01.07.2013 Views

Deutsche Geophysikalische Gesellschaft e.V. ... - Bibliothek - GFZ

Deutsche Geophysikalische Gesellschaft e.V. ... - Bibliothek - GFZ

Deutsche Geophysikalische Gesellschaft e.V. ... - Bibliothek - GFZ

SHOW MORE
SHOW LESS

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

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

<strong>Deutsche</strong><br />

<strong>Geophysikalische</strong><br />

<strong>Gesellschaft</strong> e.V.<br />

Protokoll über das<br />

23. Schmucker-Weidelt-Kolloquium für<br />

Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See,<br />

Oliver Ritter<br />

GeoForschungsZentrum<br />

Telegrafenberg<br />

14473 Potsdam<br />

28. September - 2. Oktober 2009<br />

ISSN 0946-7467<br />

herausgegeben von<br />

Ute Weckmann<br />

GeoForschungsZentrum<br />

Telegrafenberg<br />

14473 Potsdam


Widmung<br />

Wir wollen unser Kolloquium Ulrich Schmucker (1930 -2008) und Peter<br />

Weidelt (1938 – 2009) widmen und ihm den Namen „Schmucker-Weidelt-<br />

Kolloquium für Elektromagnetische Tiefenforschung“ geben.<br />

Unser Kolloquium soll Gespräche zwischen allen fördern, die mit<br />

elektromagnetischen Methoden das Innere der Erde erforschen wollen.<br />

Ulrich Schmucker und Peter Weidelt haben in diesem Kolloquium seit<br />

seiner Gründung neue und anspruchsvolle Methoden entwickelt, dort<br />

vorgestellt, oft sogar nur in diesem Kolloquium. Sie haben damit den<br />

internationalen Ruf dieses Kolloquiums begründet und das hohe<br />

intellektuelle Niveau dieser Forschungsrichtung weltweit bestimmt.<br />

Wenn wir dieses Kolloquium, das eigentlich auch ungenannt schon immer<br />

das Schmucker-Weidelt-Kolloquium war, nun auch so nennen wollen, dann<br />

hat das drei tiefe Gründe:<br />

Es sollen diese beiden Wurzeln unseres Kolloquiums und unserer<br />

Forschungsrichtung im Gedächtnis bleiben.<br />

Das Kolloquium soll uns Teilnehmer erinnern, auch in Zukunft die<br />

methodenorientierte Grundlagenforschung und ihre praktische Erprobung<br />

auf hohem Niveau als Ideal und Ziel dieses Kolloquiums anzustreben.<br />

Eine besondere Botschaft von Ulrich Schmucker und Peter Weidelt war es,<br />

gerade junge Studierende der elektromagnetischen Tiefenforschung von<br />

Anfang an in einen ernsthaften und lebhaften Kontakt mit allen<br />

Teilnehmern einzubinden, ihnen zuzuhören und zu helfen, sie zu kritisieren<br />

und Neues von ihnen zu lernen. Wenn wir diese Botschaft auch in Zukunft<br />

zum Arbeitsstil unseres Kolloquiums machen werden, wird die Entwicklung<br />

der elektromagnetischen Tiefenforschung so voller Dynamik bleiben, wie<br />

sie es bisher war.


I<br />

Vorwort<br />

Mit Erscheinen dieses Bandes und im Einvernehmen mit den Hinterbliebenen wollen wir von nun an<br />

das Kolloquium Prof. Ulrich Schmucker (1930 -2008) und Prof. Peter Weidelt (1938 – 2009) widmen.<br />

Den Wortlaut der Widmung haben wir an den Beginn des Bandes gestellt. Beide waren für die<br />

Entwicklung der elektromagnetischen Tiefenforschung weltweit von herausragender Bedeutung. Ihre<br />

Persönlichkeit und Schaffenskraft hatten aber vor allem in Deutschland einen maßgeblichen Anteil an<br />

der überaus positiven Entwicklung, die die Elektromagnetische Tiefenforschung genommen hat.<br />

Veranstalter des 23. Kolloquiums war das <strong>Deutsche</strong> GeoForschungsZentrum - <strong>GFZ</strong>. Der Tagungsort,<br />

die Heimvolkshochschule am Seddiner See, befand sich in unmittelbarer Nähe zur<br />

brandenburgischen Landeshauptstadt Potsdam, direkt am Großen Seddiner See, in einer für<br />

Brandenburg typischen Umgebung mit vielen Seen, Laub- und Kiefernwäldern. Am Kolloquium hatten<br />

sich 88 Teilnehmer angemeldet, größtenteils aus dem deutschsprachigen Raum, aber wie auch schon<br />

in den Jahren zuvor gab es auch diesmal eine ganze Reihe von Anmeldungen aus dem europäischen<br />

Ausland. Insgesamt waren 15 deutsche und 8 europäische Institutionen vertreten.<br />

Im Gedenken an Ulrich Schmucker hatten seine ehemaligen Schüler bereits ein Kolloquium vom 26 -<br />

28 Juni 2009 im Herz-Jesu Kloster in Neustadt ausgerichtet. Peter Weidelt wurde in einer Reihe von<br />

Beiträgen im Rahmen des 23. Kolloquiums gedacht und besonders gefreut hat uns, dass auch Heidi,<br />

Anne und Jan Weidelt anwesend waren. Im vorliegenden Tagungsband sind die meisten dieser<br />

Beiträge protokolliert. Allerdings haben wir uns entschlossen, sie nicht besonders hervorzuheben,<br />

sondern sie wie immer thematisch einzuordnen.<br />

Ulrich Schmucker und Peter Weidelt hatten von Beginn an in diesem Kolloquiumsband publiziert, oft<br />

sogar nur in diesem. Der diesjährige Band enthält 32 Beiträge. Es gab die übliche große Bandbreite an


Beiträgen die theoretisch- methodische Aspekte und experimentelle Untersuchungen im Hinblick auf<br />

geodynamische und angewandte Fragestellungen abdeckten. Generell war vielleicht eine Zunahme<br />

der mehr angewandten Arbeiten festzustellen, was sicherlich auch darauf zurückzuführen ist, dass<br />

die elektromagnetischen Verfahren mittlerweile fester Bestandteil der Exploration im off-shore aber<br />

auch on-shore Bereich geworden sind.<br />

Quicklebendige Debatten gab es auch im Rahmen unserer neuen Rubrik „Was Sie schon immer über<br />

die Elektromagnetik wissen wollten, sich bisher aber nicht zu fragen trauten.“, die Ute Weckmann<br />

angeregt hatte. Insbesondere sollten damit die jüngeren Teilnehmer angesprochen werden, jedoch<br />

hatten offensichtlich besonders viele der Ü40 Teilnehmer noch jahrelang unbeantwortet gebliebene<br />

Fragen. Manchmal war für die Beantwortung der Fragen auch etwas mehr Zeit zum Nachdenken<br />

nötig, weshalb wir eine nach dem Kolloquium entstandene Antwort in diesen Band ebenfalls<br />

aufgenommen haben.<br />

Für die finanzielle Unterstützung danken wir auch unseren Sponsoren: den Firmen KMS und<br />

Metronix. Herzlich bedanken möchten wir uns bei Frau Hannelore Gendt (<strong>GFZ</strong>) für die Hilfe bei<br />

administrativen und finanziellen Aufgaben. Herrn Palm und dem Daten- und Rechenzentrum sei für<br />

die Einrichtung des Internetportals gedankt. Roxana Barth und Gregor Willkommen haben für den<br />

reibungslosen Ablauf während des Kolloquiums gesorgt und auch tatkräftig bei der<br />

Zusammenstellung dieses Bandes geholfen. Natürlich ließ sich ein solches Kolloquium nur mit der<br />

Unterstützung der gesamten MT Arbeitsgruppe des <strong>GFZ</strong> durchführen. Ein herzliches Dankeschön<br />

dafür; ebenso an Herrn Bertelmann von der <strong>Bibliothek</strong> für das Hosting der Protokollbände.<br />

Oliver Ritter und Ute Weckmann<br />

Potsdam, 6. Juni 2010


II<br />

Inhaltsverzeichnis / table of contents<br />

Rita Streich & Michael Becken, EM fields generated by finite-length wire sources in 1D media:<br />

comparison with point dipole solutions ............................................................................................1<br />

M. Afanasjew, R.-U. Börner, M. Eiermann, O. G. Ernst, S. Güttel & K. Spitzer, 2D Time Domain TEM<br />

Simulation Using Finite Elements, an Exact Boundary Condition, and Krylov Subspace Methods 16<br />

Tilman Hanstein, TEM with anomalous diffusion in fractal conductive media. .................................... 25<br />

K. M. Bhatt, A. Hördt & T. Hanstein, Analysis of seafloor marine EM data with respect to motioninduced<br />

noise ................................................................................................................................ 33<br />

K. M. Bhatt, A. Hördt, P.Weidelt & T. Hanstein, Motionally Induced Electromagnetic Field within the<br />

Ocean................... .......................................................................................................................... 46<br />

S. Kütter, A. Franke-Börner, R.-U. Börner & K. Spitzer, Three-dimensional finite element simulation of<br />

magnetotelluric fields incorporating digital elevation models ...................................................... 60<br />

M. Becken, R. Streich & O. Ritter, Establishing Controlled Source MT at <strong>GFZ</strong> ...................................... 71<br />

Johannes Kenkel, Andreas Hördt & Andreas Kemna, 2D-SIP-Modellierung mit anisotropen<br />

Widerständen. ................................................................................................................................ 77<br />

Xiaoming Chen, Ute Weckmann, Kristina Tietze, Towards a 2D anisotropic inversion. ....................... 88<br />

Kristina Tietze, Oliver Ritter & Ute Weckmann, Substitute models for static shift in 2D. ..................... 97<br />

Andreas Hördt, Peter Weidelt & Anita Przyklenk, Die Übergangsimpedanz einer kapazitiv<br />

angekoppelten Elektrode ............................................................................................................. 102<br />

Häuserer & Junge, Eine Methode gegen auslaufende Ag/AgCl//KCl(H2O) Elektroden ...................... 115<br />

Johannes B. Stoll, Celle, Christopher Virgil, Attitude Algorithm Utilised in Mobile Geophysical<br />

Measuring Systems ...................................................................................................................... 120<br />

G. Muñoz, K. Bauer, I. Moeck, O. Ritter, Joint Interpretation of Magnetotelluric and Seismic Models<br />

for Exploration of the Gross Schoenebeck Geothermal Site ......................................................... 126<br />

Ulrich Kalberkamp, Magnetotelluric measurements to explore for deeper structures of the Tendaho<br />

geothermal field, Afar, NE Ethiopia.. ............................................................................................ 131<br />

J. H. Börner, V. Herdegen, R.-U. Börner, K. Spitzer, Electromagnetic Monitoring of CO2 Storage in Deep<br />

Saline Aquifers - Numerical Simulations and Laboratory Experiments. ....................................... 137<br />

K. Lippert, B. Tezkan, R. Bergers, M. Gurk, M. v. Papen, P. Yogeshwar, Erkundung eines Aquifers unter<br />

dem Mittelmeer vor der israelischen Küste mit LOTEM. .............................................................. 143<br />

M. von Papen, B. Tezkan, On the analysis of LOTEM time series from Israel and the preliminary 1D<br />

inversion of data ........................................................................................................................... 149<br />

P. Yogeshwar, B. Tezkan, M. Israil, Grundwasserkontamination bei Roorkee/Indien: 2D Joint Inversion<br />

von Radiomagnetotellurik und Gleichstromgeoelektrik Daten .................................................... 153<br />

Widodo, Marcus Gurk, Bülent Tezkan, Site Effect Assessment in the Mygdonian Basin (EUROSEISTEST<br />

area, Northern Greece) using RMT and TEM Soundings .............................................................. 164<br />

Gerlinde Schaumann, Annika Steuer, Bernhard Siemon, Helga Wiederhold & Franz Binot, Die<br />

deutsche Nordseeküste im Fokus von aeroelektromagnetischen Untersuchungen Teilgebiete<br />

Langeoog mit Wattenmeer und Elbemündung ............................................................................ 177


Annika Steuer, Bernhard Siemon and Michael Grinat, The German North Sea Coast in Focus of<br />

Airborne Electromagnetic Investigations: The Freshwater Lenses of Borkum ............................. 188<br />

Gerhard Kapinos & Heinrich Brasse, Some notes on bathymetric effects in marine magnetotellurics,<br />

motivated by an amphibious experiment at the South Chilean margin ...................................... 198<br />

Dirk Brändlein, Oliver Ritter, Ute Weckmann, A permanent array of magnetotelluric stations located<br />

at the South American subduction zone in Northern Chile. ......................................................... 210<br />

D. Eydam & H. Brasse, Discussion on backarc mantle melting in the central Andean subduction zone,<br />

based on results of magnetotelluric studies ................................................................................. 216<br />

Václav Červ, Světlana Kováčiková, Michel Menvielle and Josef Pek, Thin Sheet Conductance Models<br />

from Geomagnetic Induction Data: Application to Induction Anomalies at the Transition from the<br />

Bohemian Massif to the West Carpathians .................................................................................. 232<br />

Anne Neska, Subsurface Conductivity Obtained from DC Railway Signal Propagation with a Dipole<br />

Model ........................................................................................................................................... 244<br />

Anja Schäfer, Heinrich Brasse, Norbert Hoffmann and EMTESZ working group, Magnetotelluric<br />

investigation of the Sorgenfrei-Tornquist Zone and the NE German Basin ................................. 252<br />

P. Sass, O. Ritter, A. Rybin, G. Muñoz, V. Batalev and M. Gil, Magnetotelluric data from the Tien Shan<br />

and Pamir continental collision zones, Central Asia ..................................................................... 263<br />

Alexander Löwer, Audiomagnetotellurik im Hohen Vogelsberg ......................................................... 269<br />

Ute Weckmann, Carsten Scholl, Dumme Fragen ................................................................................ 277


III<br />

Teilnehmerverzeichnis<br />

YassineAbdelfettah LaboratoiredeGéothermie,Neuchâtel,Suisse<br />

FilipeAdão CentrodeGeofísicadaUniversidadedeLisboa,Lisboa,Portugal<br />

JulianeAdrian UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

MartinAfanasjew TUBergakademieFreiberg<br />

KatharinaBairlein TUBraunschweig<br />

RoxanaBarth UniversitätPotsdam<br />

MichaelBecken <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

K.MangalBhatt TUBraunschweig<br />

DirkBrändlein <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

AnneBublitz J.W.GoetheUniversität,FrankfurtamMain<br />

MatthiasBücker TUBraunschweig<br />

XiaomingChen <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

JinChen IFMGEOMAR,Kiel<br />

DanielDiaz FUBerlin<br />

SebastianEhmann TUBraunschweig<br />

DianeEydam <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

JohannesGeiermann igem,Bingen<br />

MichaelGrinat LeibnizInstitutfürAngewandteGeophysik,Hannover(LIAG)<br />

HannahGroßbach UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

MarkusGurk UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

VolkerHaak Blankenfelde<br />

TilmanHanstein KMSTechnologiesGmbH,Köln<br />

MichaelHäuserer J.W.GoetheUniversität,FrankfurtamMain<br />

BjörnHenningHeincke IFMGEOMAR,Kiel<br />

WiebkeHeise CentrodeGeofísicadaUniversidadedeLisboa,Lisboa,Portugal<br />

StefanHendricks AlfredWegenerInstitut,Bremerhaven<br />

PaulHofmeister TUBraunschweig<br />

SebastianHölz IFMGEOMAR,Kiel


AndreasHördt TUBraunschweig<br />

AndreasJunge J.W.GoetheUniversität,FrankfurtamMain<br />

UlrichKalberkamp FederalInstituteforGeosciencesandNaturalResources,Hannover(BGR)<br />

ThomasKalscheuer ETHZurich,Zurich,Switzerland<br />

JochenKamm UppsalaUniversity,DepartmentofEarthSciences,Uppsala,Sweden<br />

GerhardKapinos FUBerlin<br />

JohannesKenkel TUBraunschweig<br />

JudithKirchner TUBergakademieFreiberg<br />

FlorianLePape DIAS,Dublin,Ireland<br />

JochenLehmannHorn ETHZurich,Zurich,Switzerland<br />

MartinLeven InstitutfürGeophysik,UniversitätGöttingen<br />

ShengjunLiang TUBergakademieFreiberg<br />

MaikLinke TUBergakademieFreiberg<br />

KlausLippert UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

AlexanderLöwer J.W.GoetheUniversität,FrankfurtamMain<br />

StephanMalecki TUBergakademieFreiberg<br />

EricMandolesi DIAS,Dublin,Ireland<br />

NaserMeqbel <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

MarionMiensopust DIAS,Dublin,Ireland<br />

MaxMoorkamp IFMGEOMAR,Kiel<br />

DanaMoritz TUBergakademieFreiberg<br />

GerardMuñoz <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

LutzMütschard FUBerlin<br />

AnneNeska InstituteofGeophysics,PolishAcad.Sci.,CentralneObserwatorium<br />

Geofizyczne,BelskDuzy,Poland<br />

LaustB.Pedersen UppsalaUniversity,DepartmentofEarthSciences,Uppsala,Sweden<br />

JosefPek InstituteofGeophysics,Acad.Sci.CzechRep.,Prague,CzechRepublic<br />

AnitaPrzyklenk TUBraunschweig<br />

LasseRabenstein AWIBremerhaven<br />

TinoRadic RadicResearch,Berlin<br />

ZhengyongRen ETHZurich,Zurich,Switzerland<br />

OliverRitter <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong>


AnnikaRödder UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

EstelleRoux DIAS,Dublin,Ireland<br />

AnjaSchäfer FUBerlin<br />

GerlindeSchaumann LeibnizInstitutfürAngewandteGeophysik,Hannover(LIAG)<br />

CarstenScholl FugroElectroMagnetic,Berlin<br />

KatrinSchwalenberg FederalInstituteforGeosciencesandNaturalResources,Hannover(BGR)<br />

KlausSpitzer TUFreibergTUBergakademieFreiberg<br />

AnnikaSteuer LeibnizInstitutfürAngewandteGeophysik,Hannover(LIAG)<br />

JohannesStoll Celle<br />

KurtStrack KMSTechnologies,HoustonTX,USA<br />

RitaStreich <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

BülentTezkan UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

KristinaTietze <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

AndreaTreichel WestfälischeWilhelmsUniversität,Münster<br />

ChristopherVirgil TUBraunschweig<br />

MichaelvonPapen UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

UteWeckmann <strong>Deutsche</strong>sGeoForschungsZentrumPotsdam<strong>GFZ</strong><br />

JuliaWeißflog TUBergakademieFreiberg<br />

Widodo UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

GregorWillkommen UniversitätPotsdam<br />

HelmuthWinter J.W.GoetheUniversität,FrankfurtamMain<br />

TamaraWorzewski IFMGEOMAR,Kiel<br />

PritamYogeshwar UniversitätzuKöln,InstitutfürGeophysikundMeteorologie<br />

FanZhang TUBergakademieFreiberg


EM fields generated by finite-length wire sources in 1D<br />

media: comparison with point dipole solutions<br />

Rita Streich ∗† and Michael Becken ∗† ,<br />

∗ Potsdam University, Institute of Geosciences,<br />

Karl-Liebknecht-Str. 24, 14476 Potsdam-Golm, Germany<br />

† <strong>GFZ</strong> German Research Centre for Geosciences,<br />

Telegrafenberg, 14473 Potsdam, Germany<br />

ABSTRACT<br />

In present-day land and marine controlled-source electromagnetic (CSEM) surveys, EM<br />

fields are commonly generated using wires that are hundreds of meters long. Nevertheless,<br />

simulations of CSEM data often approximate these sources as point dipoles.<br />

Although this is justified for sufficiently large source-receiver distances, many real surveys<br />

include frequencies and distances at which the dipole approximation is inaccurate.<br />

For 1D layered media, EM fields for point dipole sources can be computed using<br />

well-known quasi-analytical solutions, and fields for sources of finite lengths can be<br />

synthesized by superposing point dipole fields. However, the calculation of numerous<br />

point dipole fields is computationally expensive, requiring a large number of numerical<br />

integral evaluations. We combine a more efficient representation of finite-length sources<br />

in terms of components related to the wire and its end points with very general expressions<br />

for EM fields in 1D layered media. We thus obtain a formulation that requires<br />

fewer numerical integrations than the superposition of dipole fields and permits source<br />

and receiver placement at any depth within the layer stack. Complex source geometries,<br />

such as wires bent due to surface obstructions, can be simulated by segmenting<br />

the wire and computing the responses for each segment separately. We first describe<br />

our finite-length wire expressions, and then present examples of EM fields due to finitelength<br />

wires for typical land and marine survey geometries and discuss differences to<br />

point dipole fields.<br />

INTRODUCTION<br />

Controlled-source electromagnetic (CSEM) surveys are a useful exploration tool, applicable,<br />

e.g., for exploring hydrocarbon reservoirs, geothermal reservoirs, or for characterizing<br />

and potentially monitoring sites considered for carbon sequestration. A multitude of active<br />

electromagnetic sources are available, including magnetic loops (Frischknecht et al., 1991;<br />

Spies and Frischknecht, 1991) and long wires, grounded in land-based surveys (Strack, 1992;<br />

Wright et al., 2002; Ziolkowski et al., 2007) and deployed at the seafloor (Edwards, 2005) or<br />

towed through the water (e.g., Constable and Srnka, 2007) in marine surveys. Wire sources<br />

with lengths of several 100 m are most commonly used in commercial hydrocarbon exploration<br />

because of their capability to generate three-dimensional electrical current systems<br />

sensitive to both resistive and conductive, relatively deep targets (Spies and Frischknecht,<br />

1991; Constable and Srnka, 2007).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

1


In processing, inversion and interpretation of CSEM data, the sources are commonly<br />

approximated as point dipoles (e.g., Edwards, 1997; Johansen et al., 2005; Ziolkowski et al.,<br />

2007). This is adequate for sufficiently large source-receiver distances. However, real surveys<br />

targeting increasingly deep and small structures may include distances and frequencies at<br />

which the inaccuracy of the dipole approximation is on the order of (or larger than) the<br />

target-related anomalies. In such cases, it is vital to consider the actual source geometry.<br />

For horizontally layered media, EM fields can be computed using well-known quasianalytical<br />

solutions that involve numerical evaluations of Bessel function integrals (e.g.,<br />

Weidelt, 2007; Løseth and Ursin, 2007). Using such point dipole solutions, EM fields due<br />

to finite-length wire sources can be synthesized by representing the wire as a line of point<br />

dipoles and summing the dipole fields. However, this procedure is computationally expensive,<br />

requiring many numerical integrations to calculate all of the point dipole contributions.<br />

To improve the efficiency of long-wire source simulations, Soerensen and Christensen<br />

(1994) derived integrated Hankel integral expressions for finite-length wires that barely<br />

require more integral evaluations than the computation of point dipole fields. However, their<br />

approach includes elaborate computations of filter coefficients for every source-receiver pair.<br />

Another approach that reduces the number of integrations from that required for summing<br />

dipole fields is the separation of EM fields due to long wires into contributions from the wire<br />

and its end points (Ward and Hohmann, 1987). This technique works with standard Hankel<br />

filters. Whereas previous application of this approach considered subsurface receivers and<br />

sources located at the air-ground interface (Ward and Hohmann, 1987), we have applied<br />

it to a more general 1D field formulation that permits source and receiver positions at<br />

arbitrary depths within the layer stack (Løseth and Ursin, 2007). This allows simulations<br />

of finite marine sources in addition to land sources.<br />

In this contribution, we first describe our finite-length wire representation. We then show<br />

examples of EM fields due to finite-length wires in representative land and marine settings<br />

over the frequency and distance ranges of typical CSEM surveys, and discuss differences to<br />

the fields due to infinitesimal dipoles.<br />

EM FIELD EXPRESSIONS FOR FINITE-LENGTH WIRES<br />

Expressions for the electromagnetic field due to a finite-length wire can be obtained by<br />

representing the wire as a line of infinitesimal dipoles and integrating the dipole expressions<br />

along the length of the wire. Using the nomenclature of Løseth and Ursin (2007), the EM<br />

fields for an x-directed point dipole embedded in a 1D layered medium can be expressed in<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

2


a form convenient for later integration as (see Appendix A)<br />

Ex = Idx<br />

⎛<br />

∞<br />

−μ0 ⎝ √<br />

4π pr zp 0<br />

s R<br />

z<br />

A ⎧<br />

⎨<br />

x<br />

11κJ0(κr)dκ+∂x<br />

⎩<br />

r−xs∞ <br />

μ0<br />

√<br />

r pr zp 0<br />

s R<br />

z<br />

A <br />

pr zp 11−<br />

s z<br />

˜ε r ˜ε s RA ⎫⎞<br />

⎬<br />

22 J1(κr)dκ ⎠, (1a)<br />

⎭<br />

Ey = Idx<br />

4π ∂x<br />

⎧<br />

⎨<br />

y<br />

⎩<br />

r − ys∞ <br />

μ0<br />

√<br />

r pr zp<br />

0<br />

s R<br />

z<br />

A <br />

pr zp 11 −<br />

s z<br />

˜ε r ˜ε s RA ⎫<br />

⎬<br />

22 J1(κr)dκ , (1b)<br />

⎭<br />

Ez = jIdx<br />

⎧<br />

⎨∞<br />

˜ε<br />

∂x<br />

4πω˜ε r ⎩<br />

0<br />

r ˜ε s<br />

pr zps R<br />

z<br />

D 22κJ0(κr)dκ ⎫<br />

⎬<br />

, (1c)<br />

⎭<br />

Hx = Idx<br />

4π ∂x<br />

⎧<br />

⎨<br />

y<br />

⎩<br />

r − ys ∞<br />

<br />

p<br />

r<br />

0<br />

r z<br />

ps R<br />

z<br />

D 11 −<br />

<br />

˜ε rps z<br />

˜ε spr R<br />

z<br />

D ⎫<br />

⎬<br />

22 J1(κr)dκ , (1d)<br />

⎭<br />

Hy = Idx<br />

⎛<br />

∞<br />

⎝<br />

p<br />

4π<br />

r z<br />

ps R<br />

z<br />

D 11κJ0(κr)dκ ⎧<br />

⎨<br />

x<br />

− ∂x<br />

⎩<br />

r−x s ∞<br />

p<br />

r<br />

r z<br />

ps R<br />

z<br />

D 11− <br />

˜ε rps z<br />

˜ε spr R<br />

z<br />

D ⎫⎞<br />

⎬<br />

22 J1(κr)dκ ⎠, (1e)<br />

⎭<br />

Hz = jIdx<br />

4πωμ0<br />

0<br />

yr − ys <br />

r<br />

0<br />

∞<br />

0<br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A 11κ2J1(κr)dκ, (1f)<br />

where I is the source current, dx is the length of the source dipole, κ is the horizontal<br />

wavenumber, ω is the angular frequency, μ0 is the vacuum magnetic permeability,<br />

˜ε {s,r} = ε {s,r} + jσ {s,r} /ω are the dielectric permittivity and electric conductivity of the<br />

source and receiver layer, respectively, p {s,r}<br />

z<br />

= μ0˜ε {s,r} − κ 2 /ω 2 are frequency-normalized<br />

vertical wavenumbers, r = (x r − x s ) 2 +(y r − y s ) 2 is the horizontal source-receiver distance,<br />

and J0 and J1 are the zero- and first-order Bessel functions. R A 11 , RA 22 , RD 11 and RD 22<br />

are the reflection responses of the layered medium as given by Løseth and Ursin (2007, their<br />

Equation 134). These quantities are computed recursively using the properties and thicknesses<br />

of all layers and the source and receiver depths. Subscripts 11 denote the TE-mode<br />

and subscripts 22 the TM-mode responses. Slightly different expressions apply for receivers<br />

below (Equations 134a and b) and above the source (Equations 134c and d of Løseth and<br />

Ursin, 2007). In Equations (1), a spatial derivative ∂x has been retained wherever possible.<br />

To obtain electromagnetic fields for a finite-length wire, we integrate Equations (1) over<br />

thewirelength. Assumingthatthewireisparalleltothex-axis, we express the integrals as<br />

discrete sums over N wire elements of length Δx, located at (xn,y s ,z s ). The derivatives ∂x<br />

are replaced by −∂/∂Δx (Ward and Hohmann, 1987). Then the fields at receiver location<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

3


(x r ,y r ,z r )are<br />

Ex = I<br />

4π<br />

− I<br />

4π<br />

Ey = − I<br />

4π<br />

N<br />

<br />

Δx<br />

n=1<br />

2<br />

m=1<br />

0<br />

∞<br />

−κμ0<br />

√<br />

pr zps R<br />

z<br />

A 11J0(κrn)dκ<br />

(−1) m x r − xm<br />

rm<br />

∞<br />

0<br />

∞<br />

2<br />

(−1) m yr − ys <br />

m=1<br />

Ez = − jI<br />

4πω˜ε r<br />

Hx = − I<br />

4π<br />

Hy = I<br />

4π<br />

+ I<br />

4π<br />

rm<br />

2<br />

(−1) m<br />

<br />

m=1<br />

0<br />

∞<br />

<br />

2<br />

(−1) m yr − ys m=1<br />

N<br />

<br />

Δx<br />

n=1<br />

2<br />

m=1<br />

Hz = jI(yr − y s )<br />

4πμ0ω<br />

0<br />

∞<br />

<br />

pr z<br />

ps z<br />

rm<br />

(−1) m x r − xm<br />

rm<br />

N Δx<br />

n=1<br />

rn<br />

0<br />

<br />

<br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A 11 −<br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A 11 −<br />

p r zp s z<br />

˜ε r ˜ε r RA 22<br />

p r zp s z<br />

˜ε r ˜ε r RA 22<br />

<br />

J1(κrm)dκ, (2a)<br />

<br />

J1(κrm)dκ, (2b)<br />

˜ε rps z<br />

˜ε spr R<br />

z<br />

D 22κJ0(κrm)dκ, (2c)<br />

∞<br />

0<br />

<br />

pr z<br />

ps z<br />

R D 11κJ0(κrn)dκ<br />

∞<br />

0<br />

∞<br />

0<br />

<br />

pr z<br />

ps z<br />

R D 11 −<br />

<br />

˜ε rps z<br />

˜ε spr R<br />

z<br />

D <br />

22 J1(κrm)dκ, (2d)<br />

R D <br />

11 −<br />

˜ε rps z<br />

˜ε spr R<br />

z<br />

D <br />

22 J1(κrm)dκ, (2e)<br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A 11κ 2 J1(κrn)dκ. (2f)<br />

For Ex, Hy and Hz, we obtain a contribution from each wire element. This contribution<br />

is described by the first sum in Equations 2a, 2e and 2f, with the distance between the<br />

receiver and the n th wire element given by rn = (x r − xn) 2 +(y r − y s ) 2 .Uponintegrating<br />

those terms of Equations (1) that contain derivatives ∂x, the derivatives disappear, and we<br />

obtain explicit contributions from the integration limits, i.e., the end points of the wire. In<br />

Equations (2), the end point contributions are given by the summation over m (m ∈{1, 2}),<br />

with the distances between the receiver and the wire ends denoted rm.<br />

To compute all EM field components for a finite-length wire using Equation (2), we<br />

have to evaluate three different integrals over the entire wire length. Accordingly, for a<br />

wire discretized into N elements, the number of numerical integral evaluations is approximately<br />

3N. In contrast, Equation (A-7) shows that two different integrals are required<br />

for computing only Ex for an infinitesimal dipole. The computation of all electromagnetic<br />

field components for a horizontal electric point dipole source requires numerical evaluations<br />

of eight different integrals (Løseth and Ursin, 2007). This would result in 8N numerical<br />

integrations when calculating finite-length wire fields from the contributions of N dipole<br />

elements. Compared to the simple summation of dipole fields, the separate computation of<br />

wire and end point contributions thus reduces the computational effort by more than 60%.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

4


EXAMPLES<br />

We use the expressions of EM fields due to finite-length wires given in Equation (2) to assess<br />

differences between the fields of finite sources and point dipoles for different representative<br />

experimental settings of land and marine surveys.<br />

Land survey: straight grounded wire<br />

We have simulated a land survey using the setup depicted in Figure 1. The electric conductivity<br />

model is based on the situation at the CO2 sequestration pilot site in Ketzin,<br />

Germany, using measured conductivity values and the actual depth and thickness of the<br />

layer into which CO2 is being injected (Giese et al., 2009). To simulate a realistic survey,<br />

in which sensors would typically be buried just below the surface, we placed electric and<br />

magnetic field receivers at a depth of 0.15 m. EM fields were computed for an infinitesimal<br />

dipole source at (x, y) =(0, 0) and a 1-km long finite-length wire centered at (x, y) =(0, 0).<br />

Both sources were located at a depth of 0.1 m.<br />

surface<br />

source,<br />

centered<br />

at (0,0)<br />

air<br />

1/3<br />

0.1<br />

1 S/m<br />

10 km<br />

surface<br />

receivers<br />

15 km<br />

0<br />

635<br />

650<br />

z (m)<br />

Figure 1: The 1D conductivity model and survey geometry used for simulating a land CSEM<br />

survey. The red arrow indicates the point dipole or 1000-m long wire source, centered at<br />

(x, y) =(0, 0) and located 0.1 m below the surface. Receivers are located 0.15 m below the<br />

surface.<br />

For convenience in the numerical simulations, we place the entire wire at a constant<br />

depth, either above or below the air-ground interface, although in real field surveys, grounded<br />

wires would typically be used, with the wire laid out on the surface and the end points coupled<br />

into the ground, e.g., via metal electrodes. The contribution from the wire body is<br />

typically smaller than that from the grounding points, and varies slowly as the wire depth<br />

crosses interfaces. Therefore, this simplification does not cause noticeable errors.<br />

In Figure 2, we display the electric field component Ex at a frequency of 0.1 Hz for<br />

the model depicted in Figure 1. The finite-wire and point dipole fields differ significantly<br />

withinaradiusof∼ 4 km from the source. Large relative differences also occur in the<br />

lowest-amplitude regions at oblique angles to the source; however, these are insignificant,<br />

because in these regions, Ex would not be measurable.<br />

For comparison, we show in Figure 3 the electric field for the reservoir model of Fig-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

5


y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

(a)<br />

0 3 6 9<br />

x (km)<br />

y (km)<br />

−6<br />

−3<br />

−14 −12 −10 −8<br />

log10( | Ex [V/m] |)<br />

0<br />

3<br />

6<br />

(b)<br />

0 3 6 9<br />

x (km)<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

(c)<br />

0 3 6 9<br />

x (km)<br />

| Ex wire / Ex dipole 0.8 1 1.2<br />

|<br />

Figure 2: Electric field Ex for the configuration shown in Figure 1, a frequency of 0.1 Hz<br />

and (a) a point dipole source at (0, 0, 0.1) m and (b) a 1000-m long grounded wire extending<br />

from (−500, 0, 0.1) to (500, 0, 0.1) m. (c) shows the ratio between the finite-length wire and<br />

point dipole fields.<br />

ure 1 relative to the electric field for a model that does not contain a resistive layer. This<br />

demonstrates the size of the anomaly we would attempt to detect in the CSEM survey. The<br />

reservoir-related anomaly is smaller than the relative differences between finite-length wire<br />

and dipole fields, and overlaps spatially with the region in which the response is significantly<br />

influenced by the source geometry. This clearly indicates the importance of considering the<br />

true source geometry.<br />

Similar observations can be made over a wide frequency range. In Figure 4, we compare<br />

the electric field Ex for finite-length and point dipole sources, and for the reservoir and<br />

background models, at frequencies of 0.001 Hz and 1 Hz. Significant differences between<br />

finite-length and point dipole sources occur in a similar region as for f =0.1 Hz(compare<br />

Figures 4a and b to Figure 2c). The reservoir-related anomalies are somewhat smaller than<br />

for f =0.1 Hz, underlining again that it is vital to consider the actual source length when<br />

searching for such relatively small anomalies.<br />

In Figure 5, we display the finite-length wire fields for the Ey and Ez components,<br />

and their ratios to the respective point dipole fields. Here, the relative differences between<br />

the finite-length wire and point dipole fields are of similar size as for Ex. The regions in<br />

which the responses of finite-length wires and point dipoles differ significantly are similar,<br />

or slightly larger, in extent than for Ex. However, the amplitudes of Ez are considerably<br />

smaller than those of Ex and Ey, such that nearly the entire region in which Ez would<br />

be measurable (assuming instrument detection thresholds of ∼ 10 −14 − 10 −15 , and possible<br />

measurement of Ez in shallow boreholes) is strongly influenced by the source geometry.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

6


y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

0 3 6 9<br />

x (km)<br />

Figure 3: Electric field Ex for the configuration shown in Figure 1, a 1000-m long wire source<br />

and f =0.1 Hz, normalized by the electric field for a background model not containing a<br />

resistive layer.<br />

Required wire discretization<br />

To minimize the computational effort, we have empirically investigated the effects of wire<br />

discretization on the resulting electromagnetic field values. In Figure 6, we display differences<br />

between a reference field Ex computed for a 1000-m long wire discretized using an<br />

element length of 0.1 m, and more coarsely discretized wires with element lengths varying<br />

between 1 m and 50 m. As expected, differences between the reference field and fields computed<br />

using coarser discretizations increase as the element length increases. Nevertheless,<br />

at the lower frequency of 0.1 Hz (Figure 6a), differences are small for all tested element<br />

lengths. As expected, differences are larger at f = 100 Hz, with maximum differences<br />

occurring in the vicinity of the end point of the wire at x = 500 m.<br />

From this test, we conclude that for the lower frequency, a coarse discretization of ∼<br />

20−50 m would be sufficient, whereas for the higher frequency, the wire should be discretized<br />

using elements no longer than ∼ 2 − 5 m. These results may serve as rough guidelines for<br />

further simulation studies, but actual required discretizations are likely to depend on the<br />

total wire length and the resistivity model. To exclude any discretization-related error,<br />

all results presented here were computed using somewhat too careful discretizations with<br />

element lengths of 1 m.<br />

Non-straight grounded wire<br />

In real field surveys, surface obstacles may preclude laying out the source wire in a straight<br />

line. We have therefore studied differences between the EM fields for straight and nonstraight<br />

wire sources. EM fields for non-straight wires are calculated by segmenting the<br />

wire into straight sections, and computing the responses for each segment separately using<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

7<br />

|Ex reservoir / Ex background |


y (km)<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

−6<br />

−3<br />

(a)<br />

0 3 6 9<br />

x (km)<br />

0<br />

3<br />

6<br />

(c)<br />

0 3 6 9<br />

x (km)<br />

y (km)<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

−6<br />

−3<br />

(b)<br />

0 3 6 9<br />

x (km)<br />

0<br />

3<br />

6<br />

0 3 6 9<br />

x (km)<br />

Figure 4: Ratio of Ex for a 1000-m long wire source to Ex for a point dipole for frequencies<br />

of (a) 0.001 Hz and (b) 1 Hz, and ratio of Ex for a 1000-m long wire source and the reservoir<br />

model shown in Figure 1 relative to Ex for a background model not containing the resistive<br />

layer for frequencies of (c) 0.001 Hz and (d) 1 Hz.<br />

Equations (2) with appropriate rotations. As an example, we have computed the responses<br />

for a wire that consists of two segments arranged at a 120 ◦ angle. This wire has the same<br />

grounding points as the 1000-m long straight wire used previously. Figure 7 shows the wire<br />

geometry and the electric field Ex for the bent wire relative to Ex for a straight wire at<br />

frequencies of 0.001 Hz, 0.1 Hz and 1 Hz.<br />

As expected, the Ex amplitudes for the bent wire are increased relative to the straight<br />

wire field at the side to which the wire is deviated (y >0), and decreased at y


y (km)<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

−6<br />

−3<br />

(a)<br />

0 3 6 9<br />

x (km)<br />

0<br />

3<br />

6<br />

(c)<br />

0 3 6 9<br />

x (km)<br />

−6<br />

−8<br />

−10<br />

−12<br />

−14<br />

−10<br />

−12<br />

−14<br />

−16<br />

−18<br />

log10(Ey [V/m])<br />

log10(Ez [V/m])<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

(b)<br />

0 3 6 9<br />

x (km)<br />

0 3 6 9<br />

x (km)<br />

Figure 5: Electric field components (a) Ey and (c) Ez for the model shown in Figure 1 and<br />

a 1000-m long wire source at f =0.1 Hz, and (b, d) the ratios of the finite-source Ey and<br />

Ez to the respective point dipole fields.<br />

pare Figure 7b to 3), and larger than the reservoir-related anomaly for f =1Hz(compare<br />

Figure 7c to 4d). These results demonstrate that at these frequencies, the EM fields are<br />

significantly influenced not only by the grounding point locations, but also by the entire<br />

wire layout. In contrast, at f =0.001 Hz, differences between bent and straight wire fields<br />

are barely visible (Figure 7a). At this low frequency, the electric field is similar to the potential<br />

field occurring in the DC limit (i.e., for f = 0). Here, the electric field is practically<br />

determined by the grounding point locations alone.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

9<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

(d)<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

| Ey wire / Ey dipole |<br />

| Ez wire / Ez dipole |


log 10(| (E x coarse - Ex fine ) / Ex fine |)<br />

−2<br />

−4<br />

−6<br />

−8<br />

(a)<br />

Element length (m)<br />

1<br />

2<br />

5<br />

10<br />

20<br />

50<br />

0 5 10<br />

x (km)<br />

Figure 6: Relative differences between E coarse<br />

x<br />

−2<br />

−4<br />

−6<br />

−8<br />

(b)<br />

0 5 10<br />

x (km)<br />

computed for a 1000-m long wire using wire<br />

element lengths between 1 m and 50 m, and a reference field E fine<br />

x for a wire discretized<br />

using 0.1-m long elements, for frequencies of (a) 0.1 Hz and (b) 100 Hz. The resistivity<br />

model shown in Figure 1 was used, and field values were extracted along a line parallel to<br />

the source wire at a lateral offset of 50 m.<br />

Marine survey: floating wire<br />

Finite electric dipole sources used in marine CSEM surveys for commercial hydrocarbon<br />

exploration are typically ∼ 100 − 300 m long (Constable and Srnka, 2007). Accordingly,<br />

differences between the EM fields due to such finite-length sources and point dipole fields<br />

are expected to be somewhat smaller than those observed for the 1-km long wire considered<br />

in the land survey examples.<br />

We have computed marine EM responses for the model shown in Figure 8, which contains<br />

a 200-m water column and a stack of sedimentary layers, into which a 100-m thick resistive<br />

layer, representing a hydrocarbon reservoir, is embedded at a depth of 800–900 m below<br />

the seafloor. We simulated a point dipole source located at (x, y) =(0, 0) and a 300-m long<br />

wire source, also centered at (0, 0). Both sources were located 50 m above the seafloor, and<br />

receivers were placed 0.01 m above the seafloor.<br />

Electric field Ex data for this configuration due to the finite-length wire and point dipole<br />

sources are shown in Figures 9a and b, and finite-source Ex data for the reservoir model<br />

relative to Ex for a model not containing the resistive layer are displayed in Figure 9c. The<br />

finite-length wire and point dipole fields are nearly identical at radii larger than ∼ 1km<br />

from the source center. The anomaly due to the resistive layer is several times larger<br />

than the relative differences between the finite-length wire and point dipole fields, and is<br />

largest at distances well beyond the region significantly influenced by the source geometry.<br />

Therefore, the point dipole approximation may be adequate in this case. However, taking<br />

into account the exact source geometry may again become important when searching for<br />

smaller anomalies caused, e.g., by thinner reservoirs or relatively small 3D structures.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

10


y (km)<br />

−6<br />

−3<br />

3<br />

6<br />

(a)<br />

-500 0 500<br />

120° x (m)<br />

287<br />

y (m)<br />

0 3 6 9<br />

x (km)<br />

y (km)<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

(b)<br />

0 3 6 9<br />

x (km)<br />

y (km)<br />

| Ex bent / Ex straight 0.8 0.9 1 1.1 1.2<br />

|<br />

−6<br />

−3<br />

0<br />

3<br />

6<br />

(c)<br />

0 3 6 9<br />

x (km)<br />

Figure 7: Electric field Ex for a bent wire. The inset in (a) shows the geometry of a wire<br />

that consists of two segments and has the same grounding points as the straight wire used<br />

previously. Shown is the ratio of Ebent x for the bent wire to E straight<br />

x for the straight wire at<br />

depth z =0.15 m and frequencies of (a) 0.001 Hz, (b) 0.1 Hz and (c) 1 Hz. Gray shades in<br />

(b) and (c) roughly mark very low-amplitude regions in which Ex would not be measurable.<br />

σ (S/m)<br />

0<br />

3.6<br />

0.5<br />

0.01<br />

1<br />

10 km<br />

seafloor<br />

receivers<br />

15 km<br />

0<br />

200<br />

1000<br />

1100<br />

z (m)<br />

Figure 8: The 1D conductivity model and survey geometry used for simulating a marine<br />

CSEM survey. The red arrow indicates the point dipole or 300-m long wire source, centered<br />

at (x, y) =(0, 0) and located 50 m above the seafloor. Receivers are located 0.01 m above<br />

the seafloor.<br />

CONCLUSIONS<br />

We have presented a formulation for computing electromagnetic fields due to electric dipole<br />

sources of finite extent by splitting the responses into contributions from the wire body<br />

and its end points. Being derived from a quite general representation of point dipole fields,<br />

our finite-length wire formulation allows us to compute EM fields in 1D layered media for<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

11


y (km)<br />

−4<br />

−2<br />

0<br />

2<br />

(a)<br />

4<br />

0 2<br />

x (km)<br />

4<br />

log10(Ex wire −12 −11 −10 −9 −8<br />

[V/m])<br />

y (km)<br />

−4<br />

−2<br />

0<br />

2<br />

(b)<br />

4<br />

0 2<br />

x (km)<br />

4<br />

y (km)<br />

| Ex wire / Ex dipole 0.9 0.95 1 1.05 1.1<br />

|<br />

−4<br />

−2<br />

0<br />

2<br />

(c)<br />

4<br />

0 2<br />

x (km)<br />

4<br />

| Ex reservoir / Ex background 0.5 0.75 1 1.25 1.5<br />

|<br />

Figure 9: (a) Electric field Ex for a 300-m long wire source and the marine survey configurationshowninFigure8atf<br />

=0.1 Hz. (b) The ratio between the finite-length wire field<br />

shown in (a) and the field due to a point dipole source. (c) The finite-length wire field for<br />

the reservoir model of Figure 8 relative to the background field for a model not containing<br />

the 100-m thick resistive layer. Note the different color scales in (b) and (c).<br />

sources and receivers located at any depth. Compared to direct summation of dipole fields,<br />

we gain ∼ 60% efficiency. Implementation is straightforward, as we only require integral<br />

evaluations via standard fast Hankel transforms, for which precomputed sets of coefficients<br />

are available.<br />

The utility of the finite-length wire formulation has been demonstrated by presenting<br />

EM fields for several representative CSEM survey configurations. Our tests confirm that<br />

finite-length wire and point dipole fields can differ significantly over distance ranges reaching<br />

several times the wire length. For a simulated land CSEM survey targeting a thin resistive<br />

anomaly, comparison of the responses for a 1000-m long grounded wire source to point dipole<br />

responses shows that in this case, it is crucial to consider the actual source geometry. The<br />

anomalous response of the target structure is smaller than the relative differences between<br />

finite-length wire and point dipole fields, and overlaps spatially with the region in which<br />

the response is strongly influenced by the source geometry. Qualitatively similar deviations<br />

between finite-length wire and point dipole fields are observed over relatively wide frequency<br />

ranges and for different EM field components. At frequencies above the ‘effective’ DC limit,<br />

it is also important to consider the actual wire layout; knowledge of the grounding point<br />

positions is not sufficient for computing accurate responses.<br />

In contrast, for the marine survey simulated, using a shorter wire and thicker target layer,<br />

the wire geometry only had a relatively small impact on the responses, and anomalies due<br />

to the target layer were spatially well separated from the region significantly influenced by<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

12


the source geometry. This indicates that the actual source geometry is of minor importance<br />

for the survey considered. Nevertheless, this cannot be generalized to other surveys aiming<br />

at detecting smaller target structures that generate weaker EM field anomalies.<br />

Our 1D solution for finite-length wire sources can easily be incorporated into higherdimensional<br />

CSEM modeling and inversion algorithms. Most 2D and 3D modeling schemes<br />

use a secondary field approach, in which primary fields for a simple (e.g., homogeneous or<br />

1D) model are computed analytically, and secondary fields arising from deviations of the<br />

2D or 3D resistivity model from the background model are computed numerically. Here,<br />

finite-length sources can be included into the background field computations.<br />

ACKNOWLEDGMENTS<br />

This work was funded by the German Federal Ministry of Education and Research within<br />

the framework of the GeoEn project.<br />

APPENDIX A<br />

EM FIELD EXPRESSIONS FOR POINT DIPOLES<br />

We show the derivation of expressions for dipole EM fields in a form convenient for integration<br />

over the length of a source wire using the example of the Ex component. Analogous<br />

considerations apply for the other EM field components.<br />

Using the nomenclature of Løseth and Ursin (2007), the space-frequency domain electric<br />

field Ex for an isotropic layered medium, with source and receiver embedded at arbitrary<br />

depth, is given by the double Fourier integral<br />

Ex = − Idx<br />

8π2 ∞<br />

∞<br />

1− k2 x<br />

κ2 <br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A 11 +<br />

−∞ −∞<br />

<br />

pr zps z<br />

˜ε r ˜ε s<br />

k2 x<br />

κ2 RA22 <br />

exp {j (kxx+kyy)}dkxdky, (A-1)<br />

where kx and ky are horizontal wavenumbers with κ2 = k2 x + k2 y , and the other symbols are<br />

the same as those explained for Equation (1).<br />

Transformation to cylindrical coordinates using kx = κ cos α, ky = κ sin α, x = r cos β,<br />

y = r sin β, ξ = α − β + π/2, k2 x →−∂2 x , and substituting the zero-order Bessel function,<br />

results in<br />

Ex = Idx<br />

4π<br />

⎛<br />

<br />

⎝<br />

0<br />

∞<br />

−μ0<br />

√ p r z p s z<br />

J0 (κr) = 1<br />

2π<br />

2π<br />

R A 11κJ0(κr)dκ−∂ 2 ⎧<br />

⎨∞<br />

<br />

μ0<br />

x √<br />

⎩ pr zps R<br />

z<br />

A 11−<br />

0<br />

0<br />

e jκr sin ξ dξ, (A-2)<br />

p r zp s z<br />

˜ε r ˜ε s RA 22<br />

κ J0(κr)dκ<br />

⎫⎞<br />

⎬<br />

⎠.<br />

⎭<br />

(A-3)<br />

After evaluating one of the spatial derivatives using (Ward and Hohmann, 1987)<br />

∂xJ0 (κr) =− κx<br />

r J1 (κr) , (A-4)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

13<br />

<br />

1


we obtain<br />

Ex = Idx<br />

4π<br />

⎛<br />

<br />

⎝<br />

0<br />

∞<br />

−μ0<br />

√<br />

pr zps R<br />

z<br />

A 11κJ0 (κr)dκ + ∂x<br />

⎧<br />

⎨ <br />

x<br />

⎩ r<br />

0<br />

∞<br />

<br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A 11 −<br />

p r zp s z<br />

˜ε r ˜ε s RA 22<br />

⎫⎞<br />

⎬<br />

dκ ⎠ . (A-5)<br />

⎭<br />

Equation (A-5) is used as the basis for deriving the expressions for finite-length wire fields.<br />

Further evaluation of the second spatial derivative using<br />

∂xJ1 (κr) =− x<br />

r2 J1 (κr)+ κx<br />

r J0 (κr) (A-6)<br />

results in an explicit expression for the point dipole field in terms of two different Hankel<br />

integrals:<br />

Ex = Idx<br />

4π<br />

+<br />

⎛<br />

<br />

⎝<br />

0<br />

∞<br />

x 2<br />

<br />

1 2x2<br />

−<br />

r r3 ∞<br />

− 1<br />

r2 0<br />

<br />

μ0<br />

√<br />

pr zps R<br />

z<br />

A x2<br />

11 −<br />

r2 μ0<br />

√<br />

pr zps R<br />

z<br />

A 11 −<br />

p r z p s z<br />

˜ε r ˜ε r RA 22<br />

REFERENCES<br />

p r zp s z<br />

˜ε r ˜ε r RA 22<br />

<br />

κJ0 (κr)dκ<br />

⎞<br />

J1 (κr)dκ⎠<br />

. (A-7)<br />

Constable, S., and L. J. Srnka, 2007, An introduction to marine controlled-source electromagnetic<br />

methods for hydrocarbon exploration: Geophysics, 72, WA3–WA12.<br />

Edwards, N., 2005, Marine controlled source electromagnetics: Principles, methodologies,<br />

future commercial applications: Surveys in Geophysics, 26, 675–700.<br />

Edwards, R. N., 1997, On the resource evaluation of marine gas hydrate deposits using<br />

sea-floor transient electric dipole-dipole methods: Geophysics, 62, 63–74.<br />

Frischknecht, F. C., V. F. Labson, B. R. Spies, and W. L. Anderson, 1991, Profiling methods<br />

using small sources, in Electromagnetic Methods in Applied Geophysics: Society of<br />

Exploration Geophysicists, 2, 105–270.<br />

Giese, R., J. Henninges, S. Lüth, D. Morozova, C. Schmidt-Hattenberger, H. Würdemann,<br />

M. Zimmer, C. Cosma, C. Juhlin, and CO2SINK Group, 2009, Monitoring at the<br />

CO2SINK site: A concept integrating geophysics, geochemistry and microbiology: Energy<br />

Procedia, 1, 2251–2259.<br />

Johansen, S. E., H. E. F. Amundsen, T. Røsten, S. Ellingsrud, T. Eidesmo, and A. H.<br />

Bhuyian, 2005, Subsurface hydrocarbons detected by electromagnetic sounding: First<br />

Break, 23, 3136.<br />

Løseth, L. O., and B. Ursin, 2007, Electromagnetic fields in planarly layered anisotropic<br />

media: Geophysical Journal International, 170, 44–80.<br />

Soerensen, K. I., and N. B. Christensen, 1994, The fields from a finite electrical dipole - A<br />

new computational approach: Geophysics, 59, 864–880.<br />

Spies, B. R., and F. C. Frischknecht, 1991, Electromagnetic sounding, in Electromagnetic<br />

Methods in Applied Geophysics: Society of Exploration Geophysicists, 2, 285–425.<br />

Strack, K. M., 1992, Exploration with deep transient electromagnetics: Elsevier.<br />

Ward, S. H., and G. W. Hohmann, 1987, Electromagnetic theory for geophysical applications,<br />

in Electromagnetic Methods in Applied Geophysics: Society of Exploration Geophysicists,<br />

131–311.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

14


Weidelt, P., 2007, Guided waves in marine CSEM: Geophysical Journal International, 171,<br />

153–176.<br />

Wright, D., A. Ziolkowski, and B. Hobbs, 2002, Hydrocarbon detection and monitoring<br />

with a multicomponent transient electromagnetic (MTEM) survey: The Leading Edge,<br />

21, 852–864.<br />

Ziolkowski, A., B. A. Hobbs, and D. Wright, 2007, Multitransient electromagnetic demonstration<br />

survey in france: Geophysics, 72, F197–F209.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

15


2D Time Domain TEM Simulation Using Finite Elements, an<br />

Exact Boundary Condition, and Krylov Subspace Methods<br />

M. Afanasjew 1,2 ,R.-U.Börner 1 ,M.Eiermann 2 ,O.G.Ernst 2 ,S.Güttel 2,3 , and K. Spitzer 1<br />

1 Institute of Geophysics, TU Bergakademie Freiberg, Germany, 2 Institute of Numerical<br />

Analysis and Optimization, TU Bergakademie Freiberg, Germany, 3 now at: Section of<br />

Mathematics, Université deGenève, Switzerland<br />

1 Summary<br />

In this report we present a numerical method for solving Maxwell’s equations in the time domain<br />

assuming an arbitrary two-dimensional conductivity distribution including an isolating air half-space.<br />

The method allows to carry out the computations for the subsurface only, because the air-earth interface<br />

is handled by an exact boundary condition. The spatial discretization is done with the finite element<br />

method, leading to a linear system of ordinary differential equations (ODE). We use state-of-the-art<br />

Krylov subspace methods for this ODE to advance a given initial electric field to selected times of<br />

interest. The presented theory is tested with some standard models and compared to a traditional<br />

finite difference time stepping implementation with respect to accuracy and efficiency. The results<br />

clearly demonstrate the superiority of the presented method in terms of run time given a comparable<br />

accuracy.<br />

Keywords: Analytic boundary condition, finite element method, Krylov subspace, time domain, transient<br />

EM<br />

2 Introduction<br />

The transient electromagnetic (TEM) method has become a standard technique in geophysical prospecting<br />

during the past years. It is already in wide use, e. g., for the exploration of important resources like<br />

hydrocarbons, groundwater and minerals. One important aspect here is a reliable and computationally<br />

efficient simulation of the decaying electromagnetic field, which can be leveraged to get a better understanding<br />

of field behavior in complicated real-world settings as well as a building block in inversion<br />

schemes, that ultimately aim at resolving arbitrary conductivity structures from only a few well-placed<br />

measurements.<br />

The predominant forward modeling technique in the literature is the finite difference time domain<br />

(FDTD) method, that was already introduced by Yee (1966). An explicit time-stepping technique, that<br />

already dealt with an isolating air half-space, was developed by Oristaglio and Hohmann (1984) for the<br />

two-dimensional case and later refined by Wang and Hohmann (1993) for three dimensions. Like for<br />

other explicit time-stepping methods, the size of the time steps, that the described Du Fort-Frankel<br />

scheme can stably perform, depends on the grid spacing and the lowest conductivity. Although the<br />

resistive air is already eliminated, thousands of time steps have to be performed although only a few<br />

dozen solutions are necessary to describe the decaying field.<br />

The approach taken here is based on a finite element discretization, which allows for greater flexibility<br />

when modeling complicated conductivity structures. High accuracy is obtained with less effort compared<br />

to graded tensor product grids used with finite differences. It also helps in the construction of an<br />

analytic boundary condition, avoiding a few drawbacks the implementation by Oristaglio and Hohmann<br />

(1984) has. Contrary to the finite difference approach, the matrices resulting from the discretization are<br />

symmetric and, thus, allow for a wide range of efficient and state-of-the-art time integration techniques,<br />

which can exploit this property. One such family of time integrators is based on building a Krylov<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

16


subspace and extracting approximations to the matrix exponential function from said space, thus<br />

evaluating the sought electric field directly at a given time.<br />

3 Theory<br />

3.1 Governing Equations<br />

Our governing equation derives from Maxwell’s equations in the diffusive limit, with the constitutive<br />

relations used and the magnetic field already eliminated. Thus, we write<br />

with<br />

∇× μ −1 ∇× e + ∂tσe = −∂tj e<br />

e = e(x,t) the electric field,<br />

μ = μ(x) the magnetic permeability,<br />

σ = σ(x) the electric conductivity, and<br />

j e = j e (x,t) the impressed source current density.<br />

We now restrict ourselves to the two-dimensional case (xz-plane) with infinitely long line sources<br />

perpendicular to this plane. Given these assumptions, we can express the electric field as<br />

e(x, y, z, t) =e(x, z, t) y (2)<br />

where y is the unit vector along the y-axis and e a scalar function. Equation (1) then reduces to<br />

−∇ 2 e + σμ∂te = −μ∂tj e . (3)<br />

Γ0<br />

Ω<br />

z =0<br />

Figure 1: Computational domain with an arbitrary conductivity structure. On top is the air-earth interface<br />

(bold), the remaining boundaries are subsurface boundaries.<br />

Our computational domain Ω (cf. Figure 1) is a rectangle and its top edge is aligned with the airearth<br />

interface Γ0 = {(x, z) :z =0} on which we impose an explicit boundary condition and perfect<br />

conductor boundary conditions on all other domain boundaries. It is important for these boundaries to<br />

be sufficiently far away from the sources, so that they don’t distort the propagating field.<br />

Our objective will be to compute the configuration of the electric field ei at times ti for i ∈{1, 2,...,n}<br />

givenaninitialfielde0 at t0. We use sources that are switched off at t = 0 and therefore the right hand<br />

side of (3) vanishes for t>0.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

17<br />

(1)


3.2 Air-Earth Interface<br />

We assume that e satisfies Laplace’s equation in the isolating air half-space (z 0} denotes the lower half-space.<br />

By using the exact boundary condition (6) we can write<br />

∂ta(e, ψ) σ + c(e, ψ)+<br />

+∞<br />

−∞<br />

Te(x, t) ψ(x, 0) dx = −μ<br />

+∞<br />

−∞<br />

∂tj e (x, t) ψ(x, 0) dx. (10)<br />

To be able to discretize this equation we need a computable expression for the following bilinear form<br />

b(φ, ψ) =<br />

+∞<br />

−∞<br />

Tφ(x) ψ(x) dx (11)<br />

and indeed, after some rearrangement in the Fourier domain, one obtains (Goldman et al. 1986)<br />

+∞ +∞<br />

b(φ, ψ) =− 1<br />

log(|x − y|) φ<br />

π −∞ −∞<br />

′ (x) ψ ′ (y) dxdy. (12)<br />

The discretization of this integral on the boundary of a triangulation can now be easily computed.<br />

The continuity of e allows us to use standard linear Lagrange elements on a triangulation of the<br />

computational domain. After discretizing the variational formulation we arrive at<br />

M∂te = Ke. (13)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

18


Finite Differences For comparison, we also perform a finite difference discretization and get the ODE<br />

∂te = Ae where A is a large sparse matrix representing the discrete curl-curl operator.<br />

The following procedure, that has to be repeated in every time step, outlines the implementation of an<br />

exact boundary condition in this setting: Take the field data at the air-earth interface, interpolate it to<br />

an equidistant grid, perform an FFT to transform the data into the wave number domain, multiply<br />

every wave number with a constant (this would be a convolution in the space domain), revert the<br />

FFT and finally interpolate back to the graded grid to get the field values at a negative z, thus, inside<br />

the air half-space. Use the data resulting from this classical upward continuation in the usual finite<br />

difference stencil to compute the top layer in the next step.<br />

A variation of this approach was also implemented for the numerical examples. It combines all of the<br />

above steps into a single (dense) matrix that has the same effect but can be evaluated more efficiently.<br />

This is called matrix upward continuation later on.<br />

Details about the finite difference discretization can be found in Oristaglio and Hohmann (1984).<br />

3.4 Time Integration Techniques<br />

Krylov Subspace Methods Krylov subspace methods cover a big range of applications and are in<br />

use for several decades. What we want to focus on are Krylov subspace methods for the evaluation<br />

of matrix functions, like we can see in (13), where the function is the matrix exponential and the<br />

matrices come from the spatial discretization using finite elements. A nice side effect of the origin of<br />

those matrices is, that they are usually symmetric which many algorithms can benefit from.<br />

Given a square matrix A of size n × n, a vector b of length n and a suitable scalar function f, wecan<br />

write<br />

f(A) b = p(A) b. (14)<br />

with p a polynomial that Hermite-interpolates in the eigenvalues of A. We will be focusing on the<br />

exponential function and on rational Krylov subspaces, that are defined as follows<br />

with Km(A, b) = b,Ab,A 2 b,...,A m−1 b and<br />

Q(A, b) := qm−1(A) −1 Km(A, b) (15)<br />

qm−1(z) =<br />

m−1 <br />

j=1<br />

ξj=∞<br />

(z − ξj) . (16)<br />

Such subspaces are constructed, e. g., with the rational Arnoldi method to obtain an orthonormal<br />

basis of Qm from which approximations to f(A) b can be computed with several procedures. We have<br />

to choose the poles ξj of the rational Krylov subspace. Their number and choice highly impact the<br />

quality of the approximations that can be extracted from the subspace. Luckily, there is a heuristic to<br />

determine good poles for the approximation of the exponential function, given a certain parameter<br />

interval (in our case the times we want to advance to) and accuracy requirements. These, and many<br />

more things concerning rational Krylov subspaces, are tackled in Güttel (2010), which is also a good<br />

starting place to dig deeper into the literature.<br />

Looking more closely at (13) we see, that in order to apply a Krylov subspace method we need to<br />

remove the matrix M in front of the time derivative, e. g. by moving it to the right<br />

∂te = M −1 Ke. (17)<br />

The inverse of the matrix doesn’t have to be explicitly computed, but we have to solve a system of<br />

linear equations in every iteration step. This might be considered too expensive, so there is a technique<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

19


termed mass lumping which converts M into a diagonal matrix that can be easily inverted without<br />

distorting the solution too much.<br />

However, even if we don’t have to solve linear systems with the matrix M, wedohavetosolveshifted<br />

systems with the matrix K during the construction of the rational Krylov subspace. Therefore, an<br />

efficient solver is also an important ingredient. We will see in the numerical examples what options we<br />

have.<br />

Time Stepping On the other hand there are simple, yet relatively efficient explicit time stepping<br />

techniques. One of them is the well-known Du Fort-Frankel method which is implemented here to<br />

perform the intergration in time for the finite difference discretization.<br />

4 Numerical Examples<br />

All computations were carried out on a machine with four AMD Opteron 8380 quad cores, running<br />

at 2.5 GHz with a total of 128 GiB of RAM. The algorithms were implemented in pure MATLAB<br />

code running under MATLAB 2008b, with some finite element related tasks (nothing time-critical)<br />

performed by COMSOL 3.5a, both running in a 64-bit Linux environment. To generate reproducible<br />

timing we restricted ourselves to one computational thread that was explicitly pinned to one of the<br />

cores.<br />

4.1 Model<br />

The considered models are basically those from Oristaglio and Hohmann (1984), all based on graded<br />

tensor product grids. We extended the mesh slightly so as not to get a distortion from the boundaries<br />

for late times. The resulting grid has 236 × 88 cells, is 12800 m wide and 5255 m deep, with grid spacings<br />

ranging between 10 m and 240 m. To make the computations comparable with the finite difference code<br />

we decided to use exactly the same grid for the finite element code. We simply subdivided each cell<br />

into two triangles, however, the number of unknowns or degrees of freedom and their locations are<br />

identical. Technically, this is of course not necessary and a properly adapted unstructured grid would<br />

certainly yield even better results at lower cost.<br />

The initial field e0 at t0 =10 −6 s is due to two line sources at positions (x, z) = (−500 m, 0m) and<br />

(x, z) = (0 m, 0m), the former being negative while the latter is positive with a unit source current<br />

strength. It is computed analytically for a homogeneous half-space that corresponds to the background<br />

conductivity of the model.<br />

We have looked at two models with different conductivity structures. See Figure 2 to get an approximate<br />

idea of their location inside the grid. The background has a resistivity of 300 Ωm and the conductor<br />

(red) 0.3 Ωm.<br />

Γ0<br />

Ω<br />

z =0<br />

Figure 2: Schematic view of the computed models. The models are the homogeneous half-space and a conductor<br />

(red) inside a homogeneous half-space. The conductor is located at the position 290 m ≤ x ≤ 310 m,<br />

100 m ≤ z ≤ 400 m in the grid.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

20


4.2 Methods and Setup<br />

Independently of the method of choice we start at the time t0 =10 −6 s and integrate from there till<br />

we reach tn =10 −2 s. We seek to compute the configuration of the electric field at 10 logarithmically<br />

equidistant times per decade.<br />

For the finite difference code we mostly stick to the implementation by Oristaglio and Hohmann<br />

(1984), with the only option to perform the upward continuation in the traditional way, which is a<br />

significant expense per time step, or we can decide to precompute the matrix formulation of this<br />

upward continuation and to apply it in every time step.<br />

For the finite element method we have no choice regarding the exact boundary condition. However, we<br />

can choose between different variants of the rational Krylov method. For all of them we have a total<br />

of 25 poles, which should be close to optimal for this application. For the first 24 poles we alternate<br />

between 2.7826 · 10 5 and 0.0242 · 10 5 . The last pole is set to infinity.<br />

We have four implementations, that we included in this test. Two of them use mass lumping to reduce<br />

the mass matrix to a diagonal matrix. Furthermore, for each group we can either call the standard<br />

MATLAB solver or we can exploit the fact that we have multiple identical poles for which we have to<br />

solve with different right hand sides, but identical matrices. We do this by computing a sparse LU<br />

factorization (we use UMFPACK for this) for every unique pole and then leverage that factorization to<br />

speed up the solves during the construction of the rational Krylov subspace.<br />

4.3 Results<br />

We first look at the computation times that are pictured in Figure 3 and listed in Table 1. The times<br />

shown here are for the homogeneous half-space model, but aren’t significantly different for the other<br />

models since the grid and the minimal conductivity are identical. Thus, we can use these numbers to<br />

judge the acceleration we can get by using our method.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Figure 3: Computation times split into solve and setup/post-processing step, cf. also Table 1. On the left<br />

are the computation times for the default grid and on the right for the regularly refined version<br />

of this grid with approximately four times as many nodes and degrees of freedom. The following<br />

abbreviations were used: classical upward continuation (cUC), matrix upward continuation (mUC),<br />

rational Krylov (rK), mass lumping (ML), UMFPACK solver (UMF).<br />

As is clearly visible, all FE-based methods are significantly faster than their finite difference counterparts,<br />

sometimes even 22 times faster.<br />

Looking at the transients (cf. Figure 4) at some select points inside the mesh, we see that although<br />

these newer methods are so much faster, we don’t really sacrifice accuracy. In fact, in many cases the<br />

finite element solution is more accurate than the one obtained by finite differences.<br />

We conclude our numerical examples with showing some cross-sections (cf. Figure 5), that are nothing<br />

but a few snapshots of an animation that is just not suitable for this medium, but nevertheless gives a<br />

coarse idea of how the electric field propagates with time.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

21


Figure 4: Transients of the homogeneous half-space model with relative errors (left) and the model with a<br />

conductor (right). Negative field values are denoted by dashed lines while positive values are drawn<br />

with a solid line. We can see that we have a very good agreement between the FD and FE solution.<br />

When comparing against an analytical solution in the homogeneous half-space, we see that the relative<br />

errors are often lower than those of the FD solution, despite the lower numerical effort.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

22


Figure 5: Cross-sections of a homogeneous half-space model (left) and the model with a conductor (right).<br />

These cross-sections were plotted to give a visual idea of how the fields advance in time. The common<br />

initial configuration was omitted. We decided to plot the results of the FE simulation, but there was<br />

hardly any visible difference compared to the FD solution. On the right we can nicely see how the<br />

field gets captured inside the high conductivity structure and creates an anomaly compared to the<br />

homogeneous case.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

23


Default grid Refined grid<br />

Method Solve Other Solve Other<br />

FD, time stepping, classical upward continuation 31.7s < 0.1 s 171.8s < 0.1s<br />

FD, time stepping, matrix upward continuation 8.6s 0.9s 98.4s 2.7s<br />

FE, rational Krylov 3.6s 0.6s 19.4s 1.9s<br />

FE, rational Krylov, mass lumping 3.1s 0.6s 17.7s 1.9s<br />

FE, rational Krylov, UMFPACK 1.9s 0.6s 10.4s 1.9s<br />

FE, rational Krylov, mass lumping, UMFPACK 1.4s 0.6s 7.7s 1.9s<br />

Table 1: Computation times for the default grid and one, that was obtained from this with regular refinement.<br />

The column Solve denotes the time spent for performing the integration in time, while the column<br />

Other lists the time spent in setting up the boundary condition and the computation of field values<br />

from the solution vectors in case of the FE method. Other setup times are not shown. The advantage<br />

of FE-based methods can be clearly seen.<br />

5 Conclusions<br />

We were able to leverage several state-of-the-art techniques to create a combined efficient forward<br />

modeling code that is significantly faster than traditional codes. We have also seen that we don’t have<br />

to make sacrifices regarding the accuracy of those computations. This was already achieved by using<br />

the mesh from the finite difference discretization, which was helpful for comparison, but which also<br />

has many limitations due to its regular structure. An unstructured grid—properly adapted to the<br />

problem—could be used to further increase accuracy or speed.<br />

We are working on creating a similar framework for the three-dimensional time domain TEM problem.<br />

All of the methods and tools are already existing or have a straightforward extension to 3D. Given the<br />

results from 2D we expect an even greater speed-up when applying this to three dimensions.<br />

6 Acknowledgments<br />

This work was supported by the German Research Foundation (DFG) under signature Spi 356/9.<br />

References<br />

Goldman, Y., C. Hubans, S. Nicoletis, and S. Spitz (1986). A finite-element solution for the transient electromagnetic<br />

response of an arbitrary two-dimensional resistivity distribution. Geophysics 51(7): 1450–1461.<br />

Goldman, Y., P. Joly, and M. Kern (1989). The Electric Field in the Conductive Half Space as a Model in Mining and<br />

Petroleum Prospecting. Mathematical Methods in the Applied Sciences 11: 373–401.<br />

Güttel, S. (2010). Rational Krylov Methods for Operator Functions. PhD thesis. TU Bergakademie Freiberg.<br />

Oristaglio, M. L. and G. W. Hohmann (1984). Diffusion of Electromagnetic Fields into a Two-Dimensional Earth: A<br />

Finite-Difference Approach. Geophysics 49(7): 870–894.<br />

Wang, T. and G. W. Hohmann (1993). A Finite-Difference, Time-Domain Solution for Three-Dimensional Electromagnetic<br />

Modeling. Geophysics 58(6): 797–809.<br />

Yee, K. S. (1966). Numerical Solution of Initial Boundary Problems Involving Maxwell’s Equations in Isotropic Media.<br />

IEEE Trans. Ant. Propag. 14: 302–309.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

24


Abstract<br />

TEM with anomalous diffusion in fractal conductive media<br />

Tilman Hanstein<br />

KMS Technologies GmbH, Köln<br />

The transient electromagnetic response of an inductive loop source over a half-space with<br />

fractal characteristics is simulated. The conductivity of the ground has a spatial distribution,<br />

which is described by a roughness parameter. The roughness can be related to the fractal<br />

dimension and controls the transient decay of the signal. Asymptotic limits are observed for<br />

the early and late time behaviour.<br />

The work is based on a paper by Everett (2009), which has been reviewed and the aspect of<br />

asymptotic limits in time has been further investigated. Contrary to that paper, here a<br />

relationship between roughness and the electromagnetic decay at early and late time has been<br />

deduced. The transient decay for normal diffusion is t -5/2 and for anomalous diffusion the<br />

decay is slower. The new power law for anomalous diffusion has been proven by theoretical<br />

analysis and has been verified by numerical experiments.<br />

The numerical evaluation of the inverse Laplace transform with the method of the fast Hankel<br />

transform are excellent in numerical accuracy and the method with the Gaver-Stehfest<br />

algorithm is not sufficient to estimate the power law decay at late times.<br />

Anomalous Diffusion<br />

The concept of anomalous diffusion is a useful approach for the description of diffusion<br />

process and transport dynamics in complex systems. The fractional equations are derived<br />

asymptotically from basic random walk models and become a complementary tool for<br />

handling non-exponential relaxation patterns.<br />

For transient electromagnetic diffusion Everett starts this concept with a generalized Ohm’s<br />

law (Everett 2009, Weiss and Everett 2007)<br />

j <br />

E<br />

1 t t<br />

E j E <br />

0<br />

the parameter ~t - describes the generalized electrical conductivity and is appropriate for<br />

the anomalous diffusion coefficient. The Ohm’s law becomes a convolution between the<br />

generalized conductivity and the electric field E. The roughness parameter can vary 0 1.<br />

d<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

25


After applying Ampere‘s and Faraday‘s law to the generalized current density, we get the<br />

fractional diffusion equation<br />

<br />

E 0<br />

* E 0<br />

0Dt<br />

t<br />

2 1<br />

where now the fractional derivative or Riemann-Louiville operator is introduced (Metzler and<br />

Klafter 2000)<br />

t<br />

1<br />

1 E<br />

<br />

0 Dt<br />

E t <br />

d<br />

1<br />

t<br />

t <br />

0<br />

<br />

and is the Gamma function which serves as normalization constant.<br />

The fractional diffusion equation will be solved in the Laplace domain and the transformation<br />

of the operator yield to<br />

1<br />

1<br />

~<br />

D E<br />

t s E<br />

s<br />

L 0 t<br />

with the complex Laplace variable s=i (Abramowitz & Stegun 1964).<br />

The fractional diffusion equation becomes now a simple expression<br />

2 ~<br />

E s<br />

1<br />

0<br />

<br />

This differential equation can be solved straightforward with the standard methods as used for<br />

the electromagnetic 1-D formulation in a layered medium.<br />

~<br />

E<br />

s<br />

Electromagnetic responese of a loop over rough half-space<br />

Everett (2009) has used a separate horizontal loop configuration for transmitter and receiver<br />

as it is typically used for TEM. The transmitter loops usually a square loop can be represented<br />

by a circular loop whith equivalent area. For the separate loop configuration the time<br />

derivative of the vertical magnetic field is measured outside the transmitter loop.<br />

In frequency domain the vertical magnetic field is presentated as Hankel Integrals<br />

~<br />

h z<br />

s Ia<br />

J aJr <br />

0<br />

<br />

2<br />

<br />

2<br />

k <br />

2<br />

<br />

1<br />

0<br />

E<br />

t<br />

d<br />

.<br />

The integration is over the spatial wavenumber and I is the current in the transmiter, a the<br />

transmitter loop radius, r the distance to the receiver point, J0 and J1 are Bessel functions of<br />

order 0 and 1 and k the fractional wavenumber. The integral can be solved analytically for an<br />

infinitessimally magnetic dipole, therefore the limit of the first order Bessel function for small<br />

radius is taken into account (Abramowitz & Stegun 9.1.10)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

26


a<br />

J1 a .<br />

a<br />

0<br />

2<br />

The response of a vertical magnetic dipole in frequency domain over a rough conductive<br />

media is<br />

m 1<br />

2<br />

u<br />

s 999u4uue ~ 3<br />

z<br />

3 2<br />

H<br />

2<br />

r<br />

with the fractional induction number u = k r = (s 1- 0 ) 1/2 r.<br />

Asymptotic limits in frequency and time domain<br />

u<br />

For normal diffusion with roughness equal zero the transient response can be also given in<br />

analytical expression but for a general fractional induction number with the roughness<br />

parameter the transformation to time domain has to be done by numerical techniques. Only<br />

the asymptotic limits for the high and low frequency limit can be calculated in closed form.<br />

For high frequency and early time we get<br />

~<br />

H<br />

h.<br />

f .<br />

z<br />

m 9<br />

m 9 1<br />

<br />

t<br />

2 2<br />

z<br />

3 2<br />

2<br />

r u<br />

2<br />

r r <br />

e.<br />

t.<br />

s <br />

H t<br />

The low frequency can be developed in a series<br />

n1n3 <br />

2<br />

~ m 1<br />

n<br />

H z s 1<br />

2<br />

3 <br />

2<br />

r 2 n4<br />

n!<br />

The first 3 terms are important for the late time behavior<br />

~<br />

H z<br />

m<br />

2<br />

r<br />

1 <br />

2 <br />

<br />

0<br />

2<br />

u <br />

<br />

2<br />

3<br />

s 1<br />

u u <br />

3<br />

1<br />

2<br />

4<br />

15<br />

<br />

<br />

<br />

.<br />

1 In time domain the first term is a -function without influence on late time decay. The second<br />

term is responsible for the late time behavior in a rough medium<br />

L<br />

1<br />

1<br />

s <br />

<br />

<br />

t<br />

<br />

<br />

<br />

<br />

t<br />

1<br />

2<br />

for 0<br />

for 0<br />

The third term describes the classical -5/2 decay response in a non-fractional medium and for<br />

a fractional medium this term decays faster than the previous second term.<br />

<br />

.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

27<br />

<br />

.


The late time response for the magnetic field can be summarized<br />

<br />

.<br />

H t l<br />

z<br />

t<br />

m<br />

<br />

2<br />

r<br />

3<br />

3 2<br />

5<br />

0<br />

2<br />

t<br />

10 <br />

<br />

<br />

0<br />

<br />

<br />

t<br />

<br />

2 1<br />

2<br />

for 0<br />

for 0<br />

The cause for the heavy tailed decay response in rough geological media can be explained by<br />

the additive second term which yield to a continuous transition from non-fractional to<br />

fractional diffusion as is shown in figure 1 and 2.<br />

Numerical inverse Laplace transform<br />

Numerical methods are applied for the transformation to time domain. Since the transient is a<br />

real and causal function, the inverse Fourier or Laplace transform can be calculated by a sinetransform.<br />

<br />

2<br />

h z<br />

z<br />

0<br />

t Im h ~ sint In this expression I have already considered that the current step in the transmitter decribed by<br />

1/i and the time derivative of the receiver coil by a multiplication with i cancel each other.<br />

Since for diffusion processes the kernel function is a smooth function, the technique of the<br />

Fast Hankel Transform can be applied. The sine function is experessed by Bessel function<br />

with fractional order ½<br />

2<br />

J 1 x sinx.<br />

2 x<br />

The fast Hankel transform is a well know technique for calculating the transient response, I<br />

used 250 filter coefficients, 15 /decade, calculated with the program by Christensen (1990).<br />

Everett (2009) has chosen the Gaver-Stehfest algorthim for his investigation. This method is<br />

also good for diffusion processes and has been succesfully applied. It is a favourized method<br />

in hydrology, because it needs only real arithmetic and the Laplace variable s is considered as<br />

a real variable. Knight and Raiche (1982) introduced this technique to the electromagnetic<br />

community. Stehfest (1970) published an Algol routine, which can be straight forward<br />

translated to other computer languages and Everett (2009) has tabled the coefficients for<br />

several total number of coefficients.<br />

d<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

28


Figure 1: Transient loop-loop response with early and late time approximation over halfspace<br />

with different roughness in conductivity, the offset is 100 m, = 0.1 S/m.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

29


The disadvantage of the Gaver-Stehfest algorithm is that the numerical accurracy cannot<br />

generally be increased by increasing the number of coeffcients. The accuracy is depending on<br />

the accurracy in number of digits of the kernel function and the number of digits of the<br />

maschine precision. Stehfest recommended 8 coefficients for single precision and 18<br />

coeffcients for double precision. Everett used 18 coefficients. In my experiments I have found<br />

out that 12 is an optimal number for electromagnetic application.<br />

Figure 2: Numerical evaluation of the exponent in the transient decay and theoretical<br />

asymptotic limit for different roughness<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

30


Results<br />

For the numerical experiment I used the same model as Everett, beside that here a vertical<br />

magnetic dipole instead of a horizontal circular loop is used. The base conductivity of the<br />

half-space is 0.1 S/m with combination of various roughness parameter has been applied.<br />

Figure 1 shows the transient as induced voltage due to an step response of the transmitter<br />

current for different roughness paramters. The large time range has been chosen to<br />

demonstrate the early and late time behavior in comparison with the asymptotic limits. For<br />

real measurements the time range will be much smaller. All transients - shown here - are<br />

calculated with the fast Hankel transform. As reference the transient of homogenous halfspace<br />

with normal diffusion is also shown with grey lines. For separate loop configuration the<br />

transient shows a sign change. The negative values are shown with dashed lines and positive<br />

with solid lines. The asymptotic limits agree very well with theoretical prediced limits at eary<br />

and late times shown here as straight lines. The early time behaviour is for practical pupose of<br />

minor relevance, because in real meausrements the early time is influenced by system<br />

response of the system as the ramp for the current step off. So that in the data the straight line<br />

will not be visible. But the late time behaviour will be visible if the late time data can be<br />

measured and the data quatlity is sufficient.<br />

To analyze how accurate the power law decay can be estimated from the calculated transients<br />

especialle for low roughness numbers, the transients are calculated up to extrem late times and<br />

the power law is determined by taking the numerical derivative d ln H(t) / d ln t. The result is<br />

shown in figure 2.<br />

Figure 3: Numerical evaluation of the exponent in the transient decay and theoretical<br />

asymptotic limit for different roughness calculated with Gaver-Stehfest alogorithm<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

31


The graphs show the exellent accuracy of the fast Hankel transform, even for small roughness<br />

numbers, and indicates the smooth transition, when the roughness approaches zero. The<br />

influence of the anomalous diffusion moves to later times for decreasing roughness.<br />

For very low roughness there will be a time range showing almost normal decay t -5/2 and then<br />

goint to the predicted power law decay at extrem late time – approaching infinity. Responsible<br />

for this behavior is the second term of the low frequency approximation which is added.<br />

The Gaver-Stehfest algorithm is not sufficient to estimate the power law decay. I tried<br />

different ways of programming, e.g to consider numerical accuracy the Laplace transform is<br />

done before the Hankel transform over the spatial wave number as recommended by Knight<br />

and Raiche (1982). The best results are shown in figure 3, achieved with 12 coefficients and<br />

using the analytical response for a vertical magnetic dipole in Lapace domain. Notice that the<br />

time range is shorter but still 3 decades more than in Everett’s paper.<br />

Reference:<br />

Abramowitz, M. & Stegun, I. A., 1966. Handbook of Mathematical Functions with Formulas,<br />

Graphs and Mathematical Tables. National Bureau of Standards, Applied Mathematics Series<br />

55.<br />

Christensen, N. B., 1990. Optimized Fast Hankel Transform Filters, Geophys. Prosp., 38,<br />

545-568.<br />

Everett, M., 2009. Transient electromagnetic response of a loop source over a rough<br />

geological medium, Geophys. J. Int., 177, 421-429.<br />

Knight, J. H. & Raiche, A. P., 1982. Transient electromagnetic calculation using Gaver-<br />

Stehfest inverse Laplace transform method, Geophysics, 47, 47-50.<br />

Metzler, R. & Klafter, J., 2000. The random walk’s guide to anomalous diffusion: a fractional<br />

dynamics approach, Phys. Rep.,339, 1-77.<br />

Stehfest, H., 1970. Numerical inversion of Laplace transforms, Comm. A.C.M.,13,47-49 (see<br />

also remark p.624).<br />

Weiss, C. J. & Everett, M. E., 2007. Anomalous diffusion of electromagnetic eddy currents in<br />

geological formations, J. geophys. Res. 112, B08102, doi:10.1029/2006JB004475.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

32


Analysis of seafloor marine EM data with respect to motion-induced noise<br />

K. M. Bhatt 1 , A. Hördt 1 and T. Hanstein 2<br />

1<br />

Inst. f. Geophysik u. Extraterrestrische Physik, TU Braunschweig<br />

2 KMS Technologies - KJT Enterprises Inc.<br />

Abstract<br />

Any appreciable movement of sea water induces an electromagnetic field, which acts as<br />

noise for marine controlled source electromagnetic (mCSEM) data. Noise in seafloor<br />

mCSEM data is considered small, but since the characteristic reservoir signal is also small,<br />

the understanding and possible removal of noise may be essential to increase the number<br />

of possible target reservoirs.<br />

A power spectral density plot of the time-series extracts the power of each frequency<br />

contribution, but the shortcoming is that the time information is lost. To get time<br />

information corresponding to each frequency, we plotted a spectrogram which shows how<br />

the spectral density of a signal varies with time. The collective use of these three plots, i.e.<br />

time series, power spectral density and spectrogram, helps to analyse the information<br />

concealed in the time-series. In a measured signal-free mCSEM data, we are able to<br />

identify various oceanic features like microseisms, swell etc, which play a significant role<br />

in inducing an electric field.<br />

Key words: mCSEM, Swell, Microseisms<br />

____________________________________________________________________________<br />

1. Introduction<br />

Marine controlled source electromagnetic (mCSEM) data is generally contaminated by<br />

some unwanted electromagnetic (EM) signals ambient at the seafloor. Understanding and<br />

possible removal of the noise is essential as technology is advancing from deep ocean to<br />

shallow ocean, where the noise is much more effective.<br />

In general, there could be two possible sources of oceanic noise production.<br />

Internal oceanic processes or any other external influences like ionospheric and<br />

magnetospheric current systems. Faraday (1832) reported that any appreciable<br />

movement of seafloor and sea water by induction generates electromagnetic signals with<br />

in the ocean. Motional induction got more attention after world war II, with the detailed<br />

oceanic induction study by many authors like Sanford (1971), Podney (1975), Chave and<br />

Cox (1982), Chave and Luther (1990). Similarity in all these studies is that they all<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

33


studied the theoretical aspect of motional induction occurring due to various movements<br />

in the ocean. Experimental evidence was provided by Webb and Cox (1986), who made a<br />

nice comprehensive study about the seafloor microseisms by observing pressure and<br />

electric field spectra at the seafloor.<br />

Wave progressions and oceanic processes are, usually, a function of the regional<br />

and local conditions. A variety of motional processes like swell, ripples, internal waves,<br />

microseisms, hum etc generate a range of EM noise at the seafloor, which varies from<br />

place to place. External, ionospheric and magnetospheric, current systems also<br />

contribute EM noise, but they are limited to the lower frequency range due to the<br />

conductive filtering of the ocean. Characteristics of the noise are poorly known till now,<br />

as only few studies are made. The present study is dedicated to the understanding of the<br />

noise characteristics at the sea floor.<br />

For the study, a marine EM data set recorded at the seafloor in the absence of a<br />

transmitter, is used. Recording is made for two horizontal components of electric fields<br />

i.e. Ex and Ey, which are perpendicular to each other. Visibly, Ex and Ey time series data<br />

appears different in their pattern and amplitudes despite the same oceanic environment.<br />

This suggests that a comprehensive study on the directional characteristics of the<br />

events/sources of noise could play a decisive role in identifying the noise sources.<br />

The oscillatory occurrence, due to a disturbance, can be comfortably visualised in<br />

frequency domain. The power spectral density (PSD) of each frequency contribution<br />

helps in marking the power of the individual sources. But the shortcoming of the PSD<br />

calculation is that if the source of the characteristic frequency is not known a priori, it is<br />

difficult to conclude about the source environment with the frequency information only.<br />

In this case, time information together with the PSD may help at least to distinguish<br />

between localised and ambient sources, which itself is of significant use. To overcome the<br />

PSD shortcoming to retain time information corresponding to each frequency,<br />

spectrograms are plotted, which show how the spectral density of a signal varies with<br />

time.<br />

In the signal-free mCSEM data, we have observed many features which are new<br />

for mCSEM studies. The spectrogram presents clear prints of oceanic features like<br />

microseisms and swell, which play a significant role in inducing an electric field at the<br />

seafloor.<br />

2. Marine Controlled Source Electromagnetism (mCSEM)<br />

A mCSEM data acquisition methodology is shown in Figure 1. In practice, an EM<br />

transmitter is towed close to the seafloor to maximize the coupling of electric and<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

34


magnetic fields with seafloor rocks. These fields are recorded by instruments deployed on<br />

the seafloor at some distance from the transmitter. Details about the data acquisition<br />

practice in both time and frequency domain of mCSEM is well documented in the article<br />

by Constable and Srnka, (2007).<br />

Figure 1: Schematic diagram showing field layout for mCSEM sounding.<br />

3. Sources of mCSEM noise<br />

The noise convolved in mCSEM data is capable of masking significant marine<br />

information. Broadly, two origins are expected as noise sources in mCSEM data:-<br />

a) External origin: The magnetospheric and ionospheric current system induces<br />

natural electric and magnetic (EM) fields in the conductive formations. The<br />

induced fields are signal for the magnetotelluric (MT) technique but noise for<br />

mCSEM data. The conductive ocean filters the higher frequencies. The skin depth<br />

() is given by<br />

This implies, f 76,000 / 2<br />

503<br />

1<br />

f<br />

for =3.3 S/m<br />

Therefore, at the ocean floor the external origin fields effectively contribute<br />

frequencies less than or equal to f = (76,000/ D 2 ), where D is the depth of ocean<br />

floor in meters and f is given in (1/s). As example for an ocean of depth 500 m, the<br />

external field will contribute frequencies less than 0.3 Hz at the sea floor. This<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

35


suggests that the frequencies higher than 0.3 Hz could be by internal oceanic<br />

origin.<br />

b) Internal origin: The dynamics of oceanic water with in the ambient geomagnetic<br />

field induces electric and magnetic field in the ocean. A frequency range of the<br />

fields is generally 0.3 < f (Hz) < 16 (Dahm et. al., 2006) excluding long period<br />

waves like tidal waves, tsunami, seiches and storm surge etc.<br />

Other than these noise sources, there are processes like earthquakes, volcanoes,<br />

manmade explosions etc occurring at the vibrant ocean floor, which add noise in<br />

mCSEM data.<br />

4. Motional induction<br />

The ocean is an electrically conducting fluid containing the ionic charge particles.<br />

Moving charge particles in the ambient geomagnetic field experience a deflective Lorentz<br />

force. If v is the velocity of charge particle q moving in the geomagnetic field B , then the<br />

Lorentz force is-<br />

<br />

q(<br />

v B)<br />

(1)<br />

F L<br />

The charge q experiences the deflecting force FL because of the action of an electric field<br />

which we call Lorentz electric field ( EL ),<br />

<br />

FL<br />

<br />

qEL<br />

(1-a)<br />

<br />

v B<br />

(2)<br />

E L<br />

The field EL generates a secondary electric field E , mainly by two processes: i) Galvanic<br />

process, ii) Inductive process (Bhatt et al, 2010). For a stationary frame of reference, the<br />

current density in the Ohms law is given by<br />

<br />

(<br />

E E ) (<br />

E v B)<br />

(3)<br />

J L<br />

Here, is the conductivity of the ocean, E is the current term generated by both a<br />

<br />

galvanic and an inductive process and (<br />

v B)<br />

is the source current term for the motional<br />

induction case. Under the quasi-static approximation, the set of Maxwell’s equation to be<br />

solved for the motional induction case is<br />

<br />

H (<br />

E v B0)<br />

<br />

(4-a,b)<br />

E <br />

H<br />

0<br />

t<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

36


Simplification of the source (<br />

v B)<br />

in terms of component Jx, Jy and Jz can be written as,<br />

J<br />

J<br />

J<br />

x<br />

y<br />

z<br />

(<br />

E<br />

<br />

(<br />

E<br />

(<br />

E<br />

x<br />

y<br />

z<br />

v<br />

v<br />

v<br />

y<br />

z<br />

x<br />

B<br />

B<br />

B<br />

z<br />

x<br />

y<br />

The non-divergent requirement condition of the current (i.e. . J 0<br />

<br />

) suggests that the<br />

current will close its loop with in the earth rather than closing in the ocean. Therefore,<br />

the horizontal electric currents (i.e. Jx and Jy) in the ocean are more effective than the<br />

vertical currents (i.e. Jz) as the horizontal extent of ocean is much larger than the vertical<br />

depth. In mCSEM, the data is recorded at the seafloor, where normally the vertical<br />

velocity component (i.e. vz) is very small. Therefore, terms vzBBy and vzBx B is negligible.<br />

Above argument simplifies (4),<br />

Jx<br />

(<br />

Ex<br />

v yBz<br />

)<br />

(6)<br />

J (<br />

E v B )<br />

Using (6), simplification of (4-a) gives<br />

E<br />

E<br />

y<br />

x<br />

y<br />

x<br />

x<br />

1<br />

<br />

( zH<br />

<br />

1<br />

<br />

( zH<br />

<br />

y<br />

x<br />

z<br />

v<br />

v<br />

v<br />

<br />

<br />

y<br />

x<br />

z<br />

x<br />

y<br />

B<br />

B<br />

B<br />

z<br />

z<br />

y<br />

z<br />

x<br />

)<br />

)<br />

)<br />

H ) v B<br />

y<br />

H ) v B<br />

Where the first term is the induced field and second term is the source term. Normally,<br />

the area used for mCSEM data acquisition is very small to observe horizontal variation in<br />

the vertical magnetic field Hz. Therefore, term y z H and xH z are negligible. Finally, we<br />

have a simplified equation,<br />

E<br />

E<br />

x<br />

y<br />

1<br />

<br />

zH<br />

<br />

1<br />

<br />

zH<br />

<br />

y<br />

x<br />

v B<br />

y<br />

v B<br />

Noticeably, the horizontal particle motion (i.e. vx and vy) is source for the induction of the<br />

horizontal electric field (i.e. Ex and Ey).<br />

5. The Data<br />

We have analysed horizontal component electric field (i.e. Ex and Ey) data recorded at<br />

500 m depth on the ocean floor. The recording is made in the absence of transmitter<br />

current to understand the oceanic background noise. Time series is shown in Figure 2. It<br />

is evident that the strength and pattern of Ex and Ey are different.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

37<br />

x<br />

z<br />

z<br />

<br />

<br />

<br />

<br />

<br />

<br />

x<br />

z<br />

z<br />

<br />

<br />

<br />

<br />

<br />

<br />

(5)<br />

(7)<br />

(8)


In equation (8), the first term is much smaller than the second term (i.e.<br />

( 1/<br />

)<br />

H v B ), suggesting x-and y- component of velocity principally controls the<br />

z<br />

y|<br />

x<br />

y|<br />

x<br />

z<br />

strength of the Ey and Ex component of the electric field respectively. Evidently, the<br />

strength of Ex is higher than Ey (Figure 2). This suggests that the wave velocity in the ydirection<br />

will be higher than the x-direction (i.e. vy > vx). Further, as normally the surface<br />

wave moves towards the coast. Accordingly, the velocity component pointing towards the<br />

coast has higher velocity than the other horizontal component. The present data is<br />

recorded with a setting that the y-component of the receiver points towards the cost and<br />

constructs an angle of approx. 55 with it. Thus, the data acquisition setting as well<br />

favours for vy > vx. In general, electric field measurements characterize an average<br />

motion over the length of the antenna. Therefore electric field components can be used to<br />

derive information about the average velocity of the movements.<br />

Electric field (nV/m)<br />

Electric field (nV/m)<br />

200<br />

100<br />

0<br />

-100<br />

-200<br />

100<br />

50<br />

0<br />

-50<br />

-100<br />

Time series for E x<br />

50 100 150 200 250 300<br />

Time (minute)<br />

Time series for E<br />

y<br />

50 100 150 200 250 300<br />

Time (minute)<br />

Figure 2: Five hour times-series of two horizontal components of electric fields. Ex (top) and Ey<br />

(bottom). The recording is made at the seafloor, 500 m below the sea surface.<br />

6. Power Spectral Density (PSD) and Spectrogram<br />

Power spectral density is calculated to study the power content of the frequencies. For<br />

the PSD calculation, steps followed are as follows:<br />

I. Inspection of time series (Figure 2): Inspection is done to recognize glitches or<br />

other outliers in the data that are not consistent with the rest of the time series.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

38


II. Detrending: Detrending ensures that the first Fourier coefficient does not<br />

dominate the PSD.<br />

III. Windowing: A Hanning window is applied to minimise the leakage of spectral<br />

density.<br />

IV. Calculation of discrete Fourier transform (DFT):<br />

V. Calculation of PSD: PSD is calculated by segment averaging method (Welch,<br />

1967).<br />

The PSD for Ex and Ey is shown in Figure 3. Over all spectral power is decreasing with<br />

increase in frequency. Four different slopes can be seen in the both PSD’s (Ex and Ey).<br />

Let the slope containing frequencies between 0.001 – 0.1 Hz, 0.1 - 2 Hz, 2 – 10 Hz and 10<br />

- 25 Hz represent respectively low, intermediate, sub-high and high frequency range.<br />

Evidently, in the low frequency range the power of the electric field rises sharply.<br />

Generally, external origin field mainly affects the low frequency range, as higher<br />

frequencies are filtered out by the ocean resulting in passage for low frequencies only<br />

(i.e. less than 0.3 Hz for a 500 m deep ocean) to the ocean bottom due to skin depth. The<br />

oceanic eddies (very long wavelength oceanic feature) are another low frequency source<br />

(Chave and Filloux, 1984). Here, these two sources are presumed for the sharp rise in<br />

power in the low frequency range (0.001 – 0.1 Hz). In the range of intermediate<br />

frequency, four peaks are evident, at 0.2, 0.3, 0.4 and 1 Hz, in the PSD representing that<br />

this frequency range is receptive for the various oceanic flows close to seafloor. The subhigh<br />

frequency range is nearly flat. A flat PSD, generally, corresponds to a field which<br />

contains equal power within a fixed bandwidth which resembles noise. In the high<br />

frequency range a sharp rise is evidenced which is presumed due to digitisation noise.<br />

The frequency and power information of a PSD is insufficient to characterise the<br />

source process corresponding to the frequencies. Time preservation may help in this<br />

subject. Together with the frequency and power, information about time may help in<br />

identifying and characterising the source nature of process. The localised time process<br />

may represent a source progressing in an interval of time while an ambient time process<br />

reflects a continuous process. For the purpose therefore spectrograms are plotted which<br />

preserve the time information together with frequency and power. The spectrogram is<br />

the discrete-time Fourier transform for a sequence, computed using a sliding window.<br />

For a spectrogram calculation, the time series is divided in to segments equal to the<br />

length of the hamming window. Each segment overlaps 50% of the samples with the<br />

adjacent segment and then PSD is calculated for a defined frequency length. Again and<br />

again PSD is calculated by sliding the window to build a spectrogram. The Spectrograms<br />

corresponding to time series shown in Figure 2 is shown in Figure 5. There are three<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

39


distinct anomalous features evident in spectrograms. One feature is corresponding to<br />

time approx. 165 min (like spike) and other two are at frequencies 1, 0.3 Hz. It is difficult<br />

to confidently propose a source nature for the spiky feature at approx. 165 min. This<br />

could be due to either by a regional earthquake or by a volcano like feature. In general,<br />

regional earthquakes are found in band frequencies of 0.1 to 1 Hz and here the spiky<br />

feature correspond the same range. The other two features at frequencies 1 and 0.3 Hz<br />

are respectively by the microseisms and swell, that will be discussed in the next section.<br />

E x (nV/m) 2 /Hz<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -2<br />

Power Spectral Density (E x )<br />

10 -1<br />

Frequency (Hz)<br />

0.2 & 0.4 Hz<br />

0.3 Hz<br />

10 0<br />

1 Hz<br />

10 1<br />

E y (nV/m) 2 /Hz<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -2<br />

Power Spectral Density (E y )<br />

0.2 & 0.4 Hz<br />

10 -1<br />

Frequency (Hz)<br />

Figure 3: Spectra of the two horizontal electric components Ex and Ey corresponding to time series<br />

shown in Figure 2. The graphs show the PSD vs. frequency. The anomalous peaks are visible at 0.2, 0.4 and<br />

1 Hz. There is also a weaker peak at 0.3 Hz. Four slopes can be observed in the spectra (pink dashed lines).<br />

7. Microseisms and swell<br />

Microseisms are like a soft earth tremor originating with in the ocean by non linear<br />

interaction of oceanic waves, which causes a continuing oscillation of the ocean floor. The<br />

broad frequency range for microseisms is between 0.05 to 1 Hz, which mainly depends<br />

on the ocean depth and oceanic conditions.<br />

Languet-Higgins (1950) proposed a mechanism for microseisms (0.05 -2.0 Hz) and<br />

showed that if two identical progressive waves travelling in opposite directions interact<br />

with each other, there is a second order pressure term effect which does not vanish with<br />

depth and can thus reach the deep ocean bottom (Figure 4). Consider two surface waves<br />

of frequencies f1 and f2 (f1f2), moving with approx. same velocity in opposite direction.<br />

Let the wave number of frequencies are respectively k1 and -k2, and then k1-k2.<br />

Interaction of these waves will leave behind a wave with very small wave number {i.e.<br />

k1+ (-k2) = diminutive} and very large wavelength. The large wavelength is capable of<br />

creating a pressure disturbance effectively at the ocean floor. The amplitude of the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

40<br />

0.3 Hz<br />

1 Hz<br />

10 0<br />

10 1


pressure disturbance is proportional to the product of the interacting wave heights and<br />

the frequency. These pressure fluctuations in the water column might then excite<br />

Rayleigh waves in the solid earth and be observed as microseisms.<br />

In short, the favourable oceanic conditions for microseisms generation are:<br />

1. Shoreline geological setting creating grounds for nonlinear interaction of<br />

surface waves.<br />

2. Near shore reflection of high frequency surface waves and thereafter head<br />

on interaction.<br />

3. A fast moving storm creating a sequence of wave in different directions.<br />

4. High frequency wave interaction may generate whitecaps (a wind blown<br />

wave whose crest is broken and appears white), which leads in acoustic<br />

energy transmission to ocean bottom.<br />

The depth of the ocean is another important factor in microseisms generation. The wave<br />

generation in the ocean depends on the wind velocity, which become efficient when the<br />

wave velocity is close to the phase<br />

velocity of the ocean waves. Velocity of<br />

ocean wave (V) is, roughly speaking,<br />

(gD), where D is ocean depth and g is<br />

acceleration due to gravity. Therefore,<br />

for a 4 km deep ocean and a 500 m<br />

shallow ocean, the velocity V is<br />

Figure 4: Microseisms mechanism (from Elgar approximately 200 m/s and 70 m/s<br />

(2008)). Nonlinear interaction of short opposing respectively. A typical wind velocity<br />

waves leads to a long wavelength wave generation, rarely exceeds few tens of m/s. This<br />

which reaches the seafloor and generates<br />

suggests wind velocities are quite close<br />

i i<br />

to the oceanic velocities especially in the shallow oceans. Therefore the generation of<br />

microseisms is likely to be efficient in shallow oceans (Tanimoto, 2005).<br />

Swells are a kind of oceanic surface waves. They are often created by the<br />

breaking of storms thousands of kilometres away from the seashore. The distance allows<br />

the waves comprising the swells to become more stable, clean, and continuous as they<br />

travel toward the coast. For microseisms generation, in general, higher frequency waves<br />

are more efficient than the lower frequency waves like swells. Swells are more<br />

directional and therefore chances for non linear interaction is not as much as of higher<br />

frequency gravity waves (Webb and Cox, 1986), which are generated with in the ocean by<br />

the influence of gravity at the interface involving the density contrast. In Figure 3 and 5,<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

41


the anomalous feature close to 1 Hz is likely due to a microseism. A microseism is a time<br />

localised feature, the duration of which depends on the time of effective nonlinear wave<br />

interaction. It is evident from Figure 5 that close to 1 Hz, there are four time localised<br />

features (Ey component has more clear appearance than Ex), corresponding to time 8, 40,<br />

90 and 250 min approx. Further evidence in support of microseisms is the observed<br />

frequency (i.e. 1 Hz).<br />

The acquired electric field data was recorded in a shallow ocean (500 m depth).<br />

Normally, a shallow ocean may contribute the microseisms in high frequency range (i.e<br />

close to 1 Hz) as analogous to high frequency even a comparable wavelength (not<br />

necessary a huge wavelength) may reach ocean bottom to produce microseisms. The<br />

mechanism of microseisms generation suggests that maximum electric field power will be<br />

experienced in the component perpendicular to the direction of wave propagation. Here,<br />

the velocity component Vy > Vx as amplitude of Ex > Ey. The strong power of microseisms<br />

in the Ey component compared to the Ex component further supports to interpret 1 Hz<br />

frequency as microseisms.<br />

Figure 5: Spectrogram for Ex & Ey. Electric field power is colour coded in dB and displayed as function of<br />

time and frequency. Two anomalous features are visible, one at 1 Hz and the other covering a broad<br />

frequency range, having max. power at 0.2 & 0.4 Hz.<br />

The individual feature ambient in time at 0.3 Hz (Figure 3 & 5) is interpreted here<br />

as a swell. A swell consists of long-wavelength surface waves which are more stable in<br />

their directions and frequency than normal wind waves. Swells are dispersive in nature<br />

and their frequency is given by 2 gs tanh(sD), where g is acceleration of gravity; s =<br />

2/ is wave number, is the wavelength of the swell and D is the depth of ocean. A<br />

calculation suggests that a swell of frequency 0.3 Hz corresponds to a wavelength of 18<br />

m for 500 m deep oceanic water. The obtained wavelength is well in range for the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

42


wavelengths of swells. As 0.3 Hz frequency is evident throughout the time in<br />

spectrogram (Figure 5), this is another strong support for 0.3 Hz to stand for a swell.<br />

7. Microseisms during mCSEM data recording<br />

Microseisms are a powerful feature to effectively generate EM signal at the ocean floor.<br />

They are quite capable of masking the mCSEM signals. We have an example comparing<br />

the power of microseism and mCSEM transmitter current. In Figure 5, we have seen the<br />

spectrograms of mCSEM the time-series recorded in the absence of transmitter current.<br />

On the other hand, Figure 6 represents spectrograms of the mCSEM data when<br />

transmitter was transmitting signals. Systematic decay in the power is evident in Figure<br />

6, representing transmitter is moving away from the receiver. A gathering of high power<br />

(yellow colour) patch is evident corresponding to frequency 0.1 Hz and time 200 min,<br />

may be representing a microseism. Clearly, the power of a microseism is significant<br />

enough to contaminate the mCSEM data. It is as well evident that the effect of swell is<br />

feeble in the presence of a transmitter current. For a larger transmitter receiver distance<br />

this effect may be significant.<br />

Microseisms Microseisms<br />

Figure 6: Presence of a microseism during mCSEM data recording. Electric field power is colour coded<br />

in dB and displayed as function of time and frequency. Corresponding to frequency 0.1 Hz and time 200 min<br />

(approx), a yellow coloured high power patch is evident, which may be is by a microseism. The other efficient<br />

high power patches in the spectrogram is by the transmitter current<br />

8. Conclusions<br />

By the analysis seafloor electric field data, we provide evidence for the observation of<br />

microseisms and swell in mCSEM data. Power spectral density (Figure 3) and<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

43


spectrogram (Figure 5) clearly present the evidence for the possible presence of<br />

microseism and swell corresponding to the frequencies 1 Hz and 0.3 Hz respectively.<br />

Swells are time ambient oceanic feature, which can be clearly marked in the<br />

spectrograms. The stability of 0.3 Hz in spectrograms indicates that oceanic conditions<br />

were quite stable during the recording of present mCSEM data.<br />

Figure 6 shows that the power of a microseism is significant even when the<br />

transmitter current is ‘on’. Overall, the time series analysis suggests that features like<br />

microseism and swell induce a significant electric field at the seafloor to contaminate the<br />

mCSEM data. Proper measures for removal of the noise are essential for a better<br />

interpretation of mCSEM data.<br />

Acknowledgement<br />

We thank KMS Technologies –KJT Enterprises Inc. for sponsoring the work.<br />

References<br />

1. Bhatt, K. M., A. Hördt, P.Weidelt, T. Hanstein, 2009. Motionally induced<br />

electromagnetic field within the ocean, 23. Kolloquium Electromagnetische<br />

Tiefenforschung (EMTF), Brandenburg, Germany, September-October, this issue.<br />

2. Chave, A.D., and C.S. Cox, 1982. Controlled electromagnetic sources for measuring<br />

electrical conductivity beneath the oceans, 1, forward problem and model study, J.<br />

Geophys. Res., 87, 5327-5338.<br />

3. Chave, A.D., and D.S. Luther, 1990. Low-frequency, motionally induced<br />

electromagnetic fields in the ocean, 1, theory, J. Geophys. Res., 95, 7185-7200.<br />

4. Chave, A.D., and J.H. Filloux, 1984. Electromagnetic induction fields in the deep<br />

ocean off California: oceanic and ionospheric sources, Geophys. J. R. astr. Soc., 77,<br />

143-171.<br />

5. Constable, Steven and Leonard J. Srnka, 2007. An introduction to marine controlledsource<br />

electromagnetic methods for hydrocarbon exploration, Geophysics, vol. 72, no.<br />

2, p. wa3–wa12.<br />

6. Crews, A., and J. Futterman, 1962. Geomagnetic micropulsation due to the motion of<br />

ocean waves, J. Geophys. Res., 67, 299-306.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

44


7. Dahm T, F. Tilmann, J. P. Morgan, 2006. Seismic broadband ocean-bottom data and<br />

noise observed with free-fall stations: experiences from long-term deployments in the<br />

North Atlantic and the Tyrrhenian Sea, Bull. Seismol. Soc. Am, 96, 647–664, doi:<br />

10.1785/0120040064.<br />

8. Elgar, S. ,2008. Ripples run deep, Nature, Vol 455, Page 888.<br />

9. Longuet-Higgins, M.S., E. Stern, H. Stommel, 1954. The electrical field induced by<br />

ocean currents and waves with application to the method of towed electrodes, Pap.<br />

Phys. Oceanog. Meteorol., 13(1),1-37.<br />

10. Longuet-Higgins, M.S., 1950. A theory of the origin of microseisms, Philos.Trans. R.<br />

Soc. London, Ser. A, 243, 1–35.<br />

11. Podney, W., 1975. Electromagnetic fields generated by ocean waves, J. Geophys. Res.,<br />

80, 2977-2990.<br />

12. Sanford, T.B., 1971. Motionally induced electric and magnetic fields in the sea, J.<br />

Geophys. Res., 76, 3476-3492.<br />

13. Tanimoto, T., 2005. The oceanic excitation hypothesis for the continuous oscillations<br />

of the Earth, Geophy. J. Int., 160, 276-288.<br />

14. Webb, S.S. and C.S. Cox, 1985. Observation and modelling of seafloor microseisms, J.<br />

Geophys. Res., 91, B-7, 7343-7358.<br />

15. Welch, P. D. 1967. The use of fast-Fourier transform for the estimation of power<br />

spectra: A short method based on time averaging over short, modified periodograms,<br />

IEEE Transactions of the Audio and Electroacoustics, AUlS, 70-13.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

45


Motionally Induced Electromagnetic Field within the Ocean<br />

K. M. Bhatt 1 , A. Hördt 1 , P.Weidelt 1,* and T. Hanstein 2<br />

1<br />

Inst. f. Geophysik u. Extraterrestrische Physik, TU Braunschweig<br />

2<br />

KMS Technologies - KJT Enterprises Inc<br />

Abstract<br />

The contribution of motionally induced electromagnetic (EM) fields at the seafloor is generally<br />

considered small, but since the characteristic reservoir signal in marine controlled source<br />

electromagnetic (mCSEM) data is also small, the inclusion of the motional induction<br />

contribution in modelling the signal will enhance the probability of reservoir detection. Here,<br />

we have studied the electromagnetic induction caused by ocean water flow with in earth’s<br />

magnetic field.<br />

When a charge particle moves with certain velocity in earth’s magnetic field, it<br />

experiences a Lorentz force. The action of Lorentz force generates a secondary electric field<br />

through galvanic and inductive processes. For the mathematical formulation, we considered<br />

Lorentz electric field as a source in the corresponding set of Maxwell’s equations. We solved<br />

these Maxwell's equations for a 1D model and velocity structure using two different Green’s<br />

function i.e. a two half space Green’s function and a layered Green’s function. The layered<br />

Green’s function is especially useful in studying the sensitivity of electric and magnetic field for<br />

different conductivity structures in the earth. Further, the signal variation with the conductivity<br />

of ocean, depth of ocean and wave velocity is studied to profoundly understand the effect of<br />

these parameters.<br />

________________________________________________________________________________<br />

* Deceased<br />

1. Introduction<br />

Oceanic water movements in the earth’s magnetic field induce electric and magnetic<br />

fields in the ocean. The induced fields are signal for oceanographic and seismological<br />

applications but noise for magnetotelluric (MT) and marine controlled source<br />

electromagnetic (mCSEM) applications. Oceanographers and seismologists use the fields<br />

respectively to study the velocity structure and the seismic background noise. The MT<br />

and mCSEM signals, which are generally used for lithospheric and hydrocarbon<br />

exploration studies, are contaminated by the induced field and act as noise. Our prime<br />

focus of motional induction study here is on mCSEM problems; nevertheless the results<br />

are also applicable for other scientific applications.<br />

Generally, motionally induced noise in seafloor mCSEM data is considered small,<br />

but since the characteristic reservoir signal is also small, the understanding and possible<br />

removal of noise may be essential to increase the number of possible target reservoirs.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

46


The electromagnetic induction investigation has a long history. Initially, by an<br />

observation of deflection in the galvanometer by stream waves Faraday (1832) concluded<br />

that flow of water will induce electric currents, which was afterwards measured<br />

experimentally by Young et al (1920). Later it was almost neglected and got re-attention<br />

after World War II. Longuet-Higgins et al., (1954) initiated the investigation of electric<br />

field induction by surface waves. Crews and Futterman (1962) investigated the magnetic<br />

field induction by oceanic movement. Further, Sanford, (1971) extended the theory of<br />

motional induction by considering three-dimensional water flows. A comprehensive study<br />

of the theory has been made by Podney (1975), who generalised and extended the<br />

previous work by removing the restrictive assumptions imposed on ocean-water velocity<br />

field. Chave (1983) generalised the EM induction process by considering driving electric<br />

<br />

field term (i.e. v B ) and further, Chave and Luther (1990) re-examined the motional<br />

induction problem.<br />

A primary purpose of this paper is to present a generalised simple illustrative<br />

theory for the problems of motional induction. For the purpose, we have formulated a set<br />

of Maxwell’s equation for our problem. These equations are simplified for electric and<br />

magnetic field components by considering a horizontally progressing ocean wave and<br />

then solved by using the Green’s function, with appropriate boundary conditions. Here,<br />

we solved the problem with two different Green’s functions. For a uniformly conductive<br />

earth, a two half space Green’s function is utilised and for a layered earth, a layered<br />

Green’s function is utilised. A two half space Green’s function doesn’t offer reflections<br />

because of homogeneous conductivity consideration and thus expresses only the case of<br />

downward progressive diffusive waves. The layered Green’s function includes both<br />

downward and upward propagating diffusive waves offered by the layered boundaries.<br />

2. Problem formulation in terms of Maxwell’s equation<br />

2.1 Basics<br />

An electrically conducting fluid like ocean consists of charged particles. The particles in<br />

the ambient geomagnetic field experience a deflective Lorentz force. If ‘ v ’is the velocity<br />

of charge particle ‘q’ moving in the geomagnetic field ‘ B0 ’, then the Lorentz force is:<br />

<br />

q(<br />

v B )<br />

(1)<br />

FL 0<br />

The charge ‘q’ experiences the deflecting force ‘ FL ’ because of the action of an electric<br />

field which we call as Lorentz electric field ( EL ),<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

47


Therefore, L 0<br />

<br />

FL<br />

<br />

qEL<br />

(1-a)<br />

<br />

E<br />

<br />

v B<br />

(2)<br />

The field ‘ EL ’ generates a secondary electric field ‘ E ’, mainly by two processes-<br />

(1) Galvanic process: At locations, where ‘ EL ’ has a component parallel to the<br />

conductivity gradient '<br />

'<br />

( is conductivity in S/m), space/surface charges are<br />

accumulated. The accumulated charges galvanically create a secondary field ‘ E ’,<br />

even if the wave velocity is constant.<br />

(2) Inductive process: The magnetic field ‘ HL <br />

’ is created by the current density<br />

<br />

’ i.e.<br />

‘ JL<br />

<br />

JL<br />

(<br />

v B0<br />

)<br />

<br />

H J<br />

L<br />

L<br />

when the velocity ‘ v ’ changes with time. This induces a secondary electric field E <br />

via the law of induction.<br />

We assume that all the hydrodynamics, including Coriolis force, is included in the given<br />

‘ v ’. We do not care about the sources of forces. Finally, for a stationary frame of<br />

reference, the current density in the Ohms law is given by<br />

<br />

( E E ) ( E v B )<br />

(3)<br />

J L<br />

0<br />

Here, ‘ ’ is the conductivity of the fluid, ‘ E ’ is the current term generated by both a<br />

galvanic and an inductive process and ‘<br />

<br />

( v B ) ’ is the source current term for motional<br />

0<br />

induction case. Under the quasi-static approximation, the set of Maxwell’s equation to be<br />

solved for the motional induction case is<br />

<br />

H (<br />

E v B0<br />

)<br />

<br />

(4- a, b)<br />

E <br />

H<br />

0<br />

t<br />

In an exact formulation, the ambient magnetic field is the sum of geomagnetic field ‘ B0 ’,<br />

the external magnetic field by ionospheric currents ‘ Bext <br />

’, the field generated by small<br />

local anomalies<br />

<br />

‘ B lano<br />

’ and the motionally induced field ‘ Bmi <br />

’ i.e.<br />

<br />

( B<br />

<br />

B<br />

<br />

B <br />

<br />

B ) . In general, the last three terms are orders of magnitude<br />

B 0 ext l ano mi<br />

smaller than the geomagnetic field B0 . Therefore, for computations, the local value of<br />

earth’s magnetic field would be a good choice.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

48


3. A halfspace model<br />

Let the north, east and positive downward directional depth in Cartesian coordinate<br />

system be represented by xˆ , yˆ and zˆ . Consider a two halfspace model with the boundary<br />

at the earth surface (i.e. z 0 ). Let the insulating air halfspace extend in the negative<br />

vertical upward direction while the earth halfspace extends in positive vertical downward<br />

direction from the boundary surface (Fig 1). The conductivity of the earth’s half space is<br />

=3.33 S/m, which contains an ocean and sediments below. In general, the conductivity<br />

of the ocean depends on dissolved ions content and their mobility, which is primarily a<br />

function of temperature and pressure. The conductivity in warm, top shallow water is<br />

normally high (5 S/m) compared to cold, deep water (2.5 S/m). For simplicity, oceans<br />

can be considered homogeneous as the range of conductivity difference (i.e. 2.5–5 S/m) is<br />

small. Consider with depth d=1000 m ocean moving with a velocity v in the x-direction<br />

(Fig 1). Below the ocean, a subsurface of stationary sediments (i.e. v = 0, motionless)<br />

exists.<br />

z=0<br />

+z<br />

z=d<br />

V x<br />

Ocean (v0) (v0) (v0)<br />

, , 0<br />

Sediments (v = 0)<br />

, , 0<br />

Air Half-Space<br />

x<br />

Earth Half-Space<br />

Fig 1. The two halfspace model. z = 0 is the<br />

boundary between earth and air halfspace. Depth is<br />

considered positive in the downward direction. The<br />

earth halfspace consists of two layers. The top layer<br />

represents the ocean having non-zero velocity in<br />

the x-direction. The depth of the ocean is d. Below<br />

the ocean, a second layer of static (i.e. velocity<br />

zero) sediments extends. The conductivity and<br />

magnetic permeability of earth halfspace is and<br />

0 respectively.<br />

<br />

<br />

Let B0 B0<br />

zˆ and v v x ( z,<br />

t)<br />

xˆ<br />

This will generate a horizontal electric field perpendicular to the velocity direction and a<br />

horizontal magnetic field perpendicular to the electric field, and thus we can write:<br />

<br />

<br />

E ( z,<br />

t)<br />

yˆ and H ( z,<br />

t)<br />

xˆ<br />

From eq. (4), we have<br />

E y<br />

2<br />

H x<br />

EBv Assume a harmonic time dependence for simplicity,<br />

~ it<br />

Then E ( z,<br />

t)<br />

E ( z)<br />

e , H<br />

y<br />

y<br />

z Ey 0<br />

t y 0 t x<br />

(5)<br />

x<br />

( z,<br />

t)<br />

x<br />

v<br />

x<br />

( z,<br />

t)<br />

~ it<br />

v ( z)<br />

e ,<br />

it<br />

H ( z)<br />

e<br />

~ <br />

and eq. (5) reduces to<br />

2~<br />

2 ~<br />

E k (z) E (z) - g(z)<br />

(6)<br />

z<br />

y<br />

y<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

49<br />

x


where,<br />

2<br />

g(z) k (z)<br />

~<br />

v x(z)<br />

B0<br />

is the source term, k(<br />

z)<br />

i0(<br />

z)<br />

represents the<br />

electromagnetic damping phenomenon. It is complex such that amplitude decay is<br />

associated with a phase shift with respect to the field at the surface. The Green’s function<br />

G( z | z0<br />

) corresponding to equation (6) is defined by<br />

2<br />

2<br />

G 0<br />

0<br />

0<br />

( z | z ) k (z)G(z | z ) - (z<br />

- z )<br />

(7)<br />

Crosswise multiplication with G and y E<br />

~<br />

and subtraction yields<br />

<br />

~<br />

G<br />

E<br />

~<br />

E <br />

~<br />

G)<br />

- g(z) G E (<br />

z z ) (8)<br />

z ( z y y z<br />

y 0<br />

The integration of (8) over depths z yields the solution<br />

d<br />

~<br />

E y(z0)<br />

g(z) G(z | z0<br />

) dz , for - z0<br />

(9)<br />

0<br />

i.e. a convolution of the source term with the Green’s function. The integration, which<br />

formally must be carried out from minus to plus infinity, reduces to 0 to d, because that is<br />

~<br />

y 0<br />

where the source is nonzero. For the horizontal electric field E (z ) calculation at any<br />

desired depth, knowledge of the Green’s function is required.<br />

3.1 Green’s function<br />

3.1.1. Half space<br />

The Green's function is commonly used to solve inhomogeneous boundary value<br />

problems. The solution by means of Green’s function gives a special advantage because<br />

of its reciprocity property, which states ‘relationship between a oscillating source and the<br />

resulting field at some point of observation is unchanged even if the observation and<br />

source points are interchanged’. Using the boundary conditions, the Green’s function is<br />

calculated for the halfspace. Calculation is as follows:<br />

For an arbitrary small , equation (7) follows,<br />

( z | z ) G(<br />

z | z ) 1<br />

(10)<br />

zG<br />

0 0 z 0 0<br />

The second boundary condition is that G( z | z0)<br />

is continuous at z z0<br />

i.e.<br />

G( z0<br />

| z0<br />

) G(<br />

z0<br />

| z0<br />

) 0<br />

(11)<br />

For the uniform halfspace,<br />

<br />

A e<br />

kz<br />

G ( z | z0<br />

) <br />

Be<br />

kz<br />

for z z0<br />

for z z0<br />

(12-a,b)<br />

A and B are determined from (10) and (11) which yields the Green’s function for the half<br />

space<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

50


3.1.2 Two halfspaces<br />

G(<br />

z | z<br />

0<br />

)<br />

1<br />

2k<br />

k|<br />

zz<br />

|<br />

0<br />

e<br />

(13)<br />

In order to consider the diffusive reflection and transmission effects due to boundaries,<br />

let us consider two uniform half spaces with k( z)<br />

k 0 in the air (i.e. z0). The Green’s function, for z0>0, can be written as<br />

G(<br />

z | z<br />

1 k | z z<br />

e<br />

) <br />

2k<br />

k z<br />

T e 0<br />

<br />

0<br />

|<br />

R e<br />

kz<br />

for z 0<br />

0 (14)<br />

for z 0<br />

where, T and R are respectively transmission and reflection coefficients. The continuity of<br />

G and zG<br />

at z 0 yields<br />

1 k | z z | k - k k(<br />

z z ) <br />

e<br />

0<br />

0 e<br />

0<br />

<br />

2k<br />

<br />

k k0<br />

G(<br />

z | z ) <br />

<br />

1<br />

<br />

k z kz<br />

e 0 0<br />

k k<br />

0<br />

for z<br />

0,<br />

z<br />

for z 0,<br />

z<br />

0<br />

0<br />

0 (15)<br />

In particular, for the insulating air halfspace k 0 0 , the two halfspace Green’s functions<br />

take the form<br />

1 k | z z | k(<br />

z z ) <br />

<br />

e<br />

0 e<br />

0<br />

<br />

G(<br />

z | z ) 2k<br />

1<br />

<br />

kz<br />

e 0<br />

<br />

k<br />

for z 0,<br />

z<br />

for z 0,<br />

z<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0 (16)<br />

3.1.3. A simple model and field expression<br />

Equation (9) includes the source term g(z), which depends on conductivity and velocity,<br />

both are depth dependent. For simplicity, consider a uniform horizontal velocity i.e.<br />

v(z) v0<br />

, 0 z d . Let the depth dependent conductivity be zero in the air and<br />

3.<br />

33 S / m for the earth i.e.<br />

(z)<br />

<br />

<br />

, z 0<br />

0 , z 0<br />

The equation (9), (16) and (4), yields the horizontal electric and magnetic field expression<br />

for the full space.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

51<br />

0


Electric field in full space:<br />

~<br />

Ey<br />

( z0<br />

) v<br />

~<br />

Ey<br />

( z0<br />

) v<br />

~<br />

E ( z ) v<br />

Magnetic field in full space:<br />

H ~<br />

H ~<br />

H ~<br />

y<br />

x<br />

x<br />

x<br />

0<br />

( z<br />

( z<br />

( z<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

) 0<br />

B ( 1 e<br />

B ( 1 e<br />

B<br />

0<br />

0<br />

0<br />

e<br />

kv0<br />

B<br />

) <br />

i<br />

kv0<br />

B<br />

) <br />

i<br />

kz<br />

0<br />

0<br />

0<br />

0<br />

kd<br />

kd<br />

)<br />

cosh( kz<br />

))<br />

, z<br />

0<br />

,0 z<br />

0 sinh( kd)<br />

, z d<br />

e<br />

e<br />

kd<br />

kz<br />

sinh( kz<br />

0<br />

0<br />

)<br />

, z<br />

0 sinh( kd)<br />

, z d<br />

0<br />

0<br />

0<br />

0<br />

,0 z<br />

0<br />

0<br />

0<br />

d<br />

d<br />

(17-a, b, c)<br />

(18-a, b, c)<br />

The graphical response of (17) and (18) is shown in Fig (2). The response is calculated for<br />

a halfspace model (Fig 1) of a conductivity of 3.33 S/m containing 1000 m thick layers of<br />

ocean and sediments. The magnetic permeability is kept constant for both halfspace and<br />

is equal to free space permeability i.e. 0 =410 -7 Vs/Am. The ocean has a homogeneous<br />

velocity of 10 cm/s in an ambient geomagnetic field of 510 -5 T. The responses are<br />

studied for frequencies 0.001, 0.01, 0.1, and 1 Hz. In general, the electric field becomes<br />

gradually weaker with depth (Fig 2). For 1 Hz, rather than weakening gradually, the<br />

electric field amplitude increases and becomes strongest with in the ocean (at 400 m<br />

depth). Further, the field Bx is zero at the surface and progressively becomes stronger<br />

with respect to depth. At the ocean bottom, it offers strongest amplitude. The zilch of Bx<br />

~<br />

at the earth surface is for the reason that in air halfspace field Ey<br />

is constant and<br />

~<br />

z y<br />

therefore E 0 in particular, which is BBx (eq. 4-b). Further, as far as frequency based<br />

variation are concerned, the smallest frequency Bx produces the strongest amplitude at<br />

the ocean bottom. The field Ey shows a contrary behaviour. Here the smallest frequency<br />

offers the weakest amplitude at the ocean floor. At large frequencies the field Ey can not<br />

reach to the ocean floor, because of the shallow skin depth, offers constant amplitude<br />

(Fig 3).<br />

The important result in Fig 2 & 3 is that the field Ey is smooth over the boundary<br />

between the ocean and sediment. On the other hand, the field Bx offers a sharp change in<br />

pattern there (at boundary). Note that the conductivity of the ocean and subsurface<br />

(sediments) are identical. They differ only in their velocity state (subsurface is static and<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

52


ocean is dynamic). Therefore the sensitivity of the field Bx at the boundary is interpreted<br />

as result of velocity change.<br />

Depth, (m)<br />

0<br />

-1000<br />

-2000<br />

0 1 2 3 4 5 6<br />

E (V/m)<br />

y<br />

0.001 Hz 0.01 Hz 0.1 Hz 1 Hz<br />

0<br />

Depth, (m)<br />

-1000<br />

-2000<br />

0 5 10<br />

B (nT)<br />

x<br />

15 20<br />

Fig 2. Variation of the horizontal comp. electric<br />

(Ey) and magnetic (Bx) field with respect to<br />

depth. Variation is shown for four frequencies viz.<br />

0.001, 0.01, 0.1 and 1 Hz. An ocean of conductivity<br />

=3.33 S/m extends from surface to 1000 m depth<br />

(i.e. 0 z 1000 m). Below the ocean (i.e. below<br />

1000 m depth) there is a sediment layer with the<br />

same conductivity as the ocean (i.e. =3.33 S/m).<br />

The horizontal line at 1000 m is to show the<br />

separation of these two boundaries.<br />

Depth, (m)<br />

0<br />

-1000<br />

-2000<br />

0 1 2 3 4 5 6<br />

E (V/m)<br />

y<br />

2 Hz 10 Hz 50 Hz<br />

Depth, (m)<br />

0<br />

-1000<br />

-2000<br />

0 0.5<br />

B (nT)<br />

x<br />

1 1.5<br />

Fig 3. Variation of the horizontal component<br />

electric (Ey) and magnetic (BBx) field with<br />

respect to depth. Variation is shown for three<br />

frequencies viz. 2, 10 and 50 Hz. Other information<br />

is the same as in Fig 2. It is evident thet at the ocean<br />

floor, the field Ey offers almost same amplitude for<br />

all the three frequencies. Bx is strongest and<br />

weakest respectively for the smallest and largest<br />

frequency.<br />

The field Ey (17-b) with in the ocean (i.e. 0 z0<br />

d ) consists of two terms. The first<br />

term is the kernel for the motional induction, while the second governs the depth<br />

dependent frequency based EM damping. From the expressions (17-b & 18-b), it is clear<br />

that the strength of both the fields Ey and BBx depends on the ocean depth (i.e. ocean<br />

-kd<br />

deepening) because of the factor e . Effect of the ocean deepening is studied for oceans<br />

of thickness (depth) varying from 1000 m to 9000 m. Observation depth is constant for all<br />

practical cases and is 1000 m (i.e. z0<br />

= 1000 m). Results are shown in Fig (4). Ey is<br />

strongest for the thickest ocean and weakens down gradually as ocean shallows up. The<br />

reverse is observed for the Bx case with weakest strength in the deepest ocean which<br />

progressively becomes stronger as the ocean gradually shallows up. The sensitivity for<br />

the deepening effect depends on frequency and therefore on skin depths. For that reason,<br />

the deepening effect is more effective at smaller frequencies. The skin depths, for a<br />

halfspace of 3.33 S/m, at frequencies 0.001, 0.01, 0.1, 1 and 10 Hz are approximately<br />

8664, 2739, 866, 273 and 86 m, respectively.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

53


The factor which may significantly influence the motional induction is the<br />

conductivity of the ocean. The results of the conductivity variation are shown in Fig 5.<br />

The field strength is observed at 1000 m depth for conductivities varying from 3 to 5 S/m.<br />

Observations are made for five different frequencies (i.e. 0.001, 0.01, 0.1, 1 and 10 Hz).<br />

The results suggest that the field Bx is more sensitive to conductivity variation than the<br />

field Ey. Further, lower frequencies are more responsive to conductivity variation than<br />

high ones because of the thicker skin-depth.<br />

Various Ocean Depths (m)<br />

9000<br />

8000<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

10 0<br />

0.001 Hz 0.01 Hz 0.1 Hz 1 Hz 10 Hz<br />

E y (V/m)<br />

Various Ocean Depths (m)<br />

9000<br />

8000<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

10 -5<br />

B x (nT)<br />

Fig 4. Deepening effect of an ocean: Conductivity<br />

of ocean kept constant (i.e. 3.33 S/m). For 1000 to<br />

9000 m deep ocean the field strengths are calculated<br />

at frequencies 0.001, 0.01, 0.1, 1 & 10 Hz. The y- and<br />

x-axis represents ocean depths and field strength<br />

respectively. The observation depth is constant and is<br />

1000 m. The magnetic component for 1 and 10 Hz is<br />

truncated for depths greater than 5000 and 2000 m<br />

respectively to save the Figure from masking and<br />

cluttering.<br />

10 0<br />

Conductivity (S/m)<br />

5<br />

4.8<br />

4.6<br />

4.4<br />

4.2<br />

4<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

0.001 Hz 0.01 Hz 0.1 Hz 1 Hz 10 Hz<br />

3<br />

0.5 1 1.5 2 2.5 3<br />

E (V/m)<br />

y<br />

Conductivity (S/m)<br />

5<br />

4.8<br />

4.6<br />

4.4<br />

4.2<br />

4<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3<br />

0 5 10 15<br />

B (nT)<br />

x<br />

20 25 30<br />

Fig 5. Conductivity variation effect: The<br />

conductivity of the ocean is varied from 3 to 5 S/m<br />

and the effect is studies for frequencies 0.001,<br />

0.01, 0.1, 1 & 10 Hz. The horizontal component<br />

electric and magnetic field strength, observed at<br />

1000 m ocean depth is shown w.r.t. conductivities<br />

for the chosen frequencies. Electric and magnetic<br />

components both are sensitive to the conductivity<br />

of the ocean. Sensitivity increases with the<br />

decreasing frequency.<br />

Equation (17) and (18) indicate that the field BBx is more sensitive to conductivity<br />

variation in vertical direction, as conductivity is effectively convolved, than the field Ey.<br />

But it would not be appropriate to discuss conductivity variation in vertical direction<br />

using a halfspace model. Therefore, we will be back on this issue with a justified layered<br />

model and a layered Green’s function.<br />

Moreover, the equation (17) and (18) implies that the increase in the wave<br />

velocity v0 will cause a constant shift in the Ey<br />

and Bx field. Three experiments are<br />

conducted for wave velocities of 10, 1 and 0.1 cm/s to study the effects. The results are<br />

shown in the Fig 5. A log scale is used for x-axis plotting for clarity reasons. As at the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

54


surface Bx=0, therefore the log scale plot starts from 100 m depth as 100 m is the 2 nd<br />

discretized layer of the ocean. The selection of velocities is such that the 1 st selection<br />

exceeds the 2 nd by factor of 10 and same is true for 3 rd and 2 nd . Evidently, response for<br />

the field Ey and Bx also exceeds by a static shift of factors of 10. The graph clearly<br />

illustrates that the velocity is an important parameter for practical simulation, any bias in<br />

velocity leads to the same bias in EM field.<br />

D epth, (m )<br />

0<br />

-1000<br />

10 -2<br />

-2000<br />

10 -1<br />

E y (V/m)<br />

0.1 m/s 0.01 m/s 0.001 m/s<br />

10 0<br />

10 1<br />

D epth, (m )<br />

0<br />

-1000<br />

10 -2<br />

-2000<br />

10 0<br />

B x (nT)<br />

Fig 6. Wave Velocity effect. Semi-log plot of<br />

horizontal electric and magnetic field component<br />

illustrating the effect of velocity change. The<br />

strongest velocity generates the strongest EM field.<br />

The Bx plots starts from 100 m depth as the strength<br />

of Bx at z=0 is zero and at z=100 m the next layer<br />

starts.<br />

4. Layered model<br />

4.1. Layered Green’s function<br />

10 2<br />

Depth (m )<br />

-0<br />

-1.000<br />

-2.000<br />

Let us consider a layered Green’s function defined as<br />

f=0.001 Hz f=0.01 Hz f=0.1 Hz f=1 Hz<br />

-3.100<br />

0 2 4 6<br />

Ey (Volt/meter)<br />

Depth (m )<br />

-0<br />

-1.000<br />

-2.000<br />

Ocean<br />

Sediments<br />

Reservoir<br />

Sediments<br />

-3.100<br />

0 5 10<br />

Bx (nT)<br />

15 20<br />

Fig 7. Variation of the horizontal component<br />

electric (Ey) and magnetic (Bx) field with<br />

respect to depth for a layered model. Variation is<br />

shown for four frequencies viz. 0.001, 0.01, 0.1 and<br />

1 Hz. An ocean of conductivity =3.33 S/m extends<br />

from the surface to 1000 m depth (i.e. 0 z 1000<br />

m). Below the ocean there, is a 1000 m thick<br />

sediments layer with conductivity =1 S/m. A 100<br />

m thick reservoir of conductivity 0.01 S/m is<br />

embedded at 2000 m depth. The horizontal line<br />

marks the boundary of different formation.<br />

1 k | z z | k z k z<br />

G( z | z )<br />

<br />

e 1 0 <br />

R e 1 <br />

R e 1<br />

0 <br />

<br />

0<br />

d ; 0 z d,<br />

0 z0<br />

d<br />

2k<br />

<br />

<br />

<br />

<br />

1<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

55


where, k1 i0<br />

is the electromagnetic damping in the ocean, R 0 describes the field<br />

1<br />

diffusing downward by reflection from the air-earth interface and R d describes the field<br />

diffusing upward by reflection from the seafloor (i.e. z d ). In this particular case the<br />

constants R 0 and R d are determined from boundary conditions:-<br />

a)<br />

( 0 | z ) 0 , Since ( z | z ) is constant in the air halfspace<br />

zG<br />

0<br />

G 0<br />

b) ( d | z ) b ( d)<br />

G(<br />

d | z ) 0 , where, b2 ( d)<br />

is determined via continuous<br />

transfer function.<br />

They yield,<br />

R<br />

0<br />

with<br />

zG<br />

0 0 2<br />

0<br />

1 R ce<br />

<br />

1 R e<br />

R<br />

c<br />

2k<br />

( d z<br />

)<br />

c<br />

0b<br />

<br />

b<br />

0<br />

2k<br />

d<br />

2<br />

2<br />

1<br />

1<br />

1 1<br />

k<br />

k<br />

0<br />

1<br />

1<br />

e<br />

k<br />

z<br />

1 0<br />

4.2 Layered model and Response<br />

;<br />

R<br />

d<br />

R c ( 1 e<br />

<br />

1 R e<br />

c<br />

2k<br />

z<br />

1 0<br />

2k<br />

d<br />

1 1<br />

)<br />

e<br />

k<br />

( 2d<br />

z<br />

)<br />

EM field responses calculated for a layered model are shown in Fig (6). The parameters<br />

of the model are tabled in table 1. The magnetic permeability of the each layer is equal to<br />

free space permeability (0). The response is computed for four frequencies viz. 0.001,<br />

0.01, 0.1 and 1 Hz. It is evident that the field Ey varies smoothly over layer boundaries,<br />

indicating its sensitivity to the vertically averaged conductivity rather than conductivity<br />

variation at layered boundaries. However, the field Bx senses each layer and offers a<br />

change in field strength at each layer boundary. Evidently, the magnetic field and its<br />

sensitivity for conductivity variation is frequency dependent. At a small frequency the<br />

field is strong and vice versa.<br />

Table 1: Layered Model<br />

Layers<br />

Thickness<br />

(m)<br />

Conductivity<br />

(S/m)<br />

Ocean 1000 3.33<br />

Sediments 1000 1<br />

Reservoir 100 0.01<br />

Below 1000 1<br />

1<br />

Other Parameters<br />

Wave Velocity=0.1 m/s<br />

Ext. Magnetic Field=510 -5 T<br />

Magnetic Permeability of each layer<br />

1<br />

=410 -7<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

56<br />

0


5. Discussion and Conclusion<br />

No doubt, the actual earth is 3D and therefore the electromagnetic response of the earth<br />

will always be more complex than the response of any simplified model. However, a<br />

simplified model, roughly representing actual earth layers can be used to validate results.<br />

The choice of the model dimension (i.e. 1D, 2D or 3D) sometimes has a large impact on<br />

the result. In particular, for mCSEM surveys, generally 2D data acquisition methodology<br />

is followed and therefore a 2D/3D modelling may lead to a better interpretation.<br />

However, 1D model formulations are simple and give good insight into the physics of the<br />

problems.<br />

Field strength depends on their components (depending on the physical process<br />

involved), whose production mainly depends on the involvement of the source<br />

components. For example, a 1D model in motional induction study results in Ey and Bx<br />

(two) components when a 1D source (i.e. wave-velocity in x-direction) is considered.<br />

However, a 2D source (i.e. wave-velocity in x-direction and a wave wavelength in ydirection)<br />

consideration results in Ey, Bx and Ez (three) components. This suggests, even a<br />

1D model, with a proper source formulation offers significant insight to the problems.<br />

Further, the simplicity of a 1D model is an important advantage. In this paper we<br />

thoroughly looked for the sources of the motional induction in the ocean, which is<br />

incorporated in 1D model for theoretical development. The horizontal (x-direction) wave<br />

motion and a vertical (z-direction) geomagnetic field consideration lead to the excitation<br />

of the field components ‘Ey’ and ‘Bx’. These fields illustrate some of important results and<br />

effects of the ocean dynamics in varying oceanic conditions.<br />

In general, for a uniform halfspace, there is gradual reduction and increase<br />

respectively in the strength of Ey and Bx. w.r.t. ocean depth (Fig 2). This conclusion is<br />

valid for frequencies with skin depth ‘’ greater than the ocean depth ‘d’. For the case<br />

when < d, Ey may show maximum amplitude somewhere in the ocean, rather than at<br />

the ocean surface. On the other hand in the ocean the field Bx always offers maximum<br />

strength at the floor and minimum (i.e Bx = 0) at the surface. Since Bx vanishes at the<br />

surface, the mCSEM surface measurement may offer significant noise-free signals there,<br />

if MT field is avoided. At the ocean floor as Bx always has maximum strength therefore it<br />

is necessary to correct mCSEM data for motionally induced field. It is evident from Fig<br />

(2), even in a halfspace where the conductivity of ocean and subsurface is same, Bx<br />

senses the boundary of the ocean and subsurface (i.e. sediments) though Ey does not see<br />

it. Still, an important question left to answer is ‘which field is/are sensitive to layer<br />

demarcation, either Ey or Bx or both?’ This issue will be conferred below later, in the<br />

paragraph with layered earth discussion.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

57


The fields Bx and Ey both are sensitive (Fig 5) to the conductivity variation of the<br />

oceans. The observations with several oceans differing in their conductivity suggests that<br />

at frequencies with < d, in general, the variation of ocean conductivity is negligible but<br />

at frequencies with > d, the more conductive ocean increases the strength of Ey and Bx<br />

field.<br />

The average depth of Atlantic, Pacific and Indian Ocean is respectively 3926 m,<br />

4282 m and 3963 m. Moreover, from place to place the oceanic depths are quite diverse<br />

and therefore the significance of deepening of the ocean is studied (Fig 4). For a fixed<br />

observation depth in ocean, the gradual increase in the strength of Ey and decrease in the<br />

strength of Bx is the consequence of deepening of the ocean. The component Ey and Bx<br />

show strongest and weakest strength respectively for the deepest ocean and vice-versa.<br />

For frequencies for which the skin depth is smaller than the ocean depth (i.e. < d), the<br />

further deepening of the ocean (greater than skin depth) is ineffective and therefore the<br />

field value may saturate with further deepening.<br />

Another important parameter is the ocean wave velocity. The study suggests (Fig<br />

5) that any bias in velocity value may cause same bias in field strength. The change in<br />

velocity does not modify the pattern with depth, but causes a static shift in the field.<br />

Expression for layered Green’s function has allowed us to study the EM field<br />

variation for different layers with in the earth. The layered model involves both resistive<br />

and conductive layers. Clearly, electric field Ey does not see the layered boundaries (Fig<br />

6) although the magnetic field clearly sees it by offering a change in the slope of Bx at the<br />

boundaries.<br />

Acknowledgement<br />

We thank KMS Technologies - KJT Enterprises for sponsoring the work.<br />

References<br />

1. Chave, A.D., and C.S. Cox, Controlled electromagnetic sources for measuring<br />

electrical conductivity beneath the oceans, 1, forward problem and model study, J.<br />

Geophys. Res., 87, 5327-5338, 1983.<br />

2. Chave, A.D., and D.S. Luther, Low-frequency, motionally induced electromagnetic<br />

fields in the ocean, 1, theory, J. Geophys. Res., 95, 7185-7200, 1990.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

58


3. Crews, A., and J . Futterman, Geomagnetic micropulsation due to the motion of ocean<br />

waves, J. Geophys. Res., 67, 299-306, 1962.<br />

4. Faraday, M., Bakerian Lecture-Experimental researches in electricity, Phil. Trans.<br />

Roy. Soc. London, Part 1, 163-177, 1832.<br />

5. Longuet-Higgins, M.S., E. Stern, H. Stommel, The electrical field induced by ocean<br />

currents and waves with application to the method of towed electrodes, Pap. Phys.<br />

Oceanog. Meteorol., 13(1),1-37, 1954.<br />

6. Podney, W., Electromagnetic fields generated by ocean waves, J. Geophys. Res., 80,<br />

2977-2990, 1975.<br />

7. Sanford, T.B., Motionally induced electric and magnetic fields in the sea, J. Geophys.<br />

Res., 76, 3476-3492, 1971.<br />

8. Young, F.B., H. Gerrard and W. Jevons, On electrical disturbances due to tides and<br />

waves, Phil. Mag. Ser. 6, 40,149-159.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

59


Three-dimensional finite element simulation of<br />

magnetotelluric fields incorporating digital elevation models<br />

S. Kütter, A. Franke-Börner, R.-U. Börner, K. Spitzer<br />

Institut für Geophysik, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 12, 09599 Freiberg<br />

1 Introduction<br />

Pronounced topography may have an important impact on the magnetotelluric (MT) response and its<br />

neglect may lead to severe misinterpretations. Common simulation techniques based on rectangular<br />

grids are generally not well suited to deal with arbitrary geometry. We therefore use vector finite<br />

element (FE) schemes formulated on unstructured grids to cope with realistic topography and/or<br />

bathymetry. As an example, we have chosen the volcanic island of Stromboli which is located in the<br />

Mediterranean Sea off the west coast of Italy. Stromboli is an extreme electric environment with very<br />

conductive sea water surrounding the steep topography of the resistive volcanic edifice. Moreover, the<br />

varying bathymetry and the topography of the other islands of the Liparian archipelago cause distinct<br />

effects on the magnetotelluric response. Due to the large conductivity contrast and these extraordinary<br />

features it can be expected that the magnetotelluric apparent resistivities as well as the phases show a<br />

very complex behaviour. The objective of this work is to incorporate digital elevation models of this area<br />

into our numerical MT simulations allowing for a realistic look at the electromagnetic (EM) induction<br />

phenomena in such a complicated environment. On top of its challenge for numerical simulation<br />

methods this volcano has always fascinated geoscientists (Fig. 1) and future interdisciplinary studies<br />

aim at investigating the inner structure and the processes that lead to the continuous mild eruptions<br />

recorded over the last 2 000 years (the so called ’Strombolian activity’).<br />

Figure 1: Map of the Aeolian Islands (left, Wikipedia (2010)) and Stromboli (right, SwissEduc (2010))<br />

2 Physical and numerical basics<br />

Based on Maxwell’s equations and assuming a harmonic time dependency e iωt of the incoming plane<br />

wave, the equation of induction in terms of the vector potential A reads as<br />

∇×μ −1 (∇×A)+(iωσ − ω 2 ε)A = 0, (1)<br />

where μ, σ, ε, ω and i are the magnetic permeability, the electric conductivity, the permittivity, the<br />

angular frequency and the imaginary unit, respectively. The magnetic field H and the electric field E<br />

are obtained by<br />

H = μ −1 (∇×A) and E = −iωA −∇V . (2)<br />

The scalar potential V was eliminated from eq. (1) by the gauge condition ψ = −iV/ω and substituting<br />

A for A −∇ψ which leaves H and E unchanged.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

60


To calculate A in a bounded domain Ω ⊂ R 3 , electric and magnetic insulation are required for all outer<br />

boundaries parallel (Γ ||) and perpendicular (Γ⊥) to the current flow, respectively:<br />

n × H = 0 on Γ ||,<br />

n × A = 0 on Γ⊥.<br />

Depending on whether the condition of electric insulation is applied to the boundaries parallel to the<br />

x- ory-direction, xy- oryx-polarisation data are obtained, respectively. At the top (Γtop) and bottom<br />

boundaries (Γbottom) we impose boundary values in terms of the magnetic field induced in a horizontally<br />

layered half-space (Wait (1953)):<br />

H⊥ =1Am −1<br />

on Γtop,<br />

H⊥ = Hj(z), j = x, y on Γbottom.<br />

For the construction of the full 3-D MT impedance tensor, both xy- andyx-polarization data are<br />

required. Arranging the x- andy-components of the electric and magnetic fields obtained for xy- and<br />

yx-polarization such that<br />

<br />

Ex(xy) Ex(yx)<br />

E =<br />

(5)<br />

H =<br />

Ey(xy) Ey(yx)<br />

Hx(xy) Hx(yx)<br />

Hy(xy) Hy(yx)<br />

<br />

, (6)<br />

the impedance tensor Z can the be obtained by Z = EH −1 . The off-diagonal elements of the impedance<br />

tensor are therefore<br />

Zxy = Ex(yx)Hx(xy) − Ex(xy)Hx(yx)<br />

Hx(xy)Hy(yx) − Hx(yx)Hy(xy)<br />

(3)<br />

(4)<br />

, Zyx = Ey(xy)Hy(yx) − Ey(yx)Hy(xy)<br />

. (7)<br />

Hx(xy)Hy(yx) − Hx(yx)Hy(xy)<br />

For any period T =2π/ω, the apparent resistivity ρa and phase φ can then be calculated by<br />

ρxy = 1<br />

ωμ |Zxy| 2 , ρyx = 1 2<br />

|Zyx|<br />

ωμ<br />

and φxy = arg(Zxy), φyx = arg(Zyx). (8)<br />

Inthefollowing,weusethetermsZxy mode or Zyx mode to distinguish xy- andyx-polarization<br />

from quantities derived by a combination of both (Nam et al. (2007)). An important property of<br />

time-harmonic EM field is the skin depth δ ≈ 503 ρ/f, whereδ, the resistivity ρ and the frequency<br />

f are given in [m], [Ω m] and [Hz], respectively. is determined, whereas δ, the resistivity ρ and the<br />

frequency f are given in [m], [Ω m] and [Hz], respectively.<br />

To solve the boundary value problem (1) - (4), the finite element method (FEM) is applied on unstructured<br />

tetrahedral meshes (see Börner (2010); Schwarzbach (2009)). Tetrahedral meshes are well<br />

suited for the spatial discretization of arbitrary 3-D geometries which occur when digital elevation<br />

models have to be incorporated into numerical simulations. In the 3-D simulations presented here,<br />

curl-conforming Nédélec elements with second-order basis functions are employed. In all experiments the<br />

FE discretization has been carried out using the Electromagnetics Module of the COMSOL Multiphysics<br />

package.<br />

3 Description of the digital elevation model<br />

The area used in the simulations extends from 38.4° to 39.2° N and 14.7° to 15.7° E. At 39° N, the<br />

distance between two degrees of latitude is 110.95 km, whereas the distance between two degrees of<br />

longitude is 86.51 km. Hence, the model area has a length of 86.51 km in the east-west (x) and 88.76 km<br />

in the north-south (y) direction.<br />

We have used two sets of digital terrain data. The ETOPO1 data set is a 1 arc-minute model and<br />

provides elevation values for both land and sea. It is available online at the National Geophysical Data<br />

Center (http://www.ngdc.noaa.gov) and gives a good approximation for the regional bathymetry. Since,<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

61


however, its spatial data density is not sufficient to describe Stromboli’s topography, a second data set<br />

has been used. These data are available from the Shuttle Radar Topography Mission (SRTM), a project<br />

to obtain high-resolution topographic data (http://srtm.csi.cgiar.org/). The horizontal resolution of<br />

the two data sets are listed in the following table. The digital elevation model for the area considered<br />

here has been obtained by blending both ETOPO1 and SRTM data sets.<br />

resolution x-direction resolution y-direction<br />

SRTM 0.0721 km 0.0925 km<br />

ETOPO1 1.4419 km 1.8491 km<br />

Interpolating the bathymetry data and fitting the two data sets along the coastlines leads to a digital<br />

elevation model as depicted in Fig. 2.<br />

φ [ o ]<br />

39.1<br />

39<br />

38.9<br />

38.8<br />

38.7<br />

38.6<br />

38.5<br />

38.4<br />

14.8 14.9 15 15.1 15.2<br />

λ [<br />

15.3 15.4 15.5 15.6 15.7<br />

o ]<br />

Figure 2: Data points obtained from ETOPO1 (coarse grid) and SRTM (fine grid resembling the islands) for<br />

bathymetry and topography (left), digital elevation model obtained by smooth interpolation onto the<br />

SRTM grid (right). Stromboli island is located at the centre of both images.<br />

4 Numerical studies<br />

The geometry of the volcanic island of Stromboli is incorporated into the simulation models in three<br />

different levels of increasing complexity. This approach enables us to better differentiate which part of<br />

the model has an influence on the electromagnetic fields and how distinct this influence is. The first<br />

model uses a frustum as the volcanic edifice embedded in a layered background consisting of an air<br />

layer, a sea layer and a substratum. In the second model the frustum is replaced with the topography<br />

of Stromboli. The edges of this topographic surface have been adjusted to the sea-floor. The third,<br />

most complex model uses both topography and bathymetry data.<br />

For the numerical computations three different computer architectures have been used:<br />

computer name architecture RAM<br />

klio 4Intel(R)Xeon(R)@3GHz 16 GB<br />

erato 4 Quad-Core AMD Opteron @ 2.5 GHz 128 GB<br />

RM 56 nodes based on Intel Xeon (E5450 @ 3 GHz) 16 GB per 8-core blade<br />

For each of the three models, apparent resistivities ρa and phases φ were calculated along profiles<br />

running along the sea-floor and over the slopes of the volcano for a period of T =10 3 s. Furthermore,<br />

sounding curves were determined for selected sites on the sea-floor and on the volcano to demonstrate<br />

the dependency of ρa and φ on frequency.<br />

4.1 Frustum model<br />

The geometry and appropriate conductivities of the frustum model as well as a map of the locations<br />

of the sounding sites are depicted in Fig. 3. Since there are large conductivity contrasts between the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

62


frustum, the sea and the air layer, a non-trivial behaviour of the electromagnetic fields is to be expected.<br />

The electric currents are mainly induced in the conductive sea layer and are compressed above the<br />

slopes. Moreover, the currents tend to concentrate below the sea surface due to the skin effect.<br />

Due to horizontal conductivity contrasts, electric charges accumulate at the interfaces between the<br />

volcano and the sea layer. Their surface density is proportional to −E ·∇σ. Therefore, an additional<br />

electric field is generated which leads to an increase of the total electric field towards the top of the<br />

volcano (around x = ±0.5 km in the profile, Fig. 4) and to a decrease towards the sea-floor (around<br />

x = ±5.6 km in the profile). This static shift effect results in discontinuities of the apparent resistivity<br />

at the base points of the frustum on the sea-floor, at the coastline points and at the top of the volcano.<br />

Unlike the apparent resistivity, the phase has a more continuous behaviour and appears to be a<br />

rather robust parameter even for pronounced topography. 40 km away from the center of the volcano<br />

(x, y = ± 40 km), both phase and apparent resistivity reach their values expected for the half-space<br />

(45°, 100 Ω m).<br />

-20<br />

[km] 0<br />

2<br />

air: 10 −14 Sm −1<br />

volcano: 0.02 Sm −1<br />

half-space: 0.01 Sm −1<br />

sea: 5 Sm −1<br />

32<br />

-40 -5.6 0<br />

[km]<br />

5.6 40<br />

<br />

<br />

<br />

<br />

Figure 3: Vertical cross-section through the frustum model (left) and plan view of the model domain (right)<br />

showing the locations of the sounding curves (T, N40, E40).<br />

To verify our results, two independent numerical codes have been used to provide reference datasets<br />

for the frustum model. More precisely, data have been obtained using a finite element package (FEP)<br />

by Schwarzbach (2009), and a finite difference package (FD) by Mackie et al. (1994).<br />

A comparison of the apparent resistivities with those obtained by FEP reveals that they slightly differ<br />

near the boundaries of the model and at the lower part of the volcano slopes (Fig. 4, left), where ρa is<br />

lower for FEP. Moreover, the peaks at the base of the frustum, at the coastline points and at the edges<br />

of the plateau are even more pronounced for FEP. Here, we assume an influence of the different spatial<br />

discretization of the model. The curves of the phase are nearly congruent for both FE codes.<br />

Apparent resistivities and phases obtained by FD (Fig. 4, right) show similar features in general.<br />

However, there are significant differences along the slopes of the volcano. We attribute this to the<br />

fact that in the FD code considered here electric field components are located perpendicular to cell<br />

faces of a rectangular cartesian grid. Problems in interpolating the discrete electric field components at<br />

the desired positions along the volcano slopes are thus to be expected. Finite element techniques on<br />

unstructured grids clearly show their superiority in this respect.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

63


[km]<br />

ρ a [Ω m]<br />

φ [ o ]<br />

−2<br />

0<br />

2<br />

4<br />

10 2<br />

10 1<br />

10 0<br />

50<br />

COMSOL<br />

FEP<br />

0<br />

−40 −30 −20 −10 0<br />

x [km]<br />

10 20 30 40<br />

Figure 4: Comparison of the profile curves of the frustum model, T =10 3 s.<br />

ρ a [Ω m]<br />

φ [ o ]<br />

10 2<br />

10 1<br />

10 0<br />

50<br />

COMSOL<br />

FD<br />

0<br />

−40 −30 −20 −10 0<br />

[km]<br />

10 20 30 40<br />

Fig. 5 depicts three sounding curves for the sites on top (point T in Fig. 3) and 40 km away from the<br />

volcano (E40, N40 in Fig. 3).<br />

Sounding curves for the volcano site T are displayed for periods between 10 −3 and 10 3 s (Fig. 5, left).<br />

For short periods, where the skin depth is small (δ ≈ 355 m for 50 Ω m and 10 −2 s), the resistivity of the<br />

volcano is reproduced and the phase reaches values around 45°. For periods between 10 −1 sand10 2 s,<br />

the influence of the conductive sea layer is dominant and therefore, the apparent resistivity decreases<br />

whereas the phase increases. For longer periods the resistive half-space dominates the curves (δ ≈ 16 km<br />

for 100 Ω m and 10 3 s). The sounding curves for the two sites on the sea-floor are displayed for a subset<br />

of the full period range (between 10 1 and 10 3 s) because of the strong attenuation within the very<br />

conductive sea water (Fig. 5). High frequencies generally require a much finer spatial discretization in<br />

regions where the skin depth is small. However, the computational experiments have been carried out<br />

on a mesh well suited for mid to low frequencies. Results for high frequencies were not satisfactory for<br />

the coarse meshes used here, and thus have not been included in the comparison.<br />

The long period ends of the sounding curves nicely approach the values of the lower half-space, i.e.<br />

100 Ωm for ρa and about 45° for φ. For short periods there are deviations from these values, which are<br />

larger for the points on the profiles oriented perpendicular to the component of the driving electric<br />

field, i.e. for point E40 at y =40kmforρxy and the point at x =40kmforρyx (cross markers in Fig. 5,<br />

cf. Fig. 3, right). As mentioned above, these deviations are due to the insufficient spatial discretization<br />

for short periods in regions of strong attenuation.<br />

ρ a [Ω m]<br />

φ [ o ]<br />

10 2<br />

10 1<br />

10 0<br />

100<br />

50<br />

10 −3<br />

0<br />

10 −2<br />

10 −1<br />

10 0<br />

T [s]<br />

10 1<br />

xy−polarisation<br />

yx−polarisarion<br />

10 2<br />

10 3<br />

ρ a [Ω m]<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

100<br />

50<br />

10 1<br />

0<br />

10 2<br />

T [s]<br />

xy−pol. (x)<br />

xy−pol. (y)<br />

yx−pol. (x)<br />

yx−pol. (y)<br />

Figure 5: Sounding curves for the volcano site T (left) and the sea-floor sites E40 and N40 (right) for the<br />

frustum model. Here, xy-pol. and yx-pol. refer to ρxy and ρyx, resp.<br />

The following table summarizes important computational parameters for the three different codes,<br />

such as polynomial degree of the basis functions (BFD) for the FE method and degrees of freedom<br />

(DOF). The overall CPU time required to obtain the result for one frequency was in the order of two<br />

minutes. The total time required to compute all field components for one polarization sums up to<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

64<br />

φ [ o ]<br />

10 3


about two hours for all 49 frequencies for the FE codes and 30 minutes for the FD code. Note that the<br />

3-D impedance requires the calculation of both polarizations (see eq. (7)).<br />

computer CPUs BFD elements DOFs time [min]<br />

COMSOL Multiphysics <br />

xy-pol. klio 4 quadratic 82 661 523 580 119,56<br />

FEP<br />

xy-pol. erato 8 quadratic 57 666 434 456 119,07<br />

FD code<br />

FD code 131 x 131 x 154 grid cells + 15 air layers ≈ 30<br />

4.2 Stromboli-Topography model<br />

In the second level of complexity, a realistic topography of Stromboli is incorporated. Since this<br />

topography is smoothly adjusted to the flat sea-floor, the edifice is larger in this model than in reality.<br />

Furthermore, the conductivity of the volcano has been reduced to 0.01 Sm −1 .<br />

Figs 6 and 7 illustrate the model with the positions of the profiles and the locations for which sounding<br />

curves have been calculated. Sounding curves are shown for sites a, N and E only.<br />

Figure 6: Flat sea-floor model including the real topography of Stromboli island. Red and blue lines indicate<br />

the x- and y-profiles, respectively.<br />

Figure 7: Locations for which sounding curves are computed are denoted by a, b, c, d on the island (left) and<br />

N,E,S,W on the sea-floor (right).<br />

Fig. 8 displays the apparent resistivities ρxy and phase Φxy for T =10 3 s along two profiles aligned with<br />

the x-axis (x-profile) and the y-axis (y-profile). The apparent resitivity for the x-profile is associated<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

65


with the xy-impedance tensor elements (Zxy mode) while the apparent resistivity for the y-profile<br />

is derived from the yx-elements (Zyx mode). The curves for both Zxy and Zyx modes show similar<br />

features. This can be explained by the similar topography along the two different profiles. Nevertheless,<br />

the extent of the volcano’s edifice in y-direction is larger than in x-direction. Hence, the corresponding<br />

minima and maxima of the curves are shifted relative to each other.<br />

Due to the smoother transition of the topography from the sea-floor to the volcano slopes the peaks at<br />

the base points of the edifice are not as pronounced as for the frustum model. Since there is no volcano<br />

plateau in the x-profile, there is just one minimum at x ≈ 44 km for ρxy.<br />

As for the frustum model, the sounding curves for the volcano (e.g. for site a in Fig. 9, left) reflect<br />

the layering of the earth with regard to the electrical conductivity whereas long periods correspond<br />

to large skin depths. Since the same conductivity is assigned to the volcano and the half-space the<br />

sounding curves reach values around 100 Ω m for short periods.<br />

Due to the asymmetry of the model, offsets occur between the curves derived from Zxy and Zyx that<br />

are higher for ρa than for φ. Hence, it can be assumed that the phase is less affected by these two<br />

aspects.<br />

The sounding curves for the sea-floor are displayed in Fig. 9 (right). Since the edifice is larger than the<br />

one represented by the frustum the long period apparent resitvities deviate more significantly from the<br />

undisturbed half-space values.<br />

z [km]<br />

ρ a [Ω m]<br />

φ [°]<br />

−5<br />

0<br />

5<br />

10 4<br />

10 2<br />

10 0<br />

40<br />

20<br />

xy−pol. x profile<br />

yx−pol. y profile<br />

0 10 20 30 40 50 60 70 80<br />

[km]<br />

Figure 8: Topography (top), apparent resistivity (center), phase (bottom), derived from Zxy and Zyx (xy- and<br />

yx-pol., resp.) along the x- andthey-profile for T =10 3 s.<br />

The differences in the short period range may again be attributed to the insufficient spatial discretization.<br />

ρ a [Ω m]<br />

φ [°]<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

100<br />

50<br />

10 −3<br />

0<br />

10 −2<br />

10 −1<br />

10 0<br />

T [s]<br />

10 1<br />

10 2<br />

xy−pol. a<br />

yx−pol. a<br />

10 3<br />

ρ a [Ω m]<br />

φ [°]<br />

10 3<br />

10 2<br />

10 1<br />

80<br />

60<br />

40<br />

10 1<br />

20<br />

10 2<br />

T [s]<br />

xy−pol. N<br />

yx−pol. N<br />

xy−pol. E<br />

yx−pol. E<br />

Figure 9: Sounding curves for the island site a (left) and for the sea-floor sites N and E (right). xy-pol. and<br />

yx-pol. refer to Zxy and Zyx, resp.<br />

The incorporation of the detailed Strombolian topography leads to a massively increasing number of<br />

mesh points. In order to keep computational costs low the meshes for computing the full sounding<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

66<br />

10 3


curves were coarsened.<br />

Computer CPUs BFD No. of elements DOFs Time [min]<br />

profile curves<br />

xy-pol. erato 8 quadratic 443 563 2 823 488 114.78<br />

sounding curves<br />

xy-pol. erato 8 quadratic 164 662 1 048 116 ≈ 721<br />

4.3 Bathymetry-Topography model<br />

In the third level of complexity, the topography of the island is incorporated as well as the regional<br />

bathymetry including the other Liparian islands. This allows to analyse the effect on the electromagnetic<br />

fields not only due to the volcano itself but also due to the variable thickness of the sea layer.<br />

Figs 10 and 11 illustrate the model and the locations of the profiles and points for which sounding<br />

curves have been calculated. As an example, apparent resistivities and phases for a period of 1000 s<br />

are shown for the xN- and xS-profiles as well as the full sounding curves for points a, c, N and S.<br />

Figure 10: Model including the regional bathymetry and the topography of Stromboli and the Liparian Islands.<br />

Figure 11: Left: Elevation map inferred from digital terrain data showing the locations of seafloor sounding<br />

points N, E, S, W and the profile lines xN, xS in the EW-direction and yE, yW in the NS-direction.<br />

Right: Coast line of Stromboli and locations of sounding points a, b, c, d.<br />

The xN and xS profiles running East-West in parallel (Fig. 12) are close to each other. Hence, the<br />

features of the respective apparent resistivity and phase curves are similar. As for the previous models,<br />

peaks at the coastline and at the top of the volcano can be observed. However, serious perturbations<br />

are provoked by the regional bathymetry. Induced currents are vertically compressed in the shallow<br />

parts of the ocean yielding higher apparent resistivities. Vice versa, apparent resistivites are decreased<br />

in deep water areas. The soundings at site c (Fig. 13, right) again show the influence of the volcano,<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

67


z [km]<br />

ρ a [Ω m]<br />

φ [°]<br />

−5<br />

0<br />

5<br />

10 4<br />

10 2<br />

10 0<br />

200<br />

100<br />

0<br />

xy−pol. xN profile<br />

xy−pol. xS profile<br />

0 20 40<br />

x [km]<br />

60 80<br />

Figure 12: Topography (top), apparent resistivity (center) and phase (bottom) for T = 10 3 satprofilesxS and<br />

xS. xy-pol. and yx-pol. refer to Zxy and Zyx, resp.<br />

ρ a [Ω m]<br />

φ [°]<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

400<br />

200<br />

10 −3<br />

0<br />

10 −2<br />

10 −1<br />

10 0<br />

T [s]<br />

10 1<br />

10 2<br />

xy−pol. a<br />

yx−pol. a<br />

10 3<br />

Figure 13: Sounding curves for the island sites a (left) and c (right). xy-pol. and yx-pol. refer to Zxy and Zyx,<br />

resp.<br />

the sea layer and the half-space for the appropriate period ranges. However, at site a, which is located<br />

directly at the coastline, the sea layer affects the curves already for shorter periods, especially for the<br />

Zyx mode. Still, for periods longer than T =10 −1 s the shapes of the curves are similar to those at<br />

point c (although shifted to higher ρa-values).<br />

All curves derived from Zxy and Zyx show an offset which can be attributed to the asymmetric elevation<br />

data used in the MT simulations.<br />

For the two sea-floor sites N and S the topography effect varies for the two cases considered (Fig. 14).<br />

Site N (Fig. 14, left) is located at greatest distance from the islands. Therefore, the curves are similar to<br />

those obtained using the simpler models and for longer periods they are almost assume the corresponding<br />

half-space values.<br />

The sounding curves at site S (Fig. 14, right) seem to be affected by two different facts. For ρxy the<br />

graben-like structure in this area (cf. Fig. 11, left) leads to lower apparent resistivities. On the other<br />

hand, the more resistive islands affect the ρyx-curves.<br />

We note that the out-of-quadrant phases in Fig. 14 occur mainly at the high frequencies, which indicates<br />

that the mesh used for the numerical computations is still not fine enough to obtain results with the<br />

desired accuracy.<br />

The following table summarizes the mesh properties and CPU times associated with the bathymetrytopography<br />

model.<br />

ρ a [Ω m]<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

100<br />

50<br />

10 −3<br />

0<br />

10 −2<br />

10 −1<br />

10 0<br />

T [s]<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

68<br />

φ [°]<br />

10 1<br />

10 2<br />

xy−pol. c<br />

yx−pol. c<br />

10 3


ρ a [Ω m]<br />

φ [°]<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

400<br />

200<br />

0<br />

10 1<br />

−200<br />

10 2<br />

T [s]<br />

xy−pol. N<br />

yx−pol. N<br />

10 3<br />

Figure 14: Sounding curves for the sea-floor sites N (left) and S (right). xy-pol. and yx-pol. refer to Zxy and<br />

Zyx, resp.<br />

ρ a [Ω m]<br />

φ [°]<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

400<br />

200<br />

10 1<br />

0<br />

10 2<br />

T [s]<br />

xy−pol. S<br />

yx−pol. S<br />

Computer CPUs BFD No. of elements DOFs Time [min]<br />

profile curves<br />

xy-pol. erato 16 quadratic 546 759 3 479 830 44.41<br />

sounding curves<br />

xy-pol. erato 16 quadratic 312 721 1 987 514 ≈ 1103<br />

5 Conclusions<br />

In this work, distortion effects on MT data caused by topography and bathymetry are examined for<br />

the area around Stromboli. The geometry of Stromboli has been incorporated into FE simulations<br />

applying three different levels of increasing complexity.<br />

In the first model, a frustum representing the volcano has been embedded in a layered background.<br />

This model yields a non-trivial apparent resistivity profile curve which is mainly dominated by peaks at<br />

the base points of the frustum, at the coastline points and at the edges of the volcano plateau. These<br />

are caused by static shift effects and the high conductivity contrast between the volcano and the sea<br />

layer. The results of the frustum model have been verified by a finite element code by Schwarzbach<br />

(2009) and a finite difference code by Mackie et al. (1994).<br />

For the second model of Stromboli digital elevation data of the island have been included in the model<br />

geometry. Compared to the frustum model, the apparent resistivity and phase curves along the profiles<br />

and the sounding curves for the volcano show similar features.<br />

Finally, digital bathymetry data have been added to the third, most complex Stromboli model to<br />

take into account realistic topography and bathymetry features of the region. The apparent resistivity<br />

still shows the characteristic peaks at the coastline points for most of the profiles. The effects of<br />

the bathymetry and topography of the surrounding islands are superposed, which results in very<br />

complicated profile and sounding curves.<br />

These results clearly point out that topography and bathymetry can heavily distort MT data and thus<br />

need to be considered for an accurate numerical interpretation of MT measurements.<br />

Finally, we emphasize that FE methods on unstructured grids provide a sophisticated tool for treating<br />

problems associated with complex geometry.<br />

Acknowledgements<br />

We would like to thank Randall L. Mackie for providing the results for the FD simulations. Furthermore,<br />

we are thankful to Christoph Schwarzbach for providing his FE code and for his useful help. We are<br />

grateful to the German Research Foundation DFG for funding our numerical research work (Spi 356-9).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

69<br />

10 3


References<br />

Börner, R.-U. (2010). Numerical Modelling in Geo-Electromagnetics: Advances and Challenges. Surveys in Geophysics<br />

31(2): 225–245.<br />

Mackie, R. L., J. T. Smith, and T. R. Madden (1994). Three-dimensional electromagnetic modeling using finite difference<br />

equations: The magnetotelluric example. Radio Science 29(4): 923–935.<br />

Nam, M. J., H. J. Kim, Y. Song, T. J. Lee, J.-S. Son, and J. H. Suh (2007). 3D magnetotelluric modelling including<br />

surface topography. Geophysical Prospecting (55): 277–287.<br />

Schwarzbach, C. (2009). Stability of Finite Element Solutions to Maxwell’s Equations in Frequency Domain. PhDthesis.<br />

TU Bergakademie Freiberg.<br />

SwissEduc (2010). SwissEduc: Stromboli Online - Die Insel. url: http://www.swisseduc.ch/stromboli/volcano/geogr/<br />

aerial-de.html.<br />

Wait, J. R. (1953). Propagation of radio waves over a stratified ground. Geophysics 18(2): 416–422.<br />

Wikipedia (2010). Liparische Inseln. url: http://de.wikipedia.org/wiki/Liparische_Inseln.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

70


Establishing Controlled Source MT at <strong>GFZ</strong><br />

M. Becken 1,2 , R. Streich 1,2 , O. Ritter 2<br />

1 Potsdam University, Department of Geosciences, Karl-Liebknecht-Strasse 24, 14476 Potsdam, Germany<br />

2 Helmholtz Centre Potsdam <strong>GFZ</strong> German Research Centre for Geosciences, Telegrafenberg, 14473 Potsdam, Germany<br />

Summary<br />

The multidisciplinary GeoEn project (the Brandenburg pilot project in the BMBF program<br />

„Spitzenforschung und Innovation in den Neuen Ländern“) integrates research in the fields of<br />

geothermal energy, carbon dioxide capture, transport and storage (CCTS) as well as the exploration of<br />

unconventional gas reserves (shale gas). The electrical conductivity is one key parameter to characterize<br />

reservoirs and to monitor changes due to circulation or injection of fluids at reservoir depth. Bulk<br />

electrical conductivity is highly sensitive to fluids within interconnected pores, and therefore EM<br />

techniques are powerful tools for exploring and monitoring geothermal reservoirs, CO2 storage sites and<br />

shale gas reservoirs. In the framework of the GeoEn project, we establish Controlled Source MT (CSMT)<br />

at <strong>GFZ</strong> Potsdam. We aim at combining active and passive MT to image the electrical conductivity<br />

structure within the Earth, ultimately in 3D.<br />

For CSMT, we intend to use grounded electrodes to inject a frequency-dependent current into the Earth<br />

and measure the induced electric and magnetic fields at near-field to far-field distances. We will use<br />

novel transmitter systems from Metronix (Braunschweig) which are presently under development.<br />

Standard MT receivers will be utilized to measure the induced electric and magnetic fields. In the scope<br />

of the project, we will (i) assemble the transmitter system and the source dipole, (ii) optimize and design<br />

CSMT field procedures for geothermal exploration, carbon dioxide reservoir characterization and shale<br />

gas exploration (Streich et al., 2010a), (iii) develop and implement time-series processing, (iv) develop<br />

and implement 1D modeling (Streich and Becken, 2009) and inversion software and 3D modeling codes<br />

(Streich, 2009; Streich et al., 2010b).<br />

In this contribution, we describe a test of long steel electrodes as the current electrodes of the source<br />

dipole, and we examine the resolution power of CSMT using 1D inversion of synthetic data.<br />

Current electrodes of CSMT source dipole<br />

Low grounding resistances of the current electrodes are crucial for injecting strong currents into the<br />

subsurface. The Metronix 22 kVA transmitter generates currents of max. 40 A (560 V), which can,<br />

however, only be achieved if the total resistance of the dipole (1-km long cable and grounding<br />

electrodes) is less than 14 . We use a thick cable that has a resistance of 2 /km. Accordingly, to<br />

achieve maximum currents, the electrode grounding resistances should be less than 12 . At high<br />

frequencies, the maximum current will be further limited by the inductance of the cable.<br />

The grounding resistance depends primarily on the surface of the electrode that is in contact with the<br />

ground, and the surrounding resistivity within the Earth. For a homogenous earth, the grounding<br />

resistance R of a steel rod is<br />

R<br />

4L<br />

ln<br />

2 L<br />

a<br />

1<br />

, (1)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

71


Figure 1: Electrode test. We measured the ground resistance of a steel electrode (diameter 32 mm) vs. depth of<br />

the rod at two different locations, which may represent typical conditions in northern Germany. Ground resistance<br />

measurements were taken against two reference electrodes at ~30 m distance to the left and right. DC Wenner<br />

spreads were measured at the same locations (left panels). The panels to the right show the measured resistances<br />

(red and blue lines), compared to theoretical predictions (dashed black lines) based on the resistivity-depth<br />

profiles at the electrode locations shown in the rightmost plots. The sudden drop in resistance at


greater depths, we penetrate conductive glacial till at test location 1 (upper panel in Figure 1) and a<br />

groundwater layer at test location 2 (lower panel). The measured grounding resistance values are<br />

roughly consistent with predictions (dashed lines) based on the geometry and the subsurface resistivity<br />

distribution. For resistance predictions, we approximated the resistance R from a parallel circuit of steel<br />

rod segments of length L embedded into a medium with piecewise constant resistivity i .<br />

Approximate i values were extracted from the 2D DC resistivity-depth sections. Let the individual rod<br />

pieces have resistance<br />

R<br />

i<br />

i 4 L<br />

ln<br />

2 L a<br />

then the resistance of the entire rod to a depth L is given by<br />

1<br />

R<br />

1<br />

i i R<br />

1<br />

, (3)<br />

. (4)<br />

Equations 3 and 4 were used to estimate grounding resistances from the 2D resistivity models (Figure 1).<br />

This test shows that long steel rods may be suitable electrodes for CSMT source dipoles. At one test<br />

location with favorable geology (glacial till), we achieved a ground resistance of ~10 with electrodes<br />

penetrating 7.5 m deep; the second test location, a sandy soil with a freshwater layer at ~3 m depth,<br />

may require deeper electrodes. Our predictions suggest that the ground resistance of the steel rod may<br />

drop at this location to values of ~10 at ~15 m depth. The simple model used to predict grounding<br />

resistance has proven to be useful for practical purposes. Before installing electrodes, it may be<br />

advisable to investigate potential locations with DC resistivity soundings in order to predict expectable<br />

resistances and required electrode lengths.<br />

1D CSMT Inversion<br />

Vertical currents, galvanically injected into the subsurface with a grounded dipole source, exhibit<br />

sensitivity to buried resistors at depth. This makes the CSMT technique suitable for imaging resistive<br />

layers, whereas both passive MT and CSMT are sensitive to conductive layers. The effect of anomalously<br />

resistive subsurface structures on surface CSMT data is typically smaller in land applications than in<br />

deep-water marine applications. Land applications of frequency-domain CSMT suffer from energy<br />

travelling through the air even more than marine applications, which obscures the response from<br />

deeper targets. Nevertheless, forward modeling studies suggest that in many cases, the e.m. response<br />

of thin resistive layers is above noise level (see Streich and Becken, this issue).<br />

A practical question is to what extent the resistivity models giving rise to these anomalies can be<br />

recovered from measured data. We ran 1D Occam-type inversions to investigate the resolution power of<br />

CSMT data. To infer the resistivity structure from 1D inversion, we can exploit the spatial decay of CSMT<br />

fields with increasing distance to the source, and the frequency dependency of the fields. In land-based<br />

applications, we will typically have only few source locations combined with a relatively large number of<br />

receivers (say, 100), and potentially long transmitting times that allow us to cover a broad frequency<br />

range. In contrast, marine applications use a source towed continuously over a number of receivers,<br />

effectively yielding many source point locations, but only short transmitting times and thus a narrow<br />

frequency band for a given source location. Hence, land applications will primarily utilize the frequency<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

73


Figure 2: 1D Occam inversion of synthetic CSMT data. The data are the real and imaginary parts of the inline<br />

electric field, contaminated with white noise with 1% of the signal amplitude. In addition, white noise with a<br />

standard deviation of 10 -9 V/m was added. a) Inversion of the spatial decay of the field for a single frequency (1<br />

Hz), using an unconstrained Occam inversion and using an Occam variant where the deviation from an a priori<br />

model is penalized outside of the depth of interest. b) Inversion of the frequency spectrum recorded at 5 km<br />

distance from the transmitter.<br />

Figure 3: Comparison of single and joint 1D inversion models for passive MT and CSMT data. In (a), the left panel<br />

shows passive MT data (apparent resistivity and phase) for the true model and MT inversion result displayed in the<br />

right panel. In (b), CSMT data (real and imaginary part of the inline electric field) are displayed for the true<br />

resistivity model and CSMT inversion result. (c) shows the resistivity model resulting from joint MT and CSMT<br />

inversion. In this example, the MT data add information only at greater depths, where the sensitivity of CSMT<br />

sounding is low. In practice, we anticipate that MT data will help construct a (2D/3D) regional background model,<br />

and CSMT data will help refine the model at reservoir scale (1D/3D).<br />

dependency of the fields, whereas marine applications will primarily exploit spatial variations of the<br />

fields.<br />

We have implemented a 1D CSMT inversion based on Weidelt’s formulation of the forward problem<br />

(Weidelt, 2007), and analytically derived expressions for the Jacobian matrix required in the inversion.<br />

Finite source dipoles (Streich and Becken, this issue) have been incorporated into the inversion;<br />

however, in the present form of the inversion scheme, the source and receivers are placed at the<br />

surface. Here, we only consider horizontal electric point dipole sources. The inversion is similar to<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

74


Occam inversion (Constable et al., 1987); however, Occam’s second phase (i.e., the determination of the<br />

regularization parameter that yields optimal trade-off between data residuals and model norm) has not<br />

been implemented. Instead, we fix the regularization parameter for one entire inversion run.<br />

Using this inversion scheme, we found that both the inversion of multi-offset data at a single frequency<br />

and multi-frequency data at a single receiver location have the power to yield comparable inversion<br />

models, if the frequency range and the distance to the transmitter are appropriate for imaging the<br />

target depth (Figure 2). The left panels in Figure 2a and b depict the inline electric field as a function of<br />

distance from the transmitter for the frequency f = 1 Hz and as a function of frequency at 5 km distance,<br />

respectively, for a model containing a resistive layer embedded into a homogeneous half-space.<br />

Random noise of 1% of the field amplitudes and a noise floor of 10 -9 V/m were added to the data prior to<br />

inversion. The central panel in Figure 2a shows the inversion model obtained from smooth inversion<br />

using a standard Laplacian penalty on the logarithmic model resistivities (red line). A similar result (not<br />

shown) was obtained from inversion of the multi-frequency electric field data shown in Figure 2b. In<br />

both cases, the recovered 1D model contains a resistive zone that is smeared over a wide depth range,<br />

because regularization penalizes thin model layers.<br />

For monitoring applications (e.g., CO2 injection), the depth of interest is known. A focused inversion that<br />

searches for deviations from an a priori model primarily at reservoir depth may help improve the model.<br />

We have therefore incorporated a term into the Occam inversion that allows us to weight the penalty<br />

depending on the difference of every inversion model parameter to an a priori model (Key, 2009). The<br />

right panels in Figure 2a and b show the result of this variant of regularized inversion. The hashed areas<br />

correspond to depth levels where the inversion model is constrained to deviate minimally from the a<br />

priori model (which is the true model in this case), whereas the rest of the model was permitted to vary<br />

freely (in the sense of a Laplacian regularization). This approach clearly helps focusing the resistive<br />

anomaly at the true depth and may thus be adequate for a focused model search.<br />

The penetration depth of CSMT data is limited by the lowest frequencies used and the greatest offsets<br />

providing reasonable signal quality. Combing CSMT with passive MT may prove helpful to expand the<br />

model scale. Furthermore, both techniques exhibit different sensitivities to resistive and conductive<br />

structures (and to deviations from 1D media) and thus provide complementary information. We have<br />

therefore implemented a joint 1D inversion for both data types. In Figure 3, we compare the 1D<br />

inversion models for a multi-layered structure obtained from MT data and CSMT data alone and from<br />

joint CSMT and MT inversion. Here, the CSMT data are the frequency-dependent inline electric field<br />

data, measured at 5 km offset from the transmitter. The individual inversion models show that (i) both<br />

techniques resolve shallow conductive layers, (ii) the CSMT data resolve a resistive layer at 1 km depth,<br />

and (iii) the MT data resolve the conductivity of the basal layer that is too deep to be sensed with CSMT.<br />

All of these model features are revealed by joint inversion, suggesting increased resolution capabilities<br />

from combined applications of both techniques.<br />

Acknowledgements<br />

This work is funded by the German Federal Ministry of Education and Research (BMBF) within the<br />

framework of the GeoEn project. S. Costabel, TU Berlin, made the DC measurements and provided the<br />

2D inversion models.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

75


References:<br />

Constable, S. C., Parker, R. L. and Constable, C. G., 1987, Occam's inversion: A practical algorithm for<br />

generating smooth models from electromagnetic sounding data, Geophysics 75(3), 289-300.<br />

Key, K., 2009, 1D inversion of multicomponent, multifrequency marine CSEM data: Methodology and<br />

synthetic studies for resolving thin resistive layers, Geophysics 74(2), F9-F20.<br />

Streich, R., 2009, 3D finite-difference frequency-domain modeling of controlled-source electromagnetic<br />

data: direct solution and optimization for high accuracy, Geophysics 74(5), F95-F105.<br />

Streich, R., Becken, M. and Ritter, O., 2010a, Imaging of CO2 storage sites, geothermal reservoirs, and<br />

gas shales using controlled-source magnetotellurics: modeling studies, Chemie der Erde, submitted.<br />

Streich, R. and Becken, M., 2010, EM fields generated by finite-length wire sources in 1D media:<br />

comparison with point dipole solutions, Protokoll zum Kolloquium “Elektromagnetische Tiefenforschung”,<br />

Seddiner See, 28.09.-02.10.2009.<br />

Streich, R., Schwarzbach, C., Becken, M. and Spitzer, K., 2010b, Controlled-source electromagnetic<br />

modelling studies: utility of auxiliary potentials for low-frequency stabilization, EAGE 70 th Conference<br />

and Exhibition, Barcelona, Spain, submitted.<br />

Weidelt, P., 2007, Guided waves in marine CSEM: Geophysical Journal International, 171, 153–176.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

76


2D-SIP-Modellierung mit anisotropen<br />

Widerständen<br />

Johannes Kenkel a∗ , Andreas Hördt a , Andreas Kemna b<br />

a Institut für Geophysik und extraterrestrische Physik, TU Braunschweig<br />

b Fachbereich Angewandte Geophysik, Universität Bonn<br />

1 Einleitung<br />

Die Modellierung von Messdaten der spektralen induzierten Polarisation (SIP) geschieht<br />

üblicherweise unter der Annahme makroskopisch, also in der Größenordnung einer Zelle<br />

der Modellrechnung, isotroper komplexer Widerstände im Untergrund. Für diese Art der<br />

Modellierung von Daten der SIP wurde von A. Kemna das Finite-Elemente-Programm<br />

CRMod entwickelt ([1], [2]). Mit der Annahme isotroper Widerstände lassen sich viele<br />

Messdaten hinreichend erklären. Nguyen et al. ([3]) zeigten jedoch für den reellwertigen<br />

gleichstromgeoelektrischen Fall, dass, wenn der Untergrund aus vielen abwechselnd gelagerten<br />

Schichten besteht, die eine geringe räumliche Ausdehnung gegenüber der bei der<br />

Inversion benutzen Gitterzellen haben, die Annahme isotroper Leitfähigkeiten zu Fehlinterpretationen<br />

führen kann. Hier kann die Erweiterung auf anisotrope Widerstände<br />

Abhilfe schaffen, weil diese auf makroskopischer Skala, also Zellenebene, Schichtlagerungen<br />

beschreiben können. Das Auftreten von Anisotropieeffekten auch bei SIP wurde<br />

von Winchen et al. ([4]) demonstriert. Das Programm CRMod wurde zum Zweck der<br />

Modellierung anisotroper komplexer Widerstände erweitert. In diesem Artikel sollen anhand<br />

von verschiedenen modellierten Beispielen verschiedene Effekte eines Untergrund<br />

mit anisotropen komplexen Widerständen gezeigt werden.<br />

2 Methodik<br />

2.1 Spektrale Induzierte Polarisation (SIP)<br />

Die spektrale induzierte Polarisation (SIP) ist ein elektrisches Verfahren, das wie die Geoelektrik<br />

aus dem Messen von Potentialverläufen künstlicher elektrischer Quellen Rückschlüsse<br />

auf die Leitfähigkeitsverteilung im Untergrund zieht.<br />

∗ Adresse: Mendelssohnstraße 3, D-38106 Braunschweig, E-Mail: j.kenkel@tu-bs.de<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

77


Das Aufnehmen der Messwerte geschieht in einem wiederum der Geoelektrik sehr<br />

ähnlichen Aufbau, jedoch werden zusätzlich Phasenbeziehungen zwischen Sende- und<br />

Empfangssignal aufgezeichnet. Als Quellsignal werden Wechselströme unterschiedlicher<br />

Frequenzen gewählt. Die Messdaten sind dann Spannungswerte und deren Phasenverschiebungen<br />

zum Quellsignal. Aus diesen Daten und den Positionen der Elektroden ergeben<br />

sich mit den Geometriefaktoren die Messwerte als scheinbare spezifische Widerstände<br />

(Amplituden) und deren Phasen. Diese lassen sich zur Abschätzung der Eigenschaften<br />

des Profils als Pseudosektion auftragen.<br />

Das physikalische Problem der spektralen induzierten Polarisation und der Geoelektrik<br />

lässt sich für den Fall anisotroper komplexer Widerstände in x-, y- und z-Richtung in<br />

Form der Poisson-Gleichung<br />

∂x(σx∂xφ)+∂y(σy∂yφ)+∂z(σz∂zφ)+Iδ(x − xs)δ(y − ys)δ(z − zs) =0 (1)<br />

mit dem Potential φ = φ(x, y, z), dem Strom I und der Diracschen Delta-Funktion δ<br />

schreiben. Die Anisotropie beschränke sich dabei auf die Diagonalelemente des Leitfähigkeitstensors<br />

⎛<br />

σx<br />

σ = ⎝ 0<br />

0<br />

σy<br />

⎞<br />

0<br />

0 ⎠ . (2)<br />

0 0 σz<br />

Die Stromeinspeisung findet an den Koordinaten (xs,ys,zs) statt.<br />

2.2 Modellierung<br />

Für die Modellierung wird das oben aufgeführte Problem der Potentialverteilung durch<br />

einen 2D-Finite-Elemente-Algorithmus gelöst (vgl. [2]). Die Ergebnisse sind Potentialverläufe<br />

und zugehörige Phasen an jedem Gitterpunkt. Aus diesen Informationen und<br />

den Messelektrodenpositionen lassen sich damit die zum Modell gehörigen synthetischen<br />

Messdaten - Amplituden und Phasen der scheinbaren spezifischen Widerstände - berechnen<br />

und in Pseudosektionen auftragen. Das hier benutzte Finite-Elemente-Programm<br />

CRMod ist von einer bereits bestehenden Version für die Modellierung isotroper spezifischer<br />

Widerstände und Phasen ([2]) um die Möglichkeit der Modellierung von Anisotropie<br />

erweitert worden.<br />

Die Poisson-Gleichung 1 wird unter der Annahme eines konstanten Leitfähigkeitstensors<br />

in y-Richtung durch die Fourier-Transformation in den Wellenzahlbereich mit<br />

∞<br />

F (k) = f(y)e<br />

−∞<br />

iky dy<br />

∞<br />

∞<br />

= f(y) cos(ky)dy + i f(y) sin(ky)dy<br />

−∞<br />

und der Wellenzahl k transformiert. Bei Reduktion auf Funktionen, die in y-Richtung<br />

zum Nullpunkt symmetrisch sind, fällt das zum Nullpunkt y = 0 antisymmetrische<br />

Teilintegral weg, so dass sich die Fourier-Kosinus-Transformation<br />

∞<br />

F (k) = f(y) cos(ky)dy (3)<br />

−∞<br />

−∞<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

78


ergibt. Das Ergebnis der Transformation der Gleichung 1 in den Wellenzahlbereich entlang<br />

der y-Achse ist dann<br />

∂x(σx∂x ˜ φ) − σyk 2 ˜ φ + ∂z(σz ˜ φ)+ I<br />

2 δ(x − x0)δ(z − z0) =0 (4)<br />

mit ˜ φ = ˜ φ(x, k, z). Diese Gleichung wird von dem hier verwendeten Finite-Elemente-<br />

Programm auf einem zweidimensionalen Gitter in x- und z-Richtung gelöst. Aus den<br />

erhaltenen Werten für ˜ φ, bzw. nach Fourier-Kosinus-Rücktransformation φ lassen sich<br />

mit den Geometriefaktoren der Elektrodenkonfiguration die synthetischen Messwerte für<br />

Amplitude und Phase der scheinbaren spezifischen Widerstände angeben.<br />

2.3 Bewertung der Genauigkeit<br />

Das Programm muss auf seine Gültigkeit und Genauigkeit in Bezug sowohl auf analytisch<br />

berechenbare Probleme als auch auf verschiedene, mit dem bisherigen Programm zum<br />

Modellieren isotroper spezifischer Widerstandsverteilungen berechnete Modelle geprüft<br />

werden. Für einen homogenen Halbraum mit spezifischem Widerstand ρh und Phase<br />

φh erwartet man eine Pseudosektion mit dem konstanten Wert ρ = ρh des scheinbaren<br />

spezifischen Widerstands und φ = φh für die scheinbare Phase. Die Messwerte werden<br />

mit einer simulierten Dipol-Dipol-Anordnung mit Elektrodenabstand 1 m berechnet. Die<br />

Abweichungen der Modellierungsergebnisse zum analytischen Ergebnis betragen weniger<br />

als 10 % in den scheinbaren spezifischen Widerständen und weniger als 1 % in den scheinbaren<br />

Phasen. Ein homogener Halbraum mit anisotropen spezifischen Widerständen in<br />

Horizontal- und Vertikalrichtung, ρhorizontal und ρvertikal, weist nach analytischen Berechnungen<br />

einen scheinbaren spezifischen Widerstand von<br />

ρ = √ ρhori. · ρvert.<br />

(nach z.B. [5, S. 95]) auf. In Abb. 1 ist ein Modellierungsergebnis für diesen Fall dargestellt.<br />

Der spezifische Widerstand des Halbraums beträgt 100 Ωm in horizontaler Richtung<br />

(x- und y-Richtung) und 25 Ωm in z-Richtung. Dies entspricht einem scheinbaren<br />

spezifischen Widerstand von 50 Ωm in der Pseudosektion. Die Phase des Halbraums<br />

beträgt gleichmäßig (homogen und isotrop) −5 mrad. Die Abweichung des Modellierungsergebnisses<br />

zu diesem analytischen Wert beträgt an keiner Stelle mehr als 3 %<br />

sowohl für den scheinbaren spezifischen Widerstand als auch für die scheinbare Phase.<br />

3 Modellparameter<br />

Unter der Annahme einer horizontalen Wechsellagerung unterschiedlich leitfähiger Schichten<br />

wurde ein Modell mit isotropen spezifischen Widerständen erstellt und mit einem<br />

entsprechenden Modell mit anisotropen spezifischen Widerständen verglichen. Zur<br />

Abschätzung der Stärke der Anomalien wird ein Hintergrundmodell mit homogenem und<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

79<br />

(5)


(a) (b)<br />

Abbildung 1: Dipol-Dipol-Pseudosektion von spezifischem Widerstand und Phase eines<br />

homogenen Halbraums mit den anisotropen spezifischen Widerständen<br />

(Amplituden) 100 Ωm in x- und y-Richtung und 25 Ωm in z-Richtung und<br />

der isotropen Phase −5 mrad.<br />

isotropem spezifischen Widerstand angegeben. Abb. 2 zeigt die Diskretisierung des Modellraumes<br />

in 233 Zellen in x-Richtung und 114 Zellen in z-Richtung mit variablem Gitterlinienabstand<br />

in den Außenbereichen und konstantem Gitterlinienabstand von 0, 5m<br />

m in einem Bereich von 0 m bis 100 m in x-Richtung.<br />

Abbildung 2: Gittermodell des Profils mit 233x114 Gitterzellen und einem Gitterlinienabstand<br />

von 0, 5 m im linearen Bereich.<br />

Für das Hintergrundmodell wurde wie im vorherigen Abschnitt ein homogener und<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

80


isotroper spezifischer Widerstand mit Amplitude 100 Ωm und Phase −5 mrad gewählt.<br />

3.1 Wechsellagerungen mit unterschiedlichen spezifischen Widerständen<br />

Im ersten Beispiel wird eine Wechsellagerung von unterschiedlich resistiven, isotropen<br />

Schichten (100 Ωm und 1 Ωm) in einem Hintergrund mit spezifischem Widerstand von<br />

100 Ωm betrachtet. Die Phase des spezifischen Widerstands beträgt überall −5 mrad.<br />

Eine Skizze der Anordnung findet sich in Abb. 3(a). Die Schichtung beginnt 1 m unter<br />

(a) (b)<br />

Abbildung 3: Schema des Untergrundmodells: a) vertikale Schichtung mit den spezifischen<br />

Widerständen 100 Ωm in hellgrau und 1 Ωm in dunkelgrau. b) Vertikale<br />

Schichtung mit spezifischem Widerstand 1, 98 Ωm in x- und y- sowie<br />

50, 5 Ωm in z-Richtung. Der spezifische Hintergrundwiderstand beträgt<br />

100Ωm in x-, y- und z-Richtung. Die Phase beträgt überall −5 mrad.<br />

der Oberfläche und reicht bis zu einer Tiefe von 7, 5 m. Die Ausdehnung in x-Richtung<br />

beträgt 40 m.<br />

Die gleiche Schichtung wird in Abb. 3(b) durch einen Block mit anisotropem spezifischen<br />

Widerstand beschrieben. Die anisotropen spezifischen Widerstandswerte berechnen<br />

sich nach<br />

ρ = R A<br />

l<br />

⇔ R = ρ l<br />

(6)<br />

A<br />

mit dem Widerstand R eines Blocks der Länge l und der Querschnittsfläche A. Bei<br />

Reihenschaltung von zwei Schichten mit den Widerständen R1 und R2 mit je der Querschnittsfläche<br />

A und der Länge l ergibt sich aus dem Gesamtwiderstand R, der die<br />

Querschnittsfläche A und die Länge 2 · l hat:<br />

RSerie = R1 + R2<br />

2 · l l l<br />

⇔ ρSerie = ρ1 + ρ2<br />

A A A<br />

⇔ ρSerie = ρ1 + ρ2<br />

.<br />

2<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

81<br />

(7)


Bei Parallelschaltung von zwei Schichten gleicher Querschnittsfläche A und gleicher<br />

Länge l und den Widerständen R1 und R2 ergibt sich der Gesamtwiderstand R, der<br />

die Querschnittsfläche 2 · A und die Länge l hat:<br />

⇔<br />

2 · A<br />

l<br />

1<br />

RSerie<br />

1<br />

ρSerie<br />

⇔ ρP arallel =2·<br />

= 1<br />

+<br />

R1<br />

1<br />

R2<br />

= A<br />

l ρ1 + A<br />

l ρ2<br />

1<br />

1 1 + ρ1 ρ2<br />

Mit diesen Beziehungen folgt für die äquivalenten anisotropen spezifischen Widerstände<br />

50, 5Ωminx-und1, 98 Ωm in z-Richtung. In y-Richtung beträgt der spezifische Widerstand<br />

wie in x-Richtung 1, 98 Ωm. Die Phasen der spezifischen Widerstände betragen<br />

homogen und isotrop −5 mrad. Die Modellierungsergebnisse sind als Pseudosektionen<br />

einer Dipol-Dipol-Anordnung in den Abb. 4(a) bis 5(b) jeweils für Betrag und Phase des<br />

spezifischen Widerstands dargestellt.<br />

(a) (b)<br />

Abbildung 4: Pseudosektion des Betrags des scheinbaren spezifischen Widerstands. a)<br />

Vertikale Wechsellagerung isotroper Schichten mit 1 Ωm und 100 Ωm. b)<br />

Anisotrope Beschreibung der Wechsellagerung durch spez. Widerstände<br />

50, 5 Ωm in x- und 1, 98 Ωm in y- und z-Richtung. Die Phase beträgt in<br />

beiden Modellen überall −5 mrad.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

82<br />

.<br />

(8)


(a) (b)<br />

Abbildung 5: Pseudosektion der scheinbaren Phase. a) Vertikale Wechsellagerung isotroper<br />

Schichten mit 1 Ωm und 100 Ωm. b) Anisotrope Beschreibung der<br />

Wechsellagerung durch spez. Widerstände 50, 5Ωminx-und1, 98 Ωm in<br />

y- und z-Richtung. Die Phase beträgt in beiden Modellen überall −5 mrad.<br />

Beide Modelle - isotropes mit Schichten und anisotropes mit Block - zeigen nahezu<br />

gleiche Ergebnisse in den Pseudosektionen. Sehr deutlich ist die durch die Schichtlagerung<br />

bzw. durch den Block anisotroper spezifischer Widerstände erzeugte Anomalie<br />

sichtbar. Der maximale Wert für den scheinbaren spezifischen Widerstand ρa ist etwa<br />

100 Ωm, was dem Hintergrundmodell entspricht. Der minimale Wert für ρa beträgt etwa<br />

10 0,5 Ωm = 3 Ωm. Der erwartete minimale scheinbare spezifische Widerstand beträgt im<br />

Fall des anisotropen Modells gemäß Gleichung 5 etwa 10 Ωm. Dieser Wert wird im Bereich<br />

des Blocks und darunter erreicht und sogar unterschritten. Der bis in 7, 5 m Tiefe<br />

ausgedehnte Block führt einen ”Schatten” mit, der in Form eines nach unten gerichteten<br />

Dreiecks bis in ca. 20 m Tiefe reicht. An beiden Flanken des Dreiecks zeigen schräge<br />

”Schatten” nach außen, die ebenfalls einen relativ niedrigen spezifischen Widerstand von<br />

etwa 10 Ωm aufweisen. Die scheinbaren Phasen betragen gleichmäßig −5 mrad, entsprechend<br />

den Vorgaben der beiden Modelle.<br />

3.2 Wechsellagerungen mit unterschiedlichen Phasen des spezifischen<br />

Widerstandes<br />

In diesem Abschnitt soll die Schichtlagerung mit Schichten unterschiedlicher Phasen betrachtet<br />

werden. Dazu wird das Schichtmodell aus Abb. 3(a) für Schichten mit isotropen<br />

Phasen und das Blockmodell aus Abb. 3(b) als entsprechendes Modell mit anisotropen<br />

Phasen betrachtet. Die spezifischen Widerstände ρ sind in allen Schichten gleich<br />

und entsprechend auch im anisotropen Fall in allen Richtungen gleich. Die Reihenschaltung<br />

von Elementen mit unterschiedlichen Phasen - also komplexen Impedanzen<br />

X = R(cos φ1 + i sin φ1) - ergibt sich mit Realteil ℜ(X) und Imaginärteil ℑ(X) zu<br />

XSerie = X1 + X2<br />

= ℜ(X1 + X2)+iℑ(X1 + X2)<br />

=(Rcos φ1 + R cos φ2)+i(R sin φ1 + R sin φ2).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

83<br />

(9)


Für kleine Winkel von φ1 und φ2 kann man die Kosinus-Terme über eine Taylor-Entwicklung<br />

in erster Ordnung um 0 mit<br />

cos φ ≈ 1 (10)<br />

und die Sinus-Terme mit<br />

annähern. Es ergibt sich dann für die Reihenschaltung<br />

sin φ ≈ φ (11)<br />

XSerie ≈ 2 · R + iR(φ1 + φ2) (12)<br />

Diese Beziehung lässt sich zur Berechnung des komplexen spezifischen Widerstandes<br />

heranziehen. Zwei Elemente mit den Impedanzen X1 und X2 und je der Querschnittsfläche<br />

A und der Länge l ergeben in Reihenschaltung eine Impedanz X mit der gleichen<br />

Querschnittsfläche und der doppelten Länge 2 · l:<br />

ρ = X A<br />

l<br />

A<br />

= XSerie<br />

2l<br />

≈ (R + iR φ1 + φ2<br />

)<br />

2<br />

A<br />

l<br />

≈ R(cos φ1 + φ2<br />

+ i sin<br />

2<br />

φ1 + φ2<br />

)<br />

2<br />

A<br />

l .<br />

Die Reihenschaltung zweier gleich mächtiger Schichten mit gleichem spezifischen Widerstand<br />

und unterschiedlicher Phase bedeutet folglich näherungsweise eine Mittelung der<br />

Phasen, also<br />

φSerie ≈ φ1 + φ2<br />

. (14)<br />

2<br />

Die Parallelschaltung unterschiedlicher Phasen bei gleichem Betrag des spezifischen Widerstands<br />

ergibt sich mit den obigen Näherungen und φ ≪ 1zu<br />

XP arallel =<br />

≈ R<br />

1<br />

1 1 + X1 X2<br />

1<br />

1+iφ1<br />

1<br />

+ 1<br />

1+iφ2<br />

≈ R( 1 i<br />

+<br />

2 2 (φ1 + φ2<br />

)).<br />

2<br />

Werden zwei Elemente mit gleichem spezifischen Widerstand und unterschiedlicher Phase<br />

und je der Querschnittsfläche A und der Länge l parallel geschaltet, ergibt sich die<br />

Impedanz XP arallel mit der doppelten Querschnittsfläche 2 · A und der gleichen Länge.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

84<br />

(13)<br />

(15)


Der komplexe spezifische Widerstand ergibt sich damit zu<br />

ρ = X A<br />

l<br />

= XP arallel<br />

≈ R( 1 i<br />

+<br />

2<br />

2A<br />

l<br />

2 (φ1 + φ2<br />

2<br />

)) 2A<br />

l<br />

= R(1 + i( φ1 + φ2<br />

))<br />

2<br />

A<br />

l<br />

≈ R(cos φ1 + φ2<br />

2<br />

+ i sin φ1 + φ2<br />

)<br />

2<br />

A<br />

l ,<br />

und für die Phasen wie schon bei der Reihenschaltung näherungsweise:<br />

(16)<br />

φP arallel ≈ φ1 + φ2<br />

. (17)<br />

2<br />

Dieses Ergebnis macht klar, dass bei konstantem spezifischen Widerstand eine Wechsellagerung<br />

von Schichten unterschiedlicher Phasen wohl in ihrer Stärke im Mittel, nicht<br />

jedoch in ihrer Richtung erkannt werden kann. Nur in Verbindung mit unterschiedlich<br />

spezifischen Widerständen der einzelnen Schichten können auch die Phasenwerte der<br />

abwechselnden Schichten unterschieden werden.<br />

Ein Schichtmodell mit den Phasen der einzelnen Schichten von −15 mrad bzw. −5 mrad<br />

(Hintergrund) hat demnach die anisotropen Phasen −10 Ωm in x-,y- und z-Richtung.<br />

Die Ergebnisse sind in den Abb. 6(a) bis 7(b) als scheinbare spezifische Widerstände und<br />

scheinbare Phasen in Pseudosektionen dargestellt.<br />

(a) (b)<br />

Abbildung 6: Dipol-Dipol-Pseudosektion der Amplitude des scheinbaren spezifischen<br />

Widerstands. a) Vertikale Wechsellagerung isotroper Schichten mit<br />

−15 mrad und −5 mrad. b) Anisotrope Beschreibung der Wechsellagerung<br />

durch Phasen −10 mrad in x-, y- und z-Richtung<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

85


(a) (b)<br />

Abbildung 7: Dipol-Dipol-Pseudosektion der scheinbaren Phase. a) Vertikale Wechsellagerung<br />

isotroper Schichten mit −15 mrad und −5 mrad. b) Anisotrope<br />

Beschreibung der Wechsellagerung durch Phasen −10 mrad in x-, y- und<br />

z-Richtung<br />

Wie erwartet zeigt sich in den Pseudosektionen des scheinbaren spezifischen Widerstandes<br />

keine Variation mit der eingestellten Phase. Dadurch wird der in Gl. 14 und 17<br />

aufgestellte Zusammenhang bekräftigt. Die Pseudosektionen der Phasen weisen wie die<br />

Pseudosektionen im Abschnitt 3.1 starke ”Schatten” unterhalb der unteren Störkörpergrenze<br />

auf. Auch die seitlichen Flanken sind sichtbar.<br />

4 Schlussfolgerungen<br />

Es ist mit dem auf anisotrope spezifische Widerstände und Phasen erweiterten Programm<br />

CRMod nun möglich, Effekte von Anisotropie, z.B. bei Wechsellagerungen von Sedimenten,<br />

zu berücksichtigen. Die vorliegenden Ergebnisse zeigen in den möglichen Vergleichen<br />

zu analytischen Berechnungen die Gültigkeit des Programms. Es wurde hierbei insbesondere<br />

Wert auf die Vorschrift zur Berechnung des spezifischen Widerstands im Fall eines<br />

homogenen Halbraums mit anisotropen spezifischen Widerständen (vgl. Gleichung 5)<br />

gelegt. Im Zusammenhang mit der Wechsellagerung von Phasen wurde mittels der Gleichungen<br />

14 und 17 gezeigt, dass horizontale und vertikale Wechsellagerungen gleicher<br />

Mächtigkeit und gleichen spezifischen Widerstands nicht unterschieden werden können.<br />

Im Zusammenhang mit dieser Arbeit ist eine Messkampagne geplant, die auf stark<br />

anisotropem Untergrund stattfinden soll. Die Messdaten sollen mit entsprechenden Modellen<br />

mit anisotropen spezifischen Widerständen des vorliegenden Modellierungsprogramms<br />

erklärt werden. Zusätzlich zu dieser Anwendung besteht die Möglichkeit, auch<br />

das auf dem Modellierungsprogramm CRMod basierende Inversionsprogramm CRTomo<br />

([2]) auf Anisotropie zu erweitern.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

86


5 Danksagungen<br />

Diese Arbeiten werden im Rahmen des Forschungsverbunds Geothermie und Hochleistungsbohrtechnik<br />

(GEBO) vom Niedersächsischen Ministerium für Wissenschaft und<br />

Kultur und von Baker Hughes/ Celle gefördert<br />

Literatur<br />

[1] A. Kemna, ”Tomographic Inversion of Complex Resistivity - Theroy and Application”,<br />

Berichte des Instituts für Geophysik der Ruhr-Universität Bochum, Reihe A, Nr.<br />

56, Der Andere Verlag, 2000, ISBN-13: 978-3934366923.<br />

[2] A. Kemna, ”Tomographische Inversion des spezifischen Widerstandes in der Geoelektrik”,<br />

Diplomarbeit, Institut für Geophysik und Meteorologie der Universität zu<br />

Köln, 1995.<br />

[3] F. Nguyen, S. Garambois, D. Chardon, D. Hermitte, O. Bellier, D. Jongmans, ”Subsurface<br />

electrical imaging of anisotropic formations affected by a slow active reverse<br />

fault, Provence, France”, Journal of Applied Geophysics 62, Seiten 338 - 353, 2007.<br />

[4] T. Winchen, A. Kemna, H. Vereecken und J.A. Huisman, ”Characterization of bimodal<br />

facies distributions using effective anisotropic complex resistvity: A 2D numerical<br />

study based on Cole-Cole models”, Geophysics 74, Seiten A19 - A22, 2009.<br />

[5] K. Knödel, H. Krummel, G. Lange, ”Geophysik”, 2. überarbeitete Auflage, Springer-<br />

Verlag, Berlin, 2005, ISBN 3-540-22275-8.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

87


From forward modelling of MT phases over 90 ◦ towards<br />

2D anisotropic inversion<br />

X. Chen 1,2 , U. Weckmann 1 , K. Tietze 1,3<br />

1 Helmholtz Centre Potsdam - German Research Center for Geosciences,Germany<br />

2 University of Potsdam, Institute of Geosciences, Germany<br />

3Free University of Berlin, Institute of Geological Sciences, Germany<br />

xiaoming@gfz-potsdam.de, uweck@gfz-potsdam.de, ktietze@gfz-potsdam.de<br />

Abstract<br />

Within the framework of the German - South African geo-scientific research initiative Inkaba<br />

yeAfrica several magnetotelluric (MT) field experiments were conducted along the Agulhas-<br />

Karoo Transect in South Africa. This transect crosses several continental collision zones between<br />

the Cape Fold Belt, the Namaqua Natal Mobile Belt and Kaapvaal Craton. Along the profile<br />

we can identify areas (> 10km) with phases over 90 ◦ . This phenomenon usually occurs in presence<br />

of electrical anisotropy. Due to the dense site spacing we are able to observe this behaviour<br />

consistently at several sites.<br />

In this presentation we focus on the profile section between Prince Albert and Mosselbay. With<br />

isotropic 2D inversion we are able to explain most features in the MT data but not the abnormal<br />

phase behavior. With several anisotropic forward modelling studies we have tested the influence<br />

of anisotropy parameters on the MT responses. In a first step we use simple 2D models with<br />

embedded zones of electrical anisotropy to get a basic understanding of anisotropic responses. In<br />

a second step isotropic 2D inversion results serve as background models in which we included<br />

anisotropic zones, e.g. to fit the abnormal phase curves. These resolution tests are necessary and<br />

important for the future development of a 2D inversion with spatially constraint anisotropy.<br />

1 Introduction<br />

For a 2D geoelectric model with a finite system of homogeneous, but generaly anisotropic blocks the<br />

electrical conductivity is a tensor instead of a scalar quantity. It is symmetric and positive-definite.<br />

Due to its symmetry the conductivity tensor σ within each layer of model can be diagonalized and expressed<br />

by three principal conductivities σ1, σ2, σ3 and a rotation matrix, which can be decomposed<br />

into three elementary Euler’s rotations αS, αD, αL respectively.<br />

⎛<br />

⎞<br />

σ =<br />

⎝<br />

σxx σxy σxz<br />

σyx σyy σyz<br />

σzx σzy σzz<br />

⎠<br />

⎛<br />

= Rz(−αS)Rx(−αD)Rz(−αL) ⎝<br />

σ1 0 0<br />

0 σ2 0<br />

0 0 σ3<br />

⎞<br />

⎠Rz(αL)Rx(αD)Rz(αS)<br />

where Rx and Rz are elementary rotation matrices around the coordinate axis x and z, respectively.The<br />

angles αS, αD, αL are typically called anisotropy strike, dip and slant, respectively.<br />

In this work we investigate the feasibility of use of 2D forward anisotropy modelling to simulate<br />

phases over 90 ◦ which we observed in field data from the MT survey in South Africa. Using the 2D<br />

anisotropy forward modelling algorithm of Pek and Verner (1997) we vary the model parameters in<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

88


order to gain a basic understanding of the MT transfer function in presence of anisotropy. First we<br />

will give a brief introduction of the survey area and then we will show the results of 2D isotropic<br />

inversion and 2D anisotropic forward modelling.<br />

2 Survey area<br />

Within the framework of the German - South African geo-scientific research initiative Inkaba yeAfrica<br />

four magnetotelluric (MT) field experiments were conducted along the Agulhas-Karoo Transect in<br />

South Africa (Weckmann et al., 2007, Stankiewicz et al., 2008). These transects cross several continental<br />

collision zones and their respective units, such as the Cape Fold Belt, the Namaqua Natal<br />

Mobile Belt and the Kaapvaal Craton. The MT profile on which we focus in this paper is located in<br />

CFB<br />

NNMB<br />

MT2 profile<br />

SCCB<br />

BMA<br />

MT<br />

NVR<br />

WRR<br />

Offshore seismics<br />

Kaapvaal<br />

Craton<br />

Karoo<br />

Agulhas-Falkland Fracture Zone<br />

Agulhas<br />

Plateau<br />

Transkei<br />

Basin<br />

Figure 1: Left: Map of the Agulhas-Karoo Geoscience Transect with MT Profiles (red lines) across<br />

Cape Fold Belt , the Namaqua Natal Mobile Belt and the Kaapvaal Craton. Right: The profile MT2<br />

contains 46 broad band MT sites and 8 broad band / long period MT sites. It extends 120 km from<br />

Prince Albert in the North to Mosselbay in the South and covers the entire Cape Fold Belt.<br />

the Cape Fold Belt. In total, 54 MT sites were deployed along this 120 km long profile MT 2 (Fig. 1).<br />

It extends from Prince Albert in the North to Mosselbay in the South and covers the entire Cape Fold<br />

Belt (CFB), its inliers, the Oudtshoorn and the Kaaimans Basins, the Swartberg and the Outeniequa<br />

Mountain ranges and several major thrusts and faults.<br />

3 Data and isotropic 2D inversion<br />

Along the profile MT2 we acquired 5-component MT data at all stations in a period range from<br />

0.001s to 1000s using GPS synchronized S.P.A.M. MkIII (Ritter et al., 1998) and CASTLE broadband<br />

instrument. Metronix MFS06/06 induction coil magnetometer and non-polarizable Ag/AgCl telluric<br />

electrodes were used to record natural magnetic and electric field variations. The data were processed<br />

with the EMERALD software package (Ritter et al., 1998) using both robust single site and remote<br />

reference techniques. At some sites, for which a suitable reference site was not available and the<br />

data were affected by cultural noise (extensive farming), we applied the frequency domain selection<br />

scheme after Weckmann et al. (2005) to improve data quality.<br />

Mozambique Ridge<br />

SCCB<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

89


Figure 2: Pseudo-sections of TE and TM mode apparent resistivity and phase (x direction pointing<br />

west).We can observe phase values over 90 degrees in the middle of the profile in both TE and TM<br />

component (marked with black ellipses in the upper panel). Two exemplary sites (site 121 and site<br />

116) are displayed as apparent resistivity, phase and induction arrows (Wiese convention) over period.<br />

The red arrows in the pseudo-sections show the location of the selected sites. We can identify that the<br />

phases, (especially in the TE component) leave the first quadrant at a minimum period of 10 s, but<br />

typically at longer periods of 100-1000 s.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

90


After data processing, we applied different strike and dimensionality analyses, e.g. the ellipticity<br />

analysis after Becken & Burkhardt (2004), obtaining an electromagnetic strike direction of −90 ◦ , i.e.<br />

East-West direction. This is in general compatible with the tectonic grain of the CFB. Furthermore,<br />

most of the data seem to be compatible with a 2D interpretation approach. However, in a limited area<br />

we observe impedance phases exceeding 90 degrees at long periods. They typically appear in both,<br />

TE and TM mode, and persit in either component even if we rotate the coordinate system (Fig. 2).<br />

Figure 3: TE and TM mode data used for isotropic 2D inversion. Prior to inversion they were rotated<br />

into a common coordinate system according to the geoelectric strike direction of ≈ 90 ◦ (obtained by<br />

strike and phase tensor analysis). Data with phases over 90 ◦ are discarded from the data base.<br />

For an isotropic 2D inversion approach we use RLM2D (Rodi & Mackie, 2001, implemented in<br />

the WinGlink software package). Before starting a 2D inversion we have to make sure that large phase<br />

values over 90 ◦ are excluded as they cannot be fitted by a 2D inversion approach. Figure 3 shows<br />

the data which were finally used for 2D inversion. Figure 4 displays our preferred 2D conductivity<br />

image of the upper 35 km together with a section of the geological map after Hälbich et al. (1993).<br />

The inversion was started from a homogeneous half-space of 100Ωm with τ = 10 using TE and<br />

TM component, with intermediate inclusion of the vertical magnetic transfer function. Within the<br />

framework of this work, we refrain from interpreting conductivity anomalies in a geological context,<br />

but focus on the area where phases over 90 ◦ occur. One of the most prominent conductivity anomalies<br />

is a triangular shaped, highly conductive structure in the middle of the profile. In this area we also<br />

observe phases leaving the quadrant. This conductivity anomaly is required by the data which is<br />

compatible with a 2D interpretation approach; however, we believe that the phases over 90 ◦ are caused<br />

by some electrically anisotropic structures in this area (in Fig. 4 between black dashed lines). We<br />

should also note that the deep part of this area only possesses a very limited resolution (in Fig. 4<br />

marked with semitransparent mask) because most of the data which relates to this part are excluded<br />

in order to satisfy the 2D isotropic procedure (see Fig. 3). Similar phase behaviour was explained<br />

with crustal anisotropy by Weckmann et al. (2003) and Heise & Pous (2003), where data within old<br />

continental collision zones in Namibia and on the Iberian Peninsula, respectively. In both cases the<br />

data could be fit by using a shallow and a deeper electrically anisotropic zone with different anisotropy<br />

strike.<br />

4 Anisotropic forward modelling<br />

Isotropic 2D inversion is adequate to explain the data in most parts along the profile but it is very<br />

unsatisfactory not being able to include phases greater than 90 degrees and thus neglect a substantial<br />

amount of data. In order to develop a 2D inversion with spatially constraint anisotropy, resolution<br />

tests and synthetic modelling studies are necessary. In a first step towards the constraint anisotropic<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

91


Figure 4: 2D inversion model together with a section from the geological map after Hälbich et al.<br />

(1993). The area where we observe phases over 90 ◦ is located where the inversion puts a triangular<br />

shaped high conductivity anomaly (area between black dashed lines). Zones of limited resolution are<br />

also in this area and marked with a semitransparent mask.<br />

inversion we included zones with electrical anisotropy in a 2D isotropic model and calculated its<br />

forward responses which aim to understand the anisotropy effect.<br />

4.1 Lateral extension<br />

Pek & Verner (1997) have suggested that a combination of two azimuthal anisotropies with anisotropy<br />

strikes perpendicular to each other could produce phases that exceed 90 ◦ . Based on the suggestion<br />

we study at first a combination of two different anisotropic blocks. The initial model (Fig. 5, left)<br />

consists of a 300m isotropic surface layer with a resistivity of 30Ωm and an anisotropic block starting<br />

at a depth of 300m embedded in a medium of 100Ωm. The principal resistivities of the block<br />

are σ1/σ2/σ3 = 50/0.5/50Ωm and the anisotropy strike αS is 120 ◦ . The block is underlain by an<br />

isotropic layer with a resistivity of 15Ωm. Beneath the isotropic layer a second anisotropic block<br />

with σ1/σ2/σ3 = 30/0.3/30Ωm and αS = 30 ◦ (perpendicular to αS of the first block) is embedded in<br />

an isotropic half-space with 100Ωm. The second, deeper anisotropic block has a lateral extension of<br />

15km in the first (Fig. 5, left upper panel), and 80km in the second (Fig. 5, left lower panel) model,<br />

respectively.<br />

The forward responses are displayed in figure 5 (right) as apparent resistivities and phases in xy<br />

and yx component, respectively. Comparing the responses of both models we see that the phases of<br />

yx component (Φyx) for the second model (Fig. 5, left lower panel) at sites above the first anisotropy<br />

block leave the quadrant at a period of ≈ 100s (Fig. 5, right lower panel), while they are smaller than<br />

90 ◦ (Fig. 5, right upper panel) for the first model (Fig. 5, left upper panel) with the narrower deep<br />

anisotropic block.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

92


Figure 5: Models (left) and their forward responses (right). The models differ only in lateral extension<br />

of anisotropy block. The lateral extension of the second block is about 15km in the first model (left<br />

upper panel) and 80km (left lower panel) in the second model. The major difference in the responses<br />

appears in the yx component. Phases over 90 ◦ occur if the lateral extension of the deeper anisotropy<br />

block is greater than it of the shallower anisotropy block.<br />

4.2 Depth<br />

In a second step, we used the same anisotropy parameters as described above. The initial model<br />

consists of a 300m isotropic surface layer and a isotropic half-space with resisitivity of 100Ωm. Two<br />

anisotropy blocks are embedded in the half-space. The first block started at depth of 300m and shaped<br />

as a trapezoid. In this attempt we vary the depth of the second block. In the first model (Fig. 6, left<br />

upper panel) the second block started at a depth of 1.8km and in the second model (Fig. 6, left lower<br />

panel) started at a depth of 6.8km.<br />

In the forward responses (Fig. 6, right) we see that the phases Φyx for the second model become<br />

larger than 90 ◦ at periods > 10s at sites above the first block, while they for the first model are<br />

all smaller than 90 ◦ . In this attempt the second anisotropy block has always sufficiently greater<br />

lateral extension than the first anisotropy block, but we only see the phase anomalies by model with<br />

anisotropy block in adequate depth. Besides the lateral extension of the anisotropy block, also the<br />

depth is one of those key conditions under which phase anomalies appear.<br />

4.3 Rotation angle<br />

In a third test we use a similar model as described above for step two. The model contains a 300m<br />

isotropic surface layer and an isotropic half-space with resisitivity of 100Ωm. Two anisotropic blocks<br />

are embedded in the half-space. The first block starts directly beneath the surface layer and the<br />

second block in a depth of 6km. They have the principal resistivities σ1/σ2/σ3 = 50/0.5/50Ωm and<br />

σ1/σ2/σ3 = 30/0.3/30Ωm, respectively. In this attempt we vary the anisotropy strike angle αS1 and<br />

αS2 for both blocks.<br />

The forward response for the model with αS1 = αS2 = 0 ◦ is displayed in the right upper panel<br />

of figure 7. The phases of the yx component approach 90 ◦ for long periods. For αS1 = 60 ◦ and<br />

αS2 = 120 ◦ (Fig. 7, left lower panel) the yx component phases become larger than 90 ◦ for periods<br />

> 10s, while they stay below 90 ◦ for αS1 = 90 ◦ and αS2 = 120 ◦ (Fig. 7, right lower panel). Comparing<br />

the three examples, we see that model with different combination of strike angle in both shallow and<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

93


Figure 6: Models (left) and their forward responses (right). The models differ by depth of the second<br />

anisotropy block. In the first model (left upper panel) the second block starts in a depth of 1.8km<br />

and in the second model (left lower panel) the block starts in a depth of 6.8km. The phases of yx<br />

component for the model with the deeper anisotropic block become larger than 90 ◦ at periods > 10s<br />

at sites above the first block. For the model with the shallower block all phases are smaller than 90 ◦ .<br />

αS1 and αS2<br />

αS1 = 0 ◦ and αS2 = 0 ◦<br />

αS1 = 60 ◦ and αS2 = 120 ◦ αS1 = 90 ◦ and αS2 = 120 ◦<br />

Figure 7: Model (left upper panel) and its forward responses with different anisotropy strike angles.<br />

The model contains the same surface layer, background medium and resistivities for both anisotropic<br />

blocks as the models in fig. 6. We vary the strike angles αS1 and αS2 for the shallow and the deep<br />

block (left upper panel). The forward responses are displayed as apparent resistivities and phases in<br />

xy and yx component.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

94


deep blocks produce forward responses with pronounced difference. This can be observed not only in<br />

yx component (ρa(yx) and Φyx ) but also in xy component (ρa(xy) and Φxy ). A possible explanation is<br />

that a change of strike angle forces the current flow to change its preferred direction from the shallow<br />

to the deep block. According to this fact we may conclude that the anisotropy strike angle influences<br />

forward responses significantly and phase greater than 90 ◦ will appear if the strike angle of both<br />

blocks differ by an adequate amount.<br />

5 Conclusions<br />

Phases greater than 90 ◦ can only be modelled by a combination of a deep and a shallow anisotropic<br />

block. We varied several model parameters to study the changes in forward responses. For our models<br />

we can conclude that phases out of the first quadrant occur when: (i) the anisotropy ratio (σmax/σmin)<br />

is high (for both blocks); (ii) the deep anisotropic block has a much larger lateral extension than the<br />

shallow block; (iii) the deep anisotropic block is located in greater depth; (iv) the angles of anisotropy<br />

of both blocks differ by a considerable amount. In our models the difference should be at least 45 ◦ so<br />

that the preferred direction of current flow changes significantly from the shallow to the deep block.<br />

References<br />

Becken, M., & Burkhardt, H. (2004). An ellipticity criterion in magnetotelluric tensor analysis.<br />

Geophys. J. Int., 159, 69–82.<br />

Eisel, M., & Haak, V. (1999). Macro-anisotropy os the electrical conductivity of the crust: a magnetotelluric<br />

study from the German continental Deep Drilling site (KTB). Geophys. J. Int., 136, 109–<br />

122.<br />

Hälbich, I. (1993). Cape Fold Belt – Agulhas Bank Transect across Gondwana suture, southern<br />

Africa. Geosci. Transect, 9, 18 pp. AGU, Washington, D. C.<br />

Heise, W., & Pous, J. (2003). Anomalous phases exceeding 90 ◦ in magnetotellurics: anisotropic<br />

model studies and a field example. Geophys. J. Int., 155, 308–318.<br />

Kellett, R. L., Mareschal, M., & Kurtz, R. D. (1992). A model of lower crustal electrical anisotropy<br />

for the Pontiac Subprovince of the Canadian Shield. Geophys. J. Int., 111, 141–150.<br />

Mareschal, M., Kellett, R. L., Kurtz, R. D., Ludden, J. N., Ji, S., & Bailey, R. C. (1995). Archaean<br />

cratonic roots, mantle shear zones and deep electrical anisotropy. Nature, 375, 134–137.<br />

O’Brien, D. P., & Morrison, H. F. (1967). Electromagnetic fields in an N-layer anisotropic half space.<br />

Geophysics, 32, 668–677.<br />

Pek, J., & Verner, T. (1997). Finite-difference modelling of magneto-telluric fields in two-dimensional<br />

anisotropic media. Geophys. J. Int., 128, 505–521.<br />

Reddy, I. K., & Rankin, D. (1975). Magnetotelluric response of lateerally inhomogeneous and<br />

anisotropic media. Geophysics, 40, 1035–1045.<br />

Ritter, O., Junge, A., & Dawes, G. J. K. (1998). New equipment and processing for magnetotelluric<br />

remote reference observations. Geophys. J. Int., 132, 535–548.<br />

Rodi, W., & Mackie, R. L. (2001). Nonlinear conjugate gradients algorithm for 2-d magnetotelluric<br />

inversion. Geophysics, 66, 174–187.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

95


Stankiewicz, J., Ryberg, T., Schulzw, A., Lindeque, A., Weber, M., & de Wit, M. (2007). Initial<br />

results from wide-angle seismic refraction lines in the southern Cape. South African Journal of<br />

Geology, 110, 407–418.<br />

Weckmann, U., Jung, A., Branch, T., & Ritter, O. (2007). Comparison of electrical conductivity<br />

structures and 2D magnetic modelling along two profiles crossing the Beattie Magnetic Anomaly,<br />

South Africa. South African Journal of Geology, 110, 449–464.<br />

Weckmann, U., Magunia, A., & Ritter, O. (2005). Effective noise separation for magnetic single site<br />

data processing using a frequency domain selection scheme. Geophys. J. Int., 161, 635–652.<br />

Weckmann, U., Ritter, O., & Haak, V. (2003). A magnetotelluric study of the damara belt in namibia<br />

– 2. MT phases over 90 ◦ reveal the internal structure of the Waterberg Fault/Omaruru Lineament.<br />

Phys. Earth Planet. Inter., 138(2), 91–112.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

96


Substitute models for static shift in 2D<br />

Kristina Tietze 1,2 , Oliver Ritter 1,2 , Ute Weckmann 1,3<br />

1 Helmholtz Centre Potsdam - German Research Centre For Geosciences <strong>GFZ</strong>, Department 2, Section 2.2, Telegrafenberg, 14473 Potsdam<br />

2 Free University Berlin, Department of Earth Sciences, Malteserstr. 74-100, 12249 Berlin<br />

3 University of Potsdam, Department of Geosciences, Karl-Liebknecht-Str. 24, 14476 Potsdam<br />

Contact: kristina.tietze@gfz-potsdam.de<br />

Introduction<br />

MT practitioners often down-weight apparent resistivity TE mode data prior to 2D inversion to avoid<br />

problems with static shift, obviously assuming that static shift of the TM mode is handled automatically<br />

by the inversion. Static shift is caused by conservation of charges at local conductivity discontinuities<br />

which are small with respect to the inductive scale length.<br />

Here we present a class of shallow conductivity anomalies which can produce significant up- or<br />

downwards shifted TM mode apparent resistivity curves (static shift in the TE mode cannot be<br />

simulated with 2D modeling). We examine how this static shift is reproduced by 2D inversion and show<br />

that the results are strongly influenced by grid design and regularization. We conclude that modern 2D<br />

inversion packages are not optimized to handle static shift. Our results also indicate that it is no good<br />

reason to assume that static shift cancels out on average.<br />

2D substitute structures for static shift<br />

As 2D models only allow for conductivity changes in one of the horizontal directions, static shift can only<br />

be accounted for in one component (TM) and re-produced by forward modeling.<br />

Fig. 1 (right): Starting from a 1D layered halfspace<br />

(LH) model (Fig. 1a) 2D inhomogeneities<br />

were added to the resistivity model just<br />

below/at the surface (Figs 1b-g, left panel). TE<br />

and TM model responses were calculated for<br />

frequencies between 10 3 Hz and 10 -3 Hz directly<br />

above the center of the inhomogeneities. The<br />

2D forward computations are compared to the<br />

1D LH results indicated by crosses (Figs 1b-g,<br />

right panel).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

97


Fig. 1b (left): A 15 m wide 10 m conductor with a<br />

thickness varying from 1 m to 5 m causes downward shift<br />

of the TM resistivity curves. The shift increases in<br />

dependence of the block thickness, i.e. the extent of the<br />

horizontal resistivity contrast. The amount of static shift<br />

which can be generated by extending the conductive inset<br />

downwards is limited. At some point the structure is<br />

becoming inductively effective.<br />

Fig. 1d (left): Enhancing the conductivity contrast by<br />

padding the conductor with resistive (10 3 m) cells TM<br />

downward shift is increased. Structures of this type can be<br />

observed in inversion models below static shift affected<br />

sites when static shift is not taken into account during<br />

inversion (cf. pane view in Fig. 2c (2)).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

98<br />

Fig. 1c (left): For a fixed thickness of<br />

3 m the lateral dimensions of the 10<br />

m-block were altered between 15<br />

m and 65 m. With increasing block<br />

width, the static shift effect becomes<br />

smaller as the lateral resistivity<br />

contrast which is responsible for the<br />

distortion of the electrical field<br />

moves away from the site location.<br />

Fig. 1e (left): Placing a highly<br />

resistive block (10 4 m) with a width<br />

of 15 m and varying thicknesses<br />

between 5 m and 110 m beneath the<br />

site results in upward shift of the TM<br />

mode. Static shift increases with<br />

increasing vertical block dimensions.


Fig. 1f (left): For a fixed thickness of<br />

55m the lateral extent of the 10 4 mblock<br />

was altered between 15 m and 65<br />

m. As seen for the 10 m-block, the<br />

static shift effect becomes smaller with<br />

increasing block width as the resistivity<br />

contrast distorting the electrical field<br />

moves away from the site location.<br />

Fig. 1g (left): Padding the resistive block<br />

with a thin column of conductive (10<br />

m) cells increases the upward shift of<br />

the TM resistivity curve slightly. The<br />

horizontal padding cells must be very<br />

small horizontally to prevent inductive<br />

effects.<br />

As expected, the TE resistivity curves and the phase curves of both, TM and TE, modes are mostly<br />

unaffected by these very small scale structures.<br />

Fig. 1h (left): Distortion of the TM<br />

apparent resistivities is due to static<br />

shift as the skin depth exceeds 15 km<br />

for =10 m and f=0.01 Hz. The<br />

downward shifting effect above the<br />

good conductor is significantly stronger<br />

than the upward shift above the<br />

resistive block, although resistivity<br />

contrasts of the two blocks to the<br />

background are the same. Static shift<br />

values along this profile do not sum up<br />

to zero. The figures above indicate that<br />

structures causing upwards shift have to<br />

have a larger vertical extent to produce<br />

the same amount of shift.<br />

So, it is at least questionable if a zero sum/average assumption for static shift values, which is often<br />

applied in inversion schemes, is appropriate.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

99


2D Inversion<br />

Fig. 2a (left): A 2D resistivity model was<br />

created by adding a low and a high<br />

resistive block to the layered half-space<br />

model (cf. Fig. 1a). Data were calculated<br />

for 50 sites along a 50km long profile with<br />

1 km site spacing adding 2.5% Gaussian<br />

noise.<br />

Fig. 2b (above): Right: Synthetic shift values c were generated obeying a modified logarithmic Gammadistribution:<br />

log10(c) = -(0.5,2)+0.5 and shift=c. The maximum of the modified distribution is at 10 0 ,<br />

the expected value is 10 -0.5 . Left: Shift was applied randomly to 2/3 of the sites, independently for TE<br />

and TM. The left panel shows the applied shift values for each site along the profile.<br />

Fig. 2c (below): Inversion results for un-shifted (1) and shifted (2)-(4) data for 100 iterations starting<br />

from a 100m homogeneous half space, applying a uniform smoothing with =10. Error floors were 2%<br />

(0.6°) for the phases and 5% and 500% for TM and TE apparent resistivities, respectively.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

100


Fig. 2c (above, continued): The model grid was created setting the column widths to 0.5 times of the<br />

minimum skin depth pmin, for (1) and (2), and was refined to 0.1*pmin for (3) and (4) (cf. pane views).<br />

(2) & (3): As smoothing (regularization) is working against sharp resistivity contrasts, the inversion<br />

introduces complex and wide-stretched structures below the sites affected by static shift. The refined<br />

grid in (3) reduces smoothing lengths for uniform regularization. As a consequence, introduced<br />

compensatory structures appear simpler or diminish.<br />

(4) Inversion result after introducing a so-called “tear zone” (feature of WinGLink) at site 114 (red<br />

outline). Tear zone inversion does not penalize sharp conductivity contrasts at the outline of the tear<br />

zone. Now, the inversion can properly model static shift. The model can account for TM static shift<br />

within this zone while obtaining surrounding resistivity values close to the original ones.<br />

Fig. 2d (right) shows un-shifted data,<br />

shifted data as symbols and the inversion<br />

results (cf. Fig. 2c) for sites 114 and 133 as<br />

lines. For periods longer than 10 -2 s shifted<br />

data are fitted well by all inversions. For<br />

shorter periods, the inversion struggles to<br />

fit the data, especially at site 114 where<br />

TM resistivities were shifted downwards by<br />

two decades. TM phase responses below<br />

10 -2 s deviating from forward data clearly<br />

show the inductive effects of the near<br />

surface structures. Static shift effects<br />

decrease with the refined grid (3) and<br />

introduction of a tear zone (4).<br />

Summary<br />

Static shift for the TM mode can easily be produced by a range of simple structures at surface being<br />

introduced to the regional resistivity models as substitutes for natural structures causing static shift in<br />

field data.<br />

To account for TM static shift within 2D inversion, model grids have to be very fine in the vicinity of sites<br />

with column widths and thicknesses much smaller than the induction volume of the highest frequency<br />

to be analyzed.<br />

Smoothing routinely applied in minimum structure inversions does not allow for the sharp conductivity<br />

contrasts required to produce sensible static shift. It would therefore be desirable to be able to apply<br />

different smoothing factors to a static shift compensating top layer and the remaining model.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

101


Die Übergangsimpedanz einer kapazitiv angekoppelten Elektrode<br />

Andreas Hördt, Peter Weidelt, Anita Przyklenk<br />

Institut für Geophysik und extraterrestrische Physik, TU Braunschweig<br />

Vorwort von A. Hördt<br />

Die in diesem Artikel vorgestellte Theorie beruht auf einer unveröffentlichen Arbeit von Prof.<br />

Peter Weidelt, der am 1.7.2009 unerwartet verstorben ist. Die Arbeit ist ein typisches Beispiel<br />

für seine außerordentliche Hilfsbereitschaft, für die er unter Kollegen in aller Welt bekannt<br />

war. Im Jahr 2007 hatte ich ihn um Unterstützung bei der Berechnung der<br />

Übergangsimpedanz einer kapazitiven Elektrode gebeten. Einige Wochen später hatte er die<br />

vollständige Lösung hergeleitet und in ein Programm umgesetzt. Er selbst hat die Aufgabe als<br />

nicht allzu schwierig oder wichtig empfunden, und hätte die Ergebnisse vermutlich nie<br />

veröffentlicht.<br />

1 Zusammenfassung<br />

Es wird eine Möglichkeit vorgestellt, die Impedanz einer Kreisscheibe zu berechnen, die an<br />

einen Halbraum mit endlicher Leitfähigkeit angekoppelt ist. Basierend auf den<br />

Maxwellgleichungen wird eine Integralgleichung für die Ladungsdichte q auf der<br />

Kreisscheibe aufgestellt. Mit der Randbedingung, dass das Potential auf der Scheibe konstant<br />

ist, kann die Ladungsdichte gender zu angehobener Platte stetig aund damit die Gesamtladung<br />

bestimmt werden. Es zeigt sich, dass der Übergang von aufliels Funktion des Abstandes<br />

verläuft. Zudem nimmt die Impedanz als Funktion des Abstandes zu, d.h. die Ankopplung<br />

einer kapazitiv angekoppelten Elektrode kann niemals geringer sein, als die einer<br />

aufliegenden mit gleicher Fläche. Je nach Modellparametern kann die Impedanz jedoch<br />

innerhalb weniger Nanometer über Größenordnungen variieren, so dass es dennoch günstig<br />

sein kann, die Elektrode zu isolieren, um starke Schwankungen der Impedanz zu vermeiden.<br />

2 Einleitung<br />

Bei geoelektrischen Messungen über sehr schlecht leitendem Untergrund kann es von Vorteil<br />

sein, den Strom kapazitiv anzukoppeln (Kuras et al., 2006). Dabei wird ein hochfrequentes<br />

Wechselfeld an eine Elektrode angelegt, die keinen direkten Kontakt mit dem Untergrund hat.<br />

Eine genaue Berechnung der Übergangsimpedanz einer solchen Elektrode ist notwendig, um<br />

zu beurteilen, unter welchen Bedingungen eine kapazitive Ankopplung einer galvanischen<br />

überlegen ist. Für die praktische Umsetzung ist die Frage von Bedeutung, ob es günstig ist,<br />

eine kapazitive Elektrode zu isolieren und einen galvanischen Kontakt auszuschließen. Die<br />

Standardformeln für die Ankopplung kapazitiver Elektroden sind allerdings nur für hohe<br />

Leitfähigkeiten des Untergrundes gültig. Hördt (2007) hat analytische Gleichungen für eine<br />

Kugelelektrode im sphärisch geschichteten Vollraum hergeleitet und diskutiert. Hier wird nun<br />

eine Lösung für eine zylindrische Scheibe über einem Halbraum hergeleitet.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

102


3 Modellbeschreibung<br />

Abbildung 1 illustriert das zu lösende Problem. Zu berechnen ist, welcher Strom I bei einer<br />

vorgegebenen Spannung U zwischen zwei zylindrischen Scheiben über einem homogenen<br />

Halbraum fließt.<br />

Abbildung 1: Skizze der kapazitiv angekoppelten Stromquelle: Zwei zylindrische Scheiben<br />

sind kapazitiv an einen homogenen Halbraum mit Leitfähigkeit und Permittivität <br />

angekoppelt, es fließt ein Strom I bei vorgegebener Spannung U.<br />

Wenn der Untergrund ein idealer Leiter ist, gilt bei kleinem Abstand zwischen Elektrode und<br />

Untergrund für die komplexe Übergangsimpedanz (Smythe, 1968):<br />

1<br />

Z (1)<br />

iC<br />

wobei die Kreisfrequenz ist und C die Kapazität, mit<br />

A<br />

0<br />

C (2)<br />

d<br />

Dabei ist A die Fläche der Elektrode, d der Abstand zum Untergrund, und 0 die<br />

Dielektrizitätszahl des Vakuums zwischen den Elektroden. diese Formeln stellen allerdings<br />

eine Näherung für einen idealen Leiter dar. Je nach Frequenz und Leitfähigkeit gilt die<br />

Näherung nicht mehr, und die elektrischen Parameter des Untergrundes müssen explizit<br />

berücksichtigt werden. Dies erfordert eine vollständige Lösung der Maxwellgleichungen.<br />

Die geometrischen Parameter des Modelles sind in Abbildung 2 illustriert. Die ideal leitende<br />

Kreisscheibe mit Radius a befindet sich im Abstand d über dem Untergrund. Der Mittelpunkt<br />

der Scheibe ist der Ursprung des zylindrischen Koordinatensystemes. Es wird angenommen,<br />

dass alle zeitlichen Variationen harmonisch sind, die zeitliche Abhängigkeit sich also durch<br />

e it darstellen lässt. Die Leitfähigkeit des Untergrundes lässt sich dann als komplexe Größe<br />

darstellen: 1 1 i<br />

1,<br />

wobei 1 die Gleichstromleitfähigkeit ist. Das Medium zwischen<br />

Untergrund und Scheibe hat die rein imaginäre Leitfähigkeit 0 i<br />

0 .<br />

Wenn an die Scheibe ein Potential U angelegt wird, stellt sich eine radialsymmetrische<br />

Ladungsverteilung q(r) ein. Die Gesamtladung Q erhält man dann aus<br />

a<br />

<br />

Q 2 qds<br />

s<br />

(3)<br />

0<br />

U<br />

~<br />

, <br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

103<br />

I


Abbildung 2: Parameter des Modelles der ideal leitenden Kreisscheibe über homogenem<br />

Halbraum.<br />

Es muss also die Flächenladungsdichte q(r) gefunden werden, aus der dann Q und mit<br />

Q<br />

C (4)<br />

U<br />

die komplexe Kapazität, bzw. mit Gl. (1) die komplexe Impedanz der Kreisscheibe berechnet<br />

werden kann.<br />

Aus der Forderung, dass die Leitfähigkeit auf der Scheibe unendlich ist, ergibt sich, dass das<br />

elektrische Feld auf der Scheibe verschwindet. Die Lösungsstrategie besteht darin, eine<br />

Gleichung für das elektrische Feld im gesamten Raum als Funktion von q(r) zu berechnen<br />

und dann q(r) so zu bestimmen, dass die Radialkomponente Er auf der Scheibe verschwindet.<br />

4 Theorie<br />

4.1 Grundgleichungen<br />

Um das elektrische Feld zu berechnen, müssen die Maxwellgleichungen in<br />

Zylinderkoordinaten gelöst werden. Aus dem Amperegesetz folgt unter der Berücksichtigung<br />

der harmonischen Zeitabhängigkeit:<br />

<br />

E<br />

E iEErot H<br />

(5)<br />

t<br />

mit dem elektrischen Feld E und dem Magnetfeld H. Die Leitfähigkeit und die<br />

Dielektrizitätszahl hängen von z ab. Aus Symmetriegründen besitzt das elektrische Feld nur<br />

eine Vertikal- und eine Radialkomponente, die nur von z und r abhängen. Das Magnetfeld<br />

besitzt nur eine von r und z abhängige -Komponente.<br />

Damit folgt aus Gl. (5):<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

104


H<br />

<br />

Er <br />

(6a,b)<br />

z<br />

1 r<br />

H <br />

E z <br />

r r<br />

Aus dem Induktionsgesetz:<br />

B<br />

rotE<br />

t<br />

folgt:<br />

(7)<br />

E<br />

z E<br />

r<br />

i<br />

0 H <br />

(8)<br />

r<br />

z<br />

<br />

Durch Einsetzen von (6a) und (6b) in (8) erhält man eine entkoppelte Gleichung für H:<br />

1 <br />

rHzzH i<br />

H <br />

1<br />

r r<br />

0<br />

(9)<br />

r <br />

<br />

<br />

wobei die Schreibweise r angewandt wurde.<br />

r<br />

Der erste Term in Gl. 9 legt eine Trennung der Variablen mit Hilfe der Besselfunktion 1.<br />

Ordnung mit der Wellenzahl u nahe:<br />

r, zfz,<br />

uJur<br />

H 1<br />

(10)<br />

denn die Besselfunktion erfüllt die Bessel’sche Differentialgleichung:<br />

ur r J1<br />

r<br />

2<br />

r J1<br />

1<br />

2<br />

r<br />

so dass mit (10) folgt:<br />

2<br />

ur J uru J ur 2<br />

rHuH 1<br />

1<br />

r r r<br />

1<br />

<br />

(12)<br />

<br />

Bei dieser Strategie besteht eine Analogie zur Fouriertransformation, bei welcher z.B. eine<br />

zweifache zeitliche Ableitung nach Transformation in den Frequenzbereich mit der Funktion<br />

e it zu einer Multiplikation mit – 2 führt. Entsprechend führt hier der Differentialoperator im<br />

1. Term in Gl. (9) zu einer Multiplikation mit –u 2 . Die Lösung von (9) lässt sich folglich<br />

darstellen als:<br />

z, uJurdu<br />

H r,<br />

z)<br />

f 1<br />

<br />

( <br />

(11)<br />

(13)<br />

0<br />

Die Wahl der Besselfunktion 1. Ordnung ergibt sich daraus, dass aus Symmetriegründen H<br />

bei r=0 verschwinden muss, was von J1 erfüllt wird, aber nicht von J0. Aus dem<br />

Amperegesetz (6a und 6b) ergeben sich damit direkt die Gleichungen für die Komponenten<br />

des elektrischen Feldes:<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

105


z, u<br />

1<br />

f<br />

Er ( r,<br />

z)<br />

J1<br />

z<br />

1<br />

E z ( r,<br />

z)<br />

u<br />

f 0<br />

<br />

<br />

0<br />

<br />

0<br />

urdu z, uJurdu<br />

Zur Bestimmung von f(z,u) wird zunächst Gl. (13) herangezogen. Durch Einsetzen in (9) sieht<br />

man unter Benutzung von (12), dass die Funktion f(z,u) die Gleichung<br />

(14)<br />

(15)<br />

1 2<br />

z z f f<br />

<br />

<br />

erfüllen muss, mit<br />

(16)<br />

2<br />

2<br />

i 0<br />

u<br />

(17)<br />

4.2 Bestimmung von f(z,u)<br />

Gleichung (14) muss unter der Berücksichtigung der Randbedingungen an den<br />

Schichtgrenzen und der Quellbedingung gelöst werden. Die Quelle wird durch einen Sprung<br />

der Vertikalkomponente des elektrischen Feldes an der Scheibe beschrieben:<br />

r <br />

E z<br />

wobei <br />

<br />

q<br />

r<br />

<br />

<br />

(18)<br />

<br />

0<br />

die Differenz der Werte unmittelbar oberhalb und unterhalb der Scheibe, also den<br />

Sprung des Funktionswertes, kennzeichnet.<br />

<br />

<br />

<br />

Ez0Ez0<br />

E (18a)<br />

<br />

Um aus (18) eine Bedingung für f(r,z) abzuleiten, transformien wir sie in den<br />

Wellenzahlbereich mittels Gl. (15), aus der folgt:<br />

<br />

1<br />

<br />

<br />

0<br />

<br />

<br />

( r,<br />

z)<br />

ufz,<br />

u<br />

J ur E z<br />

<br />

0<br />

r<br />

q<br />

, 0 r a<br />

0 <br />

du <br />

0,<br />

r a<br />

<br />

<br />

Die Ladungsdichte q(r) läßt sich mittels der Hankeltransformation in den Wellenzahlbereich<br />

transformieren:<br />

q~ <br />

<br />

0<br />

0<br />

u uqrJurdr<br />

(19)<br />

(20)<br />

Mit q u ~ als der Ladungsdichte im Wellenzahlbereich.<br />

Dementsprechend ist q(r) ist darstellbar als:<br />

q<br />

<br />

<br />

0<br />

0<br />

r uq~<br />

u J urdu (21)<br />

Durch Einsetzen von (21) in (19) wird klar, dass<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

106


0<br />

f z, u <br />

q~<br />

u (22)<br />

0<br />

d.h.die im Raumbereich formulierte Randbedingung (18) transformiert sich gemäß (22) in den<br />

Wellenzahlbereich.<br />

Nach der Formulierung der Randbedingungen wird nun die Differenzialgleichung für f gelöst.<br />

Aus der Form von Gleichung (16) ergibt sich direkt, dass die Lösungen Funktionen der Form<br />

z<br />

e <br />

sind, wobei berücksichtigt werden muss, dass im unteren Halbraum und in der Luft<br />

unterschiedlich sind. Daraus ergibt sich folgender Ansatz für die Funktion f(z,u):<br />

f<br />

z, u<br />

wobei<br />

0 Ae<br />

<br />

0z<br />

<br />

Ae<br />

<br />

<br />

Ce<br />

<br />

z<br />

<br />

<br />

Be<br />

<br />

z<br />

1<br />

Be<br />

,<br />

<br />

z<br />

<br />

z<br />

0<br />

,<br />

0<br />

,<br />

0 z d<br />

z d<br />

z 0<br />

(23a,b,c)<br />

2<br />

2<br />

2<br />

2<br />

0 i0 0<br />

u und 1 i 0<br />

1<br />

u<br />

(24a,b)<br />

die Parameter in Luft und im unteren Halbraum sind.<br />

Die Vorzeichen der Exponenten in den Ansatzfunktionen in (23) ergeben sich jeweils aus der<br />

Forderung, dass f(z,u) im Unendlichen sowohl für z0 verschwinden muss.<br />

Die Randbedingungen an Schichtgrenzen ergeben sich aus den Stetigkeitsbedingungen für die<br />

elektromagnetischen Felder. Da die Horizontalkomponenten von E und H an Schichtgrenzen<br />

1 f<br />

stetig sind, folgt mittels (13) und (14), dass f und stetig sein müssen.<br />

z<br />

Damit und mit der Sprungbedingung für f (Gl. (22)) lassen sich nun die Konstanten A,B und C<br />

in Gl. (23) berechnen. Beispielsweise folgt aus (23a,b), dass<br />

f<br />

z 0 <br />

_ B A<br />

z0 A B<br />

f<br />

<br />

und damit für den Sprung von f bei z=0:<br />

<br />

<br />

<br />

fz, u<br />

f z0fz02A (25a,b)<br />

(26)<br />

und mit (22) erhält man hieraus für A:<br />

0<br />

A q~<br />

u (27)<br />

2 0<br />

Die Bestimmung von B und C aus den Stetigkeitsbedingungen erfolgt analog, ist aber<br />

komplexer und wird hier nicht im Einzelnen ausgeführt. Man erhält folgendes Endergebnis<br />

für f:<br />

f<br />

<br />

2<br />

0 z, u<br />

q~<br />

u 0<br />

<br />

<br />

e<br />

<br />

<br />

e<br />

<br />

T<br />

<br />

0 z<br />

0 z 2<br />

d <br />

Rue<br />

0 z<br />

0 z2d Rue<br />

,<br />

1<br />

zd <br />

0d<br />

ue e ,<br />

mit Reflektions-und Tranmissionsfaktor<br />

,<br />

z 0<br />

0 z d<br />

z d<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

107<br />

(28a,b,c)


R<br />

T<br />

u u <br />

<br />

0<br />

u11u0 u11u0 2<br />

0u1<br />

u11u0 0 (29)<br />

(30)<br />

<br />

0<br />

Mit (28-30) und (13-15) lassen sich dann alle Feldkomponenten berechnen, wobei für die<br />

Bestimmung der Ladungsdichte nur die Gleichung für Er im Bereich zwischen Halbraum und<br />

Scheibe benötigt wird. Im Folgenden wird die Näherung 0=u verwendet. Im Vakuum ist<br />

nämlich<br />

2<br />

2 2 2<br />

2 2<br />

0 u <br />

0<br />

0 u u<br />

(24c)<br />

2<br />

c<br />

Die relevanten Wellenzahlen u ergeben sich aus der Dimension des Messystemes, welches bei<br />

den hier berücksichtigten Frequenzen in jedem Fall klein ist gegen die elektromagnetische<br />

Wellenänge im Vakuum. Es ist also u c und damit der 1. Term auf der rechten Seite<br />

vernachlässigbar.<br />

Die Gleichung für Er ergibt sich aus (28b) und (14), wobei die Ableitung von f nach z einen<br />

Faktor u unter dem Integral ergibt:<br />

E<br />

u<br />

z<br />

uz2d<br />

ueRueJurdu 0 z d<br />

<br />

1<br />

r ( r,<br />

z)<br />

uq~<br />

1<br />

2<br />

0 0<br />

4.3 Bestimmung der Ladungsdichteverteilung<br />

Zur Bestimmung der Ladungsdichteverteilung wird wir oben erläutert gefordert, dass die<br />

Radialkomponente des elektrischen Feldes auf der Scheibe verschwindet. Benötigt wird also<br />

das Feld bei z=0:<br />

E<br />

1<br />

<br />

rz0uq~ u1Ru , <br />

2u<br />

d<br />

e Jurdu r 1<br />

2<br />

0 0<br />

Die Idee ist nun, eine Ladungsverteilung q(r) zu suchen, so dass Er überall verschwindet.<br />

Damit lässt sich Gleichung (32) im Prinzip in ein Gleichungssystem überführen. Setzt man<br />

die Transformation der Ladungsdichte (Gl. 20) in (32) ein, so erhält man:<br />

1<br />

2<br />

a<br />

<br />

0 0 0<br />

sq<br />

2u<br />

d<br />

s J s u1R<br />

u e Jurduds 0<br />

0<br />

1<br />

Dabei wurde die neue Variable s eingeführt, die ebenso wie r von 0 bis a läuft, allerdings<br />

verschieden von r sein muss, da ja das elektrische Feld bei r von der gesamten<br />

Ladungsdichteverteilung abhängt. Definiert man nun die Funktion:<br />

F<br />

<br />

sr u 1<br />

Ru<br />

, <br />

0<br />

2u<br />

d<br />

e JusJurdu 0<br />

1<br />

(31)<br />

(32)<br />

(33)<br />

(34)<br />

so kann man Gl. (33) schreiben als:<br />

1<br />

2<br />

0<br />

a<br />

<br />

0<br />

s q<br />

s Fs,<br />

r<br />

ds 0<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

108<br />

(35)


Die Funktion F(s,r) lässt sich nach Gl. (34) für jedes s und r direkt durch Integration über u<br />

berechnen. Die Scheibe wird also geeignet diskretisiert, und mit den diskreten Werten für s<br />

und r erhält man eine Matrix F . Gleichung 35 wird dann zur linearen Gleichung<br />

F q 0<br />

(36)<br />

in der die Einträge von q die unbekannte Ladungsdichteverteilung darstellen. Durch<br />

Gleichung (36) ist q allerdings noch nicht vollständig bestimmt, da sie die triviale Lösung<br />

q 0 besitzt.<br />

Um eine endliche Ladungsdichte zu erhalten, benötigt man eine Gleichung für das Potential<br />

der Scheibe. Eine Möglichkeit besteht darin, ein aus 2 identischen Scheiben bestehendes<br />

System zu betrachten (Abb. 3).<br />

Abbildung 3: Geometrie des aus zwei identischen Scheiben aufgebauten Sendesytemes. Der<br />

Abstand der Mittelpunkte ist L, der Radius der Scheiben a.<br />

Die Integration von Er entlang der Verbindungslinie der beiden Scheibenmittelpunkte ergibt<br />

die angelegte Spannung U. Da Er=0 auf den Scheiben, gilt:<br />

LaLa<br />

Er0ELr, 0dr<br />

2 E r, <br />

r<br />

r <br />

U , 0 dr<br />

(37)<br />

a<br />

a<br />

r<br />

Er(r,0) wurde in Gl. (32) berechnet und lässt sich direkt einsetzen. Die Integration über r kann<br />

man ausführen, und man erhält:<br />

U 2<br />

1<br />

<br />

<br />

1<br />

<br />

<br />

La<br />

<br />

<br />

0 0<br />

<br />

<br />

0 0<br />

<br />

a<br />

uq~<br />

q~<br />

1<br />

2<br />

0 0<br />

uq~<br />

u1Ru 2u<br />

d<br />

e Jur u1Ru L<br />

a<br />

2u<br />

d<br />

e J1ur 2u<br />

d<br />

e J uaJuLa u1Ru <br />

<br />

a<br />

<br />

drdu<br />

<br />

du dr<br />

du<br />

0<br />

1<br />

0<br />

Um eine explizite Diskretisierung definieren zu können, wird die Ladungsdichte wieder<br />

mittels Gl. (20) im Raumbereich aufgeschrieben. Fasst man alle Terme mit u zusammen, so<br />

erhält man:<br />

1<br />

U <br />

<br />

1<br />

<br />

<br />

mit<br />

a<br />

<br />

0 0<br />

a<br />

<br />

0 0 0<br />

sq<br />

s J usds 1<br />

R<br />

u<br />

0<br />

sq<br />

s Gds<br />

s<br />

2u<br />

d<br />

e J uaJuLa <br />

0<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

109<br />

0<br />

du<br />

(38)<br />

(39)


G<br />

<br />

<br />

0<br />

2u<br />

d<br />

e JusJuaJuLa s R<br />

u<br />

0<br />

du<br />

1 (40)<br />

0<br />

0<br />

Gleichungen (39) und (40) werden wiederum durch Diskretisierung der Scheibe in eine<br />

Gleichung der Form<br />

q G U<br />

(41)<br />

umgesetzt.<br />

Gleichung (41) ist die inhomogene Bedingung, die nötig ist, um die triviale Lösung q 0<br />

auszuschließen und eine vom Potential abbhängige Ladungsdichte zu erhalten. Man wählt<br />

also ein beliebiges U und erhält aus der Lösung von (36) unter der Bedingung (41) eine<br />

Ladungsdichteverteilung q(r). Die Gesamtladung erhält man mit Gl. (3), und die Kapazität<br />

aus Gl. (4).<br />

Für ein aus zwei Scheiben bestehendes System gilt die Lösung nur in hinreichendem Abstand<br />

der Scheiben zueinander, da die gegenseitige Beeinflussung der Ladung auf den beiden<br />

Scheiben hier nicht berücksichtigt wurde. Eine quantitative Abschätzung dieses Effektes<br />

wurde bisher nicht vorgenommen. Betrachtungen mit halbkugelförmigen Elektroden im<br />

Gleichstromfall legen nahe, dass die vernachlässigten Terme von der Ordnung (a/L) 4 sind, so<br />

dass die Näherung schon bei geringen Elektrodenabständen gut funktioniert.<br />

Die berechnete Kapazität C der Elektrode ist komplex, da der Reflektionsfaktor (Gl. 29)<br />

komplex ist. Aus ihr lässt sich mit<br />

1<br />

Z <br />

iC<br />

die komplexe Impedanz der Elektrode, bzw. deren Betrag und Phase berechnen.<br />

5 Ergebnisse<br />

Um die Genauigkeit der Lösung zu bewerten, wurden die Ergebnisse mit einer analytischen<br />

Lösung für eine Kugelelektrode verglichen, die von Hördt (2007) vorgestellt wurde.<br />

Abbildung 4 zeigt die Geometrie und die Parametrisierung. Für die mittlere Schale mit Radius<br />

r1 wird hier Vakuum gewählt werden, mit 1=0 und 1=0 (Permitticität des Vakuums).<br />

Abbildung 4: Geometrie der Kugelelektrode. Eine ideal leitende Elektrode mit Potential V<br />

und Radius r0 ist in eine Kugelschale mit Radius r1 eingebettet.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

110


Die Impedanz der Elektrode ergibt sich dann zu:<br />

* *<br />

2 2 r1<br />

1<br />

* *<br />

1<br />

<br />

1 1 r0<br />

Z<br />

* *<br />

4<br />

r0<br />

1 2 r1<br />

*<br />

1 r0<br />

wobei<br />

*<br />

i<br />

<br />

<br />

<br />

<br />

<br />

<br />

In Abbildung 5 wird der Betrag der Impedanz der Scheibenelektrode mit der einer<br />

Kugelelektrode verglichen. Die Radien wurden so gewählt, dass die Oberfläche der Kugel<br />

gleich der einseitigen Fläche der Scheibe ist. Bei der Kugelelektrode wurde für die mittlere<br />

Schale die Werte für Vakuum eingesetzt, der Abstand entspricht der Dicke der mittleren<br />

Schale.<br />

Für mittlere Abstände stimmen die Impedanzen der Kugelelektrode und der<br />

Scheibenelektrode überein, bei großen und kleinen Abständen sind die Impedanzen<br />

unterschiedlich, was auch so zu erwarten ist. Der Grenzwert großer Abstände entspricht der<br />

Impedanz der Elektrode im Vakuum. Die Impedanzen beider Elektroden in Abb. 5<br />

konvergieren gegen den jeweiligen Grenzwert.<br />

Abbildung 5: Impedanzbetrag der kapazitiven Elektrode gegen Abstand der Elektrode zum<br />

Medium für verschiedene Leitfähigkeiten des Mediums. Durchgezogene Linie:<br />

Kugelelektrode mit Radius 0.05m. Gestrichelte Linie: Scheibenelektrode mit Radius 0.1m.<br />

Die Frequenz ist f=10 kHz, die relative Permittivität des Untergrundes ist in beiden Fällen<br />

r=4.<br />

Für die Kugel mit Radius r0 beträgt dieser (Smythe, 1968):<br />

(42)<br />

(43)<br />

1<br />

Z <br />

4<br />

i<br />

0 r0<br />

(44)<br />

und für eine Scheibe mit Radius a:<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

111


1<br />

Z (45)<br />

8i<br />

0 a<br />

Wenn a=2r0, wie für die Rechnung aus Abb. 5 der Fall, entspricht die Gesamtfläche der<br />

Kugel der einseitigen Fläche der Scheibe. Dann ist die Vakuumimpedanz der Scheibe etwas<br />

kleiner, weil im Vakuum beide Seiten der Scheibe gleich wirksam sind.<br />

Für kleine Abstände ist der Vergleich von der Leitfähigkeit des Untergrundes abhängig. Bei<br />

moderaten Leitfähigkeiten (rote und grüne Kurve) konvergiert der Impedanzbetrag gegen den<br />

Grenzwert der galvanischen Ankopplung, also den Fall, dass die Scheibe aufliegt, bzw. die<br />

Kugel direkten Kontakt zum leitenden Medium hat. Der Grenzwert ist für die Kugel<br />

1<br />

Z (46)<br />

4<br />

r0<br />

und für die kreisförmige Scheibe<br />

1<br />

Z (47)<br />

4<br />

a<br />

wobei die Gleichstromleitfähigkeit ist.<br />

In diesem Fall liefert die Scheibe bei gleicher Oberfläche (mit a=2r0) die etwas größere<br />

Impedanz. Eine Gleichheit ist hier nicht zu erwarten, da nicht nur die Fläche wichtig ist,<br />

sondern auch die Stromverteilung im Medium. Bei sehr hohen Leitfähigkeiten in der<br />

Größenordnung von Metallen (blaue Kurve) wird der Grenzwert auch für extrem kleine<br />

Abstände noch nicht erreicht. Bei =0 (schwarze Kurve) sind nur die Verschiebungsströme<br />

(mit r=4) zur Ankopplung wirksam. Die Scheibe hat eine etwas höhere Impedanz verglichen<br />

mit der Kugel.<br />

Insgesamt verhalten sich die Kurven wie erwartet. Im mittleren Abstandsbereich ist nur die<br />

Ankopplungsfläche relevant und die Impedanzen von Kugel und Scheibe stimmen überein.<br />

Bei sehr großen und sehr kleinen Abständen werden jeweils die analytischen Grenzwerte<br />

erreicht, die für Kugel und Scheibe unterschiedlich sind.<br />

Der Verlauf der Impedanz in Abbildung 5 hat zwei wichtige Konsequenzen. Zum Einen<br />

nimmt die Impedanz stetig als Funktion der Höhe zu. Dies schließt den Grenzwert einer<br />

galvanisch gekoppelten Elektrode ein. Das bedeutet, dass eine kapazitiv gekoppelte Elektrode<br />

niemals eine geringere Impedanz haben kann, als eine galvanisch gekoppelte, aufliegende.<br />

Zum Anderen ist der Übergang von der aufliegenden zur abgehobenen Elektrode stetig. Dies<br />

könnte der Intuition widersprechen; man würde evtl. einen Sprung der Impedanz zwischen<br />

aufliegender und abgehobener Impedanz erwarten.<br />

Um das beobachtete Verhalten zu verstehen, ist in Abb. 6 die Impedanz in Real-und<br />

Imaginärteil aufgespalten dargestellt, mit anderem Abstandsbereich. Der Realteil (Mitte) der<br />

Impedanz ist für sehr große und sehr kleine Leitfähigkeiten verschwindend gering bzw.<br />

identisch null und wird nicht mit dargestellt. Bei moderaten Leitfähigkeiten (grüne und rote<br />

Kurve) wird die Impedanz durch den Realteil dominiert. Bei extrem niedriger und bei extrem<br />

hoher Leitfähigkeit überwiegt der Imaginärteil.<br />

Dieses Verhalten lässt sich verstehen, wenn man die Gesamtimpedanz näherungsweise als<br />

Reihenschaltung zwischen einem kapazitiven Anteil und einem Erdungswiderstand betrachtet.<br />

Der kapazitive Anteil Zc entspricht der Überwindung des Raumes zwischen Boden und Luft<br />

und ist rein imaginär, der Erdungswiderstand Ze ist komplex und entspricht dem<br />

Ankopplungswiderstand der aufliegenden Platte. Man kann daher näherungsweise schreiben:<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

112


d<br />

1<br />

Z Z c Z e <br />

(48)<br />

2<br />

i<br />

0<br />

a 4ai01<br />

Zunächst einmal erklärt das Modell, warum der Übergang zwischen aufliegender und<br />

abgehobener Platte stetig ist und kein Sprung auftritt. Wenn der Abstand d0, beginnt der 2.<br />

Term zu dominieren und „übernimmt“ stetig die Ankopplung. Allerdings ist in der Abbildung<br />

zu erkennen, dass der Übergang bei extrem kleinen Abständen erfolgt, die sogar unterhalb<br />

eines Atomdurchmessers liegen können. In der Praxis wird der stetige Übergang in der Regel<br />

also nicht zu sehen sein, sondern als Sprung erscheinen.<br />

Abbildung 6: Betrag der Impedanz (oben), Realteil (mitte) und Imaginärteil (unten) für die<br />

Scheibe für dasselbe Modell wie in Abb. 5. Der Realteil der Impedanz (mitte) für =0 ist<br />

identlisch null und nicht darstellbar, der Realteil für die sehr gut leitende Platte ist extrem<br />

klein und wird nicht dargestellt.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

113


Das Modell aus Gl. (48) erklärt auch die Zerlegung der Impedanz in Real-und Imaginärteil.<br />

Bei sehr kleiner Leitfähigkeit und kleinem Abstand dominiert der Imaginärteil der Impedanz,<br />

bei moderater und hoher Leitfähigkeit dominiert der Realteil. Bei sehr hoher Leitfähigkeit<br />

wird der Abstand, in dem der 1.Term vernachlässigbar ist, in der Abbildung nicht erreicht und<br />

es dominiert immer noch der rein imaginäre 1. Term die Gesamtimpedanz.<br />

6 Schlussfolgerungen<br />

Es wurde ein Programm entwickelt, um die Impedanz einer kreisförmigen Scheibe über einem<br />

homogenen Halbraum zu berechnen. Das Verfahren beruht auf einer vollständigen Lösung<br />

der Maxwellgleichungen. Es verwendet eine Näherung, bei der angemommen wird, dass der<br />

Radius der Scheibe klein ist gegen die elektromagnetische Wellenlänge. Die Näherung<br />

verliert bei sehr hohen Frequenzen ihre Gültigkeit, ist jedoch beispielsweise für einen Radius<br />

0.1m und eine Frequenz von 1 MHz vollkommen unproblematisch. Das Programm wurde<br />

durch Vergleich mit einer analytischen Lösung für eine Kugelelektrode und durch analytische<br />

Grenzwerte für große und kleine Abstände verifizert.<br />

Der Übergang zur Impedanz einer aufliegenden Elektrode erfolgt stetig als Funktion des<br />

Abstandes zwischen Elektrode und Halbraum. Die Impedanz der aufliegenden Scheibe ist<br />

immer kleiner als die der angehobenen. Dies bedeutet, dass die Übergangsimpedanz nicht<br />

durch eine kapazitive Elektrode gegenüber einer gleich gebauten, galvanischen verringert<br />

werden kann. Der galvanische Anteil der Impedanz ist auch bei kapazitiver Ankopplung mit<br />

enthalten.<br />

In der Praxis kann es dennoch sinnvoll sein, kapazitive Elektroden statt galvanische zu<br />

verwenden, also den galvanischen Kontakt durch Isolierung oder Anheben zu verhindern. Bei<br />

Aufliegenden Elektroden ist die Impedanz sehr schwer zu kontrollieren, da die Auflagefläche<br />

stark variieren kann. Dies kann dazu führen, dass die Impedanz um Größenordnungen<br />

schwankt, da sie sehr stark als Funktion des Abstandes variiert und die aufliegenden Teile die<br />

Ankopplung dominieren können. Eine Isolierung stabilisiert die Impedanz und verhindert<br />

starke Schwankungen, was für die Messung von Vorteil sein kann.<br />

7 Referenzen<br />

Hördt, A., 2007, Contact impedance of grounded and capacitive electrodes, in: Ritter, O.,<br />

Brasse, H., Protokoll zum 22. Kolloquium „Elektromagnetische Tiefenforschung“, 164-173.<br />

Kuras, O., Beamish, D., Meldrum, P.I., and Ogilvy, R.D., 2006, Fundamentals of the<br />

capacitive resistivity technique. Geophysics, 71, G135-G152.<br />

Smythe, W., 1968. Static and dynamic electricity, McGRaw-Hill.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

114


H2O <br />

<br />

<br />

<br />

<br />

<br />

<br />

H2O <br />

<br />

H2O <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

H2O <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

H2O <br />

<br />

<br />

<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

115


5mm <br />

<br />

<br />

<br />

<br />

<br />

<br />

±0.5mm <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

116


EG EB<br />

EN EB <br />

dx =50m<br />

<br />

By<br />

<br />

<br />

<br />

<br />

<br />

(EG)<br />

(EN) <br />

<br />

<br />

(EB) <br />

<br />

<br />

EGel <br />

Enorm <br />

<br />

(BY ) <br />

<br />

EGel Enorm BY <br />

<br />

<br />

<br />

16 2/3Hz 50Hz <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

0 − 250Hz <br />

<br />

<br />

<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

117


EGel Enorm EnormEGel <br />

EGel/BY <br />

Enorm/BY EGel Enorm<br />

<br />

0 − 250Hz <br />

EGel Enorm <br />

<br />

Txy = <br />

< |Y | ><br />

X, Y EGel, Enorm, BY <br />

EGel = TxyEnorm <br />

Enorm = TxyEGel <br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

118


204Hz<br />

8Hz <br />

<br />

<br />

EGel = TxyBy Enorm =<br />

TxyBy <br />

ρa <br />

EGel/BY Enorm/BY <br />

<br />

204Hz 1Hz <br />

<br />

<br />

<br />

H2O<br />

<br />

<br />

<br />

<br />

<br />

• <br />

• <br />

<br />

<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

119


Attitude Algorithm Utilised in Mobile Geophysical<br />

Measuring Systems<br />

Johannes B. Stoll, Celle<br />

Christopher Virgil, Institut für Geophysik und extraterrestrische Physik, TU-Braunschweig<br />

1 Introduction<br />

Accurate real-time tracking of orientation or attitude of rigid bodies has wide applications in robotics,<br />

aerospace, on-land and underwater vehicles, automotive industry, and virtual reality. A quite recent<br />

application in geophysics is the analysis of the attitude effect of geophysical sensors such as coil<br />

arrays or magnetometers attached to mobile measuring platforms e.g. helicopters and fixed wing<br />

airplanes and the correction with respect to the Earth’s reference coordinate frame (Reid etal, 2003,<br />

Yin and Fraser, 2004). Another field of application is wireline logging with a borehole tool. While<br />

measuring, the tool spins about its vertical body axis due to the torsion force of the logging cable,<br />

which results in a disorientation of geophysical sensors.<br />

In the borehole industry, it is essential to accurately monitor and guide the direction of the drill bit. It is<br />

also necessary for an oil rig to log the location of its boreholes at a regular frequency such that the oil<br />

rig can be properly monitored. To determine the location of a drill bit in a borehole, it is necessary to<br />

know the position and the attitude, which includes the vertical orientation and the North direction.<br />

In prior art systems, a magnetometer is used to determine the magnetic field direction from which the<br />

direction of North is approximated. Triple component magnetometers were widely used to sense the<br />

tool rotation during a log run. However orientation via magnetometers is subject to restrictions. The<br />

employing of magnetometers in boreholes drilled into strongly magnetized formations such as oceanic<br />

crust, volcanoes and ore deposits, make it very difficult to utilize the Earth’s magnetic field as a<br />

reference (Steveling et al, 2003). First, these systems must make corrections for magnetic<br />

interference and use of magnetic materials for the drill pipe. Second, systems that rely only on<br />

magnetometers to determine North can suffer accuracy degradation due to the Earth's magnetic field<br />

variations. Third, a magnetometer alone cannot give an unequivocal measurement of a set of sensor<br />

attitude. Measurements made with such a sensor define the angle between the Earth’s magnetic field<br />

and a particular axis of the sensor. However, this axis can lie anywhere on the surface of a cone of<br />

semi-angle equal to that angle about the magnetic vector. Hence, an additional measurement is<br />

required to determine attitude with respect to another fixed reference frame.<br />

In order to determine the direction of a spinning borehole tool another type of sensor technology<br />

must be utilised to provide orientation that is independent from the Earth’s magnetic field as a<br />

reference (Galliot et al, 2004, Stoll and Leven, 2002).<br />

The problem to be solved consists of a way to sense the attitude of a borehole tool, which are free to<br />

rotate about any direction. A good compromise between accuracy and sensor size are the fibre optical<br />

gyros. These devices measure the turning rate about the sensor axis in the tool frame. In this paper an<br />

algorithm is presented that approaches the attitude determination based on this discrete data.<br />

2 Principle of Fibre Optical Gyro (FOG)<br />

Gyroscopes are used in various applications to sense the angular rate of turn about some defined<br />

axis. The most basic and the original form of gyroscopes make use of the inertial properties of a<br />

wheel, or rotor, spinning at high speed. The wheel tends to maintain the direction of its spinning axis in<br />

space due to conservation of the angular momentum vector. These devices are susceptible to<br />

damage from shock and vibration, exhibit cross-axis acceleration sensitivity and, for the lower cost<br />

versions, have reliability problems.<br />

Another type is the vibratory gyroscope. The basic principle of operation of such sensors is that the<br />

vibratory motion of part of the instrument creates an oscillatory linear velocity. If the sensor is rotated<br />

about an axis orthogonal to this velocity, a Coriolis acceleration is induced. The acceleration modifies<br />

the motion of the vibrating element which is an indication of the magnitude of the applied rotation.<br />

However, this type of sensor tends to produce biases in the region of 1°/s.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

120


Optical gyroscopes use an interferometer to sense angular motion. An optical gyroscope, laser or<br />

fibre, measures the interference pattern generated by two light beams, traveling in opposite directions<br />

within a mirrored ring or fibre loop, in order to detect very small changes in motion. This type of<br />

gyroscope can be subdivided into fibre optical gyro (FOG), ring laser gyro (RLG), and ring resonator<br />

gyro (PARR).<br />

In order to keep track on the orientation of the geophysical sensors the utilisation of fibre optical<br />

gyroscopes are now common in navigation and replace prior art systems like mechanical gyros.<br />

Important attributes of this new technology are no moving parts, high reliability, stable performance<br />

and low costs. Many of its components are based on proven technology from the fiber optical<br />

telecommunications industry. Using optical gyros, inertial navigation is accomplished by integrating the<br />

output of a set of angular rate sensors to compute the attitude.<br />

In figure 1 a FOG manufactured by LITEF is shown exemplary. The gyro operates at an optical<br />

wavelength of 820 nanometers, with a 110 meter coil of elliptical-core polarization maintaining fiber.<br />

The low coherence reduces unwanted interference between waves reflected from the fusion splices<br />

used to join the components.<br />

Fig. 1: Example of the FORS -family (LITEF, Freiburg) (here FORS-6U)<br />

3 Attitude Computation<br />

Inertial orientation tracking of borehole tools is based upon the same methods and algorithm as<br />

those used for aircrafts, ships, and missiles. There is a large quantity of technical literature that<br />

describes the fundamentals of inertial navigation technology in great detail, e.g., Grewal et al. (2001),<br />

v. Hinüber (1993), Savage (1998a, b), Stovall (1997). If angular rates of a tool are measured<br />

constantly with depth, the orientation of the sensors of the tool with respect to inertial space can be<br />

determined by applying a suitable coordinate transformation.<br />

A transformation from one coordinate frame to another can be carried out as three successive<br />

rotations about different axes. The transformation matrix is given by the product of these three<br />

separate transformations as follows:<br />

C n<br />

b<br />

C C<br />

C<br />

(1)<br />

3<br />

Various mathematical representations can be used to define the attitude of a body with respect to a<br />

coordinate reference frame. One of the most common ways of parameterizing the transformation<br />

matrix is by use of direction cosine matrix (DCM). Other methods of describing rotations are the use of<br />

Euler Angles and Quaternions.<br />

The DCM is a 3x3 matrix, the columns of which represent unit vectors in body axes projected along<br />

n<br />

the reference axes. Matrix Cb<br />

describes the transformation from coordinate frame “b” to the frame “n”<br />

and is written here in component form as follows:<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

121<br />

2<br />

1


c<br />

<br />

c<br />

<br />

c<br />

31<br />

c<br />

32<br />

33<br />

<br />

<br />

<br />

<br />

<br />

11 12 13<br />

C 21 c22<br />

c23<br />

n<br />

b (2)<br />

c<br />

The element in the i th row and the j th column represent the cosine of the angle between the i th axis of<br />

the reference frame “n” and the j th axis of the body frame “b”.<br />

When designing a navigation system it is necessary to relate the information from the sensors to a<br />

navigation coordinate. The coordinate choice for borehole applications is a geographic or navigation<br />

frame “n” with axes {N, E, D}, (North, East and Down). The inertial measurement unit is mounted on<br />

the borehole tool constituting a new frame. This is called the body frame “b”, and has axes {X, Y, Z}.<br />

This frame will be in rotation with respect to the geographic frame. The velocity of this rotation is<br />

measured by three orthogonal gyros. The transformation matrix that relates “b” and “n” coordinate<br />

frames propagates with time in accordance with the following equation:<br />

0 <br />

z y<br />

<br />

<br />

<br />

n n b<br />

C b<br />

b Cb<br />

nb<br />

, where nb<br />

z<br />

0 <br />

x (3)<br />

<br />

<br />

<br />

y<br />

x<br />

0 <br />

b<br />

n<br />

nb is the skew symmetric matrix formed from the elements of the vector b x y z , which<br />

th th<br />

represents the turn rate of the body between the i -frame and the (i+1) -frame as measured by the<br />

gyroscopes in the body frame. In real time applications the integration is implemented with the<br />

following approximation<br />

t<br />

C<br />

n<br />

b<br />

c<br />

c<br />

n t t<br />

C t A<br />

t<br />

(4)<br />

b<br />

T , <br />

,<br />

A t<br />

A t<br />

where is the DCM which relates the b-frame at time t to the b-frame at time t . For small<br />

angle rotations, may be written as follows:<br />

<br />

where I is the 3×3 identity matrix and<br />

0<br />

<br />

<br />

<br />

<br />

<br />

t I A (5)<br />

<br />

0<br />

<br />

<br />

<br />

<br />

0 <br />

<br />

in which , and are small rotation angles through which the body-frame has rotated over the<br />

time interval t about its yaw, pitch and roll axes respectively. If the limit t approaches zero, small<br />

angle approximations are valid and the order of the rotation becomes unimportant.<br />

Hence, the minimum sampling time of the gyros is important for obtaining the transformation matrix<br />

with reasonable accuracy. This interval will be a function of the severity of manoeuvres expected from<br />

the borehole tool. In many applications the rotation velocity expected is usually less than 25<br />

degrees/sec. With a sampling time of 100 Hz the maximum angle variation will be less than 0.25<br />

degree satisfying the small angle approximation. Rotations with faster dynamics will require smaller<br />

sampling time to compute the transformation matrix appropriately.<br />

n<br />

In order to update the DCM Cb<br />

, it is necessary to solve a matrix differential equation (3). Over a<br />

single computer cycle, from time tk<br />

to tk+1, the solution of the equation may be written as:<br />

k 1<br />

(6)<br />

C C exp dt (7)<br />

k 1<br />

k<br />

Provided that the orientation of the turn rate vector remains fixed in space over the update interval,<br />

we may define:<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

122<br />

t<br />

t<br />

k


and<br />

t<br />

k 1<br />

t<br />

k<br />

dt <br />

(8)<br />

CkAk C C exp <br />

k 1<br />

k<br />

(9)<br />

Where Ck represents the direction cosine matrix which relates body to reference axes at the k th<br />

computer cycle, and Ak the direction cosine matrix which transforms a vector from body co-ordinates<br />

at the k th computer cycle to body co-ordinates at the (k+1) th computer cycle.<br />

The variable is an angle vector with direction and magnitude such that a rotation of the body frame<br />

about through an angle equal to the magnitude of will rotate the body frame from its orientation at<br />

computer cycle k to its position at the computer cycle k+1. The components of are denoted by x,<br />

y, and z and its magnitude given by:<br />

and<br />

<br />

<br />

(10)<br />

2<br />

x<br />

2<br />

y<br />

2<br />

z<br />

0 <br />

z y <br />

<br />

<br />

z 0 <br />

x (11)<br />

<br />

<br />

<br />

y x 0 <br />

Expanding the matrix exponential function in a power series gives:<br />

where<br />

Thus we may write<br />

A<br />

k<br />

2<br />

3<br />

4<br />

<br />

Ak I <br />

(12)<br />

2!<br />

3!<br />

4!<br />

2<br />

2 2 <br />

y <br />

x<br />

<br />

x<br />

y<br />

z<br />

and which may be written as follows:<br />

<br />

<br />

z<br />

<br />

<br />

2 2 <br />

x<br />

x<br />

y<br />

<br />

<br />

3 2 <br />

<br />

2 2<br />

4<br />

<br />

<br />

y<br />

z<br />

z<br />

<br />

<br />

<br />

x<br />

<br />

<br />

<br />

2 2<br />

<br />

(14)<br />

(15)<br />

I <br />

2!<br />

<br />

2 4 <br />

I <br />

1<br />

...<br />

<br />

3!<br />

5!<br />

<br />

<br />

<br />

<br />

2 2<br />

3!<br />

2<br />

<br />

1<br />

<br />

2!<br />

2<br />

<br />

4!<br />

x<br />

y<br />

z<br />

z<br />

4!<br />

4<br />

<br />

6!<br />

y<br />

<br />

(13)<br />

2 <br />

...<br />

<br />

<br />

1cos 2 <br />

sin <br />

Ak I <br />

(17)<br />

2<br />

<br />

Equation (17) is input in equation (9) that updates the frame at time t to the frame at time t+t. To<br />

compute the attitude of the tool, first the measured rotation rates are input in equation 17. Then,<br />

beginning with C0 = I the rotation matrices Ck+1 for each computer cycle k+1 are successively<br />

computed by equation (9). The tool attitude r in the external frame at the k th cycle results to<br />

n<br />

k<br />

2<br />

...<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

123<br />

<br />

<br />

<br />

(16)


n<br />

with the starting orientation .<br />

r 0<br />

<br />

<br />

(18)<br />

n<br />

n<br />

rk Ck<br />

r0<br />

4 Attitude Computation allied to the Göttingen Borehole Magnetometer (GBM)<br />

The Göttingen Borehole Magnetometer (GBM) contains a triple of FORS-36M sensors, which are<br />

attached rigidly to the borehole tool. The rotation axes of the sensors are aligned with the body axes of<br />

the tool, and thereby sense the rotation rate about the vertical axis and the tilting movement about the<br />

two horizontal axes. Angular rates are output with a sample rate of 1 second. The salient<br />

specifications of the FORS-36m are the maximum input rotation rate of ±720°/s that allows sensing a<br />

maximum rotation rate of 2 revolutions per second with a resolution of 9·10 -5 °. Moreover, the<br />

dimensions of the sensor (53mm x 58mm x 19mm) are small enough to fit to the inner diameter of<br />

65mm of the GBM. This is important to reduce the overall diameter of the tool and make it applicable<br />

in narrower boreholes. The bias drift is 6/h (1) and is reduced to


Steveling, E., J.B. Stoll, and M. Leven, 2003. Paleomagnetic age dating from magnetisation and<br />

inclination studies of the Mauna Kea/Hawaii, Geochem.Geophys. Geosystems<br />

Stoll, J.B., and M. Leven, 2002. Results of the oriented magnetic logging at Detroit Seamount. In:<br />

Tarduno, J.A., Duncan, R.A., Scholl, D.W., Proc. ODP, Initial Reports, Vol. 197. Available from<br />

http://www-odp.tamu.edu/publications/197_IR/197ir.htm.<br />

Stovall, S.H., 1997. Basic Inertial Navigation, Naval Air Warefare Center Weapons Division,<br />

http://www.fas.org/spp/military/program/nav/basicnav.pdf<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

125


Joint Interpretation of Magnetotelluric and Seismic Models for Exploration of<br />

the Gross Schoenebeck Geothermal Site<br />

G. Muñoz, K. Bauer, I. Moeck, O. Ritter<br />

<strong>GFZ</strong> <strong>Deutsche</strong>s GeoForschungsZentrum<br />

Introduction<br />

Due to the non-uniqueness of the inverse problem, the interpretation of geophysical models in<br />

terms of geological units is not always straightforward. It is common to use a combination of<br />

different geophysical methods to obtain the distribution of independent physical properties<br />

over the area of interest in order to discriminate between the different lithologies or geologic<br />

units. This kind of studies is usually limited to qualitative comparisons of the different<br />

models, which may – or may not –support a relation between the parameters in certain areas.<br />

Quantitative approaches are in general based on empirical relations between physical<br />

parameters, which often are not of universal applicability.<br />

The magnetotelluric (MT) and seismic methods resolve the physical parameters electric<br />

resistivity (ρ) and (seismic) velocity (Vp, Vs) respectively with similar spatial resolution and<br />

are often used in combination to derive earth models. By looking at both resistivity and<br />

velocity simultaneously, we can keep the strengths of both methods while avoiding their<br />

weaknesses. The problem with a joint interpretation is that there is no unique universal law<br />

linking electrical and acoustic properties. While electrical resistivity in deep sedimentary<br />

basins is mostly sensitive to the pore geometry and contents , seismic velocity is mostly<br />

imaging rock matrix properties. However, with a statistical analysis of the distributions of<br />

both resistivity and velocity, we can find certain areas of the models space where a particular<br />

relation between the physical parameters holds locally, thus allowing us to characterize this<br />

region as a particular lithology. In the present work, we use a statistical analysis, as described<br />

by Bedrosian et al. (2007) in order to correlate two independently obtained models of the<br />

Groß Schönebeck geothermal test site in the Northeast German Basin.<br />

Methodology description<br />

The methodology used in the present paper was described by Bedrosian et al. (2007) and is<br />

based on a probabilistic approach developed by Bosch (1999), in a sense that diverse<br />

geophysical parameters are represented as a probability density function (pdf) in the joint<br />

parameter space. The coincident velocity and resistivity models are first interpolated onto a<br />

common grid. Therefore, a joint parameter space is built, where each point in the modelled<br />

area is associated with a velocity – resistivity pair. By plotting one parameter against the other<br />

in a cross-plot and including the error estimates we can then construct a joint pdf in the<br />

parameter space. The areas of enhanced probability can be identified with classes represented<br />

by a certain range of values in both resistivity and velocity. By mapping back these classes<br />

onto the spatial domain they can be related to certain lithologies and/or geological units.<br />

In the present work, the resistivity model was interpolated onto the seismic mesh, given that it<br />

is uniformly spaced and finer than the magnetotelluric mesh. An inverse distance weighted<br />

interpolation scheme was used, which forms estimates from a weighted average of many<br />

samples found within a pre-defined area around the point, with decreasing weights with<br />

distance.<br />

Each element of this distribution can be interpreted as the outcome of a process defined by a<br />

probability density function (pdf). Assuming normal error distribution and independence of<br />

the data, the joint pdf is expressed as the sum of the individual pdfs (pdfi) for each data point,<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

126


according to the following expression, with the errors of the resistivity and velocity (δlog(ρi)<br />

and δVp,i) estimated from the sensitivity matrix and the hit count distribution respectively:<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

The different classes are identified as zones of enhanced probability in the joint pdf crossplot.<br />

Assuming that the geological units are characterized by uniform physical properties<br />

normally distributed, each class is defined by a mean point and a covariance matrix<br />

(representing the ~60 per cent confidence interval ellipse for the peak) in the joint parameter<br />

space.<br />

Geophysical models<br />

The Groß Schönebeck low enthalpy geothermal site, with the well doublet GrSk 3/90 and<br />

GrSk 4/05, is located in the Northeast German Basin (NEGB). MT data was collected along a<br />

40 km-long profile centred on the well doublet. The profile consists of 55 stations with a site<br />

spacing of 400 m in the central part (close to the borehole) of the profile, increasing to 800 m<br />

towards both profile ends. The period range of the observations was 0,001 to 1000 s. This<br />

profile is spatially coincident with the seismic tomography profile and most of the stations<br />

were located at the same places as the seismic shot points. At all sites, we recorded horizontal<br />

electric and magnetic field components and the vertical magnetic field.<br />

The resistivity model for the MT profile (Figure 1a) shows a shallow conductive layer<br />

extending from the surface down to depths of about 4 km, with an antiform-type shape below<br />

the central part of the profile. At a depth range of 4-5 km two conductive bodies are found,<br />

separated by a region of moderate conductivity. According to the seismic tomography, which<br />

shows high velocity values for depths greater than 4 km, a resistive basement was introduced<br />

a priori in the resistivity model (Muñoz et al., 2010).<br />

A 40 km long seismic profile was measured coincident with the MT experiment (Figure 1b).<br />

The objective was to derive a regional 2-D seismic model, which can be combined with the<br />

electrical conductivity model from the MT data analysis to study the potential reservoir layers<br />

and overlying sediments. The experimental setup was designed to provide data suitable for<br />

refraction tomography. 45 explosion shots were fired from 20 m deep boreholes with charge<br />

sizes of 30 kg. The shot spacing was 800 m on average. The recording instrumentation<br />

consisted of 4.5 Hz 3-component geophones. These were deployed as a 40 km long receiver<br />

spread with spacings of 200 m. Each shot was recorded by all receivers.<br />

The velocity model (Figure 1b) can be divided into three major sequences: The upper section<br />

(depth range 0-2 km) is characterized by low velocities (2-3.5 km/s) and a strong increase of<br />

velocity with depth. The section between 2 and 4 km depth shows velocities between 4 and<br />

4.5 km/s and is characterized by a strong topography on top which is related with salt<br />

mobility. The third, deepest section is bounded on top by a subhorizontal interface at 4.2 km<br />

depth and reveals velocities of more than 5 km/s (Bauer et al., 2010).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

127


a)<br />

b)<br />

Figure 1: Electrical resistivity model obtained from inversion of the magnetotelluric data<br />

using a priori information from the seismic velocity model for the deeper part of the model (><br />

5 km) (a) and seismic velocity model obtained from (Vp) travel time inversion (b).<br />

Joint analysis<br />

In the cross-plot of the probability density function (Figure 2a) we can identify five more or<br />

less clear peaks, or areas of enhanced probability with respect to the neighbouring region. The<br />

coloured ellipses in Figure 2a represent the ~60 per cent confidence intervals of the Gaussian<br />

peaks best fitting the pdf. The clusters were mapped back to the cross section providing a<br />

depth distribution of the classes along the profile (Fig. 2c). In order to interpret the nature of<br />

these litho-types, the model is superimposed on the stratigraphy derived from pre-existing<br />

reflection seismic data and borehole information (Moeck et al., 2008).<br />

Class 1, the shallowest, is characterised by low velocity (1.8 – 2.7 km/s) and moderate<br />

resistivity (5 – 70 Ωm) and comprises of unconsolidated sediments. Class 2, with higher<br />

velocities (2.7 – 3.9 km/s) and lower resistivities (0.5 – 3.5 Ωm) encompasses weak or soft<br />

rocks with high porosity, which are more conductive because they provide storage for a<br />

greater volumes of fluids. Class 3 coincides with successions of Middle Triassic to Lower<br />

Permian. Significantly these successions represent harder brittle rock of limestone and<br />

sandstone as indicated by increasingly higher velocities (4 – 5 km/s) and resistivities (2 – 15<br />

Ωm). This class includes also thick salt rock layers (Zechstein) which, however, yield no<br />

significant variation of resistivity or velocity within the class. Class 4, the deepest one, is<br />

characterized by the highest velocity and resistivity values (4.7 – 5.5 km/s and around 3000 -<br />

30000 Ωm). It represents the basin floor, comprising volcanic rock, quartzite and slate. This<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

128


class reflects the fact that the magnetotelluric model used seismic a priori information to<br />

introduce a resistive basin floor. Obviously this causes the very high correlation between<br />

velocity and resistivity of class 4 in Fig. 2b. Class 5 is characterized by high velocities (4.7 –<br />

5.5 km/s) but extremely low resistivities (0.1 – 0.7 Ωm).<br />

Figure 2: Cross-plot of the probability density function in gray-scale plan view (a) and threedimensional<br />

view (b). Spatial distribution of classes (c). Colours correspond with the ellipses<br />

in (a) defining the class boundaries.<br />

It is remarkable that Class 5 is restricted to salt lows where presumably anhydrites of Upper<br />

Permian age remain after salt movement. These anhydrites have a brittle behaviour and are<br />

expected to be highly fractured. In this case, an estimation of the resistivity by using Archie’s<br />

Law (Archie, 1942) for fracture-controlled porosity and assuming a formation fluid salinity of<br />

260 g/l (Giese et al., 2001), and a reasonable range of porosities and temperatures (15% and<br />

130ºC) the modelled resistivities of 0.1 – 0.7 Ωm can be explained. High velocities can be<br />

explained by the high density of anhydrite (2.9 g/cm 3 ). The classes are summarized in the<br />

Table 1, below.<br />

Table 1: Resistivity and P velocity of the classes from figure 9. Also included are lithology<br />

and stratigraphic units.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

129


Conclusions<br />

MT and seismic data were used to derive independent 2-D models of the electrical<br />

conductivity and the seismic P velocity around the geothermal research well GrSk 3/90. The<br />

resulting models were combined in a statistical analysis to determine correlating features in<br />

both models. The classification method used in this analysis revealed 5 distinct litho-types<br />

which show up as separate clusters in the underlying geophysical parameter space of Pa and<br />

VP.<br />

This study demonstrates the concert of MT, seismic and structural geologic models. Therefore<br />

the combination of different geophysical methods (MT and seismic) combined with structural<br />

geological information revealed information which was previously unknown and which could<br />

not be determined with the individual methods. Clearly, this approach is applicable in other<br />

areas too and could represent a promising new approach for geothermal exploration.<br />

Acknowledgements<br />

This work was funded within the 6 th Framework Program of the European Union (I-GET<br />

Project, Contract nº 518378). The instruments for the geophysical experiments were provided<br />

by the Geophysical Instrument Pool Potsdam (GIPP). We wish to thank Paul Bedrosian for<br />

making his statistical analysis code available to us.<br />

References<br />

Archie, G. [1942] The electrical resistivity log as an aid in determining some reservoir<br />

characteristics. Trans. Am. Inst. Min. Metall. Pet. Eng., 146, 54–62.<br />

Bauer, K., Moeck, I., Norden, B., Schulze, A., Weber, M. [2010] . Tomographic P velocity<br />

and gradient structure across the geothermal site Gross Schoenebeck (NE German Basin):<br />

Relationship to lithology, salt tectonics, and thermal regime. J. Geophys. Res., Submitted.<br />

Bedrosian, P.A., Maercklin, N., Weckmann, U., Bartov, Y., Ryberg, T., Ritter, O. [2007]<br />

Lithology-derived structure classification from the joint interpretation of magnetotelluric and<br />

seismic models. Geophysical Journal International, 170 170, 170<br />

737-748.<br />

Bosch, M. [1999] Lithologic tomography: from plural geophysical data to lithology<br />

estimation. Journal of Geophysical Research, 104, 749-766.<br />

Giese, L., Seibt A., Wiersberg T., Zimmer M., Erzinger J., Niedermann S. and Pekdeger A.<br />

[2001] Geochemistry of the formation fluids, in: 7. Report der Geothermie Projekte, In situ-<br />

Geothermielabor Groß Schönebeck 2000/2001 Bohrarbeiten, Bohrlochmessungen, Hydraulik,<br />

Formationsfluide, Tonminerale. GeoForschunsZentrum Potsdam.<br />

Moeck, I., Schandelmeier, H., Holl, H.G. [2008] The stress regime in Rotliegend reservoir<br />

reservoir of the Northeast German Basin. International Journal of Earth Sciences (Geol.<br />

Rundsch.), doi:10.1007/s00531-008-0316-1.<br />

Muñoz. G., Ritter, O., Moeck, I. [2010] A target-oriented magnetotelluric inversion scheme<br />

for characterizing the low enthalpy Groß Schönebeck geothermal reservoir. Geophysical<br />

Journal International, submitted.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

130


Magnetotelluric measurements to explore for deeper structures of the Tendaho<br />

geothermal field, Afar, NE Ethiopia<br />

Ulrich Kalberkamp<br />

Bundesanstalt für Geowissenschaften und Rohstoffe (BGR), Hannover, Germany<br />

Email: ulrich.kalberkamp@bgr.de<br />

Introduction<br />

In regions with high heat flow, like at volcanically active plate margins, high total<br />

thermodynamic energy is accumulated in the so called high enthalpy resources. The East<br />

African Rift System is one of the privileged areas to yield numerous sites for potential<br />

geothermal energy extraction including power generation. Along the Ethiopian part of the rift<br />

high temperature geothermal resources are associated with zones of quaternary tectonic and<br />

magmatic activity. Including the Afar depression at least 120 independent geothermal systems<br />

have been identified since the 1970s (UNEP 1973) about 24 of them are judged to have high<br />

enthalpy potential. To explore for these resources up to a pre feasibility stage a range of<br />

geoscientific methods is used in a defined sequence starting with regional reviews and remote<br />

sensing followed by geologic, hydrologic, geochemical and geophysical surveys.<br />

The applied geophysical methods usually comprise temperature measurements (gradient<br />

boreholes), seismology, magnetics and resistivity methods, including Magnetotellurics (MT).<br />

Geothermal surface manifestations like hot springs, fumaroles, geysers and the associated<br />

geological and geochemical settings are indicating the presence of a geothermal reservoir.<br />

Particularly resistivity methods<br />

may be used for delineating the<br />

lateral and depth extensions of<br />

such potential reservoirs. The<br />

MT method is frequently used<br />

for this purpose since it easily<br />

covers the necessary<br />

exploration depth down to<br />

approximately 10 km.<br />

In the following paper MT data<br />

and interpretation is presented,<br />

showing the already known<br />

shallow reservoir of the<br />

Tendaho geothermal field and<br />

its so far unknown deep<br />

structure, possibly feeding the<br />

shallow reservoir.<br />

Figure 1: Survey area at the SE end of the Hararo and Dabbahu<br />

magmatic segments. TGD = Tendaho-Goba’ad Discontinuity<br />

(after Ebinger et al. 2008).<br />

Survey area and tectonic<br />

setting<br />

The Tendaho geothermal field<br />

is located in the Afar<br />

depression (NE Ethiopia), at or<br />

very close to the assumed triple<br />

junction formed by the Red Sea,<br />

Gulf of Aden and East African<br />

rift arms. The extension of the Manda-Hararo axial rift zone in its south-easterly strike<br />

direction ends up at the Tendaho geothermal system (figure 1). The Red Sea and Aden rifts<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

131


are characterised by oceanic crust. The Afar triple junction is therefore a zone where thinned<br />

continental crust of the Main Ethiopian rift joins with crust of oceanic character (Barberi et al.<br />

1972).<br />

Within this area of active extensional tectonics several geothermal manifestations can be<br />

observed. Quaternary faulting and volcanism formed chains of fissural flows, basaltic cones<br />

and stratovolcanoes, usually accompanied by shallow seismicity and positive gravity<br />

anomalies. These active magmatic segments are comparable to slow spreading mid oceanic<br />

ridge segments and are thought to form new igneous crust by dyke like intrusions as<br />

demonstrated in the recent Dabbahu rift events (since 2005 up to now) just 80 km NW of the<br />

survey area. A previously mapped magmatic segment within the Manda-Hararo rift zone had<br />

been ruptured. Earthquake recordings indicated a 60 km long and about 8 m wide dyke<br />

intrusion (Ebinger et al. 2008, figures 8 & 9).<br />

Tendaho geothermal field<br />

The Tendaho geothermal field has been<br />

investigated in detail since the 1980s.<br />

Geological, geochemical as well as magnetic,<br />

seismological and resistivity data (DC<br />

soundings) were acquired by an extensive<br />

Ethiopian-Italian cooperation (Aquater, 1996).<br />

Based on this results, six exploratory wells had<br />

been drilled, three of them hitting a production<br />

horizon at approx. 300 m depth yielding steam<br />

temperatures above 250 °C (figure 2). Additional<br />

gravity and magnetic data had been acquired<br />

recently by Lemma and Hailu (2006) to delineate<br />

fractured zones which could serve as potential<br />

pathways for hydrothermal<br />

fluid flow from a deeper<br />

reservoir.<br />

To further investigate the<br />

proposed deep reservoir<br />

and/or heat source, deep<br />

reaching resistivity<br />

methods had to be applied.<br />

Due to the generally low<br />

resistivities the penetration<br />

depths of the DC<br />

soundings were limited to a<br />

few hundred meters. To<br />

extend this depth of<br />

exploration the MT method<br />

has been applied by BGR<br />

in collaboration with the<br />

Geological Survey of<br />

Ethiopia (GSE) during a<br />

field survey in 2007 (figure<br />

3). To reach the desired<br />

exploration depth of about 5<br />

km a frequency range from<br />

Figure 2: Production test of well TD5 (at well<br />

head: T >250 °C, p > 18 bar). The well is located<br />

inside the MT survey area.<br />

L1<br />

L2<br />

Kurub<br />

Figure 3: Survey area Tendaho geothermal field (red frame) projected onto<br />

satellite images (google earth). Red triangles = MT stations, blue squares =<br />

TEM stations. Line nos. L1, L2, L3 and L97 refer to MT lines in SE-NW<br />

bearing. TD5 = productive exploratory well. Yellow line = main road to<br />

Djibouti.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

132<br />

TD5<br />

<br />

L97<br />

L3


10 kHz to 0.01 Hz (100 s) has been acquired.<br />

MT resistivity signature<br />

Due to alluvial and lacustrine infill of the Tendaho graben apparent resistivity curves show<br />

generally values below 10 Ohm*m, even at high frequencies. A typical sounding curve is<br />

shown in figure 4.<br />

The minimum is<br />

reached at approximately<br />

4 Hz with<br />

apparent resistivities<br />

just below 1 Ohm*m.<br />

With decreasing<br />

frequency the<br />

apparent resistivity<br />

rises again. Source<br />

signal was generally<br />

quite low, especially<br />

within the dead band<br />

region ranging from<br />

approximately 1 to<br />

0.1 Hz (1 to 10 sec.).<br />

Figure 4: Typical sounding curve. Apparent resistivities (top) and phases Nonetheless 16 hrs<br />

(bottom). Due to low signal strength in the dead band from about 1 to 0.1 Hz of recording time<br />

unstable estimation especially of phase values, resulting in increased error bars.<br />

proved to be<br />

sufficient to yield<br />

acceptable estimates<br />

of the impedance<br />

tensor, using mainly<br />

single site processing.<br />

This apparent resistivity pattern, which is also reflected in the resistivity models (see below),<br />

is indicative for high temperature geothermal reservoirs (figure 5, see also e.g. Kalberkamp<br />

2007) where the resistivity low is interpreted as the clay cap while the increasing resistivities<br />

below the clay cap point to the core of the reservoir and represent a possible drilling target.<br />

In the 2d inverted resistivity sections, exemplarily<br />

shown for profile L1 in figure 6, this resistivity<br />

pattern can be seen quite clearly. Although<br />

resistivities are generally low, they show signatures<br />

typical for hydrothermal alteration halos (see e.g.<br />

Johnston et al. 1992, Kalberkamp 2007) and (partly)<br />

molten magma intrusions at depth below 4 km.<br />

Taking the Dabbahu rift events into account it seems<br />

to be likely that the heat source for the geothermal<br />

reservoir is formed by magma intrusions along dyke<br />

like fracture zones as it is suggested by our<br />

interpretation of the MT data. Areas with high<br />

resistivities (>300 Ohm*m) may be associated with<br />

basalts from the Afar Stratoid Series, constituting the<br />

borders of the Tendaho graben structure.<br />

Figure 5: Schema of a generalised<br />

geothermal system. The smectite cap<br />

formed exhibits resistivities in the<br />

range of 2 Ohm*m, the mixed layer<br />

around 10 Ohm*m (modified after<br />

Johnston et al. 1992).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

133


Figure 6: 2 d inverted resistivity section (L1). Lateral extension is 27 km; vertical depth covers 8 km.<br />

Interpretation and recommendations<br />

As has been shown for profile L1 exemplarily the magnetotelluric soundings show in their 2dimensional<br />

inverted resistivity sections regions with resistivities as low as 2 Ohm*m,<br />

especially in near surface layers along profiles L1, 2, and 3 as well as at greater depth (below<br />

5 km) in profiles P1, 3, and 97. When using the 2d MT sections to compile resistivity maps<br />

for constant elevations each, thus based on the inverted resistivities, we get a general view of<br />

the lateral resistivity structure as presented in figure 7 for selected elevations.<br />

Figure 7: Resistivity maps of survey area as covered by red frame in figure 3, resistivity range 2 (red) – 2048<br />

(blue) Ohm*m. Left: at 200 masl (150 m below surface). Shallow low resistive layer (red) due to sedimentary<br />

infill. The sediments are up to 1 km thick. Centre: at 1000 mbsl (1350 m below surface). Slight increase of<br />

resistivity (> 10 Ohm*m) possibly due to mixed layer clays and advancement towards deeper reservoir (blue<br />

circle). Right: at 9000 mbsl (9350 m below surface). Resistivity drop below 2 Ohm*m along NW-SE<br />

trending feature (partly molten magma dyke?). This may form the deep heat source feeding the shallow<br />

geothermal reservoir.<br />

At greater depth of around 7 km a<br />

resistivity anomaly below 2 Ohm*m<br />

appears elongating in NW-SE direction.<br />

This direction coincides with the strike<br />

direction of the Tendaho graben and rift<br />

structures further to the NW, up to the<br />

Dabbahu rift and Boina vent, where lava<br />

ascended along a fault system, almost<br />

reaching the surface (figure 8). Therefore<br />

it seems likely, that the low resistive<br />

structures at greater depth are caused by<br />

lava filled fracture zones. (figure 7 right).<br />

Ebinger et al. (2008) have presented a<br />

working model of the Dabbahu magmatic<br />

Figure 8: Volcanic vent, created during 2006 eruption<br />

event (Photo by J. Rowland).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

134


Figure 9: Working model of the Dabbahu magmatic system as<br />

presented by Ebinger et al. (2008). Pink ellipses = shallow magma<br />

chambers, connected to assumed lower crust/upper mantle feeding<br />

zones.<br />

system (figure 9) based<br />

mainly on seismological,<br />

radar interferometric and<br />

structural data. They suggest<br />

that lower crustal/upper<br />

mantle source zones are<br />

feeding the basaltic dyke<br />

and shallow magma<br />

chambers at a few kilometres<br />

depth.<br />

Magnetotelluric measurements<br />

by the Afar Rift<br />

Consortium / University of<br />

Edinburgh (Scotland, U.K.)<br />

are ongoing in that area and<br />

results thereof may help to<br />

refine the interpretation of<br />

the Tendaho geothermal<br />

field data.<br />

With the current data therefore the following preliminary conclusions may be drawn:<br />

The deep heat source feeding the currently drilled reservoir at 300 m depth is most likely<br />

(partly) molten magma along NW-SE trending dykes or faults. In the Tendaho geothermal<br />

field these structures may be as shallow as 4 km depth.<br />

A deep reservoir may be expected below 1300 m.<br />

An up flow zone could be present at the S end of the survey area indicated by low<br />

resistivities throughout the acquired depth range.<br />

To identify the proposed up flow zone at the S end of the current survey area it is<br />

recommended to apply additional TEM soundings which could enhance the resolution at<br />

shallow depth.<br />

Additional MT sounding profiles are recommended further towards the Hararo and Dabbahu<br />

magmatic segments NW of the survey area. Since rift events including ascending magma are<br />

evident further to the NW it may be assumed that high temperature reservoirs could be<br />

reached at comparatively shallow depth there. First MT results from the research work within<br />

the Afar Rift Consortium do support this assumption (Desissa et al., 2009).<br />

Also additional MT soundings beyond the Awash river up to the geothermal manifestations of<br />

Alalobad in the south would be helpful to establish either the extension of one large reservoir<br />

or the existence of a separate second geothermal system.<br />

Acknowledgements<br />

The MT survey in Ethiopia has been carried out as part of a cooperation project between the<br />

Geological Survey of Ethiopia (GSE) and BGR as part of the GEOTHERM Programme. We<br />

gratefully acknowledge the cooperation with Mohammednur Desissa, Yohannes Lemma and<br />

the Geothermal Working Group within the GSE. We also appreciate helpful communications<br />

and cooperation with Kathy Whaler and the Afar Rift Consortium.<br />

GEOTHERM (www.bgr.de/geotherm/) is a technical cooperation programme to promote the<br />

use of geothermal energy in partner countries, implemented by the BGR on behalf of the<br />

German Federal Ministry for Economic Cooperation and Development (BMZ) under contract<br />

no. 2002.2061.6.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

135


References<br />

Aquater; 1996. Tendaho geothermal project, final report. Ministry of Mines of Ethiopia and<br />

Ministry of Foreign Affairs of Italy, San Lorenzo in Campo, Italy (unpublished).<br />

Barberi, F., Tazieff, H. & Varet, J., 1972. Volcanism in the Afar depression: its tectonic and<br />

magmatic significance. Tectonophysics, 15, 19-29.<br />

Desissa; M., Whaler, K., Hautot, S., Dawes, G., Fisseha, S., Johnson, N., 2009. A<br />

Magnetotelluric study of continental lithosphere in the final stages of break-up: Afar,<br />

Ethiopia. Abstract I.06-1158, 11th IAGA Scientific Assembly, Sopron, Hungary.<br />

Ebinger, C.J., Keir, D., Ayele, A., Calais, E., Wright, T.J., Belachew, M., Hammond, J.O.S.,<br />

Campbell, E. & Buck, W.R., 2008. Capturing magma intrusion and faulting processes<br />

during continental rupture: seismicity of the Dabbahu (Afar) rift, Geophys. J. Int.<br />

Johnston, J.M., Pellerin, L. & Hohmann, G.W. (1992): Evaluation of Electromagnetic<br />

Methods for Geothermal Reservoir Detection, Geothermal Resources Council<br />

Transactions, 16, 241-245.<br />

Kalberkamp, U., 2007. Exploration of geothermal high enthalpy resources using<br />

Magnetotellurics – an Example from Chile, in: Ritter, O., Brasse, H. (eds.): Protokoll<br />

22. Kolloquium Elektromagnetische Tiefenforschung, Hotel Maxiky. Dín, Czech<br />

Republic, 1.-5. Oktober 2007, DGG, 194-198.<br />

Lemma, Y., Hailu, A., 2006. Gravity and magnetics survey at the Tendaho geothermal field.<br />

GSE, Addis Ababa, Ethiopia, 23pp (unpublished),.<br />

UNEP (ed.), 1973. Ethiopia: Investigation of geothermal resources for power development –<br />

Geology, geochemistry and hydrology of hot springs of the East African Rift System<br />

within Ethiopia, Technical Report DP/SF/UN/116, United Nations, New York.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

136


Electromagnetic Monitoring of CO2 Storage in Deep Saline<br />

Aquifers - Numerical Simulations and Laboratory<br />

Experiments<br />

J. H. Börner, V. Herdegen, R.-U. Börner, K. Spitzer<br />

Institut für Geophysik, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 12, 09599 Freiberg<br />

1 Introduction<br />

The knowledge of petrophysical parameters and their contrasts is crucial to reliably monitoring CO2<br />

storage processes. The electrical conductivity appears to be a sensitive indicator for a resistive gaseous<br />

or supercritical CO2 phase replacing a conductive pore fluid in a porous medium like a saline sandstone<br />

aquifer. However, detailed knowledge on the influence of supercritical CO2 on the electrical resistivity<br />

of a formation is not sufficiently available yet.<br />

Therefore, we have carried out laboratory experiments to predict the contrast in electrical resistivity due<br />

to the presence of CO2 in an initially water-saturated sand sample resembling the petrophysical situation<br />

typical for a reservoir. Furthermore, the expected parameter contrasts were estimated according to<br />

empirical equations and numerical simulations (Fig. 1).<br />

On a middle- to long-term perspective, we aim at developing an electromagnetic monitoring technique<br />

using a borehole transient electromagnetic sensor which offers a unique opportunity to generate<br />

enhanced sensitivity at depth with respect to detecting migrating CO2 in a reservoir. This work is<br />

therefore integrated into national CCS research programs with an interdisciplinary variety of partners.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Figure 1: Workflow of the numerical simulation steps necessary for the transformation of water saturation into<br />

electrical resistivity.<br />

2 Theory<br />

Two-phase flow is governed by two equations simultaneously enforcing continuity of the water (w) and<br />

the CO2 (co2) phase flow (Busch et al. 1993). Both equations are linked by retention curves Se(pc) and<br />

relative permeabilities kr(S), such that<br />

∇·d w [− κkw r<br />

ηw (∇pw + d w g∇D)] = −Φ ∂(dwS w e )<br />

+ d<br />

∂t<br />

w w0<br />

∇·d co2 [− κkco2 r<br />

ηco2 (∇pco2 + d co2 g∇D)] = −Φ ∂(dco2Sco2 e )<br />

+ d<br />

∂t<br />

co2 w0<br />

with d as density, κ as intrinsic permeability, η as viscosity, Φ as porosity and g as gravitational acceleration.<br />

The link between both phases has been established by experimental data or parameterization,<br />

23 . Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

137<br />

(1)


e.g., by van Genuchten (1980) or Mualem (1976) (with effective saturation Se, capillary pressure height<br />

Hc and parameters α, n,m,L):<br />

1<br />

S w e =<br />

(1 + |αHc| n ) m<br />

k w r =(S w e ) L ·<br />

<br />

1 −<br />

<br />

1 − (S w e ) 1/m m 2<br />

S co2<br />

e =1− S w e (2)<br />

k co2<br />

r<br />

=(1− S w e ) L ·<br />

<br />

1 − (S w e ) 1/m 2m<br />

The resulting distribution of water saturation S(x, y, z, t) can be transformed into electrical formation<br />

resistivity ρ using empirical resistivity models, e.g., Archie’s law (Archie 1942):<br />

ρ = a1 · Φ −a2 −a3 1<br />

· S · (4)<br />

σw<br />

with σw denoting pore water conductivity. Archie’s empirical parameters a1, a2 and a3 strongly depend<br />

on rock formation characteristics and can be combined with porosity and water saturation yielding the<br />

formation factor F , such that<br />

ρ = F · ρw. (5)<br />

For sands and sandstones with considerable clay content Waxman and Smits (1968) expanded Archie’s<br />

law to account for both electrolytic conductance and interfacial conductance:<br />

ρ =<br />

F ∗<br />

S n · (σw + B·Qv<br />

S )−1 (6)<br />

with a unique formation factor F ∗ , the counterion mobility B describing the weak dependance of the<br />

interface conductivity on pore water salinity and the shalyness parameter Qv being the cation exchange<br />

capacity normalized to the pore volume.<br />

3 Numerical simulation studies<br />

The process of CO2 sequestration as well as the indication of a successful storage by observation of<br />

changes in electrical resistivity has been simulated following the steps indicated by the workflow shown<br />

in Fig. 1.<br />

Figure 2: CO2 saturation after 6 months of constant injection of 1000 m 3 per day into a 25 m thick reservoir.<br />

The injection well is located at the geometric center. These results were obtained using the FD<br />

simulation code Mod2PhaseThermo.<br />

23 . Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

138<br />

(3)


23 . Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

139


23 . Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

140


Final experiments have demonstrated that an increase in resistivity may also be achieved under<br />

supercritical conditions. However, the sample was not homogeneously infiltrated by the CO2 and<br />

preferential flow paths were built up during the experiment. The pore water has been pressed out only<br />

partially, leaving behind tube-shaped flow channels and a very heterogeneous water distribution with<br />

average residual water contents of more than 70 %.<br />

pressure height in m<br />

500<br />

400<br />

300<br />

200<br />

100<br />

van Genuchten<br />

measured data<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

water saturation<br />

Figure 6: Relation between CO2 pressure and water saturation derived from data obtained by the initial<br />

pressure build-up for an experiment with a projected maximum pressure of 50 bar. The reference<br />

data have been calculated according to van Genuchten (1980) with α = 0.02 and n =3.9 (cf. eqs.<br />

(2)-(3)).<br />

The complications are due to the density of supercritical CO2 which is large compared to the density<br />

of CO2 in its gaseous state. Consequently, the current experimental assembly causes the high density<br />

to result in low flow velocities within the measuring cell. In addition to the flow channels, this effect<br />

prevents an effective replacement of the pore water by supercritical CO2.<br />

The experiments have shown that theoretical pressure-saturation relations can generally be verified in<br />

practice using our set-up (Fig. 6). All experimental data sufficiently agree with Archie’s law. However,<br />

we have to carry out further investigations on the effects of CO2 dissolving in the pore water. There<br />

are indications that this has an important impact on the pore water resistivity depending on pressure,<br />

temperature, and brine salinity.<br />

6 Conclusions<br />

Numerical simulation studies and laboratory experiments show that the electrical resistivity of a<br />

fluid-saturated porous medium is highly sensitive to the presence of CO2. Geo-electromagnetic methods<br />

are therefore considered as a promising approach for monitoring CO2 storage.<br />

Still, the laboratory set-up has to be improved further to provide reliable results at high pressures.<br />

Subject to these prerequisites, well-founded simulations of electromagnetic monitoring scenarios can be<br />

carried out.<br />

Further laboratory experiments will also aim at quantifying the influence of dissolved CO2 on pore<br />

water and formation resistivity. A feasibility study could show whether electromagnetic methods are<br />

able to monitor, e.g., an expanding plume of CO2-rich formation water during CO2 injection. Finally,<br />

the applicability of the laboratory assembly for measuring reliable retention curves of unconsolidated<br />

sedimentary rocks will be further tested in the near future.<br />

23 . Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

141


References<br />

Archie, G. E. (1942). The electrical resistivity log as an aid in determining some reservoir characteristics. Trans. Americ.<br />

Inst. Mineral. Met. (146): 54–62.<br />

Busch, K.-F., L. Luckner, and K. Tiemer (1993). Geohydraulik. Edited by G. Matthess. 3rd edition. Volume 3. Lehrbuch<br />

der Hydrogeologie. Gebrüder Borntraeger. isbn: 3-443-01004-0.<br />

Häfner, F. and S. Boy (2009). Mod2PhaseThermo Nutzermanual. IBeWa.<br />

Mualem, Y. (1976). A new Model for predicting the hydraulic condyctivity of unsaturated porous media. Water Resour.<br />

Res. 12: 283–291.<br />

Schön, J. (1996). Physical Properties of Rocks - Fundamentals and Principles of Petrophysics. 1st edition. Volume 18.<br />

Handbook of Geophysical Exploration. Section I, Seismic Exploration. Pergamon.<br />

van Genuchten, M. (1980). A Closed-form Equation for Predicting the Hydraulic Conductivity of Unsaturated Soils. Soil<br />

Science Society of America Journal 44(5): 892–898.<br />

Waxman, M. H. and L. J. M. Smits (1968). Electrical conductivities in oil-bearing shaly sands. Society of Petroleum<br />

Engineers Journal 8: 107–122.<br />

23 . Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

142


Erkundung eines Aquifers unter dem Mittelmeer<br />

vor der israelischen Küste mit LOTEM.<br />

K. Lippert 1 , B. Tezkan, R. Bergers, M. Gurk, M. v. Papen, P. Yogeshwar<br />

Institut für Geophysik und Meteorologie, Universität zu Köln<br />

Abstract<br />

Im Rahmen dieses BMBF-geförderten Projektes 2 kommt die Long-Offset Transient<br />

Elektromagnetik (LOTEM) Methode zum ersten Mal in mariner Umgebung zur Erkundung<br />

von Grundwasseraquiferen zum Einsatz. Hauptziel des Projektes ist die Detektion<br />

der Süßwassergrenze unter dem Mittelmeer an der Küste Israels. Zu diesem Zweck wurden<br />

Hardware-Modifikationen durchgeführt und bereits in einer ersten Testmessung in Israel<br />

erprobt. Des weiteren werden die Interpretationen der ersten Testmessung sowie die Modellierungsergebnisse<br />

des Küstenabschnitts vorgestellt, welche zur Planung der Hauptmessung<br />

wichtig sind.<br />

Einleitung<br />

Die Bedeutung von Offshore-Süßwasserleitern<br />

für das Grundwassermanagment nahm in den<br />

letzten Jahren stark zu. Das Vorhandensein<br />

dieser Aquifere, die sich bis zu mehreren<br />

Kilometern unter dem Meeresboden erstrecken<br />

können, wurde in der Fachliteratur beschrieben:<br />

z.B. an der Küste von Guyana [Arad, 1983]<br />

oder vor den Niederlanden [Groen et al.,<br />

2005]. Der israelische Küstenaquifer ist eine<br />

der Hauptgrundwasserresourcen des Landes<br />

(vgl. Fig. 1). Aufgrund der intensiven<br />

Nutzung verschlechtert sich allerdings die<br />

Wasserqualität zumehmend. Die Gründe hierfür<br />

liegen sowohl in der anthropogenen Verschmutzung<br />

vom Land aus, als auch im Eindringen<br />

von Salzwasser.<br />

1 E-mail: lippert@geo.uni-koeln.de<br />

2 BMBF-Förderkennzeichen: 02WT0987<br />

Figure 1: Die großen Aquifere der Region.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

143


Geologie und Fragestellung<br />

Der Küstenaquifer besteht hauptsächlich aus<br />

kalkhaltigem Sandstein und Sanden und ist<br />

selbst noch in vier Sub-Aquifere unterteilt<br />

(vgl. Fig.2). Die oberen zwei grundwasserführenden<br />

Schichten werden vom Land<br />

aus anthropogen und vom Meer aus durch<br />

Salzwassereindringen verunreinigt. Bisher<br />

wurde davon ausgegangen, daß die unteren Sub-<br />

Aquifere stellenweise vom Meerwasser getrennt<br />

sind und dort keine Salzwasserintrusion stattfindet.<br />

Neuere Untersuchungen ([Kafri, U.,<br />

Goldman, M., 2006],[Yechieli et. al, 2009])<br />

hingegen deuten jedoch auf das Fortsetzen des<br />

Aquifers unter dem Meer hin. So zeigen<br />

z.B. TDEM-Messungen auf Land, nahe der<br />

Küste, einen schlechteren elektrischen Leiter<br />

10 (Ωm) eingebetet zwischen zwei guten elektrischen<br />

Leitern (∼ 2 Ωm). Diese Schicht,<br />

in der Tiefe der unteren Sub-Aquifere, wird<br />

als Grundwasserführenden Schicht interpretiert<br />

(vgl. Fig.3). Die Fragestellung an die LOTEM-<br />

Methode ist nun 1. die Prüfung der Existenz<br />

des unteren Sub-Aquifers in etwa 100m Tiefe<br />

und 2. dessen Ausbreitung unter dem Mittelmeer.<br />

Figure 2: Möglicher Verlauf des unteren Subaquifers an der Küste [Yechieli et. al, 2009]: a)<br />

Der untere Subaquifer (CD) hat keine Verbindung zum Meer, b): Das Salzwasser dringt in den<br />

unteren Aquifer ein.<br />

Figure 3: Onshore SHOTEM-Ergebnisse [Kafri, U., Goldman, M., 2006].<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

144


Hardwaremodifikationen<br />

Das Kölner LOTEM-Equipement ist für<br />

Landmessungen konzipiert. Deswegen waren<br />

einige Hardware-Modifikationen für die Offshoremessung<br />

nötig. Als Sender konnte aus<br />

logistischen Gründen nur der kleinere Kölner<br />

Sender 3 verwendet werden. Da dieser bei<br />

verwendeten Stromstärken von ∼ 9A zuviel<br />

Wärme entwickelt, war es nötig eine zusätzliche<br />

Wasserkühlung einzubauen. Für die Einspeisung<br />

des Sendesignals wurden neue Elektroden<br />

entwickelt (vgl. Fig.4), deren Form,<br />

durch Maximierung des angekoppelten Wasservolumes,<br />

elektrochemische Vorgänge minimiert.<br />

Zusätzlich kann das Sendesignal Stromgeregelt<br />

werden, um einen konstanten Stromverlauf zu<br />

erreichen. In Fig.5 sind Sendesignale gezeigt.<br />

Links wird mit konstanter Spannung gesendet.<br />

Bei sich verändernder Ankopplung ergibt sich<br />

kein konstanter Stromverlauf. Deswegen wurde<br />

der Sendestrom nachträglich auf einen konstan-<br />

Figure 5: Aufgezeichnete Sendesignale bei<br />

einem Test in Wilhelmshaven.<br />

ten Wert heruntergeregelt (Fig.5 rechts).<br />

Die verwendeten Magnetfeldsensoren 4 wurden<br />

in Druck- und Salzwasserbeständige Behälter<br />

verpackt (Fig.6).<br />

Figure 4: Neuentwickelte Elektroden für die<br />

Einspeisung des Sendesignals.<br />

Figure 6: Seewasserfestes Gehäuse für die<br />

Magnetfeld-Sensoren.<br />

3 NT-20 von Zonge Engineering & Research Organization, Inc., Tucson AZ, USA<br />

4 TEM/3-Spulen von Zonge Engineering & Research Organization, Inc., Tucson AZ, USA<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

145


Ergebnisse Messung 2008<br />

Anfang November 2008 wurden die ersten Messung<br />

vor Ort durchgeführt. Hierbei wurden verschiedene<br />

Messkonfigurationen gestestet:<br />

• BroadSide: Sender an der Küstenlinie,<br />

Empfänger an Land. Offset 200m. vgl.<br />

Fig.7: Tx2 und Rx3.<br />

• BroadSide: Sender an der Küstenlinie,<br />

Empfänger im Wasser. Offset 300m. vgl.<br />

Fig.7: Tx2 und Rx2.<br />

• BroadSide: Sender im Wasser,<br />

Empfänger an Land. Offset 500m. vgl.<br />

Fig.7: Tx1 und Rx1.<br />

Figure 7: Setups der Messung 2008.<br />

Das Setup “Sender an der Küstenlinie,<br />

Empfänger an Land” ist das einzige, welches<br />

annehmbar 1D interpretiert werden konnte.<br />

Als Beispiel ist hier die dHz/dt - Komponente<br />

gezeigt. Das Endmodell einer Marquardt-<br />

Inversion ist in Fig.9, die zugehörige Datenanpassung<br />

in Fig.8 zu sehen. Das Startmodell<br />

wurde anhand der Vorinformationen und<br />

dem Ergebnis einer Occam-Inversion gewählt 5 .<br />

Figure 8: Datenanpassung des “besten” Modells<br />

(rot) in Fig. 9.<br />

Figure 9: 1D-Inversionsergebnis mit<br />

Äquivalenzmodellen.<br />

Das Ergebnis ist gut mit den SHOTEM-<br />

Ergebnissen (Fig. 3) vergleichbar. Der gesuchte<br />

Aquifer ist deutlich in einer Tiefe ab ca. 90m<br />

zu erkennen. Sowohl die Schichtdicke mit ca.<br />

100m als auch viele<br />

Äquivalenzmodelle mit<br />

einem Widerstand von ∼ 10 Ωm sind vergleichbar<br />

und glaubwürdig.<br />

Die anderen beiden Messkonfigurationen<br />

5Vgl. M.v.Papen, B.Tezkan: “On the analysis of LOTEM time series from Israel and preliminary 1D<br />

inversion of data.”, Poster EMTF 2009.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

146


können nicht als eindimensionales Problem<br />

angesehen und somit auch nicht 1D ausgewertet<br />

werden. Die gemessenen Transienten<br />

wurden mit synthetischen Daten verglichen.<br />

Das 2D Leitfähigkeitsmodell ist in Fig. 11<br />

dargestellt. Mit diesem synthetischen Modell<br />

war eine qualitative Anpassung möglich. Exemplarisch<br />

ist in Fig.10 die Datenanpassung für<br />

die Ex-Komponente gezeigt.<br />

Figure 10: Setup Tx2 - Rx3 (vgl.Fig.7): Anpassung<br />

der Messdaten der Ex-Komponente mit<br />

modellierten Daten.<br />

Vorab-Modellierungen<br />

Für die Messung im November 2009 sind die<br />

Empfängerlokationen von Bedeutung. Eine<br />

Frage hierbei ist u.a. ob das Target auch mit<br />

einem Sender auf dem Meer und Empfängern<br />

auf dem Land auflösbar ist. Diese Konfiguration<br />

ist logistisch relativ einfach durchführbar.<br />

Es wurde also ein 2D-Küstenmodell erstellt,<br />

mit welchem synthetische Daten mittels<br />

SLDMEM3T [Druskin, Knizhnermann, 1988]<br />

erzeugt wurde. Als Datenbeispiel sind in Fig.12<br />

die Ex-Komponenten dargestellt. Hierfür werden<br />

im Modell (Fig.11) Daten mit Targetlayer<br />

und einmal ohne Targetlayer unter dem Meeresboden<br />

verglichen. “Ohne Targetlayer” bedeutet<br />

in diesem Fall, daß der Targetlayer 300m im<br />

Landesinneren endet (nicht dargestellt). Die<br />

darüberliegende Schicht (1.2 Ωm, gelb) wurde<br />

nach unten erweitert.<br />

Modelliert wurden verschiedene Senderpositionen<br />

in unterschiedlichen Entfernungen zur<br />

Küstenlinie. Im Beispiel beträgt die Entfernung<br />

1500m. Betrachtet werden nun 2<br />

Empängerpositionen: einmal 275m von der<br />

Küstenlinie entfernt auf dem Meeresboden und<br />

einmal 95m im Landesinneren. Modelliert<br />

wurde die BroadSide Konfiguration (Senderrichtung<br />

und Profilrichtung senkrecht zueinander)<br />

und die InLine Konfiguration (Senderrichtung<br />

und Profilrichtung auf einer Linie). Die<br />

System-Antwort wurde für diese Modellierungen<br />

nicht berücksichtigt.<br />

Figure 11: Verwendetes Küstenmodell. Das<br />

gesuchte Target ist die 15 Ωm-Schicht (rot).<br />

Blau = Wasser.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

147


Sowohl bei der InLine als auch bei der Broad-<br />

Side Konfiguration ist in dieser Komponente<br />

der Targetlayer klar auflösbar: die Transienten<br />

unterscheiden sich deutlich in den Amplituden<br />

von den Transienten ohne Targetlayer. Bei<br />

der BroadSide Konfiguration verschiebt sich der<br />

auftretende Vorzeichenwechsel durch das Target<br />

zu späten Zeiten. Ein Vorzeichenwechsel ist<br />

bei der Auswertung immer schwer interpretierbar.<br />

Aufgrund der Unterscheidbarkeit der Transienten<br />

sind Empfänger auf dem Land für diese<br />

Fragestellung sinnvoll.<br />

Figure 12: Modellierte Transienten der Ex-<br />

Komponente an den Positionen, die in Fig.11<br />

als Kreise markiert sind.<br />

Zusammenfassung und Ausblick<br />

Aufgrund der Ergebnisse und der Erfahrungen<br />

der ersten Messung (2008) und der Modellierungen<br />

sind die Setups für die erste Hauptmessung<br />

im November 2009 festgelegt worden. Die Messung<br />

ist inzwischen durchgeführt worden. Es<br />

wurden die Konfigurationen InLine und Broad-<br />

Side gemessen. Der Sender befand sich dabei<br />

immer auf dem Meer in verschiedenen Abständen<br />

zur Küste. Pro Senderposition gab es mehrere<br />

Empfängerpositionen, sowohl auf dem Wasser<br />

als auch auf Land.<br />

Literatur<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

148<br />

Arad, A.: “ A summary of artesian coastal<br />

basin of Guyana.”, J.Hydrol. 63:299-313, 1983.<br />

Druskin, V.L., Knizhnermann, L.A.: ”A<br />

spektral semi-discrete method for the numerical<br />

solution of 3d nonstationary problems in electrical<br />

prospecting.”, Phys. Solid Earth, 24:641-<br />

648, 1988.<br />

Groen J., Groen M., Post V., Kooi H.,<br />

Mendizabal I. and Groot S.M.: ” Seaward<br />

continuation of groundwater flow systems along<br />

the coast of the Netherlands.”, GSA Salt Lake<br />

Ann.Meet. Abstract Paper No.211-9., 2005.<br />

Kafri U. and Goldman M.: “ Are the lower<br />

subaquifers of the Mediterranean coastal aquifer<br />

blocked to seawater intrusion? Results of a TDEM<br />

(time domain electromagnetic) study.”, Isr.J.Earth<br />

Sci., 2006.<br />

Yechieli et. al: “ The inter-relationship between<br />

coastal sub-aquifers and the Mediterranean<br />

Sea, deduced from radioactive isotopes analysis.”<br />

Hydrogeology Journal Vol 17 Num 2 /<br />

2009.


On the analysis of LOTEM time series from Israel and the preliminary 1D inversion of<br />

data<br />

M. von Papen, B. Tezkan<br />

Institute of Geophysics and Meteorology, University of Cologne, Germany<br />

Abbildung 1: Overview of measurement setups (A,B,C from left to<br />

right)<br />

1. Data Processing of TEM time series from the<br />

first Israel survey<br />

Abbildung 2: Power spectrum of Ex at station A/B before (blue<br />

asterisks) and after application of different filters<br />

The measurement took place at the coast near Ashdod,<br />

Israel (fig.1), where the task is the detection of fresh<br />

groundwater bodies within the mediterranean submarine<br />

aquifers [3]. For that matter a 380 m long dipole emitting<br />

in a 50% duty cycle was used, located on the shoreline for<br />

stations A and C and at -300m offshore on the seabottom<br />

for station B. Receiver stations recorded four components<br />

of the EM field (Ex,Ey, ˙ Hy, ˙ Hz) with Summit receivers. Receivers<br />

at station C were set up 5mbeneath the water on<br />

the sea bottom.<br />

• The measured transients consist of 4096 data points<br />

(DP) with a sample rate of 1/16 ms and 256 DP as<br />

onset. Switch time of Tx is t0 = 1500ms.<br />

Abbildung 3: Parts of transient [5]<br />

Noise measurements showed a dominant 50 Hz noise<br />

together with multiples (fig.2). The average noise<br />

power for each component is: PN,Ex = 0.2V 2 ,PN,Ey =<br />

0.02V 2 ,P N, ˙ Hy = 4.8V 2 ,P N, ˙ Hz = 0.2V 2 (measured on<br />

land). Due to relative short offsets of 200-500 m between<br />

receiver and transmitter the signal is - except for Ey component<br />

- clearly visible in the raw transients.<br />

Abbildung 4: Stacked transients of only on-switches and only offswitches<br />

2. Application of a segmented lockin filter<br />

First step of the processing was to filter out the 50 Hz<br />

noise. This was done using a segmented lockin filter (LOF)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

149


Abbildung 5: Quantiles of normal distribution for stations A,B,C<br />

from top to bottom<br />

[1]. It fits harmonics in segments of 60 ms to the data and<br />

subtracts them. Amplitude and phase are computed with<br />

Fourier series expansion. Regarding the high dynamic of<br />

the transient at the time of the switch (e.g. switch on in<br />

part B of fig.3) the fitting is being done in parts A and<br />

C only [5]. In addition a 3rd degree polynomial is fitted<br />

in part C to simulate the descent of the transient. Other<br />

filters like Delay Lower Nyquist frequency (DLN), Delay<br />

filter or (not segmented) Lockin-Filter [1, 5] remove the<br />

noise less efficient for this dataset (fig.2).<br />

• Application of the filter reduced the noise power, computed<br />

at the end of the time series, to PN,Ex =<br />

0.2nV 2 ,P N, ˙ Hy =0.2V 2 and P N, ˙ Hz =1.4nV 2 .<br />

• The DC bias is removed by levelling the data in the<br />

onset after filtering.<br />

Due to the short onset and the resulting error in determining<br />

the DC bias the inverted conductivities will exhibit<br />

a slight shift [4]. Higher accuracy can be achieved by averaging<br />

over a complete transmitter period.<br />

3. Analysis of data distribution<br />

The data measured with 50% duty cycle consists of two<br />

different sets of transients. One set is created when switching<br />

the current off and another when switching on. Because<br />

the given transmitter system shows a much shorter<br />

ramp time for off-switches, these transients are generally<br />

less broad and have a higher peak than those created by<br />

switch-on (fig.4). Cluster analysis [2] in the blue area of<br />

fig.4 resulted in two equally populated groups identified as<br />

on and off switches by cross checking the Ex component.<br />

The distribution of the data in terms of probability can<br />

be visualized with quantiles of gauss distribution (fig.5).<br />

The mean values of all transients in a certain time segment<br />

are sorted by ascending value and are then plotted<br />

against an axis, which is scaled in such a way that normal<br />

distributed data will show as a line of slope 1. This has<br />

been done for each component over a 62 ms window at the<br />

end of each switch-off time series, which have been filtered<br />

and levelled beforehand.<br />

The top graph shows that the distribution of Hz (and<br />

- to a smaller extent - Ex) at station A is dominated by<br />

a step. However, cluster analysis could not sort out these<br />

time series. The other stations show a much better result<br />

with normal distributed data for all components.<br />

4. Stacking and 1D Occam inversion<br />

After the data is analysed and if it needs no further<br />

editing (like correction of the right switch moment) it is<br />

ready to be stacked and smoothed.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

150<br />

• The data was stacked selectively omitting the highest<br />

and lowest 5% of the values per DP.


Abbildung 6: Normalized stacked transients sorted by component<br />

Abbildung 7: Normalized system responses measured broadside<br />

Abbildung 8: Result of Occam inversion of data at station A for 1st<br />

and 2nd order roughness<br />

• Afterwards it was smoothed with a dynamic Hanning<br />

window function.<br />

Fig.6 shows the normalized transients of the stations<br />

sorted by components. Field setup is according to fig. 1.<br />

Ex shows a noticeable “bump” in early times at receiver<br />

station C, which could be a 2D effect of the water-land<br />

interface. More on interpretation of this data can be found<br />

in Lippert et al., 2010 .<br />

System responses measured on land are shown in fig.7<br />

and are used for convolution in forward calculation. Fig.8<br />

shows the Occam inversion results for station A. All components<br />

show a good conductor at about 80 m depth between<br />

two more resistive layers at approximately 10 m and<br />

150 m depth, which represent the sought fresh water bodies.<br />

Hz ˙ could resolve a second good conductor, which is<br />

not visible in ˙ Hy and only indicated in Ex. These results<br />

are in good agreement with the SHOTEM measurements<br />

of this area.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

151


5. Future research<br />

Some system responses measured broadside include negative<br />

values. This is supposed to be a geometric phenomenom<br />

related to the ’air wave’. Future measurements of<br />

the system response will therefore be conducted in the laboratory.<br />

For the geometries of stations B and C there are at the<br />

moment no inversion codes available, therefore they will be<br />

interpreted using 2D forward modelling. Joint inversions<br />

using different LOTEM components will be realized.<br />

6. References<br />

Literatur<br />

[1] T. Hanstein. Digitale Optimalfilter für LOTEM Daten. Protokoll<br />

über das 16. Kolloquium EMTF, pages 320–328, 1996.<br />

[2] S. Helwig. Clusteranalyse als Tool zur Selektion und Verarbeitung<br />

elektromagnetischer Daten. Protokoll über das 16. Kolloquium<br />

EMTF, pages 156–161, 1996.<br />

[3] K. Lippert et al. Erkundung eines Aquifers unter dem Mittelmeer<br />

vor der israelischen Küste mit LOTEM. Protokoll über das 23.<br />

Kolloquium EMTF, 2010.<br />

[4] A. Osman. Interpretation der LOTEM-Daten in näherer Umgebung<br />

von den Bohrungen des KTB. Diploma Thesis, Institute<br />

for Geophysics, University of Cologne, 1995.<br />

[5] C. Scholl. Die Periodizität von Sendesignalen bei LOTEM. Diploma<br />

Thesis, Institute for Geophysics, University of Cologne,<br />

2001.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

152


Grundwasserkontamination bei Roorkee/Indien:<br />

2D Joint Inversion von Radiomagnetotellurik und Gleichstromgeoelektrik<br />

Daten<br />

P. Yogeshwar 1 , B. Tezkan 1 , M. Israil 2<br />

1: Institut für Geophysik und Meteorologie, Universität zu Köln, Email: yogeshwar@geo.uni-koeln.de<br />

2: Earth Science Departement, Indian Institute of Technology, Roorkee<br />

In der landwirtschaftlich geprägten Region um Roorkee in Nordindien ist der Einsatz von Düngemittel<br />

sowie die Bewässerung der Felder mit Abwasser gängige Praxis, wobei das Abwasser durch<br />

offene, teilweise unbefestigte Kanäle transportiert wird. Die umliegende Landbevölkerung bezieht<br />

ihr Frischwasser oftmals aus Brunnen, die aus einem oberflächennahen Aquifer gespeist werden.<br />

Eine Gefährdung dieser oberflächennahen Aquifersysteme durch den praktizierten Umgang mit<br />

Abwasser und Düngemittel ist nicht auszuschließen.<br />

Im Rahmen der hier durchgeführten Untersuchungen wurden ein landwirtschaftlich genutztes und<br />

stark mit Abwasser bewässertes Gebiet nahe einer Mülldeponie und ein als kontaminationsfrei<br />

angenommenes Referenzprofil mit den Methoden der Gleichstromgeolelektrik (DC) und der Radiomagnetotellurik<br />

(RMT) vermessen.<br />

Die einzelnen Profile wurden neben Benutzung von gängiger 1D und 2D Auswertesoftware mittels<br />

eines neu entwickelten 2D Joint Inversionsprogrammes invertiert und interpretiert.<br />

Verglichen mit den Ergebnissen des Referenzprofils zeigen die Ergebnisse auf dem kontaminierten<br />

Gebiet nahe der Deponie deutlich erhöhte Leitfähigkeitswerte bis zu einer Tiefe von ca. 15 m. Der<br />

Tiefenbereich von 5 m bis ungefähr 15 m konnte mit dem oberflächennahen Aquifer identifiziert<br />

werden. Darüber hinaus ergibt sich eine Abnahme der Leitfähigkeitswerte mit dem Abstand von<br />

der Mülldeponie und den Abwasserauslässen.<br />

Durch die Anwendung einer 2D Joint Inversion konnten die Datensätze beider Methoden durch<br />

einheitliche Untergrundmodelle angepasst werden.<br />

Einleitung<br />

Roorkee befindet sich in den Vorläufern des Himalaja<br />

am rechten Ufer des Solani Flusses im Distrikt Haridwar<br />

des nordindischen Staates Uttharakand (Geographische<br />

Breite: 29 ◦ 50 ′ bis 29 ◦ 56 ′ N, Geographische<br />

Länge: 77 ◦ 48 ′ bis 77 ◦ 56 ′ E). Der Solani ist ein Nebenfluss<br />

des Ganges. Der Ganges ist in dem Gebiet<br />

kanalisiert und teilt Roorkee in „Old“- und „New“-<br />

Roorkee. Die Stadt hat ca. 100000 Einwohner, mit<br />

einer jedoch wachsenden Bevölkerung. Die Region ist<br />

stark landwirtschaftlich geprägt, wobei sich dort vermehrt<br />

Industrie entwickelt.<br />

Die Untersuchungen fanden in Zusammenarbeit<br />

mit dem Indian Institute of Technology in Roorkee/Nordindien<br />

(IIT-Roorkee) im Rahmen eines<br />

deutsch-indischen Partnerprojektes statt. Das Hauptanliegen<br />

dieses Projektes besteht in der Anwendung<br />

geophysikalischer Methoden zur Abschätzung der<br />

Gefährdung der Aquifersysteme der Region um Roorkee.<br />

In der landwirtschaftlich geprägten Region um Roorkee<br />

ist der Einsatz von Düngemittel sowie die Bewässerung<br />

der Felder mit Abwasser gängige Praxis,<br />

wobei das Abwasser durch offene, teilweise nicht befestigte,<br />

Kanäle transportiert wird. Gleichzeitig bezieht<br />

die umliegende Landbevölkerung ihr Frischwasser<br />

oftmals aus Brunnen, die aus oberflächennahen<br />

Aquiferen gespeist werden. Eine Gefährdung dieser<br />

oberflächennahen Aquifersysteme durch den praktizierten<br />

Umgang mit Abwasser und Düngemittel ist<br />

nicht auszuschließen.<br />

Ein weiteres Risiko geht von Mülldeponien aus, die<br />

zu Grundwasserkontamination in deren unmittelbaren<br />

Umgebung führen können und deren schädlicher<br />

Einfluss auf die Aquifersysteme der Region ebenfalls<br />

nicht auszuschließen ist. Die Grundwasserproblematik<br />

wird, durch den Frischwasserbedarf der zunehmenden<br />

Bevölkerung und der wachsenden Industrie,<br />

verschärft.<br />

<strong>Geophysikalische</strong> Erkundungsmethoden haben sich<br />

in ihrer Anwendung auf hydrogeologische Fragestellungen,<br />

sowie zur Untersuchung von Altlasten mehrfach<br />

bewährt ([Tezkan, 1999], [Recher, 2002] und [Seher,<br />

2005]). Insbesondere eignet sich die Kombination<br />

von DC und RMT, da sich beide Methoden in ihrem<br />

Erkundungstiefenbereich und ihrer Sensitivität<br />

gegenüber leitfähigen Strukturen ergänzen.<br />

Im Rahmen des Projektes wurde das umliegende Gebiet<br />

einer Mülldeponie nahe dem Dorf Saliyar mit diesen<br />

Methoden profilweise vermessen. Seit 1975 wird<br />

dieses Gebiet zur landwirtschaftlichen Nutzung an<br />

Kleinbauern verpachtet und intensiv mit Abwasser<br />

bewässert, welches von Roorkee durch Rohrleitungen<br />

auf die Felder transportiert wird.<br />

In Abbildung 1 ist eine Übersichtskarte der Region<br />

dargestellt. Das Messgebiet bei Saliyar befindet sich<br />

ca. 6 km nordwestlich von Roorkee auf der rechten<br />

Uferseite des Solani. Das ungeklärte Abwasser der<br />

Stadt wird von der Mahingram Pumpstation über<br />

eine Kanalleitung zur Mülldeponie geleitet. Das vermessene<br />

Referenzgebiet befindet sich auf der linken<br />

Uferseite des Solani und ist ca 10 km von Saliyar entfernt.<br />

Das ganze Gelände hat ein leichtes Gefälle in<br />

Richtung Südosten.<br />

Der Einfluss von Abwassernutzung auf die Grund-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

153


googlemaps<br />

Abbildung 1: Übersicht der Region um Roorkee. Das Messgebiet Saliyar ist rot, das Referenzgebiet grün<br />

umrandet. Weitere wichtige Plätze sind farblich hervorgehoben und beschriftet. Brunnen, die überwacht und<br />

untersucht wurden sind als weiße Kringel mit roten Punkten dargestellt (Rechte Abbildung: [Singhal et al.,<br />

2003]).<br />

wasserqualität wurde bisher überwiegend unter Berücksichtigung<br />

von hydrochemischen Daten untersucht.<br />

Die chemischen und bakteriellen Analysen des<br />

Grundwassers zeigen bisher nur erhöhte Werte für die<br />

Proben aus der unmittelbarem Umgebung von Saliyar,<br />

wo sich die Mülldeponie befindet. Singhal et al.<br />

[2003] schließen eine Gefährdung der Aquifersysteme<br />

weiter flussabwärts in Richtung Roorkee auf Grund<br />

der angenommenen Grundwasserfließrichtung. Sudha<br />

et al. [2010] haben bereits in den Gebieten nördlich<br />

von Roorkee, in Saliyar und in Khanjapur, eine Vielzahl<br />

von DC-Profilen und „Time Domain Electromagnetic“<br />

(TEM) Stationen vermessen, um den Einfluss<br />

des Abwassers auf das oberflächennahe Aquifer zu detektieren<br />

und die Ausmaße des kontaminierten Bereichs<br />

zu bestimmen.<br />

Ziel dieser Arbeit ist es ebenfalls, den Einfluss der<br />

Kontamination auf die Leitfähigkeitsverteilung der<br />

oberen 20 − 25 m zu bestimmen. Hierzu wurden die<br />

auf dem kontaminierten Gebiet vermessenen Profile<br />

mit dem Referenzprofil verglichen, auf dem keine<br />

Kontamination erwartet wird. Durch die Vielzahl der<br />

vermessenen Profile soll auch eine Aussage über die<br />

laterale Verbreitung des kontaminierten Bereichs, sowie<br />

die Ausbreitung von Kontaminationsfahnen getroffen<br />

werden.<br />

Zusätzlich wurden die Datensätze beider Methoden<br />

mittels des 2D Joint Inversionsprogrammes<br />

RMTDC2D von Candansayar und Tezkan [2008] einzeln<br />

und gemeinsam invertiert. Es soll dabei untersucht<br />

werden, inwiefern die Anwendung der 2D<br />

Joint Inversion verlässlichere und einheitlichere Untergrundmodelle<br />

zu liefern vermag.<br />

Geologie des Messgebiet<br />

Der indische Subkontinent lässt sich geologisch in drei<br />

Bereiche unterteilen: den Himalaja im Norden, die<br />

Gangesebene und die indische Halbinsel im Süden.<br />

Die Gangesebene stellt dabei das Entwässerungsbecken<br />

der großen indischen Flüsse Ganges, Indus,<br />

Brahmaputra und Yamuna dar. Das Messgebiet befindet<br />

sich im nördlichen Teil der Gangesebene (Abbildung<br />

1). Da es sich um Schwemmland des Ganges<br />

und des Solani handelt, weist der Boden überwiegend<br />

ungehärtete alluviale Sedimente jüngeren Alters<br />

mit einer einfachen eindimensionalen Schichtung auf.<br />

Diese Sedimente bestehen aus abwechselnden Schichten<br />

von Ton, Sand, Kunkur und in manchen Gebieten<br />

Kies, wobei Kunkur ein nodulares Kalziumkarbonat<br />

ist, welches sich häufig in Semi-ariden Regionen<br />

bildet. Lithologische Diagramme bei [Singhal et al.,<br />

Bohrlochinformation DC Ergebnisse<br />

Lithologie z (m) ρ (Ωm) z (m)<br />

Sandiger Lehm 0-4 >100 0-7<br />

Sand 4-32 40-100 7-38<br />

Ton + Kies 32-45 30-50 38-45<br />

Sand 45-105 — —<br />

Tabelle 1: Bohrlochinformationen verglichen mit den<br />

DC Ergebnissen in Sherpur von Sudha et al. [2010].<br />

2003] zeigen eine oberflächennahe Schicht aus sehr<br />

feinem bis sandigem Lehm von 3 − 6 m Mächtigkeit.<br />

Die darunterliegende Sandschicht in 3 − 27 m Tiefe<br />

bezeichnet den ersten unbegrenzten, oberflächennahen<br />

Aquifer. Darunter befindet sich eine nichtpermeable<br />

Ton-Kies-Schicht von ca. 13 m Mächtigkeit<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

154


googlemaps<br />

Abbildung 2: Die vermessenen Profile sind rot dargestellt und tragen die Bezeichnung der Profilnummer<br />

(PR1 bis PR13) sowie den Zusatz RMT oder RMT-DC, je nachdem ob nur RMT oder zusätzlich DC gemessen<br />

wurde. (Ab-)Wasserkanäle sind blau und Hochspannungsleitungen schwarz gestrichelt dargestellt. Der<br />

Hauptabwasserkanal verläuft von 1 nach 2. Auslässe für das Abwasser befinden sich entlang des Hauptabwasserkanals<br />

ca. alle 50m. An den Orten S1, S2 und S3 wurden Wasserproben entnommen.<br />

und trennt den oberen Grundwasserleiter von dem<br />

tieferen, der überwiegend aus Sand und Kies besteht.<br />

In der Tabelle 1 sind die DC-Inversionsergebnisse von<br />

Sudha et al. [2010] mit den Bohrlochdaten verglichen.<br />

Die Bohrlochdaten wurden von einem Litholog auf<br />

dem Referenzprofil bei Sherpur abgeleitet und zeigen<br />

gute Übereinstimmung mit den DC-Ergebnissen.<br />

Messung bei Roorkee<br />

Insgesamt wurden 13 RMT und 8 DC Profile bei einer<br />

zweiwöchigen Kampagne im Februar 2009 vermessen<br />

(Abbildung 2). Zu dieser Jahreszeit war das Messgebiet<br />

nur teilweise kultiviert und daher gut zugänglich.<br />

Die Profile sind, ausgehend von der Mülldeponie<br />

am südlichen Bildrand bis zum Flussbett des Solani,<br />

überwiegend parallel angeordnet. Der Abstand der<br />

Profile betrug ca. 50 m, je nach verfügbarem Platz.<br />

Damit wurde eine ca. 200 × 600 m 2 große Fläche abgedeckt.<br />

Das ganze Gebiet nordwestlich vom Hauptabwasserkanal<br />

hat ein leichtes Gefälle in Richtung<br />

Nordosten zum Solani hin. Die Mülldeponie im unteren<br />

Bildteil umfasst eine Fläche von 170 × 100 m 2 .<br />

Eine Messung direkt auf der Mülldeponie war nicht<br />

möglich, deswegen wurden Profil 8 und Profil 10 seitlich<br />

positioniert um einen etwaigen Einfluss der Mülldeponie<br />

durch z.B. Sickerwasser zu ermitteln.<br />

Der Hauptabwasserkanal verläuft von der Mülldepo-<br />

nie nordöstlich in Richtung des Solani. Ungefähr alle<br />

50 m befinden sich Wasserauslässe, um die Felder<br />

zu fluten. Der kleinere offene Kanal (hellblau dargestellt)<br />

im linken Bildbereich verläuft parallel zum<br />

großen und wurde von den Profilen gekreuzt. Der Bereich<br />

oberhalb von Profil 6 war stark bewachsen und<br />

konnte nicht vermessen werden. Das oberste Profil<br />

(Profil 7) verläuft im trockenen Flussbett des Solani<br />

und ist ca. 600 m von der Kontaminationsquelle entfernt,<br />

weswegen hier eine Abnahme der Kontamination<br />

zu erwarten ist. Der Bereich zwischen Profil 3 und<br />

Profil 6 war stark bewässert, konnte aber dennoch<br />

vermessen werden. Profil 1, welches im Folgenden explizit<br />

vorgestellt wird, wurde am ersten Tag mit einer<br />

Länge von 710 m und einem Stationsabstand von<br />

10 − 50 m vermessen, um den Einfluss der Kontamination<br />

mit der Entfernung von der Quelle abzuschätzen.<br />

Die RMT-Messung wurde mit dem RMT-F<br />

Gerät der Universität zu Köln durchgeführt. Es arbeitet<br />

im erweiterten Frequenzbereich von 10 kHz bis<br />

1 MHz und ermöglicht dadurch ein verbessertes Auflösungsvermögen<br />

oberflächennaher Schichten [Wiebe,<br />

2007]. Der Messpunktabstand für die RMT-Stationen<br />

betrug bei allen anderen Profilen 10 m und die Anzahl<br />

in der Regel 21 Stationen. Gemessen wurden beide<br />

horizontalen Komponenten des elektrischen und<br />

des magnetischen Feldes zur Bestimmung der spezifischen<br />

Widerstands- und Phasenwerte der TE- und<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

155


(a) (b)<br />

(c) (d)<br />

Abbildung 3: Inversionsmodelle des Referenzprofils 11 bei Sherpur der Daten der Wennerauslage (a), der<br />

Daten der TM-Mode (b), der der TE-Mode (c), sowie beider Datensätze (TE+TM) gemeinsam invertiert<br />

(d). Die RMT Stationen sind als schwarze Dreiecke dargestellt. Die 0.2 Isolinie des DOI ist als schwarz<br />

gestrichelte Linie geplottet und stellt die maximale Erkundungstiefe dar.<br />

der TM-Mode. Für die DC wurde eine ABEM Terrameter<br />

SAS 1000 benutzt. Als Messauslage wurde<br />

eine Wenner und eine Schlumberger „Short-Long“-<br />

Auslage gewählt. Durch den geringeren Elektrodenabstand<br />

im mittleren Bereich der Auslage ist eine<br />

oberflächennah bessere Auflösung gewährleistet.<br />

Eine geologische Streichrichtung zur Ausrichtung der<br />

Profile lässt sich anhand der Kontaminationsausbreitung<br />

nur ungenügend festlegen. Zu erwarten wäre<br />

eine Ausbreitung der Kontamination von den Abwasserauslässen<br />

Richtung Nordwesten oder auch eine<br />

Verbreitung der Kontamination von der Mülldeponie<br />

ausgehend.<br />

Demzufolge und in Anbetracht der nahezu eindimensionalen<br />

Struktur des Untergrundes wurden die Profile<br />

anhand der verfügbaren Sender für die RMT und<br />

der Verfügbarkeit von Platz ausgerichtet.<br />

Inversionsergebnisse Referenzprofils 11 bei<br />

Sherpur<br />

In den Abbildungen 3 sind die Inversionsmodelle der<br />

einzelnen 2D Inversionen für die Wennerauslage und<br />

für die RMT Daten mit DC2DINVRES [Günther,<br />

2004] bzw. RUND2_NLCG2_FAST (RUND2) [Rodi<br />

und Mackie, 2001] dargestellt. Die nicht-sensitiven<br />

Bereiche sind in den Inversionsergebnissen ausgeblendet<br />

und der „Depth Of Investigation“-Index (DOI)<br />

nach Oldenburg und Li [1998] ist zur Abschätzung der<br />

maximalen Erkundungstiefe in den RMT-Modellen<br />

dargestellt.<br />

Das DC Inversionsergebnis zeigt wie erwartet eine<br />

hochohmige Deckschicht mit einem spezifischen Widerstand<br />

von ρ>200 Ωm und einer Mächtigkeit von<br />

5 − 7 m, welche als sandige Lehmschicht identifiziert<br />

werden kann. Im Bereich von 0 − 60 m war das Profil<br />

sehr trocken, da dieser Bereich nicht als landwirtschaftliche<br />

Nutzfläche diente. Der Bereich von 60-<br />

120 m war kultiviert und wurde wahrscheinlich regelmäßig<br />

künstlich bewässert, wodurch sich der erniedrigte<br />

Widerstand der Deckschicht bis zu einer<br />

Tiefe von ungefähr 2 m erklären lässt. Der Bereich<br />

danach zeigt wieder eine durchgehende hochohmige<br />

Deckschicht, obwohl dieser ebenfalls kultiviert war.<br />

Entsprechend den lithologischen Informationen aus<br />

Tabelle 1 und den DC Ergebnissen von Sudha et al.<br />

[2010] zeigt das Inversionsmodell unter der Deckschicht<br />

eine leitfähige Schicht mit einem spezifischen<br />

Widerstand von 30−50 Ωm, welches das oberflächennahe<br />

Aquifer, also eine wassergesättigte Sandschicht,<br />

darstellt und als die aufzulösende Zielstruktur aufgefasst<br />

wird.<br />

Die Inversionsmodelle der Daten der TM-Mode, der<br />

TE- Mode und der Joint Inversion beider Datensätze<br />

zeigen qualitativ eine identische Leitfähigkeitsstruktur.<br />

Auch die Abnahme des Widerstands im mittleren<br />

Teil des Profils ist identisch zum DC Inversionsmodell.<br />

Die hochohmige Deckschicht wird allerdings mit einem<br />

spezifischen Widerstand von 100 − 200 Ωm rekonstruiert,<br />

welcher verglichen mit dem DC-Ergebnis<br />

zu gering ist. Die darunterliegende Zielstruktur wird<br />

wiederum mit einem spezifischen Widerstand von<br />

20 − 40 Ωm aufgelöst und ist vergleichbar mit dem<br />

DC-Ergebnis. Zusätzlich wurde ein zweites RMT Profil<br />

in ca. 200 m Entfernung von Profil 11 vermessen.<br />

Die geologische Struktur ist hier identisch und der<br />

spezifische Widerstand der Zielstruktur wird ebenfalls<br />

mit einem Wert von 20 − 40 Ωm gut aufgelöst<br />

Yogeshwar [2010].<br />

Eine Erklärung für den verminderten Widerstand der<br />

Deckschicht ist die geringe Sensitivität der RMT für<br />

hochohmige Bereiche. Desweiteren könnte ein statischer<br />

Versatz der ρa-Daten den geringen Widerstand<br />

der Deckschicht erklären. Oberflächennahe Inhomogenitäten<br />

in der Nähe der Empfängerelektroden kön-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

156


nen z.B. einen zusätzlichen Versatz auf die ρa-Daten<br />

bewirken.<br />

Die Unterkante des Aquifers wird in den Inversionsmodellen<br />

nicht aufgelöst. Zum einen hängt das<br />

mit der maximalen Erkundungstiefe zusammen, zum<br />

anderen ergibt sich wahrscheinlich kein entscheidender<br />

Leitfähigkeitskontrast zwischen dem oberflächennahen<br />

Aquifer und der darunterliegenden nichtpermeablen<br />

Ton-Kies-Schicht, deren spezifischer Widerstand<br />

ebenfalls ungefähr 40 − 50 Ωm beträgt.<br />

Um die maximale Erkundungstiefe abzuschätzen<br />

wurde der DOI nach Oldenburg und Li [1998] berechnet.<br />

Für die Wennerauslage ergibt sich bei [Yogeshwar,<br />

2010] eine maximale Erkundungstiefe von 20 m<br />

bis 25 m für den mittleren Bereich des Profils.<br />

Für die RMT-Modelle ist der DOI als schwarz gestrichelte<br />

Linie in den Modellen dargestellt. Für das<br />

Inversionsmodell der Daten der TM-Mode ist die Erkundungstiefe<br />

ungefähr 15 − 20 m, wobei im Bereich<br />

von x = 100 − 200 m der DOI, auf Grund der erhöhten<br />

Leitfähigkeit, etwas abnimmt. Für die beiden<br />

anderen RMT-Inversionsmodelle in Abbildung 3 ergibt<br />

sich eine ähnliche Erkundungstiefe von ca. 15 m.<br />

Inversionsergebnisse des kontaminierten Profils 1<br />

bei Saliyar<br />

Das Profil 1 befindet sich in ca 100 m Entfernung<br />

zur Mülldeponie und wurde am ersten Messtag vermessen<br />

(Abbildung 2). Es war nicht bewässert und<br />

kaum kultiviert. Benutzt wurde eine Wenner Long-<br />

Auslage mit einer Länge von 120 m, im Gegensatz zu<br />

den anderen DC Profilen, die ausschließlich eine Profillänge<br />

von 200 m haben. Das RMT-Profil wurde bis<br />

zu einer Entfernung von 710 m vom Hauptabwasserkanal<br />

(Abbildung 2) fortgesetzt um den Einfluss der<br />

Kontamination mit dem Abstand zur Quelle zu bestimmen,<br />

wobei der Stationsabstand im Bereich von<br />

x =0− 200 m zehn Meter betrug und dann bis zu<br />

50 m vergrößert wurde. Ausgewertet wurden nur die<br />

Daten der TE-Mode, da die der TM-Mode viele Senderlücken<br />

und teilweise negative Phasenwerte aufwiesen.<br />

In den Abbildungen 4(a) und 4(b) sind die Inversionsmodelle<br />

der einzelnen 2D Inversionen für die Wennerauslage<br />

und für die Daten der TE- Mode dargestellt.<br />

Das DC-Modell zeigt eine einfache, eindimensionale<br />

geologische Schichtung, wobei die Deckschicht im<br />

Vergleich zum Referenzprofil in Abbildung 3 wesentlich<br />

weniger mächtig ist und einen geringeren Widerstand<br />

von ca. 30 Ωm aufweist. Eine mögliche Ursache<br />

für die insgesamt weniger hochohmige Deckschicht<br />

mit ca. 2-3 m Mächtigkeit ist wahrscheinlich die sonst<br />

häufige Bewässerung mit Abwasser. Im rechten Bereich<br />

nahe des Hauptabwasserkanals befindet sich ein<br />

hochohmiger Bereich der Deckschicht, welcher von einer<br />

Erdaufschüttung herrühren könnte, und der in<br />

dem RMT Inversionsmodell identisch rekonstruiert<br />

ist. Darunter folgt das kontaminierte Aquifer, bzw.,<br />

verglichen mit dem Ergebnis für das Referenzprofil,<br />

ein Bereich stark erhöhter Leitfähigkeit bis zu<br />

einer Tiefe von ungefähr 12 m. Dies ist im RMT-<br />

(a)<br />

(b)<br />

Abbildung 4: Inversionsergebnis des Profils 1 bei Saliyar<br />

der Daten der Wennerauslage (a) und der Daten<br />

der TE-Mode (b). Die RMT Stationen sind als<br />

schwarze Dreiecke dargestellt, der DOI als schwarz<br />

gestrichelte Linie.<br />

Profil ebenfalls identisch wiedergegeben. Der spezifische<br />

Widerstand fällt im DC-Modell mit 6 − 15 Ωm<br />

etwas höher aus als der des RMT-Modells. Die Anpassung<br />

beider Datensätze ist mit einem RMS < 2.5<br />

gut.<br />

Die maximale Erkundungstiefe für das DC-Modell ergibt<br />

sich mit dem DOI-Index bei Yogeshwar [2010]<br />

zu 15 − 20 m im mittleren Bereich des Profils. Die<br />

schwarze Linie in Abbildung 4 stellt die maximale<br />

Erkundungstiefe von 10 − 15 m für das RMT-Modell<br />

dar. Mit dieser Erkundungstiefe für die RMT wird<br />

die Unterkante des kontaminierten Bereichs nur mit<br />

der DC aufgelöst.<br />

Auf den vermessenen Profilen in Saliyar wurde mit<br />

dem Programm EMUPLUS [Wiebe, 2007] eine 1D<br />

RMT-DC und eine 1D RMT-DC Joint Inversion im<br />

jeweiligen Mittelpunkt des Profils durchgeführt um<br />

die Leitfähigkeit des kontaminierten Aquifers genauer<br />

zu bestimmen. Die 1D Marquardt Inversionen für<br />

Profil 1 (Abbildung 5) entsprechen sehr gut den 2D<br />

Ergebnissen im Mittelpunkt des Profils. Das Modell<br />

der Geoelektrik zeigt einen Dreischichtfall mit gut<br />

aufgelösten Schichtgrenzen und spezifischen Widerständen,<br />

außer für den der obersten Schicht. Die äquivalenten<br />

Modelle in Abbildung 5 unterscheiden sich<br />

also kaum. Ebenso zeigt das Ergebnis für die RMT-<br />

Daten nahezu identische äquivalente Modelle, was auf<br />

eine gute Auflösung beider Schichten schließen lässt.<br />

Beide Inversionsergebnisse stimmen gut überein, wobei<br />

die RMT die Unterkante des Aquifers nicht auflösen<br />

kann.<br />

Das Modell der 1D Joint Inversion passt die Datensätze<br />

beider Methoden mit einem RMS von 1.65 gut<br />

an und löst die Unterkante der zweiten Schicht durch<br />

Hinzunahme der DC Daten auf. Der Widerstand des<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

157


z (m)<br />

0<br />

5<br />

10<br />

15<br />

Profile 1, Schlumberger, VES at x=60m<br />

equi. models<br />

RMS=1.39<br />

20<br />

0 10 20 30 40 50 60 70<br />

ρ (Ωm)<br />

(a)<br />

z (m)<br />

0<br />

5<br />

10<br />

15<br />

Profile 1, te, sounding at x=60m<br />

equi. models<br />

RMS=1.75<br />

20<br />

0 10 20 30 40 50 60 70<br />

ρ (Ωm)<br />

(b)<br />

z (m)<br />

Profile 1, x=60m, Joint Inversion RMT/DC<br />

0<br />

5<br />

10<br />

15<br />

RMS all =1.65,<br />

RMS dc =1.4,<br />

RMS rmt =2.1, cf=0.82<br />

20<br />

0 10 20 30 40 50 60 70<br />

ρ (Ωm)<br />

Abbildung 5: 1D Marquardt Inversionsmodelle für die Sondierungskurve der Schlumbergerauslage (a) und<br />

der Daten der TE-Mode (b), sowie der Joint Inversion beider Datensätze (c) bei x =60m in der Mitte des<br />

Profils 1 bei Saliyar. Aufgetragen ist der spezifische Widerstand gegen die Tiefe. Äquivalente Modelle sind<br />

grün und der Mittelwert der äquivalenten Modelle sind rot dargestellt.<br />

kontaminierten Bereichs ist mit ρ ≈ 10 Ωm ziemlich<br />

genau bestimmt. Da weitaus mehr DC-Daten vorhanden<br />

sind, stellt das Inversionsergebnis zwar einen Informationsgewinn<br />

für die RMT dar, entspricht aber<br />

ziemlich genau dem 1D Modell der DC Inversion.<br />

Nur oberflächennah liefern die hochfrequenten RMT-<br />

Daten Zusatzinformation zur Verbesserung der Auflösung<br />

der obersten Schicht.<br />

2d Joint Inversion<br />

In der Abbildung 6 sind die Inversionsmodelle der<br />

Wenner Long Auslage, der Daten der TE-Mode und<br />

das Ergebnis der 2D Joint Inversion beider Datensätze<br />

mit RMTDC2D dargestellt. Die Leitfähigkeitsstruktur<br />

der Endmodelle ist gut vergleichbar mit den<br />

Inversionsmodellen der einzelnen 2D Inversionen mit<br />

DC2DINVRES und RUND2.<br />

Die Modelle der einzelnen Inversionen sind bis zu einer<br />

Tiefe von ungefähr 10 m gut vergleichbar, wobei<br />

die Anpassung der RMT-Daten mit einem RMS von<br />

1.9 besser ist als die der DC-Daten. Der kontaminierte<br />

Bereich wird im RMT Inversionsmodell etwas leitfähiger<br />

rekonstruiert. Eine Aussage über die Unterkante<br />

des Aquifers lässt sich auf Grund der mangelnden<br />

Eindringtiefe der RMT nicht treffen, wohingegen<br />

die Unterkante im DC-Modell aufgelöst wird. Die 2D<br />

Joint Inversion in Abbildung 6 passt beide Datensätze<br />

mit einem RMS von 5.0 zufriedenstellend an. Die<br />

Unterkante und der Widerstand des Aquifers sind gut<br />

aufgelöst.<br />

Die mit dem mittleren Tiefenbereich korrespondierenden<br />

DC-Daten in Abbildung 7(b) sind über die<br />

ganze Profillänge relativ gleichmäßig,<br />

z (m)<br />

z (m)<br />

−5<br />

0<br />

5<br />

10<br />

15<br />

(c)<br />

Profile 1/Saliyar, Wenner long, ρ 0 =9 Ω m, It.=9 Alpha=5 RMS=3.0816<br />

20<br />

0 10 20 30 40 50<br />

x (m)<br />

60 70 80 90 100<br />

5<br />

−5<br />

0<br />

5<br />

10<br />

15<br />

(a)<br />

Profile 1/Saliyar, TE, ρ 0 =9 Ω m, It.=30 Alpha=5 RMS=1.9128<br />

20<br />

0 10 20 30 40 50<br />

x (m)<br />

60 70 80 90 100<br />

5<br />

(b)<br />

Profile 1/Saliyar, TE/Wenner long, ρ =9 Ω m, It.=12 Alpha=10 RMS=5.0686<br />

0<br />

−5<br />

100<br />

z (m)<br />

0<br />

5<br />

10<br />

15<br />

20<br />

0 10 20 30 40 50<br />

x (m)<br />

60 70 80 90 100<br />

5<br />

(c)<br />

Abbildung 6: Inversionsergebnis der Wenner Long-<br />

Auslage (a), der TE-Mode (b) und der 2D Joint Inversion<br />

(c) des Profils 1. Die RMT-Stationen sind als<br />

schwarze Dreiecke dargestellt.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

158<br />

100<br />

65<br />

42<br />

28<br />

18<br />

12<br />

8<br />

65<br />

42<br />

28<br />

18<br />

12<br />

8<br />

ρ (Ω m)<br />

100<br />

65<br />

42<br />

28<br />

18<br />

12<br />

8<br />

ρ (Ω m)<br />

ρ (Ω m)


ρ a (Ωm)<br />

ρ a (Ωm)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

45 °<br />

Profile 1, TE−mode,RMS=6.52, f=18000 Hz<br />

ρ a,mess<br />

ρ a,calc<br />

φ mess<br />

φ calc<br />

20<br />

0 20 40 60<br />

x (m)<br />

80 100 120<br />

Profile 1, TE−mode,RMS=6.52, f=666000 Hz<br />

ρ<br />

a,mess<br />

ρ<br />

a,calc φ<br />

mess<br />

φ<br />

calc<br />

80<br />

45 °<br />

0 20 40 60<br />

x (m)<br />

(a)<br />

80 100 120<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

φ (°)<br />

φ (°)<br />

n−level<br />

n−level<br />

n−level<br />

1<br />

3<br />

6<br />

8<br />

10<br />

12<br />

1<br />

3<br />

6<br />

8<br />

10<br />

12<br />

1<br />

3<br />

6<br />

8<br />

10<br />

12<br />

Meas. Pseudo−section:Wenner<br />

Calc. Pseudo−section: Wenner<br />

0 20 40 60<br />

x (m)<br />

(b)<br />

80 100 120<br />

ρ a (Ω m)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

ρ a (Ω m)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

%Misfit Section: Wenner (RMS=4.0584) %MISFIT<br />

10<br />

Abbildung 7: Daten und Datenanpassung der Daten der TE-Mode (a) und der Wenner Long Auslage (b) für<br />

das Ergebnis der 2D Joint Inversion mit RMTDC2D. Die gemessenen ρa und ϕ Werte sind als schwarze,<br />

bzw. rote Punkte und die Modellantwort als schwarze, bzw. rote Kreuze für f =18kHz, 666 kHz gegen die<br />

Station aufgetragen. Die DC-Daten und die Modellantwort sind als Pseudosektionen dargestellt, wobei die<br />

untere Abbildung (b) die relative Differenz beider zeigt.<br />

wohingegen die RMT Daten der 18 kHz Frequenz in<br />

Abbildung 7(a) und das Inversionsmodell der einzelnen<br />

2D RMT Inversion in Abbildung 6(b) auch Bereiche<br />

erhöhter Leitfähigkeit zeigen. Bei der 2D Joint<br />

Inversion werden die RMT-Daten daher nicht so gut<br />

wie die DC-Daten angepasst. Die Modellantwort der<br />

RMT ist durch die Regularisierung und dem Einfluss<br />

der insgesamt gleichmäßigeren DC-Daten sehr glatt.<br />

Der RMS für die RMT-Daten ergibt sich zu 6.5. Die<br />

Hinzunahme der RMT-Daten führt zu einer besseren<br />

Bestimmung der Deckschicht vor allem im etwas<br />

hochohmigeren Bereich von x =80− 100 m. Die DC-<br />

Daten in diesem Bereich sind nicht so gut angepasst,<br />

wie in der Darstellung der relativen Differenz der Daten<br />

und der Modellantwort in Abbildung 7(b) erkennbar<br />

ist. Eine Verringerung des Glättungsparameters<br />

passt diese zwar besser an, führt allerdings zu Inversionsartefakten<br />

für die tieferen Bereiche.<br />

Qualitativ zeigen die Auflösungsmatrizen der 2D<br />

Joint Inversion bei Yogeshwar [2010] eine verbesserte<br />

Auflösung im Vergleich zur RMT-Einzelinversion,<br />

wobei die Randbereiche im Vergleich zur DC-<br />

Einzelinversion besser aufgelöst sind, da die DC dort<br />

keine Tiefeninformation liefert. Die maximale Erkundungstiefe<br />

wird durch die DC-Daten bestimmt, weswegen<br />

keine erhöhte Sensitivität der tieferen Strukturen<br />

durch Hinzunahme der RMT-Daten erreicht wird.<br />

Abschätzung der Kontaminationsverbreitung<br />

Um die lateralen Ausmaße des kontaminierten Bereichs<br />

abzuschätzen, wurden insgesamt 9 parallele<br />

RMT und 5 parallele DC Profile auf einem ca 200 ×<br />

600 m 2 großen Gebiet vermessen (Abbildung 2). Die<br />

pr6<br />

pr5<br />

pr4<br />

pr3<br />

pr2<br />

pr1<br />

pr9<br />

pr10<br />

pr7<br />

Waste<br />

Solani<br />

Kanal<br />

0<br />

−10<br />

Abbildung 8: Ansicht aller RMT-Profile von der<br />

Mülldeponie bis zum Solani. Der Hauptabwasserkanal<br />

verläuft von der Mülldeponie in Richtung des Solani<br />

und ist rot dargestellt.<br />

RMT-Profile sind in Abbildung 8 als Flächenschnitte<br />

der xz-Ebene dargestellt. Die Profile verlaufen parallel<br />

zwischen der Mülldeponie und dem Solani. Die<br />

RMT-Modelle liefern auf Grund der teilweise mäßigen<br />

Datenqualität nicht so glatte und gleichmäßige<br />

Modelle wie die DC-Modelle bei Yogeshwar [2010].<br />

In der Darstellungen 8 ist eine leichte Abnahme der<br />

Leitfähigkeit zum Solani hin zu beobachten. Das Profil<br />

7 am Solani in Abbildung 8 zeigt insgesamt hochohmigere<br />

Bereiche im Inversionsmodell für die RMT.<br />

Die Inversionsergebnisse der DC-Daten sind bei Yo-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

159


z (m)<br />

0<br />

5<br />

10<br />

15<br />

20<br />

Profile 1/3/5/11, Joint Inversion model<br />

1D Joint Pr11<br />

1D Joint Pr1<br />

1D Joint Pr3<br />

1D Joint Pr5<br />

reference site ρ=50 Ωm<br />

10 100<br />

(a)<br />

ρ (Ωm)<br />

z (m)<br />

0<br />

5<br />

10<br />

15<br />

20<br />

Profile 7/8/9/13/11, Joint Inversion model<br />

1D Joint Pr11<br />

1D Joint Pr7<br />

1D Joint Pr8<br />

1D Joint Pr9<br />

1D Joint Pr13<br />

reference site ρ=50 Ωm<br />

10 100<br />

(b)<br />

ρ (Ωm)<br />

z (m)<br />

Profile 2/4/6/10/11, RMT Inversion modells<br />

0<br />

5<br />

10<br />

15<br />

20<br />

1D Joint Pr11<br />

1D RMT Pr2<br />

1D RMT Pr4<br />

1D RMT Pr6<br />

1D RMT Pr10<br />

reference site ρ=50 Ωm<br />

10 100<br />

Abbildung 10: 1D Marquardt Joint Inversionsmodelle der DC- und RMT-Daten im jeweiligen Mittelpunkt<br />

der Auslage für Profil 1,3,5 (a) und Profil 7,8,9,13 (b), sowie einzelne RMT 1D Inversionsmodelle für Profil<br />

2,4,6,10 (c). Das 1D Joint Inversionsmodell des Referenzprofils (Profil 11) ist rot dargestellt.<br />

geshwar [2010] mit denen der RMT sehr gut vergleichbar,<br />

wobei die Unterkante des Aquifers jedoch mit der<br />

DC aufgelöst wird. Um eine genauere Vorstellung<br />

z (m)<br />

z (m)<br />

Relative difference between model result of pr11 and pr1: Wenner ρ 01 =120 Ωm, ρ 02 =15 Ωm<br />

−5<br />

0<br />

5<br />

10<br />

15<br />

0.81<br />

0.64<br />

0.46<br />

0.29<br />

0.11<br />

−0.06<br />

20<br />

−0.24<br />

−0.41<br />

25<br />

−0.59<br />

30<br />

0 20 40 60<br />

x (m)<br />

80 100<br />

−0.76<br />

120<br />

−5<br />

0<br />

5<br />

10<br />

15<br />

20<br />

25<br />

(a)<br />

Relative difference between model results of pr11 and pr1: ρ 01 =50 Ωm, ρ 02 =10 Ωm<br />

30<br />

0 20 40 60 80 100 120<br />

x (m)<br />

(b)<br />

0.99<br />

0.79<br />

0.63<br />

0.47<br />

0.31<br />

−0.51<br />

relative difference<br />

relative difference<br />

0.14<br />

−0.02<br />

−0.18<br />

−0.34<br />

Abbildung 9: Dargestellt ist die relative Differenz<br />

der RMT- (a) und der DC-Inversionsmodelle (a) des<br />

Referenzprofils 11 und des kontaminierten Profils 1<br />

(Relative Differenz > 0 ⇒ ρref >ρpr1).<br />

des Leitfähigkeitsunterschiedes zwischen dem Inversionsmodell<br />

des Referenzprofils und den Profilen auf<br />

dem kontaminierten Bereich zu erhalten, wurde für<br />

(c)<br />

ρ (Ωm)<br />

jedes Profil die relative Differenz beider Modelle berechnet.<br />

Die relative Differenz des Profils 1 mit dem<br />

Referenzprofil 11 in Abbildung 9(a) zeigt für die Geoelektrik<br />

über die gesamte Auslagenlänge einen wesentlich<br />

geringeren spezifischen Widerstand. In einer<br />

Tiefe von ca. 15 m befindet sich die Unterkante des<br />

kontaminierten Aquifers und die Widerstände werden<br />

vergleichbar, bzw. der des Profils 1 wird größer. Für<br />

die RMT zeigt die Abbildung 9(b) ein identisches Ergebnis.<br />

Auch hier ist der spezifische Widerstand des<br />

Inversionsmodells für das Referenzprofil weitaus höher.<br />

Bei beiden Ergebnisses zeigen nur kleine oberflächennahe<br />

Bereiche der Deckschicht ein anderes Verhalten.<br />

In der Darstellung der relativen Differenz bei<br />

Yogeshwar [2010] zeigen die vermessenen Profile auf<br />

dem kontaminierten Gebiet in Saliyar im Vergleich<br />

mit dem Referenzprofil ausschließlich erhöhte Leitfähigkeitswerte<br />

bis zu einer Tiefe von ca. 15 − 20 m, bis<br />

auf das Profil 7, welches sich am Ufer des Solani befand<br />

und am weitesten von der Kontaminationsquelle<br />

entfernt war.<br />

In Abbildung 10(a) und 10(b) sind die 1D Joint Inversionsmodelle<br />

der Daten aller Profilmittelpunkte in<br />

Saliyar mit dem des Referenzprofils verglichen. Insgesamt<br />

zeigen alle Profile unterhalb einer etwas hochohmigeren<br />

Deckschicht einen spezifischen Widerstand<br />

von ρ ≈ 10 Ωm für das kontaminierte Aquifer, wiederum<br />

gefolgt von einer Schicht mit einen erhöhten spezifischen<br />

Widerstand. Das in grün geplottete Modell im<br />

rechten Teil der Abbildung 10 entstammt dem Pro-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

160


y (m)<br />

y (m)<br />

600<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

600<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

xy−planeview of RMT profiles at the depth of z=1m<br />

0 50 100<br />

x (m)<br />

150 200<br />

(a)<br />

xy−planeview of RMT profiles at the depth of z=10m<br />

0 50 100<br />

x (m)<br />

150 200<br />

(c)<br />

y (m)<br />

y (m)<br />

600<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

600<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

xy−planeview of RMT profiles at the depth of z=5m<br />

0 50 100<br />

x (m)<br />

150 200<br />

(b)<br />

xy−planeview of RMT profiles at the depth of z=15m<br />

pr7<br />

pr6<br />

pr5<br />

pr4<br />

pr3<br />

pr2<br />

pr1<br />

pr9<br />

pr10<br />

canal<br />

waste<br />

0 50 100<br />

x (m)<br />

150 200<br />

2<br />

Abbildung 11: xy-Schnitte der RMT Inversionsmodelle von RUND2 für verschiedene Tiefen: 1 m (a), 5 m<br />

(b),10 m (c) und 15 m (d). Die Kreuze bezeichnen die RMT-Stationen.<br />

fil 7 am Solani und zeigt eine etwas andere Struktur<br />

mit höheren Widerständen, wie schon in Abbildung<br />

8 ersichtlich war. Die Modelle der 1D RMT Inversionen<br />

in Abbildung 10(c) zeigen ebenfalls geringere<br />

spezifische Widerstände für das Aquifer. Das Ergebnis<br />

von Profil 4 in Abbildung 5(c) zeigt den Beginn<br />

einer hochohmige Schicht in ca. 8 m Tiefe, die auch im<br />

2D Modell der TM-Mode zu finden ist, aber nicht in<br />

dem der TE-Mode, und auf mangelnde Datenqualität<br />

oder die Streichrichtung zurückzuführen ist [Yogeshwar,<br />

2010]. Die Ergebnisse der einzelnen 2D Inversionsmodelle<br />

wurden für beide Methoden linear auf ein<br />

dreidimensionales Gitter interpoliert. Um eine Aussage<br />

über die laterale Verbreitung der Kontamination<br />

mit der Tiefe zu treffen, sind in Abbildung 11, am<br />

Beispiel der RMT, Flächenschnitte in der xy-Ebene<br />

für vier verschiedene Tiefen, z=1 m, 5 m, 10 m und<br />

1 5m, dargestellt.<br />

In einer Tiefe von einem Meter erkennt einen sehr<br />

hochohmigen Bereich direkt in der Umgebung der<br />

(d)<br />

ρ(Ωm)<br />

300<br />

172<br />

Mülldeponie. Das gesamte Profil 10, Profil 9 und die<br />

ersten 50 m direkt am Kanal von Profil 1 zeigen diese<br />

hochohmige Deckschicht. Dies könnte von einer Erdaufschüttung<br />

im Zusammenhang mit der Mülldeponie<br />

herrühren. Genaueres ist jedoch nicht bekannt.<br />

Für z = 5 m zeigt die Flächendarstellung spezifischen<br />

Widerstände um die 10 Ωm im Bereich von y =0m<br />

bis 300 m. Die Leitfähigkeit ist in der Nähe der Mülldeponie<br />

erhöht und nimmt dann in Richtung des Solani<br />

etwas ab. Das Profil 7 am Solani ist wie bereits<br />

gesagt über den gesamten Tiefenbereich hochohmiger<br />

als die anderen Profile.<br />

In einer Tiefe von 10 m und 15 m lässt die Kontamination<br />

etwas nach und es zeichnet sich die Unterkante<br />

des Aquifers ab, wobei der Bereich nahe der Mülldeponie<br />

im Vergleich zum Rest leitfähiger ist. Dies lässt<br />

wieder auf einen Einfluss der Mülldeponie schließen,<br />

vor allem da die Leitfähigkeit der Profile in der unmittelbaren<br />

Nähe der Deponie bis zu einer größeren<br />

Tiefe erhöht ist. So ist für Profil 9 in 15 m keine Un-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

161<br />

99<br />

56<br />

32<br />

19<br />

11<br />

6<br />

3


terkante des kontaminierten Aquifers erkennbar. Die<br />

DC-Ergebnisse bei Yogeshwar [2010] bestätigen die<br />

der RMT, wobei die Unterkante des kontaminierten<br />

Bereichs sich in 15 m wesentlich deutlicher bei den<br />

DC-Ergebnissen abzeichnet.<br />

Eine solche flächenhafte Darstellung ist auf Grund<br />

der Interpolation zwischen den Profilen kritisch zu<br />

bewerten. Zwischen dem letzten Profil und dem Ufer<br />

des Solani lässt sich keine verlässliche Aussage treffen,<br />

da sich die Entfernung auf ca. 300 m beläuft. Problematisch<br />

sind auch Inversionsartefakte, die bei der<br />

Interpolation zwischen den Profilen überakzentuiert<br />

werden. Trotzdem eignet sich die Flächendarstellung<br />

zur Visualisierung der lateralen Leitfähigkeitsverteilung<br />

und insbesondere, um ein räumliches Bild des<br />

Untergrundes zu gewinnen.<br />

Diskussion und Ausblick<br />

Die Inversionsmodelle, der bei der Mülldeponie aufgenommenen<br />

Datensätze, haben in der Regel einander<br />

ähnliche Strukturen. Unterhalb einer etwas hochohmigeren<br />

Deckschicht zeigen diese, im Vergleich zu<br />

dem Inversionsmodell des Referenzprofils, Werte des<br />

spezifischen Widerstands kleiner als 10 Ωm bis zu einer<br />

Tiefe von ungefähr 15 m. Dieser Bereich konnte<br />

mit dem oberflächennahen Grundwasserleiter identifiziert<br />

werden und der im Vergleich geringere Wert<br />

des spezifischen Widerstands kann auf den Einfluss<br />

des Abwassers zurückgeführt werden.<br />

Aus der flächenhaften Darstellung der Inversionsmodelle<br />

konnte die laterale Verbreitung der Kontamination<br />

abgeleitet werden. Erkennbar ist eine Zunahme<br />

des spezifischen Widerstands mit der Entfernung<br />

von der Mülldeponie, so dass diese vermutlich auch<br />

als Kontaminationsquelle angenommen werden kann.<br />

Auch reichen die Werte verringerter spezifischer Widerstände<br />

nahe der Deponie vergleichsweise tief.<br />

Darüber hinaus sind die spezifischen Widerstände der<br />

RMT-Inversionsmodelle in der Nähe des Hauptabwasserkanals,<br />

wo sich die Wasserauslässe befinden,<br />

ebenfalls verringert. In ungefähr 600 m Entfernung<br />

von der Mülldeponie im Flussbett des Solani sind die<br />

Werte des spezifischen Widerstands wiederum deutlich<br />

höher und vergleichbar mit denen des Referenzprofils.<br />

Insgesamt lassen sich die Mülldeponie und das Abwasser<br />

als Kontaminationsquellen bestätigen und es<br />

kann eine Systematik der Kontaminationsverbreitung<br />

mit dem Abstand von den Quellen abgeleitet werden.<br />

Das Programm RMTDC2D [Candansayar und Tezkan,<br />

2008] lieferte vergleichbare Inversionsmodelle<br />

wie die einzelnen Inversionen der Datensätze mit<br />

DC2DINVRES [Günther, 2004] und RUND2 [Rodi<br />

und Mackie, 2001]. Insbesondere die Inversionsmodelle<br />

der einzelnen Inversionen mit RMTDC2D waren<br />

den Inversionsmodellen der anderen Programmen<br />

sehr ähnlich.<br />

Bei Inversionsmodellen, die in ihrer Struktur starke<br />

Unterschiede zeigten, war die Joint Inversion schwierig,<br />

da beide Datensätze nicht gleichzeitig anzupassen<br />

waren. Trotzdem konnte mit der Jointinversion<br />

ein einheitliches Modell beider Datensätze gefunden<br />

werden, was eine nützliche Hilfe bei der Interpretation<br />

der Ergebnisse darstellt.<br />

Vor allem konnte die Unterkante des kontaminierten<br />

Bereichs durch Hinzunahme der DC-Daten aufgelöst<br />

werden und gleichzeitig lieferten die RMT-Daten<br />

Zusatzinformation an den Rändern der Profile, wo<br />

mit der DC keine Tiefenaussage zu treffen möglich<br />

war. Die Joint Inversion führte unter diesem Gesichtspunkt<br />

zu verbesserten Endmodellen.<br />

Problematisch gestaltete sich allerdings die Auswirkung<br />

des Regularisierungsparameters. Bei einem zu<br />

geringen Wert ergaben sich stark überstrukturierte<br />

Inversionsmodelle und Inversionsartefakte.<br />

Schwierig war auch die Bewertung der Inversionsmodelle<br />

anhand der Auflösungsmatrizen. Die Auflösungsmatrizen<br />

der Joint Inversion waren mit denen<br />

der einzelnen Inversionen von RMTDC2D nicht<br />

vergleichbar und eine quantitative Aussage über das<br />

verbesserte Auflösungsvermögen der Joint Inversion<br />

konnte daher nicht getroffen werden.<br />

Im Rahmen des DFG-DST Projektes, welches diese<br />

Arbeit ermöglicht hat, ist eine weitere Messung auf<br />

dem selben Gebiet vorgesehen. Hierbei soll die „Transient<br />

Electromagnetic“ Methode (TEM) zum Einsatz<br />

kommen. Mit dieser Methode besteht die Möglichkeit<br />

tiefer liegende Strukturen aufzulösen. Wünschenswert<br />

wäre mit dieser Methode den Einfluss der<br />

Kontamination auf den tiefer liegenden Grundwasserleiter<br />

zu untersuchen, da sich die hier vorliegende<br />

Arbeit auf die Erkundung des oberen Grundwasserleiter<br />

konzentriert hat.<br />

Des Weiteren eignen sich die verwendeten Methoden<br />

gut für die Erkundung von kontaminierten Böden<br />

und könnten gerade in Indien vermehrt zum Einsatz<br />

kommen um damit zur Verbesserung der Grundwasserversorgung<br />

beizutragen.<br />

Interessant wäre auch die Einbeziehung der geophysikalischen<br />

Ergebnisse in eine hydrogeologische Interpretation,<br />

um die Auswirkung der Kontamination auf<br />

größeren räumlichen und auch zeitlichen Skalen abzuschätzen.<br />

Die 2D Joint Inversion stellt ein zusätzliches Hilfsmittel<br />

zur Interpretation von RMT- und DC-<br />

Inversionsmodellen dar. Die einzeln erhaltenen Inversionsmodelle<br />

lassen sich durch die Hinzunahme der<br />

Joint Inversionsmodelle leichter qualitativ und quantitativ<br />

bewerten. Eine weitere Anwendung dieses Programmes<br />

ist daher wünschenswert.<br />

Literatur<br />

Candansayar, M. und B. Tezkan, Two-dimensional<br />

joint inversion of radiomagnetotelluric and direct current<br />

resistivity data, Geophysical Prospecting, 56, 737–<br />

749, 2008.<br />

Günther, T., Inversion Methods and Resolution Analysis<br />

for the 2D/3D Reconstruction of Resistivity Structures<br />

from DC Measurements, Dissertation, Technischen<br />

Universität Bergakademie Freiberg, 2004.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

162


Oldenburg, D. W. und Y. Li, Estimating the depth<br />

of investigation in DC resistivity and IP surveys, Geophysics,<br />

64, 403–416, 1998.<br />

Recher, S., Dreidimesionale Erkundung von Altlasten<br />

mit Radiomagnetotellurik - Vergleiche mit geophysikalischen,<br />

geochemischen und geologischen Analysen an<br />

Bodenproben aus Rammkernsondierungen, Dissertation,<br />

Universität zu Köln, Institut für Geophysik und<br />

Meteorologie, 2002.<br />

Rodi, W. und R. L. Mackie, Nonlinear conjugate gradients<br />

algorithm for 2D magnetotelluric inversion, Geophysics,<br />

66, (1), 174–187, 2001.<br />

Seher, T., Untersuchung von Feuchtbiotopen in Ostfriesland:<br />

Gefährdungsabschätzung mit Multielektroden-<br />

Geoelektrik und Radiomagnetotellurik, Diplomarbeit,<br />

Universität zu Köln, Institut für Geophysik und Meteorologie,<br />

2005.<br />

Singhal, D., T. Roy, H. Joshi und A. Seth, Evaluation<br />

of Groundwater Pollution in Roorkee Toen, Uttaranchal,<br />

Journal Geological Society of India, 62, 465–<br />

477, 2003.<br />

Sudha, B.Tezkan, M.Israil, D. Singhal und J.Rai,<br />

Geoelektrical mapping of aquifer contamination: a case<br />

study from Roorkee, India, Near Surface Geophysics, 8,<br />

33–42, 2010.<br />

Tezkan, B., A review of environmental applications of<br />

quasi-stationary electromagnetic techniques, Surveys<br />

in Geophysics, 20, 279–308, 1999.<br />

Wiebe, H., 1D-Joint-Inversion von Geoelektrik und Radiomagnetotellurik,<br />

Diplomarbeit, Universität zu Köln,<br />

Institut für Geophysik und Meteorologie, 2007.<br />

Yogeshwar, P., Grundwasserkontamination bei Roorkee/Indien:<br />

2D Joint Inversion von Radiomagnetotellurik<br />

und Gleichstromgeoelektrik Daten, Diplomarbeit,<br />

Universität zu Köln, Institut für Geophysik und Meteorologie,<br />

2010.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

163


Site Effect Assessment in the Mygdonian Basin (EUROSEISTEST area, Northern Greece)<br />

using RMT and TEM Soundings<br />

Widodo 1,2) , Marcus Gurk 2) , Bülent Tezkan 2)<br />

1) Adhi Tama Institute of Technology Surabaya (ITATS), Indonesia<br />

2) Institute of Geophysics and Meteorology, University of Cologne, Germany<br />

Abstract<br />

During the project “Euroseistest Volvi-Thessaloniki”, a strong-motion test site<br />

(EUROSEISTEST) for Engineering Seismology was installed in the Mygdonian Basin between the<br />

two lakes Volvi and Lagada ca. 45 km northeast of Thessaloniki (Northern Greece). The basin itself<br />

is a neotectonic graben structure (5 km wide) with increased seismic activity along distinct normal<br />

fault patterns. Fluvioterrestrial and lacustrien sediments (approximately 350-400 m thick) are<br />

overlying the basement consisting of gneiss with schist. To improve the seismic wave propagation<br />

model it is vital to know about site effects, e.g. the geotectonic properties of the area such as the topof-basement,<br />

vertical tectonic boundaries (faults and basement fracturation) and the geothermal<br />

regime. Therefore, we carried out near surface EM studies to understand the distribution of the<br />

active faulting and the top of basement structure of this particular area.<br />

The RMT (Radiomagnetotelluric) and TEM (Transient electromagnetic) measurements were<br />

carried out on three profiles and thirty sounding. The inverted RMT and TEM data show generally a<br />

four layer model. The layers are indicated as metamorphic and sediment rocks, which are in detail:<br />

marly silty sand with gravel (>> 100 Ωm), marly silty sand with clay (50 - 100 Ωm), sandy clay<br />

(30 – 50 Ωm) and silty sand (10 - 30 Ωm). Due to the high resistivity of the top layer, the skin depths<br />

of the RMT soundings are around 35 m. The TEM data gives detail information of the lower structure<br />

down to a depth of 200 m. According to our analysis, a normal fault next to the Euroseistest could be<br />

located having a strike direction of N 60 E. The joint interpretation of RMT and TEM data proves to<br />

be an effective tool to investigate complex geology structures.<br />

Introduction<br />

This study refers to the Thessaloniki area, which has recently been affected by the 1978<br />

destructive earthquake sequence [Papazachos and Papazachou, 1997]. It has been well established<br />

that the strong ground motion of such a seismic event causes irregularly distributed modification to<br />

the local geology [Tranos and Mountrakis, 1998], [Raptakis, et. al., 1999]. Different geophysical<br />

methods have been applied in this area [Thanassoulas, et.al.,1987], [Raptakis,et.al.,1999],<br />

[Savvaidis,et.al.,1999], [Tranos, et.al, 2003], [Gurk, et.al., 2007], however detail information of fault<br />

structures has not been verified so far. Ambient noise measurements from the area east of the<br />

Euroseistest experiment give strong implication for a complex 3-D tectonic setting. Therefore we<br />

carried out near surface EM studies to understand the distribution of the active faulting and the top of<br />

basement structure of this particular area.<br />

Joint TEM and RMT inversion has been successfully applied to geological and engineering<br />

problems in the past [Tezkan et.al., 1995], [Schwinn, 1999], [Steuer, 2002]. The RMT method has<br />

low penetration depth and therefore gives information of the surface layers, whereas the TEM<br />

method gives detail information of the lower structure of the investigated area. Hence, a joint<br />

interpretation of RMT and TEM data will produce a good resolution of resistivities and thicknesses, in<br />

shallow and in deeper parts of the subsurface.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

164


Test Area and Geological Setting<br />

The investigated sites are located between the Lagada and the Volvi Lake in Northern Greece. The<br />

area is now covered with alluvial deposits as shown in (Fig.1). The local geological mapping in this<br />

area was done by the Geology Institute the University of Cologne [Maith, 2009]. The Mygdonian<br />

system corresponds to various types of sediments that were deposited in a quaternary graben<br />

structure. Due to its hydrothermal activity, the top was intensively covered with travertine/tufa<br />

deposits, now mostly removed by erosion [Jongmans, et.al. 1998]. The local geological map of the<br />

area is shown in Fig.1. We find four major units: metamorphic basement formed by schist and<br />

gneisses in the deeper depths, lower terrace deposit was deposited on the top of the basement<br />

metamorphic, lacustrine and deltaic sediments including conglomerate, gravel and sand. Fans<br />

comprise of soil and silt are located between lower terrace deposit and the Holocene deposit is<br />

composed of sand, silt and clay which are deposited on the top of these sediments. To obtain detail<br />

information of the geology in this area, TEM and RMT data were observed on all various types of the<br />

sediments (Fig.1).<br />

Lower<br />

Terrace Dep.<br />

Fans<br />

Holocene<br />

Deposit<br />

Profile Direction of RMT<br />

Profile 1 of RMT<br />

Gneiss-Schist<br />

Fig. 1. Site Location, geology of the study area, location of RMT station ( ), TEM station ( ) and<br />

location of boreholes ( ) are also displayed in the figure GTEM-1 at the location of TEM-1<br />

sounding.<br />

RMT Measurements and Interpretation<br />

GTEM-1<br />

S-1<br />

Profile 3 of RMT<br />

Profile 2 of RMT<br />

The RMT measurements were carried out on three profiles as indicated in Figure 1. Parameter of<br />

the survey design is listed in Table 1. The RMT of profile 1 is along 1600 m with direction N 0 S,<br />

whereas the direction of profile 2 is located N60 E, which are parallel with transmitters. The direction<br />

of profile 3 in this area is N60W with length of 1400 m, which are perpendicular with transmitter. Due<br />

to partial inaccessibility of the area the profiles could not be set up in a straight line. The RMT-F<br />

system consists out of two magnetic sensors (induction coils, 30 cm length), a preamplifier for the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

165


electrical channels, two electrical antennae and a recorder. It is a four channel instrument (Ex, Ey,<br />

Hx, Hy) with the capability of estimating the full impedance tensor [Tezkan and Saraev, 2008].<br />

Table1. Parameter of RMT survey design<br />

RMT Length Site distance [m] Orientation Transmitter<br />

Profile [m]<br />

azimuth<br />

Profile 1 1600 25 N0oS N60oE Profile 2 1000 25 N60oE N60oE Profile 3 1400 25 N60oW N60oE A special feature is the way how the electric field is sampled. Instead of grounded dipoles,<br />

the device uses symmetrical electrical dipoles (e.g. two arms of 20 m length) that are capacitively<br />

coupled to the ground. The system records time series in two bands: D2 band (10-100 kHz) and D4<br />

band (100 kHz—1 MHz). The coherency level is important to prevent signals with high noise level,<br />

generally we used for this study a coherency value of 0.8.<br />

N S<br />

Silty Sand<br />

Marly Salty Sand<br />

Fault structure<br />

(????)<br />

Sandy clay<br />

Silty Sand<br />

Fig 2. The RMT 2-D inversion model of profile 2 indicates metamorphic rock (marly salty<br />

sand) in a depth of 0 -5 m. Stations 1 – 16 and stations 25 – 38 show a very low resistivity (


Fault Structure<br />

(???)<br />

Fig.3. The Fault structure direction (black arrow) derived from 2-D RMT inversions<br />

ρ (Ωm)<br />

Fig.4. Correlation between the 1-D inversion of the TEM data (GTEM1) at the reference site and<br />

borehole S-1 information<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

167


In order to calibrate the geophysical measurements with the lithology, TEM data at our<br />

reference site was correlated with borehole data in S-1 (Fig.1). This analysis allows identifying<br />

characteristic strata based on the electrical conductivity. The modeling of RMT data was done by<br />

using 2-D inversion techniques (Fig. 2 and Fig.3). The 2-D inversion was performed with the 2-D<br />

Mackie code [Mackie, et.al, 1997], the 1-D inversion was done with the Marquardt [Cerv and Pek,<br />

1979] and Occam algorithm [Constable, 1978]. Prior to any 2-D modeling, the penetration depth<br />

have been estimated using the ρ* (z*) transformation [Schmucker, 1979]. Penetration depths are<br />

found to be around 35 m.<br />

As an example, we explain the RMT model of profile 2 in detail. This model (Fig.2) shows<br />

high resistivities on the top layer (more than 100 Ω m) which correlates to the marly silty with clay of<br />

the borehole data (Fig.4). For stations 1 -16 and stations 25-38 there is a low resistive structure<br />

beneath the surface layer. Between stations 17-24 there is a high resistive structure with the same<br />

resistivities as the surface layer. This region is interpreted as a fault structure (Fig.2), which is filled<br />

with sedimentary rock (sandy clay).<br />

The data quality and the consistency between ρa and φ could be checked in the apparent<br />

resistivity curves, which are shown in Fig.5. RMT station 8 on profile 2 show that the resistivity<br />

values decrease from highest resistivity (> 80 Ωm) to low resistivity (30 Ωm) with phase values less<br />

than 50 °. The sounding curves of station 8 and 13 (figure 5) and station 16 (figure 5) on profile 2<br />

shows that the phases exceeds value more than 45° indicating a low resistive (sediment) structure<br />

at larger depths, meanwhile the good conductive metamorphic series is indicated by the phase<br />

values lower than 45. Fig. 6 shows a comparison between measured and calculated data at<br />

selected frequencies (79 kHz, 387 kHz and 784 kHz). Measured and synthetic data fit well together<br />

keeping in mind the geological complexity and inhomogeneous in this survey area, however some<br />

misfits which are associated as 3-D effect.<br />

The direction of the fault can be constructed using the 2-D inversion models of profile 1-3 as<br />

shown in Fig.3. The direction of the fault structure is found to be N60E.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

168


N60E<br />

N60E<br />

N60E<br />

Fig.5. Comparison of measured and calculated RMT data of stations 8, 13 and 16 on profile 2<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

169


Fig.6. Comparison of measured and calculated RMT data at frequencies 79, 387 and 784 kHz<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

170


TEM soundings<br />

In order to overcome the resolution problem of TEM soundings for near surface structures,<br />

we carried out RMT soundings at each TEM location to use this additional information in the<br />

following joint inversion. The distance between the TEM stations is depending on the accessibility of<br />

the area. Thirty loop-loop measurements (Tx: 50 m x 50 m and Rx: 10 m x 10 m) were performed<br />

during this campaign. TEM data were obtained using the Zonge NT 20 transmitter and the Zonge<br />

GDP32 receiver.<br />

c<br />

a b<br />

Time (s)<br />

Time (s)<br />

0.01<br />

0.0001<br />

1e-06<br />

1e-08<br />

1e-10<br />

e<br />

Fig.7. The segmentation recording scheme for the in-loop Zonge manufactured TEM system;<br />

low-gain Nano TEM (a), high-gain Nano TEM (b), automatic-gain Zero TEM (c), combination of alltransient<br />

(d), TEM sounding of station 1 was processed with deconvulation and inversion (e).<br />

The circle indicate unreliable decay curves for very late times that we address to be an A/D<br />

conversion problem of the Zonge device.<br />

d<br />

Time (s)<br />

Time (s)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

171


The system allows to measure TEM data in two distinct modes, according to different<br />

investigation depths: Nano TEM with low and high gain and Zero TEM measurements (Fig.7). Both<br />

modes give unreliable decay curves for very late times that we address to be an A/D conversion<br />

problem of the Zonge device. To cope with this problem we simply changed polarity of the receiver<br />

loop. Consequently a time consuming data acquisition procedure was required.<br />

The first measurement for shallow investigation was done in NANO TEM low gain mode.<br />

This mode uses low currents (about 0.3 A) which is switched of quickly. Moreover, the ramp time<br />

were taken from 0.3 μs to 2.5 ms with manual low-gain settings. Whereas, Nano TEM high-gain uses<br />

a higher transmitter current (about 3 A), which gives a higher penetration depth with a saturation for<br />

earlier time. Nano TEM low-gain and high-gain uses 12 Volt power supply. To obtain higher<br />

penetration depths, Zero TEM measurements were carried out from 31 μs to 6 ms time window with<br />

accompanying automatic-gain settings. It uses a relatively high transmitter current (10 A) at 24 Volts<br />

and 50-55 μs turn-off ramp time.<br />

The penetration depth of TEM methods depends on the time after the transmitters current is<br />

switched off [Parasnis, 1986]. The diffusion process of the transient electromagnetic induction field<br />

can be visualized using the smoke ring concept [Nabighian, 1979]. Due to the low conductivity<br />

(metamorphic rock), the maximal penetration depth (δT) of TEM sounding is approximately 200-300<br />

meter. The next step is a deconvolution of the data that was done by the EADEC algorithm (Lange,<br />

2002) followed by a 1-D inversion with EMUPLUS (e.g. Scholl, 2005) (Fig.7 (e)). The 1-D<br />

interpretation of the TEM data also shows the lateral boundary of the fault structure at stations 12- 17<br />

on profile 1 (Fig.8).<br />

Fig.8. One-dimensional TEM model on profile 1<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

172


Joint Interpretation of RMT and TEM data<br />

A joint interpretation can be used to get a better resolution of the model parameter and this<br />

will enhance the possibility to detect the active fault structure. Generally we conducted RMT<br />

soundings at each TEM site location so that we can estimate a jointly inverted model at each TEM<br />

location (Fig.9).<br />

a).<br />

b).<br />

c).<br />

Table2. The importance parameter of station 10<br />

RMT Imp TEM Imp Joint Imp<br />

ρ1[Ωm] 322.46 0.99 367.98 0.57 247.99 0.99<br />

ρ2[Ωm] 158.17 0.99 367.98 0.88 344.23 0.27<br />

h1[m] 5.56 0.97 9.75 0.97 9.15 0.84<br />

RMS[%] 4.90 3.10 3.20<br />

Table 3. Model resistivies obtained from the<br />

RMT and TEM data for the selected stations<br />

Station RMT TEM Joint<br />

10 322.46 367.98 344.23<br />

11 332.20 273.45 303.06<br />

12 450.73 372.20 616.36<br />

13 562.36 271.88 531.33<br />

14 363.93 420.46 865.49<br />

17 199.91 260.49 208.14<br />

26 4.47 23.48 26.93<br />

25 28.71 15.83 32.98<br />

Fig. 9. (a) One-dimensional model section derived by RMT data (b). One-dimensional model<br />

section of TEM data (c) Joint interpretation of TEM and RMT data. The comparison between<br />

TEM, RMT and Joint on station 10, 11,12,14,17 and 25 are shown with ellipses.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

173


The 2-D model of the RMT data (Fig.2) gives information about the top structure of this area,<br />

but the deeper structures and the fault system cannot be resolved in detail solely based on RMT<br />

soundings. On the other hand, TEM shows a good resolution of deeper structure. The joint<br />

interpretation/inversion process was done using the Marquardt algorithm. The interpretation of the 1-<br />

D RMT model between station 13 – 25 resolves of the top layer (Fig 9 (a)) and the TEM 1-D model<br />

(Fig.9b) shows the bottom of boundary layers of fault structure (Fig.8) in this area more distinctly,<br />

however the top layers of TEM stations 13-25 could not be resolved well because the top layer has<br />

high resistivity value (> 80 Ω m) addressed as metamorphic rock from our reference borehole.<br />

Joint inversion can increase the number of model parameter that includes some of which the<br />

methods cannot resolve separately (Fig.9c). The importance parameters were calculated for the<br />

model resistivities and thickness [Jupp and Vozoff, 1975] [Tezkan et.al., 1995], [Schwinn, 1999],<br />

which are plotted in Table 2. It shows the parameter (ρ1, ρ2, h1) as a result of the individual<br />

single/joint inversions. Values between 0 (unimportant) and 1 (important) are calculated. From this<br />

analysis we conclude that the resistivity values and thicknesses of the fault structure (Fig.9) is well<br />

resolved at station 10, 11,12,14,17 and 25; however, at station 13, it is hardly resolved (importance<br />

0.65). The resistivity value at station 26 is weakly resolved with importance parameter 0.12. The<br />

model resistivities of the second layer (between 5-20 m) obtained from 1-D models of TEM and RMT<br />

data are in good agreement with the joint inversion models (Table 3).<br />

Conclusion<br />

The Inversion of RMT and TEM data indicates a normal fault structure with a strike direction<br />

of approximately N 60 E. The RMT and TEM models generally show four layers that can be<br />

separated according to their resistivity values. We address them as to be metamorphic and sediment<br />

rocks, which are marly silty sand with gravel (>> 100 Ω m), marly silty sand with clay (50 - 100 Ω m),<br />

sandy clay (30 – 50 Ω m) and silty sand (10-30 Ω m). The RMT data was interpreted using 2-D<br />

inversions technique that results in a good fitting between observed and calculated data. Due to the<br />

high resistivity of the top layer, the skin depths of the RMT soundings are around 35 m. The TEM<br />

data gives detail information of the lower structure down to a depth of 200 m.<br />

This study was financed by the Marie Curie project: IGSEA – Integrated Nonseismic<br />

Geophysical Studies to Assess the Site Effect of the EUROSEISTEST Area in Northern Greece –<br />

PERG03-GA-2008-230915 {REF RTD REG/T.2 (2008)D/596232}.<br />

References:<br />

Constable C.Steven, Parker L.R.,Constable G.C., Occam’s Inversion: A practical algorithm for generating<br />

smooth models from electromagnetic sounding data, Geophysics vol.52 no.3 P280-300, March, 1978.<br />

Cerv, V. and Pek, J., Solution of one-dimensional magnetotelluric problem. Stud. Geophys. Geod., 23:349,<br />

1979.<br />

Gurk.M, Savvaidis A.S., Bastani M., Tufa Deposit in the Mygdonian Basin (Northern Greece) studied with<br />

RMT /CSTAMT, VLF & Self-Potential, EMTF Kolloqium, June, 2007.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

174


Jongmans, D., Patilakis,K., Demanet, D., Raptakis, D., Hor-rent, C., Tsokas, G., Lontzetidis, K., Reipl, J.,<br />

EURO-SEISTEST: Determination of The geological Stucture of The Volvi Basin and Validation of The Basin<br />

Respone. Bull. Seismoll.Soc. Am. 88, 473 – 487, 1998.<br />

Jupp, D.L.B. and K. Vozoff, stable iterative methods for the inversion of geophysical data,<br />

Geophys.J.R.astr.soc., 42,957-976, 1975.<br />

Lange J., Joint Inversion von central Loop TEM und Long TEM Transient am Beispiel von Messdaten aus<br />

Israel , Diplomarbeit,Universität zu Köln, Institut fur Geophysik und Meteorologie (2003), 16,31,51, 2002.<br />

Maith, I., Erläuterungen zur Kartierung der kristallinen Randbereiche der paläo-mesozoischen Abfolge des<br />

quartären-Makedonischen Massivs und der quartären Beckenfüllung im Bereich von Stivos (Mygdonisches<br />

Becken, NE-Greichenland), Diplomkartierung, der Mathematisch-Naturwissenschaftlichen Fakultät der<br />

Universität zu Köln, 2009.<br />

Mackie R., Rieven S., Rodi W., User Manual and Software Documentation for two-dimensional of<br />

magnetotelluric data, Cambridge Massachuesetts, USA, 1997.<br />

Nabighian, M.N., Quasy-static transient response of a conducting half space: An approximate representation.<br />

Geophysichs, 44:1700-1705, 1979.<br />

Papazachos,B.C., Papazachou, C., The Earth quake of Greece, Ziti Publications , Thessaloniki, 1997.<br />

Parasnis, D. S., Principle of applied geophysics (4 th ed.): Chapman and Hall, London, 402p, 1986.<br />

Raptakis D., F.J. Chavez-Garcia, Makra K., Pitilakis K., Site efecst at Eurositetest – I. Determination of the<br />

valley structure and confrontation of observations with 1-D analysis, Soil dynamic and earthquake<br />

Engineering, 19 1 -22, 1999.<br />

Savvaididis , A., Pedersen, L.B., Tsokas, G.N.,Dawes, G.J., Structure of the Mygdonian Basin (N.Greece)<br />

inferred from MT and gravity data, Tectonophysics, 317, 171-1886, 2000.<br />

Scholl C., The influence of multidimensional structures on the interpretation of LOTEM data with onedimensional<br />

models and the application to data from Israel, Inaugural Dissertation, Institute Geophysik und<br />

Meteorologie Uni Zu Koln, 2005.<br />

Schwinn W., 1-D Joint Inversion Radiomagnetotellurik (RMT) und Transientelektromagnetik Daten (TEM):eine<br />

Anwendung zur Grundwasser exploration in Grundfor, Danemark, Diplomarbeit, Institute Geophysik und<br />

Meteorologie Uni Zu Koln, Juli, 1999.<br />

Steuer, A., Kombinierte Auswertung von Messungen mit Transient-Elektromagnetik und Radio Magnetotellurik<br />

zur Grundwassererkundung im Bechen von Quarzazate (Maroko), Diplomarbeit, Institute Geophysik und<br />

Meteorologie Uni Zu Koln, June, 2002.<br />

Schmucker, U., Erdmagnetische Variationen und die elektrische Leitfahigkeit in tieferen Schichten der Erde.<br />

Sitzungsber. Mitt. Braunschw Wiss. Ges. Sonderh., 4:45-102, 1979.<br />

Tranos, M.D., Mountrakis D.M., Neotectonic joints of the northern Greece; their significance on the<br />

understanding of the active deformation. Bulletin of Geological Soceitey of Greece 32,209-219, 1998.<br />

Thanassoulas, C., Tselentis, G-A., Traganos, G., A preliminary resistivity investigation (VES) of the Lagada<br />

hot springs area in northern Greece, Gheothermics, Vol. 16, NO. 3, pp. 227-238, 1978.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

175


Tranos D.M., Papadimitriou E.E., Kilias A.A.,Thessaloniki-Gerakarou Fault Zone (T<strong>GFZ</strong>): the western<br />

extension of the 1978 Thessaloniki earthquake fault (Northern Greece) and seismic hazard assessment,<br />

Journal of structural geology, 25 2109-2123, 2003.<br />

Tezkan, B., M. Goldman, S. Greinwald, A.Hordt, I.Muller, F.M. Neubauer and H.G. Zacher, A joint application<br />

of radio magnetotellurics and transient electromagnetic to the investigation of a waste deposit in Cologne<br />

(Germany), Applied Geophysics, 34, 199-212,1995.<br />

Tezkan, B., and Saraev, A., A new broadbrand Radiomagnetotelluric instrument: Application to near surface<br />

investigation, Near surface Geophysichs, 6.243-250, 2008.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

176


Die deutsche Nordseeküste im Fokus von aeroelektromagnetischen<br />

Untersuchungen<br />

Teilgebiete Langeoog mit Wattenmeer und Elbemündung<br />

Gerlinde Schaumann 1 , Annika Steuer 2 , Bernhard Siemon 2 , Helga Wiederhold 1 & Franz Binot 1<br />

1. Einleitung<br />

1 Leibniz-Institut für Angewandte Geophysik (LIAG), Hannover<br />

2 Bundesanstalt für Geowissenschaften und Rohstoffe (BGR), Hannover<br />

gerlinde.schaumann@liag-hannover.de<br />

Hubschrauberelektromagnetische (HEM) Untersuchungen bieten ein großes Potential für die flächendeckende<br />

Kartierung der Sedimente der ersten hundert Meter des Untergrundes (Siemon et al., 2009).<br />

Sie sind für hydrogeologische Fragestellungen von großer Bedeutung, da mit Hilfe des spezifischen<br />

Widerstands die Verteilung sandiger und tonhaltiger Sedimente im Untergrund sowie Versalzungszonen<br />

und Süßwasserbereiche ermittelt werden können.<br />

In den Jahren 2008 und 2009 wurden in Kooperation von LIAG und BGR im Rahmen des LIAG-<br />

Projektes zur „Flächenhaften Befliegung“ und des „D-AERO“-Projektes der BGR gemeinsam aerogeophysikalische<br />

Erkundungen zu Salz-/Süßwasserfragestellungen in insgesamt fünf Messgebieten<br />

im Bereich der deutschen Nordseeküste durchgeführt (Wiederhold et al. 2008, Steuer et al. 2009).<br />

Das Messgebiet Langeoog umfaßt die Ostfriesischen Inseln Langeoog und Spiekeroog und das angrenzende<br />

Wattenmeer, das Messgebiet Glückstadt umfaßt den Bereich der Elbemündung nordwestlich<br />

von Hamburg (Abb. 1).<br />

Abbildung 1: Lage der Messgebiete Langeoog und Glückstadt (nordwestlich von Hamburg).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

177


Für die Insel Langeoog waren die Lage und die Ausdehnung der Süßwasserlinse zu ermitteln. Im küstennahen<br />

Bereich erhoffte man im Wattenmeer mutmaßliche Süßwasseraustritte zu finden. Im Bereich<br />

der Elbemündung sollten die grundwasserführenden Schichten und mögliche Versalzungszonen<br />

kartiert werden.<br />

Die Daten dienen als Grundlage für die Planung und Arbeit in vielfältigen ökonomischen und ökologischen<br />

Bereichen, wie z.B. Raumplanungen und der Entwicklung von Wassernutzungs- und Wasserschutzkonzepten.<br />

Sie werden über das Fachinformationssystem (FIS) Geophysik des LIAG nutzbar<br />

sein.<br />

Das eingesetzte Messsystem wird bei Steuer et al. (2010) beschrieben. Die Karten des scheinbaren<br />

spezifischen Widerstands a und der Schwerpunktstiefe z* (Siemon, 2009) für die verschiedenen<br />

Messfrequenzen geben einen ersten Überblick über die Leitfähigkeitsstrukturen in den Messgebieten.<br />

Die Erkundungstiefe nimmt dabei mit abnehmender Frequenz und Leitfähigkeit zu. Die HEM-<br />

Messungen liefern die Verteilung der elektrischen Leitfähigkeit bis maximal 150 m Tiefe. Dabei können<br />

Salzwasser und Süßwasser oder ton- und sandhaltige Sedimente unterschieden werden. Hier dargestellt<br />

sind erste Ergebnisse in Form von Karten auf dem Bearbeitungsstand nach der Grundprozessierung,<br />

bei der Korrekturfaktoren und Filtereinstellungen für jeden Flug gleich angesetzt werden. Die<br />

darauf folgende Feinprozessierung behandelt jeden Flug und jede Frequenz individuell. Die Prozessierung<br />

der Daten beinhaltet: Filterung, Sprungkorrektur, Nullniveaukorrektur, Feinjustierung der Kalibrierfaktoren<br />

und Berechnung der Halbraumparameter.<br />

2. Messgebiet Langeoog<br />

Langeoog und Spiekeroog sind zwei der Ostfriesischen Inseln im Wattenmeer der Nordsee, die sich<br />

wenige Kilometer vor der Küste befinden. Sie bestehen aus quartären Sedimenten wie Sanden, Tonen<br />

und Schluffen (Abb. 3). Durch Versickerung der Niederschläge in den Dünengürteln werden dort<br />

Grundwasserreservoire mit Süßwasser aufgefüllt, die die Versorgung der Inseln mit Frischwasser sicherstellen.<br />

Weil Süßwasser ein geringeres spezifisches Gewicht als das versalzte Grundwasser hat,<br />

schwimmt es als „Linse“ auf dem Salzwasser. Verändert sich der Wasserstand des dichteren Salzwassers,<br />

ändert sich das Druckniveau und entsprechend auch Stand und Ausdehnung der Süßwasserlinse.<br />

Durch Stürme können Salzwassereinbrüche bis an den Rand des inneren Dünengürtels der<br />

Inseln gelangen und dort versickern. Solche Ereignisse, aber auch der erhöhte Wasserbedarf für den<br />

Tourismus in den Sommermonaten gefährden das Süßwasserreservoir. Noch erfolgt die gesamte<br />

Trinkwasserversorgung mit inseleigenem Grundwasser, welches in den Wintermonaten allein durch<br />

Regenfälle wieder aufgefüllt wird. Unterhalb wasserundurchlässiger Bodenschichten kann Süßwasser<br />

darüber hinaus vom Festland aus unterschiedlich weit in das Wattenmeer vordringen, um dann unter<br />

gewissen Bedingungen an die Oberfläche zu treten. Die Befliegungen wurden bei Niedrigwasser<br />

durchgeführt, um vergleichbare Bedingungen für jeden Flug zu haben und insbesondere auch, um<br />

den Untergrund des Wattenmeeres ohne die bei Flut vorhandene Meerwasserschicht besser erkun-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

178


den zu können. Aufgrund der Berücksichtigung von Naturschutzauflagen konnte die aerogeophysikalische<br />

Erkundung nur in den Wintermonaten durchgeführt werden. Das Messgebiet umfasst die gesamte<br />

Insel Langeoog, den westlichen Teil der Insel Spiekeroog und das Wattenmeer zwischen Langeoog<br />

und dem Festland (Abb. 2). Einen Überblick über die Messkampagne gibt Tabelle 1.<br />

Abbildung 2: Insel Spiekeroog mit Wattenmeer während der Befliegung im März 2009 und westlicher Teil von<br />

Langeoog mit der Dünenlandschaft, unter der sich Süßwasserlinsen verbergen (Fotos: W. Voß).<br />

Tabelle 1:<br />

Größe des Messgebietes: 259 km²<br />

Gesamtprofillänge: 1200 km<br />

Linienabstand: 67 Mess-Linien mit 250 m (N-S), 7 Kontroll-Linien mit 2000 m (W-O)<br />

Messzeitraum: Februar und November 2008, Februar und März 2009; nur bei Niedrigwasser<br />

Abbildung 3: Geologische Karte des Messgebietes aus dem FIS Geophysik des LIAG.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

179


Abbildung 4 zeigt die vorläufigen Karten des scheinbaren spezifischen Widerstands (mit a bis maxi-<br />

mal 2000 m) der verschiedenen Messfrequenzen (133400 Hz, 41520 Hz, 8390 Hz, 5345 Hz, 1823 Hz und<br />

387 Hz) für das Messgebiet. Jeder Frequenz wird dabei zu den einzelnen Messpunkten eine Schwerpunktstiefe<br />

z* zugeordnet, die in Abbildung 5 dargestellt ist. Für die höchsten Frequenzen sind dies<br />

einige wenige Meter unter der Erdoberfläche, die tiefsten Frequenzen erreichen eine Erkundungstiefe<br />

bis etwa 100 m. Diese Erkundungstiefen sind neben der Frequenz auch von den spezifischen Widerständen<br />

der Untergrundstrukturen abhängig und können somit von Messpunkt zu Messpunkt variieren.<br />

Die Abfolge einer Auswahl von Karten des scheinbaren spezifischen Widerstands in einer 3D-<br />

Darstellung gibt daher eine Abbildung der Leitfähigkeitsstrukturen mit Bezug zur Tiefe nur bedingt<br />

wieder (Abb. 7). Die tiefste Frequenz bildet die Basis der Süßwasserlinse ab.<br />

Deutlich sind die den Süßwasserlinsen zugeordneten Strukturen auf der Insel und die mit salzhaltigem<br />

Nordseewasser gefüllten Priele zu erkennen. Im Bereich vor der Festlandküste kann man Gebiete mit<br />

für Süßwasser typischen Werten der scheinbaren Widerstände erkennen. Hier ist zu klären, ob dies<br />

Austritten von Süßwasser im Watt entspricht.<br />

Abbildung 4: Vorläufige Karten des scheinbaren spezifischen Widerstands a für die Messfrequenzen<br />

133400 Hz, 41520 Hz, 8390 Hz, 5345 Hz, 1823 Hz und 387 Hz (a bis maximal 2000 m).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

180


Abbildung 5: Vorläufige Karten der Schwerpunktstiefe z* für die Messfrequenzen133400 Hz, 41520 Hz,<br />

8390 Hz, 5345 Hz, 1823 Hz und 387 Hz.<br />

Benutzt man einen anderen Farbkeil für die Darstellung der scheinbaren spezifischen Widerstände<br />

(mit a bis maximal 200 m), kann man beispielsweise die Priele noch viel deutlicher erkennen, siehe<br />

Abbildung 6. Auch die mutmaßlichen Austritte von Süßwasser im Bereich vor der Festlandküste sind<br />

wesentlich besser aufgelöst.<br />

Abbildung 6: Vorläufige Karten des scheinbaren spezifischen Widerstands a für drei ausgewählte Frequenzen<br />

(41520 Hz, 8390 Hz und 387 Hz), dargestellt mit einem anderen Farbkeil (a bis maximal<br />

200 m).<br />

In Abbildung 8 wird die topographische Karte mittels des digitalen Höhenmodells (digital elevation model,<br />

DEM) aus dem FIS Geophysik des LIAG gezeigt. Dabei wurden die Daten des Höhenmodells auf<br />

die Fluglinien übertragen. Die topographische Karte kann auch aus GPS- und Laseraltimeterdaten des<br />

Hubschraubermesssystems abgeleitet werden.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

181


Abbildung 7: 3D-Darstellung der Karten des scheinbaren spezifischen Widerstands a für die Frequenzen<br />

41520 Hz, 8350 Hz und 387 Hz. Bild aus Google-Earth, September 2009.<br />

Abbildung 8: Topographische Karte mittels des digitalen Höhenmodells (digital elevation model, DEM) aus<br />

dem FIS Geophysik des LIAG. Dabei wurden die Daten des Höhenmodells auf die Fluglinien<br />

übertragen.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

182


3. Messgebiet Glückstadt<br />

Das Messgebiet Glückstadt umfasst große Teile des Elbemündungsbereichs (Abb. 9). Teile der Ergebnisse<br />

der HEM-Befliegung werden in das Projekt KLIMZUG-NORD (www.klimzug-nord.de) einfließen,<br />

bei dem unter anderem die Auswirkungen des Klimawandels auf das Ästuar der Elbe erforscht<br />

werden. Es ist davon auszugehen, dass sich der Elbwasserstand in diesem Gebiet erhöht, was eine<br />

Intrusion von Brackwasser in den Süßwasseraquifer nach sich ziehen würde. Mit hydraulischen Strömungsmodellen<br />

können solche Fragestellungen untersucht werden. Bereits vorhandene geologische<br />

Informationen sollen mit Hilfe von elektrischen Leitfähigkeitsmodellen der HEM flächenhaft zu einem<br />

geologischen Strukturmodell ergänzt werden. Dieses wird mit hydraulischen Parametern belegt und<br />

dient damit als Grundlage für eine hydraulische Modellierung.<br />

10 km<br />

Abbildung 9: Flugplan für das Messgebiet Glückstadt, welches große Teile des Elbemündungsbereichs<br />

umfasst.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

183


Einen Überblick über den Umfang der Messkampagne gibt Tabelle 2.<br />

Tabelle 2:<br />

Größe des Messgebietes: 1949 km²<br />

Gesamtprofillänge: 8500 km<br />

Linienabstand: 209 Mess-Linien mit 250 m (W-O), 19 Kontroll-Linien mit 2500 m (N-S)<br />

Messzeitraum: Juli und Oktober 2008, März-Mai 2009<br />

Aus den GPS- und den Laseraltimeterdaten des Hubschraubermesssystems kann ein digitales Höhenmodell<br />

abgeleitet werden (Abb. 10). Hier wird besonders der Verlauf der Geest (Höhenzug bestehend<br />

aus pleistozänen Sandablagerungen von Moränen) und der Marsch (flaches holozänes<br />

Schwemmland) deutlich.<br />

10 km<br />

Abbildung 10: Digitales Höhenmodell des Elbemündungsgebietes, abgeleitet aus GPS- und Laseraltimeterdaten<br />

des Hubschraubermesssystems.<br />

Erste Ergebnisse in Form von vorläufigen Karten des scheinbaren spezifischen Widerstands und der<br />

Schwerpunktstiefe sind in Abbildung 11 dargestellt. Die Datenlücken sind teilweise auf Hochspannungsleitungen<br />

und entlang der Elbe auf Radarstationen bzw. Sperrzonen um die Atomkraftwerke<br />

herum zurückzuführen.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

184


Elbe<br />

3 2<br />

1<br />

3<br />

4<br />

41 520 Hz<br />

6<br />

8 390 Hz<br />

1 823 Hz<br />

387 Hz<br />

5<br />

2<br />

a<br />

[m]<br />

10 km<br />

a<br />

[m]<br />

a<br />

[m]<br />

a<br />

[m]<br />

a<br />

[m]<br />

41 520 Hz<br />

8 390 Hz<br />

1 823 Hz<br />

Abbildung 11: Vorläufige Karten des scheinbaren spezifischen Widerstands a und der Schwerpunktstiefe z*<br />

für vier ausgewählte Messfrequenzen.<br />

387 Hz<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

185<br />

z*<br />

[m]<br />

z*<br />

[m]<br />

z*<br />

[m]<br />

z*<br />

[m]<br />

10 km<br />

z*<br />

[m]


Mit Hilfe der geologische Karte vom Mündungsbereich der Elbe (Abb. 12a) können z.B. in der Karte<br />

des scheinbaren spezifischen Widerstands zu der Frequenz 41 kHz (Abb. 12b) folgende Leitfähigkeitsstrukturen<br />

identifiziert werden: 1. Brackwasser der Elbe, 2. Perimarine Ablagerungen: Ton, schluffig,<br />

3. Brackische Ablagerungen: Ton bis Schluff, feinsandig, 4. Moor: Torfe, 5. Rotschlammdeponie:<br />

Eisen- und Titanoxide, 6. Glazifluviatile Ablagerungen: Sand und Kies.<br />

Elbe<br />

4. Zusammenfassung und Ausblick<br />

3<br />

2<br />

3<br />

4<br />

a) b)<br />

2<br />

41 520 Hz<br />

Abbildung 12: a) Geologische Karte und b) Karte des scheinbaren spezifischen Widerstands der Frequenz<br />

41 kHz. Im Mündungsbereich der Elbe können folgende Leitfähigkeitsstrukturen identifiziert<br />

werden: 1. Brackwasser der Elbe, 2. Perimarine Ablagerungen: Ton, schluffig, 3. Brackische<br />

Ablagerungen: Ton bis Schluff, feinsandig, 4. Moor: Torfe, 5. Rotschlammdeponie: Eisen-<br />

und Titanoxide, 6. Glazifluviatile Ablagerungen: Sand und Kies.<br />

In beiden Messgebieten wurde die HEM zur Erkundung von Grundwasserstrukturen erfolgreich eingesetzt.<br />

Die hier gezeigten Karten dokumentieren die vorläufigen Ergebnisse. Nach Abschluss der Prozessierung<br />

werden die HEM-Daten in Schichtmodelle des spezifischen Widerstands invertiert.<br />

Auf den Ostfriesischen Inseln Langeoog und Spiekeroog konnten die Lage und Ausdehnung der<br />

Süßwasserlinsen erfasst werden. Daneben zeigten sich insbesondere in den Daten der tiefsten Messfrequenz<br />

Bereiche veränderter scheinbarer spezifischer Widerstände im Wattenmeer vor der Festlandküste,<br />

die die Vermutung nahe legen, dass hier Süßwasser austritt.<br />

Im Gebiet Glückstadt werden als Fragestellung mögliche Intrusionen von Brackwasser in den Süßwasseraquifer<br />

untersucht. Die bereits vorhandenen geologischen Informationen werden mit Hilfe der<br />

HEM-Modelle flächenhaft zu einem geologischen Strukturmodell ergänzt. Nach einer Belegung mit<br />

hydraulischen Parametern kann es damit als Grundlage für eine Interpretation mit Strömungsmodellen<br />

dienen.<br />

Elbe<br />

3 2<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

186<br />

1<br />

3<br />

4<br />

6<br />

5<br />

2<br />

a<br />

[Ohm*m]<br />

a<br />

[m]<br />

10 km


Danksagung<br />

Die hier vorgestellten Teilgebiete des BGR/LIAG-Projektes wurden in Zusammenarbeit mit dem Institut<br />

für Chemie und Biologie des Meeres (ICBM), der Universität Oldenburg sowie der Behörde für<br />

Stadtentwicklung und Umwelt Hamburg durchgeführt.<br />

Referenzen<br />

Siemon, B. [2009] Electromagnetic methods – frequency domain: Airborne techniques, In: Kirsch, R.<br />

(Ed.), Groundwater Geophysics – A Tool for Hydrogeology. 2nd Edition, Springer-Verlag, Berlin,<br />

Heidelberg, 155-170.<br />

Siemon, B., Christiansen, A.V. & Auken, E. [2009] A review of helicopter-borne electromagnetic<br />

methods for groundwater exploration. Near Surface Geophysics, 7, 629-646.<br />

Steuer, A., Siemon, B., Schaumann, G., Wiederhold, H., Meyer, U., Pielawa, J., Binot, F. & Kühne, K.<br />

[2009] The German North Sea Coast in Focus of Airborne Geophysical Investigations, AGU Fall<br />

Meeting 2009, San Francisco, USA.<br />

Steuer, A., Siemon, B. & Grinat, M. [2010] The German North Sea Coast in Focus of Airborne<br />

Electromagnetic Investigations: The Freshwater Lenses of Borkum, EMTF, this volume.<br />

Wiederhold, H., Binot, F., Kühne, K., Meyer, U., Siemon, B. & Steuer, A. [2008] Airborne geophysical<br />

investigation of the German North Sea Coastal Area, 20th Salt Water Intrusion Meeting 2008, Naples,<br />

USA.<br />

http://www.liag-hannover.de/forschungsschwerpunkte/grundwassersysteme-hydrogeophysik/salz-<br />

suesswassersysteme/flaechenhafte-befliegung.html<br />

http://www.bgr.bund.de/cln_101/nn_328750/DE/Themen/GG__Geophysik/Aerogeophysik/Projektbeitr<br />

aege/D-AERO/deutschlandweite__aerogeophysik__befliegung__D-AERO.html<br />

http://www.fis-geophysik.de/<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

187


The German North Sea Coast in Focus of Airborne Electromagnetic<br />

Investigations: The Freshwater Lenses of Borkum<br />

Annika Steuer 1 , Bernhard Siemon 1 and Michael Grinat 2<br />

1<br />

Federal Institute for Geosciences and Natural Resources (BGR), Hannover, Germany<br />

2<br />

Leibniz Institute for Applied Geophysics (LIAG), Hannover, Germany<br />

annika.steuer@bgr.de<br />

Abstract<br />

Helicopter-borne electromagnetic (HEM) measurements were conducted over the North Sea island of<br />

Borkum to determine the size of the freshwater lenses. Additionally, geoelectrical measurements were<br />

carried out at locations where data of 13–17 years old Schlumberger soundings exist. HEM<br />

successfully revealed the lateral extension as well as the thickness of the freshwater lenses of the<br />

island of Borkum. At several locations the depth of the freshwater/saltwater boundary determined by<br />

HEM was confirmed by the results of the Schlumberger soundings. No significant changes of the<br />

freshwater/saltwater boundary were found at the locations where old and new Schlumberger<br />

soundings exist. Regarding the results of direct-push and borehole measurements, the HEM related<br />

resistivity structure could be determined more precisely and additional to the freshwater/saltwater<br />

boundary thin clay layers were revealed. Based on the HEM inversion results the volume of the<br />

freshwater resource was estimated.<br />

North Sea<br />

Germany<br />

Figure 1: The location of the island of Borkum in the North Sea<br />

and a Google-Earth map with the flight-lines.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

188


Introduction<br />

Since 2008, the German North Sea Coast has been in the focus of airborne geophysical investigations<br />

carried out by the Federal Institute for Geosciences and Natural Resources (BGR) and the Leibniz<br />

Institute for Applied Geophysics (LIAG). Especially electromagnetics is the most versatile of the<br />

airborne geophysical methods and widely applied in hydrogeological investigations as the<br />

measurements respond to both lithologic and water-chemistry variations. The applications comprise<br />

geologic mapping and aquifer structure, delineation of soil and groundwater salinization, saltwater<br />

intrusion into coastal aquifers etc.<br />

Here, we present the HEM results for the North Sea island of Borkum, which was the first of the five<br />

HEM project areas investigated in 2008-2009, and compare them with DC Schlumberger soundings,<br />

lithological logs and direct-push measurements. These data will be used for a hydraulic modelling of<br />

the freshwater lenses of the island of Borkum in the CLIWAT project (http://cliwat.eu, EU Interreg IVB<br />

North Sea Region). Additionally we show an estimation of the volume of the freshwater lenses<br />

calculated by Siemon et al. (2009b).<br />

Other HEM results for the Wadden Sea, the North Sea islands of Langeoog and Spiekeroog, and the<br />

Elbe estuary are shown by Schaumann et al. (2010) in this volume. Wiederhold et al. (2009) presented<br />

the results of two SkyTEM surveys at the North Sea island of Föhr and the saltdome Bad Segeberg,<br />

which were conducted in 2008 by SkyTEM Aps by order of LIAG.<br />

HEM Survey<br />

The Borkum survey covers an area of 88 km² and was flown within two days in March 2008. The flightline<br />

spacing was 250 m and the tie-line spacing was 500 m, totalling to about 412 line-kilometres<br />

(Figure 1). The BGR helicopter-borne geophysical system simultaneously uses frequency-domain EM,<br />

magnetics and radiometrics (Figure 2).<br />

The HEM system, a RESOLVE bird manufactured by Fugro Airborne Surveys, operates at six<br />

frequencies ranging from 386 Hz to 133 kHz.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

189


Figure 2: BGR helicopter-borne geophysical system:<br />

Electromagnetics:<br />

RESOLVE (Fugro Airborne Surveys)<br />

Frequencies: 386, 1 823, 5 495, 8 338,<br />

41 430 and 133 300 Hz<br />

Coil separations: ~8 m<br />

Coil orientation: 5× horizontal coplanar<br />

1× vertical coaxial<br />

Investigation depth: 1–70 m (at Borkum)<br />

Magnetics: Cs magnetometer<br />

Radiometrics: 256 channel gamma-ray<br />

spectrometer<br />

Navigation/Positioning:<br />

GPS / radar and laser altimeter<br />

Bird altitude: ~ 34 m<br />

Flight speed: ~140 km/h<br />

Sampling distance: ~4 m<br />

The investigation depth increases with decreasing frequency. The EM fields of the lowest frequency<br />

penetrate the freshwater lenses nearly completely as seen in the apparent resistivity maps (Figure 3).<br />

Here, higher resistivities are represented by green and blue colours (freshwater) and lower resistivities<br />

by red colours (saltwater). The apparent resistivity maps, which are based on the half-space<br />

approximation, provide a first insight into the subsurface conductivity distribution.<br />

133 kHz 41 kHz 1.8 kHz 0.4 kHz<br />

Figure 3: Apparent resistivity maps: The investigation depth increases with decreasing frequency. The<br />

freshwater lenses – represented by higher resistivities (green & blue) – are nearly completely<br />

penetrated by the EM fields of lowest frequency.<br />

The HEM data were also inverted using Marquardt-Levenberg 1-D inversion technique based on a<br />

layered half-space model (Siemon et al., 2009a). Starting models required were derived automatically<br />

from the apparent resistivities vs. centroid depth sounding curves. Resistivity maps at selected depths<br />

and stitched together vertical resistivity sections were derived from these 1-D inversion models. The<br />

lateral extensions as well as the thicknesses of the freshwater lenses were mapped successfully with<br />

HEM, as shown in the resistivity maps and the vertical resistivity sections of Figure 4.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

190


Figure 4: Resistivity maps at 5 and 40 m bsl and vertical resistivity sections along the tie lines T13.9<br />

and T6.9.<br />

The first layer represents the resistive dry dune sands above the groundwater table (dark blue) and<br />

the bottom layer the conductive saltwater-filled sediments (red) below the freshwater lenses (blue).<br />

The two layers in-between are associated with sediments filled with fresh or brackish water. The<br />

freshwater/saltwater boundary was detected down to about 60 m depth.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

191


DC survey: Comparison of measurements 16 years apart<br />

In 2008, 36 Schlumberger soundings were<br />

carried out at sites already measured<br />

between 1991 and 1995 (Worzyk, 1995a<br />

and 1995b).<br />

All the sounding curves at Borkum have a<br />

Q-type shape with three or more layers (e.g.<br />

Figure 5). Higher resistivities are due to dry<br />

and freshwater filled sediments, whereas the<br />

lower resistivities at the end of the sounding<br />

curves are due to saltwater-filled sediments<br />

underlying the freshwater lenses.<br />

apparent resistivity [m]<br />

10000<br />

1000<br />

1 10 100 1000<br />

current electrode separation AB / 2 [m]<br />

In 1992, the freshwater filled sediments<br />

were associated with resistivities between<br />

Figure 5: DC sounding curve of site GTS 60.<br />

70 and 110 m in the water catchment area<br />

Ostland; these resistivities were almost reproduced in 2008. The underlying saltwater-filled sediments<br />

showed resistivities below 10 m, but these values were not determined exactly. Taking equivalent<br />

models into consideration the thickness of the freshwater lens can be determined with an accuracy of<br />

about 6–8 m.<br />

No significant changes of the freshwater/saltwater boundary were found at the locations where old and<br />

new Schlumberger soundings exist.<br />

100<br />

10<br />

GTS 60 (5/1992): Data + model (-)<br />

GTS 60 (10/2008): Data + model (-)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

192


Combined analysis<br />

At several locations the depth of the freshwater/saltwater boundary determined by HEM was<br />

confirmed by the results of the Schlumberger soundings (e.g., Figure 6).<br />

a)<br />

c)<br />

BR40<br />

T 7.9<br />

NE SW<br />

HEM T 7.9<br />

GTS 60<br />

BR 40<br />

HF510S3L5 D2 SF075 200<br />

The elevation map in Figure 6b shows the location of the Schlumberger sounding GTS 60 (blue circle),<br />

the drilling BR 40 (orange circle) and a part of flight-line T 7.9 (red line). The flight-line crosses a dune<br />

and the Schlumberger sounding is located about 80 m apart from the flight-line in a dune valley. The<br />

vertical resistivity section of this part in Figure 6c shows that the saltwater/freshwater boundary was<br />

determined by the HEM 1-D inversions at about 60 m depth. This is confirmed by the 1-D inversion<br />

result of the Schlumberger sounding of 2008 which has not changed significantly since 1992 (Figure<br />

6d).<br />

Regarding the results of direct-push (Winter, 2008) and borehole measurements, the HEM related<br />

resistivity structure could be determined more precisely. A more detailed 6-layer approach revealed<br />

additionally to the freshwater/saltwater boundary thin conductive layers at about 10 m depth, not seen<br />

in the 4-layer models of the standard inversion procedure (Figure 7b). Partly, they were identified as<br />

clay layers by lithological logs; and also the result of a direct-push measurement shows an increasing<br />

conductivity at 10 m depth (Figure 7c). The depth of the freshwater/saltwater boundary remains nearly<br />

constant and has a layer with little higher resistivity above, which could be interpreted as transition<br />

b)<br />

d)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

0<br />

8<br />

depth [m]<br />

HEM<br />

GTS 60 (10/2008) GTS 60 (5/1992) BR 40<br />

4560 2578 m<br />

Br unnen 40<br />

2181 m<br />

Feinsand,<br />

z.T. schluffig<br />

76<br />

Filter I<br />

69<br />

1.05<br />

92<br />

1.1<br />

87<br />

1.1*<br />

M itt elsand<br />

Mittelsand,<br />

fein- bis<br />

grobsandig<br />

Geschiebelehm,<br />

Schluff, tonig<br />

Filter II<br />

ET<br />

Figure 6: a) Resistivity map (5 mbsl) with the location of a part of flight-line T7.9 (red line), the lithological log<br />

BR40 (orange circle) and the DC site GTS 60 (blue circle). b) Digital elevation model revealed<br />

from laser scan data of NLWKN. c) Vertical resistivity section T7.9 as result of a 5-layer 1-D<br />

inversion (every 5 th model is shown). d) Comparison of 1-D inversion results of HEM and DC<br />

reveals the freshwater/saltwater boundary at a similar depth of about 60 mbsl.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

193<br />

BR40<br />

T 7.9


zone of brackish water. The data misfit q [%] of the 6-layer inversion is smaller than the misfit of the 4layer<br />

inversion, especially in the south-eastern part of the section.<br />

a)<br />

b)<br />

c)<br />

NW HEM L16.1<br />

SE<br />

DP 06<br />

DirectPush 06<br />

GTS 8<br />

m/min mS/m Ohmm<br />

10-12 m<br />

L16.1<br />

WWBO 1026 BORK II 811<br />

H_F5_10_S3_L4_D2_SF100_<br />

0<br />

10<br />

20<br />

30<br />

40<br />

50<br />

60<br />

70<br />

78<br />

37<br />

2.0<br />

80<br />

depth [m]<br />

HEM GTS 8<br />

9890<br />

80<br />

18<br />

103<br />

9.3<br />

2.0<br />

9883<br />

NW HEM L16.1<br />

SE<br />

Figure 7: a) Apparent resistivity map (41 kHz) with the location of the vertical resistivity section<br />

L16.1, the lithological logs and the direct-push measurement. b) Comparison of HEM 1-D<br />

inversion results (4- and 6-layer models) along L16.1 with c) lithological logs (WWBO 1016 and<br />

BORK II 811) (LBEG), direct-push results (DP 06) (Winter, 2008) and DC 1-D inversion results<br />

(GTS8).<br />

118<br />

61<br />

2.3<br />

DP 06<br />

1035 m<br />

GTS 8<br />

WWBO 1026<br />

WWBO 1026 BORK II 811<br />

> 8.9 m clay<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

194<br />

H_F5_10_S3_L6_D2_SF100_200<br />

BORK II 811<br />

6.3-9.5 m clay<br />

9.5-12 m peat


Estimation of the aquifer thickness<br />

The aquifer thickness was estimated using Archie’s law with the following approach: The resistivity of<br />

freshwater is assumed to be higher than 5.6 m (180 mS/m) and thereby the resistivity of freshwater<br />

saturated sediment with 33% porosity and a formation factor of 4 should be above 23 m (Siemon et<br />

al., 2009b). Based on that, the accumulated layer thickness with resistivities of 30–500 m is defined<br />

as aquifer thickness and is shown in Figure 8. From that, the freshwater volume was estimated to 175<br />

x 10 6 m³.<br />

Conclusions<br />

Figure 8: Aquifer thickness derived from DC and HEM.<br />

(1991/92)<br />

(based on 4-layer model)<br />

At several locations the depth of the freshwater/saltwater boundary as derived by the HEM inversion<br />

was confirmed by Schlumberger soundings. The DC measurements were carried out at the same sites<br />

as 16 years before, no significant changes of the freshwater/saltwater boundary were found.<br />

Regarding direct-push and borehole data, the HEM inversion provided more information about<br />

resistivity structure. A 6-layer approach revealed additionally to the freshwater/saltwater structure thin<br />

clay layers.<br />

Based on the HEM inversion results the volume of the freshwater resource was estimated.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

195


Acknowledgments<br />

The HEM survey was funded by LIAG/BGR project “Airborne Geophysical Investigation of the German<br />

North Sea Coastal Area” (Wiederhold et al., 2008):<br />

http://www.liag-hannover.de/forschungsschwerpunkte/grundwassersysteme-<br />

hydrogeophysik/salz-suesswassersysteme/flaechenhafte-befliegung.html<br />

http://www.bgr.bund.de/cln_101/nn_328750/DE/Themen/GG__Geophysik/Aerogeophysik/Proj<br />

ektbeitraege/D-AERO/deutschlandweite__aerogeophysik__befliegung__D-AERO.html<br />

The lithological logs were taken from the LBEG Map Server (http://www.geophysics-database.de/) of<br />

the Niedersächsisches Landesamt für Bergbau, Energie und Geologie (LBEG).<br />

The direct-push results were provided by the Ingenieurbüro für Hydrogeologie, Sedimentologie und<br />

Wasserwirtschaft (HSW) (Winter, 2008).<br />

The digital elevation model was provided by the Niedersächsischer Landesbetrieb für<br />

Wasserwirtschaft, Küsten- und Naturschutz (NLWKN), Forschungsstelle Küste, An der Mühle 5, 26548<br />

Norderney.<br />

References<br />

Schaumann, G., Steuer, A., Siemon, B., Wiederhold, H. & Binot, F. [2010] Die deutsche Nordseeküste<br />

im Fokus von aeroelektromagnetischen Untersuchungen, Teilgebiete Langeoog mit Wattenmeer und<br />

Elbemündung, EMTF, this volume.<br />

Siemon B. [2006] Electromagnetic methods – frequency domain: Airborne techniques, In: Kirsch, R.<br />

(Ed.), Groundwater Geophysics – A Tool for Hydrogeology. Springer-Verlag, Berlin, Heidelberg, 155-<br />

170.<br />

Siemon B., Auken E. & Christiansen A.V. [2009a] Laterally constrained inversion of frequency-domain<br />

helicopter-borne electromagnetic data, Journal of Applied Geophysics, 67, 259-268.<br />

Siemon, B., Christiansen, A.V. & Auken, E. [2009b] A review of helicopter-borne electromagnetic<br />

methods for groundwater exploration. Near Surface Geophysics, 7, 629-646.<br />

Wiederhold, H., Binot, F., Kühne, K., Meyer, U., Siemon, B. & Steuer, A. [2008] Airborne geophysical<br />

investigation of the German North Sea Coastal Area, 20th Salt Water Intrusion Meeting 2008, Naples,<br />

USA.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

196


Wiederhold, H., Kirsch, R., Schaumann, G., Scheer, W. & Steuer, A. [2009] Airborne geophysical<br />

investigations for hydrogeological purposes in Northern Germany. Extended Abstracts Book of the<br />

15th European Meeting of Environmental and Engineering Geophysics – Near Surface 2009, Dublin,<br />

Ireland, 7.-9.9.2009.<br />

Winter, S. [2008] Erkundung und Bewertung von Grundwasserbelastungen im Bereich der<br />

Süßwasserlinsen Borkum, Teilprojekt Direct-Push-Sondierungen, Ergebnisbericht, HSW, Leer.<br />

Worzyk P. [1995a] Geoelektrische Tiefensondierungen auf der Insel Borkum – Untersuchungen zur<br />

Veränderung der Süßwasserlinse im Ostland. NLfB Archives no 111751, Hannover.<br />

Worzyk P. [1995b] Geoelektrische Tiefensondierungen auf der Insel Borkum – Untersuchungen zur<br />

Struktur der Süßwasserlinse im Einzugsbereich des Gewinnungsgebietes Waterdelle. NLfB Archives<br />

no 114088, Hannover.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

197


Some notes on bathymetric effects in marine<br />

magnetotellurics, motivated by an amphibious<br />

experiment at the South Chilean margin<br />

Gerhard Kapinos, Heinrich Brasse 1<br />

Freie Universität Berlin, Fachrichtung Geophysik, Malteserstr. 74-100, 12249 Berlin,<br />

Germany<br />

1 Introduction<br />

Whilst the onshore magnetotelluric method is established as a useful method for imaging<br />

electrical conductivity structures in the deep earth’s interior, electromagnetic investigation<br />

in marine environment has attracted until recently much less attention, although<br />

the oceans cover the most part of the Earth’s surface. Beside of logistical challenges,<br />

high technical demands on the instruments and involved higher costs, physical conditions<br />

in marine environments make the acquisition and interpretation of offshore data<br />

by far more difficult than onshore data. The broad period range of usable signals utilized<br />

in the terrestrial MT is substantially limited on the seafloor. The ionospheric and<br />

magnetospheric primary source field recorded on the seafloor is at short periods strongly<br />

attenuated by the covering highly conductive ocean layer and at long periods contaminated<br />

by motionally induced secondary electromagnetic fields. Additionally, attention<br />

has to be payed to the shape of the ocean bottom, which can massively distort the<br />

electromagnetic fields particularly in coast region.<br />

2 Decay of electromagnetic fields in the ocean<br />

The decay of the electromagnetic amplitude in a high conductive homogeneous half space<br />

like ocean water (ρ =0.3Ωm) and a less conductive medium like earth (ρ = 100Ωm)<br />

shows Fig. 1. In a homogeneous half space without boundary in vertical direction the<br />

field behaviour in both media can be estimated according to the simple formula used for<br />

the calculation of the skin depth<br />

F = F0e −kz , (1)<br />

1 kapinosg@geophysik.fu-berlin.de, h.brasse@geophysik.fu-berlin.de<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

198


z/δ<br />

0<br />

1<br />

2<br />

3<br />

4<br />

Re<br />

Im<br />

magnitude<br />

land<br />

ocean<br />

5<br />

-0.5 0 0.5 1<br />

E(z)/E(0)<br />

Figure 1: Decay of the of magnitude of electromagnetic field F (green) and its real<br />

(red) and imaginary part (blue) as function of depth in a highly conductive (i.e. ocean<br />

water, dashed line) and a less conductive homogeneous half space (solid line). The<br />

decay is scaled to the surface value F0.<br />

where k = √ iμσω. At a depth of 4000 m, which corresponds to the average ocean<br />

depth, the amplitude of a field propagating in the ocean at a period of 10 s is attenuated<br />

hundredfold, i.e. the field on the ocean bottom would be approximately 1% of the surface<br />

value.<br />

This simple relationship becomes more complicated when taking into account real<br />

conditions with a highly conductive ocean layer which is limited downward by a relatively<br />

resistive basement. Analog to the N-layered half space the solution has to be upgraded<br />

by a further term according to the reflection on the ocean bottom at depth h. The<br />

solution for the x-component of the electric field in the first and second layer is<br />

E1,x = E11e −k1z + E12e k1z<br />

E2,x = E22e −k2z<br />

where E11 is the amplitude of the downgoing and E12 of the upgoing wave in the first<br />

medium and E22 the amplitude of the downgoing wave in the second medium. For the<br />

y−component of the magnetic field follows from Faraday’s law in a 1-D case<br />

B1,y = − 1 ∂E1x<br />

iω ∂z<br />

B2,y = − 1 ∂E2x<br />

iω ∂z<br />

= k1<br />

iω<br />

<br />

E11e −k1z k1z<br />

− E12e<br />

(2)<br />

(3)<br />

(4)<br />

= k2<br />

iω E22e −k2z . (5)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

199


A similar relation can be derived for the pair Ey and Bx. Using the continuity of the<br />

fields at z = h resulting from the boundary conditions<br />

and introducing the admittances<br />

ζ1 = k1<br />

iω =<br />

E1,x(h) =E2,x(h), B1,y(h) =B2,y(h) (6)<br />

<br />

iωμσ1<br />

iω =<br />

<br />

μσ1<br />

iω<br />

and ζ2 = k2<br />

iω =<br />

<br />

μσ2<br />

iω<br />

the ratio of amplitudes of the downgoing and upgoing fields in the ocean is<br />

where ζ12 = ζ1−ζ2<br />

ζ1+ζ2<br />

E12<br />

E11<br />

(7)<br />

= ζ1 − ζ2<br />

e<br />

ζ1 + ζ2<br />

−2k1h<br />

= ζ12e −2k1h<br />

, (8)<br />

is the reflection coefficient. Its value is estimated according to eq. 7 by<br />

the conductivity contrast at the boundary between layers. For instance, in case of sharp<br />

contrast like ocean water (σ1 =3.2 S/m) and basement (σ2 =0.001 S/m) ζ12 =0.965.<br />

Note, that using as solution in eq. 2 and 3 the magnetic field instead of electric field, the<br />

reflection coefficient has to be expressed by reciprocal impedances instead of admittance<br />

and yields: η12 = η1−η2<br />

η1+η2 ,withη1,2 = k1<br />

μσ1,2 =<br />

iω<br />

μσ1,2 .<br />

Combining the equations 8 and 2 the electric field at depth z ≤ h becomes<br />

E1,x(z) =<br />

−k1z<br />

E11 e + ζ12e −2k1h k1z<br />

e <br />

and at the surface z =0<br />

<br />

E1,x(0) = E11 1+ζ12e −2k1h . (10)<br />

The ratio of the equations 9 and 10 gives the attenuation of the electric field in the ocean<br />

for 0 ≤ z ≤ h normalized in terms of the total field at the surface<br />

|VE| = E1,x(z)<br />

E1,x(0) = e−k1z + ζ12e−2k1h 1+ζ12e−2k1h −2k1(h−z)<br />

1+ζ12e<br />

= e−k1z<br />

1+ζ12e−2k1h This applies accordingly for the ratio of magnetic fields<br />

B1,y(z) =<br />

−k1z<br />

B11ζ1 e − ζ12e −2k1h k1z<br />

e <br />

B1,y(0) =<br />

−2k1h<br />

B11ζ1 1 − ζ12e<br />

(12)<br />

(13)<br />

|VB| = B1,y(z)<br />

B1,y(0) = e−k1z − ζ12e−2k1h 1 − ζ12e−2k1h −2k1(h−z)<br />

1 − ζ12e<br />

=<br />

1 − ζ12e−2k1h e−k1z . (14)<br />

The behaviour of the electric and magnetic fields as function of depth in a 5 km deep<br />

ocean for two selected and representative periods (100 s and 10000 s) and realistic conductivities<br />

in the ocean (σ1 = 3.2 S/m) and basement (σ1 = 0.001 S/m) is illustrated in<br />

Fig. 2(left). The fields are normalized to their surface magnitude according to relations<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

200<br />

(9)<br />

(11)


z [km]<br />

0<br />

1<br />

2<br />

3<br />

4<br />

5<br />

3200<br />

B<br />

E<br />

E,B (hs)<br />

T=100s<br />

T=10000s<br />

0 0.2 0.4 0.6 0.8 1<br />

|V | , |V |<br />

B E<br />

T=10000s<br />

B<br />

E<br />

E,B (hs)<br />

=3.2<br />

3200<br />

0 0.2 0.4 0.6 0.8 1<br />

|V B | , |V E |<br />

Figure 2: The decay of electromagnetic fields in an ocean shows a deviation from simple<br />

half space attenuation due to boundary conditions on the seafloor and depends beside<br />

period length and ocean depth also on resistivity contrast between ocean water and<br />

basement. Attenuation of electric (blue) and magnetic (red) fields for two periods<br />

T1 = 100 s (solid line) and T1 = 10000 s (dashed line) in a 5 km deep ocean with a sea<br />

water resistivity of 0.3Ωm and seafloor resistivity of 1000 Ωm(right). Attenuation of the<br />

fields at a period of 10000 s for two different conductivities of the basement (0.001 S/m<br />

and 1 S/m) (left). See also (Brasse, 2009).<br />

11 and 14. They show different decay in dependence of period. At short periods (solid<br />

line) both the electric (blue) and magnetic (red) fields experience strong attenuation and<br />

the curve shape is determined by exponential decay. Large differences in attenuation<br />

are observed at long periods (dashed line). While the electric field penetrates the ocean<br />

layer from surface to the seafloor nearly unchanged, the magnetic field is attenuated<br />

about fifteen-fold and reaches on the ocean bottom just about 7% of its surface value.<br />

However, both fields would behave identically in an infinitely extended ocean, that would<br />

be regarded as a homogeneous half space.<br />

At first sight of the equations 11 and 14 one would think that the behaviour of the<br />

fields depends rather on the ocean thickness than on period. However, actually both<br />

is true. The decay of the fields is determined by both period T and thickness h of the<br />

ocean layer. That becomes evident considering the exponential and crucial term in the<br />

equations as function of skin depth according to the relations of skin depth δ =<br />

and complex wave number k =(1+i) ωσμ<br />

2 :<br />

2<br />

ωσμ<br />

e −2k1h = e − 2h<br />

δ 1 (i+1) . (15)<br />

The quantity h<br />

δ1 is the electrical layer thickness and via δ ∼ √ T also directly associated<br />

with the period length of the penetrating wave (McNeill and Labson, 1986). The fields<br />

behave identically if the exponent remains constant, regardless if high frequency fields<br />

propagate in a shallow ocean or low frequency fields in a deep ocean. The behaviour<br />

of the fields replicates, similar to a self-similarity principle by stimulating the exponent<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

201


h , i.e. by appropriate and simultaneous increasing or decreasing the frequency and<br />

δ1<br />

the ocean thickness and thus creating either electrical thick h >> 1 or thin conditions<br />

δ1<br />

h > 1andbothe−2k1h<br />

δ1<br />

and e−2k1(h−z) → 0 (Brasse, 2009). Thus the effect of the basement on the fields is only<br />

marginal and the behaviour in a thick electrical layer is governed by exponential decay<br />

E1,x(z)<br />

E1,x(0)<br />

≈ e−k1z<br />

respectively B1,y(z)<br />

B1,y(0) ≈ e−k1z . (16)<br />

The decay of both fields becomes more alike the more the ocean depth or the frequency<br />

increases until finally approaching the decay curve of a homogeneous half space.<br />

For low frequencies or for an electrically thin ocean, where the skin depth is significantly<br />

greater than ocean depth, h


would imply that the currents only flow in the ocean, inducing a secondary magnetic<br />

field which removes entirely the opposite primary field in agreement with the above<br />

mentioned result. In real conditions with high but not infinite basement resistivity<br />

the currents diffuse also to the seafloor, so that the the secondary magnetic field of<br />

the diminished currents in the ocean can in fact reduce but not completely cancel the<br />

primary magnetic field. Thus the ratio B1,y(z)<br />

B1,y(z)<br />

doesn’t become zero rather 0 < < 1<br />

B1,y(0) B1,y(0)<br />

depending on ζ12 and electrical conditions of the ocean layer. In other words the reflection<br />

coefficient regularizes the conditions for distribution of electrical currents in sea bottom<br />

and estimates in this manner the damping behaviour of the fields (Chave et al., 1991).<br />

The decay of the electric field in the ocean is against intuition, and negligible compared<br />

to the magnetic field. In the theoretical case with infinite conductivity contrast, (ζ12 =1)<br />

the ocean represents for long periods an electrical thin layer that the electric field passes<br />

almost intact. In real condition mentioned above ζ12 is still high and the low frequency<br />

field decays gently, in terms of an electrical thin layer. That would change assuming<br />

a layer of highly conductive sediments on the ocean bottom. Since the layer becomes<br />

thicker, the electric field decays more strongly, the magnetic field in contrast much less<br />

since a part of currents now flows in the sediments. Thus both fields would approximate<br />

the curve of attenuation in a homogeneous half space, what for short periods corresponds<br />

to exponential decay Fig. 2(right).<br />

Similar effects would be observed considering the fields at a fixed point z1, within the<br />

ocean (0


It is also instructive to consider the electric and magnetic fields in the ocean as function<br />

of periods. Fig. 3 illustrates graphically the ratio |VB| = B1,y(z)<br />

B1,y(0) and |VE| = E1,x(z)<br />

E1,x(0)<br />

for various thicknesses of the ocean layer. It becomes clear immediately that in the<br />

real ocean environment attenuation of the fields differs fundamentally over a total period<br />

range. While the electric field remains nearly unchanged in electric thin layer<br />

( h


ocean depends on pressure, salinity and temperature. While the effect of pressure and<br />

salinity is marginally, the temperature variations in ocean change resistivity of oceanic<br />

water from 0.16 Ωm at 30 ◦ C to 0.33 Ωm at 1.0 ◦ C. The estimated resistivity values about<br />

0.3Ωm corresponds very well to literature values (Filloux, 1987)<br />

3 Offshore magnetotelluric and magnetic transfer<br />

functions in presence of bathymetry<br />

The discussions about magnetotellurics on the seafloor are often dominated by attenuation<br />

of electromagnetic fields by high conductive ocean and secondary induction by<br />

sea tides. Fewer studies refer to distorting effects on electromagnetic fields caused by<br />

changes in seafloor bathymetry in presence of overlaying high conductive environment<br />

(Constable et al., 2009).<br />

For a model incorporating a homogeneous half space with resistivity of 100Ωm and<br />

conductive ocean with bathymetry that is observed at the South Chilean continental<br />

margin (Fig. 5, top), magnetotelluric transfer function were calculated with the algorithm<br />

of Mackie et al. (1994). At least two clear findings can be derived from the model<br />

0<br />

2<br />

4<br />

6<br />

10 0<br />

10 1<br />

10 2<br />

10 3<br />

104 100 10 1<br />

10 2<br />

10 3<br />

10 4<br />

ocean<br />

ρ= 0.3Ωm ρ=100Ωm TM<br />

TM<br />

0 50 100 150 200 250 300 350 400<br />

distance [km]<br />

0 50 100 150 200 250 300 350 400<br />

distance [km]<br />

Figure 5: Responses of a synthetic amphibious model shown on the top. Middle: resistivity<br />

of TM and TE mode. Bottom: phase of TM and TE mode. Dashed line separates<br />

the profile into offshore and onshore part.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

205<br />

TE<br />

TE<br />

4<br />

3<br />

2<br />

1<br />

0<br />

90<br />

75<br />

60<br />

45<br />

30<br />

15<br />

0


esponses (Fig. 5, middle and bottom). Firstly the ocean has rather a marginal effect<br />

on the onshore magnetotelluric transfer function (right of the dashed line). Actually<br />

only in TM mode of onshore stations close to the ocean slightly decreasing phases and<br />

enhanced apparent resistivity on the edge between offshore- and onshore profile marked<br />

by a dashed line, can be observed. Secondly the bathymetry produces dramatic anomalous<br />

effects in the ocean, which can be observed in both modes as well as in phases and<br />

resistivities. The resistivity in TE mode rises and falls dramatically producing cusps in<br />

the image and phases even exceed −180 ◦ and 180 ◦ .<br />

Similar anomalous features can be observed in the magnetic transfer functions (Fig.<br />

6). Slight changes in the shape of the seafloor, adumbrated by a dotted line, are overdrawn<br />

reproduced by dramatic variations in the magnitude of the induction arrows (note<br />

the smooth bathymetry change between stations A09 and A13 and their huge impact on<br />

the magnitude of the induction arrows). On the other hand, the ocean effect on onshore<br />

station is limited to long-periods and to near-coastal region, again.<br />

10 0<br />

10 1<br />

10 2<br />

10 3<br />

10 4<br />

tipper M r<br />

0 50 100 150 200 250 300 350 400<br />

distance [km]<br />

tipper Mag<br />

0 50 100 150 200 250 300 350 400<br />

distance [km]<br />

Figure 6: Real part (left) and the magnitude (right) of the tipper calculated for synthetic<br />

and amphibious model shown on the top of figure 5. The dotted line indicates<br />

the shape of the bathymetry of the model.<br />

Comparing these synthetic responses with the real induction vectors observed at the<br />

only station at the continental slope, ob7, shows that the bathymetry indeed affects the<br />

measured data, as shown in Fig. 7. The huge induction vectors (over 1) at middle<br />

and long periods correspond roughly with magnitude values calculated from the simple<br />

amphibious model. For short periods (until 100 s) a comparison is impossible due to<br />

Re<br />

1.0<br />

0.5<br />

0.0<br />

10 2<br />

ob7<br />

10 3<br />

Figure 7: Real part of the observed induction vectors at station ob7, deployed on the<br />

continental margin off Southern Chile. Induction vectors observed land-side are presented<br />

by Brasse (2009).<br />

10 4<br />

10 5<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

206<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0


unavailable data (note the different period range of the images).<br />

The anomalous responses can be elucidated considering electric and magnetic fields in<br />

presence of changing seafloor shape. An enhanced concentration of electrical currents at<br />

the continental slope produces anomalous strong vertical magnetic field, that is known<br />

also on the land-side as coast effect (e.g., Fischer, 1979). Due to the high sea water<br />

conductivity the induced electric currents concentrate primarily in the ocean avoiding<br />

the much more resistive subsurface and generate a secondary magnetic field which compensates<br />

the primary opposite field, so that the magnetic field diminish considerably or<br />

disappears completely in case of insulating substrate as mentioned in 2 and shown in<br />

Fig. 2. At the continental margin, where the seafloor shallows towards land, the density<br />

of electric currents flowing in the ocean along the coast (i.e. in the TE mode) and above<br />

(instead below) measurement points increase inducing an anomalous and opposite magnetic<br />

field. This opposite, secondary field becomes predominant on the ocean bottom,<br />

where the primary field is strongly damped whereas the electric field is by the coast effect<br />

only marginally affected, and causes jumps in the phase (Weidelt, 1994). Moreover the<br />

resistivities in TE mode rise steeply generating upward cusps and the tipper gets very<br />

large as presented in Fig. 6 and 5.<br />

The periods at which the effects occur depend on position of the probe in relation to<br />

the slope and corresponds to lateral distance to which a transfer function is sensitive<br />

to bathymetry. However, at high frequencies the field is generally unaffected by the<br />

bathymetry and the results are MT-like because the fields still behave like plane waves.<br />

At low frequencies, the currents are below the ocean. At the middle frequencies, the<br />

currents are in the ocean and this causes the responses looking like those due to infinite<br />

line currents right above the station. Note that if the MT responses were recorded<br />

on the top of the sea, the responses would be perfectly fine as expected in an usual<br />

2-D case, since all the currents pass below the sounding surface instead above like in<br />

the presented marine model. This behavior is overcome or reduced by introducing a<br />

very conductive layer below the ocean bottom (e.g., sediments), which has to be at least<br />

several kilometers thick. However, the assumption of an oceanic crust being entirely well<br />

conductive seems unrealistic; this is even more the case for the continental crust below<br />

the slope: every aquiclude would prohibit the intrusion of sea water into deeper layers.<br />

Another class of models makes the effect described above disappear too: the association<br />

of the upper part of the downgoing plate with a good conductor, in accordance with<br />

standard models of subduction.<br />

4 Summary<br />

Offshore magnetotellurics is an useful method for exploration of seafloor and oceancontinent<br />

subduction zones, where fluids and melts are known to control the subduction<br />

process. A combined on- and offshore magnetotelluric transect across subduction zone<br />

is assumed to be able to resolve high conductive structures associated with dehydration<br />

processes, water migration and melts building.<br />

However special conditions on the ocean floor not allow to interpret and to deal<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

207


with the marine records in the same manner as with onshore data. The presence of<br />

bathymetry distorts the magnetic and magnetotelluric transfer functions, particularly<br />

in the TE mode. Even a gently changing seafloor shape generates an enhanced concentrations<br />

of electric currents flowing above instead below measurement station and<br />

inducing on the ocean bottom a predominant anomalous opposite magnetic field. This<br />

results in phases exceeding the quadrant and cusps in the apparent resistivity.<br />

The high conductive sea water causes a strong attenuation of electric and magnetic<br />

field at short periods. Towards long periods the decay differs clearly for both fields. The<br />

electric field penetrates the ocean layer from surface to the seafloor, counterintuitive,<br />

nearly unchanged, while the magnetic field experiences a strong decay and reaches the<br />

ocean bottom just with a fraction of its surface value. Moreover the decay depends<br />

strongly from the resistivity contrast between ocean and seafloor. Reducing the resistivity<br />

of the basement the electrical thickness of layer increases and both fields approximate<br />

a field decay like in a homogeneous half space.<br />

References<br />

Brasse, H., G. Kapinos, Y. Li, L. Mütschard, W. Soyer, D. Eydam (2009): Structural<br />

electrical anisotropy in the crust at the South-Central Chilean continental margin as<br />

inferred from geomagnetic transfer functions, Phys. Earth Planet. Inter., 173, 7-16.<br />

Brasse, H. (2009): Methods of geoelectric and electromagnetic deep sounding, Lecture<br />

Notes, Free University of Berlin.<br />

Chave, A.D., S.C. Constable, and R.N. Edwards (1991): Electrical Exploration Methods<br />

for the Seafloor, in: Electromagnetic Methods in Applied Geophysics, Vol. 2 (Ed. M.N.<br />

Nabighian), Soc. Expl. Geophys., Tulsa, 931-966.<br />

Constable, S.C., K. Key, and, L. Lewis (2009): Mapping offshore sedimentary structure<br />

using electromagnetic methods and terrain effects in marine magnetotelluric data,<br />

Geophys. J. Int., 176, 431-442.<br />

Filloux, J.H. (1987):Instrumentation and experimental methods for oceanic studies, in:<br />

Geomagnetism, (Eds. J.A. Jacob), Academic Press,1987,143–248.<br />

Fischer, G. (1979): Electromagnetic induction effects on the ocean coast, Proc. IEEE,<br />

67, 1050-1060<br />

McNeill, J.D., and V. Labson (1986), Geological mapping using VLF radio fields. Electromagnetic<br />

Methods in Applied Geophysics. Volume 2, Appliction, Parts A and B,<br />

(Ed. M.N. Nabighian), SEG, Tulsa, 521-639.<br />

Mackie, R., J. Smith, and T.R. Madden (1994), Three-dimensional modeling using finite<br />

difference equations: The magnetotelluric example, Radio Science, 29, 923-935.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

208


Weidelt, P. (1994), Phasenbeziehungen für die B-Polarisation, in: Protokoll über das<br />

Kolloquium ”Elektromagnetische Tiefenforschung” (Eds. K. Bahr and A. Junge),<br />

Höchst, Odenwald, 60-65.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

209


A permanent array of magnetotelluric stations located at the<br />

South American subduction zone in Northern Chile.<br />

Introduction<br />

Dirk Brändlein 1,2 , Oliver Ritter 1,2 , Ute Weckmann 1,3<br />

1 <strong>GFZ</strong> German Research Centre for Geosciences, Potsdam, Germany<br />

2 Freie Universität Berlin, Fachrichtung Geophysik, Berlin, Germany<br />

3 University of Potsdam, Institute of Geosciences, Potsdam, Germany<br />

Monitoring the dynamic behavior of an active deep<br />

subduction system is focus of the Integrated Plate Boundary<br />

Observatory Chile (IPOC), a permanent array of combined<br />

geophysical and geodetic stations in Northern Chile which is<br />

operated since 2006 by the <strong>GFZ</strong> German Research Centre for<br />

Geosciences. Magnetotelluric (MT) data is gathered at seven<br />

out of a total of eleven observation sites.<br />

The MT set up consists of three component long period<br />

fluxgate magnetometers (GeoMagnet) and non-polarizing<br />

Ag/AgCl electrodes to measure all three components of the<br />

magnetic field and both horizontal components of the<br />

electric field. The signals of the electromagnetic fields are<br />

continuously sampled at a rate of 20 Hz and at four sites<br />

transferred via satellite link to the <strong>GFZ</strong> in Germany. The<br />

objective of the project is to monitor and analyze<br />

electromagnetic data to decipher possible changes in the<br />

subsurface resistivity distribution, e.g. as a consequence of<br />

large scale fluid relocation.<br />

Here, we present vertical magnetic transfer functions as time<br />

series over a time span of more than two years for the<br />

period range from 10 -1 to 10 4 seconds 1 . These vertical<br />

magnetic transfer functions are sensitive to lateral changes<br />

of electric conductivity in the subsurface. Some components<br />

of these transfer functions show frequency dependent<br />

variations with a periodicity of roughly one year. These<br />

effects are observed at all sites of the array.<br />

Figure 1: The IPOC-MT array in<br />

Northern Chile showing sites PB01<br />

to PB07 (red symbols). Blue stars<br />

indicate major earthquakes.<br />

Due to the extreme dry ground of the Atacama desert<br />

continuous monitoring the electric field is difficult. Contact<br />

resistances are on the order of MΩ and electrolyte is leaking. Different types of electrodes are<br />

currently being tested.<br />

1 Periods shorter than 10 seconds are not considered since this is the shortest period fluxgate magnetometers are able to resolve.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

210


Periodic variations in the MT monitoring data<br />

To identify changes in the subsurface we examine time series of vertical magnetic transfer functions<br />

Tx(ω) and Ty(ω):<br />

Bz(ω) = Tx(ω) Bx(ω) + Ty(ω) By(ω).<br />

These quantities are sensitive to lateral changes of electric conductivity in the subsurface.<br />

The value of each day and component of the transfer function time series is obtained by processing<br />

magnetic field time series of 3 days (using the day before and after). Subsequently the 3 day time<br />

window is forwarded by one day and the procedure is repeated until the entire time series is<br />

processed. The geomagnetic transfer functions were obtained using the robust processing described<br />

in Ritter el al. (1998), Weckmann et al. (2005), and Krings (2007).<br />

Figure 2: Time series showing the differences between daily values of Ty and the median of the<br />

entire data section of 730 days at each period for site PB03. For monitoring purposes it is<br />

desirable to amplify variations in the vertical magnetic transfer function components. Therefore,<br />

the median is calculated for each period and component and subtracted from the value of each<br />

day. The median of each frequency band is plotted on the right hand side. The median is more<br />

robust against outliers than the arithmetic average. Vertical red lines indicate major earthquakes<br />

(see Fig. 1), vertical blue lines indicate turn of the year. Interestingly, the ocean effect, which<br />

would be indicated by a large positive value of the median of Re[Ty], is rather small at site PB03.<br />

The time series of the real and imaginary parts of the Ty (east-west) component at PB03 show a<br />

remarkable frequency dependent feature (Fig. 2). At periods between approximately 40 and 1000 s<br />

the observed Re[Ty] values show a continuous variation of the amplitudes with a periodicity of<br />

roughly one year. The amplitudes vary in the range ± 0.1 and reach maximum values in the austral<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

211


winter before the winter solstice. In the austral summer, the minimum occurs before the summer<br />

solstice. Similar effects can be observed consistently at all sites of the array (Fig. 3).<br />

Figure 3: Time series of real (blue) and imaginary (red) parts of vertical magnetic transfer<br />

functions Tx and Ty of around 600 days at a period of 1000 seconds of sites PB01, PB02, PB03<br />

and PB05.<br />

Possible causes for the periodic variations of Ty<br />

Possible causes for this periodic variation of the Ty component could be source field<br />

inhomogeneities. The sources for the MT-method are the natural variations of the earth's magnetic<br />

field which are assumed to be far away from the observer so that electromagnetic waves penetrate<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

212<br />

Figure 4: Time series<br />

of the magnetic auto<br />

spectra of PB03,<br />

extracted from the 3day-average<br />

processing described<br />

in section 2. The<br />

ByBy * spectra show<br />

the same variations<br />

with a periodicity of<br />

one year as Ty.


the subsurface as plane waves.<br />

To examine if the variations in the time series of vertical magnetic transfer functions are caused by<br />

seasonal alterations of source fields we plotted the time series of magnetic auto spectra of PB03 (Fig.<br />

7) for the same time window as in Figure 2. These magnetic auto spectra show a clear correlation<br />

between the By-auto spectrum (east-west direction) and the Ty data (Fig. 8). This would support the<br />

idea that the results are influenced by source field effects.<br />

Sq variations and equatorial electrojet<br />

There exist several effects on the geomagnetic field<br />

which have a wide spectrum of variations<br />

(Onwumechili, 1997). Annual variations of the<br />

geomagnetic solar quiet daily variation (Sq) show the<br />

same periodicity as the measured variations of the<br />

magnetic auto spectra at PB03. The geomagnetic Sq<br />

field has a variation with a dominant periodicity of 1<br />

day. Solar radiation generates ionized molecules in<br />

parts of the ionosphere (around 80 to 300 km height)<br />

which produces variable charged particles and,<br />

consequently, conducting air. Additionally, the solar<br />

radiation causes thermo-tidal winds which move the<br />

conducting ionosphere through the geomagnetic field.<br />

The result is a system of electric currents depending<br />

on the position of the sun. Horizontal and vertical<br />

components of Sq tend to be predominantly semiannual<br />

in the zone of Equatorial Electrojet (EEJ) but<br />

predominantly annual at other latitudes (Onwumechili,<br />

1997).<br />

The IPOC MT array (21 - 23° south) in<br />

Northern Chile is located near the<br />

tropic of Capricorn (23° 26' south). This<br />

means, it is influenced by the<br />

Equatorial Electrojet (EEJ), a varying<br />

electric current flowing eastward in the<br />

ionosphere at a height of 100 - 130 km.<br />

Over South America the EEJ is warped<br />

to the south because it follows the<br />

magnetic dip equator (Fig. 8). The<br />

amplitudes are aligned to the north<br />

and to the south by return currents<br />

(Lühr et al.,2004) with peak values at<br />

latitudes some 5° away from the<br />

magnetic dip equator.<br />

Figure 5: Exemplary electric currents in<br />

the ionosphere cause Sq variations in<br />

the northern summer.<br />

http://geomag.usgs.gov/images/ionosp<br />

heric_current.jpg (10.02.2010)<br />

Figure 6: Electrojet current densities inferred from 2600<br />

passes of the CHAMP satellite over the magnetic<br />

equator between 11:00 and 13:00 local time.<br />

http://www.geomag.us/info/equatorial_electrojet.html<br />

(10.02.2010)<br />

Annual Sq variations are most likely the cause of the variations of the Y-component of the vertical<br />

magnetic transfer functions at the MT-array in Northern Chile.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

213


Non-periodic variations in the MT monitoring data<br />

No periodic long term variations can be identified in the Tx component of the vertical magnetic<br />

transfer function time series (Fig. 7). A continuous variation at a period of approximately 20 s can be<br />

observed in the Re[Tx] component. It has a high amplitude variation between -0.2 and +0.2.<br />

Figure 7: Time series of differences between daily values of Tx and median of the entire data<br />

section at each period at site PB03. Vertical red lines indicate major earthquakes, vertical blue<br />

lines indicate turn of the year. Values of median versus period are plotted on the right hand side.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

214<br />

Figure 8: Time series of<br />

differences between daily<br />

values of vertical magnetic<br />

transfer functions and<br />

median at site PB03 versus<br />

period. Length and period<br />

range correspond to the<br />

black box in figure 6.<br />

Vertical red lines indicate<br />

major earthquakes (Fig. 1),<br />

vertical blue lines indicate<br />

turn of the year.


Figure 9: Time series of differences between daily values of<br />

vertical magnetic transfer functions and median at site PB05<br />

versus period. Length and period range correspond to the black<br />

box of Fig. 6. Vertical red lines indicate major earthquakes (Fig.<br />

1), vertical blue lines indicate turn of the year.<br />

Note that the curve of the<br />

median, which represents the<br />

average transfer function,<br />

varies smoothly and<br />

consistently in this period<br />

range.<br />

A remarkable feature can be<br />

observed between two major<br />

earthquakes in December<br />

2007 (Fig. 7, black box). The<br />

area of interest shows a<br />

sudden rise in amplitudes up<br />

to ±0.25 (Fig. 8), which<br />

corresponds to 100% of the<br />

median value of Tx (Fig. 7,<br />

right hand side). A similar but<br />

less pronounced effect is<br />

observed at site PB05 (Fig. 9).<br />

More modeling is necessary<br />

to quantify the scale and location of structural changes of the electrical conductivity in the<br />

subsurface which could explain these non-periodic variations in the vertical magnetic transfer<br />

function time series. Calculation and modeling of inter station transfer function time series should<br />

also give more detailed information about changes in the conductivity structure of the underground.<br />

Acknowledgements<br />

We acknowledge funding from the <strong>GFZ</strong> German Research Centre for Geosciences. For continuous<br />

logistic help in Chile we want to thank Prof. Guillermo Chong and the people from the Universidad<br />

Catolica del Norte in Antofagasta. We are grateful to Günter Asch for his cooperation and help to<br />

install and maintain the technical infrastructure in the Atacama desert. We are very thankful for<br />

support in the field by Kristina Tietze and Thomas Krings.<br />

References<br />

Krings, T., 2007. The influence of Robust Statistics, Remote Reference, and Horizontal Magnetic Transfer<br />

Functions on data processing in Magnetotellurics. Diploma Thesis, WWU Münster ― <strong>GFZ</strong> Potsdam.<br />

Lühr, H., S. Maus, and M. Rother, 2004. Noon-time equatorial electrojet: Its spatial features as determined by<br />

the CHAMP satellite, Journal of Geophysical Research, 109, A01306, doi:10.1029/2002JA009656.<br />

Onwumechili, C. A., 1997. The Equatorial Electrojet, Gordon and Breach, Newark, N. J.<br />

Ritter, O., Junge, A. and Dawes, G., 1998. New equipment and processing for magnetotelluric remote<br />

reference observations. Geophysical Journal International, 132, 535-548.<br />

Weckmann, U., Magunia, A. and Ritter, O., 2005. Effective noise separation for magnetotelluric single site<br />

data processing using a frequency domain selection scheme. Geophysical Journal International, 161,<br />

635-652.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

215


Discussion on backarc mantle melting in the central Andean<br />

subduction zone, based on results of magnetotelluric studies<br />

D. Eydam ∗ & H. Brasse<br />

Free University of Berlin, Malteserstr. 74-100, 12249 Berlin<br />

Abstract<br />

Long-period magnetotelluric measurements in the Bolivian orocline revealed a large<br />

conductor in the backarc upper mantle of the central Andean subduction zone at 18◦S. Data analysis and interpretation by 2-D inversion has already been described by Brasse<br />

& Eydam (2008a) who interpreted this conductor as an image of partial melt triggered<br />

by fluid influx from the subducting slab.<br />

Remarkable are its high conductivities ranging up to above 1 S/m and implying high<br />

melting rates of more than 6vol%. However, regarding the distribution of intermediate<br />

depths seismicity which marks dehydration reactions in the subducting slab, one has<br />

to note the large distance of more than 60 km between fluid source and the assumed<br />

fluid-triggered partial melting of mantle peridotite. How do actual concepts of subduction<br />

zone dynamics fit to this resistivity image?<br />

There are at least two approaches which are both followed by in the subsequent discussion:<br />

first is to study the mobility of deep fluids in subduction zones, second is to<br />

study the mobility of the slab itself.<br />

Introduction<br />

Ocean-continent subduction zones are earth regions of major raw material recyclings.<br />

Hot magmas erupt nearby where cold oceanic lithosphere subducts. This apparently<br />

contradictory observation is explainable by consumption of deep slab-derived fluids for<br />

flux melting of the peridotitic mantle in the asthenospheric wedge:<br />

Hydrous minerals in the oceanic lithosphere are progressively dehydrating during subduction.<br />

The hereby released water may trigger partial melting of the overlying asthenospheric<br />

wedge by lowering mantle solidus generally at depths below 80 km. Melts<br />

and fluids ascent with some intermediate storage at density contrasts at the Moho<br />

or intra-crustal boundaries and give rise to arc and backarc volcanism. Due to their<br />

∗ now at <strong>GFZ</strong> Potsdam, Telegraphenberg, 14437 Potsdam<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

216


Eydam & Brasse: Discussion on backarc mantle melting<br />

enhanced conductivities, saline fluids and (hydrous) melt should principally be detectable<br />

by deep electromagnetic sounding methods, as was demonstrated, e.g., in the<br />

Cascadia/Juan de Fuca subduction zones (Kurtz et al., 1986; Wannamaker et al., 1989;<br />

Varentsov et al., 1996).<br />

Hydrous phases in the oceanic crust relevant for sub-arc water release are amphibole<br />

and lawsonite. Amphibole is abundant in the basaltic oceanic crust and decays only<br />

slightly pressure-dependent at 65-80 km depths, marking the transition from blueschist<br />

to eclogite facies. In contrast, lawsonite can be stable up to high pressures of 8.4 GPa<br />

(z ≈ 250 km) and dehydrates when temperatures exceed 700-800◦C. Mantle water release<br />

is mainly controlled by serpentinite dehydration between 500-700◦C. Serpentinite<br />

can be stable up to high pressures of 6 GPa where it transforms through a waterconserving<br />

reaction to the hydrous DHMS (Dense Hydrous Magnesium Silicat) -phase<br />

A.<br />

The amount of released water thus strongly depends on the thermal structure as well<br />

as on subduction geometry. Young and slowly subducting slabs may dehydrate completely<br />

whereas old and rapidly subducting lithospheres may import large amounts of<br />

mantle water into the deeper mantle, making them crucial for the global water cycle.<br />

The definitive amount of deep fluids in subduction zones is still uncertain because<br />

of the dependency of dehydrations from slab heating via mantle convection which in<br />

turn, strongly depends on water content in the wedge, besides other mostly poorly<br />

constrained parameters.<br />

Geological setting<br />

The central Andean subduction zone<br />

Along 7000 km length the oceanic Nazca plate subducts beneath the South American<br />

continental margin where the Andes evolved as a Neogene volcanic chain embedded in<br />

a compressive backarc tectonic setting. Holocene volcanism is seperated in four active<br />

segments, the austral, southern, central and northern volcanic zone (fig. 1).<br />

Characteristic for the whole subduction system are low slab dips of less than 30◦ and inactive volcanism in regions where aseismic ridges subduct. Those hot oceanic<br />

lithospheres probably flatten at main depths of sub-arc water release due to a retard<br />

of dehydration and thus density increase of the slab which may lead to ’horizontal subduction’<br />

over several hundreds of kilometres (Gutscher et al., 2000). Therefore, slab<br />

pull and buoyancy forces should be quite equilibrated probably making the slab mobile<br />

in geological time scales.<br />

The Bolivian orocline (13-28◦S) encompasses the central volcanic zone (CVZ) where<br />

the oldest part of the Nazca plate (55 Ma) subducts obliquely with an angle of N77◦E and with a current velocity of 6.5 cm/a (Klotz et al., 2006). Convergence rate has<br />

continuously slowed down from high 15 cm/a since the breakup of the oceanic Farallon<br />

plate in Nazca and Cocos plate 26 Ma ago (Somoza, 1998).<br />

During the highly compressive Miocene regime, stresses are mainly relaxed by extreme<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

217


Eydam & Brasse: Discussion on backarc mantle melting<br />

10°S<br />

20°S<br />

30°S<br />

40°S<br />

Nazca Rise<br />

Iquique Rise<br />

Juan-Fernández Ridge<br />

Chile Rise<br />

7.8 cm/a<br />

SVZ<br />

Easte<br />

Altiplano<br />

CVZ<br />

Puna<br />

80°W 75°W 70°W 65°W<br />

Figure 1: South American subduction zone with structural units of the central volcanic zone (CVZ)<br />

backarc and foreland. Hachures: Altiplano-Puna transition between 22-24 ◦ S; SB, SP - thick-skinned<br />

folding and thrusting in the Santa-Bárbara system and Sierras Pampeanas; P - Pica-Gap of Holocene<br />

volcanism (19.2-20.7 ◦ S) where Iquique Rise subducts (Wörner et al., 2000); SVZ - southern volcanic<br />

zone. Convergence rate after Somoza (1998).<br />

crustal shortening, thickening and uplift of the entire Andean arc and backarc region<br />

(e.g., Allmendinger et al., 1997; Scheuber et al., 1994; Elger et al., 2005). As a consequence<br />

the Altiplano-Puna high plateau developed, encompassing the highest volcanoes<br />

of the world (e.g., 6500 m for Sajama volcano) in the Western Cordillera to the west,<br />

rough mountain ranges in the Eastern Cordillera at the eastern boundary and the intramontanous<br />

drained Altiplano and Puna basins with thick Miocene stratas in-between.<br />

The up to 1800 km long and 350 km wide Altiplano-Puna plateau is the worldwide<br />

greatest high plateau in a subduction context. Crustal thickness is enormous (up to<br />

75 km; e.g., Wigger et al., 1994) and decreases beyond the plateau extension to normal<br />

values for continental lithosphere of 35-40 km. Likewise volcanism in the CVZ is highly<br />

crustally contaminated against more island-arc denoted lavas in the SVZ.<br />

Altiplano and Puna form an unique tectonomorphic entity within the orogen although<br />

their morphology differs. While the Altiplano in the north is rather flat and tectonically<br />

quiet, the Puna in the south is pervaded by many deep reaching, active horst- and<br />

graben structures. Average elevations are around 3700 m in the Altiplano and 4500 m<br />

in the Puna.<br />

Actual compressive tectonics are mainly restricted to the Andean foreland, where<br />

westvergent thin-skinned thrusting in the Subandean to the north change to thickskinned<br />

thrusting in the Santa Bárbara system and Sierras Pampeanas to the south<br />

(Allmendinger et al., 1997). At Altiplano latitudes, the comparatively fast westward<br />

subduction of the Brazilian Shield enhances the compressive backarc regime.<br />

P<br />

rn<br />

Cordi<br />

llera<br />

SP<br />

Suba<br />

SB<br />

nde<br />

an<br />

Chaco<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

218


Eydam & Brasse: Discussion on backarc mantle melting<br />

50 km<br />

75 km<br />

Pacific Ocean<br />

100 km<br />

125 km<br />

Argentina<br />

100 km<br />

CC<br />

LV<br />

150 km<br />

175 km<br />

125 km<br />

PC<br />

200 km<br />

Western Cordillera<br />

San Andrés fault<br />

P S<br />

225 km<br />

Coniri<br />

fault<br />

system<br />

Corque Basin<br />

Altiplano<br />

Rio Desaguadero<br />

Chuquichambi<br />

Eastern Cordillera<br />

Laurani fault<br />

Figure 2: Survey area with locations of the 30 MT sites in the Bolivian orocline. Profile trends roughly<br />

perpendicular to the strike of main morphological units and to the contour lines of the Wadati-Benioff<br />

zone (dotted lines). CC denotes Coastal Cordillera, LV Longitudinal Valley and PC Precordillera; P,<br />

S are Parinacota, Sajama volcanoes; mN is magnetic north.<br />

The survey area<br />

The magnetotelluric profile transverses the central Bolivian orocline from the Coastal<br />

Cordillera in northermost Chile, to the active volcanic arc and the Altiplano in central<br />

Bolivia and ends in the Eastern Cordillera. The profile follows a general trend of N48◦S which is roughly perpendicular to main structural units as well as to contour lines of<br />

the Wadati-Benioff zone (fig. 2).<br />

Main morphological units in the forearc are paralleling the trench and correspond to an<br />

ancient magmatic arc which shifted eastward through time, recordable since the Jurassic<br />

in the Coastal Cordillera, afterwards in the Longitudinal Valley, in the Precordillera<br />

and until the Neogene in the Western Cordillera (Scheuber et al., 1994). The exposure<br />

of the forearc crust to periods of extensive and long-lived magmatism gave rise to the<br />

formation of deep, long and wide fault zones (sub-)paralleling the trench among which<br />

some are still active like the Precordillera Fault System at 19.2-21◦S. Here, large copper<br />

deposits evidence strong fluid circulation and mineralisation after the eastward shift of<br />

the arc 32 Ma ago.<br />

In the closer survey area, the forearc is incised by deep, E-W running valleys which<br />

open to the ocean and may have served as Miocene drainage systems. Some of them<br />

may be associated with active faults (Wörner et al., 2000). A system of west vergent,<br />

steeply dipping thrust faults in the eastern Precordillera, the West-Vergent-Thrust-<br />

System, marks the transition to the high plateau and is thought to be the location of<br />

major plateau uplift (Muñoz & Charrier, 1996).<br />

The smooth monoclinal western slope of the high plateau is locally shaped by huge<br />

land slides, e.g., the Oxaya collaps 12 Ma ago (Wörner et al., 2002). In the West-<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

219<br />

thrust<br />

275 km<br />

250 km<br />

mN N


Eydam & Brasse: Discussion on backarc mantle melting<br />

ern Cordillera, Plio-Pleistocene to recent, andesitic to rhyodacitic stratovolcanoes are<br />

build on thick layers of ignimbrites which are widespread throughout the entire forearc<br />

and are dated between 22-19 Ma (Oxaya Ignimbrites) and to 2.7 Ma (Lauca-Pérez<br />

Ignimbrite) (Wörner et al., 2000).<br />

The Altiplano can be separated into the volcanic Mauri region to the west and in huge<br />

sedimentary basins easterly. The San-Andrés fault marks the boundary in-between and<br />

limits the Arequipa terrane to the east, a component of the Arequipa-Antofalla massif<br />

which comprises up to 2 Ga old rocks and underpins the volcanic arc probably along<br />

the whole plateau (see in James & Sacks, 1999). The most prominent feature in the<br />

Altiplano is the 80 km wide Corque basin, a deep asymmetric sedimentary basin with<br />

thick and folded tertiary strata, exceeding 10 km thickness (Hérail et al., 1997) and<br />

evidencing orogeny dynamics during Miocene’s pronounced compressive regime.<br />

To the east the basin is controlled by the Chuquichambi thrust system. The Coniri-<br />

Laurani fault system marks the transition to the Eastern Cordillera, the rough eastern<br />

flank of the plateau. Farther to the east the Interandean Zone and Main Andean Thrust<br />

separate the Altiplano plateau from the actual deformation front in the Subandean,<br />

with thin-skinned folding and thrusting holding the record of shortening rates within<br />

the study area (Allmendinger et al., 1997).<br />

After the Oligocene eastward shift of the magmatic arc, Neogene volcanism first relived<br />

in the Eastern Cordillera 27 Ma ago and afterwards encompassed the whole plateau region.<br />

Neogene volcanism of the Altiplano plateau is composed of three main units:<br />

Pliocene to recent stratovolcanoes are erupting in the main arc and Miocene to recent<br />

backarc calderas and mafic mongenetic pulses are distributed in the arc and backarc.<br />

Intensive ignimbritic volcanism erupted 10-4 Ma ago at the Altiplano-Puna transition<br />

between 21-24◦S and formed one of the greatest ignimbrite provinces of the earth, the<br />

Altiplano-Puna-Volcanic-Complex. Recent ignimbrite eruptions are older than 1 Ma.<br />

The high water, phenocrystal and silicate content signify long storage in huge crustal<br />

magma chambers with pronounced fluidal circulation.<br />

Holocene volcanic activity is less than in the Puna where the lithosphere probably is<br />

50 km thinner (Whitman et al., 1996). Volcanism in the closer survey area is categorised<br />

from dormant (Parinacota) to solfataras state with intensive fumarolic activity<br />

at Guallatiri volcano. The profile volcano Parinacota may possess a stable magma<br />

chamber with only minor magmatic input.<br />

Glance on data analysis and 2-D inversion<br />

We measured horizontal components of the electromagnetic field as well as the vertical<br />

magnetic field for long periods between 10 and 20 000 s. Data analysis and its interpretation<br />

by 2-D inversion has already been described by Brasse & Eydam (2008a) and<br />

will not be replicated in this article.<br />

Some important results are just mentioned: We could not resolve subduction-related<br />

features near the slab, like fluid curtains or fluid and melt pathways, due to significant<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

220


Re<br />

Period: 993 s<br />

Eydam & Brasse: Discussion on backarc mantle melting<br />

Figure 3: Real part induction arrows in<br />

Wiese convention for one chosen period<br />

demonstrating the remarkable good alignement<br />

of arrows in profile direction. Arrows<br />

in the forearc are deflected; they are pointing<br />

parallel to the coast.<br />

3-D effects in the forearc. This can easily be seen in the behavior of real part induction<br />

arrows pointing parallel to the coast (fig. 3) which is also observed in other coastal regions<br />

in Chile so that one might call them ’Chile arrows’. Brasse et al. (2008b) showed<br />

that they can be explained by anisotropy in the forearc crust. Seven forearc station<br />

data were thus discarded from further interpretation by 2-D inversion.<br />

Induction arrows as well as the analysis of impedance data, which reveals phasesensitive<br />

skew values below 0.3 for plateau sites, suggest fairly good electrical 2-D<br />

approximation for the plateau region. We observe a remarkable good alignement of<br />

middle to long period real part induction arrows in profile direction, as seen in figure 3.<br />

Correspondingly, electrical strike derived by impedance data is fairly perpendicular<br />

to the profile. Real arrows disappear in the central Altiplano where resistivities are<br />

signifcantly reduced for all sites and rotation angles and follow a well marked downward<br />

trend for longer periods indicating assumably well conducting features at greater<br />

depths.<br />

2-D inversions were performed with the inversion program of Rodi & Mackie (2001).<br />

Accounting for the consistency of induction arrows in the Altiplano, tipper weight was<br />

set high by assigning a small (absolute) error floor of 0.02. Error floors for resistivities<br />

and phases were set to 20% and 5% in order to overcome the static shift problem.<br />

Reliable models are achieved after numerous experiments inculding tests of starting<br />

models, specifying the regularisation term and further resolution tests (see Eydam,<br />

2008).<br />

Discussion on backarc mantle melting<br />

The final model, shown in figure 4, is obtained by jointly inverting tippers, TE and<br />

TM mode apparent resistivities and phases. Static effects were accounted for by using<br />

the program internal adjustment. Fair regularisation parameters lie around 20. Model<br />

RMS is 1.80, with larger misfits at the electrically less 2-D plateau borders (see Brasse<br />

& Eydam, 2008a).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

221


depth (km)<br />

Eydam & Brasse: Discussion on backarc mantle melting<br />

SW NE<br />

Moho<br />

600°C<br />

W. Cordillera Altiplano<br />

E. Cordillera<br />

fluids<br />

&<br />

melt<br />

A1 A2 A3<br />

C1<br />

B C2<br />

MASH - zone<br />

D<br />

mantle convection<br />

6-7vol% melt<br />

WB-zone<br />

San-Andrés<br />

Fault<br />

lithospheric<br />

mantle layer?<br />

not<br />

resolved<br />

distance (km)<br />

Coniri-Laurani<br />

Fault<br />

fluids & melt?<br />

Figure 4: Final resistivity model with suspectable fluid & melt dynamics - Ai: Corque and minor<br />

basins, B: Arequipa block(?), Ci: structures beneath the Eastern Cordillera, D: backarc mantle magma<br />

reservoir plus rise of fluids and melts. Circles indicate location of mayor earthquakes (M > 4.5, Engdahl<br />

& Villaseñor, 2002). WB denotes Wadati-Benioff zone. MASH-zone stands for melting, assimilation,<br />

storage and homogenisation zone of arc magmas. Below: RMS for the inverted station data.<br />

The large upper mantle conductor D is interpreted as an image of partial melts triggered<br />

by fluid influx from the slab. Conductivities range up to above 1 S/m and are<br />

definitely required by data down to 115 km depth, particularly to fit the downward<br />

trends in long period app. resistivities; they are consistent with data down to slab<br />

depths. This implies high melting rates of more than 6vol% (Eydam, 2008).<br />

Lithospheric mantle material should stabilise magma and trap hot fluids which otherwise<br />

would rapidly trigger wide scale crustal melting like this is the case in the southern<br />

Altiplano (e.g., ANCORP Working Group, 2003). This is consistent with seismic data<br />

from Dorbath & Granet (1996) who resolve lithospheric characteristics just beneath<br />

the Altiplano Moho.<br />

The source region of subduction related magmatism is offset from the arc by almost<br />

100 km, implying fluid and melt storage near the Moho (in so called MASH-zones) and<br />

non-vertical rise of only few fluidal and molten material towards the volcanoes. This<br />

accords to low volcanic productivity in the closer survey area, while lavas are bearing<br />

highly modified mantle signatures (Wörner et al., 1992). Magma and fluid lateral motion<br />

should be enforced by internal convections.<br />

This ’scenario’ implies a widely hydrated crust and upper mantle beneath the high<br />

plateau which is consistent with other geophysical anomalies, like high heat flow densities<br />

(e.g., Springer & Förster, 1998), neagtive Bouguer anomalies (Tassara et al., 2006),<br />

slow seismic velocities in a thick and highly absorbing crust (e.g., Beck & Zandt, 2002)<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

222<br />

?<br />

C3<br />

<br />

4<br />

3<br />

lg m 2<br />

1<br />

0


component of<br />

oceanic plate<br />

sediments<br />

crust<br />

mantle<br />

total<br />

Eydam & Brasse: Discussion on backarc mantle melting<br />

water import<br />

per m<br />

Nazca plate [kg]<br />

2<br />

1 700<br />

170 000<br />

250 000<br />

1 000 000<br />

421 700<br />

1 171 700<br />

in percentage of<br />

imported water<br />

20<br />

55<br />

60<br />

water release in 80-180 km depth<br />

per subducted<br />

meter [kg]<br />

340<br />

93 500<br />

150 000<br />

600 000<br />

243 840<br />

693 840<br />

per Ma [10 kg] 2) 6<br />

32<br />

8 883<br />

14 250<br />

57 000<br />

23 165<br />

65 915<br />

3)<br />

in neogene cycle,<br />

6<br />

WC = VA [10 kg]<br />

832<br />

230 958<br />

1) 1) 1) 1)<br />

2) 2) 2) 2)<br />

370 500<br />

1 482 000<br />

602 290<br />

1 713 790<br />

Table 1: Water budget for main dehydration depths in the central Andean subduction zone calculated<br />

by using 1) a standardised mantle hydration model after Rüpke et al. (2004), 2) a mantle hydration<br />

model deduced from local data after Ranero & Sallarès (2004) and 3) by using averaged subduction<br />

rates after Somoza (1998). WC - Western Cordillera, VA - main volcanic arc.<br />

and a strong mantle signature of geothermal fluids (Hoke et al., 1994).<br />

Estimation of arc and backarc water release<br />

A critical mind might state that the ability of water import is probably not sufficient<br />

to nurture such wide scale melting of mantle material. Table 1 shows the water balance<br />

for slab dewatering at main dehydration depths between 80-180 km, deduced from intermediate<br />

depths microseismicity published in David (2007) and seen in figure 7.<br />

For the calculation we used hydration estimates after Rüpke et al. (2004) modified by local<br />

data (fig. 5). Mantle hydration is poorly definable and depends on depth and density<br />

of bending related faults as well as on the thermal structure of the oceanic lithosphere.<br />

Therefore estimates based on regional seismic and gravimetric data evaluated by Ranero<br />

& Sallarès (2004) are four times higher than those assumed by Rüpke et al. (2004) who<br />

standardised hydration levels with regards to the age of the lithosphere. Water balances<br />

are presented for both models.<br />

The presented water releases are based on results of chemo-thermo-mechanical modellings<br />

performed by Rüpke et al. (2004) who solved for slab dehydration during sub-<br />

moho<br />

fault<br />

pelagic<br />

sediments<br />

sheeted dykes<br />

& pillow lavas<br />

1)<br />

100 m 7.3wt% H2O gabbro intrusions 4 km ~ dry<br />

serpentinized<br />

peridotite<br />

1 km 2.7wt% H O<br />

2km 1wt%HO<br />

2<br />

2)<br />

~23 km<br />

2<br />

2.5wt% H O<br />

Figure 5: Oceanic lithosphere with estimated water content in sediments, crust and mantle after<br />

Rüpke et al. (2004) modified by local data: 1) the thickness of sedimentary deposits after von Huene<br />

et al. (1999) and 2) mantle hydration after Ranero & Sallarès (2004).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

223<br />

2<br />

2)


Eydam & Brasse: Discussion on backarc mantle melting<br />

duction regarding latent heat consumption and neglecting shear heating1 (fig. 6).<br />

Thereafter, one column of one square metre Nazca lithosphere releases 700 t water at<br />

main dehydration depths (80-180 km) which sums up to 1.7 Gt water regarding the entire<br />

Neogene cycle of Nazca plate subduction starting 26 Ma ago (Sempere et al., 1990).<br />

Using the molar volume of water at 120 km depths, this mass fills a cross-section of<br />

more than 1500 km2 which finally lies in the dimension of the mantle magma reservoir.<br />

Phases of oroclinal bending might have been important for the water budget, because<br />

bending related faults should have opened parallel as well as perpendicular to the trench<br />

so that large amounts of water could have been fixed in the mantle.<br />

Slab-derived water in the mantle wedge should thus be sufficient to trigger a wide<br />

scale melting of mantle material, well noticing the presumably prior hydration of the<br />

Western Cordillera as pre-Neogene backarc. If the residual water, retained in the<br />

oceanic lithosphere, is released as well, the cross-section will expand for additional<br />

170 km 2 in case of complete crustal dehydration resp. 880 km 2 for complete mantle<br />

depth [km]<br />

retained water [%]<br />

100<br />

200<br />

100<br />

50<br />

T(°C)<br />

1300<br />

1100<br />

900<br />

700<br />

500<br />

300<br />

100<br />

100 200 300 400<br />

distance [km]<br />

crust<br />

sediments<br />

mantle<br />

50 100 150 200 250<br />

depth [km]<br />

Figure 6: Dehydration of oceanic lithosphere<br />

during subduction, presented above) as main<br />

depths’ water release of sediments, crust and<br />

mantle and below) as percentage of retained water<br />

respectively; after Rüpke et al. (2004).<br />

a water-conserving reaction to mantle phase A (Eydam, 2008).<br />

dehydration which points out the role<br />

of oceanic mantle for subduction zone<br />

dynamics.<br />

The double seismic layer shown in figure<br />

7 allows allocating the events to stability<br />

limits of mayor hydrous minerals in the<br />

oceanic lithosphere which can be used to<br />

predict slab temperatures at depths.<br />

Crustal seismicity stops around 140 km<br />

depth, whereas seismicity in the oceanic<br />

mantle is ongoing until the stability limit<br />

of serpentine in 180 km. Lawsonite is the<br />

only hydrous mineral in the crust which is<br />

stable down to 140 km depth. Lawsonite<br />

decay starts at the hot interplate boundary<br />

at shallower dephts and proceeds during<br />

subduction to deeper slab layers where temperatures<br />

at 140 km depth should exceed<br />

730◦C. Serpentine as the main hydrous<br />

mantle phase should still be present at<br />

180 km depth where it transforms through<br />

These estimates correspond to a fairly warm subduction system although the Nazca<br />

plate is quite old (55 Ma). Shear heating at the interplate boundary may elevate<br />

temperatures in the oceanic crust, because lubricates like sediments are sparse in the<br />

1 Modifying the mantle hydration model to higher estimates after Ranero & Sallarès (2004) will change<br />

the modelling results to even larger water releases than calculated in table 1, considering that dehydrations<br />

are mostly exothermic reactions.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

224


Eydam & Brasse: Discussion on backarc mantle melting<br />

central Andean subduction zone as a consequence of its arid climate (Lamb & Davis,<br />

2003). Thus more water than listed in table 1 may be released into the mantle wedge.<br />

Dehydrations finish below 200 km depth, suggesting that slab-derived fluids supply<br />

partial melting of mantle peridotite more than 60 km laterally away from their source<br />

region. Might this be possible?<br />

Discussion on fluid mobility<br />

There are several models describing fluid mobility in subduction zones like fluid channeling<br />

along cracks opened by dehydration induced hydrofracturing, lateral transports<br />

in the coupled mantle wedge, porous flows and thermally induced convections in the<br />

asthenospheric wedge.<br />

At considered depths hydrofractured faults shouldn’t reach far into the mantle and<br />

a coupled mantle wegde should be unrealistically thick to explain the requested<br />

transportation lengths. Porous flows start when interstitial fluids or melts interconnect<br />

depending on temperature, fluid salinity and melt viscosity. Deep slab-derived fluids<br />

are highly saline (Scambelluri & Philippot, 2001) and temperatures below 140 km<br />

depth definitively exceed 620◦C (see above), the minimum temperature of pure water<br />

interconnection in an olivine matrix (Mibe et al., 1999).<br />

Water enhances mantle convection via lowering the peridotitic solidus in the mantle<br />

wedge. This effect may be more pronounced than expected so far, considering the<br />

very slow reaction kinetics of olivine melting in laboratory experiments (Grove et al.,<br />

2006). The correction of peridotitic solidus to much lower temperatures, e.g. to 860◦C at 100 km depth, implies that first water-enriched melts may be generated just above<br />

the slab. The melt fraction should be minor due to fast segregation of the low-viscous<br />

melts into lower and hotter mantle regions.<br />

Additionally, water import in the central Bolivian orocline might have been temporarily<br />

elevated in phases of its oroclinal bending leading to a significant reduction of<br />

mantle viscosity and enabling effective material transports via thermal convections,<br />

like it is modelled for some general case studies by Gerya et al. (2006). Saline<br />

fluids and melts may be transported far into the hotter backarc mantle where melt<br />

fractions rise and segregation rates decrease and fluids and melts become detectable<br />

by magnetotellurics.<br />

Models of mantle convection deal with numerous low constrainted parameters, like<br />

the degree of mantle hydration, shear heating at the interplate boundary, latent<br />

heat produced by mineral reactions or advective heat transports by fluids and melts.<br />

Therefore mantle rheology is only poorly definable and viscosity values may differ over<br />

several magnitudes.<br />

Water release at intermediate depths plays thus an important role in subduction zone<br />

dynamics. Currie & Hyndman (2006) even postulate that hot backarc mantles are a<br />

fundamental characteristic of an ocean-continent subduction zone. And indeed, other<br />

magnetotelluric images from the central Andean backarc mantle seem to concur with<br />

this: Schwarz & Krüger (1997) modelled 0.02 S/m conductive mantle features at the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

225


Eydam & Brasse: Discussion on backarc mantle melting<br />

Profil<br />

E4<br />

E5<br />

Depth [km]<br />

Depth [km]<br />

0<br />

100<br />

Amp<br />

Ser<br />

600°C<br />

< 40°<br />

Law<br />

730°C<br />

E4<br />

200<br />

0 200 400<br />

0<br />

100<br />

E5<br />

200<br />

0 200<br />

Distance [km]<br />

400<br />

Figure 7: Local microseismic events (M > 1) after David (2007) with map of projection lines E4, E5<br />

and MT profile. Above: Events allocated to decay temperatures of hydrous minerals, this are in the<br />

crust: Amp - amphibole and Law - lawsonite and in the oceanic mantle: Ser - serpentine. Note the<br />

double seismic layer.<br />

transition to the Puna around 23 ◦ S and northwards at 20-21 ◦ S upper mantle material<br />

seems to be quite conductive as well with values over 0.05 S/m (Lezaeta, 2001), though<br />

electromagnetic fields are severely attenuated by a crustal high conductivity zone here<br />

(Brasse et al., 2002).<br />

Discussion on the mobility of the slab itself<br />

The Nazca plate subducts along the whole margin with small dip angles between 25-30 ◦<br />

or even horizontally in greater depths at some segments (so called ’flat subduction’)<br />

where volcanism above is inactive.<br />

Depth [km]<br />

0<br />

100<br />

Nazca Plate<br />

A (25-20 Ma)<br />

Flat to normal slab transition (25-20 Ma)<br />

uplifting<br />

volcanic front<br />

migration<br />

hydrated lithosphere<br />

200<br />

0 200 400<br />

Distance [km]<br />

Eastern<br />

Cordillera<br />

melt<br />

asthenospheric<br />

influx & slab -<br />

steepening<br />

600<br />

Figure 8: ’Mobile slab model’ for the central Andes after<br />

James & Sacks (1999). Due to retarded dehydration at<br />

intermediate depths, slab flattens and dewaters sparsely,<br />

insufficient to melt but to hydrate the base of the overlying<br />

lithosphere. Slab compacts and steepens again when<br />

hot asthenospheric material influx and triggers effective<br />

dehydrations; magmatism and volcanism above develop.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

226<br />

Thus, slab pull and buoyancy<br />

forces are fairly equilibrated and<br />

the slab may react sensible to density<br />

changes due to retarded dehydrations;<br />

it gets mobile in geological<br />

time scales.<br />

Such ’chemically controlled slab<br />

mobility’ was already discussed by<br />

James & Sacks (1999) and applied<br />

on central Andean history by supposing<br />

a period of flat subduction<br />

35-25 Ma ago, basically due to<br />

absent lava dates for this period<br />

(fig. 8).<br />

Remarkable is the slight slab steepening<br />

in the central orocline (fig. 7).


Depth [km]<br />

Depth [km]<br />

Depth [km]<br />

0<br />

100<br />

A<br />

Eydam & Brasse: Discussion on backarc mantle melting<br />

Nazca Plate<br />

Western<br />

Cordillera<br />

melt<br />

200<br />

oroclinal bending (16-10 Ma)<br />

0 200 400<br />

0<br />

100<br />

B<br />

formation of backarc<br />

melt reservoire (< 10-6 Ma)<br />

huge mantle<br />

water release<br />

200<br />

0 200 400<br />

0<br />

100<br />

C<br />

Altipano<br />

pulse of mafic<br />

volcanism<br />

melt thermal weakening<br />

melt<br />

fluid-melt-path<br />

melt<br />

200<br />

isolation of backarc<br />

melt reservoire (< 6 Ma)<br />

slab steepening<br />

0 200<br />

Distance [km]<br />

400<br />

Figure 9: Attempt to reconstruct Neogene Nazca<br />

plate subduction in the Bolivian orocline:<br />

A) Increased hydration of oceanic Nazca plate<br />

mantle due to significant oroclinal bending.<br />

B) 2-4 Ma later, highly hydrated mantle reaches<br />

dehydration depths of serpentine (100-180 km) where<br />

the released fluids trigger high-grade partial melting<br />

in the backarc. Backarc melts rise vertically as well as<br />

non-vertically to the arc via internal convections.<br />

C) Slab steepens during dehydration. The voluminous<br />

backarc reservoir persists but is maintained by<br />

less fluids. Volcanic productivity decreases.<br />

Tracing the regression line of shallow Wadati-Benioff events (z < 100 km) to greater<br />

depths, this line cuts the mantle conductor at dehydration depths of serpentine<br />

(z = 180 km), the most important hydrous mantle mineral which may bound huge<br />

amounts of water. Therefore backarc peridotitic melting may have been triggered<br />

just above the slab by enormous fluid influx whereby the slab was compacting and<br />

steepening slowly to its actual position (fig. 9). Today, the magma reservoir should be<br />

maintained by much less fluidal material and probably cools down.<br />

This scenario implies a prior period of extensive oceanic mantle hydration which might<br />

have occured during phases of oroclinal bending. Bending dates are hard to fix, probably<br />

block rotations in the closer survey area give a hint, they were dated at 12 Ma<br />

(Wörner et al., 2000).<br />

Using an average for ancient subduction rates of 10 cm/a, evaluated by Somoza (1998),<br />

and a constant slab dip of 25◦ , those huge amounts of fixed water will have started<br />

to deliberate after 2.4 Ma of subduction (z ≈ 100 km) and will have reached depths<br />

of high-grade partial backarc melting (z ≈ 180 km) further 1.9 Ma later. Surficial expression<br />

of high-grade mantle melting are Pliocene to recent pulses of mafic andesites<br />

erupting in the Bolivian Altiplano backarc (Davidson & de Silva, 1994). Furthermore<br />

backarc mantle magma might have been transported via internal convections to the<br />

sub-arc region and have also erupted at the main volcanic arc. However, the basal input<br />

into the deep magma system of the recent strato-volcanoes is only minor (Wörner<br />

et al., 2000).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

227


Eydam & Brasse: Discussion on backarc mantle melting<br />

Conclusion<br />

The huge backarc mantle conductor imaged by 2-D inversion of magnetotelluric data<br />

from the central Bolivian orocline inspired reflections about fluid and melt dynamics<br />

in the central Andean subduction zone. Among the two approaches considered here<br />

to explain the remarkable offset between intermediate depths mircoseismic events and<br />

the imaged magma reservoir, sub-arc and -backarc water release plays a key role for<br />

subduction zone dynamics.<br />

In the central Bolivian orocline huge amounts of water might have been released at<br />

main slab dehydration depths some million years after oroclinal bending occurred where<br />

faults should have opened not only parallel but also perpendicular to the trench allowing<br />

deep water infiltration and fixation in the oceanic mantle. This implies on the one hand<br />

that mantle viscosity is definitely lowered and leads to significant material transports<br />

due to thermal convections, whereby fluids and melts may be widespreaded in the<br />

arc and backarc mantle. On the other hand, enhanced slab dehydration and slab<br />

compacting may have been followed by slab steepening and the imaged mantle melting<br />

process isn’t supplied by much fluidal material anymore and therefore probably cools<br />

down.<br />

References<br />

Allmendinger, R.W., Jordan, T.E., Kay, S.M., & Isacks, B.L. 1997. The evolution of<br />

the Altiplano-Puna plateau of the Central Andes. Annual Review of Earth and Planetary<br />

Sciences, 25, 139–174.<br />

ANCORP Working Group. 2003. Seismic imaging of a convergent continental margin<br />

and plateau in the Central Andes (ANCORP’96). Journal of Geophysical Research, 108<br />

(B7), 2328, doi:10.1029/2002JB001771.<br />

Beck, S.L., & Zandt, G. 2002. The nature of orogenic crust in the Central Andes. Journal<br />

of Geophysical Research, 107(B10), doi:10.1029/2000JB000124.<br />

Brasse, H., & Eydam, D. 2008a. Electrical conductivity beneath the Bolivian Orocline and<br />

its relation to subduction processes at the South American continental margin. Journal of<br />

Geophysical Research, 113(B07109), doi:10.1029/2007JB005142.<br />

Brasse, H., Lezaeta, P., Rath, V., Schwalenberg, K., Soyer, W., & Haak, V.<br />

2002. The Bolivian Altiplano conductivity anomaly. Journal of Geophysical Research,<br />

107(B5), doi:10.1029/2001JB000391.<br />

Brasse, H., Kapinos, G., Li, Y., Mütschard, L., Soyer, W., & Eydam, D. 2008b.<br />

Structural electrical anisotropy in the crust at the South-Central Chilean continental margin<br />

as inferred from geomagnetic transfer functions. Physics of the Earth and Planetary<br />

Interiors, 173, 7–16.<br />

Currie, C.A., & Hyndman, R.D. 2006. The thermal structure of subduction zone back<br />

arcs. Journal of Geophysical Research, 111, doi:10.1029/2005JB004024.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

228


Eydam & Brasse: Discussion on backarc mantle melting<br />

David, C. 2007. Comportement actuel de l’avant arc et de l’arc du coude de Arica dans<br />

l’orogénèse des Andes Centrales. PhD thesis at university Toulouse III.<br />

Davidson, J.P., & de Silva, S.L. 1994. Late Cenozoic magmatism of the Bolivian Altiplano.<br />

Contributions to Mineralogy and Petrology, 119 (4), 387–408.<br />

Dorbath, C., & Granet, M. 1996. Local earthquake tomography of the Altiplano and<br />

the Eastern Cordillera of northern Bolivia. Tectonophysics, 259, 117–136.<br />

Elger, K., Oncken, O., & Glodny, J. 2005. Plateaustyle accumulation of deformation:<br />

Southern Altiplano. Tectonics, 24, doi:10.1029/2004TC001675.<br />

Engdahl, E.R., & Villaseñor, A. 2002. Global Seismicity: 1900-1999. In: Int. Handbook<br />

of Earthquake and Engineering Seismology, Part A (eds. Lee, W.H.K., Kanamori, H.,<br />

Jennings, P.C. & Kisslinger, C.). Academic Press, 665–690.<br />

Eydam, D. 2008. Magnetotellurisches Abbild von Fluid- und Schmelzprozessen in Kruste und<br />

Mantel der zentralen Anden. Diploma thesis at Freie Universität Berlin.<br />

Gerya, T.V., Connolly, J.A.D., Yuen, D.A., Gorczyk, W., & Capel, A.M. 2006.<br />

Seismic implication of mantle wedge plumes. Earth and Planetary Science Letters, 156,<br />

59–74.<br />

Grove, T. L., Chatterjee, N., Parman, S. W., & Médard, E. 2006. The influence of<br />

water on mantle wedge melting. Earth and Planetary Science Letters, 249, 74–89.<br />

Gutscher, M.-A., Spakman, W., Bijwaard, H., & Engdahl, E. R. 2000. Geodynamics<br />

of flat subduction: Seismicity and tomographic constraints from the Andean margin.<br />

Tectonics, 19(5), 814–833.<br />

Hoke, L., Hilton, D., Lamb, S., Hammerschmidt, K., & Friedrichsen, H. 1994. 3 He<br />

evidence for a wide zone of active mantle melting beneath the Central Andes. Earth and<br />

Planetary Science Letters, 128, 341–355.<br />

Hérail, G., Rochat, P., Baby, P., Aranibar, O., Lavenu, A., & Masclez, G.<br />

1997. El Altiplano Norte de Bolivia, evolución geológica terciaria. In: El Altiplano: ciencia<br />

y conciencia en los Andes, Actas 2. Simposio Internacional Estudios Altiplánicos, Arica<br />

1993. R. Charrier, P. Aceituno, M. Castro, A. Llanos and L.A. Raggi (eds). Universidad<br />

de Chile: Santiago. 33-44.<br />

James, D.E., & Sacks, I.S. 1999. Cenozoic formation of the Central Andes: A geophysical<br />

perspective. Society of Economic Geology, Special Publication, 7, 1–26.<br />

Klotz, J., Abolghasem, A., Khazaradze, G., Heinze, B., Vietor, T., Hackney,<br />

R., Bataille, K., Maturana, R., Viramonte, J., & Perdomo, R. 2006. Longterm<br />

signals in the present-day deformation field of the central and southern Andes and<br />

constraints on the viscosity of the earths upper mantle. In: The Andes: Active Subduction<br />

Orogeny. O. Oncken et al. (Eds.). Berlin: Springer. 65–89.<br />

Kurtz, R.D., Delaurier, J.M., & Gupta, J.C. 1986. A magnetotelluric sounding across<br />

Vancouver Island detects the subducting Juan de Fuca plate. Nature, 321, 596–599.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

229


Eydam & Brasse: Discussion on backarc mantle melting<br />

Lamb, S., & Davis, P. 2003. Cenozoic climate change as a possible cause for the rise of the<br />

Andes. Nature, 425, 792–797.<br />

Lezaeta, P. 2001. Distortion analysis and 3-D modeling of magnetotelluric data in the<br />

Southern Central Andes. PhD thesis at Freie Universität Berlin.<br />

Mibe, K., Fujii, T., & Yasuda, A. 1999. Control of the location of the volcanic front in<br />

island arcs by aquous fluid connectivity in the mantle wedge. Nature, 401, 259–262.<br />

Muñoz, N., & Charrier, R. 1996. Uplift of the western border of the Altiplano on a<br />

west-vergent thrust system, Northern Chile. Journal of South American Earth Sciences,<br />

9, 171–181.<br />

Ranero, C.R., & Sallarès, V. 2004. Geophysical evidence for alteration of the crust<br />

and mantle of the Nazca Plate during bending at the North Chile trench. Geology, 32(2),<br />

549–552.<br />

Rodi, W., & Mackie, R.L. 2001. Nonlinear conjugate gradients algorithm for 2-D magnetotelluric<br />

inversions. Geophysics, 66, 174–187.<br />

Rüpke, L.H., Morgan, J.P., Hort, M., & Connolly, J.A.D. 2004. Serpentine and the<br />

subduction zone water cycle. Earth and Planetary Science Letters, 223, 17–34.<br />

Scambelluri, M., & Philippot, P. 2001. Deep fluids in subduction zones. Lithos, 55,<br />

213–227.<br />

Scheuber, E., Bogdanic, T., Jensen, A., & Reutter, K.J. 1994. Tectonic development<br />

of the North Chilean Andes in relation to plate convergence and magmatism since the<br />

Jurassic. In: Tectonics of the Southern Central Andes. K.J. Reutter, E. Scheuber and P.J.<br />

Wigger (eds). Berlin: Springer. 121–139.<br />

Schwarz, G., & Krüger, D. 1997. Resistivity cross section through the southern Central<br />

Andes as inferred from magnetotelluric and geomagnetic deep soundings. Journal of<br />

Geophysical Research, 102, 11957–11978.<br />

Sempere, T., Hérail, G., Oller, J., & Bonhomme, J. 1990. Late Oligocene - early<br />

Miocene major tectonic crisis and related basins in Bolivia. Geology, 18, 946–949.<br />

Somoza, R. 1998. Updated Nazca (Farallon)-South America relative motions during the<br />

last 40 My: Implications for the mountain building in the central Andean region. Journal<br />

of South American Earth Sciences, 11(3), 211–215.<br />

Springer, M.H., & Förster, A. 1998. Heatflow density across the Central Andean subduction<br />

zone. Tectonophysics, 291, 123–139.<br />

Tassara, A., Götze,H.J.,Schmidt,S.,&Hackney,R.2006. Three-dimensional density<br />

model of the Nazca plate and the Andean continental margin. Journal of Geophysical<br />

Research, 111(B6), doi:10.1029/2005JB003976.<br />

Varentsov, I.M., Golubev, N.G., Gordienko, V.V., & Sokolova, E.Y. 1996. Study<br />

of deep geoelectrical structure along EMSLAB Lincoln-Line. Phys. Solid Earth, 32, 375–<br />

393.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

230


Eydam & Brasse: Discussion on backarc mantle melting<br />

von Huene, R., Weinrebe, W., & Heeren, F. 1999. Subduction erosion along the North<br />

Chile margin. Journal of Geodynamics, 27, 345–358.<br />

Wannamaker, P.E., Booker, J.R., Jones, A.G., Chave, A.D., Filloux, J.H., Waff,<br />

H.S., & Law, L.K. 1989. Resistivity cross-section through the Juan de Fuca subduction<br />

system and its tectonic implications. Journal of Geophysical Research, 94, 14127–14144.<br />

Whitman, D., Isacks, B.L., & Kay, S.M. 1996. Lithospheric structure and along-strike<br />

segmentation of the Central Andean Plateau: seismic Q, magmatism, flexure, topography<br />

and tectonics. Tectonophysics, 259, 29–40.<br />

Wigger, P.J., Schmitz, M., Araneda, M., Asch, G., Baldzuhn, S., Giese, P.,<br />

Heinsohn, W.D., Martinez, E., Ricaldi, E., Röwer, P., & Viramonte, J. 1994.<br />

Variation in the crustal structure of the Southern Central Andes deduced from seismic<br />

refraction investigations. In: Tectonics of the Southern Central Andes. K. J. Reutter and<br />

E. Scheuber and P. J. Wigger (eds). Berlin: Springer. 23–48.<br />

Wörner, G., Uhlig, D., Kohler, I., & Seyfried, H. 2002. Evolution of the West<br />

Andean Escarpment at 18 ◦ S (N. Chile) during the last 25 Ma: Uplift, erosion and collapse<br />

through time. Tectonophysics, 345, 183–198.<br />

Wörner, G., López-Escobar, L., Moorbath, S., Horn, S., Entenmann, J., &<br />

Davidson, J.D. 1992. Variaciones geoquímicas, locales y regionales, en el arco volcánico<br />

Andino del norte de Chile (17 ◦ 30’S-22 ◦ 00’S). Revista Geológica de Chile, 19(1), 37–56.<br />

Wörner, G., Hammerschmidt, K., Henjes-Kunst, F., Lezaun, J., & Wilke, H.<br />

2000. Geochronology ( 40 Ar/ 39 Ar, K-Ar and He-exposure ages) of Cenozoic magmatic<br />

rocks from Northern Chile (18-22 ◦ S): Implications for magmatism and tectonic evolution<br />

of the Central Andes. Revista Geológica de Chile, 27, 205–240.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

231


Thin Sheet Conductance Models from<br />

Geomagnetic Induction Data:<br />

Application to Induction Anomalies at the Transition from<br />

the Bohemian Massif to the West Carpathians<br />

VclavČerv a (vcv@ig.cas.cz), Světlana Kov čikov a , Michel Menvielle b and Josef Pek a<br />

a Inst. Geophys., Acad. Sci. Czech Rep., v.v.i., Prague, Czech Rep.<br />

b CEETP CNRS, Saint Maur des Fosses, France<br />

Abstract<br />

Thin sheet approximation of Earth’s conductive structures made it possible to quantitatively estimate effects of<br />

lateral conductivity variations in the Earth long before full 3-D electromagnetic modeling was practicable. The<br />

thin sheet approach is still useful when induction data with limited vertical resolution are to be interpreted on a<br />

surface. It especially refers to collections of long period induction arrows across large areas and geological units.<br />

For this purpose, inverse procedures, both linearized and stochatic, for a conductance distribution in a thin sheet have<br />

been suggested recently. We present a stochastic Monte Carlo inversion of geomagnetic induction data based on<br />

the bayesian formulation of the inverse problem. An example of the inversion of practical induction data from the<br />

transition zone between the Bohemian Massif and the West Carpathians suggests that an SW-NE anomalous induction<br />

zone observed above the eastern slopes of the Bohemian Massif admits explanation in terms of a phantom effect due<br />

to the superposition of fields of the strong SE-NW Carpathian conductivity anomaly to the east with NW-SE to W-E<br />

trending conductivity zones to the west that conform with the fault pattern of the eastern Bohemian Massif.<br />

1 Introduction<br />

Regional geoelectrical information on the contact zone between the Variscean Bohemian Massif and the Alpine Western<br />

Carpathians is mainly available from long-period geomagnetic transfer functions. Magnetotelluric data and broadband<br />

electromagnetic induction experiments are scarce in the region, mainly because of high level of the civilization<br />

noise all over the area, and also because of the lack of instrumentation in the past. Only recently, magnetotelluric<br />

experiments have been carried out in some subareas of the region, mainly for commercial targets (e.g., Voz r, 2005).<br />

Long-period geomagnetic induction data covering a period band of about one decade in the range of thousands<br />

of seconds cannot provide detailed geoelectrical information on structures beneath the region of interest. They are,<br />

however, suitable characteristics to be used to model large-scale horizontal conductivity distribution in the Earth’s<br />

crust or lithosphere, and thus to indicate regional lateral conductivity anomalies that are responsible for the observed<br />

induction pattern over the area under study. With data at periods of the order of thousands of seconds, with penetration<br />

depths starting at several tens of km for standard conditions of the continental lithosphere, a quasi 3-D thin sheet<br />

approximation of crustal conductivity structures is a reasonable induction model (e.g., Vasseur and Weidelt, 1977).<br />

The thin sheet approximation largely reduces the computational demands of the modeling procedure as compared with<br />

a full 3-D treatment and, moreover, bypasses some intrinsic difficulties of dealing with the geomagnetic induction data<br />

alone, especially their low sensitivity with respect to the normal layered background of the model.<br />

Recently, two methods of inversion of geomagnetic induction data for conductances in a thin sheet have been<br />

developed, one based on the non-linear conjugate gradient technique (Wang and Lilley, 2002) and the other on a<br />

Bayesian approach with Markov chain Monte Carlo (MCMC) method used for a stochastic sampling process (Grandis,<br />

2002). In this contribution, we are using the latter approach to analyze conductance distributions which are compatible<br />

with the geomagnetic induction data in the West Carpathians region and at its transition to the Bohemian Massif on<br />

the territories of the Czech Republic, Slovakia and Poland.<br />

More specifically, the data base of this study are induction response data at 150 field stations that were published<br />

earlier (Praus and Pěčov , 1991) covering the Bohemian Massif (BM), the Brunovistulicum (BV), and the West<br />

Carpathian region (WCP) and re-analyzed in (Pěčov and Praus, 1996). Spatial distribution of both the in-phase and<br />

the out-of phase induction vectors, contour maps of individual transfer functions (TF) and contour maps of anomalous<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

232


Latitude (deg)<br />

54<br />

53<br />

52<br />

51<br />

50<br />

49<br />

48<br />

47<br />

Mid-G<br />

Saxo-Thuringian<br />

Rheno-Hercynian<br />

Moldanubian region<br />

North-German-Polish Caledonides<br />

ine<br />

erma n C rystall R<br />

ise<br />

BM<br />

it of Alpi atio<br />

Alpine Front<br />

ne deform<br />

o r li<br />

ute<br />

outer limn<br />

mit of Varisceandefor tion<br />

East European<br />

Platform<br />

Holy Cross<br />

Mts<br />

8 10 12 14 16 18 20 22<br />

Longitude (deg)<br />

T<br />

eisseyre rnquist Line<br />

Vienna<br />

Basin<br />

ma<br />

-To<br />

Midmountains<br />

Pannonian<br />

Basin<br />

Bohemian Massif (BM ) with uni ts:<br />

Moldanubicum<br />

Teplá-Barrandian region<br />

Krušné hory-Thuringian<br />

region<br />

Bohemian Cretaceous<br />

basin<br />

West-Sudetic region<br />

Brunovistulicum (BV)<br />

including Silesicum, Moravicum<br />

and Brno Massif<br />

West Carpathians (WCP) with units:<br />

Tertiary basins<br />

(Carpathian foredeep,<br />

Pannonian basin)<br />

Outer (Flysch) Carpathians<br />

Klippen Belt<br />

Central (Inner) Carpathians<br />

Neovolcanites<br />

Figure 1: Geological scheme of the region of Central Europe relevant to our modeling study. The dashed-line rectangle<br />

shows the region in which the inversion for a laterally variable conductance distribution in a thin sheet is carried out.<br />

vertical field that were generated from the TF’s by the hypothetical field of different polarizations and systems of internal<br />

anomalous currents indicate the existence of two zones of anomalous induction at the eastern margin of the BM<br />

and near the boundary of the Carpathian plate (Kov čikov et al., 1997). The previous analyses, and the recent one<br />

performed by a new approach to imaging the induction data of the Wiese vectors at almost 1800 localities covering<br />

mainly the Central European area (Wybraniec et al., 1999) suggest that these anomalies might be connected with the<br />

North-German-Polish anomalous zone, representing an important part of the Trans-European Suture Zone (TESZ).<br />

The analysis of certain models of electrical conductivity distribution is performed to fit the anomalous features of<br />

the induction response data over the Central European area, specifically zones of anomalous induction in the eastern<br />

margin of the BM, across the entire block of the BV and near the margin of the Carpathian tectonic plate.<br />

The structure of the article is as follows: In Section 2, we give a short overview of the principal geological units of<br />

the region under study. Section 3 briefly summarizes the geoelectrical features of the region previously inferred from<br />

the geomagnetic induction data in the region. In Section 4, we formulate the thin sheet model used in the inversion.<br />

The principles of the Bayesian inversion of the geomagnetic induction data for the conductance distribution in the<br />

thin sheet are presented in Section 5. Section 6 then summarizes outputs of the Bayesian inversion for the induction<br />

data over the BM/BV/WCP region with indications on possible correlations of the conductance model with regional<br />

geological structures.<br />

2 Geological context of the induction studies<br />

Fundamental elements of the central European geological structure that are relevant to our modeling experiment are<br />

schematically displayed in Fig. 1 together with the major geological units over the Czech and Slovak Republics covered<br />

by induction response data involved in the modeling process.<br />

(i) The Tornquist-Teisseyre tectonic zone (TTZ) constitutes one substantial part of a number of fault zones and<br />

sutures found within the Trans-European Suture Zone (TESZ) that represents the most important geological boundary<br />

in Europe separating mobile Phanerozoic western terranes (Meso-Europe) from the Precambrian east European Craton.<br />

It is as clearly defined in the deep lithosphere as in the upper crust, Moho depths increase across this zone from 30 km<br />

beneath Variscan Europe to 45 km beneath the East European Craton. In contrast to the relatively cold eastern craton<br />

relatively high heat flow characterises Western Europe.<br />

(ii) The Bohemian Massif (BM) represents the easternmost consolidated block of the Variscan branch of the European<br />

Hercynides (Meso-Europe) that builds up Bohemia and the western part of Moravia. The major elements of<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

233


the BM are several SW-NE trending zones separated usually by deep-seated faults, which are reflected in the results<br />

of the deep seismic profiling, in the gravity, geomagnetic and heat flow maps (Suk et al.,1984). The fault structures<br />

are essentially parallel with the boundaries of individual Hercynian zones. The intersection of these zones with the<br />

second order system of NW-SE trending faults is responsible for the complicated block structure of the BM.<br />

(iii) The Carpathians belong to the young Tertiary Alpine-orogenic belt and they constitute the NE branch of the<br />

Alpides. Their boundary with the Eastern Alps runs along the Danube Valley. The northern boundary with the East<br />

European Platform is defined by the erosive margin of the flysch nappes. Our model covers the West Carpathians<br />

(WCP) and we distinguish here the inner (central) Carpathians, the Klippen zone, the outer (Flysch) Carpathians and<br />

the Carpathian foredeep (Fig. 1, dashed-line rectangle).<br />

(iv) The Brno unit, termed recently the Brunovistulicum (BV) (Dudek, 1980; Suk, 1995) lies in between those<br />

previously mentioned structural elements. The BV unit is assumed to be an independent geological structure forming<br />

the Precambrian (Cadomian) basement (Palaeo-Europe) of both the eastern part of the Hercynides (Variscides) of the<br />

BM and the Alpides, i.e. of the WCP in Moravia. Recent geophysical and geological data have shown that the BV<br />

and the whole Moravian block occupy a highly independent position and form a separate geological unit belonging<br />

probably to the Fenno-Sarmatian Platform (Dudek, 1980).<br />

3 Geoelectrical data and features of the region<br />

3.1 The data<br />

The experimental data were collected in 1970’s and 1980’s in a series of field experiments organized in co-operation<br />

with Czech, Slovak and Polish colleagues. The field measurements consisted in analogue recordings of magnetic<br />

transient variations at alltogether 150 temporary field sites over an area of about 500 × 250 km 2 . They were analysed<br />

in terms of induction arrows (Wiese, 1962). At each station, in-phase and out-of phase induction arrows have been<br />

estimated for periods in the range of 1200 to 5840 s. They correspond to the real and imaginary components of singlestation<br />

transfer functions between the horizontal and vertical components of the transient magnetic field at the station.<br />

A sample of the real and imaginary induction arrows at a particular period of T = 3860 s is presented in Fig. 2.<br />

3.2 Major conductivity anomalies<br />

As the magnetic Z-component is known to be highly sensitive to laterally inhomogeneous distribution of the internal<br />

electrical conductivity, maps of these vectors provide us with a view of the changing anomalous behaviour of the<br />

Z-variation as function of frequency and location. Reversals of the arrows distinguish zones of anomalous induction,<br />

which often mark important geological features such as contacts between blocks with different geological histories of<br />

development, zones of past and recent tectonic activities, collision zones and etc. A qualitative analysis of the maps<br />

of induction arrows led to evidence two major conductivity anomalies in the area under study. Quantitative modeling<br />

allowed to characterize the conductivity distribution in the lithosphere that accounts for the observed induction arrows.<br />

The Carpathian anomaly, WCA The induction vectors across the WCP region show a clear perpendicular orientation<br />

with respect to a general trend of the anomalous induction zone localized along the external margin of the<br />

Carpathians Mts. chain (see WCA in Fig. 2. They show almost a perfect 180 o reversals of their azimuths above the<br />

anomalous induction zone. In the WCP, modules of the induction vectors situated to the south from the zero line are<br />

by about 25-50% larger than the corresponding induction vectors located to the north of the anomaly. The Carpathian<br />

14˚ 15˚ 16˚ 17˚ 18˚ 19˚ 20˚ 21˚ 22˚ 23˚<br />

51˚ 51˚<br />

Real arrow 0.5<br />

50˚ 50˚<br />

49˚<br />

WCA<br />

49˚<br />

EBMA<br />

48˚ 48˚<br />

14˚<br />

15˚<br />

16˚<br />

?<br />

17˚<br />

18˚<br />

19˚<br />

20˚<br />

21˚<br />

22˚<br />

23˚<br />

14˚ 15˚ 16˚ 17˚ 18˚ 19˚ 20˚ 21˚ 22˚ 23˚<br />

51˚ 51˚<br />

Imag arrow 0.5<br />

50˚ 50˚<br />

49˚ 49˚<br />

48˚ 48˚<br />

Figure 2: Sample of experimental real and imaginary induction arrows in the BM/BV/WCP region for the period of<br />

3860 s, with two main regional conductivity anomalies indicated, the West Carpathian conductity anomaly (WCA)<br />

and the conductivity anomaly on the eastern margin of the Bohemian Massif (EBMA).<br />

14˚<br />

15˚<br />

16˚<br />

17˚<br />

18˚<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

234<br />

?<br />

19˚<br />

20˚<br />

21˚<br />

22˚<br />

23˚


geoelectrical anomaly is constrained by the presence of high-conductivity rock series at depth of 10-25 km in different<br />

segments of the orogen. Different geological models explaining the presence of highly conductive rocks or solutions<br />

as sources of the anomalies have been suggested. The role of altered and/or fractured rocks, saturated with hot mineral<br />

waters, as well as the proximity of anomaly sources to boundaries of different crustal blocks, including those between<br />

the Carpathians and adjacent platforms, have been taken into account ( d m and Posp il, 1984; Jankowski et al.,<br />

1984; 1991; 2008; Hvo dara and Voz r, 2004). Other authors discuss possible connection between the Carpathian<br />

anomaly and the presence of metamorphosed coal bearing Carboniferous strata beneath the orogen or relation to the<br />

graphitized rocks occurring close to the boundaries of crustal blocks (Glover and Vine, 1992; Glover and Vine, 1995;<br />

´Zytko, 1997).<br />

The EBMA anomaly The induction vectors on the eastern margin of the BM are distinguished by the vectors<br />

oriented predominantly parallel to the general SW-NE trend of the anomalous zone. Also the reversal of the azimuths<br />

at the anomalous zone is rather poorly developed, only sudden changes of azimuths are observed at individual profiles.<br />

These facts are clear indications of a 3-D character of the conductivity distribution. The EBMA has been generally<br />

attributed to processes in the the subduction of the eastern margin of the BM beneath the WCP plate, but precise<br />

induction processes for generating the peculiar 3-D features of this anomaly are still unknown (e.g. Kov čikov et al.,<br />

2005). Speculating on the physical/geological sources of the anomaly may thus be premature.<br />

3.3 Results from numerical modeling<br />

From the equivalent current systems (Červ et al., 1997) it was concluded that the source depth of the induction<br />

anomalies can be about 18 km in the WCP region and about 10 km in the EBM/BV. These estimates are suggesting the<br />

source of the anomalies at shallower depth than those obtained previously by separating the magnetic field variations<br />

into internal and external parts (Pěčov and Praus, 1996) and applying the line current approximation (Jankowski et<br />

al., 1985).<br />

2-D models for induction vectors along the profiles crossing the WCP are summarized in (Jankowski et al., 1985).<br />

The models featured anomalous bodies with a cross-section× conductivity parameter of the order of 10 7 to 10 8 Sm,<br />

with the top of the bodies at depths beneath 12-15 km. In (Jankowski et al., 1991) the 2-D modeling was used for<br />

simultaneous modeling of the induction vectors and apparent resistivities, collected in a series of sites in the Polish<br />

section of the Carpathian Foredeep. In the latter models, the source of the WCA was situated at shallower depths, less<br />

than about 10 km, and hypothesized to be related to deep sediments of the Carpathian Foredeep. The 2-D inversion on<br />

the Carpathian data was firstly used in (Červ and Pek, 1981). The geoelectrical structural model along the DSS profile<br />

No VI, crossing the BM, BV and WCP, based on the MT and MV results was presented in (Červ et al., 1984).<br />

3.4 Depth of the asthenosphere<br />

The electrical asthenosphere, if present, is an additional structural feature that can affect the induction data especially<br />

at long periods. In the model derived from the P-wave residuals for the Bohemian Massif the depth of the lithosphereasthenosphere<br />

transition zone are between 90 and 140 km (Babu ka et al., 1988). From MT sounding a layer of<br />

increased electrical conductivity attributable to the asthenosphere was interpreted at depth between 100 and 150 km<br />

(Červ et al., 1984).<br />

The most likely depth of the conducting layer in the upper mantle in the Pannonian Basin region are between<br />

60 and 85 km in the central part of the depression. The depths seem to increase towards the flanks of the basin to<br />

about 100 km ( d m, 1976). In some parts (R ba-Ro nava tectonic line) the thickness of the litosphere reduces to<br />

even less than 60km ( d m, 1988). The regions of lithosphere thinning penetrate from the Pannonian Basin into<br />

inner parts of the WCP in several promontories. In the thinned part of the lithosphere in the West Carpathians the<br />

lithosphere-asthenosphere transition zone is at the depth 90-120 km (Červ et al., 1984).<br />

A gradual increase of the lithosphere thickness to 140-180 km occurs in the Outer Carpathians and father towards<br />

the margin of the East European Platform (Praus et al., 1990). In the Alps the depth of the asthenosphere varies from<br />

100 to 200 km (Praus et al., 1990).<br />

4 A possible thin sheet model<br />

In the subsequent inversion for a laterally non-uniform conductance, we will use a thin sheet model formally defined<br />

as follows:<br />

Let us consider a model consisting in a heterogeneous thin sheet at the surface of or embedded in a 1-D medium,<br />

hereafter called normal model. The normal model is defined by the conductivities σn(r) at any point r. In the thin<br />

sheet, the actual conductivity σ(r) differs from the normal one in a domain of interest Ω. InsideΩ, σa(r) denotes the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

235


difference σ(r) − σn(r): σa(r) is the anomalous conductivity that is zero outside Ω. Solutions of the forward and<br />

inverse problems have already been published for such a model. They are briefly recalled in the next section.<br />

With regard to the experimental geomagnetic TF’s available, several aspects of the thin sheet model must be<br />

considered. First, the validity of the thin sheet hypothesis must be assessed. We assume that the anomalous electrical<br />

targets are localized mainly in the crust. Penetration depths for typical continental crustal sections should not be less<br />

than about 50 km in the period range of thousands of seconds (corresponds to the resistivity of 10 Ωm and period<br />

of 1000 s). Thus, the anomalous zone of 20-30 km below the Earth’s surface can be reasonably (though, may be,<br />

not completely unquestionably for highly conductive anomalous zones) approximated by a thin sheet. According to<br />

Bruton (1994), the size of the square tiles a used to discretize the anomalous subdomain Ω of the sheet should meet<br />

the condition aωμSmax ≪ 1. For the most extreme parameters in our models, say Tmin = 1200 sandSmax =<br />

5000 siemens, the size of the tiles should be much less than about 30 km. We typically use tiles with a =20to 25 km<br />

in our models, which meets the Bruton’s condition for most of the model situations.<br />

The second aspect is the depth of the thin sheet. As we are not able to employ a multiple sheet model in the<br />

inversion at present, we use a single thin sheet that integrates all the conductivity anomalies from the surface down to<br />

the crust. Thus, the sheet is situated on the surface of the Earth in our experiments. This may result in increased misfit,<br />

especially in imaginary induction arrows, if the anomalous current flow deep in the crust.<br />

The third aspect is the embedding normal structure. Due to large differences of the depth vs. resistivity sections<br />

of the two main region of our study, the BM and the WCP with the adjacent Pannonian Basin, we cannot suggest<br />

any single 1-D normal model down to the asthenospheric depths for the whole region. Therefore, we simplified the<br />

normal model into a two-layer structure with a poor conductor resistivity of several hundreds of Ωm up to 1000 Ωm)<br />

underlaid by a more conductive asthenosphere with resistivity within the range of 100-500 Ωm. The effect of the<br />

topography of the asthenospheric layer was checked independently by a 3-D modeling experiment with seismic data<br />

(Praus et al., 1990) taken to approximate the top of the asthenosphere. The checks showed that the effect of the<br />

asthenosphere is negligible in induction arrows for periods of the order of thousands of seconds unless the resistivity<br />

of the asthenosphere is less than about 10 Ωm.<br />

5 Bayesian Monte Carlo Markov Chain thin sheet inversion<br />

We present in this section the Bayesian Monte Carlo Markov Chain (MCMC) method we used to solve the inverse<br />

problem. For the sake of self-completeness, let us first briefly recall basic notions concerning Bayesian inversion and<br />

Markov chains behaviour. The readers are referred to Roussignol et al. (1993), Menvielle and Roussignol (1995), and<br />

Grandis et al. (1999; 2002), and references therein for more details.<br />

5.1 The Bayesian approach<br />

Let the a priori knowledge be the information available on the model before processing the data, and the a posteriori<br />

knowledge the information available after processing the data. A priori and a posteriori distributions account in a<br />

probabilistic way for the a priori and a posteriori knowledge respectively. In the Bayesian context, solving the inverse<br />

problem thus comes down to determining the a posteriori knowledge by updating the a priori knowledge with the<br />

information gleaned from the data (Box et Tiao, 1973; Berger, 1985; Press et al., 1989).<br />

Solving the inverse problem first requires the direct problem to be solved. Let F be the direct problem function<br />

which enables computation of the observations d for a model m. Assuming that the error δd is only related to the<br />

data acquisition, it becomes<br />

d = F (m)+δd<br />

The a priori knowledge is given by a probability distribution function (pdf ) P0(M = m | M ∈M) defined on<br />

the set of possible models M. The a posteriori probability for the parameter vector M to take the value m given the<br />

observations d is given by Bayes’ formula (Bayes, 1763; also see, e.g., Bolstad, 2004, for a more modern treatment).<br />

Noting P (M = m | D = d, M ∈M) the conditional probability of M given D, and assuming that the error δd is<br />

gaussian with standard deviation τ, it becomes<br />

<br />

<br />

||F (m) − d)||2<br />

exp − P0(m)<br />

P (M = m|D = d, M ∈M)=<br />

<br />

m∈M<br />

exp<br />

<br />

−<br />

2τ 2<br />

||F (m) − d)||2<br />

2τ 2<br />

<br />

P0(m)<br />

The value d of the random vector D corresponds to the observed data. Since d is not modified during the inversion<br />

process, we will from now on omit D in the expression of the probability distributions and denote the a posteriori pdf<br />

P (M = m|D = d, M ∈M) simply by Π(m).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

236<br />

(1)


The normalization constant, which appears in the denominator of eq. (1) is a sum over the set of possible models,<br />

the dimension of which is very large: for instance, it is equal to M × L for models with M parameters that can each<br />

take on L different values. It is actually very difficult, if not impossible to estimate directly. We choose to use a Monte<br />

Carlo Markov Chain (MCMC) simulation method to achieve this estimation.<br />

5.2 Markov chains<br />

In order to estimate the pdf Π(m), let us define a Markov Chain on the set M of possible models that have Π(m)<br />

as invariant probability. The marginal pdf ’s of Π(m) will be estimated by empirical averages from simulations of the<br />

Markov Chain.<br />

Consider a system which can take a certain number of states and evolves at random with time. At a given time,<br />

the state of the system can be described as a random variable M, the values of which, m, belong to M. The system is<br />

a Markovian process if, at any time, its future evolution depends only on its present state. It means that a Markovian<br />

system depends only on its past through its present state. Markov chains are a particular class of Markovian processes<br />

such that (i) the set M is finite or countable, and (ii) the successive times for the evolution are denoted by integers.<br />

Aseries{M(n),n =0, 1, 2,...,N} of random variables with values in a finite or countable space is then a Markov<br />

chain if the state of the system at the time n +1, M(n +1), depends only on its past through the state of the system<br />

at time n, M(n). The behaviour of a Markov chain is defined by the set of its transition probabilities,<br />

Pn(m, m ′ )=P (M(n +1)=m ′ | M(n) =m) ,<br />

where m and m ′ belong to M. If (i) the transition probabilities do not depend on time n, (ii) M is finite, and (iii) each<br />

possible state can be reached from any other, the chain is a homogeneous ergodic aperiodic Markov chain, and there<br />

exists one, and only one probability density function Π on M which is invariant for the Markov chain. When n<br />

increases towards infinity, the behavior of ergodic Markov chains is such that the average fraction of time at which the<br />

chain is at a state m (m ∈M) tends towards the invariant probability of m,i.e.Π(m).<br />

Methods for Bayesian inversion with Markov chains has been proposed for the 1-D magnetotelluric (Grandis et<br />

al., 1999) and DC problems (Schott et al., 1999) with the a priori of smooth variation of resistivity with depth, and for<br />

the thin sheet approximation (Roussignol et al. 1993; Grandis, 1994; Grandis et al., 2002). The present analysis of<br />

crustal conductivity over the transition from the Bohemian Massif to the West Carpathians relies on the inversion of<br />

induction arrows using the latter method, which is briefly described in the following sections.<br />

5.3 The thin sheet forward problem<br />

Let us consider a model consisting of a heterogeneous thin sheet at the surface of or embedded in a 1-D medium,<br />

specified in detail in the first two paragraphs of Section 4 above. Let E(r,ω,σ) and H(r,ω,σ) be the time Fourier<br />

transforms of the electric and magnetic field at point r and circular frequency ω, ω =2π/T, T period, for the actual<br />

conductivity structure σ. Using the Green kernel method, we can rewrite basic equations of electromagnetism as<br />

(Weidelt, 1975)<br />

<br />

E(r,ω,σ) = En(r,ω) − iωμ σa(r<br />

Ω<br />

′ ) G(r, r ′ ,ω) E(r ′ ,ω,σ(r ′ )) d 3 r ′ , (2)<br />

<br />

H(r,ω,σ) = Hn(r,ω)+ σa(r ′ )curl(G(r, r ′ ,ω)) E(r ′ ,ω,σ(r ′ ) d 3 r ′ , (3)<br />

Ω<br />

at any point r and any frequency ω. The quantity μ is the magnetic permeability of the vacuum in our models.<br />

G(r, r ′ ,ω) is a 3 × 3 complex matrix, named Green kernel, and represents the Fourier transform at frequency ω of the<br />

electric field created at point r by a unit dipole δ(r ′ ) located at point r ′ (Morse and Feshbach, 1953), curl(G(r, r ′ ,ω))<br />

is a 3 × 3 complex matrix obtained by taking the curl at point r of the field of the column vectors of the matrix<br />

G(r, r ′ ,ω).<br />

A numerical solution of the forward problem has been first proposed by Vasseur and Weidelt (1977) for a superficial<br />

thin sheet. The Vasseur and Weidelt’s (1977) approach can be extended to a thin sheet at any depth below the surface.<br />

The code was made operational by Tarits (1989). These solutions are based upon a digitization of Ω in K small square<br />

cells Pk, k = 1,...,K, in which conductivities, Green kernels, electric and magnetic fields are nearly constant.<br />

Let S(Pk) and Sn(Pk) be the actual and normal integrated conductivities (i.e. conductances) of the cell Pk, and<br />

Sak = S(Pk) − Sn(Pk), k =1,...,K, the anomalous conductance of Pk. Then basic equations become, using the<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

237


same notations for the sake of simplicity,<br />

E(r,ω,S) = En(r,ω) − iωμ <br />

H(r,ω,S) = Hn(r,ω)+ <br />

k<br />

k<br />

Sak G(r,Pk,ω) E(Pk,ω,S(r ′ )) |Pk|, (4)<br />

Sak curl(G(r,Pk,ω)) E(Pk,ω,S(r ′ )) |Pk| (5)<br />

at any point r and any frequency ω.<br />

Let us consider the linear system made up of equation (4) written at each point Pk of Ω, k =1,...,K.Given<br />

the distribution of anomalous conductances Sak, k =1,...,K, the quantities E(Pk,ω,S(r ′ )) are solutions of this<br />

system. For each frequency, the system has 6 × K equations and 6 × K real unknowns in the general case. Solving<br />

this system allows us to compute the electric field E(r,ω,S(r)) at any point r and frequency ω. The magnetic field<br />

H(r,ω,S(r)) is then derived using (5) written at any point r and frequency ω.<br />

5.4 The inverse problem<br />

The structure of the embedding 1-D model, i.e., Sn(r), the depth of the heterogeneous thin sheet, the domain Ω, and<br />

the cells Pk are first defined, and remain fixed during the inversion. Their values are to be deduced from previous<br />

investigations in the studied area (geology, other geophysical methods, ...) and from laboratory measurements of<br />

rock conductivity. The conductances Sk, k =1,...,K, are the parameters, and solving the inverse problem consists<br />

in estimating their a posteriori pdf given the data d =(dij), i =1,...,I, j =1,...,J, and the a priori pdf P0.<br />

The data d are actually estimates of a function T [ E(r,ω,S(r)), H(r,ω,S(r)) ] at points ri, i =1,...,I,ofthe<br />

surface and frequencies ωj, j =1,...,J, plus an experimental noise δd,<br />

dij = T [ E(ri,ωj,S(r)), H(ri,ωj,S(r)) ] + δdij. (6)<br />

The δdij, and therefore the dij are assumed to be independent gaussian random variables with zero mean value and<br />

standard deviation τ.<br />

In situations for which no particular information is available, we choose as a priori distribution the product of a<br />

uniform pdf on each layer. In this case P0 is a uniform distribution over all the possible models and where the different<br />

parameters are independent. P0 is digitised over a set of conductance values, hereafter called possible conductance<br />

values, the choice of which depends on the a priori knowledge of the considered medium. There is no constraint on<br />

this choice, but it is clear that the larger this number the better the determination of the a posteriori distribution, but<br />

the greater the computer time. The possible conductance values may depend on the cell. We will only consider here<br />

situations with the same number L of possible values Sl, l =1,...,L, for each cell. The a posteriori distribution will<br />

accordingly be expressed as the a posteriori probability of these possible conductance values.<br />

The a posteriori pdf of the parameters S is estimated by means of a Markov chain. Let an algorithm that considers<br />

each cell Pk successively and updates the value of the parameter for this cell with a transition probability equal to the<br />

conditional probability distribution of Sl, l =1,...,L, given the actual values of the parameters for the other cells,<br />

the data d and the a priori pdf P0(S). The sequence of images thus obtained after each scanning of the whole set of<br />

K cells is a homogeneous Markov chain, because the probability of an image depends only on the previous image,<br />

and does not change with step n. This process is a Markov chain on M called the Gibbs sampler (Robert, 1996).<br />

It is ergodic, and its invariant probability Π is the a posteriori probability is the conditional probability Π(S) of the<br />

parameters given the data d and the a priori pdf P0(S) (Roussignol et al., 1993; Roussignol and Menvielle, 1995;<br />

Robert, 1996; Grandis et al., 1999; 2002).<br />

In the present case, the required L × K solutions of the forward problem that are necessary for computing the<br />

transition probability of the Markov chain (in the present case, the K conditional pdf ’s corresponding to the whole set<br />

of cells) requires a very long CPU time. In order to limit the CPU time, the forward problem is not fully solved at each<br />

scanning, and the electric field is estimated by the Markov chain. The Markov chain is then a two-dimensional one,<br />

which estimates a quantity (the conductance) distributed over a finite set of real values and another one (the electric<br />

field) distributed over a continuous set of real vectors.<br />

The asymptotic behavior of such a Markov chain has first been studied empirically by Jouanne (1991), then Grandis<br />

(1994) for situations representatives of those encountered in geophysics. Using synthetic models, Grandis (1994)<br />

evidenced that the empirical average of the transition pdf of this Markov chain tends towards an invariant pdf that corresponds<br />

to the a posteriori pdf of the parameters, provided the estimate of the electric field at each scanning is precise<br />

enough for the transition probabilities to be reasonably estimated. Later on, Touijar (1994) gave the first demonstration<br />

that such a chain converges under given conditions. The physical interpretation of Touijar’s mathematical results<br />

suggests that the empirical results established by Jouanne (1991) and Grandis (1994) are likely to be valid for most of<br />

geophysical situations (Menvielle and Roussignol, 1995).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

238


6 The inversion<br />

6.1 The description of the model and data<br />

In accord with the above discussion, we chose the following particular parameters for the thin sheet model used in the<br />

MCMC inversion of the long-period geomagnetic induction data over the BM/BV/WCP region:<br />

(i) The anomalous domain of the superficial thin sheet is divided into 21(N)×29(E) square cells, each 25×25 km 2<br />

in size, covering thus a region of about 500(N)×750(E) km 2 . This gives in total 609 variable parameters (cell conductances)<br />

for the inversion.<br />

(ii) The normal model was composed of a two-layer background model with the first layer of 100 km/1000 Ωmand<br />

a uniform basement of 300 Ωm. The basement was used to simulate the asthenospheric layer. The topography of the<br />

asthenosphere was neglected. The layered model was overlain by a uniform infinite thin sheet with the conductance<br />

of 300 siemens.<br />

(iii) For the inversion purposes, the 150 experimental geomagnetic TF’s available were interpolated into the centers<br />

of the sheet cells. As reliable data variances were not available for the experimental TF’s, we assumed the data to be<br />

contamined with Gaussian errors with the standard deviation of 0.03.<br />

(iv) For the MCMC run, data for only one period, T = 3860 s, were used to keep the computation times in<br />

reasonable limits. Fit of the data and model arrows for further periods available were only checked by direct modeling<br />

tests.<br />

6.2 The MCMC procedure, convergence, output models, post-processing<br />

For the MCMC procedure (Gibbs sampler), the conductance of each of the 609 variable cells could take on one of<br />

12 predefined values from the interval of 3 to 5000 siemens. The Markov chains were typically running for 1000 cycles<br />

of the Gibbs sampler, which took about 10 hours of the CPU time on a standard PC station (Intel Xeon 3 GHz, 1 MB<br />

RAM). Repeated MCMC runs were carried out to simulate parallel chains and check the convergence of the MCMC<br />

process. First 500 cycles of each of the MCMC chains were discarded as a burn-in period during which the chain<br />

approaches a vicinity of the solution and stabilizes. In our runs, the chains did not change substantially after the burnin<br />

period, and simulated parallel chains converged to similar conductance distributions for the same cells. From this<br />

rough diagnostics, we can conclude that the chains converged and can provide reasonable estimates of the means of<br />

the cell conductances within the selected model class M. As the number of MCMC iteration cycles is relatively small<br />

in our experiments, it is not realistic to obtain reasonably accurate estimates of higher moments of the aposteriori pdf ’s<br />

for the cell conductances. In fact, only highly qualitative conclusions may be made as regards the uncertainties of the<br />

variables.<br />

The MCMC procedure gives a series of models which should be probabilistically distributed in accord with the<br />

experimental data. We obtain a whole histogram of conductance values for each cell, which shows how much likely<br />

are the particular conductance values in the cell considered. After long enough iteration process, the histograms<br />

approximate the marginal aposteriori probabilities of the discrete conductance values in the individual sheet cells. We<br />

can estimate the mean values of the cell conductances by integrating over the histogram, as well as the most probable<br />

cell conductances (maximum aposteriori probabability, or MAP estimates) by taking the most frequent value from the<br />

histogram.<br />

Fig. 3 displays sheet models over the BM/BV/WCP region set together from the mean values of the conductance<br />

in the individual sheet cells and from the MAP estimates of the cell conductances. These models do not represent<br />

the average or MAP models from the chain, which would both require integration of the multi-dimensional empirical<br />

posterior pdf over the whole parameter space. The presented models are put together from estimates based on individual<br />

marginal pdf ’s for the individual sheet cells. It is clear that these models may differ considerably from each<br />

other, especially in regions where the cell conductances are poorly constrained by the data. Nevertheless, the models<br />

in Fig. 3 represent simple integrated information on the conductance structure projected from a 3-D probabilistic (histogram)<br />

image of the cell conductances provided by the whole underlying Markov chain. Fit of the model data from<br />

the ‘average’ model in Fig. 3 (left) and the experimental induction arrows is shown in Fig. 4 for the inverted period<br />

of 3860 s. Model vs. experiment fit for longer periods (not shown) is very similar, larger discrepancies are observed<br />

for the period T = 1200 s with model induction arrows substantially smaller than the experimental data. This is most<br />

likely due to a misspecified depth of the sheet in our model (superficial sheet).<br />

Qualitatively, we can also estimate the uncertainty of the conductance by simply observing the ‘flatness’ or ‘peakiness’<br />

of the histograms for the cell conductances. In Fig. 5 (left), we show the ‘average’ model from Fig. 3 (left)<br />

modified by keeping in it only those cells for which 90% of the cell conductance values in the MCMC histograms fall<br />

within an interval of a width of 2/3 of a decade. All remaining cells, with flatter conductance histograms, have been<br />

discarded. The cells in Fig. 5 (left) thus indicate regions of the model in which the conductance is reasonably well constrained<br />

by the data. More precise, quantitative uncertainty estimation would be perhaps possible from longer MCMC<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

239


chains which would allow us to integrate directly for variance-covariance functions estimates for the parameters.<br />

Because of high computational demands, thin sheet models for the MCMC inversion are mostly designed with<br />

only a coarse tiling. In the model post-processing stage, spatial smoothing often results in a more realistic image of<br />

the conductance distribution. The smoothed image has to be checked by an additional direct modeling test. We show<br />

a smoothed version of the ‘average’ conductance model from Fig. 3 (left) in the right-hand side panel of Fig. 5.<br />

Figure 3: Left: Conductance model over the BM/BV/WCP region designed from average cell conductances obtained<br />

form the stabilized part of the Markov chain. Right: Conductance model designed from maximum aposteriori conductances<br />

in the individual sheet cells. The color scale corresponds to the logarithm of the conductances. Period of<br />

inverted data was 3860 s.<br />

14˚ 15˚ 16˚ 17˚ 18˚ 19˚ 20˚ 21˚ 22˚ 23˚<br />

51˚ 51˚<br />

50˚ 50˚<br />

49˚ 49˚<br />

48˚ 48˚<br />

14˚<br />

15˚<br />

16˚<br />

17˚<br />

18˚<br />

19˚<br />

20˚<br />

21˚<br />

22˚<br />

23˚<br />

14˚ 15˚ 16˚ 17˚ 18˚ 19˚ 20˚ 21˚ 22˚ 23˚<br />

51˚ 51˚<br />

50˚ 50˚<br />

49˚ 49˚<br />

48˚ 48˚<br />

Figure 4: Fit of the real (left) and imaginary (right) induction arrows for the period of 3860 s. White arrors show the<br />

input data obtained by interpolating the experimental TF’s into the centers of the sheet cells. Orange-yellow arrows<br />

were generated by the thin sheet model designed from average cell conductivities from the MCMC chain (model to<br />

the left in Fig. 3).<br />

7 Tectonic correlation<br />

Interpretation of narow-band long-period geomagnetic data by the MCMC sampling can only provide a large-scale<br />

image of the conductance distribution which conforms with the observed TF’s. The conductance pattern restored from<br />

the induction arrows collected above the eastern slopes of the BM and its transition to the WCP gives a geologically<br />

plausible image of the region. The Carpathian conductivity anomaly is reconstructed in detail, suggesting possible<br />

weakening/local interruption at the crossing with the Central Slovakia fault zone (near 19.2 o E in Fig. 5 right) where<br />

other geophysical fields and geological indicators show discontinuous features on a presumably strike-slip fault (e.g.,<br />

Kov č and H t, 1993). Though in coarse mesh, the high conductance anomaly well fits the position of a low resistivity<br />

layer from the magnetotelluric data collected in the esternmost sector of the model, close 22 o E (Voz r, 2005).<br />

The induction anomaly at the eastern margin of the Bohemian Massif seems to have a more complex explanation<br />

than that of a quasi-linear conductor along the SW-NE zone of a rapid change of direction of the induction arrows<br />

(see Fig. 2). Two features may be observed in the model in Fig. 5 (right), which were already discussed earlier by<br />

Kov čikov et al. (2005). First, it is a conductive zone in the NE of the BM, which may generate induction arrows<br />

directed towards the SW, as clearly observed all across the area to the west of the EBMA. Second, it is an alteration<br />

of conductive and resistive E-W zones in the eastern part of the BM. Though still not clear how much of this effect<br />

14˚<br />

15˚<br />

16˚<br />

17˚<br />

18˚<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

240<br />

19˚<br />

20˚<br />

21˚<br />

22˚<br />

23˚


14˚ 15˚ 16˚ 17˚ 18˚ 19˚ 20˚ 21˚ 22˚ 23˚<br />

51˚ 51˚<br />

50˚ 50˚<br />

49˚ 49˚<br />

48˚ 48˚<br />

14˚<br />

15˚<br />

16˚<br />

17˚<br />

18˚<br />

19˚<br />

20˚<br />

21˚<br />

22˚<br />

23˚<br />

14˚ 15˚ 16˚ 17˚ 18˚ 19˚ 20˚ 21˚ 22˚ 23˚<br />

51˚ 51˚<br />

50˚ 50˚<br />

49˚ 49˚<br />

48˚ 48˚<br />

14˚<br />

15˚<br />

16˚<br />

17˚<br />

18˚<br />

19˚<br />

20˚<br />

21˚<br />

22˚<br />

23˚<br />

4<br />

3<br />

2<br />

1<br />

0<br />

log10(S)<br />

Figure 5: Left: Model from Fig 3 (left) with only those cells of the original ‘average’ model shown that have 90% of<br />

all conductance values within 2/3 of a decade. Cells with flat histograms are eliminated. Right: Spatially smoothed<br />

‘average’ conductance model. The dashed lines indicate main regional fault zones.<br />

may be caused by the profile arrangement of the data, this quasi-anisotropic domain is actually required by the data. It<br />

helps in feeding the induction process all across the Moravia region and in keeping the relatively large modules of the<br />

induction arrows in that region. If these induction sources could be verified, the regional induction in the eastern part<br />

of the BM would be mainly driven by the NW-SE to W-E pattern of tectonic zones (Elbe fault zone, Sudety faults,<br />

Poˇr č -Hronov fault zone, Odra fault zone, etc.) rather than by the SW-NE zones which conform with the tectonics of<br />

the transition to the WCP.<br />

8 Conclusion<br />

From the experience accumulated with the MCMC inversion we can conclude that for moderately sized thin sheet<br />

models (tens of cells in each direction), the MCMC inverse procedure is practicable with standard computer facilities.<br />

Speeding-up the forward solutions by using a 2-D Markov chain derived from the integral equation numerical approach<br />

by Vasseur and Weidelt (1977) is essential for the algorithm.<br />

Though computationally still demanding, the advantage of the MCMC is that it provides a probabilistic output<br />

for the inverse solution. Parameters can be assessed with respect to their values and uncertainty ranges, though true<br />

uncertainties (variance-covariance matrices) are hard to obtain in problems with demanding direct solutions. Effective<br />

visualization of complete probabilistic outputs from the MCMC algorithm may still be one of its weak points.<br />

The conductance pattern restored by the MCMC sampling from the induction arrows collected above the eastern<br />

slopes of the BM and its transition to the WCP gives a geologically plausible image of the region. The Carpathian<br />

conductivity anomaly is reconstructed in detail. The induction anomaly suggested at the eastern margin of the BM<br />

admits an alternative explanation in terms of a NW-SE to W-E conductance patterns, hypothetically coinciding with<br />

the fault pattern of the eastern Bohemian Massif. A detailed verification of the structural hypotheses in this region is,<br />

however, hardly possible by employing the long period geomagnetic induction data alone.<br />

Acknowledgements<br />

Financial assistance of the Czech Sci. Found., contract No. 205/07/0292, and of the Grant Agency Acad. Sci. Czech<br />

Rep., contract No. IAA300120703, is highly acknowledged.<br />

References<br />

d m, A., 1976. Results of deep electromagnetic investigations. In: d m (Editor), Geoelectric and Geothermal<br />

Studies, KAPG Geophysical Monograph, Akad miai Kiad , Budapest, 547-559.<br />

d m, A. and Posp il, L., 1984. Crustal conductivity anomalies in the Carpathian region. Acta Geodaet. Geophys.<br />

Montanist. Hung., 19, 19-34.<br />

d m, A., Landy, K. and Nagy, Z., 1988. New evidence for the special distribution of the electric conductivity in the<br />

Earth’s crust and upper mantle in the Pannonian basin as a ‘hot spot’, Tectonophysics, 164, 361–368.<br />

Babu ka, V., Plomerov , J. and Pajdu k, P., 1988. Lithosphere-asthenosphere in Central Europe: Models derived<br />

from P-residuals, in Proceedings of the Fourth Workshop on the European Geotraverse (EGT) project: The Upper<br />

Mantle, G. Nolet and B. Dost (Eds.), ESF, Strasbourg, 37-48.<br />

Berger, J. O., 1985. Statistical Decision Theory and Bayesian Analysis, Springer-Verlag, New York.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

241


Bolstad, W. M., 2004. Introduction to Bayesian Statistics, John Wiley & Sons, 2004.<br />

Box, G. and Tiao, G., 1973. Bayesian Inference in Statistical Analysis, Addison-Wesley, Reading.<br />

Bruton P., 1994. Testing Vasseur and Weidelt’s thin sheet algorithm, in Proceedings of the 15th Colloquium ‘Electromagnetic<br />

Depth Research’, Bahr, K. and Junge, A. (Eds.), DGG, pp. 54-59.<br />

Červ, V. and Pek, J., 1981, Numerical solution of the two-dimensional inverse geomagnetic induction problem. Studia<br />

geoph. et geod., 25, 69-80.<br />

Červ, V., Pek, J. and Praus, O., 1984. Model of geoelectrical anomalies in Czechoslovakia. J. Geophys., 55, 161-168.<br />

Červ, V., Pek, J. and Praus, O., 1987.Numerical modelling of geoelectrical structures in Czechoslovakia. Phys. Earth<br />

Planet. Inter., 45, 170-178.<br />

Červ, V., Pek, J. and Praus, O., 1990. Limitations of the unimodal thin sheet modelling of near-surface inhomogeneities<br />

in Czechoslovakia, Phys. Earth Planet. Inter., 60, 100-114.<br />

Červ, V., Kov čikov , S., Pek, J., Pěčov , J. and Praus, O., 1997. Model of Electrical Conductivity Distribution across<br />

Central Europe, J. Geomag. Geoelectr., 49, 1585-1600.<br />

Dudek, A., The crystalline basement block of the Outer Carpathians in Moravia: Brunovistulicum, Ropr. Cs. Akad.<br />

Ved, R. mat. a pr. ved, Praha, 90, 1980, 1-85.<br />

Glover, P.W.J. and Vine, F.J., 1992. Electrical conductivity of carbon-bearing granulite at raised temperatures and<br />

pressures, Nature 360, 723-726.<br />

Glover, P.W.J. and Vine, F.J., 1995. Beyond KTB-Electrical conductivity of the deep continental crust, Surveys in<br />

Geophysics, 16, 5-36.<br />

Grandis, H., 1994. Imagerie electromagnetique Bayesienne par la simulation d’une chaine de Markov, Doctorat<br />

d’Universit , Universit Paris VII.<br />

Grandis, H., Menvielle, M. and Roussignol, M., 1999. Bayesian inversion with Markov chains—-I. The magnetotelluric<br />

one-dimensional case, Geophys. J. Int., 138, 757–768.<br />

Grandis, H., Menvielle, M. and Roussignol, M., 2002. Thin-sheet electromagnetic inversion modeling using Monte<br />

Carlo Markov Chain (MCMC) algorithm, Earth Planets Space, 54, 511–521.<br />

Hvo dara, M. and Voz r, J., 2004. Laboratory and geophysical implications for explanation of the nature of the<br />

Carpathian Conductivity Anomaly. Acta Geophysica Polonica, Vol. 52, No. 4, pp. 497-508.<br />

Jankowski, J., Petr, V., Pěčov , J. and Praus , O., 1984. Geoelectric anomaly in the Czechoslovak-Polish section of the<br />

Carpathians on the basis of geomagnetic and magnetotelluric soundings, Acta Geodaet., Geophys. Montanist. Hung.,<br />

19, 81-91.<br />

Jankowski, J., Tarlowski, Z., Praus, O., Pěčov , J. and Petr, V., 1985. The results of Deep Geomagnetic Soundings in<br />

the West Carpathians, Geophys. J. Royal Astron. Soc., 80, 561-574.<br />

Jankowski, J., Pawliszyn, J., J wiak, W. and Ernst, T., 1991. Synthesis of electric conductivity surveys performed<br />

on the Polish part of the Carpathians with geomagnetic and magnetotelluric sounding methods, Publs. Inst. Geophys.<br />

Pol. Ac. Sci., A-19(236),183-214.<br />

Jankowski, J., J ´zwiak, W. and Voz r, J., 2008. Arguments for ionic nature of the Carpathian electric conductivity<br />

anomaly, Acta Geophys., 56, 455-465.<br />

Jouanne, V., 1991. Application des techniques statistiques Bayesiennes a l’inversion de donn es lectromagn tiques,<br />

Doctorat d’Universit , Universit Paris VII.<br />

Kov č, P. and H k, J., 1993. The Central Slovak Fault System-=field evidence of a strike slip. Geol. Carpath., 44,<br />

133-159.<br />

Kov čikov , Červ, V., Praus, O., 1997. Modelling of geomagnetic transfer functions on the eastern margin of the<br />

Bohemian Massif, Annales Geophysicae, 15, Suppl. I, C 159.<br />

Kov čikov , S., Červ, V. and Praus, O., 2005. Modelling of the conductance distribution at the eastern margin of the<br />

European Hercynides, Studia geophys. geod., 49, 403-421.<br />

Menvielle, M. and Roussignol, M., 1995. Imagerie lectromagn tique bayesienne. In: Munschy, M., Sauter, D.,<br />

Schlich, R. (Eds.), Comit National Fran ais de G od sie et G ophysique, Rapport Quadriennal 1991–1994, Paris,<br />

France, pp. 177–190.<br />

Pěčov , J. and Praus, O., 1996. Anomalous induction zones in the Czech republic in relation to large scale European<br />

anomalies, Studia geoph. et geod., 40, 50-76.<br />

Praus, O., Pěčov , J., 1991. Anomalous geomagnetic fields of internal origin in Czechoslovakia, Studia geoph. et<br />

geod., 35, 81-89.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

242


Praus, O., Pěčov , J., Petr, V., Babu ka, V. and Plomerov , J., 1990. Magnetotelluric and seismological determination<br />

of the lithosphere-asthenosphere transition in Central Europe, Phys. Earth Planet. Inter., 60, 212-228.<br />

Press, S.J., Flannery, B.P., Teukolsky, S.A. and Vetterling, W.T., 1989. Numerical Recipes: the Art of Scientific<br />

Computing, Cambridge University Press, Cambridge.<br />

Robert, C., 1996. M thodes de Monte Carlo par Chaines de Markov, Economica, Paris.<br />

Roussignol, M., Jouanne, V.,Menvielle, M. and Tarits, P., 1993. Bayesian electromagnetic imaging, in Computer<br />

Intensive Methods, pp. 85–97, Hardle, W. and Siman, L. (Eds.), Physical Verlag.<br />

Schott, J.-J., Roussignol, M., Menvielle, M. and Nomenjahanary, F. R., 1999. Bayesian inversion with Markov<br />

chains—-II. The one-dimensional DC multilayer case, Geophys. J. Int., 138, 769–783.<br />

Suk, M., Bl kovsk , Buday, M., Chlup č, I., Cicha, I., Dudek, A., Dvoˇr k, J., Eli , M., Holub, V., Ibrmajer, J.,<br />

Kodym, O., Kukal, Z., Malkovsk , M., Menc k, E., M ller, V., Tyr ček, J., Vejnar, Z., Zeman, A., 1984. Geological<br />

history of the Czech Socialist Republic, Publ. of stˇredn stav geologick , 17-86.<br />

Suk, M., 1995. Regional geological division of the Bohemian Massif, Exploration Geophysics, Remote Sensing and<br />

Environment, II. 2., 25-30<br />

Touijar D., 1994. Stochastic algorithms and application to Bayesian electromagnetic imaging, PhD Thesis, Travaux<br />

Universitaires–Thčse nouveau doctorat, Universit de Lille 1, Villeneuve-d’Ascq, France, 104 pp.<br />

Vasseur, G. and Weidelt, P., 1977. Bimodal EM induction in non-uniform thin sheets with an application to the<br />

Northern Pyrenean Anomaly, Geophys. J. R. astr. Soc., 51, 669-690.<br />

Voz r, J., 2005. Magnetotelluric Soundings on the Territory of Slovakia, PhD Thesis, GFU SAV Bratislava, 2005, 108<br />

pp. (in Slovak).<br />

Wang, L. J. and Lilley, F. E. M., 2002. Inversion of magnetometer array data by thin-sheet modelling, Geophys. J.<br />

Int., 137, 128-138.<br />

Weidelt, P., 1975, Electromagnetic induction in three-dimensional structures, J. Geophys., 41, 85-109.<br />

Wiese, H., 1962. Geomagnetische Tiefentellurik Teil II: Die Streichrichtung der Untergrundstrukturendes elektrischen<br />

Widerstandes, erschlossen aus geomagnetischen Variationen, Geofisica pura e applicata, 52, 83-103.<br />

Wybraniec, S., Jankowski, J., Ernst, T., Pěčov , J., Praus, O., 1999. A new method for presentation of induction vector<br />

distribution in Europe, Acta geophys. polonica, XLVII, 323-334.<br />

´Zytko, K., 1997, Electrical conductivity anomaly of the Northern Carpathians and the deep structure of the orogen,<br />

Annales Societatis Geologorum Poloniae, 67, 25-43.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

243


Subsurface Conductivity Obtained from DC Railway<br />

Signal Propagation with a Dipole Model<br />

Anne Neska<br />

Institute of Geophysics PAS, Ul. Ks. Janusza 64, 01-452 Warszawa, Poland, email: anne@igf.edu.pl<br />

Abstract<br />

In the presented study, an attempt is made to model the propagation behavior of signals<br />

emitted by a Polish DC railway line with a number of grounded horizontal electric dipoles on<br />

the surface of a homogeneous half-space. The signals were measured on a magnetotelluric<br />

profile perpendicular to the railway line mainly in its near-field and transition zone, and they<br />

were separated from the natural electromagnetic variations by means of a reference site.<br />

Fitting the DC signals to the dipole model yields the conductivity of the homogeneous halfspace.<br />

Its value for the vertical and for the perpendicular horizontal magnetic component is<br />

confirmed by the usual 2D MT model obtained from the profile.<br />

Introduction<br />

The electrified part of the Polish railway network (fig. 1) is run by DC current. This makes<br />

magnetotelluric (MT) studies in this country difficult. On the other hand, the “disturbing” (in<br />

terms of MT) signals emitted by this system propagate according to a dipole model (Oettinger et<br />

al. 2001 and citations therein) which contains, like MT, the electrical conductivity of the<br />

subsurface as a parameter. So, after designing and performing a MT field experiment<br />

comprehending railway signals in an appropriate way, it should be possible to obtain information<br />

about the subsurface conductivity by both the dipole model and the MT approach. By this idea<br />

the present study has been motivated.<br />

The methodology of this study lies somewhere in-between two other domains encountered in<br />

geophysics. One is certainly the electromagnetic methods of the controlled-source near-field<br />

branch. The difference to our case is that the source is more under control there, but in return not<br />

free of experimental effort. The second related domain is techniques for modeling the<br />

propagation behavior of disturbances from DC railways and cables, applied especially to<br />

investigate their influence on magnetic measurements run at observatories (e.g. Pirjola et al.<br />

2007, Lowes 2009, Maule et al. 2009). However, the induction character of this propagation<br />

including phenomena like frequency-dependent damping, phase shift, and the electric<br />

conductivity as a propagation parameter is almost completely omitted there in contrast to the<br />

approach presented here.<br />

In the following, there will be considered the data processing including the separation of the<br />

natural electromagnetic variations from the railway signals and the estimation of the transfer<br />

functions between different stations. The dipole model will be described and the data be inverted<br />

according to it. The conductivity result will be compared to the output of a usual 2D MT model.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

244


Fig. 1 The electrified part of the Polish railway network (red lines) with a profile of MT sites for<br />

investigation of one line (zoomed-in part). Green squares denote positions of reference stations.<br />

Measurement and data processing<br />

As shown in fig. 1, a profile of six MT sites has been installed perpendicular to an isolated railway<br />

line. The distances of sites to the line were 0.8, 7, 16, 25, 50, and 75 km. Two remote reference<br />

sites were set up at relatively noise-free places, and the whole array was running synchronously.<br />

The obtained data were processed with codes by Neska 2006, but not only for MT purposes.<br />

By means of its correlation to the reference sites, the natural part of the electromagnetic<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

245


signal could be removed from the measured data using a method by Larsen et al. 1996 (fig. 2). In<br />

this way a dataset for analysis of the railway signals was established (cf. fig. 3). It has to be emphasized<br />

that it can be problematic to work with such “deduced” data, since distortions introduced<br />

to it during the separation, e.g. due to uncorrelated noise in the reference site, can bias the<br />

subsequent results. It proved useful to separate the data of the station closest to the railway which<br />

plays a special role in the following with one reference site and the rest with the second one.<br />

Fig. 2 Separation of time series into a “natural” and a “railway” part (site s02, see fig. 1).<br />

Fig. 3 The propagation/decay of railway signals (component Bx) over the profile. Note that a rest of<br />

natural variations remained in the time series in spite of separation.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

246


All following considerations refer to “railway data” separated from natural signals in the way<br />

indicated above.<br />

The formulas for the dipole response are given in frequency domain, so our calculation has to<br />

take place there. Furthermore, in these formulas there occur factors like current I and dipole<br />

length L (see table 1), i.e. technical parameters of railway traction which we cannot measure<br />

immediately. So it appeared reasonable to remove these factors by normalizing each field<br />

component to that of the railway-nearest site which represents the strongest, most pronounced, or<br />

least attenuated railway signal, respectively.<br />

To perform this normalization in a convenient way, the algorithm for calculation of the interstation<br />

transfer function or Horizontal Magnetic Tensor (HMT) was used, which played an<br />

essential role during the separation already. The site nearest to the railway takes the function of<br />

the reference, i.e. its data are input channels, and the data of the remaining sites are the output<br />

channels. So Bx and By data of “local” and “reference” sites lead to normalized values for both<br />

horizontal magnetic components, Ex and Ey to those for the electric components (even if the<br />

analogy to the HMT is overstretched here since it refers to magnetic data only, the formalism is<br />

the same), and Bz and some other, uncorrelated channel (e.g. of one of the real, off-profile<br />

references, just to fit the requirement of the HMT formalism that there must be two input<br />

channels) to the normalized value for the vertical magnetic component. This description remains<br />

a bit vague because it cannot be recommended for imitation for the following reason:<br />

For this study, it has been noticed too late that this approach is not quite correct. The problem is<br />

that the HMT formalism bases on bivariate statistics (i.e. there are two independent variables or<br />

input channels, respectively), whereas the railway provides only one independent source<br />

polarization (pers. comm. K. Nowożyński). A similar problem occurs in Controlled-Source MT,<br />

where only one scalar transfer function instead of the usual 2x2 impedance tensor can be<br />

obtained if only one transmitter is used (pers. comm. M. Becken). So it has to be expected that<br />

our normalized values are influenced in an unfavorable way. This impact is expected to be very<br />

small in the case of Bz, since the warranted uncorrelatedness of the second input channel should<br />

make the difference between the univariate and the bivariate result vanish. This will be<br />

confirmed in the following section (cf. fig. 4: better result for Bz). However, the direction of the<br />

railway line under investigation is almost East-West (fig. 1), so its signal has a polarization<br />

almost coinciding with one of the directions we measure and calculate in with the MT or HMT<br />

approach. So there is hope that the error introduced due to this wrong statistic approach is<br />

somehow bearable if the diagonal elements of the “HMT” are used.<br />

Finally, the normalized values are displayed over period and distance in 3D-plots (red lines in<br />

fig. 4). The decay of amplitudes with distance as well as their period-dependency (stronger<br />

damping at low periods) becomes clearly visible for Bx and Bz components. Bx has partly<br />

unsystematic and, especially at long periods, rather high values that make the impression to be<br />

distortions, maybe due to incomplete separation or the wrong statistic approach mentioned.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

247


Hence the possibility to include data errors to the business was provided, since they have an<br />

important meaning as weights during the inversion.<br />

The dipole model and its application<br />

The railway line was simulated by a chain of horizontal grounded electric dipoles following its<br />

(not constant) bearing within a distance of some dozens of km around the profile. The distance<br />

between single dipole centers was 250 m. The solutions for the electromagnetic field components<br />

for the quasi-static approximation are taken from Zonge & Hughes 1987 and hold for one dipole<br />

on the surface of a homogeneous half-space (see table 1). These values are transformed from<br />

cylindrical to Cartesian coordinates to match the measuring system of MT and summarized over<br />

all dipoles subsequently. Then the values for each component and station were divided by the<br />

value of the corresponding component of the station closest to the railway. Since we look only<br />

for one model parameter (i.e. the conductivity σ), the inversion consists just of testing the whole<br />

model space and selecting the value with minimum RMS. The relevant software related to the<br />

dipole model was created by K. Nowożyński.<br />

The components behaving best in our study were Bx and Bz. The inversion of Bx alone gave a<br />

resistivity of 5 Ωm, that of Bz alone 14 Ωm, and the joint inversion 10 Ωm. The model response<br />

for the latter is shown in fig. 4 as blue lines.<br />

Table 1<br />

Solution for horizontal grounded electric dipole on HH surface<br />

according to Zonge & Hughes, 1987<br />

Where I/K_0/1 - modified Bessel functions and<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

248


Fig. 4 Railway signal data (red) and dipole model response for a 10 Ωm homogeneous half-space (blue)<br />

over period and distance. Upper part for Bx, lower part for Bz component amplitude. The values on the<br />

vertical axis are normalized to that of site s01 nearest to the railway line (cf. fig. 1).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

249


Fig. 5 2D MT model (TE and TM mode) of the profile on fig. 1 obtained with the REBOCC code<br />

(Siripunvaraporn & Egbert 2000). The shallow cells beneath the northernmost sites have resistivity values<br />

between 9 and 80 Ωm.<br />

Discussion and conclusions<br />

The MT model yields resistivities of 9-80 Ωm beneath the northernmost sites close to the railway<br />

(fig. 5). This is consistent with the dipole model result, so there is evidence that the dipole<br />

approach is not unreasonable.<br />

Nevertheless, there are unexplained features in this model and one has to be critical with it. Not<br />

shown here is the behavior of phases and of electric components, where a similarity between data<br />

and model response is much harder to find for reasons still unclear.<br />

However, taking into account first, the similarity of data and model response (fig. 4), and second,<br />

the compatibility of MT and dipole modeling results in spite of some quite principal problems in<br />

this implementation (i.e. input data biased during separation, errors due to application of<br />

bivariate instead of univariate statistics, limitation to the oversimplified case of a homogeneous<br />

half-space), there can be stated that the dipole model is not only a valid approach to model the<br />

propagation of DC railway signals, but even a relatively robust one.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

250


Acknowledgements<br />

This work would not exist without the help of Paweł Czubak, Tomasz Ernst, Jerzy Jankowski, Waldemar<br />

Jóźwiak, Krzysztof Kucharski, Janusz Marianiuk, Mariusz Neska, Krzysztof Nowożyński, Jan Reda,<br />

Michał Sawicki, and Józef Skowroński.<br />

The project is funded by the Ministry of Science and Higher Education of the Republic of Poland (grant<br />

number N N307 2496 33).<br />

Literature<br />

Larsen, J. C., R. L. Mackie, A. Manzella, A. Fiordelisi, and S. Rieven. Robust smooth magnetotelluric<br />

transfer functions. Geophysical Journal International, 124:801819, 1996.<br />

Lowes, F. J., DC railways and the magnetic fields they produce - the geomagnetic context, Earth Planets<br />

Space, 61, i–xv, 2009<br />

Maule, C.F., P. Thejll, A. Neska, J. Matzka, L.W. Pedersen, and A. Nilsson, Analyzing and correcting for<br />

contaminating magnetic fields at the Brorfelde geomagnetic observatory due to high voltage DC power<br />

lines, Earth Planets Space, 61, 1233–1241, 2009<br />

Neska, A., Remote Reference versus Signal-Noise Separation: A least-square based comparison between<br />

magnetotelluric processing techniques, PhD thesis, Freie Universität Berlin, Fachbereich Geowissenschaften,<br />

available at http://www.diss.fu-berlin.de/2006/349, 2006.<br />

Oettinger, G., V. Haak, and J.C. Larsen. Noise reduction in magnetotelluric time-series with a new signalnoise<br />

separation method and its applications to a field experiment in the Saxonian Granit Massif.<br />

Geophysical Journal International, 146:659669, 2001.<br />

Pirjola, R., L. Newitt, D. Boteler, L.a Trichtchenko, P. Fernberg, L. McKee, D. Danskin, and G. J. van<br />

Beek, Modelling the disturbance caused by a dc-electrified railway to geomagnetic measurements, Earth<br />

Planets Space, 59, 943–949, 2007.<br />

Siripunvaraporn, W., and G. Egbert, An efficient data-subspace inversion for two-dimensional magnetotelluric<br />

data, Geophysics, 65, 791-803, 2000.<br />

Zonge, K.L. and L.J. Hughes, Controlled source audio-frequency magnetotellurics, In: M.N. Nabighian<br />

(editor), Electromagnetic Methods in Applied Geophysics, Volume 2, Application, Parts A and B, pp.<br />

713809, SEG, Tulsa, 1987.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

251


Magnetotelluric investigation of the<br />

Sorgenfrei-Tornquist Zone and the NE German Basin<br />

Anja Schäfer 1∗ , Heinrich Brasse 1 , Norbert Hoffmann 2<br />

and EMTESZ working group<br />

1 Freie Universität Berlin, Fachrichtung Geophysik, Malteserstr. 74-100, 12249 Berlin, Germany<br />

2 Wilhelm-Külz-Straße 70, 14532 Stahnsdorf, Germany<br />

Abstract<br />

The Sorgenfrei-Tornquist Zone (STZ) is the northwestern branch of the Trans-European Suture<br />

Zone (TESZ). The TESZ runs along more than 2 000 km through the North Sea, South Scandinavia,<br />

the Baltic Sea, and Poland into the Black Sea. It divides old Precambrian lithosphere in<br />

the northeast from younger, Caledonian and Variscan one in the southwest. Data, dimensionality<br />

analysis and 2D models of a long-period magnetotelluric (LMT) profile crossing this prominent<br />

tectonic border in South Scandinavia are presented and a possible interpretation of conductivity<br />

anomalies below Rügen Island and south of Strelasund Basin is given. Furthermore these models<br />

map the saline aquifer of the Northeast German Basin. The ”North German Conductivity<br />

Anomaly” is perhaps mainly due to the effect of the basin edges.<br />

Introduction<br />

The EMTESZ project (Electromagnetic Study of the Trans-European Suture Zone) was a multinational<br />

research project to study the electrical conductivity of the Trans-European Suture Zone<br />

(TESZ), which is one of the largest tectonic boundaries in Europe. It separates the East European<br />

Platform from the Paleozoic mobile belt of central and western Europe and is traced from the<br />

Black Sea through Poland and Southern Scandinavia into the North Sea (Gee (1996); Pharaoh<br />

(1999)). In the northeast the margin is marked by the lineaments of the Sorgenfrei-Tornquist<br />

Zone (STZ) and the Teisseyere-Tornquist Zone (TTZ, which runs in a SE-NW direction through<br />

Poland and the Baltic Sea, fig. 1). Several seismic refraction and reflection experiments have been<br />

carried out (BABEL and BASIN 9601 in Northeast Germany, POLONAISE and the LT surveys<br />

in Poland). Major results include a sedimentary thickness in the Northeast German Basin and<br />

below the TTZ of 4 to more than 11 km. Also sharp lateral boundaries, a Moho from 32-35 km<br />

in the SW to 40-45 km in the NE and a reflector below the STZ/TTZ at depth of 50-55 km are<br />

known. Most of the earlier EMTESZ measurements were carried out in 2003 to 2005 (e.g., Brasse<br />

et al. (2006), Ernst et al. (2008)). Profiles MVB and MVS were conducted in 2006 and 2009.<br />

The long period magnetotelluric profile shown here is one of these EMTESZ profiles. The direction<br />

of this profile was chosen because of earlier measurements by the Bundesanstalt für Geowissenschaften<br />

und Rohstoffe (BGR) in the 1990ies. Due to tipper data along this profile and of site<br />

MAT deployed in 2005 on Rügen Island (MT working group Free University of Berlin) a direction<br />

∗ e-mail: schaefer@geophysik.fu-berlin.de<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

252


OST<br />

DAB<br />

BAK<br />

MIR<br />

SWI<br />

NIE<br />

KOC<br />

SCZ WIA<br />

DRE<br />

PIE<br />

GAT<br />

WIE<br />

ZAR<br />

ZAL<br />

GRA<br />

KUS<br />

KNY<br />

PRZ<br />

KUZ<br />

SIE<br />

SLA<br />

B010<br />

ZOL ZOL<br />

B020<br />

POM<br />

GLE<br />

B030<br />

A010 END<br />

REC<br />

B040<br />

A020<br />

RAD<br />

B050<br />

A030 CRI<br />

LUB<br />

KUK<br />

B060<br />

TEE<br />

LYS<br />

B070<br />

A040 LOE<br />

MOC<br />

LES<br />

LAW<br />

B080<br />

A050 OLE<br />

CHO<br />

WAR<br />

B090<br />

A060<br />

LAK<br />

B100<br />

PAS<br />

MOS<br />

A070<br />

STO<br />

SAR<br />

B110<br />

A080<br />

B120<br />

NKU BLE<br />

GOR<br />

B130<br />

VEV<br />

A090<br />

FRI<br />

B140<br />

A100<br />

HOB FAL<br />

B150<br />

A110<br />

MAR MAD<br />

B160<br />

B170<br />

AGO KUN<br />

B180<br />

A120<br />

B190<br />

DAM<br />

BRI<br />

B200<br />

A130<br />

GLI<br />

B210<br />

A140<br />

B220<br />

BUS<br />

A150<br />

TOR<br />

B230 A160<br />

B240<br />

HBO<br />

A170<br />

B250 A180<br />

B260<br />

B270<br />

B280<br />

56°N<br />

55°N<br />

54°N<br />

53°N<br />

52°N<br />

VDF<br />

TEF<br />

Rostock<br />

AAF<br />

STZ<br />

BGR-B<br />

KIS<br />

ROH<br />

TET<br />

WIL<br />

BRE<br />

DOL<br />

Magdeburg<br />

Trelleborg<br />

CDF<br />

IBI<br />

DAR<br />

LEL<br />

WEN<br />

ROT<br />

JAB<br />

MVS<br />

NGK<br />

KRU<br />

PAP<br />

k31<br />

MAT<br />

PAR<br />

TAN<br />

BAR<br />

Germany<br />

BGR-A<br />

LOV<br />

TOM<br />

BOO<br />

k25<br />

Berlin<br />

KOP<br />

NYT<br />

HOR<br />

AND<br />

Rügen<br />

KRI<br />

BER<br />

REG<br />

BLU<br />

MVB<br />

BS1<br />

DZI<br />

FEL<br />

BS3<br />

Odra R.<br />

Bornholm<br />

Baltic Sea<br />

Poland<br />

0 50<br />

11°E 12°E 13°E 14°E 15°E 16°E 17°E 18°E 19°E<br />

BOR<br />

Szczecin<br />

Kostrzyn<br />

Frankfurt<br />

CYB<br />

NIN<br />

Koszalin<br />

KAR<br />

WLO<br />

ROT<br />

TAC<br />

BLA<br />

MAS<br />

BS2<br />

ZAS<br />

JAN<br />

Poznan<br />

LT-7<br />

VDF<br />

km<br />

study area<br />

Figure 1: Measured profiles of the EMTESZ project and profile B from the Bundesanstalt für Geowissenschaften<br />

und Rohstoffe (BGR-B). Location of MT sites in Poland, Germany, Sweden and Bornholm<br />

(Denmark); STZ-Sorgenfrei-Tornquist-Zone; TTZ-Tornquist-Teiseyre-Zone; TEF-Trans European Fault;<br />

CDF-Caledonian Deformation Front; VDF-Variscan Deformation Front; AAF-Amorica Avalonia Fault.<br />

perpendicular to the assumed strike-direction of the Sorgenfrei-Tornquist Zone (see fig. 1) was<br />

chosen. It runs over about 370 km with a site spacing from 6 to 12 km.<br />

Data examples<br />

Stable transfer functions result from remote reference analysis according to Egbert & Booker<br />

(1986). The remote data is from the observatories of Belsk and Niemegk and simultaneous measuring<br />

sites. In figure 2 three representative data examples for sites of profile MVS are shown.<br />

Site ROT, which is located in the southern part of the profile, shows the characteristic curves for<br />

the North German Basin. Apparent resistivities at short periods are very low until 100 s (1 to<br />

3 Ωm), pointing to a very good conductor near the surface. With increasing periods app. resistivity<br />

is growing, accompanied by a splitting of impedance components Zxy and Zyx, reflecting<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

253<br />

STZ<br />

P2<br />

Gdansk<br />

P08<br />

TTZ<br />

TTZ<br />

Wisła R.<br />

HLP<br />

E


ρ a [Ωm]<br />

10 4<br />

10 3<br />

10 2<br />

10 1 10 1<br />

rot<br />

Zxy<br />

Zyx<br />

10 0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

10<br />

90<br />

0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

10<br />

90<br />

0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

90<br />

φ ο<br />

Re<br />

Im<br />

45<br />

10 4<br />

10 3<br />

10 2<br />

10 1 10 1<br />

45<br />

boo<br />

0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

↑Ν<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

0<br />

10<br />

T [s]<br />

1 10 2 10 2 10 3 10 4 10 5<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

↑Ν<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

0<br />

10<br />

T [s]<br />

1 10 2 10 2 10 3 10 4 10 5<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

↑Ν<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

10 1 10 2 10 2 10 3 10 4 10 5<br />

T [s]<br />

Figure 2: Examples of transfer functions for sites of the profile (ρa-apperent resistivity, φ-Phase and Re-<br />

Real part, Im- imaginary part of induction arrows as function of period). Site ROT as an example for the<br />

southern part of the profile shows low resitivity in short periods, which shows the thick sedimentary covers<br />

of North German Basin (ρa at low periods around 1-3 Ωm), for longer periods it shows higher resitivity.<br />

BOO is the site at the most southern part of the swedish part and shows the margin of the sedimentary<br />

basin. NYT is an example for the northern part of the profile in Southscandinavia. There the profile<br />

reaches the crystalline basement (ρa has higher values even for low periods).<br />

multidimensional structures at greater depths. Induction arrows are very small. For sites in the<br />

southern part of the profile a change of direction for induction arrows is observed. Site BOO is<br />

located at the edge of the East European Craton. Here real parts of induction arrows are very<br />

large (up to an absolute value greater than 0.8) and are perpendicular to the strike direction of the<br />

Sorgenfrei-Tornquist Zone. They point to a very good conductor in southeastern direction. Apparent<br />

resistivities for this site are even for short periods much higher than in the North German<br />

part of the profile and increase rapidly for large periods, which points to the resistive basement of<br />

the East European Craton. NYT is an example for the northernmost part of the profile and shows<br />

high apparent resistivity even for the lowest periods (about 1 000 Ωm), reflecting the crystalline<br />

basement of Baltica.<br />

Magnetic transfer functions<br />

Induction arrows at long periods increase from south to the north, hinting at well-conductive<br />

structures in the south, i.e., below the Baltic Sea. The apparent resistivities ρa in the north of<br />

the profile are relatively high and get lower at the southern sites, which agrees with the geological<br />

structure. The high resistivities in the north are caused by the crystalline basement and the low<br />

resistivities in the south by the sedimentary cover.<br />

Figure 3 shows a map of the real part of induction arrows at a period 1 820 s for the EMTESZ<br />

profiles LT-7, MVB and the discussed profile MVS, according to Wiese-Convention (Wiese, 1962).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

254<br />

Zxy<br />

Zyx<br />

10 4<br />

10 3<br />

10 2<br />

10 1 10 1<br />

45<br />

nyt<br />

Zxy<br />

Zyx


For profile MVS arrows are mainly perpendicular to the NW-SE striking direction. With values<br />

greater than 0.8, the unusually large induction arrows in the northern part of the profile indicate<br />

a good conductor below Rügen Island. Real induction arrows at the southernmost station of<br />

Sweden, the sites in the Baltic Sea and at the northernmost point of Rügen Island are largest.<br />

The ”flip-around” of induction arrows in Northern Germany and Poland was already recognized<br />

in the initial days of electromagnetic deep sounding. The underlying cause – a large conductivity<br />

anomaly at depth – was termed the North German-Polish Conductivity Anomaly (e.g.,<br />

Schmucker (1959); Untiedt (1970); Jankowski (1967)). Until today, however, the depth extent<br />

of this anomaly is controversially discussed and models range from an upper mantle high<br />

conductivity zone to the simple effect of the basin edges. The magnitude of the inductive effect<br />

at the basin margins and their rapid decrease seem to favor the second explanation, which will<br />

also become evident from 2-D inversion (see later).<br />

56°N<br />

55°N<br />

54°N<br />

53°N<br />

52°N<br />

12°E 13°E 14°E 15°E 16°E 17°E 18°E 19°E<br />

km<br />

0 50<br />

1820 sec<br />

Figure 3: Induction arrows for profiles MVS, MVB and LT-7 (real part) at a period of 1 820 s. In South<br />

Sweden and north of Rügen Island induction arrows are very large. They obviously result from the edge<br />

of the Baltic Shield (STZ). The change of direction of the real parts in the center of the NE German and<br />

Polish Basins shows the effect of the so-called ”North German-Polish Conductivity Anomaly”. It can be<br />

seen in all profiles. Profile MVS shows this effect (sign reversal of real part) approximately at the location<br />

of the Stralsund-Anklam-Fault (SAF). In the northern part of profile MVS the induction arrows are very<br />

large. The highest absolute value with more then 0.8 is reached in the southern part of Sweden and 0.7 in<br />

the northernmost part of Rügen Island.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

255


Dimensionality analysis<br />

Strike angles were calculated for single and multi sites after Smith (1995), which allows to sort<br />

out bad data with a weighting matrix (see figure 4). The strike angle for profile MVS has a strong<br />

variation especially in the northern part; the average for the profile is around N67 ◦ W.<br />

WEST<br />

NORTH<br />

SOUTH<br />

EAST<br />

Figure 4: Average strike angle after Smith (1995) for profile MVS: N67 ◦ W.<br />

Figure 5 shows the phase tensor (Caldwell et al., 2004) for each site of the profile for two<br />

different periods. The orientation of the ellipses point to the strike direction of the conductive<br />

structures. The northwest-pointing direction of the main axis of the ellipse corresponds with the<br />

calculated strike direction.<br />

a) b)<br />

56˚<br />

55˚<br />

54˚<br />

53˚<br />

PAP<br />

IBI<br />

DAR<br />

LEL<br />

WEN<br />

ROT<br />

JAB<br />

KIS<br />

ROH<br />

TET<br />

WIL<br />

BRE<br />

DOL<br />

MAT<br />

PAR<br />

TAN<br />

BAR<br />

KRU<br />

KOP<br />

NYT<br />

HOR<br />

BER<br />

REG<br />

LOV<br />

TOM<br />

BOO<br />

KRI<br />

0 50<br />

12˚ 13˚ 14˚ 15˚ 12˚ 13˚ 14˚ 15˚<br />

-16 -12 -8 -4 0 4 8 12 16<br />

β [ o ]<br />

PAP<br />

IBI<br />

DAR<br />

LEL<br />

WEN<br />

ROT<br />

JAB<br />

KIS<br />

ROH<br />

TET<br />

WIL<br />

BRE<br />

DOL<br />

MAT<br />

PAR<br />

TAN<br />

BAR<br />

KRU<br />

KOP<br />

NYT<br />

HOR<br />

BER<br />

REG<br />

LOV<br />

TOM<br />

BOO<br />

KRI<br />

0 50<br />

Figure 5: Phase tensor plot for periods of 992 s and 1 820 s calculated after Caldwell et al. (2004).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

256


Skew angle β is small throughout the study area with the exception of the northernmost sites in<br />

Sweden. Summarizing, it suffices to carry out a 2-D modeling of the data.<br />

2-D modeling<br />

All presented 2-D inversion models were calculated with the CG-FD inversion algorithm of Rodi<br />

&Mackie(2001). First the data were rotated by -67 ◦ . A homogeneous halfspace of 100 Ωm<br />

and a fine grid for the start model was chosen. The smoothing factor was set to τ=15, which is<br />

the best trade-off between model roughness and data misfit (Hansen & O’Leary, 1993). For<br />

the inversion of the 2D-models in fig. 6 all components (TE-, TM-mode and tipper) were used.<br />

In these models all resolved structures can be seen; these models have the best fitting model<br />

response to the data (see examples in fig. 7). For a better overview of the modeled features, the<br />

profile intersecting geological structures in the respective locations are marked. Furthermore, the<br />

location of the basement is shown. Figure 6a shows the result of inversion of TE-, TM-mode and<br />

tipper data after 200 iterations. The resulting RMS is 2.27. The model shown in figure 6b also<br />

results from inversion of all components (TE-, TM-mode and tipper), furthermore a horizontal<br />

weighting function was used (β=1). The resulting RMS of this inversion after 200 iterations is<br />

with 2.67 worse then the RMS of the model without weighting function, but the model is more<br />

consistent with geological assumptions (Hoffmann & Franke, 2008).<br />

a)<br />

b)<br />

z [km]<br />

z [km]<br />

0<br />

20<br />

40<br />

60<br />

80<br />

0<br />

20<br />

40<br />

60<br />

80<br />

S<br />

VDF<br />

dol<br />

bre<br />

wil<br />

tet<br />

roh<br />

kis<br />

jab<br />

rot<br />

wen<br />

lel<br />

dar<br />

ibi<br />

pap<br />

kru<br />

OA<br />

OA<br />

A<br />

A<br />

East-Avalonia<br />

Paleozoic Platform<br />

SAF CDF<br />

Strelasund<br />

-Basin<br />

D<br />

D<br />

bar<br />

Wiek<br />

trough<br />

tan<br />

par<br />

mat<br />

k31<br />

C<br />

Baltica<br />

Baltica<br />

k25<br />

boo<br />

STZ<br />

tom<br />

lov<br />

reg<br />

and<br />

ber<br />

hor<br />

nyt<br />

kop<br />

Baltica<br />

0 50 100 150 200 250 300 350<br />

distance [km]<br />

C<br />

Baltica<br />

Baltica<br />

East European Craton<br />

Figure 6: 2-D models for profile MVS inverted by applying Rodi and Mackie’s algorithm starting from<br />

a homogeneous halfspace. a) For this model all components were inverted (TE-, TM-mode and tipper<br />

data). The resulting RMS after 200 iterations was 2.27. b) For inversion of this model all components<br />

were used (TE-, TM-mode and Tipper). A horizontal weighting function was used. Resulting RMS after<br />

200 iterations was 2.67. For both models the smoothing factor was set to τ=15. Letters sign resolved<br />

conductivity structures: A-saline aquifer in Northeast Germany, Baltica-basement of the Baltic Continent,<br />

OA-basement, associated with the basement of the East-Avalonian plate. Also the assumed location of<br />

plates and fault zones is marked.<br />

In both models one can see the underlying plate of Baltica in the north as a poor conductor, the<br />

sediments in Northeast Germany as a very good conductor (structure A) and two conductivity<br />

anomalies below Rügen Island (structure C) and south of Stralsund (structure D) in depths from<br />

8 to 30 km. Figure 7 shows examples of model response of the inversion shown in figure 6a.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

257<br />

kri<br />

N<br />

4<br />

3<br />

log ρ [Ωm]<br />

2<br />

1<br />

0


Rho App. (ohm.m)<br />

Phase (deg)<br />

Tip Mag<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

90<br />

60<br />

30<br />

0<br />

1,0<br />

,5<br />

,0<br />

dol<br />

10 1<br />

TE measured<br />

TE model response<br />

rms= 1.2261<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

90<br />

60<br />

30<br />

0<br />

1,0<br />

,5<br />

lel rms= 1.2746<br />

10 2<br />

10 3<br />

10 4<br />

10 1<br />

10 2<br />

10 3<br />

,0<br />

10 4 10 1<br />

10 2<br />

10 3<br />

10 4<br />

,0<br />

4<br />

10<br />

Period(sec)<br />

1<br />

10 2<br />

,0<br />

Period(sec) Period(sec)<br />

Period(sec)<br />

TM measured<br />

TM model response<br />

Hz measured<br />

Hz model response<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

90<br />

60<br />

30<br />

0<br />

1,0<br />

,5<br />

mat rms= 1.6917<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

90<br />

60<br />

30<br />

0<br />

1,0<br />

,5<br />

boo rms= 2.1131<br />

Figure 7: Examples of model responses for the 2-D inversion result as shown in fig. 6a.<br />

Geological Interpretation<br />

Structure A seen in the models of fig. 6 is characteristic for the North German Basin; it represents<br />

the saline aquifer and is already known from previous measurements. This structure is resolved in<br />

all inversions of this profile and could be seen in all components (TE-, TM-mode and tipper) and<br />

with its vertical elongation and extension to depths of 3 to 4 km it is consistent with simulations<br />

of Magri et al. (2007) and models of the BGR-profiles in Hoffmann et al. (1997). This<br />

structure is resolved as a surface conductor in all other sections of the EMTESZ project (e.g.,<br />

Brasse et al. (2006), Ernstetal.(2008)) and in the models of Houpt (2008) as well. It is<br />

caused by huge deposits of Zechstein salt, which is dissolved by deeper ground water. Through<br />

thermal and hydraulic transport mechanisms it reaches in some cases up to the surface (Magri<br />

et al., 2005). In some parts of the North German Basin salinities up to 350 g/l are observed<br />

(Hoth et al., 1997). The conductivity of such layers depends on the salinity of the fluid, the<br />

size and connection of the pore spaces. High salt contents and pore sizes of sediment layers in<br />

North Germany may cause conductivities well above 1 S/m.<br />

Structure ”Baltica” is also resolved in all inversions of profile MVS - a very poor conductor in the<br />

north of the profile. It represents the crystalline Precambrian basement of Baltica. This basement<br />

of Baltica is with an age of 1.9 billion years one of the oldest rocks in Europe. Its high resistivity<br />

is due to the strong compression of rocks at depth with almost no interconnected fluid inclusions<br />

left.<br />

Extension and value of the conductivity of the good conductor below Rügen Island (structure<br />

C) can also be regarded as assured due to its shallow depth and the sensitivity tests carried out<br />

(Schäfer, 2010). Its boundary to the north is indicated by the size of the induction arrows at<br />

MAT, the northernmost station on Rügen Island. Structure C is located at about 8 to 20 km<br />

depth. This good conductor is depicted in all inversions as a structure separated from the surface<br />

conductor (structure A).<br />

Also structure D, with a depth of 10 to 30 km is resolved in all models of the joint inversion of TE-,<br />

TM-mode and tipper, and also proofed by sensitivity tests as a separate structure from the surface<br />

conductor. Conductors in these depths were explained with the occurrence of highly-carbonated<br />

paleozoic black shales, the so-called Scandinavian alum shales (e.g., Hoffmann et al. (1998);<br />

Hengesbach (2006)).<br />

These alum shales crop out at the Andrarum quarry in the southern Swedish province of Scania,<br />

near MT sites AND and REG, where they have been mined since centuries. In order to test<br />

the hypothesis of high conductivity, several DC-geoelectric array measurements were conducted<br />

in this area during a student field trip of the Free University of Berlin (Field Report Scania,<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

258<br />

10 3<br />

10 4


2006). Fig. 8 displays an example. It clearly shows the black shale (blue colors) as an intermediate<br />

conductor only, with resistivities in the range of 100 to 150 Ωm. Nevertheless, this does not exclude<br />

high conductivities of deeply buried shales; at or near the surface, the relatively high resistivity<br />

may be due to strong weathering, destroying the conductive paths, which are otherwise continuous<br />

in deeper layers.<br />

Figure 8: Geoelectric section near the Andrarum black shale quarry (South Sweden). Note the reverse<br />

color scale.<br />

From laboratory studies and drilling black or alum shale at greater depths and without weathering<br />

it is obtained with conductivities of less than 1 Ωm (Duba et al., 1988). Black shale has been<br />

investigated by drilling in the well of G14 (north of Rügen Island) and Rügen 5, where it reaches<br />

thicknesses of 50 to 70 m. In the well Pröttlin I further south, black shales were found, too, but<br />

could not be fully intersected. In addition, studies demonstrated that the necessary pore size for<br />

an electrolytic conduction mechanism not exists in deeper sediment layers (Jödicke, 1991). This<br />

would exclude the interpretation of structure D as a deep sedimentary trough.<br />

Thus, the conductive ”layer” in the model of figure 6b can be seen as a reference to alum shale occurrence.<br />

However, this layer can be even more conductive because of hydrothermal waters, which<br />

are often rising in geological fault zones. Under high pressure and special temperature conditions<br />

in highly-carbonated black shales also graphite-structures may form, according to estimates from<br />

Duba et al. (1988) and Raab et al. (1998). With a layer of mostly continuous black shale<br />

or graphite surface, a relatively low-friction thrust faulting with little deformation of the plate<br />

boundaries in the collision of Eastern Avalonia and Baltica at the time of the Ordovician could<br />

be explained.<br />

Furthermore, Raab et al. (1998) could demonstrate in laboratory experiments that pyrite inclusions<br />

(which often occur in black shale layers) may convert to much more conductive pyrrhotite<br />

at temperature and pressure conditions as encountered in the middle crust. Pyrrhotite usually<br />

occurs in dendritic form and increases conductivity even through small volumes.<br />

Model experiments and sensitivity tests show that a 70 m thick alum shale layer at such depths is<br />

not sufficient, even if modeled with bulk resistivities as low 0.1 Ωm. Another possible explanation<br />

is the occurrence of intrusive bodies, which rose in the time of Rotliegend (Upper Carboniferous<br />

to Middle Permian). Their contact aureoles usually consist of highly conductive material (such as<br />

ilmenite or even graphite), and can extend over larger areas. So the high conductive structures C<br />

and D might be explained by a combination of highly conductive layers of black shales and such<br />

dykes. Motivation for this hypothesis emerged from the well Greifswald I, where intrusive granite<br />

bodies from the Rotliegend with apophyses could be found. This granite body is also known as<br />

Südrügen Pluton (Hoth et al., 1993) and is possibly extending to the south of Rügen Island.<br />

In fact, the combination of these two interpretations, seems to be the best explanation of the high<br />

conductivities of structures C and D at depth. Thus, it is very likely a combination of graphitized<br />

black shale and intrusive bodies (e.g., Südrügen Pluton) with the accompanying conductive material<br />

and the associated apophyses. This interpretation would explain the very good conductivity<br />

in such depths and the resolution of the structures as an almost vertical body in the mid crust.<br />

Unfortunately no better resolution of these structures is possible due to the strong shielding effect<br />

of the saline aquifer.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

259


The good conductor of model 6b ends in the south of the Stralsund Anklam Fault (SAF). The trend<br />

of the SAF, which is also indicated by magnetic transfer functions and their graphic presentation<br />

in fig. 3 fits very well to the results of Houpt (2008) and Ernst et al. (2008), which were<br />

obtained further east. According to these models and the induction arrows shown in fig. 3, the<br />

SAF appears to be a fundamental crustal boundary of Northeastern Europe which can be traced<br />

to the southeast of Szczecin in Poland.<br />

Figure 9 shows a comparison with an interpretation of the BASIN 9601 project and 2 prolonging<br />

profiles in the Baltic Sea from McCann & Krawczyk (2001). There we observe a good correlation<br />

with underlying Baltica and the Zechstein base (strong reflector), and the well-conducting<br />

sediments in the upper layers of the Northeast German Basin.<br />

z [km]<br />

0<br />

20<br />

40<br />

60<br />

80<br />

S VDF<br />

SAF CDF<br />

dol<br />

bre<br />

wil<br />

tet<br />

roh<br />

kis<br />

jab<br />

A<br />

OA<br />

Moho<br />

Baltica<br />

Baltica<br />

0 50 100 150 200 250 300 350<br />

distance [km]<br />

STZ<br />

Grimmen High Baltic Sea<br />

G14<br />

ber<br />

hor<br />

nyt<br />

kop<br />

Figure 9: 2-D models for profile MVS with an overlayed interpretation of the seismic reflection/refraction<br />

profile BASIN 9601 and 2 prolonging profiles in the Baltic Sea from McCann & Krawczyk (2001).<br />

References<br />

Brasse, H., Červ, V., Ernst, T., Jó´zwiak, W., Pedersen, L., Varentsov, I. & Pomerania<br />

working group, E. (2006). Probing the electrical conductivity structure of the Trans-<br />

European Suture Zone. EOS Trans. AGU, 87, 29, doi:10.1029/2006EO290002.<br />

Caldwell, T., Bibby, H.M. & Brown, C. (2004). The magnetotelluric phase tensor. Geophys.<br />

J. Int., 158, 457–469.<br />

Duba, A., Huenges, E., Nover, G., Will, G. & Jödicke, H. (1988). Impedance of black<br />

shale from Münsterland 1 borehole: An anomalously good conductor? Geophys. J., 94, 413–419.<br />

Egbert,G.D.&Booker,J.R.(1986). Robust estimation of geomagnetic transfer functions.<br />

Geophysical Journal of the Royal Astronomical Society, 87, 173–194.<br />

Ernst, T., Brasse, H., Červ, V., Hoffmann, N., Jankowski, J., Jó´zwiak, W., Kreutzmann,<br />

A., Neska, A., Palshin, N., Pedersen, L., Smirnov, M., Sokolova, E. &<br />

Varentsov, I. (2008). Electromagnetic images of the deep structure of the Trans-European<br />

Suture Zone beneath Polish Pomerania. Geophys. Res. Lett., 35, doi:10.1029/2007GL034610.<br />

Field Report Scania (2006). Investigation of near-surface structures of the Sorgenfrei-Tornquist-<br />

Zone (STZ) at the margin of Baltica in Scania (Sweden). Geophysical Excursion to Scania<br />

(Sweden)-Field Report, FU Berlin, Fachrichtung Geophysik, Berlin.<br />

Gee, D. G. & Zeyen, H. (1996). Europrobe 1996-lithosphere dynamics. origin and evolution of<br />

continents. EUROPROBE Secretariat, Uppsala University, EUROPROBE 1996 (eds.), pp.138.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

260<br />

kri<br />

N<br />

4<br />

3<br />

log ρ [Ωm]<br />

2<br />

1<br />

0


Hansen,P.C.&O’Leary,D.P.(1993). The use of the L-curve in the regularization of discrete<br />

ill-posed problems. SIAM J. Sci. Comput., 14, 1487–1503.<br />

Hengesbach, L. (2006). Magnetotellurische Studien im Nordwestdeutschen Becken: Ein Beitrag<br />

zur paleaogeographischen Entwicklung des Unterkarbons. Dissertation, Universität Münster.<br />

Hoffmann, N. & Franke, D. (2008). The Avalonia-Baltica suture in NE Germany - new constraints<br />

and alternative interpretations. Abschlußbericht, Bundesanstalt für Geowissenschaften<br />

und Rohstoffe, Hannover.<br />

Hoffmann, N., Fluche, B., Jödicke, H., Jording, A., Kallaus, G., Müller, W. & Pasternak,<br />

G. (1997). Erforschung des tieferen Untergrundes der Nordwestdeutschen Senke - ein<br />

Beitrag der Magnetotellurik zur Untersuchung des präwestfalen Muttergesteinpotentials. Abschlußbericht,<br />

Bundesanstalt für Geowissenschaften und Rohstoffe, Hannover.<br />

Hoffmann, N., Jödicke, H., Fluche, B., Jording, A. & Müller, W. (1998). Modellvorstellungen<br />

zur Verbreitung potentieller präwestfalischer Erdgas-Muttergesteine in Norddeutschland<br />

– Ergebnisse neuer magnetotellurischer Messungen. Z. angew. Geol., 44, 140–158.<br />

Hoth, K., Rusbühlt, J., Zagora, K., Beer, H. & Hartmann, O. (1993). Die tiefen Bohrungen<br />

im Zentralabschnitt der Mitteleuropäischen Senke. Dokumentation für den Zeitabschnitt<br />

1962 bis 1990. In: Schriftenr. Geowiss. Berlin, 7–145.<br />

Hoth, P., Seibt, A., Kellner, T. & Huenges, E. (1997). Geowissenschaftliche Bewertungsgrundlagen<br />

zur Nutzung hydrogeothermaler Ressourcen in Norddeutschland. <strong>GFZ</strong> Potsdam - Sc.<br />

Techn. Report, 15.<br />

Houpt, L. (2008). Neue elektromagnetische Untersuchungen zur norddeutschen Leifähigkeitsanomalie.<br />

Diploma thesis, Fachrichtung Geophysik, FU Berlin.<br />

Jankowski, J. (1967). The Marginal Structures of the East European Platform in Poland on<br />

Basis of Data on Geomagnetic Field Variations. Polish Scientific Publishers, 93-102.<br />

Jödicke, H. (1991). Zonen hoher elektrischer Krustenleitfäigkeit im Rhenoherzynikum und<br />

seinem nördlichen Vorland. Dissertation, Universität Münster, Münster, Hamburg.<br />

Magri, F., Bayer, U., Jahnke, C., Clausnitzer, V., Diersch, H. J., Fuhrmann, J.,<br />

Möller, P., Pekdeger, A., Tesmer, M. & Voigt, H. J. (2005). Fluid-dynamics driving<br />

saline water in the North East German Basin. Int. J. Earth Sci., 94, 1056–1069.<br />

Magri,F.,Bayer,U.,Tesmer,M.,Möller, P. & Pekdeger, A. (2007). Salinization problems<br />

in the NEGB: results from thermohaline simulations. Int. J. Earth Sci., 97, 1075–1085.<br />

McCann, T. & Krawczyk, C. M. (2001). The Trans-European Fault: a critical reassessment.<br />

Geol. Mag., 138 (1), 19–29.<br />

Pharaoh, T. C. (1999). Paleozoic terranes and their lithospheric boundaries within the Trans-<br />

European Sutur Zone (TESZ): a review. Tectonophysics, 314, 17–41.<br />

Raab, S., Hoth, P., Huenges, E. & Müller, H. J. (1998). Role of sulfur and carbon in the<br />

electrical conductivity of the middle crust. J. Geophys. Res., 103, 9681–9689.<br />

Rodi,W.&Mackie,R.L.(2001). Nonlinear conjugate gradients algorithm for 2-D magnetotelluric<br />

inversions. Geophysics, 66, 174–187.<br />

Schäfer, A. (2010). Magnetotellurische Untersuchung der Transeuropäischen Suturzone auf einer<br />

Traverse von Südschweden nach Nordostdeutschland. Diploma thesis, Fachrichtung Geophysik,<br />

FU Berlin.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

261


Schmucker, U. (1959). Erdmagnetische Tiefensondierung in Deutschland 1957 – 59; Magnetogramme<br />

und erste Auswertung. Abh. Akad. Wiss., Göttingen. Math.–Phys. Kl. Heft 5.<br />

Smith, J. T. (1995). Understanding telluric distortion matrices. Geophysical Journal International,<br />

122, 219–226.<br />

Untiedt, J. (1970). Conductivity anomalies in Central and Southern Europe. J. Geomag. Geoel.,<br />

22, 131–149.<br />

Wiese, H. (1962). Geomagnetische Tiefentellurik Teil II: Die Streichrichtung der Untergrundstrukturen<br />

des elektrischen Widerstandes, erschlossen aus geomagnetischen Variationen. Geofis.<br />

Pura e Appl., 52, 83–103.<br />

Acknowledgement<br />

Other members of the EMTESZ Working Group include T. Ernst, V. Červ, J. Jankowski, W.<br />

Jó´zwiak, A. Kreutzmann, A. Neska, N. Palshin, L. Pedersen, M. Smirnov, E. Sokolova and I.<br />

Varentsov. We thank Belsk and Niemegk Observatories for supplying data and German Science<br />

Foundation (DFG) for funding.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

262


Magnetotelluric data from the Tien Shan and Pamir<br />

continental collision zones, Central Asia<br />

P. Sass*, O. Ritter*, A. Rybin**, G. Muñoz*, V. Batalev** and M. Gil*<br />

*Helmholtz Centre Potsdam <strong>GFZ</strong>, German Research Centre for Geosciences<br />

**Research Station of the Russian Academy of Sciences, Bishkek, Kyrgyzstan<br />

1 Intoduction<br />

We present magnetotelluic (MT) data obtained within the framework of the multidisciplinary<br />

Tien Shan - Pamir Geodynamic Program (TIPAGE). The dynamics of the<br />

Tien Shan and Pamir orogenic belts are dominated by the collision of the Indian and<br />

Eurasian continental plates. With the geophysical components, we intend to image the<br />

deepest active intra-continental subduction zones on Earth (the N-dipping Hindu Kush<br />

and the S-dipping Pamir zones) and to establish how the highest strain over the shortest<br />

distance that is manifested in the India-Asia collision zone is accommodated structurally.<br />

The Tien Shan-Pamirs mountain knot forms the northwestern corner of the India-Asia<br />

Figure 1: Central and eastern Asia orogens. TIPAGE and INDEPTH surveys are<br />

marked on the map. TIPAGE investigation area lies inside the blue circle.<br />

collision zone and the Pamir-Tibet Plateau (Fig. 1). Note that N - S shortening in<br />

the Pamir and western Tien Shan is absorbed in less than 50% of the distance than<br />

further east. Basins denote areas little affected by intra-continental shortening. The<br />

Pamir excels over the adjacent Tibet by including the most active areas of intermediatedepth<br />

seismicity in the world and by far the most active one not associated with oceanic<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

263


subduction, three coevally active, thick-skinned, high-strain contractional belts and four<br />

distinct intra-continental magmatic belts.<br />

2 Measurements<br />

The MT data were recorded in summer 2008 at 80 stations in the Pamir mountain<br />

ranges in Tajikistan and in summer 2009 at 98 stations in the Pamir and the southern<br />

Tien Shan in Kyrgyzstan and Tajikistan(Fig. 2). A typical spacing was approximately<br />

2 km between BB-only sites and 14 km for the combined BB+LMT sites. The stations<br />

form an approximately 340 km long profile from Osh in Kyrgyzstan via Sarytash, the<br />

Kyrgyz-Tajik border, Karakul and Murgab to Zorkul in southern Tajikistan.<br />

72˚<br />

73˚<br />

74˚<br />

75˚<br />

41˚ 41˚<br />

Uzbekistan<br />

Osh<br />

Kyrgyzstan<br />

40˚ 40˚<br />

39˚<br />

Kara-Kul<br />

Lake 186L<br />

39˚<br />

Tajikistan<br />

Sary-Tash<br />

Afganistan<br />

China<br />

7000<br />

Sarez<br />

Lake<br />

6000<br />

Murghab<br />

38˚<br />

5000<br />

38˚<br />

m<br />

134<br />

116L<br />

102L<br />

088L<br />

074L<br />

km<br />

356<br />

354L<br />

340L<br />

326L<br />

312L<br />

298L<br />

270L<br />

256L<br />

242L<br />

228L<br />

214L<br />

284L<br />

4000<br />

3000<br />

2000<br />

1000<br />

37˚<br />

0 50<br />

0<br />

37˚<br />

72˚<br />

73˚<br />

74˚<br />

75˚<br />

200L<br />

172L<br />

158L<br />

144L<br />

130L<br />

060L<br />

046L<br />

000<br />

032L<br />

018L<br />

004L<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

264<br />

Figure 2: Topographic map with<br />

stations and profile of the TIPAGE<br />

survey. Numbers of stations with<br />

broad band and long period measurements<br />

(L) are marked. The station<br />

alignement follows the a single<br />

road in that remote area with rough<br />

terrain.


3 Preliminary Data Analysis<br />

Data examples<br />

Some representative graphs of apparent resistivities and phases are shown in Fig. 3.<br />

The curves were obtained after a recording time of three days, using robust single site<br />

processing. Varying shapes of the curves indicate considerable changes in the underlying<br />

conductivity structure. The distances between the shown stations are approximately<br />

50 km. For station locations, see Fig. 2. The overall data quality is superb, especially<br />

in the very remote southern parts of the profile.<br />

Figure 3: Exemplary apparent resistivity and phase curves of four broad-band stations<br />

from the Pamir. For station locations, see Fig. 2.<br />

Pseudosections<br />

Apparent resistivities and impendance phases of all TIPAGE stations are plotted as<br />

pseudo sections in Fig. 4. In the southern part of the profile, the TE and TM modes<br />

apparent resistivities show values below 10 Ωm at higher periodes and high phase values<br />

reaching 90 ◦ . There are more low-resistivity features further north at middle to high<br />

periods, which are more present in the TE mode. This hints at a deep conducting<br />

structure in the southern profile part and a complex distribution of conducting features<br />

in the northern segment of the profile.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

265


Figure 4: Pseudosections of apparent resistivity and phase data, in TE and TM modes.<br />

In the left part of each pseudosection the southern profile stations are plotted, in the<br />

right part the northern stations. For stations location, see Fig. 2.<br />

Induction arrows<br />

The presence or absense of lateral variation in conductivity can be inferred from induction<br />

arrow maps. In Fig. 5 induction arrows are plotted using the Wiese convention for<br />

periods of 1, 32, 128, and 1024 s. The arrow distribution indicates several ’reversals’,<br />

which suggest the presence of elongated good conductors below the profile. Alltogether<br />

the induction arrow plots exhibit a complex distribution of the subsurface conductivity<br />

structure.<br />

Strike analysis<br />

Regional strike analysis (Becken & Burkhard, 2003) of sites recorded in the Pamir and<br />

Tien Shan is displayed in Fig. 6. The analysis of surface sensitive higher frequencies (Fig.<br />

6 left) shows a E - W (N - S) strike direction. This is consistent with the predominant<br />

E - W distribution of geological structures. The lower frequencies (Fig. 6 right) reveal a<br />

pronounced geoelectric strike of approximately −17 ◦ to −20 ◦ . This strike direction may<br />

be in agreement with structures of the India-Asian collision zone.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

266


Figure 5: Maps of induction vectors (Wiese convention) for periods of 1, 32, 128, and<br />

1024 s. Each arrow starts at one station location, compare Fig. 2.<br />

Figure 6: Regional strike analysis (Becken & Burkhard, 2003) of all sites recorded in<br />

the years 2008 and 2009. Left for the period range 0.001 s - 10 s, right for the period<br />

range 10 s - 10000 s.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

267


Figure 7: The Kyrgyz/Russian - Tajik - German MT team of 2008 (left image) and<br />

2009 (right image). The Research station of the Russian Academy of Sciences in Bishkek<br />

provided most of the field logistics, including 4WD vans and 4WD trucks.<br />

Figure 8: The amazing hospitality of local people during the field campaigns was a<br />

memorable experience.<br />

4 Acknowledgments<br />

The MT field team: Employees and students of the international Research Center,<br />

Russian Academy of Sciences, Bishkek and from Germany: D. Brändlein, X. Chen,<br />

T. Krings, A. Nube, M. Schüler, K. Tietze, C. Twardzik and G. Willkommen.<br />

Dr. V. E. Minaev, Dr. N. Radjabov, Dr. I. Oimahmadov, Prof. A. R. Faiziev: Institute<br />

of Geology, Academy of Sciences of the Republic of Tajikistan, Dushanbe.<br />

Dr. B. Moldobekov, Prof. H. Echtler, Dr. A. Mikolaichuk: Central Asian Institute for<br />

Applied Geosciences, Bishkek.<br />

Dr. S. K. Negmatullaev: PMP International / Seismic Monitoring Network in Tajikistan,<br />

Dushanbe.<br />

All other German colleagues participating in the TIPAGE collaborative research project.<br />

We very gratefully acknowledge substantial funding which we received from the <strong>GFZ</strong><br />

and the DFG. The magnetotelluric instruments were provided by the <strong>GFZ</strong> Geophysical<br />

Instrument Pool (GIPP).<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

268


23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

269


23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

270<br />

16 2/ 3


23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

271<br />

E x<br />

E y<br />

= Z xx<br />

Z yx<br />

Z xy<br />

Z yy<br />

B x<br />

B y


23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

272


Z B =0.5· (Z xy − Z yx )<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

273


23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

274


·<br />

·<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

275


23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

276


In our session "What I always wanted to know ..." we had the question<br />

How would you choose the period for LOTEM?<br />

Carsten Scholl continued thinking about it even after the colloquium and send an email with some<br />

additional thoughts. Here is his email:<br />

Well, in the discussion I said that it is not necessary to wait until the signal is decayed into the noise<br />

level before switching again, which is totally true. Actually, this would be to some extent a circular<br />

argument, because the noise level depends on the number of stacks, and the number of stacks<br />

depends on the period.<br />

So, one should check before with synthetic models, which time range is required to resolve a certain<br />

feature. Note, the time range to RESOLVE a feature, is not the same (and typically significantly larger)<br />

than the time range up to the point where deviations between the “target” and “no target” curves<br />

appear. I recommend doing 1D inversions with reasonable noise estimates for resolution studies… .<br />

Further, it is advisable to increase this time, because the true background resistivity might deviate<br />

from the synthetic und thus delay the features representing the target.<br />

The time saved by using a higher period and thus collecting the required number of stacks faster can<br />

be used by measuring at more sites.<br />

Often, in academic LOTEM campaigns, the target depth is not well defined but you’d like to get as<br />

deep as possible. Further, setting up LOTEM stations is tedious and the number of stations is limited<br />

anyways because of limited equipment. Often, the time spent on deploying and packing up the<br />

sensors exceeds the actual measurement time. So, in this case, there is not really the option to<br />

measure more sites on the same day (well, I always envision a roll-along scheme, where you start to<br />

measure at one station while building up the next site. When the final site is set up, the first site is<br />

redeployed, while the TX is still running…).<br />

In this case, I recommend to take some generous period the first day. In the evening the data for the<br />

day should be processed to see what the latest usable time is (this should be done with either a<br />

switch-off E-field or a magnetic component). For the subsequent days, the period should be set to<br />

something slightly longer (say, a factor of 2) longer than this time.<br />

Another aspect of this question: You should check, whether your time domain forward solver actually<br />

can handle higher switching times by taking into account previous transients. Otherwise, it might be<br />

better to stick to a more conservative period as well.<br />

23. Schmucker-Weidelt-Kolloquium für Elektromagnetische Tiefenforschung,<br />

Heimvolkshochschule am Seddiner See, 28 September - 2 Oktober 2009<br />

277

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!