22.07.2013 Views

Dynamics of the Eurasian Plate

Dynamics of the Eurasian Plate

Dynamics of the Eurasian Plate

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Dynamics</strong> <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> <strong>Plate</strong><br />

DIPLOMA THESIS<br />

presented to <strong>the</strong><br />

Department <strong>of</strong> Physics<br />

<strong>of</strong> <strong>the</strong><br />

Swiss Federal Institute <strong>of</strong> Technology Zurich<br />

by<br />

Mark T. Sargent<br />

March 2004<br />

Advisors:<br />

Pr<strong>of</strong>. Dr. Saskia Goes<br />

Gabriele Morra<br />

Institut für Geophysik<br />

ETH Hönggerberg<br />

Schaffmattstr. 30 (HPP)<br />

CH-8093 Zürich<br />

Switzerland


Acknowledgements<br />

It is a pleasure to thank Saskia Goes for guiding me both skillfully and patiently<br />

through this project and also Gabriele Morra for sharing his expertise in ABAQUS<br />

with me. They and <strong>the</strong> many o<strong>the</strong>r friendly and helpful people I met at <strong>the</strong> Institute<br />

<strong>of</strong> Geophysics provided me with a hospitable atmosphere in which I enjoyed working.<br />

My thanks also to Pr<strong>of</strong>essor Rice <strong>of</strong> <strong>the</strong> Institute <strong>of</strong> Theoretical Physics for agreeing<br />

to be <strong>the</strong> co-referee for this <strong>the</strong>sis.<br />

I am especially indebted to my mo<strong>the</strong>r and fa<strong>the</strong>r whose support and encouragement<br />

throughout my studies I deeply appreciate. In particular I am grateful for <strong>the</strong><br />

subscription to Scientific American <strong>the</strong>y gave me a couple <strong>of</strong> years ago, thanks to<br />

which I happened across <strong>the</strong> article on plate tectonics that awakened my interest in<br />

geophysics.<br />

I am also thankful to my fa<strong>the</strong>r for reading through this manuscript and reducing<br />

its linguistic shortcomings.<br />

Zurich, March 2004 Mark Sargent<br />

iii


Contents<br />

Acknowledgements iii<br />

1 Introduction 1<br />

2 The rectangular models 5<br />

2.1 The ABAQUS model . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2.2 Kinematic and dynamic boundary conditions . . . . . . . . . . . . . . 10<br />

2.3 The effect <strong>of</strong> lateral strength variations . . . . . . . . . . . . . . . . . 13<br />

2.4 The effect <strong>of</strong> different rheologies . . . . . . . . . . . . . . . . . . . . . 18<br />

3 The model <strong>of</strong> Eurasia 25<br />

3.1 The properties <strong>of</strong> <strong>the</strong> ABAQUS model . . . . . . . . . . . . . . . . . 26<br />

3.2 Forces revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

3.3 Rheology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.4 A comparison with measurements . . . . . . . . . . . . . . . . . . . . 30<br />

4 Conclusions 49<br />

Appendix A 53<br />

v


vi CONTENTS<br />

Appendix B 63<br />

Bibliography 79


THE CAUSE IS HIDDEN.<br />

THE EFFECT IS VISIBLE TO ALL.<br />

Ovid<br />

Chapter 1<br />

Introduction<br />

The past three decades have seen numerous attempts to numerically model stress<br />

patterns in <strong>the</strong> lithosphere <strong>of</strong> <strong>the</strong> Earth on both global and regional scales. These<br />

efforts have been indispensable in identifying <strong>the</strong> features we need to include in our<br />

endeavour to develop better models <strong>of</strong> our planet’s lithosphere and <strong>the</strong>y have also<br />

raised our awareness for <strong>the</strong> many unresolved issues that need to be addressed in<br />

<strong>the</strong> future.<br />

One such issue is our generally still very modest understanding <strong>of</strong> <strong>the</strong> forces driving<br />

plate tectonics. In principle <strong>the</strong> forces that play a role are those due to density<br />

contrasts within <strong>the</strong> lithosphere and <strong>the</strong> mantle. Tectonic plates are considered <strong>the</strong><br />

surface manifestation <strong>of</strong> mantle convection, but no mantle convection model to date<br />

has been able to satisfactorily reproduce plate tectonics. Work conducted so far<br />

indicates that nearly all aspects <strong>of</strong> plate generation require complex lithospheric<br />

rheologies invoking elasticity, viscosity and plasticity (Bercovici [3]). However, current<br />

technical and computational limitations do not allow <strong>the</strong> implementation <strong>of</strong><br />

complicated rheological behaviour on <strong>the</strong> scale <strong>of</strong> <strong>the</strong> entire mantle. Fur<strong>the</strong>rmore,<br />

<strong>the</strong> relevant parameters (e.g. viscosity or <strong>the</strong> importance <strong>of</strong> <strong>the</strong> water content in <strong>the</strong><br />

lithosphere and mantle) are still subject to considerable uncertainties.<br />

Studies including complex rheologies become feasible if <strong>the</strong> effect <strong>of</strong> forces due to<br />

density contrasts are parametrised on <strong>the</strong> scale <strong>of</strong> <strong>the</strong> lithosphere. The results <strong>of</strong><br />

such modeling have <strong>the</strong> advantage that <strong>the</strong>y can be compared with actual observations<br />

and thus increase our understanding <strong>of</strong> surface deformation and its evolution.<br />

They may also be useful in <strong>the</strong> development <strong>of</strong> self-consistent dynamic mantle models.<br />

In 1975 an article by Forsyth and Uyeda [11] listed <strong>the</strong> potential plate-driving<br />

forces that should be reflected in such a parametrisation and (by assuming that <strong>the</strong>ir<br />

values are <strong>the</strong> same worldwide) gave an estimate <strong>of</strong> <strong>the</strong>ir relative importance on a<br />

global scale. These include:<br />

1


2 CHAPTER 1. INTRODUCTION<br />

1. Forces due to internal lithospheric density contrasts:<br />

(a) <strong>the</strong> ridge push force FRP acting at divergent plate boundaries,<br />

(b) <strong>the</strong> slab pull force FSP , which pulls oceanic plates towards <strong>the</strong> trench <strong>of</strong> a<br />

subduction zone due to <strong>the</strong> tendency <strong>of</strong> <strong>the</strong> cold and heavier old oceanic<br />

lithosphere to sink into <strong>the</strong> mantle, and<br />

(c) <strong>the</strong> force FCM due to <strong>the</strong> difference in gravitational potential energy<br />

across continental margins.<br />

2. Forces due to (viscous) resistance<br />

(a) <strong>the</strong> mantle drag force FDF which is due to <strong>the</strong> viscous coupling between<br />

plates and <strong>the</strong> mantle beneath <strong>the</strong>m,<br />

(b) an additional drag force FCD beneath continental plates to account for<br />

<strong>the</strong> depth-dependent rheological properties <strong>of</strong> <strong>the</strong> mantle immediately<br />

underlying thin oceanic and thicker continental lithosphere,<br />

(c) <strong>the</strong> collisional resistance FCR acting on plate boundaries <strong>of</strong> converging<br />

continental lithosphere,<br />

(d) <strong>the</strong> slab resistance FSR, due to <strong>the</strong> viscous resistance <strong>the</strong> subducting slab<br />

encounters as it plunges into <strong>the</strong> mantle, and<br />

(e) <strong>the</strong> transform fault resistance FT F which counteracts strike-slip displacement<br />

on faults joining <strong>of</strong>fset mid-oceanic ridge segements.<br />

3. Viscous driving forces:<br />

(a) <strong>the</strong> trench suction force FSU which draws <strong>the</strong> overriding plate towards<br />

subduction zones because <strong>of</strong> regional mantle flow patterns induced by<br />

<strong>the</strong> subducting slab, and<br />

(b) <strong>the</strong> forces FDF and FCD if one assumes that mantle flow drives plate<br />

motions at <strong>the</strong> surface <strong>of</strong> <strong>the</strong> Earth ra<strong>the</strong>r than resisting <strong>the</strong>m (c.f. 2.(a)<br />

& 2.(b), above).<br />

Only <strong>the</strong> ridge push force, which reflects lateral density variations caused by<br />

spreading and cooling oceanic lithosphere, is understood well enough (see, for instance,<br />

Artyushkov [2] or Turcotte and Schubert [33]) to allow a quantitatively sound<br />

application as a boundary constraint in models <strong>of</strong> individual plates. In addition to<br />

<strong>the</strong>se forces that generate what is sometimes called <strong>the</strong> first order stress field, more<br />

recent approaches have also included <strong>the</strong> contributions <strong>of</strong> topography and lithospheric<br />

density variations to <strong>the</strong> stress field (e.g., Bird [4] or Lithgow-Bertelloni and<br />

Guynn [22]).<br />

The second important issue is that <strong>of</strong> <strong>the</strong> lithosphere’s rheological properties.<br />

Often <strong>the</strong> lithosphere is modeled as a purely elastic (or even rigid, as in <strong>the</strong> case <strong>of</strong><br />

Forsyth and Uyeda [11]) thin shell, following <strong>the</strong> argument that plates as a whole


ehave elastically away from <strong>the</strong> boundaries, and that viscous effects are negligible<br />

on <strong>the</strong> short time scales needed to compute <strong>the</strong> current stress field (e.g., Gölke and<br />

Coblentz [15], Grünthal and Stromeyer [16] or Lithgow-Bertelloni and Guynn [22]).<br />

On <strong>the</strong> o<strong>the</strong>r hand, in <strong>the</strong>ir publications investigating plate motions from <strong>the</strong> Cenozoic<br />

to <strong>the</strong> present epoch, Lithgow-Bertelloni and Richards [23] and Conrad and<br />

Lithgow-Bertelloni [8] resort to a purely viscous rheology for <strong>the</strong>ir lithosphere.<br />

The concept <strong>of</strong> plasticity is used by Regenauer-Lieb and Petit [28] and Hochstein<br />

and Regenauer-Lieb [19] in <strong>the</strong>ir models <strong>of</strong> <strong>the</strong> Alpine and Himalayan collisions and<br />

Bird [4] also introduces a plastic yield limit. Fur<strong>the</strong>rmore Lithgow-Bertelloni and<br />

Guynn [22] maintain that, by limiting <strong>the</strong> maximum harmonic degree <strong>of</strong> <strong>the</strong>ir fluid<br />

velocity field, <strong>the</strong>y too implicitly specify a yield strength.<br />

Inevitably perhaps, given <strong>the</strong> complexity <strong>of</strong> <strong>the</strong> task, up until now all modelers<br />

have decided to focus on those factors <strong>the</strong>y expect to contribute most to <strong>the</strong>ir subject<br />

<strong>of</strong> investigation. Lithgow-Bertelloni and Richards [23], Conrad and Lithgow-<br />

Bertelloni [8] and Lithgow-Bertelloni and Guynn [22] base <strong>the</strong>ir global models <strong>of</strong><br />

lithospheric stress patterns and plate motions primarily on models <strong>of</strong> density inhomogeneities<br />

in <strong>the</strong> mantle that generate mantle flow and thus traction on <strong>the</strong> base <strong>of</strong><br />

<strong>the</strong> lithosphere. Grünthal and Stromeyer [16] and Gölke and Coblentz [15], on <strong>the</strong><br />

o<strong>the</strong>r hand, neglect viscous coupling to <strong>the</strong> mantle altoge<strong>the</strong>r in <strong>the</strong>ir models <strong>of</strong> <strong>the</strong><br />

European part <strong>of</strong> Eurasia, because <strong>the</strong>y expect it to be negligible due to Eurasia’s<br />

slow motion.<br />

In yet ano<strong>the</strong>r approach Bird [4] imposes plate velocity patterns as <strong>the</strong> boundary<br />

condition on <strong>the</strong> base <strong>of</strong> <strong>the</strong> lithosphere. This in turn raises <strong>the</strong> question <strong>of</strong> whe<strong>the</strong>r<br />

it is more appropriate to apply kinematic or dynamic boundary conditions in <strong>the</strong>se<br />

kinds <strong>of</strong> models.<br />

Concerning <strong>the</strong> forces acting on <strong>the</strong> surface and edges <strong>of</strong> <strong>the</strong> plates, Bird [4] and<br />

Lithgow-Bertelloni and Guynn [22], for instance, incorporate <strong>the</strong> effects <strong>of</strong> topography<br />

and lateral density contrasts in <strong>the</strong> lithosphere and thus account for ridge push,<br />

slab pull and topographically induces stresses, but in <strong>the</strong> case <strong>of</strong> <strong>the</strong> latter study<br />

<strong>the</strong> collisional forces between <strong>the</strong> plates are omitted. The enigmatic trench suction<br />

force is invoked by Loohuis et al. [24] and Conrad and Lithgow-Bertelloni [8], but<br />

nei<strong>the</strong>r <strong>of</strong> <strong>the</strong>se projects accounts for topography.<br />

In <strong>the</strong> studies presented here we have tried to shed light on some <strong>of</strong> <strong>the</strong> open<br />

questions pertaining to stress modeling. These range from issues as fundamental as<br />

why it remains necessary to isolate <strong>the</strong> lithosphere from mantle flow in models <strong>of</strong><br />

plate dynamics, to specific questions concerning regional stress patterns. In <strong>the</strong> case<br />

<strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate one would like to know if <strong>the</strong> collision with India is responsible<br />

for extensional tectonics in <strong>the</strong> region <strong>of</strong> Lake Baikal, how in China compression<br />

can be aligned east-west when we would expect trench suction to lead to extension<br />

along that axis and if <strong>the</strong> <strong>Eurasian</strong> plate is a single unit or ra<strong>the</strong>r made up <strong>of</strong> smaller<br />

plates. In Western Europe itself it is not clear why stress directions trend NW-SE<br />

when previous models (e.g., Goes et al. [14] and Loohuis et al. [24]) predict <strong>the</strong>m<br />

to be roughly east-west, and why <strong>the</strong> same area experiences not only compressional<br />

3


4 CHAPTER 1. INTRODUCTION<br />

tectonics as expected from its setting between ridge push and Alpine collision, but<br />

also normal and strike-slip faulting.<br />

We have chosen to concentrate on <strong>the</strong> importance <strong>of</strong> rheology and lateral strength<br />

variations for lithospheric stress patterns and use our findings to build a model <strong>of</strong> <strong>the</strong><br />

<strong>Eurasian</strong> plate. In doing so we want to go beyond purely elastic models <strong>of</strong> <strong>the</strong> Central<br />

European stress field and to develop a model for this area which is not artificially<br />

cut at its eastern border as has been done by Gölke and Coblentz [15], Grünthal and<br />

Stromeyer [16] or Regenauer-Lieb and Petit [28], who argue that <strong>the</strong> presence <strong>of</strong> <strong>the</strong><br />

East European platform permits such a simplification. By investigating <strong>the</strong> influence<br />

<strong>of</strong> lateral strength contrasts such as cratons 1 on stress trajectories we should be able<br />

to determine if this technique is indeed justified or not.<br />

We begin with a finite element analysis <strong>of</strong> rectangular pieces <strong>of</strong> elastic, plastic,<br />

viscoelastic or elastoviscoplastic lithosphere in which we place regions <strong>of</strong> thinner or<br />

thicker lithosphere in various locations and observe <strong>the</strong>ir response to ei<strong>the</strong>r kinematic<br />

or dynamic boundary constraints. By proceeding in this manner we hope to identify<br />

<strong>the</strong> “fundamental” features <strong>of</strong> <strong>the</strong> stress field before we embark on a model <strong>of</strong> <strong>the</strong><br />

<strong>Eurasian</strong> plate in which geometrical effects may also contribute to <strong>the</strong> stress pattern.<br />

1 A craton is a stable area <strong>of</strong> continental crust that has not undergone much plate tectonic or<br />

orogenic activity for a long period. A craton includes a crystalline basement <strong>of</strong> rock (commonly<br />

Precambrian) called a shield, and a platform in which flat-lying or nearly flat-lying sediments or<br />

sedimentary rock surround <strong>the</strong> shield. (The Schlumberger Oilfield Glossary [31])


OUR MIDDLE CLASS CONSISTS OF EQUILATERAL OR EQUAL-SIDED TRIANGLES.<br />

OUR PROFESSIONAL MEN AND GENTLEMEN ARE SQUARES [...] OR PENTAGONS.<br />

Edwin A. Abbott, ’Concerning <strong>the</strong> Inhabitants <strong>of</strong> Flatland’ [1]<br />

Chapter 2<br />

The rectangular models<br />

2.1 The ABAQUS model<br />

Our first object <strong>of</strong> study provides us with a setting to explore <strong>the</strong> behaviour <strong>of</strong> <strong>the</strong><br />

stress field for different kinds <strong>of</strong> boundary conditions and rheological properties <strong>of</strong><br />

<strong>the</strong> lithosphere when we place structures <strong>of</strong> varying strength in it. By working with a<br />

geometrically simple outline, in our case a rectangle, we do not have to worry about<br />

<strong>the</strong> diverging stress concentrations that can occur in <strong>the</strong> corners <strong>of</strong> more complex<br />

shapes in finite element modeling and which might mask <strong>the</strong> effects we are interested<br />

in.<br />

To enable rough comparisons <strong>of</strong> <strong>the</strong> stress patterns in our initial models with those<br />

actually measured in <strong>the</strong> <strong>Eurasian</strong> plate, our rectangle approximately covers <strong>the</strong><br />

area <strong>of</strong> <strong>the</strong> latter. Figure 2.1 shows <strong>the</strong> rectangular finite element mesh used in<br />

ABAQUS [18], superimposed on an oblique Mercator projection 1 <strong>of</strong> Eurasia, <strong>the</strong><br />

boundaries <strong>of</strong> which are based on <strong>the</strong> NUVEL1 model <strong>of</strong> plate velocities (DeMets,<br />

Gordon, Argus and Stein [10]).<br />

The grid is composed <strong>of</strong> 1800 elements with an area <strong>of</strong> 7.1 × 10 10 m 2 each, giving<br />

<strong>the</strong> whole rectangle a total area <strong>of</strong> 1.28 × 10 8 km 2 . The distance between two neighbouring<br />

nodes is 300 km which corresponds to about 4.2 ◦ at 50.3 ◦ N, <strong>the</strong> latitude<br />

<strong>of</strong> <strong>the</strong> oblique Mercator projection’s origin. The two-dimensional elements making<br />

up <strong>the</strong> grid are so called plane stress elements that can be used when <strong>the</strong> thickness<br />

<strong>of</strong> a body is small relative to its lateral (in-plane) dimensions. The stresses<br />

1 As we will be working with two-dimensional models, it is essential to find a projection that<br />

changes <strong>the</strong> plate’s area as little as possible when going from <strong>the</strong> Earth’s spherical geometry to a<br />

flat surface. The requirement <strong>of</strong> minimal distortion determined our choice <strong>of</strong> an oblique Mercator<br />

projection with origin at 98.8 ◦ E / 50.3 ◦ N. (In an ordinary Mercator projection, for instance,<br />

<strong>the</strong> importance <strong>of</strong> ridge push along Eurasia’s nor<strong>the</strong>rn boundary in <strong>the</strong> Arctic ocean would be<br />

exaggerated).<br />

5


6 CHAPTER 2. THE RECTANGULAR MODELS<br />

130˚W<br />

120˚W<br />

40˚N<br />

110˚W<br />

100˚W<br />

90˚W<br />

80˚W<br />

70˚W<br />

60˚W<br />

40˚N<br />

50˚W<br />

40˚W<br />

30˚N<br />

30˚W<br />

20˚W<br />

10˚W<br />

30˚N<br />

0˚<br />

10˚E<br />

20˚E<br />

30˚N<br />

Figure 2.1: The finite element mesh used for <strong>the</strong> rectangular models, superimposed on <strong>the</strong><br />

<strong>Eurasian</strong> plate (oblique Mercator projection with origin at 98.8 ◦ E / 50.3 ◦ N).<br />

are functions <strong>of</strong> planar coordinates alone, out-<strong>of</strong>-plane normal and shear stresses<br />

are equal to zero and all loading and deformation are also restricted to this plane.<br />

Hence, with our models, we will not be able to predict <strong>the</strong> height <strong>of</strong> mountains<br />

forming in regions <strong>of</strong> continental collision and compare <strong>the</strong>m with actual elevations.<br />

In contrast to many earlier studies employing triangular elements (e.g., Bird [4],<br />

Gölke and Coblentz [15], Loohuis et al. [24]) we work with four-sided elements. In<br />

<strong>the</strong> Lagrangian formulation ABAQUS 2 uses, deformation is always underestimated<br />

(c.f. Getting Started with ABAQUS/Standard, p. 4-4) because shear locking causes<br />

<strong>the</strong> elements to be too stiff. Shear locking is not a problem for reasonably regular<br />

four-sided elements since <strong>the</strong>ir edges are able to curve, yet it does affect <strong>the</strong> results<br />

if one wants to investigate deformation with triangular elements 3 .<br />

Eurasia borders on <strong>the</strong> North American, <strong>the</strong> Pacific, <strong>the</strong> Philippine, <strong>the</strong> Indian,<br />

2 Appendix A contains a basic introduction into <strong>the</strong> <strong>the</strong>oretical concepts and <strong>the</strong> procedures <strong>of</strong><br />

<strong>the</strong> finite element analysis performed in ABAQUS.<br />

3 It should be mentioned that <strong>the</strong> triangular elements used by Bird [4], Gölke and Coblentz [15]<br />

and Loohuis et al. [24] can be expected to perform well if only used to model stresses.<br />

30˚E<br />

20˚N<br />

40˚E<br />

10˚N<br />

0˚<br />

10˚S


2.1. THE ABAQUS MODEL 7<br />

Figure 2.2: Map <strong>of</strong> <strong>the</strong> major tectonic plates <strong>of</strong> <strong>the</strong> world. (Courtesy <strong>of</strong> <strong>the</strong> U.S. Geological<br />

Survey.)<br />

<strong>the</strong> Australian, <strong>the</strong> Arabian and <strong>the</strong> African plates (see Figure 2.2) 4 . Its boundaries<br />

with <strong>the</strong>m include divergent margins along <strong>the</strong> Mid-Atlantic Ridge and <strong>the</strong> Arctic<br />

Mid-Ocean Ridge, collisional boundaries along most <strong>of</strong> Eurasia’s sou<strong>the</strong>rn border<br />

from <strong>the</strong> Mediterranean Sea to <strong>the</strong> Himalayas, subduction zones in Sou<strong>the</strong>ast Asia<br />

and west <strong>of</strong> Japan and, finally, two segments with mostly strike-slip displacement<br />

that link <strong>the</strong> Mid-Atlantic Ridge with <strong>the</strong> Mediterranean and <strong>the</strong> Arctic Mid-Ocean<br />

Ridge with <strong>the</strong> Pacific subduction zones.<br />

In order to make <strong>the</strong> preliminary models as realistic as possible in spite <strong>of</strong> <strong>the</strong>ir<br />

simplified geometry, <strong>the</strong> different boundary conditions were applied to <strong>the</strong> sides <strong>of</strong><br />

<strong>the</strong> rectangle as follows:<br />

1. ridge push FRP along <strong>the</strong> left edge (Figure 2.3) and along three quarters <strong>of</strong> <strong>the</strong><br />

upper edge, <strong>of</strong> a magnitude appropriate to <strong>the</strong> mean age <strong>of</strong> <strong>the</strong> oceanic lithosphere<br />

along Europe’s western continental margin 5 (approximately 80 million<br />

years) and in <strong>the</strong> Arctic ocean (50 million years),<br />

4 Recently scientists (e.g., Bird [5]) have suggested <strong>the</strong> existence <strong>of</strong> ano<strong>the</strong>r 38 small plates, <strong>of</strong><br />

which eight are within <strong>the</strong> area labeled as “<strong>Eurasian</strong> plate” in <strong>the</strong> map.<br />

5 Note that we are not applying ridge push as a pressure distributed through <strong>the</strong> entire oceanic<br />

lithosphere, as would be strictly correct, because <strong>the</strong> implementation <strong>of</strong> such a force in our two-


8 CHAPTER 2. THE RECTANGULAR MODELS<br />

Boundary force Forsyth & Uyeda Loohuis et al. Coblentz et al.<br />

FSU<br />

∼ 3<br />

2FRP 0.8 − 2.1 × 1012 FCC<br />

∼<br />

-<br />

1<br />

2FRP 2.1 − 3.1 × 1012 2 × 1012 Table 2.1: Estimated values for <strong>the</strong> trench suction and continental collision forces FSU and FCC<br />

(in Nm −1 ) as given by <strong>the</strong> three listed authors. Coblentz’ value for FCC is for <strong>the</strong> Himalayas, <strong>the</strong><br />

one <strong>of</strong> Loohuis for Eurasia as a whole. The range over which <strong>the</strong> values <strong>of</strong> FRP vary can be found<br />

in <strong>the</strong> following table.<br />

2. continental margin forces FCM opposing ridge push along <strong>the</strong> same segments,<br />

3. a free border along what remains <strong>of</strong> <strong>the</strong> upper edge, representing a zone <strong>of</strong><br />

probably mostly strike-slip displacement (Gaina, Roest and Müller [12]) across<br />

Siberia where Eurasia borders on <strong>the</strong> North American plate,<br />

4. trench suction FSU, drawing Eurasia towards <strong>the</strong> subduction zones in <strong>the</strong><br />

Pacific and Indian Oceans, implemented along <strong>the</strong> rectangle’s right and a<br />

segment <strong>of</strong> its lower edge,<br />

5. continental collision where <strong>the</strong> Indian, Arabian and African plates converge<br />

on Eurasia (applied to most <strong>of</strong> <strong>the</strong> lower edge).<br />

To date, exact calculations <strong>of</strong> most <strong>of</strong> <strong>the</strong>se boundary forces have not been carried<br />

out because many <strong>of</strong> <strong>the</strong> physical processes <strong>the</strong>y involve are not well known 6 .<br />

However, several authors (Forsyth and Uyeda [11], Loohuis et al. [24] and Coblentz,<br />

Sandiford, Richardson, Zhou and Hillis [7]) have published estimates <strong>of</strong> <strong>the</strong>ir values<br />

which <strong>the</strong>y obtained from global or regional models <strong>of</strong> plate dynamics. Table 2.1<br />

lists <strong>the</strong> values <strong>the</strong>se authors obtained from <strong>the</strong>ir models and Table 2.2 provides<br />

an overview <strong>of</strong> <strong>the</strong> dependence <strong>of</strong> <strong>the</strong> ridge push force - <strong>the</strong> only numerically well<br />

constrained boundary force - on <strong>the</strong> age <strong>of</strong> <strong>the</strong> oceanic lithosphere.<br />

Apart from <strong>the</strong> forces acting on <strong>the</strong> plate’s edges <strong>the</strong>re is also <strong>the</strong> mantle drag<br />

force, caused by <strong>the</strong> viscous coupling between lithosphere and <strong>the</strong> underlying mantle.<br />

In all our models it resists plate motion, i.e. we assume that, even if mantle flow<br />

patterns can play a role locally in trench suction, it is not driving <strong>the</strong> <strong>Eurasian</strong><br />

plate on a large scale. To implement basal drag with ABAQUS we assign so called<br />

’dashpot elements’ to <strong>the</strong> nodes <strong>of</strong> <strong>the</strong> finite element mesh. Based on observations,<br />

modelers <strong>of</strong> plate dynamics and kinematics <strong>of</strong>ten use <strong>the</strong> working hypo<strong>the</strong>sis that<br />

plate motions are not accelerated over long periods and ensure this by balancing<br />

dimensional ABAQUS model is not convenient. Instead, we chose to take <strong>the</strong> edge <strong>of</strong> <strong>the</strong> slab<br />

as <strong>the</strong> border <strong>of</strong> Eurasia’s continental lithosphere and apply to it (as a line force) <strong>the</strong> value <strong>of</strong><br />

ridge push as integrated over <strong>the</strong> whole oceanic lithosphere between <strong>the</strong> Mid-Oceanic ridge and<br />

<strong>the</strong> continental margin.<br />

6 Scholz and Campos [32] note that “The state <strong>of</strong> understanding <strong>of</strong> this topic [<strong>the</strong> nature <strong>of</strong><br />

forces in subduction zones] at present is probably best described as confused.”


2.1. THE ABAQUS MODEL 9<br />

Age (in millions <strong>of</strong> years) Ridge push FRP [Nm −1 ]<br />

10 0.484×10 12<br />

20 0.968×10 12<br />

30 1.452×10 12<br />

40 1.937×10 12<br />

50 2.421×10 12<br />

60 2.905×10 12<br />

70 3.389×10 12<br />

80 3.873×10 12<br />

90 4.357×10 12<br />

100 4.841×10 12<br />

110 5.326×10 12<br />

120 5.810×10 12<br />

Table 2.2: The dependence <strong>of</strong> <strong>the</strong> magnitude <strong>of</strong> FRP on <strong>the</strong> age <strong>of</strong> <strong>the</strong> oceanic lithosphere (after<br />

Turcotte and Schubert [33])<br />

<strong>the</strong> torques <strong>of</strong> <strong>the</strong> forces involved (e.g., Forsyth and Uyeda [11], Loohuis et al. [24]<br />

or Lithgow-Bertelloni and Guynn [22]). The dashpots exert a resistive force on <strong>the</strong><br />

nodes that is proportional to <strong>the</strong> velocity 7 <strong>of</strong> <strong>the</strong> lithosphere at that point and our<br />

tests show that this leads to a steady state velocity field in our models as well, even<br />

if we do not explicitly apply a torque balance.<br />

A thorough understanding <strong>of</strong> <strong>the</strong> relevant boundary forces is mandatory for a realistic<br />

model <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate, but <strong>the</strong>ir precise values are <strong>of</strong> no great importance<br />

for <strong>the</strong> rectangular models, <strong>the</strong> aim <strong>of</strong> which is to gain a qualitative understanding<br />

<strong>of</strong> <strong>the</strong> relevance <strong>of</strong> rheology and lateral strength variations for <strong>the</strong> characteristics<br />

<strong>of</strong> lithospheric stress. Actual values <strong>of</strong> <strong>the</strong> forces applied in <strong>the</strong> various rectangular<br />

models <strong>of</strong> <strong>the</strong> coming sections are listed toge<strong>the</strong>r with all o<strong>the</strong>r relevant model parameters<br />

for each model in Appendix B.<br />

While <strong>the</strong> multifaceted issue <strong>of</strong> boundary forces will be discussed in more detail in<br />

<strong>the</strong> next chapter, <strong>the</strong> subject <strong>of</strong> kinematic and dynamic boundary conditions will<br />

be dealt with in <strong>the</strong> following section.<br />

7 The input for <strong>the</strong> dashpot elements is <strong>the</strong> coefficient <strong>of</strong> friction that relates <strong>the</strong> force to <strong>the</strong><br />

velocity. It is a function <strong>of</strong> viscosity, which, in turn, determines <strong>the</strong> value <strong>of</strong> <strong>the</strong> shear stress τLA<br />

between lithosphere and upper mantle (Turcotte and Schubert [33], p.230):<br />

τLA = −2ηvref<br />

h<br />

· (2 + 3 hL<br />

), (2.1)<br />

h<br />

where η ≈ 10 19 − 10 20 P a · s and h ≈ 220 km are <strong>the</strong> viscosity and thickness <strong>of</strong> <strong>the</strong> low-viscosity<br />

upper mantle and hL <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> lithosphere. By multiplying this basal shear stress with<br />

<strong>the</strong> area <strong>of</strong> an element one obtains <strong>the</strong> force acting on such an element and, after dividing <strong>the</strong><br />

result by vref (<strong>the</strong> average plate velocity), <strong>the</strong> value <strong>of</strong> <strong>the</strong> coefficient <strong>of</strong> friction that needs to be<br />

assigned to <strong>the</strong> dashpots.


10 CHAPTER 2. THE RECTANGULAR MODELS<br />

2.2 Kinematic and dynamic boundary conditions<br />

While <strong>the</strong> magnitudes <strong>of</strong> boundary forces are subject to considerable uncertainties,<br />

current plate velocities are quite well known. One could thus be tempted to use kinematic<br />

ra<strong>the</strong>r than dynamic boundary conditions, but considerable caution should<br />

be exercised in doing so. The relative velocities at plate boundaries do not contain<br />

any information on <strong>the</strong> prevailing mechanisms and <strong>the</strong> forces definitely do not scale<br />

directly with <strong>the</strong> velocities.<br />

The inconsistencies between <strong>the</strong> dynamic and <strong>the</strong> kinematic approach are clearly<br />

visible when one compares two different segments where Eurasia possesses a convergent<br />

plate boundary: (1) <strong>the</strong> subduction zones in <strong>the</strong> Far East and (2) <strong>the</strong> Himalayas.<br />

In <strong>the</strong> latter location <strong>the</strong> relative rate <strong>of</strong> convergence is 5 centimeters per year on<br />

average, whereas in <strong>the</strong> former it is almost twice as high. Yet, if one assumes that<br />

<strong>the</strong> <strong>Eurasian</strong> plate is being drawn towards <strong>the</strong> Pacific trenches by suctional effects<br />

(Conrad and Lithgow-Bertelloni [8]), <strong>the</strong>n <strong>the</strong> relevant quantity is <strong>the</strong> absolute and<br />

outwardly directed velocity <strong>of</strong> Eurasia (approx. 1 centimeter per year). Due to <strong>the</strong><br />

high viscosity <strong>of</strong> <strong>the</strong> mantle, this modest speed implies a large force FSU acting on<br />

<strong>the</strong> eastern margin <strong>of</strong> Eurasia and a simple calculation shows it to lie between one<br />

and two times <strong>the</strong> magnitude <strong>of</strong> ridge push FRP (in approximate agreement with<br />

Forsyth and Uyeda [11] and Loohuis et al. [24], see Table 2.1 above):<br />

FSU ≈ η × (velocity <strong>of</strong> <strong>the</strong> overriding plate at trench)<br />

= 4 × 10 22 P a · s × 0.94cm/yr = 3.17 × 10 12 Nm −1 . (2.2)<br />

The force FCC acting in areas <strong>of</strong> continental collisions, however, is estimated<br />

(Forsyth and Uyeda [11], Loohuis et al. [24]) to be only half <strong>the</strong> size <strong>of</strong> <strong>the</strong> ridge<br />

push force FRP . One <strong>the</strong>refore faces <strong>the</strong> paradox that in <strong>the</strong> Himalayas one has lower<br />

forces in spite <strong>of</strong> quite high velocities, but near <strong>the</strong> trenches exactly <strong>the</strong> opposite is<br />

<strong>the</strong> case.<br />

Depending on <strong>the</strong> kind <strong>of</strong> boundary conditions applied to <strong>the</strong> model slab, <strong>the</strong><br />

stress pattern in its interior could thus vary significantly, as is indeed <strong>the</strong> case<br />

in Figures 2.3 and 2.4, which show <strong>the</strong> stress intensity 8 (top) and stress direction<br />

(bottom) in a purely elastic rectangle for <strong>the</strong> two cases described. When <strong>the</strong> relative<br />

velocities <strong>of</strong> <strong>the</strong> neighbouring plates with respect to Eurasia are applied as boundary<br />

8 The quantity plotted is <strong>the</strong> von Mises stress (a measure <strong>of</strong> stress intensity <strong>of</strong>ten used in engi-<br />

neering):<br />

σν =<br />

where <strong>the</strong> σi are <strong>the</strong> principal stresses.<br />

1<br />

2 [(σ1 − σ2) 2 + (σ2 − σ3) 2 + (σ1 − σ3) 2 ], (2.3)


2.2. KINEMATIC AND DYNAMIC BOUNDARY CONDITIONS 11<br />

Figure 2.3: Stress pattern for <strong>the</strong> model kin23-02-04 1, in which kinematic boundary conditions<br />

were applied. The values next to <strong>the</strong> colour scale in <strong>the</strong> upper half are in units <strong>of</strong> Pa. Yellow lines<br />

in <strong>the</strong> lower half <strong>of</strong> <strong>the</strong> picture give <strong>the</strong> orientation <strong>of</strong> extensional principal stress, red lines that <strong>of</strong><br />

compressional principal stress.<br />

constraints, one receives an almost purely compressional stress pattern, whereas in<br />

<strong>the</strong> case <strong>of</strong> dynamic boundary conditions one finds that extensional stresses dominate<br />

on <strong>the</strong> right side <strong>of</strong> <strong>the</strong> plate, due to <strong>the</strong> application <strong>of</strong> trench suction. A comparison<br />

<strong>of</strong> <strong>the</strong> stress fields <strong>of</strong> both models shows not only a different stress pattern, but<br />

also substantially higher stresses in <strong>the</strong> kinematic case. Velocities are very strong<br />

constraints forcing a region to move as prescribed, even if this induces unrealistically<br />

high stress values. Fur<strong>the</strong>rmore, being instantaneous quantities and probably only<br />

valid in <strong>the</strong> crust <strong>of</strong> <strong>the</strong> Earth, velocities are less suited as long term boundary<br />

conditions (we run <strong>the</strong> majority <strong>of</strong> our models for 1.5 million years) than are forces<br />

which are less susceptible to rapid changes (leaving aside events like <strong>the</strong> onset <strong>of</strong><br />

continental collision). The additional benefit <strong>of</strong> dynamic boundary constraints is<br />

that - at least in <strong>the</strong> case <strong>of</strong> ridge push and slab pull - <strong>the</strong>y are averages over <strong>the</strong><br />

entire thickness <strong>of</strong> <strong>the</strong> lithosphere, or are held to be <strong>of</strong> <strong>the</strong> same order <strong>of</strong> magnitude<br />

as FRP and FSP . Hence, forces differ among each o<strong>the</strong>r at most by a factor <strong>of</strong> two<br />

or three, which is a smaller range than with plate velocities, that can vary over two


12 CHAPTER 2. THE RECTANGULAR MODELS<br />

Figure 2.4: Stress pattern for <strong>the</strong> model dsm elastic, in which dynamic boundary conditions were<br />

applied.<br />

orders <strong>of</strong> magnitude; from 1 millimeter per year to 10 centimetres a year.<br />

In spite <strong>of</strong> this, <strong>the</strong> application <strong>of</strong> kinematic boundary conditions should still be<br />

permissible if dynamic consistency is ensured. A way to achieve this is to verify<br />

that <strong>the</strong>y do not put more energy into a system than can be extracted from it<br />

(see Han and Gurnis [17] for a discussion <strong>of</strong> this issue in <strong>the</strong> context <strong>of</strong> subduction).<br />

Never<strong>the</strong>less, we believe that modeling stresses with <strong>the</strong> help <strong>of</strong> dynamic constraints<br />

is more consistent than doing so by applying kinematic boundary conditions and<br />

mixing kinematics and dynamics in <strong>the</strong> process. This choice is reflected throughout<br />

<strong>the</strong> rest <strong>of</strong> this <strong>the</strong>sis, which is dominated by models to which boundary forces,<br />

ra<strong>the</strong>r than velocities, have been applied.<br />

The finite element program ABAQUS <strong>of</strong>fers two different ways to implement<br />

boundary forces; one can apply <strong>the</strong>m as concentrated loads to <strong>the</strong> nodes on <strong>the</strong> edge<br />

<strong>of</strong> <strong>the</strong> plate or as distributed loads to <strong>the</strong> edge <strong>of</strong> an element. For <strong>the</strong> concentrated<br />

loads a direction may be specified while <strong>the</strong> distributed loads are basically a pressure<br />

that acts perpendicularly to <strong>the</strong> boundary surface. By applying <strong>the</strong> concentrated<br />

loads at right angles to <strong>the</strong> boundary (at closely spaced nodes) and comparing <strong>the</strong>


2.3. THE EFFECT OF LATERAL STRENGTH VARIATIONS 13<br />

Figure 2.5: Map <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate (adapted from Villaseñor et al. [34]), showing <strong>the</strong> location<br />

<strong>of</strong> <strong>the</strong> two major cratons in <strong>the</strong> region studied; <strong>the</strong> East European platform (EEP) and <strong>the</strong> Siberian<br />

craton (SC). The o<strong>the</strong>r labels are: AP - Arabian peninsula; AR - Andaman ridge; C - Caucasus;<br />

H - Himalayas; HD - Hangay dome; HK - Hindu-Kush; IS - Indian shield; TB - Tarim basin; TIP<br />

- Turkish-Iranian <strong>Plate</strong>au; TP - Tibetan <strong>Plate</strong>au; TS - Tien-Shan; U - Urals; Z - Zagros.<br />

result with a model applying distributed loading, we found that <strong>the</strong> two alternatives<br />

are equivalent.<br />

2.3 The effect <strong>of</strong> lateral strength variations on <strong>the</strong><br />

stress field<br />

In <strong>the</strong> past several studies <strong>of</strong> <strong>the</strong> European stress field have truncated <strong>the</strong>ir modeled<br />

area at approximately 40 ◦ East longitude (e.g., Gölke and Coblentz [15], Grünthal


14 CHAPTER 2. THE RECTANGULAR MODELS<br />

Figure 2.6: Stress pattern for <strong>the</strong> model dsm elasticrat. The two cratons are visible as dark blue<br />

patches <strong>of</strong> lower stress intensity.<br />

and Stromeyer [16] or Regenauer-Lieb and Petit [28]) because <strong>the</strong>y believe that a<br />

region <strong>of</strong> old and stable continental crust known as <strong>the</strong> East European platform -<br />

a so called craton - effectively shields Europe from <strong>the</strong> processes in <strong>the</strong> remaining<br />

parts <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate (see Figure 2.5). By placing regions <strong>of</strong> thicker or thinner<br />

lithosphere in our rectangular slab and observing <strong>the</strong> reaction <strong>of</strong> <strong>the</strong> stress field to<br />

<strong>the</strong>se inhomogeneities we can determine if: (1) stress levels change (get reduced in<br />

<strong>the</strong> case <strong>of</strong> thicker and thus mechanically stronger cratonic lithosphere) in <strong>the</strong> ’lee’<br />

<strong>of</strong> such structures, (2) stress directions get bent and (3) this happens in any setting<br />

or if <strong>the</strong> forces prevailing in a given region play a role too.<br />

We can also investigate if raising <strong>the</strong> dashpots’ coefficients <strong>of</strong> friction beneath <strong>the</strong><br />

cratons, thus simulating <strong>the</strong> higher drag that <strong>the</strong> deeper lithospheric roots <strong>of</strong> a<br />

craton experience due to viscosity growing with depth, anchors <strong>the</strong> plate at that<br />

point, <strong>the</strong>reby justifing <strong>the</strong> method <strong>of</strong> Gölke and Coblentz [15] who pin Europe’s<br />

Eastern margin.<br />

If one alters <strong>the</strong> model shown in Figure 2.4 by assigning a thickness <strong>of</strong> 150 km<br />

(compared with 60 km for <strong>the</strong> surrounding continental lithosphere) to <strong>the</strong> rectangle’s


2.3. THE EFFECT OF LATERAL STRENGTH VARIATIONS 15<br />

Figure 2.7: Stress pattern for <strong>the</strong> model dsm elastohypocrat1, which contains an ’artificial’<br />

craton next to <strong>the</strong> boundaries experiencing trench suction instead <strong>of</strong> <strong>the</strong> East European platform<br />

and <strong>the</strong> Siberian craton.<br />

elements that coincide with <strong>the</strong> East European platform and <strong>the</strong> Siberian craton on<br />

<strong>the</strong> map in Figure 2.1, <strong>the</strong>n <strong>the</strong> stress field depicted in Figure 2.6 results. The<br />

stress intensity map in <strong>the</strong> upper half <strong>of</strong> <strong>the</strong> pictures reveals more structure than in<br />

Figure 2.4. Both cratons are visible as dark blue patches <strong>of</strong> lower stress intensity in<br />

<strong>the</strong> upper half <strong>of</strong> <strong>the</strong> rectangle, a bit to <strong>the</strong> left and right <strong>of</strong> <strong>the</strong> centre. The East<br />

European platform is also clearly visible as a region <strong>of</strong> compressive stress oriented<br />

at roughly 45 ◦ from <strong>the</strong> upper left to <strong>the</strong> lower right in <strong>the</strong> lower half <strong>of</strong> <strong>the</strong> figure.<br />

In Figure 2.4 this area was dominated by a stress field at right angles to <strong>the</strong> lower<br />

border. On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong> Siberian craton, while visible in <strong>the</strong> stress intensity<br />

plot, cannot be distinguished in <strong>the</strong> perpendicular stress pattern generated by Arctic<br />

ridge push and continental collision <strong>of</strong> India with Eurasia.<br />

The same observation can be made in Figure 2.7; in this model <strong>the</strong>re is only one<br />

craton, located in <strong>the</strong> area that is experiencing extensional stresses due to trench<br />

suction. It too shows up in <strong>the</strong> stress intensity map but much less so in <strong>the</strong> lower part<br />

<strong>of</strong> <strong>the</strong> figure. Why does <strong>the</strong> influence <strong>of</strong> cratons vary in what, at first glance, might<br />

seem to be an unpredictable fashion? The results suggest that thicker and thus


16 CHAPTER 2. THE RECTANGULAR MODELS<br />

Figure 2.8: Stress pattern for <strong>the</strong> model dsm elastithincrat, which contains thinner ra<strong>the</strong>r than<br />

thicker lithosphere in <strong>the</strong> areas <strong>of</strong> <strong>the</strong> East European platform and <strong>the</strong> Siberian craton.<br />

stronger lithosphere always reduces stress values in <strong>the</strong> area <strong>of</strong> <strong>the</strong> craton. However<br />

it affects stress orientation only if <strong>the</strong> structure is located in an area under <strong>the</strong><br />

influence <strong>of</strong> more than one source <strong>of</strong> stress (in Figure 2.6 ridge push and continental<br />

collision) that lead to two competing fields <strong>of</strong> comparable magnitude. In this case,<br />

stress directions do tend to change at <strong>the</strong> edge <strong>of</strong> <strong>the</strong> craton and <strong>the</strong>reby induce<br />

so called ’stress bending’. Such bending <strong>of</strong> stress orientations has been invoked by<br />

Grünthal and Stromeyer [16] and Müller et al. [26] to explain European variations<br />

in stress style.<br />

The models indicate that both orientation and magnitude <strong>of</strong> <strong>the</strong> stress fields on<br />

<strong>the</strong> two sides <strong>of</strong> <strong>the</strong> structure are not necessarily independent. While cratons do lead<br />

to reduced stress intensities within <strong>the</strong> craton, <strong>the</strong>y do not seem to create a stress<br />

shadow, in which <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> stress field is significantly lower on <strong>the</strong> edge<br />

far<strong>the</strong>r from <strong>the</strong> source. As can be seen in Figure 2.6, <strong>the</strong> cratons are symmetrically<br />

surrounded by regions <strong>of</strong> higher stress values, and Figure 2.7 reveals that <strong>the</strong> stress<br />

on <strong>the</strong> left <strong>of</strong> <strong>the</strong> hypo<strong>the</strong>tical craton reaches <strong>the</strong> same values as to <strong>the</strong> right <strong>of</strong> it.<br />

In view <strong>of</strong> <strong>the</strong>se findings we are led to <strong>the</strong> conclusion that <strong>the</strong> valitity <strong>of</strong> Gölke and


2.3. THE EFFECT OF LATERAL STRENGTH VARIATIONS 17<br />

Figure 2.9: Full displacement field <strong>of</strong> <strong>the</strong> model dsm creepcrat after 32000 years; no discontinuities<br />

are observed at <strong>the</strong> edges <strong>of</strong> <strong>the</strong> cratons.<br />

Coblentz’ [15], Grünthal and Stromeyer’s [16] and Regenauer-Lieb and Petit’s [28]<br />

method mentioned above, in which <strong>the</strong>y cut <strong>the</strong> <strong>Eurasian</strong> lithosphere at <strong>the</strong> East<br />

European platform, is questionable.<br />

Having looked at <strong>the</strong> result <strong>of</strong> thickening a region <strong>of</strong> <strong>the</strong> lithosphere, we now<br />

consider thinner pieces <strong>of</strong> lithosphere. In Figure 2.8 <strong>the</strong> areas formerly making up<br />

<strong>the</strong> East European platform and <strong>the</strong> Siberian craton now have a thickness <strong>of</strong> 50<br />

km, while <strong>the</strong> rest <strong>of</strong> <strong>the</strong> plate is 100 km thick. (The 40 km increase in lithosphere<br />

thickness results in <strong>the</strong> stress levels in Figure 2.8 being generally lower than in <strong>the</strong><br />

preceding models). The regions that displayed lower stress intensities in Figure 2.6<br />

now have values higher than <strong>the</strong> surroundings. As before, <strong>the</strong> thickness variation<br />

in <strong>the</strong> area <strong>of</strong> <strong>the</strong> Siberian craton does not appear in <strong>the</strong> diagram showing stress<br />

directions, but <strong>the</strong> area <strong>of</strong> <strong>the</strong> East European platform tends to deflect stresses.<br />

Still, <strong>the</strong> stress pattern resembles <strong>the</strong> model without lateral structure in Figure 2.4<br />

more than <strong>the</strong> one with <strong>the</strong> cratons in Figure 2.6. The main effect <strong>of</strong> <strong>the</strong> thinner<br />

lithosphere in Figure 2.9 is a less abrupt change from horizontal to vertical stress<br />

orientations in <strong>the</strong> area above and below its left edge. (’Horizontal’ and ’vertical’<br />

refer to alignment along <strong>the</strong> long and short edge <strong>of</strong> <strong>the</strong> rectangle, respectively.)<br />

In a fur<strong>the</strong>r model we tested <strong>the</strong> effect <strong>of</strong> both increasing <strong>the</strong> thickness <strong>of</strong> a piece <strong>of</strong><br />

lithosphere and raising <strong>the</strong> as<strong>the</strong>nospheric drag forces beneath it. Keeping in mind<br />

that Gölke and Coblentz [15] fix <strong>the</strong> eastern edge <strong>of</strong> <strong>the</strong>ir model <strong>of</strong> Eurasia, we are<br />

especially interested if raising <strong>the</strong> drag forces below <strong>the</strong> cratons might anchor <strong>the</strong>m<br />

or induce jumps in <strong>the</strong> displacement field at <strong>the</strong>ir edges. The boundary constraints<br />

<strong>of</strong> <strong>the</strong> model shown in Figure 2.9 are identical with those <strong>of</strong> dsm elasticrat in Figure<br />

2.6 and <strong>the</strong> absence <strong>of</strong> discontinuities at <strong>the</strong> craton’s edges after a period <strong>of</strong> 32000<br />

years suggests that it would be more realistic not to pin <strong>the</strong> eastern margin <strong>of</strong> <strong>the</strong><br />

finite element mesh when modeling <strong>the</strong> stress field in Europe.


18 CHAPTER 2. THE RECTANGULAR MODELS<br />

If one wishes to make a more detailed model <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate to <strong>the</strong> west <strong>of</strong><br />

<strong>the</strong> Urals, one should begin by determining <strong>the</strong> stresses along <strong>the</strong> foreseen cut using<br />

a model <strong>of</strong> <strong>the</strong> whole <strong>Eurasian</strong> plate and <strong>the</strong>n apply <strong>the</strong>se stresses as boundary<br />

constraints on <strong>the</strong> eastern border <strong>of</strong> <strong>the</strong> refined model.<br />

As a final comment it should be emphasized that, although <strong>the</strong> conclusions <strong>of</strong> this<br />

section were reached assuming a purely elastic lithosphere, <strong>the</strong> observations remain<br />

correct in <strong>the</strong> viscoelastic, elastoplastic and elastoviscoplastic rheologies discussed<br />

in <strong>the</strong> upcoming section.<br />

2.4 The effect <strong>of</strong> different rheologies on <strong>the</strong> stress<br />

field<br />

To date <strong>the</strong> majority <strong>of</strong> <strong>the</strong> models <strong>of</strong> <strong>the</strong> Earth’s lithosphere published in scientific<br />

literature have employed ei<strong>the</strong>r a purely elastic (e.g., Gölke and Coblentz [15],<br />

Grünthal and Stromeyer [16], Lithgow-Bertelloni and Guynn [22]) or viscous (e.g.,<br />

Bird [4], Lithgow-Bertelloni and Richards [23], Conrad and Lithgow-Bertelloni [8])<br />

material behaviour. In <strong>the</strong> previous section I presented a model for a rectangular<br />

piece <strong>of</strong> purely elastic lithosphere that was run for approximately 32000 years<br />

and contains two regions <strong>of</strong> thicker continental crust, representing <strong>the</strong> East European<br />

platform and <strong>the</strong> Siberian craton (Figure 2.6). This section deals with three<br />

additional rheologies, namely viscoelasticity, plasticity and elastoviscoplasticity to<br />

see how <strong>the</strong>se material properties modify <strong>the</strong> stress pattern obtained in <strong>the</strong> purely<br />

elastic case by applying ridge push and continental margin forces, as well as trench<br />

suction and continental collision along <strong>the</strong> boundaries.<br />

First we will consider a lithosphere under uniaxial stress that behaves like a<br />

viscoelastic (Maxwell) body with <strong>the</strong> rheological equation (c.f. Turcotte and Schubert<br />

[33])<br />

˙ɛij = ˙σij<br />

E<br />

σij<br />

+ , (2.4)<br />

2η<br />

where E is Young’s modulus and η is <strong>the</strong> viscosity 9 . Under constant strain <strong>the</strong><br />

9 In an elastic (or Hooke) body stress is proportional to strain<br />

σij = Eɛij, (2.5)<br />

while in a linearly viscous (or Newton) body stress is proportional to <strong>the</strong> strain rate:<br />

σij = 2η ˙ɛij. (2.6)<br />

By taking <strong>the</strong> derivative <strong>of</strong> Equation 2.5 with respect to time and combining it with 2.6 one gets<br />

Equation 2.4


2.4. THE EFFECT OF DIFFERENT RHEOLOGIES 19<br />

Figure 2.10: Stress pattern for <strong>the</strong> model dsm viscoelasticrat, in which <strong>the</strong> lithosphere behaves<br />

like a Maxwell body with a viscosity <strong>of</strong> ηLIT H = 10 23 P a · s.<br />

solution <strong>of</strong> equation 2.4 becomes<br />

σij = σij,init · exp<br />

<br />

− E<br />

<br />

· t , (2.7)<br />

2η<br />

where σij,init is <strong>the</strong> initial stress. Equation 2.7 defines <strong>the</strong> so called Maxwell time<br />

τM = η<br />

, above which <strong>the</strong> viscous effects in a Maxwell body begin to dominate <strong>the</strong><br />

E<br />

elastic ones. The Maxwell time is important for our calculations, since we have to<br />

know for how long our models should be run to investigate <strong>the</strong> effects <strong>of</strong> viscosity on<br />

<strong>the</strong> stress field. Using <strong>the</strong> common values for Young’s modulus (E = 1011 Pa) and<br />

viscosity (η = 1023 P a · s) in <strong>the</strong> lithosphere to estimate <strong>the</strong> Maxwell time one finds<br />

it to be <strong>of</strong> <strong>the</strong> order <strong>of</strong> 1012 seconds or around 32000 years (i.e. <strong>the</strong> computation<br />

time <strong>of</strong> <strong>the</strong> purely elastic models in <strong>the</strong> preceding section) and I have raised <strong>the</strong><br />

computation time to 1.5 × 1013 seconds (approximately 1.5 million years) for <strong>the</strong><br />

models in this section.<br />

The stress intensity distribution and <strong>the</strong> stress directions for a viscoelastic rheology<br />

are given in Figure 2.10 (apart from rheology <strong>the</strong> same input parameters were<br />

used as in Figure 2.6). A comparison <strong>of</strong> <strong>the</strong>se two figures clearly shows that both


20 CHAPTER 2. THE RECTANGULAR MODELS<br />

Figure 2.11: Stress pattern for <strong>the</strong> model dsm elastoplasticrat, which has a plastic rheology with<br />

a yield stress <strong>of</strong> 100 MPa.<br />

stress levels and orientations are very similar. According to Equation 2.7 Maxwell<br />

bodies exhibit exponential stress relaxation, with <strong>the</strong> relaxation time τM. Once <strong>the</strong><br />

viscoelastic plate has reached a steady state - as we expect it to do after running <strong>the</strong><br />

model for 15 Maxwell times - stresses relax at <strong>the</strong> same rate as new ones build up<br />

due to <strong>the</strong> flowing material, so stress values should no longer change. Fur<strong>the</strong>rmore,<br />

because viscoelastic bodies behave elastically on Maxwell-time scales <strong>the</strong> residual<br />

stress levels should be those <strong>of</strong> a purely elastic model that is already at equilibrium<br />

at <strong>the</strong> Maxwell time. The features <strong>of</strong> <strong>the</strong> stress field in Figure 2.10 are thus<br />

consistent with viscoelastic lithosphere in a steady state.<br />

In keeping with <strong>the</strong>se considerations we do not anticipate any changes in <strong>the</strong><br />

stress field <strong>of</strong> <strong>the</strong> purely elastic plate after 32000 years, even if <strong>the</strong> forces were<br />

applied for 1.5 million years. A closer look at Figure 2.11 appears to reveal a flaw<br />

in this reasoning. It depicts <strong>the</strong> stress pattern for a rectangle <strong>of</strong> material with a<br />

yield threshold <strong>of</strong> 100 MPa. Since yielding is limited to small areas in three <strong>of</strong> <strong>the</strong><br />

four corners, <strong>the</strong> largest part <strong>of</strong> <strong>the</strong> plate should have behaved purely elastically and<br />

thus look exactly like in Figure 2.6. Yet <strong>the</strong> stress intensity over much <strong>of</strong> <strong>the</strong> plate


2.4. THE EFFECT OF DIFFERENT RHEOLOGIES 21<br />

Figure 2.12: Stress pattern for <strong>the</strong> model dsm elastoviscoplasticrat, which has a viscoelastic<br />

rheology with a yield stress <strong>of</strong> 100 MPa.<br />

has reached ∼50 MPa, 30 MPa higher than in Figure 2.6. This is most likely caused<br />

by <strong>the</strong> dashpots. Since <strong>the</strong>y add a viscous component to <strong>the</strong> a priori purely elastic<br />

piece <strong>of</strong> lithosphere, <strong>the</strong>y delay <strong>the</strong> transmission <strong>of</strong> stresses into <strong>the</strong> plate’s interior<br />

and cause stress values to continue growing. There is fur<strong>the</strong>r evidence to support<br />

this hypo<strong>the</strong>sis. First <strong>of</strong> all, in a snapshot <strong>of</strong> <strong>the</strong> elastoplastic model in Figure 2.11<br />

after 32000 years stress values are still <strong>the</strong> same as in Figure 2.6. Fur<strong>the</strong>rmore,<br />

increasing <strong>the</strong> frictional coefficients <strong>of</strong> <strong>the</strong> dashpots produces even more extensive<br />

areas <strong>of</strong> relaxed lithosphere in <strong>the</strong> plate’s interior, compatible with a still slower<br />

transmission <strong>of</strong> <strong>the</strong> stresses towards <strong>the</strong> inside.<br />

While <strong>the</strong> characteristics <strong>of</strong> <strong>the</strong> dashpots probably influence stress levels, <strong>the</strong>y do<br />

not alter stress directions significantly. One <strong>of</strong> <strong>the</strong> main issues <strong>of</strong> my <strong>the</strong>sis was to<br />

find how one might alter <strong>the</strong> stress field in Europe, which in purely elastic models is<br />

dominated by Mid-Atlantic ridge push and continental collision in <strong>the</strong> Himalayas,<br />

by prescribing different rheological properties to <strong>the</strong> lithosphere. We supected that<br />

plasticity might <strong>the</strong> key to <strong>the</strong> question. Yielding should localize all deformation<br />

and, because only <strong>the</strong> yield stress can be propagated beyond <strong>the</strong> area <strong>of</strong> highest<br />

deformation, reduce stress levels in <strong>the</strong> remaining plate. In this manner <strong>the</strong> rest <strong>of</strong>


22 CHAPTER 2. THE RECTANGULAR MODELS<br />

Figure 2.13: Stress pattern for <strong>the</strong> model ksm elastoplasticrat, which has <strong>the</strong> same plastic rheology<br />

as <strong>the</strong> rectangle in Figure 2.11, but which uses kinematic ra<strong>the</strong>r than dynamic boundary<br />

conditions.<br />

Eurasia could be decoupled to a certain extent from <strong>the</strong> stresses generated by <strong>the</strong><br />

indentation <strong>of</strong> <strong>the</strong> Indian plate. This would allow <strong>the</strong> weaker collisional forces from<br />

<strong>the</strong> convergence <strong>of</strong> Africa on Europe to gain importance in influencing <strong>the</strong> orientation<br />

European stress field.<br />

So far <strong>the</strong>re is no grounds to believe that more complex rheologies resolve <strong>the</strong> problem.<br />

From Figures 2.6, 2.10 and 2.11, <strong>the</strong> stress in <strong>the</strong> left quarter <strong>of</strong> <strong>the</strong> rectangle,<br />

a region we can identify with Europe in our simple model, is seen to be horizontal.<br />

There is thus no indication whatsoever <strong>of</strong> <strong>the</strong> African collision acting on <strong>the</strong> lower<br />

edge in that portion <strong>of</strong> <strong>the</strong> lithosphere. As illustrated in Figure 2.12 <strong>the</strong> situation is<br />

no different in an elastoviscoplastic rheology; indeed, one can barely distinguish any<br />

differences in <strong>the</strong> stress patterns <strong>of</strong> <strong>the</strong> elastic, viscoelastic and elastoviscoplastic<br />

rectangles.<br />

For comparison, consider a kinematically constrained rectangle again. The model<br />

<strong>of</strong> Figure 2.13 has <strong>the</strong> same plastic rheology as <strong>the</strong> one in Figure 2.11, but <strong>the</strong><br />

boundary forces for trench suction and continental collision have been replaced by


2.4. THE EFFECT OF DIFFERENT RHEOLOGIES 23<br />

Figure 2.14: Stress pattern for <strong>the</strong> model ydsm elastoplasticrat0. The collisional forces have<br />

been chosen such that <strong>the</strong>y lead to roughly <strong>the</strong> same extent <strong>of</strong> yielding as in Figure 2.13.<br />

kinematic constraints. Now <strong>the</strong> formerly horizontal stress orientations are replaced<br />

by a field at roughly 45 ◦ , consistent with <strong>the</strong> large scale stress orientations in Europe<br />

which are aligned NW-SE, apparently reflecting <strong>the</strong> influence <strong>of</strong> both ridge push and<br />

African convergence. Could this <strong>the</strong>n imply that we have been applying <strong>the</strong> wrong<br />

forces in <strong>the</strong> dynamic models?<br />

In <strong>the</strong> section on kinematic and dynamic boundary conditions we found that velocities<br />

lead to (too) high boundary loads which, in <strong>the</strong> purely elastic models, created<br />

unrealistically high stress intensities throughout <strong>the</strong> plate. By prescribing a yield<br />

stress in <strong>the</strong> model <strong>of</strong> Figure 2.13, <strong>the</strong> deformation is localized and stress levels<br />

outside <strong>the</strong> area <strong>of</strong> yielding (in grey, in Figures 2.13 and 2.14) are as low as in<br />

<strong>the</strong> models with dynamic constraints, with <strong>the</strong> important difference that <strong>the</strong> stress<br />

patterns seem to be closer to reality than in <strong>the</strong> dynamic models.<br />

In view <strong>of</strong> this discovery we now are led to <strong>the</strong> conclusion that plasticity is <strong>the</strong><br />

key to creating a realistic stress map <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate, because it allows <strong>the</strong><br />

application <strong>of</strong> sufficiently large forces without raising <strong>the</strong> stress levels in <strong>the</strong> interior<br />

<strong>of</strong> <strong>the</strong> plate to unrealistically high values. Figure 2.14 shows a dynamical model


24 CHAPTER 2. THE RECTANGULAR MODELS<br />

in which <strong>the</strong> boundary forces have been chosen such that <strong>the</strong> characteristics <strong>of</strong> <strong>the</strong><br />

stress field are quite similar to those in <strong>the</strong> preceding kinematically constrained<br />

model. To achieve this, forces with a magnitude <strong>of</strong> 1.5 × 10 18 N, 2 × 10 18 N and<br />

3 × 10 18 N, for <strong>the</strong> African, Arabian and Indian collision, respectively, were applied.<br />

Atlantic ridge push on <strong>the</strong> o<strong>the</strong>r hand has a magnitude <strong>of</strong> 1.34 × 10 18 N. On page 8<br />

it was mentioned that previous studies predict collisional forces to be about half as<br />

large as ridge push, while in our case <strong>the</strong>y are equal or larger. How can <strong>the</strong>se two<br />

observations be reconciled? We believe that <strong>the</strong> forces listed in Table 2.1 are <strong>the</strong><br />

stresses that are transmitted to <strong>the</strong> interior <strong>of</strong> <strong>the</strong> plate at <strong>the</strong> border <strong>of</strong> <strong>the</strong> area<br />

<strong>of</strong> yield. Since perfectly plastic materials do not <strong>of</strong>fer any resistance to deformation<br />

once <strong>the</strong>ir yield limit has been attained, this yield stress acts on <strong>the</strong> border and <strong>the</strong><br />

rest <strong>of</strong> <strong>the</strong> plate and by doing so contributes to its torque balance. We assume that<br />

this yield force, ra<strong>the</strong>r than <strong>the</strong> actual collisional forces, was used for <strong>the</strong> calculation<br />

<strong>of</strong> <strong>the</strong> relative force magnitudes in <strong>the</strong> results listed in Table 2.1.<br />

With <strong>the</strong> insights on rheology, strength variations and forces gained in <strong>the</strong> models<br />

presented in this section we will move on to consider <strong>the</strong> ’real’ <strong>Eurasian</strong> plate now.<br />

In <strong>the</strong> next chapter we investigate whe<strong>the</strong>r <strong>the</strong> promising results live up to our<br />

expectations and are able to produce a model <strong>of</strong> Eurasia that satisfactorily matches<br />

actual data.


MOTIONS UP IN THE HEAVENS ARE ORDERLY, PRECISE, REGULAR AND MATHEMATICAL, THOSE<br />

DOWN ON EARTH MESSY AND IRREGULAR AND CAN BE DESCRIBED ONLY QUALITATIVELY...<br />

The Aristotelian view <strong>of</strong> <strong>the</strong> universe [9]<br />

Chapter 3<br />

The model <strong>of</strong> Eurasia<br />

Before discussing more realistic models <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate than those treated in<br />

Chapter 2, we should sumarize which <strong>of</strong> <strong>the</strong> features observed in our best rectangular<br />

model (Figure 2.14) already agree with actual measurements.<br />

Introducing plasticity causes stress directions to be determined by <strong>the</strong> nearest sources<br />

<strong>of</strong> stress; in <strong>the</strong> portion <strong>of</strong> <strong>the</strong> rectangle that we identify with Europe this leads to a<br />

field matching <strong>the</strong> NW-SE maximal horizontal stress in Western Europe as given by<br />

The 2003 release <strong>of</strong> <strong>the</strong> World Stress Map [30]. The area east <strong>of</strong> <strong>the</strong> Indian continental<br />

collision displays extensional stresses which are consistent with <strong>the</strong> extrusion<br />

<strong>of</strong> large blocks <strong>of</strong> lithosphere along <strong>the</strong> major Chinese strike-slip fault systems, such<br />

as <strong>the</strong> Red River fault and <strong>the</strong> Altyn Tagh fault that is located on <strong>the</strong> border between<br />

<strong>the</strong> Tibetan <strong>Plate</strong>au and <strong>the</strong> Tarim Basin. In fact, even in our rectangle we<br />

predict strike-slip faulting along <strong>the</strong> band where both compressional (in red) and<br />

extensional (in yellow) stresses prevail. It lies in a line linking <strong>the</strong> right edge <strong>of</strong> <strong>the</strong><br />

Indian indenter to <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> subduction zone “fur<strong>the</strong>r north” and thus<br />

agrees quite well with <strong>the</strong> orientation <strong>of</strong> a region <strong>of</strong> high seismic activity (c.f. The<br />

Global Seismic Hazard Map [13]) running from Afghanistan north to Lake Baikal.<br />

Finally, <strong>the</strong> models predict reasonable stress values (between 20 and 60 MPa) in<br />

much <strong>of</strong> <strong>the</strong> plate’s interior, and <strong>the</strong> displacement velocities (between 1.9 cm and 8<br />

mm a year) also coincide well with <strong>the</strong> range <strong>of</strong> velocities measured in Eurasia.<br />

The results in a model using <strong>the</strong> <strong>Eurasian</strong> plate’s real outline will have to stand<br />

up to more trying tests than <strong>the</strong> rectangle, which allows qualitative comparisons at<br />

best. In <strong>the</strong> new setting we will be able to compare our findings with actual data<br />

for stress, strain rate and deformation velocities. The prerequisites for achieving <strong>the</strong><br />

higher level <strong>of</strong> accuracy made necessary by <strong>the</strong> increased complexity <strong>of</strong> <strong>the</strong> problem<br />

are: (1) an improved understanding <strong>of</strong> <strong>the</strong> boundary conditions in <strong>the</strong> areas <strong>of</strong><br />

continental collision and subduction, and probably also (2) a rheological model that<br />

is closer to reality than perfect plasticity.<br />

25


26 CHAPTER 3. THE MODEL OF EURASIA<br />

EEP<br />

INT<br />

SC<br />

Figure 3.1: The finite element mesh used for <strong>the</strong> models <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate (same projection<br />

as in Figure 2.1). Areas <strong>of</strong> thicker continental lithosphere are marked in blue (150 km) and green<br />

(200 km): EEP - East European Platform; INT - stable platform in central Eurasia; SC - Siberian<br />

craton.<br />

3.1 The properties <strong>of</strong> <strong>the</strong> ABAQUS model<br />

The whole mesh as shown in Figure 3.1 has 4013 elements with an average area <strong>of</strong><br />

1.719 × 10 10 m 2 . The nodes are separated by a distance <strong>of</strong> approx. 130 km which,<br />

at <strong>the</strong> latitude <strong>of</strong> <strong>the</strong> projection’s origin, corresponds to 2.1 ◦ . Building <strong>the</strong> mesh for<br />

a complex outline like <strong>the</strong> <strong>Eurasian</strong> plate is trickier than in a simple and symmetric<br />

shape like a rectangle, because all elements should be as regular as possible for a<br />

stable and fast finite element analysis. To construct a mesh fulfilling this requirement<br />

we used <strong>the</strong> mesh building program PATRAN [25], which generates a file that can<br />

be read by ABAQUS. As in <strong>the</strong> rectangular models we employ four sided continuum<br />

elements with plane stress characteristics.<br />

The <strong>Eurasian</strong> plate has an area <strong>of</strong> 69×10 6 km 2 <strong>of</strong> which 51×10 6 km 2 are continental<br />

lithosphere. Due to <strong>the</strong> limitations <strong>of</strong> <strong>the</strong> two-dimensional elements (c.f. footnote<br />

on page 7) that keep us from applying ridge push as a distributed load we will


3.2. FORCES REVISITED 27<br />

restrict our analysis to continental Eurasia. This leaves us with 3723 elements with<br />

a thickness <strong>of</strong> 200 km in <strong>the</strong> East European platform and <strong>the</strong> Siberian craton, and<br />

thicknesses <strong>of</strong> 150 km in <strong>the</strong> quiescent area between <strong>the</strong>se two continental shields<br />

and 100 km in <strong>the</strong> rest <strong>of</strong> <strong>the</strong> plate.<br />

3.2 Forces revisited<br />

In <strong>the</strong> rectangular models we implemented four forces, namely: basal drag, ridge<br />

push, trench suction and continental collision 1 . The first two <strong>of</strong> <strong>the</strong>se we will apply<br />

in <strong>the</strong> same conceptual way as before, albeit in a slightly refined manner. Thus we<br />

refer to <strong>the</strong> Digital age map <strong>of</strong> <strong>the</strong> ocean floor [27] for detailed data on <strong>the</strong> age <strong>of</strong><br />

<strong>the</strong> ocean floor along Eurasia’s Atlantic and Arctic continental margins to calculate<br />

a realistic ridge push field. From it we <strong>the</strong>n subtract <strong>the</strong> continental margin force<br />

(which we assume to be <strong>of</strong> <strong>the</strong> same magnitude along all relevant border segments)<br />

to get <strong>the</strong> boundary loads along <strong>the</strong> perimeter all <strong>the</strong> way from <strong>the</strong> Algarve to <strong>the</strong><br />

Laptev Sea. We no longer apply basal drag as a uniform force but, accounting for<br />

<strong>the</strong> depth dependence <strong>of</strong> viscosity, calculate it individually for <strong>the</strong> cratons, continental<br />

lithosphere and <strong>the</strong> region between <strong>the</strong> cratons. The values <strong>of</strong> shear stress acting<br />

on <strong>the</strong> base <strong>of</strong> <strong>the</strong> lithosphere <strong>the</strong>n range from 2.42 MPa underneath <strong>the</strong> continent<br />

to 34.04 MPa beneath <strong>the</strong> cratons.<br />

As far as trench suction and continental collision are concerned, our lack <strong>of</strong> understanding<br />

<strong>of</strong> <strong>the</strong> underlying physical processes has so far prevented quantitative<br />

predictions <strong>of</strong> <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> forces involved. In <strong>the</strong> following paragraphs I<br />

will discuss concepts that could help estimate <strong>the</strong> values <strong>of</strong> <strong>the</strong> boundary constraints<br />

in <strong>the</strong>se environments.<br />

As continental collision is due to interaction between <strong>the</strong> plates involved, it should<br />

act on both plates with equal strength. This does not mean that <strong>the</strong> same amount<br />

<strong>of</strong> deformation need occur on each <strong>of</strong> <strong>the</strong>m; if one <strong>of</strong> <strong>the</strong> plates participating in <strong>the</strong><br />

event is stronger - as for instance India probably is, because <strong>of</strong> <strong>the</strong> presence <strong>of</strong> a<br />

craton - it will likely suffer less deformation.<br />

Forsyth and Uyeda [11] argue that during collision a higher strain rate does not<br />

increase <strong>the</strong> stress, but merely reduces <strong>the</strong> length <strong>of</strong> time required to reach <strong>the</strong><br />

level <strong>of</strong> stress at which brittle failure or plastic yielding occurs. Thus, <strong>the</strong>y take<br />

<strong>the</strong> average stress over a period <strong>of</strong> time to be independent <strong>of</strong> <strong>the</strong> relative motion<br />

at <strong>the</strong> plate boundary. I believe this point <strong>of</strong> view can be reconciled with our<br />

assumption that, for <strong>the</strong> reasons set forth on page 24, <strong>the</strong> stress acting on <strong>the</strong><br />

plates’ interior is <strong>the</strong> yield stress. However, <strong>the</strong> varying sizes <strong>of</strong> Eurasia’s orogens is<br />

a strong indication that <strong>the</strong> relative rate <strong>of</strong> convergence never<strong>the</strong>less is an important<br />

1 We continue to neglect <strong>the</strong> resistance along strike-slip boundary segments (which also includes<br />

transform fault resistance on <strong>the</strong> strike-slip faults joining different segments <strong>of</strong> <strong>the</strong> mid-oceanic<br />

ridges).


28 CHAPTER 3. THE MODEL OF EURASIA<br />

factor. India, Arabia and Africa are, on average, moving at speeds <strong>of</strong> 4.6 cm, 2.9<br />

cm and 0.6 cm with respect to Eurasia today and <strong>the</strong> mountainous regions along<br />

<strong>the</strong>ir borders cover a surface <strong>of</strong> 19.2 × 10 5 km 2 (Himalayas and Pamir), 7.2 × 10 5<br />

km 2 (Zagros) and 1.2 × 10 5 km 2 (Alps). Consequently it seems appropriate to scale<br />

<strong>the</strong> collisional forces that induce yielding with velocities.<br />

Laboratory results suggest that rocks under lithospheric conditions are most likely<br />

to obey a power law dislocation creep relationship unless <strong>the</strong>y behave in a brittle<br />

manner (Brace and Kohlstedt [6]):<br />

˙ɛ = Aτ n exp(−Q/RT ), (3.1)<br />

where ˙ɛ is <strong>the</strong> strain rate, τ is <strong>the</strong> deviatoric stress, T is <strong>the</strong> absolute temperature,<br />

R <strong>the</strong> universal gas constant and A, Q and n are material-dependent constants.<br />

The activation energy, Q, ranges from ca. 500 kJ mol −1 for dry olivine to ca. 160<br />

kJ mol −1 for quartz and <strong>the</strong> power-law exponent, n, is in <strong>the</strong> range 3-5 (Kirby<br />

and Kronenburg [20]). When relating stresses to <strong>the</strong> velocities <strong>of</strong> <strong>the</strong> neighbouring<br />

plates we thus assume that <strong>the</strong> following relation holds (strain rate being <strong>the</strong> spatial<br />

gradient <strong>of</strong> velocity):<br />

σ = σref 3<br />

v<br />

vref<br />

. (3.2)<br />

The value for <strong>the</strong> reference stress σref can be gained by considering <strong>the</strong> European<br />

stress field. Judging by its orientation it is governed by ridge push and a force <strong>of</strong><br />

about <strong>the</strong> same magnitude acting along its sou<strong>the</strong>rn perimeter. Hence, we take σref<br />

to be equal to <strong>the</strong> value <strong>of</strong> ridge push transmitted by <strong>the</strong> piece <strong>of</strong> oldest oceanic<br />

crust along Europe’s western continental margin. The reference velocity vref going<br />

with σref corresponds to <strong>the</strong> convergence rate <strong>of</strong> <strong>the</strong> African plate on Europe in <strong>the</strong><br />

Central Mediterranean Sea. Having fixed <strong>the</strong>se values we can deduce <strong>the</strong> stresses<br />

along all collisional borders by scaling <strong>the</strong>m with <strong>the</strong> local velocities according to<br />

Equation 3.2. The actual forces that get applied parallel to <strong>the</strong> velocity at each<br />

boundary node <strong>the</strong>n are computed by multiplying <strong>the</strong> stresses by <strong>the</strong> thickness <strong>of</strong><br />

<strong>the</strong> lithosphere and <strong>the</strong> distance between <strong>the</strong> individual nodes.<br />

The source <strong>of</strong> trench suction is a secondary hydrodynamic flow induced in <strong>the</strong><br />

upper mantle above a sinking slab, and this flow exerts shear tractions at <strong>the</strong> base<br />

<strong>of</strong> nearby plates. Since <strong>the</strong> subduction <strong>of</strong> <strong>the</strong> slab gives rise to <strong>the</strong> secondary flow it<br />

appears reasonable to assume that <strong>the</strong> forces acting on <strong>the</strong> overriding plate scale with<br />

its rate <strong>of</strong> descent. The basal drag forces mentioned at <strong>the</strong> beginning <strong>of</strong> this section<br />

are proportional to <strong>the</strong> velocity at which <strong>the</strong> plate drifts over <strong>the</strong> upper mantle.<br />

Whe<strong>the</strong>r <strong>the</strong> mantle moves with respect to <strong>the</strong> lithosphere (as is <strong>the</strong> case in trench<br />

suction) or vice versa (as happens in basal drag) is merely a question <strong>of</strong> reference<br />

frame and thus we expect trench suction to vary linearly with <strong>the</strong> velocity <strong>of</strong> <strong>the</strong>


3.3. RHEOLOGY 29<br />

subducting plate at <strong>the</strong> trench. Such a linear dependence is also found by Scholz<br />

and Campos [32], although it is not a dependence on <strong>the</strong> velocity <strong>of</strong> <strong>the</strong> subducting<br />

plate alone, but a combination <strong>of</strong> overriding and subducting plate velocities and<br />

trench roll back speed instead.<br />

Just as in <strong>the</strong> case <strong>of</strong> continental collision we still need a reference force for <strong>the</strong><br />

scaling relationship. The estimates listed in Table 2.1 on page 8 allow for values<br />

<strong>of</strong> FSU between 0.8 × 10 12 Nm −1 and 6 × 10 12 Nm −1 . Our own estimate given in<br />

equation 2.2 lies about half way between <strong>the</strong>se two values and logically <strong>of</strong>fers itself as<br />

a compromise. We chose it to correspond to <strong>the</strong> mean subduction velocity along all<br />

<strong>Eurasian</strong> trenches in <strong>the</strong> Far East and apply <strong>the</strong> forces so calculated perpendicular<br />

to <strong>the</strong> plate boundary in <strong>the</strong> corresponding regions.<br />

3.3 Rheology<br />

The perfect plasticity used in <strong>the</strong> rectangular models is an idealisation. In reality,<br />

two mechanisms, diffusion creep and dislocation creep, are active during <strong>the</strong> purely<br />

elastic phase below <strong>the</strong> yield stress <strong>of</strong> 100 MPa. Diffusion creep only occurs if <strong>the</strong><br />

rocks are made up <strong>of</strong> very small grains or at stresses significantly lower than those in<br />

our models. We can <strong>the</strong>refore ignore this kind <strong>of</strong> creep and restrict ourselves to <strong>the</strong><br />

implementation <strong>of</strong> dislocation creep. Laboratory results suggest that rocks under<br />

lithospheric conditions are most likely to obey a power law creep as in Equation 3.1<br />

(commonly with n = 3), provided <strong>the</strong>y do not behave in a brittle manner.<br />

Dislocation creep breaks down when stress and strain rates are increased sufficiently<br />

so that dislocations start to appear in considerable number inside <strong>the</strong> individual<br />

grains <strong>of</strong> <strong>the</strong> material, <strong>the</strong>reby ending <strong>the</strong> domain <strong>of</strong> low power viscous creep. At<br />

stresses higher than <strong>the</strong> yield limit σY <strong>the</strong> activation <strong>of</strong> different dominant glide<br />

planes leads to an exponential dependence <strong>of</strong> strain rate on stress known as Peierl’s<br />

creep or low temperature plasticity (see Regenauer-Lieb and Yuen [29]):<br />

˙ɛ ∝ exp(σ) or ˙ɛ ∝ sinh(σ). (3.3)<br />

In purely fluid dynamic models <strong>the</strong> laws describing this regime are frequently<br />

approximated by a power law dislocation glide <strong>of</strong> <strong>the</strong> form:<br />

˙ɛ = ˙ɛY + A · (σ − σY ) n , (3.4)<br />

where ˙ɛY is <strong>the</strong> value <strong>of</strong> <strong>the</strong> strain rate at σY , attained in <strong>the</strong> rheological regime<br />

valid in <strong>the</strong> stress range below <strong>the</strong> yield limit. The larger <strong>the</strong> exponent n is, <strong>the</strong><br />

more <strong>the</strong> material behaves like a perfectly plastic body. Thus in our elastoplastic<br />

models we calculate <strong>the</strong> response <strong>of</strong> a solid with a very high value <strong>of</strong> n at stresses


30 CHAPTER 3. THE MODEL OF EURASIA<br />

larger than σY and for which viscous creep below this limit is neglected. Strictly<br />

taken this is probably only appropriate if <strong>the</strong> model is run for less than <strong>the</strong> Maxwell<br />

time introduced in Equation 2.7 <strong>of</strong> <strong>the</strong> preceding chapter, and not for 1.5 million<br />

years as we do.<br />

Power law creep and plasticity are at opposite ends <strong>of</strong> <strong>the</strong> rheological spectrum<br />

in that plasticity can be described by a viscosity that jumps from infinity below σY<br />

to zero after yielding whereas <strong>the</strong> nonlinear (or effective) viscosity η = σ/2˙ɛ varies<br />

continuously between <strong>the</strong>se two values.<br />

In <strong>the</strong> next section we present two models <strong>of</strong> Eurasia, one using a viscoelastic power<br />

law rheology and <strong>the</strong> o<strong>the</strong>r using plasticity.<br />

3.4 A comparison with measurements<br />

To begin with we investigate how well a model with a perfectly elastoplastic rheology<br />

matches actual data for stress directions, displacement velocities and strain rates<br />

within Eurasia.<br />

The stress field <strong>of</strong> continental Eurasia, as depicted in Figures 3.2 to 3.4, results<br />

from <strong>the</strong> following boundary conditions:<br />

1. ridge push counteracted by continental margin forces along <strong>the</strong> Atlantic and<br />

Arctic continental margin <strong>of</strong> Eurasia; <strong>the</strong> ridge push is largest <strong>of</strong>f <strong>the</strong> coast<br />

<strong>of</strong> Portugal, where <strong>the</strong> oceanic lithosphere is about 100 million years old and<br />

smallest in <strong>the</strong> Laptev Sea where it is only half as old,<br />

2. forces due to continental collision with <strong>the</strong> African, Arabian and Indian plates<br />

along <strong>the</strong> sou<strong>the</strong>rn border <strong>of</strong> Eurasia and <strong>the</strong> North American plate in Siberia<br />

(scaled with relative velocities according to Equation 3.2),<br />

3. trench suction scaled linearly with <strong>the</strong> convergence rate <strong>of</strong> <strong>the</strong> neighbouring<br />

slab being subducted in all Far Eastern subduction zones and in <strong>the</strong> Aegean<br />

Sea where <strong>the</strong> African plate is diving beneath Eurasia,<br />

4. strike-slip segments are left free; <strong>the</strong>y include<br />

(a) <strong>the</strong> segment running from <strong>the</strong> coast <strong>of</strong> Burma to <strong>the</strong> Kingdom <strong>of</strong> Bhutan,<br />

(b) <strong>the</strong> segment joining <strong>the</strong> coast <strong>of</strong> Pakistan to Nepal,<br />

(c) <strong>the</strong> segment between <strong>the</strong> Strait <strong>of</strong> Gibraltar and <strong>the</strong> Atlantic continental<br />

margin,<br />

(d) a short segment in nor<strong>the</strong>rn Siberia


3.4. A COMPARISON WITH MEASUREMENTS 31<br />

Figure 3.2: Stress intensity map for <strong>the</strong> elastoplastic model eurasia27-01-04 6 (von Mises stress,<br />

c.f. Equation 2.3 on page 10). In <strong>the</strong> grey areas <strong>the</strong> yield limit <strong>of</strong> 100 MPa has been exceeded.


32 CHAPTER 3. THE MODEL OF EURASIA<br />

Figure 3.3: Orientation <strong>of</strong> compressional principal stress for <strong>the</strong> model eurasia27-01-04 6.


3.4. A COMPARISON WITH MEASUREMENTS 33<br />

Figure 3.4: Orientation <strong>of</strong> extensional principal stress for <strong>the</strong> model eurasia27-01-04 6.


34 CHAPTER 3. THE MODEL OF EURASIA<br />

5. an unconstrained segment approximately 300 km in length along <strong>the</strong> coast <strong>of</strong><br />

Pakistan where Arabian oceanic lithosphere is being subducted underneath<br />

Eurasia (as mantle flow is probably not important in that area due to <strong>the</strong><br />

surrounding regions undergoing continental collision, we do not apply trench<br />

suction to it).<br />

Looking at Figure 3.2 one can easily recognize <strong>the</strong> two cratons as bluish regions <strong>of</strong><br />

low stress intensity. The two areas in grey mark those regions where <strong>the</strong> yield limit<br />

has been exceeded and orogeny should occur. In <strong>the</strong> case <strong>of</strong> <strong>the</strong> Himalayas <strong>the</strong> extent<br />

<strong>of</strong> this area is consistent with <strong>the</strong> presence <strong>of</strong> actual mountains and, as a fascinating<br />

detail, <strong>the</strong>re is an area within it where <strong>the</strong> lithosphere has not yet yielded and that<br />

one could tentatively identify with <strong>the</strong> Tibetan <strong>Plate</strong>au. This feature could be due<br />

to <strong>the</strong> shape <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> border which does not present <strong>the</strong> indenting Indian<br />

plate with a straight border, but one that has a convex bend to it. To <strong>the</strong> north<br />

<strong>of</strong> <strong>the</strong> boundary with Arabia <strong>the</strong> area <strong>of</strong> yield is far too large compared with real<br />

mountain ranges. Forces along this zone are applied more obliquely to <strong>the</strong> border<br />

<strong>of</strong> <strong>the</strong> plate than in <strong>the</strong> Indian collision and we suspect this induces an unrealistic<br />

reaction <strong>of</strong> <strong>the</strong> boundary elements that seem to deform too easily when sheared.<br />

Since <strong>the</strong> boundary loads for trench suction applied to Sou<strong>the</strong>ast Asia are weaker<br />

than <strong>the</strong> collisional forces by a factor <strong>of</strong> up to three, that area <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate<br />

also has quite low stress levels.<br />

Let us now turn to <strong>the</strong> orientation <strong>of</strong> <strong>the</strong> compressional and extensional stresses<br />

in <strong>the</strong> elastoplastic model, given in Figure 3.3 and 3.4. They can be compared with<br />

<strong>the</strong> maps in Figures 3.5 to 3.7 that show data ga<strong>the</strong>red in <strong>the</strong> framework <strong>of</strong> <strong>the</strong><br />

World Stress Map Project [30].<br />

Beginning in <strong>the</strong> east, Figure 3.5 displays compressional stresses parallel to <strong>the</strong> motion<br />

<strong>of</strong> India in <strong>the</strong> central Himalayas, and stresses that rotate to WNW-ESE in<br />

<strong>the</strong> Hindu-Kush. Our model in Figure 3.3 matches <strong>the</strong>se observations well. To <strong>the</strong><br />

east <strong>of</strong> <strong>the</strong> Himalayas and all <strong>the</strong> way to <strong>the</strong> Japan and Ryukyu Trenches compressional<br />

stresses are aligned roughly E-W. However, as can be seen in Figure 3.4, we<br />

predict extensional stresses to have that orientation and compressional stresses run<br />

N-S. Similarly, our compression is perpendicular to <strong>the</strong> observed directions along<br />

<strong>the</strong> Philippine Trench. Along <strong>the</strong> Java Trench we do predict extensional stresses at<br />

90 ◦ to <strong>the</strong> observed compression, but this is probably irrelevant since our models<br />

show vanishingly small compression in that area. In fact, although our model predicts<br />

that <strong>the</strong> whole <strong>of</strong> Sou<strong>the</strong>astern Asia will be dominated by extension, <strong>the</strong>re is<br />

no indication <strong>of</strong> such a regime in <strong>the</strong> World Stress Map (henceforth referred to as<br />

WSM).<br />

A comparison with Iranian WSM data in Figure 3.6 reveals that <strong>the</strong> orientation<br />

<strong>of</strong> compressional stresses in Figure 3.3 are generally correct in this region. The<br />

prominent extensional stresses along <strong>the</strong> plate boundary from Nepal to Iraq are<br />

probably an artifact <strong>of</strong> <strong>the</strong> too easy deformability <strong>of</strong> <strong>the</strong> elements under shearing<br />

loads, so we need not be concerned that we predict strike slip faulting ra<strong>the</strong>r than


3.4. A COMPARISON WITH MEASUREMENTS 35<br />

60˚<br />

60˚<br />

Method:<br />

focal mechanism<br />

breakouts<br />

80˚<br />

80˚<br />

100˚<br />

drill. induced frac.<br />

60˚ 60˚<br />

borehole slotter<br />

overcoring<br />

hydro. fractures<br />

geol. indicators<br />

Regime:<br />

NF SS TF U<br />

Quality:<br />

A<br />

B<br />

C<br />

! (2003) World Stress Map<br />

100˚<br />

120˚<br />

120˚<br />

140˚<br />

40˚ 40˚<br />

20˚ 20˚<br />

0˚ 0˚<br />

World Stress Map Rel. 2003<br />

Heidelberg Academy <strong>of</strong> Sciences and Humanities<br />

Geophysical Institute, University <strong>of</strong> Karlsruhe<br />

Projection: Mercator<br />

Figure 3.5: World Stress Map data for Asia. The lines indicate <strong>the</strong> direction <strong>of</strong> maximal compression.<br />

Stress indicators in red, green and blue are based on focal mechanism <strong>of</strong> earthquakes: NF<br />

- normal faulting; SS - strike-slip faulting; TF - thrust faulting.<br />

<strong>the</strong> predominantly observed thrust faulting.<br />

In Europe, <strong>the</strong> region to which we wanted to pay special attention, data is abundant<br />

(Figure 3.7). In its western part and in <strong>the</strong> British Isles compressional stresses<br />

trend NW-SE. There is evidence that <strong>the</strong>ir orientation rotates around <strong>the</strong> arc defined<br />

by <strong>the</strong> Alps to NNE-SSW, although this tendency is masked to a certain extent by<br />

140˚


36 CHAPTER 3. THE MODEL OF EURASIA<br />

45˚<br />

45˚<br />

50˚<br />

40˚ 40˚<br />

35˚ 35˚<br />

30˚ 30˚<br />

Method:<br />

focal mechanism<br />

breakouts<br />

drill. induced frac.<br />

borehole slotter<br />

overcoring<br />

hydro. fractures<br />

geol. indicators<br />

Regime:<br />

NF SS TF U<br />

Quality:<br />

25˚ A<br />

25˚<br />

B<br />

C<br />

! (2003) World Stress Map<br />

World Stress Map Rel. 2003<br />

Heidelberg Academy <strong>of</strong> Sciences and Humanities<br />

Geophysical Institute, University <strong>of</strong> Karlsruhe<br />

50˚<br />

55˚<br />

55˚<br />

Figure 3.6: World Stress Map data for Iran.<br />

60˚<br />

60˚<br />

Projection: Mercator<br />

regional patterns which are probably due to topographical effects. In sou<strong>the</strong>astern<br />

Europe stress directions and focal mechanisms reflect <strong>the</strong> motion <strong>of</strong> Turkey towards<br />

Greece and in Scandinavia compression appears to be parallel to ridge push.<br />

Our model contains evidence <strong>of</strong> a rotation <strong>of</strong> stresses into NNE-SSW to <strong>the</strong> east<br />

<strong>of</strong> <strong>the</strong> Alps, albeit only in <strong>the</strong> proximity <strong>of</strong> <strong>the</strong> sou<strong>the</strong>rn border, and we successfully<br />

predict compressional stress directions in <strong>the</strong> interior <strong>of</strong> western Europe as far<br />

north as Great Britain. However along <strong>the</strong> continental margin and in Scandinavia<br />

we fail to match <strong>the</strong> observations. Here <strong>the</strong> predicted compressional stresses are<br />

parallel to <strong>the</strong> continental margin. Although <strong>the</strong> East European platform is in <strong>the</strong><br />

vicinity, we do not believe it to be <strong>the</strong> source <strong>of</strong> <strong>the</strong> phenomenon since we have no<br />

evidence <strong>of</strong> such abrupt stress bending from <strong>the</strong> rectangular models presented in <strong>the</strong><br />

last chapter. As seen in to Figure 3.4, extensional stresses are negligible in Europe,<br />

65˚<br />

65˚


3.4. A COMPARISON WITH MEASUREMENTS 37<br />

Method:<br />

focal mechanism<br />

breakouts<br />

drill. induced frac.<br />

borehole slotter<br />

overcoring<br />

hydro. fractures<br />

geol. indicators<br />

Regime:<br />

NF SS TF U<br />

Quality:<br />

A<br />

B<br />

C<br />

! (2003) World Stress Map<br />

0˚<br />

60˚ 60˚<br />

40˚ 40˚<br />

0˚<br />

World Stress Map Rel. 2003<br />

Heidelberg Academy <strong>of</strong> Sciences and Humanities<br />

Geophysical Institute, University <strong>of</strong> Karlsruhe<br />

20˚<br />

20˚<br />

Figure 3.7: World Stress Map data for Europe.<br />

Projection: Mercator<br />

suggesting that <strong>the</strong> strike-slip faulting observed in central Europe may result from<br />

local effects superimposed on <strong>the</strong> large scale stress distribution. Such a change in<br />

40˚<br />

40˚


38 CHAPTER 3. THE MODEL OF EURASIA<br />

Figure 3.8: Strain rates in <strong>the</strong> Himalayas and <strong>the</strong> Zagros mountains for <strong>the</strong> model eurasia27-01-<br />

04 6. The values <strong>of</strong> <strong>the</strong> second invariant <strong>of</strong> <strong>the</strong> strain rate tensor are given in s −1 .<br />

regime would probably only be possible if stress levels are generally low in this part<br />

<strong>of</strong> Europe and indeed <strong>the</strong> calculations presented in Figure 3.2 show <strong>the</strong>m to be in<br />

<strong>the</strong> range <strong>of</strong> 30 to 40 MPa.<br />

For much <strong>of</strong> nor<strong>the</strong>rn and central Eurasia stress measurements are sparse, making<br />

it difficult to verify <strong>the</strong> validity <strong>of</strong> our model in <strong>the</strong>se settings.<br />

In summary we can say that, with respect to stresses, our elastoplastic model does<br />

well in predicting <strong>the</strong> stress orientations in much <strong>of</strong> Europe, as well as from Iran to<br />

Tibet. On <strong>the</strong> o<strong>the</strong>r hand <strong>the</strong> inability to predict <strong>the</strong> correct stress patterns along<br />

much <strong>of</strong> <strong>the</strong> eastern margin is a serious failure which raises <strong>the</strong> question if we are<br />

applying <strong>the</strong> correct forces. This suspicion is supported by <strong>the</strong> fact that Figure 3.2<br />

implies high stress levels and even smallish areas <strong>of</strong> yield long <strong>the</strong> eastern boundary,<br />

whereas in reality one does not observe any mountain building; <strong>the</strong> mountains in<br />

<strong>the</strong>se parts are <strong>of</strong> volcanic origin.<br />

One might argue that <strong>the</strong> issue could be resolved by integrating <strong>the</strong> high Himalayan


3.4. A COMPARISON WITH MEASUREMENTS 39<br />

Figure 3.9: Global strain rates published by Kreemer et al. [21], based on a compilation <strong>of</strong><br />

geodetically measured plate velocities.<br />

topography which leads to gravitational spreading and thus induces compression<br />

at <strong>the</strong> foot <strong>of</strong> <strong>the</strong> mountain range. While <strong>the</strong> negligence <strong>of</strong> topographical effects<br />

admittedly is a shortcoming <strong>of</strong> our simple model (indeed, Lithgow-Bertelloni and<br />

Guynn [22] have shown that topography may even be dominant in certain places),<br />

it is unlikely that this would influence <strong>the</strong> stress field all <strong>the</strong> way to <strong>the</strong> eastern<br />

plate boundary. Since <strong>the</strong>re is no obvious way to get E-W compression in <strong>the</strong> Far<br />

East while trench suction is pulling on <strong>the</strong> eastern margin <strong>of</strong> Eurasia, one is led to<br />

<strong>the</strong> conclusion that trench suction is not <strong>the</strong> appropriate boundary condition. In<br />

<strong>the</strong> second model presented after <strong>the</strong> next paragraph we will investigate <strong>the</strong> effect<br />

<strong>of</strong> leaving those boundaries unconstrained.<br />

Having treated <strong>the</strong> stress field we now discuss <strong>the</strong> strain rate. Figure 3.8 is a<br />

contour plot <strong>of</strong> values <strong>of</strong> <strong>the</strong> second invariant <strong>of</strong> <strong>the</strong> strain rate tensor<br />

I2 = ˙ɛ 2 12 + ˙ɛ 2 23 + ˙ɛ 2 31 − (˙ɛ11 ˙ɛ22 + ˙ɛ11 ˙ɛ33 + ˙ɛ22 ˙ɛ33) (3.5)<br />

in <strong>the</strong> Himalayas and <strong>the</strong> Zagros mountains. A comparison with <strong>the</strong> global strain<br />

rate distribution calculated by Kreemer et al. [21] (Figure 3.9) reveals that <strong>the</strong> values<br />

in <strong>the</strong> model are <strong>of</strong> <strong>the</strong> correct magnitude, but <strong>the</strong>ir distribution is not consistent<br />

with actual data. In <strong>the</strong> Himalayas <strong>the</strong> areas with <strong>the</strong> range <strong>of</strong> strain rates expected<br />

in regions undergoing continental collision are far smaller than in Figure 3.9 and<br />

also smaller than those <strong>of</strong> <strong>the</strong> model in Iran. The latter fact once again points to


40 CHAPTER 3. THE MODEL OF EURASIA<br />

unrealistic deformation caused by excessive, local shearing <strong>of</strong> <strong>the</strong> elements <strong>of</strong> <strong>the</strong><br />

finite element mesh and seems to indicate that, without this artificial phenomenon,<br />

strain rates would be lower than are observed.<br />

Figures 3.10 to 3.12 show <strong>the</strong> stress levels and orientations in a viscoelastic<br />

<strong>Eurasian</strong> plate in which dislocation creep is described by a power law rheology<br />

<strong>of</strong> order three 2 . To improve <strong>the</strong> faulty features in <strong>the</strong> stress and strain maps <strong>of</strong> <strong>the</strong><br />

elastoplastic model we change <strong>the</strong> following boundary constraints:<br />

1. Sections formerly experiencing trench suction (i.e. <strong>the</strong> trenches in <strong>the</strong> Far East<br />

and in <strong>the</strong> Aegean Sea) are now left free. Doing so could be justified if <strong>the</strong><br />

collisional resistance - a quantity we neglected so far - across <strong>the</strong> interface <strong>of</strong><br />

<strong>the</strong> subducting slab and <strong>the</strong> overriding plate is comparable in magnitude to<br />

trench suction, so that <strong>the</strong> two forces effectively cancel out.<br />

2. In Siberia, <strong>the</strong> convergence <strong>of</strong> North America on Eurasia happens at quite an<br />

oblique angle. We thus leave this segment unconstrained, interpreting it as a<br />

strike-slip boundary ra<strong>the</strong>r than a collisional zone.<br />

3. In an attempt to match <strong>the</strong> lateral extent <strong>of</strong> <strong>the</strong> areas with realistic strain<br />

rates in <strong>the</strong> Himalayas we apply collisional forces corresponding to an Indian<br />

indendation velocity <strong>of</strong> 15 cm per year. Albeit completely arbitrary on <strong>the</strong><br />

grounds <strong>of</strong> current observations this value is <strong>the</strong> estimated velocity <strong>of</strong> <strong>the</strong><br />

Indian plate at <strong>the</strong> onset <strong>of</strong> continental collision about 40 million years ago.<br />

There are two major differences between <strong>the</strong> von Mises stress maps <strong>of</strong> <strong>the</strong> elastoplastic<br />

(Figure 3.2) and <strong>the</strong> power-law model (Figure 3.10). First <strong>of</strong> all, both areas<br />

<strong>of</strong> yield have shrunk considerably, with <strong>the</strong> result that Iran now has a mountain<br />

range consistent with <strong>the</strong> lateral dimensions <strong>of</strong> <strong>the</strong> Zagros. Since boundary conditions<br />

in this part <strong>of</strong> <strong>the</strong> plate were not adapted it is obviously a consequence <strong>of</strong><br />

<strong>the</strong> change <strong>of</strong> rheology. The lithosphere, which is perfectly elastic below σY in <strong>the</strong><br />

elastoplastic model, propagates stresses into <strong>the</strong> interior more easily than <strong>the</strong> lithosphere<br />

described by a viscous power law relationship. For this latter case, some <strong>of</strong><br />

<strong>the</strong> work done by <strong>the</strong> boundary loads is dissipated by frictional effects. Secondly,<br />

stress values east <strong>of</strong> 95 ◦ E have decreased now that we are not pulling on <strong>the</strong> eastern<br />

margin and a glance at Figure 3.11 reveals that leaving it unconstrained results in<br />

E-W directed compression in China, in agreement with WSM data plotted in Figure<br />

3.5. The model even predicts <strong>the</strong> orientation <strong>of</strong> compressional stresses around 20 ◦<br />

2 For a power law as in Equation 3.1 <strong>the</strong> constant A has been determined experimentally. If<br />

we take <strong>the</strong> law to be <strong>of</strong> <strong>the</strong> form ˙ɛ = Aσ n instead, <strong>the</strong>n A contains <strong>the</strong> dependence <strong>of</strong> strain<br />

rate on temperature and pressure that was formerly described by <strong>the</strong> exponential factor. Since A<br />

is no longer constant in this case and our models do not incorporate <strong>the</strong> effects <strong>of</strong> pressure and<br />

temperature we were forced to calculate it with <strong>the</strong> observation that strain rates in areas <strong>of</strong> yield<br />

is on <strong>the</strong> order <strong>of</strong> 10 −15 s −1 . Inserting <strong>the</strong>se two values into <strong>the</strong> power law gives us A ≈ 10 −39<br />

P a −3 · s −1 .


3.4. A COMPARISON WITH MEASUREMENTS 41<br />

Figure 3.10: Stress intensity map for <strong>the</strong> model eurasia01-03-04 1 which obeys a power law<br />

rheology with a stress exponent <strong>of</strong> three.


42 CHAPTER 3. THE MODEL OF EURASIA<br />

Figure 3.11: Orientation <strong>of</strong> compressional principal stress for <strong>the</strong> model eurasia01-03-04 1.


3.4. A COMPARISON WITH MEASUREMENTS 43<br />

Figure 3.12: Orientation <strong>of</strong> extensional principal stress for <strong>the</strong> model eurasia01-03-04 1.


44 CHAPTER 3. THE MODEL OF EURASIA<br />

N/110 ◦ E, where <strong>the</strong>y rotate south into <strong>the</strong> sou<strong>the</strong>ast Asian protrusion. Sou<strong>the</strong>ast<br />

Asia itself has low stress levels in <strong>the</strong> new rheology, making it difficult to predict<br />

compressional stress directions based on Figure 3.11 or 3.12. Never<strong>the</strong>less extensional<br />

stresses do run parallel to <strong>the</strong> Java trench as <strong>the</strong> WSM tells us it should.<br />

Along much <strong>of</strong> <strong>the</strong> nor<strong>the</strong>rn perimeter <strong>of</strong> Eurasia, <strong>the</strong> modification <strong>of</strong> <strong>the</strong> boundary<br />

conditions results in compressional stresses lying parallel to <strong>the</strong> plate boundary<br />

although one would expect <strong>the</strong>m to be perpendicular to <strong>the</strong> edge because <strong>of</strong> ridge<br />

push. We have no explanation for this observation, however <strong>the</strong> fact that in Figure<br />

3.11 it also happens in areas where no cratons are found nearby supports our assumption<br />

that <strong>the</strong>y are not responsible for it. The effect appears to occur in regions<br />

<strong>of</strong> low stress intensities where a change <strong>of</strong> boundary conditions (e.g. <strong>the</strong> higher<br />

collisional forces in <strong>the</strong> Himalayas in <strong>the</strong> power-law model) might be able to rotate<br />

stress directions.<br />

In Figure 3.13 we plot <strong>the</strong> spatial variation <strong>of</strong> <strong>the</strong> values <strong>of</strong> <strong>the</strong> second invariant <strong>of</strong><br />

<strong>the</strong> strain rate tensor. Here power law rheology is responsible for a much shallower<br />

gradient <strong>of</strong> strain rate values, with <strong>the</strong> consequence that, for instance, <strong>the</strong> area<br />

bounded by <strong>the</strong> 5.8 × 10 6 s −16 contour (18.3 × 10 −9 in <strong>the</strong> more common units <strong>of</strong><br />

yr −1 ) agrees well with <strong>the</strong> extent <strong>of</strong> similar values in Figure 3.9. It is interesting<br />

to note that <strong>the</strong> sou<strong>the</strong>rn edges <strong>of</strong> both cratons within Eurasia define <strong>the</strong> trend <strong>of</strong><br />

<strong>the</strong> 8.3 × 10 −17 s −1 contour, supporting <strong>the</strong> notion that <strong>the</strong>y are not as easy to<br />

deform as normal continental lithosphere. The isolines <strong>of</strong> second invariant strain<br />

rate are much closer toge<strong>the</strong>r in Figure 3.8 than in Figure 3.13 because power laws<br />

with higher exponents (<strong>of</strong> which perfect plasticity is just <strong>the</strong> most extreme version)<br />

result in deformation which is more localized around <strong>the</strong> stress sources.<br />

The introduction <strong>of</strong> new boundary conditions has improved <strong>the</strong> stress orientations<br />

to such an extent that it is appears justified to claim that <strong>the</strong> power-law model agrees<br />

well with <strong>the</strong> large scale features <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> stress field. It has also lead to a<br />

realistic distribution <strong>of</strong> strain rates. The third comparison we can make is to test if<br />

<strong>the</strong> deformation velocities can be matched. Unfortunately I am not able to present<br />

a figure <strong>of</strong> <strong>the</strong> deformation fields <strong>of</strong> ei<strong>the</strong>r <strong>the</strong> elastoplastic or <strong>the</strong> power-law model<br />

due to technical problems encountered as this manuscript was nearing completion.<br />

Instead Figure 3.14 depicts <strong>the</strong> velocity field during <strong>the</strong> last final 160000 years <strong>of</strong><br />

ano<strong>the</strong>r model <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate (remember that <strong>the</strong> model is run for a total<br />

<strong>of</strong> approximately 1.5 million years). The boundary conditions are identical to those<br />

applied in <strong>the</strong> power-law model, but <strong>the</strong> rheology is such that it behaves elastically<br />

up to <strong>the</strong> yield limit and, if it is exceeded, obeys a power law creep <strong>of</strong> order three.<br />

In calculating <strong>the</strong> velocity field two important issues arise: (1) Is <strong>the</strong> velocity field<br />

in a steady state 3 and (2) in which reference frame should <strong>the</strong> field be plotted?<br />

An inversion for a component <strong>of</strong> rigid rotation was performed for eight consecutive<br />

periods in <strong>the</strong> second half <strong>of</strong> <strong>the</strong> computation time <strong>of</strong> <strong>the</strong> model. The pole <strong>of</strong><br />

3 If we are to interpret <strong>the</strong> velocity as a physically meaningful quantity arising from <strong>the</strong> applied<br />

boundary constraints, <strong>the</strong> motion <strong>of</strong> <strong>the</strong> plate should not be accelerating. If this requirement is<br />

satisfied <strong>the</strong> torques <strong>of</strong> <strong>the</strong> forces acting on our plate are balanced.


3.4. A COMPARISON WITH MEASUREMENTS 45<br />

Figure 3.13: Values <strong>of</strong> <strong>the</strong> second invariant <strong>of</strong> <strong>the</strong> strain rate tensor for <strong>the</strong> model eurasia01-03-<br />

04 1.


46 CHAPTER 3. THE MODEL OF EURASIA<br />

0.4 mm/yr<br />

0.08 mm/yr<br />

0.1 mm/yr<br />

0.3 mm/yr<br />

0.4 mm/yr<br />

0.5 mm/yr<br />

0.3 mm/yr<br />

0.5 mm/yr<br />

0.2 mm/yr<br />

0.5 mm/yr<br />

0.7 mm/yr<br />

1.2 mm/yr<br />

Figure 3.14: Deformation velocity during <strong>the</strong> last 160000 years <strong>of</strong> <strong>the</strong> model eurasia02-02-04 1.<br />

rotation varied over an area corresponding to 1/100000 <strong>of</strong> <strong>the</strong> area <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong><br />

plate and <strong>the</strong> angular frequency strayed from its mean by at most four percent.<br />

Having confirmed that our plate is in a steady state we filtered out <strong>the</strong> component<br />

<strong>of</strong> rigid rotation for <strong>the</strong> time span <strong>of</strong> interest and chose a reference frame such that<br />

a node between <strong>the</strong> East European platform and <strong>the</strong> Siberian craton is at rest.<br />

The result obtained does not agree with <strong>the</strong> data from GPS and o<strong>the</strong>r measurements.<br />

For example <strong>the</strong> magnitudes <strong>of</strong> <strong>the</strong> velocities are all too small by a factor <strong>of</strong> ten. Such<br />

an error might be caused by too high coefficients <strong>of</strong> friction in <strong>the</strong> dashpots used<br />

to simulate basal drag forces and could be overlooked if <strong>the</strong> pattern <strong>of</strong> <strong>the</strong> velocity<br />

field matched measurements. Yet this is not <strong>the</strong> case. Referring to Bird’s [5] plate<br />

boundary model we see that only <strong>the</strong> directions <strong>of</strong> velocities east <strong>of</strong> <strong>the</strong> Himalayas<br />

are in rough agreement with observations. In Europe we predict motion to <strong>the</strong> west<br />

whereas according to Bird it is towards <strong>the</strong> nor<strong>the</strong>ast, while in <strong>the</strong> Himalayas our<br />

deformation velocities are lower by a factor <strong>of</strong> thirty than suggested by Kreemer et<br />

al. [21].<br />

Since we work with a simple two dimensional model, we cannot expect to repro-


3.4. A COMPARISON WITH MEASUREMENTS 47<br />

duce <strong>the</strong> detailed properties <strong>of</strong> <strong>the</strong> stress, strain and velocity fields. The velocity<br />

field depends strongly on <strong>the</strong> location <strong>of</strong> plate boundaries. Bird [5] proposes <strong>the</strong><br />

existence <strong>of</strong> several smaller plates within <strong>the</strong> area <strong>of</strong> our model, in view <strong>of</strong> which<br />

it is not surprising that we fail to predict a small scale feature such as <strong>the</strong> motion<br />

<strong>of</strong> <strong>the</strong> Aegean plate along <strong>the</strong> North Anatolian fault in ei<strong>the</strong>r <strong>the</strong> velocity or <strong>the</strong><br />

stress field. Similarly accounting for topographical forces o<strong>the</strong>r than <strong>the</strong> continental<br />

margin force would lead to local corrections <strong>of</strong> our generally satisfactory stress field.<br />

Finally it should be mentioned that, as far as we are aware, none <strong>of</strong> <strong>the</strong> studies to<br />

date has even attempted to match <strong>the</strong> whole set <strong>of</strong> stress, strain rate and velocity<br />

data with a single model. Our efforts, even if carried out at a simple two-dimensional<br />

level, have rewarded us with a model (presented in <strong>the</strong> Figures 3.10 to 3.13) that<br />

makes satisfactory to good predictions with respect to large scale properties <strong>of</strong> stress<br />

and strain rate within <strong>the</strong> <strong>Eurasian</strong> plate. It will be <strong>the</strong> subject <strong>of</strong> fur<strong>the</strong>r research<br />

to determine if <strong>the</strong> velocity field can also be made to agree with observations.


48 CHAPTER 3. THE MODEL OF EURASIA


THE CASE HAS, IN SOME RESPECTS, BEEN NOT ENTIRELY DEVOID OF INTEREST.<br />

Sherlock Holmes, ’A Case <strong>of</strong> Identity’<br />

Chapter 4<br />

Conclusions<br />

The main aim <strong>of</strong> <strong>the</strong> work presented in <strong>the</strong> preceding chapters was to determine how<br />

different rheological models and lateral strength variations influence <strong>the</strong> characteristics<br />

<strong>of</strong> <strong>the</strong> lithospheric stress field. We limited our research to two dimensional finite<br />

element models carried out in <strong>the</strong> program ABAQUS, with which <strong>the</strong> concepts <strong>of</strong><br />

viscoelasticity, plasticity and more complex material behaviour like power-law rheologies<br />

can be implemented. The <strong>Eurasian</strong> plate is <strong>the</strong> setting underlying our tests<br />

and we ultimately propose a model for this area based on <strong>the</strong> experience ga<strong>the</strong>red<br />

in geometrically simple rectangles.<br />

The rectangular models <strong>of</strong> Chapter 2 allowed us to first investigate <strong>the</strong>se issues<br />

on a qualitative basis, without obliging us to match <strong>the</strong> details <strong>of</strong> <strong>the</strong> stress field in<br />

Eurasia. We found that thicker regions <strong>of</strong> lithosphere reduce stress levels within <strong>the</strong><br />

structure and that, at <strong>the</strong> edges <strong>of</strong> such regions, stress orientations can change. Such<br />

stress bending, however, occurs only if <strong>the</strong> structure is situated in an area where<br />

different sources <strong>of</strong> stress (e.g. boundary conditions along different edge segments<br />

<strong>of</strong> <strong>the</strong> plate) contribute roughly equally to <strong>the</strong> local stress field. Regions <strong>of</strong> thinner<br />

lithosphere than <strong>the</strong>ir surroundings display higher stress intensities and <strong>the</strong>y also<br />

tend to bend stress directions, but less so than thicker lithosphere. Fur<strong>the</strong>rmore<br />

<strong>the</strong>re is no evidence that thicker regions such as cratons act as stress barriers in<br />

that <strong>the</strong>y shield areas on opposite sides <strong>of</strong> <strong>the</strong> structure from <strong>the</strong> influence <strong>of</strong> one<br />

ano<strong>the</strong>r.<br />

While cratons do not decouple different parts <strong>of</strong> <strong>the</strong> lithosphere, introducing plasticity<br />

by imposing a yield strength to <strong>the</strong> lithosphere does. Stress patterns for purely<br />

elastic and viscoelastic rheologies are generally very similar in that <strong>the</strong> stress pattern<br />

is dominated by <strong>the</strong> largest boundary load (e.g., <strong>the</strong> collision with India in<br />

Eurasia) in both cases. Only in an elastoplastic rheology can weaker stress sources<br />

(an example <strong>of</strong> which is <strong>the</strong> collision with Africa) make a noticeable contribution to<br />

<strong>the</strong> stress orientations in <strong>the</strong> regions where <strong>the</strong>y act.<br />

With respect to boundary conditions we maintain that dynamic boundary conditions<br />

49


50 CHAPTER 4. CONCLUSIONS<br />

are <strong>the</strong> better approach for models <strong>of</strong> <strong>the</strong> lithosphere than kinematic constraints.<br />

The latter are based on measurements <strong>of</strong> relative plate motions which are instantaneous<br />

quantities and probably also only are valid at <strong>the</strong> surface <strong>of</strong> <strong>the</strong> Earth.<br />

Fur<strong>the</strong>rmore, <strong>the</strong> relative motion <strong>of</strong> <strong>the</strong> plates contains no information <strong>of</strong> <strong>the</strong> processes<br />

involved at <strong>the</strong> interfaces <strong>of</strong> tectonic plates. Concerning dynamic boundary<br />

conditions we believe that collisional forces are significantly higher than proposed<br />

by previous studies which calculate <strong>the</strong> relative magnitudes <strong>of</strong> boundary forces by<br />

assuming that <strong>the</strong>ir torques are balanced. In keeping with this we suggest that <strong>the</strong><br />

forces contributing to <strong>the</strong> torque balance are probably <strong>the</strong> yield forces acting on <strong>the</strong><br />

interior <strong>of</strong> a plate at <strong>the</strong> boundary <strong>of</strong> an area <strong>of</strong> yield. While this distinction may<br />

not be crucial to global models seeking to explain large scale dynamics, it is likely<br />

to be important when it comes to setting up detailed models <strong>of</strong> individual plates or<br />

specific regions within <strong>the</strong>m.<br />

Models employing <strong>the</strong> real shape <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate entail a higher level <strong>of</strong><br />

sophistication. The prediction <strong>of</strong> a stress field that agrees well with observations necessitates<br />

a quantitative understanding <strong>of</strong> <strong>the</strong> boundary forces. Based on rheological<br />

and dynamic arguments we developed scaling relations between relative plate velocities<br />

and <strong>the</strong> corresponding forces in collisional and subduction zones. Throughout<br />

most <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate <strong>the</strong> application <strong>of</strong> <strong>the</strong> boundary constraints according to<br />

our assumptions leads to satisfactory results. Contrary to our expectations, stress<br />

orientations in <strong>the</strong> Far East are at odds with observations if we apply trench suction.<br />

Leaving <strong>the</strong> eastern margin <strong>of</strong> Eurasia unconstrained results in a good match with<br />

World Stress Map data for <strong>the</strong> region and probably indicates that collisional resistance<br />

along <strong>the</strong> interface <strong>of</strong> <strong>the</strong> overriding and <strong>the</strong> subducting plate approximately<br />

cancels trench suction.<br />

Models <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate were run using ei<strong>the</strong>r (1) a purely elastic plate in which<br />

<strong>the</strong> material deforms without resistance once <strong>the</strong> plastic yield limit is exceeded, or<br />

(2) a third order viscous power law rheology. While stress orientations are generally<br />

insensitive to <strong>the</strong> different material behaviour if identical boundary constraints<br />

are applied, stress levels, and consequently <strong>the</strong> extent <strong>of</strong> lithospheric yielding, do<br />

change. Due to frictional dissipation, which drains <strong>the</strong> system <strong>of</strong> some <strong>of</strong> <strong>the</strong> energy<br />

furnished by <strong>the</strong> boundary loads, stress levels remain lower in <strong>the</strong> interior <strong>of</strong><br />

<strong>the</strong> model described by <strong>the</strong> power law rheology. At <strong>the</strong> same time, prescribing a<br />

power law behaviour seems to remove a tendency <strong>of</strong> purely elastoplastic lithosphere<br />

to deform too easily under shear (e.g., oblique boundary loads), resulting in a more<br />

realistic prediction <strong>of</strong> <strong>the</strong> size <strong>of</strong> orogens.<br />

The analysis <strong>of</strong> our rectangular models provided us with strong evidence that <strong>the</strong><br />

implementation <strong>of</strong> plasticity is <strong>the</strong> prerequisite for a good stress model <strong>of</strong> Eurasia.<br />

Interestingly, a power law rheology with appropriately chosen parameters seems<br />

to have very similar effects on <strong>the</strong> outcome <strong>of</strong> <strong>the</strong> stress pattern and it thus also<br />

appears to be suited for <strong>the</strong> task. To determine whe<strong>the</strong>r this is generally valid or<br />

simply a result <strong>of</strong> <strong>the</strong> geometry <strong>of</strong> <strong>the</strong> <strong>Eurasian</strong> plate we will need to run additional<br />

rectangular models employing <strong>the</strong> same viscoelastic power-law behaviour.


We believe that realistic models <strong>of</strong> continental lithosphere require <strong>the</strong> application <strong>of</strong><br />

some kind <strong>of</strong> nonlinear rheological behaviour. For models <strong>of</strong> <strong>the</strong> more rigid oceanic<br />

lithosphere on <strong>the</strong> o<strong>the</strong>r hand, an elastic rheology might suffice.<br />

Our best model can account for <strong>the</strong> observed directions <strong>of</strong> maximal horizontal<br />

compression and for stress levels in <strong>the</strong> <strong>Eurasian</strong> plate. It also predicts a reasonable<br />

distribution <strong>of</strong> strain rates, but is inadequate when it comes to forecasting <strong>the</strong><br />

deformation velocities. It is possible that a calculation <strong>of</strong> <strong>the</strong> velocity field for our<br />

favoured model presented on pages 41 et sqq. will allow us to modify our judgement<br />

in this respect. If this is not <strong>the</strong> case, <strong>the</strong> mismatch might be due to ei<strong>the</strong>r a<br />

conceptual or technical flaw in our extraction <strong>of</strong> <strong>the</strong> actual deformation from <strong>the</strong><br />

whole displacement. It might also be a hint that by rejecting an active role <strong>of</strong><br />

<strong>the</strong> mantle <strong>the</strong>re may be regions where we have not applied <strong>the</strong> correct boundary<br />

conditions at <strong>the</strong> base <strong>of</strong> <strong>the</strong> lithosphere.<br />

51


52 CHAPTER 4. CONCLUSIONS


Appendix A<br />

The contents <strong>of</strong> this appendix is based on <strong>the</strong> first two chapters <strong>of</strong> <strong>the</strong> ABAQUS<br />

Theory Manual.<br />

A.1 Theory<br />

In structural analysis one is interested in <strong>the</strong> deformation <strong>of</strong> an initial configuration<br />

throughout <strong>the</strong> history <strong>of</strong> loading. A material particle initially located at some position<br />

Xi (i=1, 2, 3) in space will move to a new position xi (“current configuration”):<br />

if one assumes material cannot appear or disappear, <strong>the</strong>re will be a one-to-one correspondence<br />

between Xi and xi, so it is always possible to write <strong>the</strong> history <strong>of</strong> a<br />

particle’s location as<br />

xi = xi(Xn, t) (A.1)<br />

and this relationship can be inverted. Two neighbouring particles, located at Xi<br />

and at Xi + dXi in <strong>the</strong> initial configuration, must satisfy<br />

dxi = ∂xi<br />

∂Xj<br />

· dXj = Fij · dXj<br />

(A.2)<br />

in <strong>the</strong> current configuration. The matrix Fij is called <strong>the</strong> deformation gradient<br />

matrix.<br />

The velocity <strong>of</strong> a material particle is vi = ∂xi , where <strong>the</strong> partial derivative with<br />

∂t<br />

respect to time t means <strong>the</strong> rate <strong>of</strong> change <strong>of</strong> <strong>the</strong> spatial position, xi, <strong>of</strong> a particular<br />

particle. ABAQUS thus takes a Lagrangian viewpoint: it follows a material particle<br />

through <strong>the</strong> motion, ra<strong>the</strong>r than looking at a fixed point in space and watching<br />

<strong>the</strong> material flow through this point. The Lagrangian perspective makes it easy to<br />

record and update <strong>the</strong> state <strong>of</strong> a material point since <strong>the</strong> mesh is embedded in <strong>the</strong><br />

53


54 APPENDIX A<br />

material.<br />

The velocity difference between two neighbouring particles in <strong>the</strong> current configuration<br />

is<br />

dvi = ∂vi<br />

∂xj<br />

· dxj = Lij · dxj<br />

(A.3)<br />

where Lij is <strong>the</strong> velocity gradient. It is composed <strong>of</strong> a rate <strong>of</strong> deformation plus a<br />

rate <strong>of</strong> rotation and can be split into a symmetric strain rate matrix<br />

˙ɛij = 1<br />

2 (Lij + Lji) = 1<br />

<br />

T<br />

∂vi ∂vi<br />

+<br />

2 ∂xj ∂xj<br />

and an antisymmetric rotation rate matrix<br />

(A.4)<br />

Ωij = 1<br />

2 (Lij − Lji) = 1<br />

<br />

T<br />

∂vi ∂vi<br />

− . (A.5)<br />

2 ∂xj ∂xj<br />

Many <strong>of</strong> <strong>the</strong> problems to which ABAQUS is applied involve finding an approximate<br />

(finite element) solution for <strong>the</strong> displacements, deformations, stresses and<br />

forces in a solid body subjected to some history <strong>of</strong> loading. The exact solution <strong>of</strong><br />

such a problem requires that both force and moment equilibria be maintained at all<br />

times over any arbitrary volume <strong>of</strong> <strong>the</strong> body.<br />

Let V denote <strong>the</strong> volume occupied by a part <strong>of</strong> <strong>the</strong> body in <strong>the</strong> current configuration<br />

and S be <strong>the</strong> surface bounding this volume. Force equilibrium for <strong>the</strong> volume<br />

states that <strong>the</strong> integral <strong>of</strong> <strong>the</strong> surface tractions ti a over S and <strong>the</strong> integral <strong>of</strong> <strong>the</strong><br />

body forces fi over <strong>the</strong> volume be equal but opposite in magnitude. For <strong>the</strong> i-th<br />

component this yields:<br />

<br />

S<br />

<br />

ti dS +<br />

The surface traction is related to <strong>the</strong> stress tensor σij by<br />

V<br />

fi dV = 0. (A.6)<br />

ti = σijnj, (A.7)<br />

where nj is <strong>the</strong> unit outward normal to S. Using this definition and Gauss’ <strong>the</strong>orem<br />

to rewrite <strong>the</strong> surface integral as a volume integral we get


A.1. THEORY 55<br />

<br />

V<br />

∂σij<br />

∂xj<br />

<br />

+ fi dV = 0. (A.8)<br />

Since <strong>the</strong> volume is arbitrary, <strong>the</strong> integrand must vanish everywhere, thus providing<br />

three differential equations <strong>of</strong> translational equilibrium.<br />

Moment equilibrium is most simply written by taking moments about <strong>the</strong> origin.<br />

The vector product <strong>of</strong> Equation A.6 with xk is:<br />

<br />

S<br />

<br />

ɛlkixkti dS +<br />

V<br />

ɛlkixkfi dV = 0. (A.9)<br />

With <strong>the</strong> definition <strong>of</strong> <strong>the</strong> stress tensor σij in A.7 and <strong>the</strong> help <strong>of</strong> Gauss’ <strong>the</strong>orem<br />

one can prove that <strong>the</strong> stress tensor is symmetric:<br />

σij = σji. (A.10)<br />

By chosing <strong>the</strong> stress matrix to be symmetric, moment equilibrium is satisfied automatically<br />

and one only needs to consider translational equilibrium when explicitly<br />

writing <strong>the</strong> equilibrium equations.<br />

ABAQUS approximates <strong>the</strong> equilibrium requirement by replacing it with a weaker<br />

requirement, that equilibrium must be maintained in an average sense over a finite<br />

number <strong>of</strong> divisions <strong>of</strong> <strong>the</strong> volume’s body. To develop such an approximation <strong>the</strong><br />

three equations represented by Equation A.8 are replaced by an equivalent “weak<br />

form” - a single scalar equation over <strong>the</strong> entire body. It is obtained by multiplying<br />

<strong>the</strong> pointwise differential equations by an arbitrary, vector-valued test function,<br />

defined over <strong>the</strong> entire volume, and integrating. The test function can be imagined<br />

to be a “virtual” velocity field, δvi, which is completely arbitrary except that it<br />

must obey any prescribed kinematic constraints and have sufficient continuity: <strong>the</strong><br />

scalar product <strong>of</strong> this test function with <strong>the</strong> equilibrium force field <strong>the</strong>n represents<br />

<strong>the</strong> “virtual” 1 work rate.<br />

Taking <strong>the</strong> scalar product <strong>of</strong> <strong>the</strong> equation describing translational equilibrium with<br />

δvi and integrating over <strong>the</strong> entire body gives<br />

<br />

V<br />

∂σij<br />

∂xj<br />

<br />

+ fi · δvi dV = 0. (A.11)<br />

From this expression we need to derive a basic equilibrium statement for <strong>the</strong> finite<br />

element formulation that will be introduced in <strong>the</strong> next section (“Procedures”). The<br />

1 Virtual quantities are infinitesimally small variations <strong>of</strong> physical measurements.


56 APPENDIX A<br />

chain rule allows us to write<br />

so that<br />

<br />

V<br />

∂σij<br />

∂xj<br />

∂<br />

∂xj<br />

· (σijvi) = ∂σij<br />

· δvi dV =<br />

=<br />

=<br />

<br />

<br />

<br />

V<br />

S<br />

S<br />

∂xj<br />

· δvi + σij · ∂δvi<br />

, (A.12)<br />

∂xj<br />

<br />

∂<br />

· (σijvi) − σij ·<br />

∂xj<br />

∂δvi<br />

<br />

dV<br />

∂xj<br />

<br />

njσijδvi dS − σij ·<br />

V<br />

∂δvi<br />

dV<br />

∂xj<br />

<br />

tiδvi dS − σij · ∂δvi<br />

dV,<br />

∂xj<br />

where Gauss’ <strong>the</strong>orem and <strong>the</strong> definition <strong>of</strong> <strong>the</strong> stress tensor were applied in <strong>the</strong><br />

first and second equalities respectively. Thus, <strong>the</strong> virtual work statement, Equation<br />

A.11, can be written<br />

<br />

S<br />

<br />

tiδvi dS +<br />

V<br />

<br />

fiδvi dV =<br />

V<br />

V<br />

σij · ∂δvi<br />

∂xj<br />

dV. (A.13)<br />

The quantitiy ∂δvi is <strong>the</strong> virtual version δLij <strong>of</strong> <strong>the</strong> velocity gradient introduced<br />

∂xj<br />

in Equation A.3. It too may be expressed as <strong>the</strong> sum<br />

δLij = δDij + δΩij,<br />

<strong>of</strong> its symmetric and antisymmetric parts<br />

With <strong>the</strong>se definitions<br />

and since σij is symmetric,<br />

δDij = 1<br />

2 (δLij + δLji)<br />

δΩij = 1<br />

2 (δLij − δLji).<br />

σijδLij = σijδDij + σijδΩij,


A.1. THEORY 57<br />

σijδΩij = 1<br />

2 (σijδLij − σijδLji) = 1<br />

2 (σijδLij − σjiδLji) = 0.<br />

Thus, in its classical form, <strong>the</strong> virtual work statement that will be used for <strong>the</strong><br />

finite element analysis is<br />

<br />

V<br />

<br />

σijδDij dV =<br />

S<br />

<br />

tiδvi dS +<br />

Recall that ti, fi, and σij are an equilibrium set:<br />

ti = σijnj,<br />

∂σij<br />

∂xj<br />

and that δDij and δvi are compatible:<br />

+ fi = 0, σij = σji;<br />

δDij = 1<br />

<br />

∂δvi<br />

+<br />

2 ∂xj<br />

∂δvj<br />

<br />

.<br />

∂xi<br />

V<br />

fiδvi dV. (A.14)<br />

The virtual work statement has a simple physical interpretation: <strong>the</strong> rate <strong>of</strong><br />

work done by <strong>the</strong> external forces ti and fi subjected to any virtual velocity field<br />

δvi is equal to <strong>the</strong> rate <strong>of</strong> work done by <strong>the</strong> equilibrium stresses σij, acting at <strong>the</strong><br />

rate <strong>of</strong> deformation δDij <strong>of</strong> <strong>the</strong> same virtual velocity field. The advantage <strong>of</strong> this<br />

formulation is that it is cast in <strong>the</strong> form <strong>of</strong> an integral over <strong>the</strong> volume <strong>of</strong> <strong>the</strong> body:<br />

it is possible to introduce approximations by choosing test functions for <strong>the</strong> virtual<br />

velocity field that are not entirely arbitrary, but <strong>the</strong> variation <strong>of</strong> which is restricted<br />

to a finite number <strong>of</strong> nodal values. This approach provides a stronger ma<strong>the</strong>matical<br />

basis for studying <strong>the</strong> approximation than <strong>the</strong> alternative <strong>of</strong> direct discretisation<br />

<strong>of</strong> <strong>the</strong> derivative in <strong>the</strong> differential equation <strong>of</strong> equilibrium at a point, which is <strong>the</strong><br />

typical starting point for a finite difference approach to <strong>the</strong> same problem.


58 APPENDIX A<br />

A.2 Procedures<br />

In Equation A.14 <strong>the</strong> internal virtual work rate<br />

<br />

σijδDij dV<br />

V<br />

was expressed directly in terms <strong>of</strong> <strong>the</strong> current volume V. The elasticity <strong>of</strong> a<br />

material is derivable from a <strong>the</strong>rmodynamic potential written about a reference<br />

state to which it returns upon unloading. Therefore, for iso<strong>the</strong>rmal deformations,<br />

<strong>the</strong>re will be a potential function for <strong>the</strong> elastic strain energy per unit <strong>of</strong> <strong>the</strong> natural<br />

reference volume. The internal virtual work rate may be rewritten as an integral<br />

over <strong>the</strong> natural reference volume:<br />

<br />

V<br />

<br />

σijδDij dV =<br />

V 0<br />

J · σijδDij dV 0<br />

where <strong>the</strong> Jacobian J = dV/dV 0 is <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> volume <strong>of</strong> <strong>the</strong> material in <strong>the</strong><br />

current and <strong>the</strong> natural configurations. It is <strong>the</strong>n convenient to define <strong>the</strong> stress<br />

measure<br />

τij = Jσij<br />

(A.15)<br />

as <strong>the</strong> work conjugate to <strong>the</strong> strain measure, <strong>the</strong> rate <strong>of</strong> which is <strong>the</strong> rate <strong>of</strong> deformation,<br />

Dij. Employing <strong>the</strong> natural reference volume V 0 , Equation A.14 becomes<br />

<br />

V 0<br />

τijδ ˙ɛij dV 0 =<br />

<br />

S<br />

<br />

tiδvi dS +<br />

V<br />

fiδvi dV, (A.16)<br />

where τij and ɛij are any conjugate pairs <strong>of</strong> material stress and strain measures.<br />

In a first discretisation step a finite element interpolator is introduced, <strong>the</strong> i th component<br />

<strong>of</strong> which is:<br />

ui = <br />

N nodes<br />

N N i u N ,<br />

where <strong>the</strong> u N are nodal variables (e.g. displacement or temperature), <strong>the</strong> N N i are<br />

interpolation functions that depend on some material coordinate system and <strong>the</strong>


A.2. PROCEDURES 59<br />

summation is carried out over all nodes in <strong>the</strong> finite element mesh. For example,<br />

<strong>the</strong> vector u might represent <strong>the</strong> displacement field <strong>of</strong> <strong>the</strong> solid, in which case <strong>the</strong><br />

individual values <strong>of</strong> u N would be <strong>the</strong> magnitudes <strong>of</strong> <strong>the</strong> displacement at <strong>the</strong> different<br />

nodes. The contribution <strong>of</strong> each node (in direction and amplitude) to <strong>the</strong> total<br />

displacement field is determined by <strong>the</strong> N N i , which can be regarded as directional<br />

weighting functions 2 .<br />

The virtual field, δvi, must be compatible with all kinematic constraints. Introducing<br />

<strong>the</strong> above interpolation implies that it must have an identical spatial form:<br />

δvi = <br />

N nodes<br />

N N i δv N .<br />

Now δ ˙ɛij is <strong>the</strong> virtual rate <strong>of</strong> material strain associated with δvi and, because it<br />

is a rate form, it must be linear in δvi. Hence, <strong>the</strong> interpolation assumption gives 3<br />

δ ˙ɛij = <br />

N nodes<br />

β N ij δv N ,<br />

) is a matrix that depends, in general, on <strong>the</strong> current<br />

where βN ij = βN ij (xk, N N l<br />

position xi and <strong>the</strong> interpolation functions N N i . The equilibrium equation A.16 is<br />

approximated as<br />

δv N<br />

<br />

V 0<br />

β N ij τij dV 0 = δv N<br />

<br />

tiN<br />

S<br />

N i dS +<br />

<br />

fiN<br />

V<br />

N i dV<br />

Since <strong>the</strong> δv N are independent variables, each one can be chosen to be nonzero<br />

and all o<strong>the</strong>rs zero in turn, to arrive at a system <strong>of</strong> nonlinear equilibrium equations:<br />

<br />

V 0<br />

β N ij τij dV 0 =<br />

<br />

tiN<br />

S<br />

N i dS +<br />

<br />

<br />

.<br />

fiN<br />

V<br />

N i dV. (A.17)<br />

This system <strong>of</strong> equations forms <strong>the</strong> basis for <strong>the</strong> finite element analysis procedure.<br />

However, for <strong>the</strong> Newton algorithm used in ABAQUS/Standard, one needs to know<br />

2 Consider <strong>the</strong> following example which shows that <strong>the</strong> contributions <strong>of</strong> <strong>the</strong> different nodes must<br />

be weighted by <strong>the</strong> N N i : let a circle be approximated by an octagon and each node be displaced by<br />

<strong>the</strong> same magnitude and in <strong>the</strong> same direction. The octagon as a whole moves by <strong>the</strong> same distance<br />

as <strong>the</strong> individual nodes, so one cannot simply sum all nodal displacements but must divide <strong>the</strong><br />

sum by eight to obtain <strong>the</strong> correct displacement. By doing so one effectively introduces a weighting<br />

factor <strong>of</strong> 1<br />

8 for each node.<br />

3Only <strong>the</strong> sums over <strong>the</strong> nodal variables are written out explicitly. For <strong>the</strong> subscripts Einstein’s<br />

summing convention is applied.


60 APPENDIX A<br />

<strong>the</strong> Jacobian <strong>of</strong> <strong>the</strong> finite element equilibrium equations. It can be developed by<br />

taking <strong>the</strong> variation <strong>of</strong> Equation A.16, giving<br />

<br />

V 0<br />

(dτijδ ˙ɛij + τijdδ ˙ɛij) dV 0 <br />

−<br />

S<br />

<br />

dtiδvi dS −<br />

<br />

<br />

− dfiδvi dV − fiδvidJ<br />

V<br />

V<br />

1<br />

J<br />

S<br />

tiδvidAr<br />

1<br />

Ar<br />

dS<br />

dV = 0, (A.18)<br />

where d( ) represents <strong>the</strong> linear variation <strong>of</strong> <strong>the</strong> quantity ( ) with respect to <strong>the</strong><br />

basic variables (<strong>the</strong> degrees <strong>of</strong> freedom <strong>of</strong> <strong>the</strong> finite element model). In <strong>the</strong> above<br />

expression J = |dV/dV 0 | is <strong>the</strong> volume change between <strong>the</strong> reference and <strong>the</strong> current<br />

volume occupied by a piece <strong>of</strong> <strong>the</strong> solid and, likewise, Ar = |dS/dS 0 | is <strong>the</strong> surface<br />

ratio between <strong>the</strong> reference and <strong>the</strong> current configuration. The Jacobian is obtained<br />

by allowing only variations <strong>of</strong> <strong>the</strong> nodal variables u N in Equation A.18, and after a<br />

lengthy calculation one obtains<br />

K MN <br />

=<br />

Here <strong>the</strong> two quantities<br />

<br />

−<br />

V 0<br />

β M ij Hjkβ N ki dV 0 +<br />

S<br />

N M i Q N <br />

S,i dS −<br />

Q N S,i = ∂ti<br />

+ ti<br />

∂uN Q N V,i = ∂fi<br />

+ fi<br />

∂uN <br />

V 0<br />

∂β<br />

τij<br />

M ij 0<br />

dV<br />

∂uN N<br />

V<br />

M i Q N V,i dV. (A.19)<br />

1<br />

Ar<br />

∂Ar<br />

∂u N<br />

1 ∂J<br />

J ∂uN stand for <strong>the</strong> variation <strong>of</strong> <strong>the</strong> load vectors with nodal variables, and based on<br />

mechanical constitutive <strong>the</strong>ory it is assumed that dτij can be expressed as<br />

dτij = Hikdɛkj + gij,<br />

where <strong>the</strong> matrices Hij and gij depend on <strong>the</strong> properties <strong>of</strong> <strong>the</strong> material being<br />

loaded.<br />

ABAQUS/Standard uses <strong>the</strong> Newton incremental method for solving <strong>the</strong> nonlinear<br />

equilibrium equations A.17 and A.19. They can symbolically be written as


A.2. PROCEDURES 61<br />

F N (u M ) = 0, (A.20)<br />

where F N is <strong>the</strong> force component conjugate to <strong>the</strong> N th variable in <strong>the</strong> problem<br />

and u M is <strong>the</strong> value <strong>of</strong> <strong>the</strong> M th nodal variable.<br />

The basic idea <strong>of</strong> Newton’s method is <strong>the</strong> following. Let u M n be an approximation<br />

to <strong>the</strong> solution <strong>of</strong> Equation A.20 reached after <strong>the</strong> n th iteration and let c M n+1 denote<br />

<strong>the</strong> difference between this solution and <strong>the</strong> exact solution, meaning that<br />

F N (u M n + c M n+1) = 0.<br />

The left hand side <strong>of</strong> this equation can be expanded in a Taylor series about<br />

<strong>the</strong> incremental solution u M n . If u M n is a good approximation <strong>of</strong> <strong>the</strong> solution, <strong>the</strong><br />

magnitude <strong>of</strong> each c M n+1 will be small, and it is thus possible to neglect terms <strong>of</strong><br />

second and higher orders in <strong>the</strong> series expansion, such that:<br />

F N (u M n ) + <br />

P nodes<br />

∂F N<br />

∂u P (uM n ) · c P n+1 = 0.<br />

One thus is left with a linear system <strong>of</strong> equations<br />

where KNP n = ∂F N<br />

solution is <strong>the</strong>n<br />

c P n+1 = <br />

P nodes<br />

<br />

− K NP −1 N<br />

n Fn (u M <br />

n ) , (A.21)<br />

∂u P is <strong>the</strong> Jacobian matrix. The next approximation to <strong>the</strong><br />

u M n+1 = u M n + c M n+1<br />

and <strong>the</strong> iteration continues until <strong>the</strong> solution to Equation A.20 converges.


62 APPENDIX A


Appendix B<br />

Appendix B gives an overview <strong>of</strong> all relevant input parameters for each <strong>of</strong> <strong>the</strong> models<br />

mentioned in this <strong>the</strong>sis. The models are listed in <strong>the</strong> order in which <strong>the</strong>y appear<br />

in <strong>the</strong> text. (E stands for Young’s modulus, ν for <strong>the</strong> Poisson ratio and σY for <strong>the</strong><br />

plastic yield limit).<br />

kin23-02-04 1 (Fig. 2.3, p. 11):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 1 × 10 12 s (approx. 32000 years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. Rheology: purely elastic with E = 1 × 10 11 Pa & ν = 0.3<br />

8. Lithospheric thickness (uniform): hL = 60 km<br />

9. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

10. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

11. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

63


64 Appendix B<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

12. Kinematic boundary constraints: relative velocities <strong>of</strong> neighbouring plates<br />

from Siberia to Gibraltar<br />

dsm elastic (Fig. 2.4, p. 12):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 1 × 10 12 s (approx. 32000 years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. Rheology: purely elastic with E = 1 × 10 11 Pa & ν = 0.3<br />

8. Lithospheric thickness (uniform): hL = 60 km<br />

9. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

10. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

11. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m


dsm elasticrat (Fig. 2.6, p. 14):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 1 × 10 12 s (approx. 32000 years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton<br />

8. Rheology (identical for cratons and continental lithosphere): purely elastic<br />

with E = 1 × 10 11 Pa & ν = 0.3<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

dsm elastohypocrat1 (Fig. 2.7, p. 15):<br />

1. Rectangle length: 16000 km (60 elements)<br />

65


66 Appendix B<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 1 × 10 12 s (approx. 32000 years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. Craton in <strong>the</strong> region dominated by <strong>the</strong> subduction zones’ stress field<br />

8. Rheology (identical for cratons and continental lithosphere): purely elastic<br />

with E = 1 × 10 11 Pa & ν = 0.3<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

dsm elastithincrat (Fig. 2.8, p. 16):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2


4. Computation time: 1 × 10 12 s (approx. 32000 years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton with thinner lithosphere<br />

8. Rheology (identical for cratons and continental lithosphere): purely elastic<br />

with E = 1 × 10 11 Pa & ν = 0.3<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 100 km<br />

10. Thickness <strong>of</strong> thinned lithosphere: hCRAT = 50 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

dsm creepcrat (Fig. 2.9, p. 17):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

5. Minimal time step: 1 × 10 5 s<br />

67


68 Appendix B<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton<br />

8. Rheology (identical for cratons and continental lithosphere): diffusion creep<br />

until <strong>the</strong> yield stress <strong>of</strong> 50 MPa is reached, <strong>the</strong>n power-law creep (dislocation<br />

glide), E = 1 × 10 11 Pa & ν = 0.3<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 80 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle under continental lithosphere:<br />

ηCONT = 4 × 10 19 P a · s<br />

12. Viscosity <strong>of</strong> <strong>the</strong> upper mantle beneath cratons: ηCRAT = 5 × 10 20 P a · s<br />

13. Coefficient <strong>of</strong> friction in continental dashpots: kCONT = 7.98 × 10 25 kg/s<br />

14. Coefficient <strong>of</strong> friction in dashpots under cratons: kCRAT = 1.31 × 10 27 kg/s<br />

15. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

dsm viscoelasticrat (Fig. 2.10, p. 19):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 5 × 10 13 s (approx. 1.5 million years)


5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton<br />

8. Rheology (identical for cratons and continental lithosphere): viscoelastic with<br />

E = 1 × 10 11 Pa, ν = 0.3 & ηLIT H = 10 23 P a · s<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

dsm elastoplasticrat (Fig. 2.11, p. 20):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

69


70 Appendix B<br />

7. East European platform and Siberian craton<br />

8. Rheology (identical for cratons and continental lithosphere): elastoplastic with<br />

E = 1 × 10 11 Pa & ν = 0.3 and <strong>the</strong> yield stress σY = 100 MPa<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

dsm elastoviscoplasticrat (Fig. 2.12, p. 21):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton


8. Rheology (identical for cratons and continental lithosphere): viscoelastic with<br />

a plastic yield stress <strong>of</strong> σY = 100 MPa (E = 1 × 10 11 Pa, ν = 0.3 & ηLIT H =<br />

10 23 P a · s<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 9.08 × 10 17 N (line force: 2.5 × 10 12 N/m)<br />

(e) FCC due to Arabia: 3.55 × 10 17 N (line force: 1.67 × 10 12 N/m)<br />

(f) FCC due to Africa: 1.74 × 10 17 N (line force: 0.56 × 10 12 N/m)<br />

(g) FSU line force: 3 × 10 12 N/m<br />

ksm elastoplasticrat (Fig. 2.13, p. 22):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton<br />

8. Rheology (identical for cratons and continental lithosphere): elastoplastic with<br />

E = 1 × 10 11 Pa & ν = 0.3 and <strong>the</strong> yield stress σY = 100 MPa<br />

71


72 Appendix B<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km<br />

10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

14. Kinematic boundary constraints:<br />

(a) average collisional velocity <strong>of</strong> <strong>the</strong> Indian plate: 1.43 × 10 −9 m/s<br />

(b) average collisional velocity <strong>of</strong> <strong>the</strong> Arabian plate: 9.51 × 10 −10 m/s<br />

(c) average collisional velocity <strong>of</strong> <strong>the</strong> African plate: 3.17 × 10 −10 m/s<br />

(d) velocity due to trench suction: 2.91 × 10 −10 m/s<br />

ydsm elastoplasticrat0 (Fig. 2.14, p. 23):<br />

1. Rectangle length: 16000 km (60 elements)<br />

2. Rectangle width: 8000 km (30 elements)<br />

3. Area <strong>of</strong> ABAQUS-element: 7.1 × 10 10 m 2<br />

4. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

5. Minimal time step: 1 × 10 5 s<br />

6. Maximal time step: 1 × 10 11 s<br />

7. East European platform and Siberian craton<br />

8. Rheology (identical for cratons and continental lithosphere): elastoplastic with<br />

E = 1 × 10 11 Pa & ν = 0.3 and <strong>the</strong> yield stress σY = 100 MPa<br />

9. Thickness <strong>of</strong> continental lithosphere: hL = 60 km


10. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle (at 100 km depth, <strong>the</strong> average thickness <strong>of</strong><br />

oceanic, continental and cratonic lithosphere): ηCONT = 4 × 10 19 P a · s<br />

12. Coefficient <strong>of</strong> friction in dashpots: kCONT = 9.86 × 10 25 kg/s<br />

13. Dynamic boundary constraints:<br />

(a) FRP along atlantic continental margin: 1.34 × 10 18 N (line force for 100<br />

million year old oceanic crust: 4.841 × 10 12 N/m)<br />

(b) FRP along arctic continental margin: 6.31 × 10 17 N (line force for 50<br />

million year old oceanic crust: 2.421 × 10 12 N/m)<br />

(c) FCM opposing ridge push: 10 12 N/m<br />

(d) FCC due to India: 3 × 10 18 N<br />

(e) FCC due to Arabia: 2 × 10 18 N<br />

(f) FCC due to Africa: 1.5 × 10 18 N<br />

(g) FSU line force: 3 × 10 12 N/m<br />

eurasia27-01-04 6 (Figs. 3.2 to 3.5 & Fig. 3.8, p. 31 et sqq.):<br />

1. No. <strong>of</strong> elemements in continental lithosphere: 3723<br />

2. Area <strong>of</strong> ABAQUS-element: 1.719 × 10 10 m 2<br />

3. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

4. Minimal time step: 1 × 10 5 s<br />

5. Maximal time step: 1 × 10 11 s<br />

6. East European platform and Siberian craton<br />

7. Rheology (identical for cratons and continental lithosphere): elastoplastic with<br />

E = 1 × 10 11 Pa & ν = 0.25 and <strong>the</strong> yield stress σY = 100 MPa<br />

8. Thickness <strong>of</strong> continental lithosphere: hL = 100 km<br />

9. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 200 km<br />

10. Thickness <strong>of</strong> lithosphere between cratons: hINT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle under continental lithosphere:<br />

ηCONT = 4 × 10 19 P a · s<br />

73


74 Appendix B<br />

12. Viscosity <strong>of</strong> <strong>the</strong> upper mantle beneath cratons: ηCRAT = 5 × 10 20 P a · s<br />

13. Viscosity <strong>of</strong> <strong>the</strong> upper mantle at 150 km depth: ηINT = 2.5 × 10 20 P a · s<br />

14. Coefficient <strong>of</strong> friction in continental dashpots: kCONT = 2.627 × 10 25 kg/s<br />

15. Coefficient <strong>of</strong> friction in dashpots under cratons: kCRAT = 3.695 × 10 26 kg/s<br />

16. Coefficient <strong>of</strong> friction in dashpots between cratons: kINT = 1.580 × 10 26 kg/s<br />

17. Dynamic boundary constraints:<br />

(a) ridge push field (DLOAD) along <strong>the</strong> atlantic & arctic continental margins<br />

(line forces for 100 × 10 6 , 75 × 10 6 , 55 × 10 6 & 50 × 10 6 years old oceanic<br />

lithosphere, averaged over <strong>the</strong> corresponding lithospheric thicknesses and<br />

acting perpendicular to <strong>the</strong> margin)<br />

(b) continental margin line forces opposing ridge push: 10 12 N/m (also averaged<br />

over <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> oceanic lithosphere)<br />

(c) collisional forces scaled with velocity ranging from 2 × 10 18 N (for a velocity<br />

<strong>of</strong> 5.25 cm/year) in <strong>the</strong> Eastern Himalayas to 5.48 × 10 16 N (for a<br />

velocity <strong>of</strong> 0.39 cm/year) at Gibraltar (CLOADs); <strong>the</strong> scaling follows <strong>the</strong><br />

rule (strainrate) ∝ (stress) 3 , i.e. <strong>the</strong> stresses scale as <strong>the</strong> velocities to<br />

<strong>the</strong> 3rd root<br />

(d) collisional forces due to <strong>the</strong> North American plate in Siberia<br />

(e) trench suction scaled with subduction velocity (taking <strong>the</strong> average velocity<br />

<strong>of</strong> 8.29 cm/year to correspond to a line force <strong>of</strong> 3 × 10 12 N/m) and<br />

averaged over <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> continental lithosphere (DLOADs)<br />

(f) trench suction in <strong>the</strong> Aegean Sea (DLOAD)<br />

(g) free segments:<br />

i. <strong>the</strong> stretch between Gibraltar and <strong>the</strong> continental margin<br />

ii. <strong>the</strong> line linking Burma and Bhutan<br />

iii. <strong>the</strong> coast <strong>of</strong> Pakistan<br />

iv. part <strong>of</strong> <strong>the</strong> line linking Pakistan’s shore with Nepal<br />

eurasia01-03-04 1 (Figs. 3.10 to 3.13, p. 41 et sqq.):<br />

1. No. <strong>of</strong> elemements in continental lithosphere: 3723<br />

2. Area <strong>of</strong> ABAQUS-element: 1.719 × 10 10 m 2<br />

3. Computation time: 5 × 10 13 s (approx. 1.5 million years)


4. Minimal time step: 1 × 10 5 s<br />

5. Maximal time step: 1 × 10 11 s<br />

6. East European platform and Siberian craton<br />

7. Rheology (identical for cratons and continental lithosphere): viscoelastic with<br />

E = 1 × 10 11 Pa & ν = 0.25 and a power law (n = 3 & A = 10 −39 P a −3 s −1 )<br />

describing viscosity<br />

8. Thickness <strong>of</strong> continental lithosphere: hL = 100 km<br />

9. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 200 km<br />

10. Thickness <strong>of</strong> lithosphere between cratons: hINT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle under continental lithosphere:<br />

ηCONT = 4 × 10 19 P a · s<br />

12. Viscosity <strong>of</strong> <strong>the</strong> upper mantle beneath cratons: ηCRAT = 5 × 10 20 P a · s<br />

13. Viscosity <strong>of</strong> <strong>the</strong> upper mantle at 150 km depth: ηINT = 2.5 × 10 20 P a · s<br />

14. Coefficient <strong>of</strong> friction in continental dashpots: kCONT = 2.627 × 10 25 kg/s<br />

15. Coefficient <strong>of</strong> friction in dashpots under cratons: kCRAT = 3.695 × 10 26 kg/s<br />

16. Coefficient <strong>of</strong> friction in dashpots between cratons: kINT = 1.580 × 10 26 kg/s<br />

17. Dynamic boundary constraints:<br />

(a) ridge push field (DLOAD) along <strong>the</strong> atlantic & arctic continental margins<br />

(line forces for 100 × 10 6 , 75 × 10 6 , 55 × 10 6 & 50 × 10 6 years old oceanic<br />

lithosphere, averaged over <strong>the</strong> corresponding lithospheric thicknesses and<br />

acting perpendicular to <strong>the</strong> margin)<br />

(b) continental margin line forces opposing ridge push: 10 12 N/m (also averaged<br />

over <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> oceanic lithosphere)<br />

(c) collisional forces (CLOADs) scaled with current velocities along <strong>the</strong> borders<br />

with Africa and Arabia; <strong>the</strong> scaling follows <strong>the</strong> rule (strainrate) ∝<br />

(stress) 3 , i.e. <strong>the</strong> stresses scale as <strong>the</strong> velocities to <strong>the</strong> 3rd root<br />

(d) collisional forces (CLOADs) scaled as above but corresponding to a relative<br />

velocity <strong>of</strong> 15 cm/yr in India<br />

(e) free segments:<br />

i. <strong>the</strong> stretch between Gibraltar and <strong>the</strong> continental margin<br />

ii. <strong>the</strong> line linking Burma and Bhutan<br />

iii. <strong>the</strong> coast <strong>of</strong> Pakistan<br />

75


76 Appendix B<br />

iv. part <strong>of</strong> <strong>the</strong> line linking Pakistan’s shore with Nepal<br />

v. <strong>the</strong> boundary with North America running across Siberia<br />

vi. <strong>the</strong> trenches in <strong>the</strong> Far East and <strong>the</strong> Aegean Sea<br />

eurasia02-02-04 1 (Fig. 3.14, p. 46):<br />

1. No. <strong>of</strong> elemements in continental lithosphere: 3723<br />

2. Area <strong>of</strong> ABAQUS-element: 1.719 × 10 10 m 2<br />

3. Computation time: 5 × 10 13 s (approx. 1.5 million years)<br />

4. Minimal time step: 1 × 10 5 s<br />

5. Maximal time step: 1 × 10 11 s<br />

6. East European platform and Siberian craton<br />

7. Rheology (identical for cratons and continental lithosphere): elastoplastic with<br />

E = 1 × 10 11 Pa & ν = 0.25 and <strong>the</strong> yield stress σY = 100 MPa, followed by<br />

a power law (n = 3) for dislocation creep above σY<br />

8. Thickness <strong>of</strong> continental lithosphere: hL = 100 km<br />

9. Thickness <strong>of</strong> lithosphere within cratons: hCRAT = 200 km<br />

10. Thickness <strong>of</strong> lithosphere between cratons: hINT = 150 km<br />

11. Viscosity <strong>of</strong> <strong>the</strong> upper mantle under continental lithosphere:<br />

ηCONT = 4 × 10 19 P a · s<br />

12. Viscosity <strong>of</strong> <strong>the</strong> upper mantle beneath cratons: ηCRAT = 5 × 10 20 P a · s<br />

13. Viscosity <strong>of</strong> <strong>the</strong> upper mantle at 150 km depth: ηINT = 2.5 × 10 20 P a · s<br />

14. Coefficient <strong>of</strong> friction in continental dashpots: kCONT = 2.627 × 10 25 kg/s<br />

15. Coefficient <strong>of</strong> friction in dashpots under cratons: kCRAT = 3.695 × 10 26 kg/s<br />

16. Coefficient <strong>of</strong> friction in dashpots between cratons: kINT = 1.580 × 10 26 kg/s<br />

17. Dynamic boundary constraints:<br />

(a) ridge push field (DLOAD) along <strong>the</strong> atlantic & arctic continental margins<br />

(line forces for 100 × 10 6 , 75 × 10 6 , 55 × 10 6 & 50 × 10 6 years old oceanic<br />

lithosphere, averaged over <strong>the</strong> corresponding lithospheric thicknesses and<br />

acting perpendicular to <strong>the</strong> margin)


(b) continental margin line forces opposing ridge push: 10 12 N/m (also averaged<br />

over <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> oceanic lithosphere)<br />

(c) collisional forces (CLOADs) scaled with current velocities along <strong>the</strong> borders<br />

with Africa and Arabia; <strong>the</strong> scaling follows <strong>the</strong> rule (strainrate) ∝<br />

(stress) 3 , i.e. <strong>the</strong> stresses scale as <strong>the</strong> velocities to <strong>the</strong> 3rd root<br />

(d) collisional forces (CLOADs) scaled as above but corresponding to a relative<br />

velocity <strong>of</strong> 15 cm/yr in India<br />

(e) free segments:<br />

i. <strong>the</strong> stretch between Gibraltar and <strong>the</strong> continental margin<br />

ii. <strong>the</strong> line linking Burma and Bhutan<br />

iii. <strong>the</strong> coast <strong>of</strong> Pakistan<br />

iv. part <strong>of</strong> <strong>the</strong> line linking Pakistan’s shore with Nepal<br />

v. <strong>the</strong> boundary with North America running across Siberia<br />

vi. <strong>the</strong> trenches in <strong>the</strong> Far East and <strong>the</strong> Aegean Sea<br />

77


78 Appendix B


Bibliography<br />

[1] E.A. Abbot. Flatland: a romance <strong>of</strong> many dimensions by A. Square. Number<br />

ISBN 0 14 04.3531 X. Penguin Books Ltd., 1884.<br />

[2] E.V. Artyushkov. Stresses in <strong>the</strong> lithosphere caused by crustal thickness inhomogeneities.<br />

J. Geophys. Res., 78:7675–7708, 1973.<br />

[3] D. Bercovici. The generation <strong>of</strong> plate tectonics from mantel convection. Earth<br />

Planet. Sci. Lett., 205:107–121, 2003.<br />

[4] P. Bird. Testing hypo<strong>the</strong>ses on plate-driving mechanisms with global lithosphere<br />

models including topography, <strong>the</strong>rmal structure and faults. J. Geophys. Res.,<br />

103:10115–10129, 1998.<br />

[5] P. Bird. An updated digital model <strong>of</strong> plate boundaries. Geochem. Geophys.<br />

Geosyst., 4(3), 2003.<br />

[6] W.F. Brace and D.L. Kohlstedt. Limits on lithospheric stress imposed by laboratory<br />

experiments. J. Geophys. Res., 85:6248–6252, 1980.<br />

[7] D.D. Coblentz, M. Sandiford, R.M. Richardson, S. Zhou, and R. Hillis. The<br />

origins <strong>of</strong> <strong>the</strong> intraplate stress field in continental Australia. Earth Planet. Sci.<br />

Lett., 133:299–309, 1995.<br />

[8] C.P. Conrad and C. Lithgow-Bertelloni. How mantle slabs drive plate tectonics.<br />

Science, 298:207–209, 2002.<br />

[9] R.P. Crease. The prism and <strong>the</strong> pendulum: <strong>the</strong> ten most beautiful experiments<br />

in science. Number ISBN 1-4000-6131-8. Random House, 2003.<br />

[10] C. DeMets, R.G. Gordon, D.F. Argus, and S. Stein. Current plate motions.<br />

Geophys. J. Int., 101:425–478, 1990.<br />

[11] D. Forsyth and S. Uyeda. On <strong>the</strong> relative importance <strong>of</strong> <strong>the</strong> driving forces <strong>of</strong><br />

plate motion. Geophys. J. R. astr. Soc., 43:163–200, 1975.<br />

[12] C. Gaina, W.R. Roest, and R.D. Müller. Late Cretaceous-Cenozoic deformation<br />

<strong>of</strong> nor<strong>the</strong>ast Asia. Earth Planet. Sci. Lett., 197:273–286, 2002.<br />

79


80 BIBLIOGRAPHY<br />

[13] The Global Seismic Hazard Assessment Program. Global Seismic Hazard Map.<br />

Available online at seismo.ethz.ch/GSHAP/, 1999.<br />

[14] S. Goes, J.J.P. Loohuis, M.J.R. Wortel, and R. Govers. The effect <strong>of</strong> plate<br />

stresses and shallow mantle temperatures on tectonics <strong>of</strong> northwestern Europe.<br />

Global Planet. Change, 27:23–38, 2000.<br />

[15] M. Gölke and D. Coblentz. Origins <strong>of</strong> <strong>the</strong> European regional stress field.<br />

Tectonophysics, 266:11–24, 1996.<br />

[16] G. Grünthal and D. Stromeyer. The recent crustal stress field in central Europe:<br />

trajectories and finite element modellling. J. Geophys. Res., 97:11805–11820,<br />

1992.<br />

[17] L. Han and M. Gurnis. How valid are dynamic models <strong>of</strong> subduction and<br />

convection when plate motions are prescribed? Phys. Earth Planet. Inter.,<br />

110:235–246, 1999.<br />

[18] Hibbit, Karlsson, and Sorenson Inc. ABAQUS/Standard, version 6.3, 2002.<br />

[19] M.P. Hochstein and K. Regenauer-Lieb. Heat generation associated with <strong>the</strong><br />

collision <strong>of</strong> two plates: The Himalya Geo<strong>the</strong>rmal Belt. J. Volc. Geo<strong>the</strong>rm. Res.,<br />

83:75–92, 1998.<br />

[20] S.H. Kirby and A.K. Kronenburg. Rheology <strong>of</strong> <strong>the</strong> lithosphere: selected topics.<br />

Rev. Geophys., 25:1219–1244, 1987.<br />

[21] C. Kreemer, W.E. Holt, and A.J. Haines. An integrated global model <strong>of</strong> presentday<br />

plate motions and plate boundary deformation. Geophys. J. Int., 154:8–34,<br />

2003.<br />

[22] C. Lithgow-Bertelloni and J.H. Guynn. Origin <strong>of</strong> <strong>the</strong> lithospheric stress field.<br />

J. Geophys. Res., 109:B014008, doi:10.1029/2003JB002467, 2004.<br />

[23] C. Lithgow-Bertelloni and M.A. Richards. The dynamics <strong>of</strong> Cenozoic and Mesozoic<br />

plate motions. Rev. Geophys., 36:27–78, 1998.<br />

[24] J.J.P. Loohuis, M.J.R. Wortel, and P.Th. Meijer. A first order stress model for<br />

<strong>the</strong> <strong>Eurasian</strong> plate with an assessment <strong>of</strong> <strong>the</strong> role <strong>of</strong> <strong>the</strong> Alpine collision zone.<br />

1998.<br />

[25] The MacNeal-Schwendler Corporation. MSC/Patran, version 7, 1997.<br />

[26] B. Müller, M.L. Zoback, K. Fuchs, L. Mastin, S. Gregersen, N. Pavoni,<br />

O. Stephansson, and C. Ljunggren. Regional patterns <strong>of</strong> tectonic stress in<br />

Europe. J. Geophys. Res., 97:11783–11803, 1992.<br />

[27] R.D. Müller, W.R. Roest, J.-Y. Royer, L.M. Gahagan, and J.G. Sclater. Digital<br />

isochrons <strong>of</strong> <strong>the</strong> world’s ocean floor. J. Geophys. Res., 102:3211–3214, 1997.


BIBLIOGRAPHY 81<br />

[28] K. Regenauer-Lieb and J.-P. Petit. Cutting <strong>of</strong> <strong>the</strong> European continental lithosphere:<br />

plasticity <strong>the</strong>ory applied to <strong>the</strong> present Alpine collision. J. Geophys.<br />

Res., 102:7713–7748, 1997.<br />

[29] K. Regenauer-Lieb and D.A. Yuen. Modeling shear zones in geological and<br />

planetary sciences: solid- and fluid-<strong>the</strong>rmal-mechanical approaches. Earth Sci.<br />

Rev., 63:295–349, 2003.<br />

[30] J. Reinecker, O. Heidbach, and B. Mueller. The 2003 release <strong>of</strong> <strong>the</strong> World Stress<br />

Map. Available online at www.world-stress-map.org, 2003.<br />

[31] Schlumberger. The Oilfield Glossary: Where <strong>the</strong> Oil Field Meets <strong>the</strong> Dictionary.<br />

Available online at www.glossary.oilfield.slb.com, 2004.<br />

[32] C.H. Scholz and J. Campos. On <strong>the</strong> mechanism <strong>of</strong> seismic decoupling and back<br />

arc spreading at subduction zones. J. Geophys. Res., 100:22103 – 22115, 1995.<br />

[33] D.L. Turcotte and G. Schubert. Geodynamics, applications <strong>of</strong> continuum<br />

physics to geological problems. Number ISBN 0-521-66186-2. Cambridge University<br />

Press, 2002.<br />

[34] A. Villaseñor, M.H. Ritzwoller, A.L. Levshin, M.P. Barmin, E.R. Engdahl,<br />

W. Spakman, and J. Trampert. Shear velocity structure <strong>of</strong> central Eurasia from<br />

inversion <strong>of</strong> surface wave velocities. Phys. Earth Planet. Sci. Int., 123:169–184,<br />

2001.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!