abstract<br />

Several of Whistler's paintings were<br />

examjned <strong>and</strong>, where possible, analyzed;<br />

specific examples of pigment<br />

<strong>and</strong> media analyzes are given. Literature<br />

sources were searched fo r references<br />

to the artist's materials <strong>and</strong><br />

methods. This paper describes Whistler's<br />

interest in texture, his use of<br />

dark gray grounds, his limited color<br />

ranges <strong>and</strong> careful preparation of the<br />

palette, his frequent erasure of his<br />

unsuccessful work, <strong>and</strong> details of his<br />

studio practice <strong>and</strong> portrait techniques.<br />

Probable changes in the<br />

appearance of his work are also discussed.<br />

Art for Art's Sake: The <strong>Materials</strong> <strong>and</strong> <strong>Techniques</strong> of<br />

James McNeill Whistler (1834-1903)<br />

Stephen Hackney<br />

Conservation Department<br />

Tate Gallery<br />

Millbank<br />

London SW 1P 4RG<br />

United Kingdom<br />

Introduction<br />

Several of Whistler's oil paintings were studied in detail; works in the Tate<br />

Gallery, in particular, were analyzed <strong>and</strong> examined before conservation treatment<br />

using a fu ll range of analytical techniques. <strong>Painting</strong>s in the H unterian<br />

Art Gallery, Glasgow from the Birnie Philip bequest were also examined in<br />

detail <strong>and</strong> in some cases analyzed (1). <strong>Painting</strong>s in the National Gallery of<br />

Art, Washington were examined in conjunction with their conservation records.<br />

Several other works were examined or viewed in isolation; finally, the<br />

paintings in the Whistler Exhibition (1994) were inspected on arrival at the<br />

Tate Gallery (2). It has therefore been possible to make some interesting<br />

observations about Whistler's materials <strong>and</strong> techniques, based on detailed examinations<br />

of selected works <strong>and</strong> supported by the wider survey, <strong>and</strong> to compare<br />

them with the literature on Whistler's methods. This interpretation may<br />

be useful in underst<strong>and</strong>ing the present condition of his works <strong>and</strong> in helping<br />

to decide appropriate conservation treatments. It is also of general interest for<br />

the viewer wishing to appreciate a particular painting.<br />

Supports<br />

From the 1880s, Whistler frequently carried small panels for sketching, focusing<br />

particularly on street scenes, views of the sea, <strong>and</strong> figure studies. Some<br />

of his nocturnes <strong>and</strong> early studies were also on larger panels. The scale of<br />

these works was small because he painted distant views approximately the<br />

size that they appeared to him. For his major finished paintings <strong>and</strong> portraits,<br />

however, he worked on canvas, which provided a texture he liked. He often<br />

chose quite heavy canvases <strong>and</strong> applied thin grounds in order to preserve<br />

their texture. On other occasions he might paint on a fine canvas, then have<br />

the painting glue-lined onto a coarser one very soon after completion or<br />

even during the time of painting. He did this quite deliberately <strong>and</strong> there is<br />

evidence that he wished to imitate the lined appearance of old master paintings<br />

(3). His later work, such as Mother of Pearl <strong>and</strong> Silver: The Andalusian<br />

(1888-1900), illustrates this desire to express the canvas texture.<br />

Grounds<br />

Feurc 1. James Whistler, Note in Red:<br />

The Siesta, 1883- 1884. Oil 011 palle/.<br />

Daniel j. Terra Collcetioll.<br />

Whistler's grounds were crucial to his methods. He came to London shortly<br />

after the Pre-Raphaelites had taken the use of smooth white grounds <strong>and</strong><br />

intense color to extremes; some of his early grounds appear to be the white,<br />

commercially available grounds of the mid-century. Yet Whistler's training in<br />

Paris with Charles Gleyre (an academician who specialized in subjects taken<br />

from his travels in the Middle East), his early interest in the methods of the<br />

French Realists, <strong>and</strong> his experience with pastel, etching, <strong>and</strong> drypoint were<br />

more important influences. Whistler frequently applied a light gray imprimatura<br />

of oil paint on top of the ground to allow him to paint directly in a<br />

mid-ground technique. This technique is most easily observed in his small<br />

sketches on panel, for example, Note in Red: The Siesta (1883-884) (Fig. 1).<br />

After 1871 he increasingly used darker gray, exploiting the ground to develop<br />

full chiaroscuro effects in his later works, often setting his figure against a<br />

dark background. A photograph of his studio reveals that he had a black cloth<br />

186<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, alld <strong>Studio</strong> <strong>Practice</strong>


Figure 2. James Whistler, Nocturne in Blue <strong>and</strong> Silver: Cremorne Lights, 1872. Oil on canvas.<br />

The Tate Gallery, Londol1.<br />

to drape over an easel behind the sitter. Dark grounds were first successfully<br />

exploited by Whistler to produce his nocturnes of the 1870s. Similarly, fo r<br />

his watercolors, designs, <strong>and</strong> etchings, Whistler also sought old mid-toned or<br />

dark papers to form the basis of his images.<br />

To make dark grounds, he mixed ivory black <strong>and</strong> lead white, frequently modified<br />

with other pigments. In his Chelsea studio, the Greaves brothers helped<br />

him prepare his materials. Walter Greaves describes his ground as a very absorbent<br />

distemper, indicating that at this stage Whistler was not only modifYing<br />

commercial grounds but also preparing his own (4). The production of<br />

nocturnes was a period of experimentation. Nocturne in Blue <strong>and</strong> Silver: Cremarne<br />

Lights (1872) (Fig. 2) is on a partially scraped down, reused canvas from<br />

an earlier series with an unmodified, lighter commercial ground, whereas<br />

Nocturne in Blue Green: Chelsea (1871) has a dark ground applied over a white<br />

commercial priming on a mahogany panel. Nocturne in Black <strong>and</strong> Gold: The<br />

Fire Wheel (1875) has no commercial ground, only an artist's priming over<br />

the sized canvas. This ground consists of ivory black, chrome or cadmium<br />

yellow, <strong>and</strong> small amounts of lead white.<br />

Palette<br />

Figure 3. Brushes, paints, palettes, <strong>and</strong> other<br />

painting materials. Courtesy of the Hunterian<br />

Art Gallery, University of Glasgow, Birnie<br />

Philip Bequest.<br />

Several of Whistler's palettes survive, along with his brushes, charcoal, engraving<br />

tools <strong>and</strong> tube paints, presumably dating from the last period ofWhistler's<br />

life before they were donated to the Hunterian Art Gallery (Fig. 3).<br />

These have been examined <strong>and</strong> analyzed during the present study (5). For a<br />

short time (1898-1901), Whistler taught at the Academie Carmen. Reports<br />

by students of his methods demonstrate Whistler's concern fo r the preparation<br />

<strong>and</strong> layout of his palette. Colors were laid out in a specific order across the<br />

top of the palette from left to right: Prussian blue, cobalt blue, raw umber,<br />

burnt sienna, raw sienna, yellow ochre, a large blob of lead white, vermilion,<br />

Venetian red, Indian red, <strong>and</strong> black. Flesh tones were mixed just below the<br />

white, using the appropriate surrounding colors, <strong>and</strong> in turn these tones were<br />

modified by the black which was spread in a broad b<strong>and</strong> curving downward.<br />

A preparation for the background color was mixed at the left. He then<br />

worked out all the colors <strong>and</strong> tones for his composition on his palette before<br />

placing any paint on his canvases. So obsessed was he with this "scientific"<br />

method that he would frequently examine his students' palettes yet ignore<br />

their paintings, wishing (he claimed) not to interfere with their individuality<br />

<strong>and</strong> free expression (6) .<br />

Hackney 187


Although this is a description of Whistler's late work, the artist acknowledges<br />

that he learned his approach from Gleyre <strong>and</strong> therefore would have employed<br />

similar methods throughout his career, with either a wider range of pigments<br />

or less mixing of pigments in his earlier work. Menpes confirms the attention<br />

to the palette but attributes to him the use of more intense lemon <strong>and</strong> cadmium<br />

yellows placed next to the ochre, <strong>and</strong> the use of rose madder instead<br />

of the Venetian red (7).<br />

Whistler kept a great number of brushes, many of which he used at a single<br />

sitting in order to prevent his dominant hues from becoming fu rther mixed.<br />

He spent much time cleaning <strong>and</strong> preparing these brushes, sometimes trimming<br />

<strong>and</strong> changing their shape. He had various types, including decorator's<br />

brushes to lay in grounds <strong>and</strong> backgrounds <strong>and</strong> extremely long-h<strong>and</strong>led<br />

brushes for his portrait technique.<br />

Pigments analyzed<br />

Figure 4. James Whistler, Harmony in<br />

Grey <strong>and</strong> Green: Miss Cicely Alex<strong>and</strong>er,<br />

1872- 1874. Oil on canvas. The Tate Gallery,<br />

London.<br />

Analysis of pigments fo und on Whistler's paintings confirm the use of complex<br />

mixtures of pigments usually involving at least some ivory black.<br />

Frequently, the same pigments are found mixed in different proportions<br />

throughout a painting. In nocturnes, mixtures of lead white, ivory black, cobalt,<br />

Prussian blue, <strong>and</strong> ultramarine have been found. By these means, Whistler<br />

produced his harmonies. For instance, Harmony in Grey <strong>and</strong> Green: Miss Cicely<br />

Alex<strong>and</strong>er (1872) has a gray background, but this is an optical mixture of<br />

several colors based on lead white <strong>and</strong> ivory black (Fig. 4) . The same colors<br />

occur in the superficially white dress. By restricting the saturation <strong>and</strong> hue<br />

of his colors throughout the scheme, Whistler could model his fo rms without<br />

introducing discordant hues typical of the works of many nineteenth-century<br />

painters influenced by the Pre-Raphaelites <strong>and</strong> exploiting a range of unmixed<br />

pigments. This restriction allowed him to explore subtle color effects; fo r<br />

example, he could choose to highlight a red as in the Arrangement in Flesh<br />

Color <strong>and</strong> Black: Portrait


5). There are many such examples of scraped down paint that have not been<br />

entirely painted over (i.e., Cicely Alex<strong>and</strong>er) <strong>and</strong> there must be even more that<br />

are no longer visible. Although tiresome for the sitter, this approach was<br />

essential to the artist. The harmony would not survive if the pigments were<br />

not all mixed in a single operation on the same palette. His paint is essentially<br />

opaque with little or no deliberate glazing. Discolored glazes would also destroy<br />

the harmony. Therefore, as a rule, all his transparent pigments were<br />

mixed with opaque.<br />

A further constraint in the studio is described by Walter Sickert (8):<br />

Figure 5. Detail from Cicely Alex<strong>and</strong>er,<br />

showing areas scraped down to the canvas.<br />

Whistler was a great portrait painter, <strong>and</strong> prided himself on the precision<br />

oj his portraiture. Whistler also had extensive knowledge, <strong>and</strong> knew that<br />

the eye could only see at a glance an object which in size is one-third oj<br />

the distance between the eye <strong>and</strong> that object. In other words, if you are<br />

painting a man six Jeet high you should be 18 Jeet away Jrom him.<br />

Whistler had a very long studio, <strong>and</strong> he was accustomed to place his model<br />

against a black velvet background, <strong>and</strong> alongside his model he placed his<br />

canvas. His painting table was 18 Jeet away. He would st<strong>and</strong> at the<br />

painting table, carifully survey the model, then charging his brush with the<br />

requisite pigment he ran at Ju ll tilt up to the canvas <strong>and</strong> dropped it on the<br />

spot.<br />

This unlikely scenario is consistent with the small scale of much of his work<br />

<strong>and</strong> the flatness of the space around <strong>and</strong> behind his models.<br />

In his nocturnes, Whistler's paint was extremely dilute. He mixed his oil paint<br />

with large amounts of turpentine <strong>and</strong> also added mastic varnish to produce<br />

a paint that could be brushed freely <strong>and</strong> did not dry too matte. He called it<br />

his "sauce" <strong>and</strong> analyses at the Institute for Atomic <strong>and</strong> Molecular Physics in<br />

Amsterdam confirm its formulation from mastic <strong>and</strong> a drying oil. In his later<br />

work he may have used newly available petroleum oil, as recommended to<br />

him by Sickert in 1885. In portraits such as Arrangement in Yellow <strong>and</strong> Grey:<br />

Effi e Deans (1876-1878), he has allowed paint to run down the canvas, indicating<br />

just how diluted it was. His nocturnes were painted quite quickly<br />

(in a day) with minor modifications the next day, as Whistler freely admitted<br />

in his libel trial with Ruskin. The dark passages were first brushed in umber<br />

<strong>and</strong> black, or were simply the part of the imprimatura left exposed. Then a<br />

layer of "sauce" was applied using a large brush to scumble over the shadows<br />

<strong>and</strong> develop the lighter parts. The color <strong>and</strong> tone were controlled mainly by<br />

the thickness of application. The paint was worked wet-in-wet, scraped,<br />

rubbed, <strong>and</strong> even dragged across the wet surface, as in the reflections in Nocturne:<br />

Blue <strong>and</strong> Silver (1872) (Plate 38). Finally, details such as highlights were<br />

applied, after which the nocturne was put outside to dry in the sun. When<br />

it was dry, perhaps the next day, fu rther details could be applied, preferably<br />

from the same palette, but reworking was not possible at this stage.<br />

Condition <strong>and</strong> changes with time<br />

Changes in appearance have occurred on many of his works <strong>and</strong>, despite his<br />

frequent scraping down, minor adjustments to compositions-such as the<br />

positions of arms <strong>and</strong> feet-are now evident. Most significantly, many works<br />

have become darker <strong>and</strong> cooler as the thin paint has increased in transparency<br />

over the dark ground. Similarly, the small sketches with medium-toned<br />

grounds have lost contrast in the mid-tones. His earlier, more decoratively<br />

colorful work on light grounds or thickly painted work, although flawed in<br />

detail, has preserved its appearance much better <strong>and</strong> is more readily appreciated.<br />

The darkness of much of Whistler's later work is due, in part, to<br />

changes resulting from dark grounds <strong>and</strong> also possibly from the darkening of<br />

medium, but to some extent the effect was intended.<br />

Whistler appears to have always varnished his work <strong>and</strong> now many of his<br />

varnishes are excessively darkened <strong>and</strong> yellowed. In particular the discolora-<br />

Hackney 189


tion of varnish has dramatically affected the cool tones of nocturnes <strong>and</strong> some<br />

portrait backgrounds.<br />

Acknowledgments<br />

This research has been a collaboration between the author <strong>and</strong> Dr. Joyce Townsend,<br />

who also carried out all the pigment analysis <strong>and</strong> investigated the media. Medium<br />

analysis was also carried out by the Institute fo r Atomic <strong>and</strong> Molecular Physics,<br />

Amsterdam. The author expresses his gratitude to Prof. Jaap Boon <strong>and</strong> Jos Pureveen<br />

fo r this work. Appreciation is also extended to the Hunterian Art Gallery, Glasgow,<br />

<strong>and</strong> to the National Gallery of Art, Wa shington, fo r allowing access to their collections;<br />

<strong>and</strong> to colleagues at the Tate GaLlery.<br />

Notes<br />

1. McLaren Young, A., M. F MacDonald, <strong>and</strong> R. Spencer, with the assistance of<br />

A. Miles. 1990. The Pailltings oj james McNeill Whistler. New Haven <strong>and</strong> London.<br />

2. Dorment, R., <strong>and</strong> M. MacDonald. 1994. james McNeill Whistler. Exhibition catalogue.<br />

London: Tate Gallery.<br />

3. Hackney, S. 1994. Colour <strong>and</strong> tone in Whistler's nocturnes <strong>and</strong> harmonies. TIle<br />

Burlington Magazine. Vol. CXXXVI, (1099): 695-99.<br />

4. Pennell, E. R.., <strong>and</strong> J. Pennell. 1911. The Life oj james McNeill Whistler. 5th rev. ed.<br />

London: Heinemann, 113-17.<br />

5. Townsend, J. H. 1994. Whistler's oil painting materials. The Burlington Magazine.<br />

Vol. CXXXVI (1099):690-95.<br />

6. Pennell, op. cit., 375-78.<br />

7. Menpes, M. 1904. Whistler as I Knew Him. London: Adam <strong>and</strong> Charles Black, 69-<br />

71.<br />

8. Sickert, W. 1930. Walter Sickert's class: methods of instruction. Manchester Guardian,<br />

31 November.<br />

190<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Abstract<br />

This paper discusses aspects of paintings<br />

executed on a photographic<br />

substrate. Inspired by Mervyn Ruggles's<br />

research into the use of this<br />

practice in the nineteenth century,<br />

this discussion presents examples of<br />

materials <strong>and</strong> techniques that have<br />

been proposed, manufactured, or<br />

employed fo r this purpose over the<br />

last 100 years. Information on this<br />

subject was gathered from the available<br />

literature <strong>and</strong> through personal<br />

communication with artists engaged<br />

in producing this type of artwork, as<br />

well as through discussions with<br />

manufacturers <strong>and</strong> museum personnel.<br />

North American <strong>and</strong> European<br />

instances are included, <strong>and</strong> details regarding<br />

the works of contemporary<br />

artists, such as Lynton Wells, Shirley<br />

Wiitasalo, Anselm Kiefer, <strong>and</strong> James<br />

Turrell, are noted.<br />

<strong>Painting</strong> on a Photographic Substrate: Notes Regarding<br />

<strong>Materials</strong> <strong>and</strong> <strong>Techniques</strong> over the Past 100 Ye ars<br />

John R. Gayer<br />

Koivikkotie 3 I 76<br />

00630 Helsinki<br />

Finl<strong>and</strong><br />

Introduction<br />

The idea of painting over photographic images has been present since the<br />

invention of photography. Sources regarding materials <strong>and</strong> techniques appropriate<br />

to this process have varied; some information has appeared in scientific<br />

<strong>and</strong> photographic publications. Communication with a number of contemporary<br />

artists, though, has shown that such literature rarely directly influenced<br />

their work. The principal factors guiding their production tended to consist<br />

of a willingness to experiment, a general awareness that such media may be<br />

combined, <strong>and</strong> the fact that this combination may serve their particular aesthetic<br />

or intellectual aims. The photographic image has also not only served<br />

as a replacement fo r the underdrawing, as was noted by Mervyn Ruggles in<br />

cases of nineteenth-century portraiture, but has been employed for a variety<br />

of purposes (1).<br />

The following presents a brief discussion outlining the evolution of materials<br />

<strong>and</strong> techniques over the last 100 years. The discussion culminates in a focus<br />

on contemporary practice since the 1960s, a period in which the painter's<br />

use of a photographic substrate became more widespread.<br />

From the late nineteenth century to 1950<br />

In the late nineteenth century, improvements in photographic technology <strong>and</strong><br />

the availability of commercial products made it far easier for the artist to<br />

utilize such means in the painting studio. M. L. Winter of Vienna, for example,<br />

established an operation in 1877 fo r the extensive production of enlarged<br />

photographs on linen, <strong>and</strong> proprietary br<strong>and</strong>s of gelatin-silver emulsion-coated<br />

linen were available from the 1890s (2, 3) . The introduction of faster<br />

bromide emulsions in the 1880s also greatly reduced the difficulties associated<br />

with producing enlargements, thus de-emphasizing the need for specialists<br />

(4).<br />

Oils were frequently used for painting on photographic images, but other<br />

materials were also proposed. Instructions published during the 1890s <strong>and</strong><br />

thereafter commonly referred to the coloring of photographs on paper supports<br />

<strong>and</strong> were directed at those lacking artistic skills. Details regarding the<br />

media proposed for this purpose have been included here because this information<br />

provides a more complete range of the materials that may be encountered,<br />

as well as indicating the interests <strong>and</strong> concerns of the time.<br />

A number of proposals fo r the coloring of photographs appeared in Scientific<br />

American during the 1890s. An 1894 article, reprinted from Anthony's Bulletin,<br />

recommended the use of transparent <strong>and</strong> covering colors. These colors were<br />

to be made by mixing dry powdered pigments with a medium consisting of<br />

100 cc filtered albumin, 5 g ammonium carbonate, 3 cc glycerin, 4 cc liquid<br />

ammonia, <strong>and</strong> 25 cc water (5). Another article, deriving from Photographisches<br />

Archiv, noted the use of aniline dyes. These were dissolved in alcohol <strong>and</strong><br />

applied on the reverse (6) . Another article noted the use of oils, watercolors,<br />

<strong>and</strong> pastels. These materials were only to be applied over a preparatory layer.<br />

Gelatin was recommended for oils; shellac was recommended fo r watercolors<br />

<strong>and</strong> pastels (7).<br />

A proposal, reprinted from the British Journal of Photography, for the use of<br />

wax media appeared in a 1919 issue of Scientific American Supplement (8). Two<br />

Gayer 191


ecipes were described. The first recipe suggested combining 1 oz. white wax<br />

(bleached beeswax) , 1 oz. carbon tetrachloride, 1 oz. turpentine, 1 oz. benzene<br />

(refined naphtha) <strong>and</strong> 1 dr (118 oz.) 0.88 ammonia. The second medium was<br />

described as an improved version consisting of 114 oz. white wax (Cera Alba),<br />

112 oz. spike oil of lavender, 1 dr hard primrose soap, 2 dr gum elemi <strong>and</strong> 3<br />

112 oz. turpentine. The author proposed using wax, based on the fact that<br />

wax, employed since ancient times, had proven its durability. The author's<br />

rationale was also based on the notion that many of the negative effects<br />

resulting from the use of oils <strong>and</strong> varnishes, such as darkening <strong>and</strong> yellowing,<br />

would be avoided.<br />

A 1934 edition of a manual describing the use of various media for coloring<br />

prints on paper noted that enlargements were to be mounted on beaver board,<br />

three-ply board, or linen canvas, <strong>and</strong> then stretched (9). Adhesive residue <strong>and</strong><br />

dirt were to be removed before the application of a preparatory layer. This<br />

layer was to consist of either a 25% solution of glacial acetic acid, gelatin, <strong>and</strong><br />

Lepage's Liquid Glue in water (1:20), or a mixture consisting of 3 oz. of paste<br />

(made from 1 oz. pure casein, 180 g powdered borax, 3 oz. water), 3 oz.<br />

alcohol, S drops glycerin, <strong>and</strong> 5 drops carbolic acid. The image was then<br />

ready to be painted with oils, watercolors, or a medium consisting of tempera<br />

colors <strong>and</strong> the casein mixture.<br />

The 1936 edition of a manual on retouching stated that the best results for<br />

a portrait in oil were to be obtained by painting over a carbon print on<br />

canvas. A priming, consisting of a starch solution to which some mucilage<br />

had been added, was to be applied to the print before it could be painted.<br />

Reference was also made to the "Russian method" of coloring, in which<br />

layers of transparent oils were rubbed on with cotton (10). <strong>Materials</strong> recommended<br />

for this process included Marshall's Transparent Photo Oil Colors,<br />

Roehrig's Photo Oil Colours, or a combination of artists' oils <strong>and</strong> megilp.<br />

The 1930s were also a period in which photographic developments exp<strong>and</strong>ed<br />

into an architectural context. Articles in architecture journals noted that photographic<br />

murals could be realized through the use of photo-mural paper or<br />

the spray application of light sensitive emulsions directly onto architectural<br />

surfaces (11, 12). The notion that a variety of materials could be employed<br />

as supports fo r photographic images was also reaffirmed by new developments<br />

in industry. The Glenn L. Martin Company of Baltimore, for example, reduced<br />

aircraft production time by developing an emulsion that enabled the<br />

full-scale reproduction of designs on aircraft materials (13). A commercial<br />

form of this emulsion was marketed following World War II (14). Articles<br />

outlining the use of such materials frequently noted the need for preparatory<br />

<strong>and</strong> protective layers, should the images be colored with oils or other media<br />

(15, 16).<br />

Mter 1950<br />

At the middle of this century, painters' attitudes toward the photographic<br />

image began to change. Manifestations of this attitude change included the<br />

Photo-Realist movement <strong>and</strong> Robert Rauschenberg's <strong>and</strong> Andy Warhol's use<br />

of photomechanical processes. The American artist James Couper, who has<br />

made use of 3M's Scanamural process (a computerized spray painting technique)<br />

to generate large-scale underdrawings fo r his paintings, has stated that<br />

he was drawn to this technique after learning how some Photo-Realist painters<br />

transferred images to canvas through an emulsion process (17). Studies<br />

published in the United States on the work of such painters, though, have<br />

contradicted this view. Patton, for example, has held that the intention of<br />

these painters was to comment on photography, not include it, <strong>and</strong> interviews<br />

conducted with a number of these artists have also not confirmed the use of<br />

any light sensitive coatings (18, 19).<br />

This change in attitude toward the photographic image was also affected by<br />

a new interest in early <strong>and</strong> alternative processes that occurred during the<br />

192<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


1960s in North America. Requiring both a h<strong>and</strong>s-on approach <strong>and</strong> minimal<br />

resources, these processes appealed to those who felt that the elements of<br />

urban life, such as high technology <strong>and</strong> mass-produced goods, were worthy<br />

of rejection. Interest in these processes became widespread <strong>and</strong> resulted in a<br />

number of publications that provided information on various techniques (20,<br />

21, 22, 23).<br />

Commercial products, such as emulsion-coated canvas <strong>and</strong> ready-to-use<br />

emulsion, became more widely available at about the same time. Argenta,<br />

which operated in Munich, Germany, developed an emulsion-coated canvas<br />

in the late 1950s <strong>and</strong> early 1960s (24). Designed for theater use, advertising<br />

purposes, <strong>and</strong> the reproduction of stitches, this Photoleinen, or photo linen,<br />

was first marketed in 1962. The product was actually made from a cotton<br />

fabric prepared with a pigmented gelatin layer. The pigmenting agent comprised<br />

a mixture of baryta sulfate <strong>and</strong> titanium oxide. The photographic<br />

emulsion applied to the gelatin-coated fabric was the same as that used in<br />

the manufacture of baryta papers of medium gradation. It was also unwashed,<br />

that is, all superfluous salts were left in the emulsion. A 1991 product list from<br />

Luminos Photo Corporation of the United States listed the availability of<br />

sheet <strong>and</strong> roll forms of a similar product. The sheet form had been impregnated<br />

with a bromide emulsion, <strong>and</strong> the roll form with a chlorobromide<br />

emulsion (25).<br />

New lines of ready-to-apply photographic emulsions also became, <strong>and</strong> continue<br />

to be, available. One version, developed by Argenta of Munich, was<br />

described as suitable for most surfaces (26). Poor adhesion was to be remedied<br />

with a preparatory layer of varnish; metallic surfaces were to be precoated<br />

with gelatin. Print-E-Mulsion, a version developed in the United States, first<br />

appeared in the mid-1970s (27). At the end of the 1970s, the name was<br />

changed to Liquid Light.<br />

Contemporary painters<br />

Several contemporary painters have employed such materials in their work.<br />

The American artist Lynton Wells generally used a canvas manufactured by<br />

Argenta. This was used in the production of paintings <strong>and</strong> sculptures from<br />

the late 1960s until about 1983. In some of these paintings, a single image<br />

spanned multiple panels, where the total length could exceed four meters. To<br />

create these works, the artist tacked the photo linen to the studio wall <strong>and</strong><br />

exposed the material in situ. To ensure that the image was properly aligned<br />

when the canvases were stretched, the tacking margin of each section was<br />

carefully folded under, before the edges of the canvas were put in contact.<br />

Processing was done in homemade developing trays <strong>and</strong> the images finished<br />

with one or more media. Oils, acrylics, aniline dyes, pastels, <strong>and</strong> charcoal were<br />

applied in varying densities. Frequently, these additions mimicked elements<br />

present in the photographic portion of the image. Wells also normally applied<br />

two to three layers of an acrylic polymer before painting with oils (28).<br />

Other painters who have made use of photographic bases in their work have<br />

included Arnulf Rainier of Austria, Shirley Wiitasalo <strong>and</strong> Kathleen Vaughan<br />

of Canada, Anselm Kiefer of Germany, <strong>and</strong> Fariba Hajamadi <strong>and</strong> James Turrell<br />

of the United States. From the late 1960s through the 1970s, Arnulf Rainier<br />

used photographs, photographs mounted onto wood or aluminum, <strong>and</strong> photo<br />

linen (29). The range of media employed in these works included oils, oil<br />

crayon, pencil, <strong>and</strong> ink. Portions of the photographic image remain visible in<br />

many of his works, but in some it has been completely negated by the thick<br />

application of paint.<br />

Figure 1. Shirley Wiitasalo, Interior,<br />

1981. Oil <strong>and</strong> photo emulsion on canvas.<br />

Photograph courtesy of National Gallery of<br />

Canada, Ottawa.<br />

Shirley Wiitasalo created the 1981 painting Interior by applying Liquid Light<br />

over a stretched cotton canvas prepared with an acrylic gesso (Fig. 1). Exposure<br />

was carried out using a slide projector. The image was then finished<br />

with Bellini oils <strong>and</strong> left unvarnished (30). As one of a series of paintings<br />

dominated by the television screen <strong>and</strong> featuring ambiguous <strong>and</strong> distorted<br />

Gayer 193


imagery, it was the only work in which a photographic emulsion had been<br />

employed (31).<br />

Kathleen Vaughan's painting technique has also involved the use of Liquid<br />

Light. In some of her work, the emulsion was applied directly to the canvas<br />

<strong>and</strong> processed by sponging chemicals over the support. Areas of the canvas<br />

that the artist intended to appear similar to the photographic portions were<br />

stained with a distemper medium. These layers were then isolated with a<br />

coating of an acrylic varnish; texture was applied with acrylic media <strong>and</strong> the<br />

final details in oil. Due to allergies, the artist has only used linseed oil to thin<br />

her oil paints (32).<br />

Anselm Kiefer has become well known for his robust paintings in which<br />

photographic enlargements, mounted on canvas, support oil, acrylic, shellac,<br />

straw, s<strong>and</strong>, <strong>and</strong> lead additions. Kiefer's photographic underdrawings have been<br />

printed on Dokumentenpapier P 90. The harsh or catastrophic appearance<br />

of these images was created through the manipulation of lighting or processing<br />

methods (33). Although the photographic image remains visible in<br />

some of the paintings, a number of them have been completely overpainted.<br />

In such cases, only the excess of support material on the reverse of the stretcher<br />

may provide evidence as to the presence of a photographic substrate (34).<br />

Since the late 1980s, Fariba Hajamadi, an Iranian-born painter living in New<br />

York, has combined paint <strong>and</strong> photographic images to produce fictitious interiors<br />

(35). These paintings have been made by brushing a commercially<br />

available photographic emulsion directly onto cotton canvas or wood. This<br />

method was selected to enable the texture or grain of the support to contribute<br />

to the composition. Following exposure <strong>and</strong> processing, an airbrush<br />

was used to apply color. The paint, a commercially available transparent oil<br />

paint, had been thinned to the appropriate consistency with lacquer thinner<br />

(36).<br />

In the planning for his "Roden Crater Project," James Turrell has produced<br />

studies on frosted drafting mylar in which wax, photographic emulsion, <strong>and</strong><br />

various paint <strong>and</strong> graphic media have been combined (37). These studies were<br />

created in the following way: A coating of hot beeswax was first sprayed onto<br />

the mylar. This wax layer was then coated with a photographic emulsion <strong>and</strong><br />

the desired image of the crater exposed <strong>and</strong> processed. The image was then<br />

manipulated <strong>and</strong> elements may have been removed with an eraser or by<br />

scraping with a knife. Frequently, wax pastels were used to replace removed<br />

portions, make additions, or enhance or blur particular details. The colors<br />

used were carefully chosen so that the additions might st<strong>and</strong> out or coalesce<br />

with the existing image. In some cases, a type of s<strong>and</strong>wich was made by dry<br />

mounting drafting vellum to the emulsion-coated surface. Further details were<br />

then added to the front or back of these studies with ink, paint, graphite, or<br />

wax pastels (38).<br />

Conclusion<br />

<strong>Painting</strong> <strong>and</strong> photography are techniques that have frequently been used in<br />

conjunction with each other. For the most part, this relationship has been<br />

based on visual <strong>and</strong> intellectual concerns, not the material union of media.<br />

This paper has attempted to illustrate that although painters may not always<br />

have utilized photographic materials in their work, information regarding this<br />

possibility remained available <strong>and</strong> contributed to related <strong>and</strong>/ or contemporary<br />

forms of art production. The genre of traditional portraiture was excluded<br />

from the discussion; while some portraitists may have utilized such materials<br />

in the past, recent examples of this were not found.<br />

Notes<br />

1. Ruggles, M. 1983. A study paper concerning oil paintings on a photographic<br />

base. AIC Preprints, 104-12.<br />

194<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


2. Eder, ]. M. 1945. History of Photography. New York: Columbia University Press,<br />

791, note 4.<br />

3. Coe, B., <strong>and</strong> M. Haworth-Booth. 1983. A Guide to Early Photographic Processes.<br />

London: Hurtwood Press, 19.<br />

4. Gernsheim, H., <strong>and</strong> A. Gernsheim. 1969. The History of Photography. London:<br />

Thames <strong>and</strong> Hudson, 315.<br />

5. Joe,]. 1894. Coloring photographs. Scientijic American 71 (11):163.<br />

6. Scientijic American. 1895. Coloring photographs. Scientijic American 73 (1):27.<br />

7. Scientijic American. 1899. Coloring bromide prints. Scientijic American 81 (21):323.<br />

8. Godbold, A. V 1919. A wax medium for the permanent coloring of photographs.<br />

Scientijic American Supplement 87 (2248):74-75.<br />

9. Tobias, ]. C. 1934. The art of coloring photographic prints in transparent water-color,<br />

tempera, opaque <strong>and</strong> transparent pastel, wax crayons, opaque <strong>and</strong> transparent oils, chemical<br />

coloring, <strong>and</strong> coloring lantern slides. Boston: American Photographic Publishing<br />

Co., 54, 57-58, 60-61.<br />

10. Johnson, R. 1936. The art of retouching photographic negatives <strong>and</strong> photographic enlargements,<br />

etc. Revised <strong>and</strong> rewritten by T. S. Bruce <strong>and</strong> A. Braithwaite, 13th ed.<br />

(2nd American, revised <strong>and</strong> enlarged by A. Hammond. Boston: American Photographic<br />

Publishing Co., 143, 148-150, 152.<br />

11. Duryea, D. 1936. Notes on murals by photography. Architectural Record 80 (1):57-<br />

58, 60, 62.<br />

12. Architectural Review. 1937. The true photo-mural: a new technique of decoration.<br />

Architectural Review 81 (February) :86-87.<br />

13. Perry, H. W 1940. Photography speeds up aircraft production. American Photography<br />

34 (12): 886, 888.<br />

14. Progressive Architecture. 1946. Photographic emulsion makes any base possible. Progressive<br />

Architecture 27 (110):110.<br />

15. Duryea, op. cit., 60.<br />

16. Hochman, L. 1947. Print your photos anywhere. Popular Science 150 Oune):196.<br />

17. Corbino, M. 1986. James Couper. American Artist 50 (529):66, 68-69, 84-86.<br />

18. Patton, P. 1976. Super realism: a critical anthology. Arifo rum 14 (5):52.<br />

19. Chase, L., N. Foote, <strong>and</strong> T. McBurnett. 1972. The Photo-Realists: twelve interviews.<br />

Art in America 60 (6):73-89.<br />

20. Wade, K. E. 1976. Alternative Photographic Processes. Dobbs Ferry, New York: Morgan<br />

<strong>and</strong> Morgan, Inc.<br />

21. Howell-Koehler, N. 1980. Photo Art Processes. Worcester, Massachusetts: Davis<br />

Publications, Inc.<br />

22. Eastman Kodak Co. 1982. Photographic sensitizer for cloth <strong>and</strong> paper. Photo<br />

information bulletin PIB-6. December 1982.<br />

23. Eastman Kodak Co. 1983. Making a photographic emulsion. Photo information<br />

bulletin PIB-7. February 1983.<br />

24. Hegenbart, W 1994. Personal communication. Division Manager, Finishing Europe,<br />

Tetenal Vertriebs GmbH, Norderstedt, Germany.<br />

25. Luminos Photo Corp. 1991. B & W Products fo r the Professional. Yonkers, New<br />

York 10705.<br />

26. Hegenbart, op. cit.<br />

27. Cone, R. 1993. Personal communication. Rockl<strong>and</strong> Colloid Corp., Piermont,<br />

New York 10968.<br />

28. Wells, L. 1991. Personal communication. New York, New York.<br />

29. Fuchs, R. H. 1989. Arnulf Rainer. Vienna: ARGE Gabriele Wimmer & John<br />

Sailer, 147-48.<br />

30. Gagnier, R. 1991. Personal communication. Conservator of Contemporary Art,<br />

National Gallery of Canada, Ottawa.<br />

31. Monk, P. 1987. Shirley Wiitasalo. Toronto: Art Gallery of Ontario, 13-14.<br />

32. Vaughan, K. 1993. Personal communication. Toronto, Ontario.<br />

33. Winter, P. 1989. Manipulieren, ironisieren, maltratieren: Das Foto unter der H<strong>and</strong><br />

des Kiinstlers. Kunstwerk 42 (March): 11.<br />

34. Penn, S. P. 1993. Personal communication. Associate Conservator of <strong>Painting</strong>s,<br />

Philadelphia Museum of Art, Philadelphia, Pennsylvania 19101.<br />

35. Cotter, H. 1990. Fariba Hajamadi at Christine Burgin. Art in America 78 (7):169.<br />

36. Hajamadi, F. 1991. Personal communication. New York, New York.<br />

37. Adcock, C. 1990. James Turrell: The Art if Light <strong>and</strong> Space. Berkeley <strong>and</strong> Los<br />

Angeles: University of California Press, 187-96.<br />

38. Turrell,]. 1993. Personal communication. Flagstaff, Arizona.<br />

Gayer 195


Abstract<br />

It is argued here that painters of the<br />

Baroque adhered to the die-hard tradition<br />

of loading their palettes with<br />

a limited number of tints, suitable<br />

only for painting the passage they<br />

planned to finish in that stage of the<br />

work. Support for this proposition<br />

comes from various directions: written<br />

sources, studio representations,<br />

<strong>and</strong> scientific research generated by<br />

different methods. An example of a<br />

specific studio practice is used to<br />

demonstrate the much discussed interrelation<br />

of technique <strong>and</strong> style.<br />

Reflections on the Relation between Te chnique<br />

<strong>and</strong> Style: The Use of the Palette by the<br />

Seventeenth-Century Painter<br />

Ernst van de Wetering<br />

Kunsthistorisch Instituut<br />

Herengracht 286<br />

1016 BX Amsterdam<br />

The Netherl<strong>and</strong>s<br />

Introduction<br />

In the Vitae, as is widely known, Giorgio Vasari attempted to describe the<br />

development of Italian art from Giotto onward as a process of continuous<br />

progress culminating in the work of Michelangelo. The idea that art changes<br />

because better solutions appear was certainly not restricted to Vasari. While<br />

stylistic developments in Western art were viewed as a matter of progress, it<br />

was inevitable that the characteristics of earlier styles would be explained in<br />

relation to problems that had meanwhile been solved. This could apply to<br />

the invention of perspective <strong>and</strong> to the invention of oil painting-which,<br />

according to Vasari, "softens <strong>and</strong> sweetens the colors <strong>and</strong> renders them more<br />

delicate <strong>and</strong> more easily blended than do the other mediums" (1). Finally, the<br />

development of style can be applied within oil painting to the development<br />

of the technique fo r achieving a "'glowing" incarnate, as Karel van M<strong>and</strong>er<br />

termed it, by means of an underpainting in vermilion that "glows more<br />

fleshy" (2).<br />

Viewed thus, there can be said to exist a clear relation between style <strong>and</strong><br />

technique. During the nineteenth century, however, this was a hotly debated<br />

issue among art theorists. Some of these, especially writers on architecture<br />

such as Gottfried Semper (1803-1 879) <strong>and</strong> Viol1et-le-Duc (1814-1 879),<br />

aimed to demonstrate a fu ndamental interrelation of style <strong>and</strong> technique in<br />

the arts (3). Twentieth-century developments, such as the theories underlying<br />

the Bauhaus, continued to build on the same ideas.<br />

Not everyone agrees that style <strong>and</strong> technique are interrelated. A parallel<br />

stream of art history adheres to the idea, current since the Romantic period,<br />

that every fo rm of art had its own fo rmal legitimacy. In this way of thinking,<br />

it was not necessary to explain styles in terms of technical limitations <strong>and</strong><br />

possibilities. This line of thought culminated in the concept of Kunstwollen,<br />

originated by Alois Riegl (1858-1 905), which was seen as the manifestation<br />

of an "urge to fo rm," independent of the restrictive influence of such factors<br />

as function, materials, <strong>and</strong> technique. All the same, Riegl did not deny the<br />

influence of technique. He believed, however, that Kunstwollen overcame the<br />

technical limitations. Technical frontiers, considered by Semper to play a positive<br />

part in the creative process, constituted in Riegl's view a "coefficient of<br />

friction" within the Gesamtprodukt of Kunstwollen (4).<br />

It could be argued that Riegl's notion of Kunstwollen was partly responsible<br />

for the fact that "style" has so long remained one of the main domains of art<br />

historical research, with art historians such as Wollflin <strong>and</strong> Focillon being<br />

prominent representatives of this direction. It is worth noting here that the<br />

stylistics ofWollflin were dominated by an outlook in which the autonomous<br />

or even abstract qualities of the visual vocabulary took priority over the<br />

pictorial means employed to achieve a convincing representation of reality.<br />

Since Wollflin's times, research into artistic techniques has become more <strong>and</strong><br />

more detached from stylistic considerations. Owing to the shift towards sci-<br />

This article was previously published, in a different form, in Oud Holl<strong>and</strong> 107: 1, 1993,<br />

137-51; <strong>and</strong> in KM, vakil'!for111atie voor beeldende Kunstenaars en res fa rato ren, Fall 1994,<br />

28-31; reprinted, with changes, by permission of the author.<br />

196<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


Figure 1. Jan Baptiste Collaert, Color Olivi, 1566- 1628. After Joharll1es Stradanus. Courtesy of<br />

the Rij"ksmuseum-Stichting, Amsterdam.<br />

entific research methods, research into historic techniques has now come to<br />

play a central part in research on authenticity <strong>and</strong> workshop practice, as well<br />

as in conservation <strong>and</strong> restoration.<br />

Research into historic techniques has made considerable progress in these<br />

areas during recent decades, so it now makes sense for current applications<br />

<strong>and</strong> future research to turn renewed attention to the interrelation between<br />

technique <strong>and</strong> style. This interrelation is examined here through the study of<br />

developments affecting the painter's palette.<br />

Palettes<br />

One of the most fascinating <strong>and</strong> complete documents concerning the history<br />

of the art of oil painting is a well-known engraving by Jan Baptist Collaert,<br />

after Stradanus, which dates from the end of the sixteenth century (Fig. 1).<br />

This engraving gives a highly detailed picture of an idealized painter's studio:<br />

the master is working on a history piece, while an assistant is occupied with<br />

painting a portrait. Two other assistants are grinding <strong>and</strong> preparing colors.<br />

The engraving shows countless details that provide valuable hints about dayto-day<br />

practice in the late-sixteenth-century painter's studio. In the fo reground,<br />

three boy apprentices can be seen; the smallest is practicing the<br />

rudiments of drawing, <strong>and</strong> the more advanced apprentice on the left is drawing<br />

from plaster casts.<br />

For our present purposes, we are solely concerned with the apron-clad apprentice<br />

st<strong>and</strong>ing next to the master. He is setting out a palette of small shells,<br />

presumably containing colors prepared by the assistants, <strong>and</strong> holds a palette<br />

similar to that of the master. As it will soon be made clear, the arrangement<br />

of the master's palette with so few colors is far from arbitrary. Like so many<br />

other details, this must be regarded as a faithful representation of sixteenthcentury<br />

practice. The youth in the foreground has applied a limited number<br />

of colors to the palette in his h<strong>and</strong>. As on his master's palette, they are spread<br />

out over the surface of the palette.<br />

There are generally no written sources to be fo und on the most routine<br />

activities of the painter's studio through history. Incidental evidence on certain<br />

aspects of painterly practice can be gleaned from documents, but we have to<br />

live with the fact that such sources are extremely sporadic in both time <strong>and</strong><br />

place. In the present case, a late-seventeenth-century Italian text can be shown<br />

van de Wetering 197


to relate to the activities depicted on the previously mentioned engraving by<br />

Collaert, which originated a hundred years earlier in the Netherl<strong>and</strong>s. Using<br />

available sources in this way is, of course, only justifiable when the phenomenon<br />

being investigated is widespread <strong>and</strong> displays a certain constancy. The<br />

considerable mobility of painters in the sixteenth <strong>and</strong> seventeenth centuries<br />

provided many opportunities for the spreading of painterly techniques <strong>and</strong><br />

procedures. Many young painters traveled across Europe <strong>and</strong> worked in the<br />

studios of various masters for shorter or longer periods; a similar situation still<br />

exists in the international art restorers' world. This form of mobility ensured<br />

a rapid dissemination of knowledge <strong>and</strong> experience, leading to a high level<br />

of international uniformity in knowledge <strong>and</strong> craft practices. Bearing this in<br />

mind, one might venture a cautious guess that the question of what the youth<br />

with the apron is doing in the studio of Stradanus could be answered by a<br />

passage from the Volpato Manuscript, a seventeenth-century Italian document.<br />

This text by Giovanni Battista Volpato (born in 1633) must date from<br />

somewhere around 1680, <strong>and</strong> was written in the fo rm of a series of dialogues.<br />

It contains the following exchange between F., an older painter's apprentice,<br />

<strong>and</strong> Silvio, a younger apprentice (5):<br />

Silvio: Tell me, if you will, whether you set your master's palette.<br />

F: Surely . ... It suffices Jor him to tell me what he intends to paint, Jor<br />

I then know which colors I must place on the palette.<br />

The engraving after Stradanus <strong>and</strong> this late-seventeenth-century text provide<br />

two of the few hints-which until now have not been given any consideration<br />

in the literature on art history <strong>and</strong> painting techniques-that painters<br />

formerly used palettes which were set with groups of colors specifically for<br />

certain parts of the painting, <strong>and</strong> which thus did not include all the available<br />

pigments. This underst<strong>and</strong>ing of the situation has a far-reaching consequence:<br />

one must then see the seventeenth-century (but also earlier or later) painting<br />

as a composite image made up of interlocking passages, comparable to the<br />

giornate, the successively executed "daily portions," of fresco painting, although<br />

in the case of oil painting, a number of passages would have generally been<br />

completed on a given day (6).<br />

In using such a method, the seventeenth-century way of painting differed<br />

fu ndamentally from the approach of late nineteenth- <strong>and</strong> twentieth-century<br />

artists, namely in developing the painting as a tonal entity. For example, the<br />

Hague School painter Jozef IsraeIs used a palette with a full range of colors<br />

<strong>and</strong> with a mixing area covered with patches of mixed paint, of which the<br />

tone <strong>and</strong> color could be further modified. A palette of this kind enabled the<br />

artist to continue working over the whole area of the painting simultaneously,<br />

with an eye to controlling the tonal consistency of the painting.<br />

In the first instance, the idea of earlier artists working in giornate as described<br />

above may sound highly exaggerated. After all, we know that the great majority<br />

of seventeenth-century painters did in fact conceive their painting as<br />

a tonal unity, as is evidenced by the practice of starting with a largely monochromatic<br />

or "dead color" underpainting (7) . The point of importance here<br />

is that after laying in the underpainting, the artist developed the composition<br />

further by successively adding isl<strong>and</strong>s of modulated local color. Once this is<br />

understood, it becomes clear that a painting such as Rembr<strong>and</strong>t's Jewish Bride<br />

has far more in common with work of predecessors such as Raphael than<br />

with the paintings of Jozef IsraeJs <strong>and</strong> his contemporaries, however much<br />

inspiration IsraeIs drew from that painting.<br />

A method such as that described is intimately connected with the material,<br />

technical, <strong>and</strong> economic constraints that were inherent to oil painting, constraints<br />

that only vanished (<strong>and</strong> were subsequently forgotten) with the introduction<br />

of ready-to-use, "mutually compatible" tube colors (8).<br />

The early palettes were small. Only in the course of the nineteenth century<br />

did they grow to the size of small tabletops. It will be clear that this relatively<br />

198<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


sudden increase in the dimensions of the palette is part of the argument put<br />

forth here (9).<br />

When one studies the distribution of various tints on the countless palettes<br />

that appear in paintings, it is impossible to avoid the impression that the<br />

painters adhered strictly to a certain set of rules. This is certainly true for the<br />

period after about 1600. Prior to that time, the palette was ostensibly set with<br />

relative freedom. But it is precisely in those earliest representations of palettes<br />

that it is sometimes obvious that the palettes were set up specifically for each<br />

passage to be painted.<br />

Niklaus Deutsch's painting St. Luke <strong>Painting</strong> the Madonna (1515) is typical of<br />

a group of paintings completed in the fifteenth <strong>and</strong> sixteenth centuries, in<br />

which St. Luke is in the process of painting the Madonna's robe. The depicted<br />

artists' palettes show a limited number of patches of paint, with various shades<br />

of blue as well as black <strong>and</strong> a little white. These are exactly the colors needed<br />

to render the modeling in the drapery of the blue robe-distributed apparently<br />

at r<strong>and</strong>om over the surface of the palette (10). In paintings of the same<br />

period in which St. Luke is painting the naked Christ Child or the face of<br />

Maria, the palette carries the range of colors needed to mix the various tints<br />

of the flesh: white, yellow ochre, vermilion, red lake, various browns, black,<br />

<strong>and</strong> sometimes terre verde (11).<br />

As explained in the following discussion, the flesh tint, like blue, had an<br />

important status. The colors for painting the human skin were not yet systematically<br />

arranged on the palette before 1600, but appear to be distributed<br />

at r<strong>and</strong>om. For example, the position of the white paint differs from one case<br />

to another on earlier palettes. From 1600 onward, studio scenes <strong>and</strong> self portraits<br />

depict palettes with a row of lumps of paint spaced evenly along the<br />

top edge. The range of colors depicted normally runs from a somewhat larger<br />

portion of white near the thumb, to yellow ochre, vermilion, red lake, <strong>and</strong><br />

then through a series of progressively darker browns to black. Alternately, the<br />

vermilion is sometimes placed between the white <strong>and</strong> the thumb. This arrangement<br />

agrees with a passage in the Mayerne Manuscript: "It is to be<br />

observed, moreover, that in setting the palette, the lightest tints must always,<br />

without exception, be placed at the top <strong>and</strong> the darker tints lower down"<br />

(12). It is striking, <strong>and</strong> very significant in the context of this article, that the<br />

author has found no instance of the depicted st<strong>and</strong>ard palettes from either<br />

before 1600 or afterward that includes any intense green, bright yellow, or<br />

blue paint. These are what are called, in the Mayerne Manuscript of 1630,<br />

"the strong colors" (13). The reason for the absence of these colors is clarified<br />

by the following.<br />

In the Volpato Manuscript, the older apprentice tells the younger that his<br />

master merely has to indicate what passage is to be painted in order to lay<br />

out an appropriate palette. This statement implies the existence of fixed recipes<br />

for reproducing the various elements of nature. Around the same time,<br />

the Dutch painter Willem Beurs wrote down such recipes for the benefit of<br />

both "students of the Noble Painterly Art" <strong>and</strong> interested "amateurs" (14).<br />

As an example, a recipe for painting a white horse follows (15):<br />

One paints the illuminated side using white, light ochre <strong>and</strong> black, with<br />

pure white for the highlights; light ochre is recommended for the intermediate<br />

color, <strong>and</strong> it is advisable there to be rather sparing with white. For the<br />

shadow, black <strong>and</strong> light ochre must be mixed together with a little white;<br />

the niflection under the belly should be mostly light ochre, with sparing use<br />

of black <strong>and</strong> white. The hooves can sometimes be painted with black, white<br />

<strong>and</strong> light ochre, with a touch of vermilion; <strong>and</strong> sometimes with black, white,<br />

<strong>and</strong> umber. The color of the nose is the same as that of the hoofs. But as<br />

fo r the eyes: the pupil should be painted with bone black, <strong>and</strong> the rest with<br />

umber, black <strong>and</strong> white.<br />

It is clear that if an artist plans to paint a white horse, the only pigments<br />

needed on the palette are white lead, yellow ochre, vermilion, umber, <strong>and</strong><br />

van de Wetering 199


one black; <strong>and</strong> if the painter expects painting the horse to be a day's work,<br />

no other colors need be prepared for that day, apart from the five pigments<br />

just mentioned.<br />

It is notable that throughout hundreds of recipes of this kind, only a severely<br />

limited range of pigments is prescribed. This provides a piece of information<br />

that is of critical importance to the picture of the artists' practice being<br />

sketched here: paint was only ground <strong>and</strong> prepared when it was needed. It is<br />

precisely because painters wished to keep working without unnecessary delays<br />

<strong>and</strong> therefore found it advantageous to use paint that dried quickly (thus<br />

containing strong drying oils <strong>and</strong> other drying agents), that paint could not<br />

be kept for long. If the complete range of pigments, each already ground<br />

with oil, had to be available, this would mean that much paint would have<br />

to be thrown away unused. Hence the economic reasons for the method of<br />

working with restricted, specific palettes are clear. The technical background<br />

of this method will be discussed briefly in this paper.<br />

It is significant that the series of recipes recorded by Willem Beurs concludes<br />

with flesh colors (16). Beurs writes as follows: "Just as we humans consider<br />

ourselves the foremost among animals; so, too, are we the foremost subject of<br />

the art of painting, <strong>and</strong> it is in painting human flesh that its highest achievements<br />

are to be seen" (17). The palette Beurs gives fo r painting human flesh<br />

comprised mixtures of the pigments lead white, light ochre, schijtgeel (an organic<br />

yellow), vermilion, red lake, tawny ochre, terre verde, umber, <strong>and</strong> "coal<br />

black" (probably ground charcoal which gives a bluish black). It was the<br />

palette for what was considered to be the summit of creation, <strong>and</strong> the most<br />

difficult subject of all to paint, the human figure. It is almost invariably the<br />

palette for flesh colors that is depicted in self portraits <strong>and</strong> studio scenes after<br />

1600, paintings that are generally intended to represent the art of painting at<br />

its noblest.<br />

It will come as no surprise to anyone who has examined fifteenth- or sixteenth-century<br />

paintings to any depth as material objects, that every passage<br />

was executed as a separate entity. The additive character of the painting as a<br />

whole is generally plain to see, despite the psychological compulsion perceptually<br />

acting on the viewer to transform the painting into a Gestalt.<br />

Even though revolutionary developments during the seventeenth century<br />

brought the pursuit of pictorial unity to an unprecedented level, the sources<br />

quoted above suggest that no change had taken place in the tradition of using<br />

recipes for various components of the painting. The economic rationale for<br />

this approach has already been mentioned above. Originally, however, technical<br />

reasons may have provided an even stronger motivation for executing a<br />

painting as a series of successive passages.<br />

For the twentieth-century painter, who normally regards paint as a pasty<br />

substance of a certain color that can be squeezed out of a tube, it is hard to<br />

imagine that to artists of not only the fifteenth <strong>and</strong> sixteenth centuries, but<br />

also through the first half of the nineteenth century, each pigment presented<br />

its own inherent possibilities <strong>and</strong> constraints (18). Some pigments could not<br />

be worked up with oil; some pigments could only safely be mixed with one<br />

or two other pigments; some pigments could only be used transparently <strong>and</strong><br />

yet others only opaquely. Other properties, too, such as color permanence,<br />

workability, drying qualities, <strong>and</strong> so on, could differ so strongly from one<br />

pigment to another that it was normal to use a given pigment either in pure<br />

form or mixed with any of a limited number of other pigments in order to<br />

somewhat modifY the tone <strong>and</strong> color (19). This helps explain why, in the<br />

work of artists such as van Eyck or Lucas van Leyden, the colors unmistakably<br />

interlock like the pieces of a jigsaw puzzle, <strong>and</strong> each color has an individual<br />

character, especially as to transparency, surface texture, <strong>and</strong> thickness of the<br />

paint layer. The most easily workable pigments were the earth colors, which<br />

ranged from yellow ochre through red ochres to the darkest brown tints <strong>and</strong><br />

were varied in tone by mixing together <strong>and</strong> by the addition of white or<br />

200<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, <strong>Materials</strong>, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


lacks. Passages painted with mixtures of these pigments are subtle in their<br />

tonal values <strong>and</strong> graduations. Whenever it was necessary to achieve strong,<br />

bright colors (i.e., red, yellow, <strong>and</strong> blue robes), it is clear that the passage<br />

concerned was executed within carefully delineated contours in accordance<br />

with a fixed recipe, involving a specific layering or a fixed type of underpainting.<br />

This also explains why these colors are usually absent from fleshcolor<br />

palettes depicted in studio scenes <strong>and</strong> self-portraits.<br />

Figure 2. Neutronactivation autoradiographic<br />

image of Rembr<strong>and</strong>t's Bellona, 1633. The<br />

Metropolitan Museum of Art, New York. In<br />

this image the activated phosphor atoms (from<br />

bone black) <strong>and</strong> mercury atoms (from vermilion,<br />

used on the figure'S mouth) have mainly<br />

blackened the film.<br />

Figure 3. The image of the autoradiograph<br />

in Fig. 2 is mainly determined by the radiation<br />

from the copper atoms (from the bluegreen<br />

parts in the painting) <strong>and</strong> again the<br />

mercury of the vermilion.<br />

That artists worked in this way during the fifteenth <strong>and</strong> sixteenth centuries<br />

can easily be discerned from the paintings themselves. But did this practice<br />

continue into the Baroque period? In the case of Rembr<strong>and</strong>t, this is by no<br />

means self-evident. There were already indications, however, that Rembr<strong>and</strong>t<br />

must have used this method; it has been observed that he completed his<br />

paintings systematically from background to foreground, passage by passage,<br />

on the basis of a monochromatic underpainting (20). This does not necessarily<br />

imply, however, that he did this using a restricted palette. An important piece<br />

of evidence for that practice is supplied by scientific studies. It was the failure<br />

of a reconstruction project that, in fact, formed the seed of the study presented<br />

here. The preparation of a dummy Rembr<strong>and</strong>t using all the procedures of<br />

Rembr<strong>and</strong>t's studio that were known at the time was unsuccessful because<br />

the painting was executed using a complete palette. All the pigments that<br />

were in use during the seventeenth century were present on the palette <strong>and</strong><br />

they were mixed on the same mixing surface, which was cleaned at intervals.<br />

Autoradiographic investigation of the dummy made it clear that the procedure<br />

used resulted in all the pigments being present to a greater or lesser<br />

degree all over the painting. By contrast, autoradiography of Rembr<strong>and</strong>t's<br />

actual paintings resulted in surprisingly "clean" images (Figs. 2, 3) (21).<br />

In Rembr<strong>and</strong>t's painting, certain pigments occur only within clearly demarcated<br />

zones, not in the rest of the painting. This observation is additionally<br />

confirmed by studies of paint samples from Rembr<strong>and</strong>t's paintings, which<br />

time <strong>and</strong> again reveal that only a limited number of pigments were used in<br />

a given passage. The mixtures found usually consist of two to four different<br />

pigments; mixtures of five-or, in very exceptional cases, six-pigments are<br />

found only incidentally (e.g., in flesh passages). To achieve the typically Baroque<br />

tonal unity using such methods implies a highly developed level of<br />

"management" in the use of colors <strong>and</strong> tones. It is therefore no coincidence<br />

that artists began theorizing about the nature of this "management" during<br />

the seventeenth century. This was made clear by Paul Taylor's study of the<br />

term houding (roughly, "disposition") that appears regularly in seventeenthcentury<br />

sources (22). Until some time into the sixteenth century, the picture<br />

space was still structured simply as a foreground with a background or campo,<br />

a concept on which Jeroen Stumpel performed an important study (23). The<br />

idea denoted by the term houding was a far more complex one, however.<br />

Seventeenth-century writers used houding to denote the spatial coherence<br />

created in the painting by the disposition of tones <strong>and</strong> colors. The viewer<br />

could "walk" through this space in his imagination. In a formulation by Willem<br />

Goeree (1668), houding is (24):<br />

. . . that which binds everything together in a Drawing or <strong>Painting</strong>, which<br />

makes things move to the front or back, <strong>and</strong> which causes everything from<br />

the foreground to the middle ground <strong>and</strong> thence to the background to st<strong>and</strong><br />

in its proper place without appearing fu rther away or closer, <strong>and</strong> without<br />

seeming lighter or darker, than its distance warrants; so that everything<br />

st<strong>and</strong>s out, without confusion, from the things that adjoin <strong>and</strong> surround<br />

it, <strong>and</strong> has an unambiguous position through the proper use of size <strong>and</strong><br />

color, <strong>and</strong> light <strong>and</strong> shadow; <strong>and</strong> so that the eye can naturally perceive the<br />

intervening space, that distance between the bodies which is lift open <strong>and</strong><br />

empty, both near <strong>and</strong> far, as though one might go there on foot, <strong>and</strong><br />

everything st<strong>and</strong>s in its proper place therein.<br />

When one considers how a painter would have been intensively concerned<br />

with obtaining a good houding while at the same time working with selective<br />

van de Wetering 201


palettes as described above, it is clear that the level of accomplishment required<br />

was of a quite different nature to the methods of painters from the<br />

later nineteenth <strong>and</strong> twentieth centuries, who, thanks to the full palette, could<br />

now work on the whole painting at once. This fact clearly explains the somewhat<br />

"muddy" effect, compared to earlier paintings, characteristic of Jozef<br />

Israels <strong>and</strong> many of his contemporaries <strong>and</strong> successors.<br />

Research into the borderline region between style <strong>and</strong> technique throws light<br />

on unexpected aspects of seventeenth-century studio practice. It is clear, for<br />

instance, that the view on the genesis of paintings presented here must have<br />

implications for scientific research into Baroque paintings. Analytical studies<br />

of the binding media used in the seventeenth century (e.g., by Rembr<strong>and</strong>t)<br />

suggest that a different blend of media is likely to be found in every giornate<br />

of the painting (25).<br />

This study has produced a picture that may be relevant to the discussion of<br />

the interrelation of style <strong>and</strong> technique, namely that of a development of<br />

Baroque painting in which artists strove for a unified tonal image that did<br />

not appear to be composed by additive methods. The fact that these additive<br />

techniques were used gives a different picture of the artist at work, one in<br />

which technical limitations actually constituted the "coefficient of friction"<br />

(to use Riegl's term) that thwarted the artists in their pursuit of a new style.<br />

It was even necessary for the seventeenth-century painters to introduce a new<br />

art-theoretical concept, namely houding, to bring the discussion of these efforts<br />

to a new level of abstraction. Only in the nineteenth century did the<br />

technical materials <strong>and</strong> means arise-namely, the full color palette, paint in<br />

tubes, <strong>and</strong> the associated large palette sizes-that made it possible to apply a<br />

technique appropriate to the painter's stylistic aspirations (as exemplified by<br />

Cezanne's remark, "You must underst<strong>and</strong> that I h<strong>and</strong>le the whole painting at<br />

once, in its totality") (26). Semper appears to be borne out, on the other<br />

h<strong>and</strong>, if one observes that a late Rembr<strong>and</strong>t, placed between a Raphael <strong>and</strong><br />

a nineteenth-century piece of "Rembr<strong>and</strong>tism" such as a late work by Jozef<br />

Israels, shows closer kinship to the Raphael than to the Israels. Indeed, the<br />

technique of additive painting has an unmistakable determinative effect on<br />

the style of a painting.<br />

Aclrnowledgll1ents<br />

Thanks are hereby expressed to Jos Koldeweij , Hayo Menso de Boer, Paul Broeckhof,<br />

Dieuwertje Dekkers, Karin Groen, <strong>and</strong> Charlie Srnid. Much of the research underlying<br />

this study was carried out within the framework of the Rembr<strong>and</strong>t Research<br />

Project, financed by the Netherl<strong>and</strong>s Organization fo r Scientific Research, <strong>and</strong> also<br />

received valuable support from the Director <strong>and</strong> the Head of the Art History Department<br />

of the Central Laboratory for Research on Objects of Art <strong>and</strong> Science,<br />

Amsterdam.<br />

Notes<br />

1. Vasari on technique. 1960. Ed. B. Brown. New York, 230.<br />

2. van M<strong>and</strong>er, K. 1604. Den grondt der ede! vry Schilderconst. Haarlem, fol. 64r. P<br />

Taylor of the Warburg Institute, London, is carrying out research on the concept<br />

"glowing" in connection with the painting of flesh in the early seventeenth<br />

century.<br />

3. Semper, G. 1860-1863. Der Stil in den technischen und tektonischen KiJnsten. Frankfurt<br />

am Main. See also Viollet-Ie-Duc, E. E. 1863-1872. L'entretien sur I'architecture.<br />

Paris.<br />

4. Riegl, A. 1893. Stilfragen. Berlin, v-xix, 20, 24. See also Kultermann, U. 1966.<br />

Geschichte der Kunstgeschichte. Vienna, 288.<br />

5. Giovanni Battista Volpato: Modo da tener nel dipingere. 1967. In Original Treatises<br />

on the Arts oj <strong>Painting</strong>. Ed. M. P Merrifield. New York, 726-55, especially<br />

746-48.<br />

6. Mora, P, L. Mora, <strong>and</strong> P Philippot. 1984. Conservation oj Wall <strong>Painting</strong>s. London,<br />

138-64.<br />

7. van de Wetering, E. 1977. De jonge Rembr<strong>and</strong>t aan het werk. In Qud Holl<strong>and</strong><br />

(91):7-65, especially 20-24.<br />

202<br />

<strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong>, Maten:als, <strong>and</strong> <strong>Studio</strong> <strong>Practice</strong>


8. It is not clear when people first started keeping prepared paint in small pouches<br />

made of pig's bladder. The pouches are occasionally illustrated in eighteenthcentury<br />

sources. According to R. G. Gettens <strong>and</strong> G. L. Stout (1942. <strong>Painting</strong><br />

<strong>Materials</strong>: A Short Encyclopedia. New York, 318), the tube came into widespread<br />

use around 1840.<br />

9. A description of the evolution of the palette can be fo und in Gettens <strong>and</strong> Stout,<br />

op. cit. (note 8), 299-304.<br />

10. Colyn de Coter, St. Luke <strong>Painting</strong> the Madonna, Vievre, Alliers (near Moulins),<br />

ca. 1490; de Meester van het Augustiner Altaar, St. Luke <strong>Painting</strong> the Madonna,<br />

Germanisches Nationalmuseum Niirnberg, 1487; Nicolaus Manuel Deutsch, St.<br />

Luke <strong>Painting</strong> the Madonna, Bern Kunstmuseum, 1514-1515; Unknown Flemish<br />

Master, St. Luke <strong>Painting</strong> the Madonna, the Grimani Brevarium, fol. 781 v, Bibliotheca<br />

Marciana, Venice.<br />

11. Hinrik Bornemann, St. Luke <strong>Painting</strong> the Madonna, Hamburg, Jakobikerk, 1499;<br />

School of Quinten Massys, St. Luke <strong>Painting</strong> the Madonna, National Gallery, London;<br />

Derick Baegert, St. Luke <strong>Painting</strong> the Madonna, L<strong>and</strong>esmuseum, Munster,<br />

1490. Belonging to the same category of palette are: Jacob Cornelisz van Oostzanen,<br />

Self Portrait of the Artist Working on a Portrait of his Wife, Museum of Art,<br />

Toledo, ca. 1530; Catharina van Hemessen, Self Portrait, Offentliche Kunstsammlung<br />

Basel, ca. 1548.<br />

12. Pierre Lebrun, Recueil des essaies des merveilles de la Peinture. In Original Treatises<br />

on the Art of <strong>Painting</strong>, Vol. II. 1967. Ed. M. P. Merrifield. New York, 770-<br />

71.<br />

13. Quellen fur Maltechnik wahrend der Renaissance und deren Folgezeit (XVI-XVII Jahrhundert),<br />

Vol. IV 1901. Ed. E. Berger. Munich, 288-89.<br />

14. Beurs, W 1698. De Groote Waereld in het Kleen geschildert. Amsterdam.<br />

15. Beurs, op. cit., 165-66.<br />

16. Beurs, op. cit., 184-87.<br />

17. Beurs, op. cit., 186.<br />

18. This st<strong>and</strong>ardization of the working properties of various colors of paint is<br />

achieved by the addition of fillers such as bentonites or aluminium hydrates, or<br />

by the use of alternative pigments. Naples yellow, for example, is substituted by<br />

a mixture of titanium white <strong>and</strong> cadmium yellow.<br />

19. See for example the specifications given for the various pigments in a section<br />

about oil paint in De volmaakte Verwer, 1770. Amsterdam, appendix.<br />

20. van de Wetering, E. 1977. De jonge Rembr<strong>and</strong>t aan het werk. In Oud Holl<strong>and</strong><br />

91:7-65, especially 20-2. See also van de Wetering, E. 1982. In A corpus of Rembr<strong>and</strong>t<br />

<strong>Painting</strong>s, Vol. I. J. Bruyn, B. Haak, S. H. Levie, P .J. J. van Thiel, E. van de<br />

Wetering. Den Haag, 25-31.<br />

21. Ainsworth, M. W et al. 1982. Art <strong>and</strong> Autoradiography: Insights into the Genesis if<br />

<strong>Painting</strong>s by Rembr<strong>and</strong>t, Van Dijck <strong>and</strong> Vermeer, New York. For activities in the<br />

autoradiographic area by the Gemalde-Galerie <strong>and</strong> the Hahn-Meitner-Instituut<br />

fUr Kernforschung in Berlin, see MaltechniklRestauro I (1987):21 et seq.<br />

22. Taylor, P. 1992. Harmony <strong>and</strong> Illusion in Golden Age Dutch Art Theory: The<br />

Concept of Houding. In Journal if Warburg <strong>and</strong> Courtauld Institutes. (LV) :210-32.<br />

23. Stumpel, J. 1990. The Province of <strong>Painting</strong>. Theories if Italian Renaissance Art.<br />

Utrecht, 131-174 (also published in: Simiolus 1989, 19 [3/4]:219-43).<br />

24. Goeree, W 1670. Inleyding tot d'Algemeene Teykenkonst. (First printing, 1668.)<br />

Middelburg, 219-43.<br />

25. c.f. the article by C. M. Groen in the still unpublished publication in connection<br />

with the restoration of the late Rembr<strong>and</strong>ts in the Rijksmuseum, Amsterdam.<br />

26. "Je mene comprenez un peu, toute ma toile a la fois, d'ensemble." Gasquet, J.<br />

1926. Cezanne. Paris, 89.<br />

van de Wetering 203


Index of Contributors<br />

J. R. J. van Asperen de Boer, 135<br />

Sylvana Barrett, 6<br />

Ulrich Birkmaier, 117<br />

Barbara H. Berrie, 127<br />

Leslie A. Carlyle, 1<br />

Dace Choldere, 155<br />

Josephine A. Darrah, 70<br />

Kate I. DufiY, 78<br />

Jacki A. Elgar, 78<br />

Molly Faries, 135<br />

Beate Federspiel, 58<br />

John R. Gayer, 191<br />

E. Melanie Gifford, 140<br />

Helen Glanville, 12<br />

Stephen Hackney, 186<br />

Erma Hermens, 48<br />

Helen C. Howard, 91<br />

Melissa R. Katz, 158<br />

Jo Kirby, 166<br />

Ann Massing, 20<br />

Catherine A. Metzger, 127<br />

Ilze Poriete, 155<br />

Andrea Rothe, 111, 117<br />

Ashok Roy, 166<br />

Zuzana Skalova, 85<br />

Eddie Sinclair, 105<br />

Christa Steinbiichel, 135<br />

Dusan C. Stulik, 6<br />

Joyce H. Townsend, 176<br />

J0rgen adum, 148<br />

Arie allert, 38, 117<br />

Ernst van de etering, 196<br />

Sally A. 00dcock, 30<br />

Renate oudhuysen-Keller, 65<br />

204<br />

Index oj Contributors


Errata for <strong>Historical</strong> <strong>Painting</strong> <strong>Techniques</strong> PDF<br />

COVER ILLUSTRATION <strong>and</strong> PLATE 8<br />

Gherardo Cibo, “Colchico,” folio 17r of Herbarium, ca. 1570. © British Library Board (ADD MS 22333).<br />

PLATES 7a, 7b <strong>and</strong> PAGE 50, Figure 1<br />

Valerio Mariani de Pesaro, Battaglia di San Fabiano, 1618-1620, GDSU. Images provided by permission<br />

of the Ministero per i Beni e le Attivita Culturali, Florence, Italy.<br />

PLATE 12<br />

Plate printed upside down in relation to Plate 11.<br />

PLATE 17: (c. 50 x 246.5 x 2.5 cm)<br />

PLATE 18: (c. 50 x 212 x 2.5 cm)<br />

PLATE 29e<br />

Plate printed upside down in relation to Plates 29a-d,f.<br />

PAGE 52<br />

Figure 3. Gherardo Cibo, “Hemionite,” folio 143r of Herbarium, ca. 1570. © British Library Board (ADD<br />

MS 22333).<br />

PAGE 53<br />

Figure 4. Gherardo Cibo, “Fusaina [. . .] nocella qui chiamata a Roccha C[ontrada],” folio 183v of<br />

Herbarium, ca. 1570. © British Library Board (ADD MS 22332).<br />

PAGE 53, CONT.


Figure 5. Gherardo Cibo, several proofs of colors <strong>and</strong> “Fusaina” flowers, from folio 184v of Herbarium,<br />

ca. 1570. © British Library Board (ADD MS 22332).<br />

PAGE 85<br />

Paragraph 2. Alfred Butler assessed the beams in 1880, not the 1870s.<br />

PAGE 86<br />

Paragraph 4. The first icon, The Virgin Mary <strong>and</strong> Child enthroned between Archangels, Prophets <strong>and</strong><br />

Egyptian, Greek <strong>and</strong> Syrian Holy Bishops <strong>and</strong> Monks, measures c. 50 x 246.5 x 2.5 cm, <strong>and</strong> is shown in<br />

Plate 17 <strong>and</strong> Figures 3 (left) <strong>and</strong> 4a, b. The second icon, Six Equestrian Saints (originally ten saints),<br />

measures c. 50 x 212 x 2.5 cm, is shown in Plate 18 <strong>and</strong> in Figures 2a,b; 3 (right); <strong>and</strong> 4c-e.<br />

PAGE 89<br />

Acknowledgments. The author…has been working since 1989 as the Field Director of the joint Egyptian-<br />

Netherl<strong>and</strong>s ‘Coptic Icons Conservation Project’ based in the Coptic Museum in Cairo, not as the Field<br />

Director of the Coptic Museum in Cairo as originally published.<br />

Note 3. The correct reference: Butler, A. J. 1884.<br />

PAGE 90<br />

Note 13. The correct reference: MacCoull, L.S.B. 1993. The Apa Apollos Monastery at Pharoou. In Le<br />

Muséon (106), note 11.<br />

Note 20. Abu al-Makarim, a Coptic priest from the church of the Holy Virgin at Harat Zuwayla, Cairo,<br />

left folios in a manuscript dating from around 1200 known as History of the Churches <strong>and</strong> Monasteries in<br />

Lower Egypt in the 13th Century. See Père Samuel du Monastère des Syriens. 1990. Icônes et<br />

iconographie en Égypte au XIIe siècle d’après le manuscrit d’Abū- El-Makārim, publié en Arabe au Caire<br />

en 1984. Le Monde Copte, No. 18. Limoges, 78.<br />

PAGE 115<br />

Figure 3. Mantegna, Andrea (1431-1506). The Presentation in the Temple. Original strainer with panel<br />

inserts seen from front. 68.9 x 86.3 cm. Gemaeldegalerie, Staatliche Museen, Berlin, Germany. Photo<br />

credit: bpk, Berlin/Art Resource, NY.<br />

PAGE 135<br />

Lines 9,12, <strong>and</strong> 14. “M<strong>and</strong>er” should read “van M<strong>and</strong>er.”<br />

PAGE 137<br />

Line 2. “No ground is used” should read “no ground appears in this sample.”<br />

PAGE 158<br />

Figure 1. Image source: Period photo reproduced in O. von Schleinitz, William Holman Hunt (Bielefeld<br />

<strong>and</strong> Leipzig: Verlag von Velhagen & Klasing, 1907).


PAGE 159<br />

Figure 2. Image source: Period photo reproduced in O. von Schleinitz, William Holman Hunt (Bielefeld<br />

<strong>and</strong> Leipzig: Verlag von Velhagen & Klasing, 1907).<br />

PAGE 160<br />

Figure 5. William Holman Hunt, Autoritratto, Galleria degli Uffizi. Image provided by permission of the<br />

Ministero per i Beni e le Attivita Culturali, Florence, Italy.<br />

Figure 6. Unknown Artist. Paint Bladder, 19th century. Oil paint contained in an animal bladder with<br />

three ivory plugs; 4.45 x 2.54 cm (1 3/4 x 1 in.) Harvard Art Museums/Fogg Museum, Straus Center for<br />

Conservation. Gift of C. Roberson <strong>and</strong> Co., London, FINV2155.<br />

Photo: Imaging Department © President <strong>and</strong> Fellows of Harvard College<br />

Brass paint tube, (not assigned). Harvard Art Museums/Fogg Museum, Straus.698<br />

Photo: Imaging Department © President <strong>and</strong> Fellows of Harvard College<br />

Empty paint tube, (not assigned). Harvard Art Museums/Fogg Museum, Straus.696<br />

Photo: Imaging Department © President <strong>and</strong> Fellows of Harvard College<br />

Figure 8. William Holman Hunt. The Miracle of the Sacred Fire, Church of the Holy Sepulchre, 1892-<br />

1899. Mixture of oil <strong>and</strong> resin on canvas; 92.1 x 125.7 cm (36 1/4 x 49 1/2 in.) Framed: 114.5 x 148.3 x 7<br />

cm (45 1/16 x 58 3/8 x 2 3/4 in.) Harvard Art Museums/Fogg Museum, Gift of Grenville L. Winthrop,<br />

Class of 1886, 1942.198.<br />

Photo: Imaging Department © President <strong>and</strong> Fellows of Harvard College<br />

PAGE 169<br />

Figure 3. The Execution of Lady Jane Grey. Pencil drawing for the central group, 18 X 16.5 cm. © The<br />

Trustees of the British Museum.<br />

PAGE 178<br />

Figure 2. J. M. W. Turner, Dolbadern Castle, North Wales, 1802. Oil on canvas, 1190 x 889 mm.<br />

Courtesy of the Royal Academy of Arts, London, www.racollection.org.uk.<br />

PAGE 182<br />

Figure 6. Detail of Anna's veil, painted over clouds <strong>and</strong> sky, from Joshua Reynolds's The Death of<br />

Dido, ca. 1781. Oil on canvas, 1473 X 2407. Supplied by Royal Collection Trust / © HM Queen<br />

Elizabeth II 2012.<br />

PAGE 187<br />

Figure 3. Brushes, paints, palettes, <strong>and</strong> other painting materials. © The Hunterian, University of Glasgow<br />

2012.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!