a comparative study of maxwell viscoelasticity at large strains … · only models that use the...

32
A comparative study of Maxwell viscoelasticity at large strains and rotations 1 2 Christoph E. Schrank *1, 2 , Ali Karrech 3 , David A. Boutelier 4 , Klaus Regenauer-Lieb 5 3 4 1) Queensland University of Technology, School of Earth, Environmental and Biological Sciences, 5 Brisbane, QLD, Australia 6 2) University of Western Australia, School of Earth and Environment, Crawley, WA, Australia 7 3) University of Western Australia, School of Civil, Environmental and Mining Engineering, 8 Crawley, WA, Australia 9 4) University of Newcastle, School of Environmental and Life Sciences, Callaghan, NSW, 10 Australia 11 5) University of New South Wales, School of Petroleum Engineering, Sydney, NSW, Australia 12 13 History 14 Accepted date: XX/YY/2017. Received date: XX/YY/2017; in original form date: XX/YY/2017 15 *Corresponding author: e-mail: [email protected], tel.: (+61 7) 3138 1583, ORCID: 16 0000-0001-9643-2382 17 18

Upload: others

Post on 04-Nov-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

A comparative study of Maxwell viscoelasticity at large strains and rotations 1

2

Christoph E. Schrank*1, 2, Ali Karrech3, David A. Boutelier4, Klaus Regenauer-Lieb5 3

4

1) Queensland University of Technology, School of Earth, Environmental and Biological Sciences, 5

Brisbane, QLD, Australia 6

2) University of Western Australia, School of Earth and Environment, Crawley, WA, Australia 7

3) University of Western Australia, School of Civil, Environmental and Mining Engineering, 8

Crawley, WA, Australia 9

4) University of Newcastle, School of Environmental and Life Sciences, Callaghan, NSW, 10

Australia 11

5) University of New South Wales, School of Petroleum Engineering, Sydney, NSW, Australia 12

13

History 14

Accepted date: XX/YY/2017. Received date: XX/YY/2017; in original form date: XX/YY/2017 15

*Corresponding author: e-mail: [email protected], tel.: (+61 7)3138 1583, ORCID: 16

0000-0001-9643-2382 17

18

Page 2: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Abbreviated title: Viscoelasticity at large transformations 19

Summary 20

We present a new Eulerian large-strain model for Maxwell viscoelasticity using a logarithmic co-21

rotational stress rate and the Hencky strain tensor. This model is compared to the small-strain model 22

without co-rotational terms and a formulation using the Jaumann stress rate. Homogeneous 23

isothermal simple shear is examined for Weissenberg numbers in the interval [0.1;10]. Significant 24

differences in shear-stress and energy evolution occur at Weissenberg numbers > 0.1 and shear 25

strains > 0.5. In this parameter range, the Maxwell-Jaumann model dissipates elastic energy 26

erroneously and thus should not be used. The small-strain model ignores finite transformations, 27

frame indifference, and self-consistency. As a result, it overestimates shear stresses compared to the 28

new model and entails significant errors in the energy budget. Our large-strain model provides an 29

energetically consistent approach to simulating non-coaxial viscoelastic deformation at large strains 30

and rotations. 31

Keywords 32

Fault zone rheology, creep and deformation, rheology: crust and lithosphere, rheology and friction 33

of fault zones 34

1. Introduction 35

The Maxwell body of a linear-elastic Hookean spring in series with a Newtonian dashpot is the 36

simplest rheological model for geological deformation (e.g., Bailey (2006)). It has been employed 37

to describe many deformation processes and structures, for example mantle convection (Beuchert 38

and Podladchikov, 2010, Mühlhaus and Regenauer-Lieb, 2005), Rayleigh-Taylor instabilities (Kaus 39

and Becker, 2007, Robin and Bailey, 2009), lithospheric deformation (Kaus and Podladchikov, 40

2006, Korenaga, 2007, Kusznir and Bott, 1977), postglacial rebound (Larsen et al., 2005, Peltier, 41

1976), shear zones (Branlund et al., 2000, Mancktelow, 2002, Sone and Uchide, 2016), and folds 42

(Liu et al., 2016, Mancktelow, 1999, Zhang et al., 2000). In spite of their simplicity, current 43

Page 3: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Maxwell models display well-known errors (e.g., Beuchert and Podladchikov, 2010, Mühlhaus and 44

Regenauer-Lieb, 2005) when the associated strains and, importantly, rotations are large. Such 45

conditions are met, for example, during plate motions (Tohver et al., 2010), in bending subducting 46

plates (Conrad and Hager, 1999), blocks of crust bounded by faults (Abels and Bischoff, 1999), 47

folds (Ramsay and Huber, 1987), shear zones (Ramsay, 1980), boudins (Mandal and Khan, 1991), 48

porphyroblasts (Johnson, 2009) and –clasts (Passchier and Simpson, 1986), and grains deforming 49

by diffusion creep (Wheeler, 2010). Large rotations pose a mathematical challenge when elasticity 50

is considered in the rheology of highly transformed materials (Dienes, 1979, Karrech et al., 2012, 51

Karrech et al., 2011, Moss, 1984, Bruhns et al., 2001a, Bruhns et al., 1999, Bruhns et al., 2001b, 52

Meyers et al., 2005, Xiao et al., 1997a, Xiao et al., 2006, Xiao et al., 1998a, Xiao et al., 1998b, 53

Colak, 2004, Lehmann, 1972a): one requires an objective formulation of the stress rate (time 54

derivative of stress). 55

If a volume element of a transforming body incurs a rigid-body rotation over an 56

infinitesimal time increment, its state of stress and strain should not change relative to a coordinate 57

system that co-rotates with this volume element. In other words, the notion of objectivity, or frame 58

indifference, requires that the co-rotational stress and strain increments become zero during an 59

instantaneous rigid-body rotation (see, for example, Chapter 1.6 of Chakrabarty (2006)). In an 60

Eulerian (stationary) reference frame, the objectivity requirement is commonly addressed with a so-61

called co-rotational objective stress rate, which contains spins that account for elastic rotations (for 62

an in-depth review, see Xiao et al., 2006). Here, we compare the self-consistent formulation of the 63

Eulerian finite-transformation model for Maxwell viscoelasticity (called FT model hereafter) to two 64

thermodynamically non-consistent Maxwell models, which are commonly used to model 65

geologically relevant strains (e.g., Kaus and Becker, 2007, Beuchert and Podladchikov, 2010, 66

Mancktelow, 1999, Zhang et al., 2000, Moresi et al., 2002): the small-strain formulation without 67

co-rotational terms, referred to as SS model, and the formulation using the Jaumann stress rate, 68

called MJ model hereafter. Self-consistency (i.e., adherence to thermodynamic principles) in the FT 69

Page 4: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

model is achieved by using the logarithmic spin introduced by Xiao and co-workers in the context 70

of finite elastoplasticity (Bruhns, 2003, Bruhns et al., 1999, Xiao et al., 1997a, Xiao et al., 2006, 71

Xiao et al., 1998a, Xiao et al., 1998b). 72

The choice of the logarithmic spin in the new FT model is motivated by two reasons. First, 73

the logarithmic co-rotational rate of the Hencky strain was proven to be the only strain-rate 74

measure, which is always identical with the stretching (Bruhns et al., 1999, Xiao et al., 1997a, Xiao 75

et al., 1997b). The stretching, the symmetric part of the velocity gradient tensor, is the most 76

commonly used and kinematically most appealing measure of deformation rates (cf. chapter 2 of 77

Xiao et al., 1997b). Second, the logarithmic stress rate is the only choice, which delivers a self-78

consistent Eulerian constitutive model (Bruhns et al., 1999). Bruhns et al. (1999) prove self-79

consistency in the co-rotational Eulerian rate model through integrating the additive deformation-80

rate decomposition for the case of purely elastic stretching. If the rate formulation of the model is 81

exactly integrable to deliver a truly elastic response, i.e., fully reversible deformation without 82

energy dissipation, then the Eulerian finite-strain model is defined as (thermodynamically) self-83

consistent. Only models that use the logarithmic spin are self-consistent. 84

To analyse the relevance of the co-rotational logarithmic spin used in the FT model, 85

homogeneous 2D isothermal simple shear is examined because of its non-zero, skew-symmetric 86

spin and high conceptual importance for understanding shear zones (e.g., Ramsay, 1980, Ramsay 87

and Graham, 1970). The relative stress and energy states are evaluated for a shear-strain range of 0 88

to 10, providing the basis for comparison of the SS, MJ and FT models. The sections to come are 89

organised as follows: first, the need for a proper finite Maxwell model is established. Second, 90

analytical solutions for simple shear with the SS and MJ models are presented. Third, the FT model 91

is derived and its numerical solution described. Fourth, all three models are compared. Finally, the 92

results and their implications are discussed. 93

2. Maxwell viscoelasticity, co-rotational stress rates, and the need for a finite Maxwell model 94

The constitutive equation for Maxwell viscoelasticity of an incompressible material is: 95

Page 5: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

96

Dij =12Lij + L

Tij( ) =

! ʹσ ij

2G+

ʹσ ij

2η (1), 97

98

whereDij is the stretching or deformation-rate tensor, defined as the symmetric part of the velocity-99

gradient tensor Lij = ∂vi / ∂x j with velocities vi and directions xi, the superscript T marks the matrix 100

transpose, the overdot denotes a derivative with respect to time t, G is the shear modulus of an 101

isotropic linear-elastic material, ʹσ ij is the deviator (indicated by the dash) of the Cauchy stress 102

tensor σ ij , and η is the Newtonian viscosity. In the limit of infinitesimal deformations, the 103

stretching tensor becomes the familiar strain-rate tensor for small strains. Equation (1) states that 104

elastic (first term) and viscous deformation rates (second term) are additive (serial coupling of an 105

elastic spring with a viscous dashpot). The term ʹ!σ ij is the stress rate, which highlights the need for 106

a co-rotational formulation in order to acknowledge rotations at finite strain. Formulations that 107

include geometrically admissible co-rotational terms are called “objective”. 108

Equation (1) is solved by direct integration for the SS model while ʹ!σ ij is replaced with rates 109

augmented by specific spins in the MJ and FT models (section 3). It is obvious that the SS model 110

ignores large rotations and strains and hence is not geometrically objective and not 111

thermodynamically self-consistent. Errors are expected at large non-coaxial deformations. The 112

magnitude of these errors can be estimated by comparing the SS model with other existing models 113

such as MJ and now FT. The MJ model employs a classical objective co-rotational stress rate, the 114

Jaumann rate (Jaumann, 1911). 115

Two important shortcomings of the MJ model under simple shear have been recognized in 116

the geodynamic literature previously (Mühlhaus and Regenauer-Lieb, 2005, Beuchert and 117

Podladchikov, 2010). To summarise these findings, we first recall the Maxwell relaxation time trel 118

and the Weissenberg number W: 119

Page 6: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

120

(2a, b), 121

122

where !γ is a characteristic shear-strain rate, which in our study corresponds to the imposed simple-123

shear rate defined in Eq. (4). Deformation observed at time scales well below trel is dominated by 124

elastic contributions to the strain rate and stresses. After about five trel , deformation is essentially 125

viscous. The Weissenberg number W measures to which degree the material behavior is nonlinear 126

and how much of the current strain is recoverable (White, 1964, Dealy, 2010). Mühlhaus and 127

Regenauer-Lieb (2005) compared Maxwell models using the Jaumann and Green-Naghdi rates, two 128

objective stress rates from the continuum-mechanics community, with the SS model frequently used 129

in geological modeling. While these objective formulations contain the necessary geometric terms 130

that ensure frame indifference, they do not satisfy the self-consistency condition of Bruhns et al. 131

(1999). Consequently, for the same material properties, the Green-Naghdi and MJ models give 132

different stress-strain curves at shear strains > 1. The SS model (used in geology) and the Green-133

Naghdi model (used in some engineering applications, e.g., for car-crash simulations 134

(Abaqus/Explicit, 2015)) produce qualitatively similar results but the MJ model (used in other 135

engineering application, e.g., for calculations of the nonlinear creep of the steel carapace of a 136

nuclear reactor under elevated temperatures (Abaqus/Standard, 2009)) exhibits spurious softening 137

during numerical simulations of simple shear for W = 1. Mühlhaus and Regenauer-Lieb (2005) 138

showed, however, that this effect vanishes if non-linear creep and/or plasticity are added to the 139

constitutive model. Later, Beuchert and Podladchikov (2010) demonstrated numerically that the MJ 140

model exhibits unphysical shear-stress oscillations at W = 10. They also showed that the choice, and 141

even the neglect, of co-rotational terms become irrelevant at W ≤ 0.1. Therefore, two regimes of 142

intermediate (0.1 < W < 1) and high (W ≥ 1) Weissenberg numbers exist, in which one can expect 143

trel =ηG

W = γtrel

Page 7: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

problems with the SS and MJ models. The question arises: where in the Earth may the intermediate 144

and high-W regimes occur? 145

Beuchert and Podladchikov (2010) and Mühlhaus and Regenauer-Lieb (2005) aiming for 146

modeling viscoelastic mantle convection argued that only the lithosphere exhibits sections with W 147

>> 0.1. At the tectonic time and length scales of interest to these studies, the lithosphere behaves 148

essentially elastic and hardly deforms. Hence, the problems of the MJ model, spurious softening 149

and stress oscillations, were predicted to be irrelevant in viscoelastic mantle simulations. This 150

argument is further supported by the notion that power-law flow and plasticity, both common in the 151

lithosphere (e.g., Karato, 2008), serve as stress limiters, which inhibit spurious softening and 152

oscillations (Beuchert and Podladchikov, 2010). 153

However, we argue that deformation problems within the lithosphere, in particular at length 154

scales ≤ 1 km, require the tackling of the problem of modeling Maxwell viscoelasticity in the 155

intermediate and high-W regimes properly. The shear moduli of most rock-forming minerals are 156

within the range of 10 to 100 GPa (Bass, 1995). The effective viscosity of lithospheric rocks varies 157

over several orders of magnitude due to its sensitivity to, e.g., material composition, temperature, 158

pressure, strain rate, stress, and grain size (e.g., Karato, 2008). We adopt lower and upper bounds of 159

1018 and 1025 Pas, respectively. Shear-strain rates also cover several orders of magnitude. A typical 160

lower bound for natural shear zones is 10-15 s-1 (Biermeier and Stüwe, 2003, Pfiffner and Ramsay, 161

1982). Estimates for the upper bound in nature are of order 10-10 s-1 for shear zones with widths ≤ 162

100 m (White and Mawer, 1992). Computing W for the above parameter ranges shows that a 163

significant subset (~ 35% to 50%) of this parameter space is within the critical regimes of W (Fig. 164

1). Moreover, one can envisage geological scenarios during which the internal deformation of the 165

lithosphere remains elastic, or transient viscoelastic, in large parts while the associated rotations 166

become large, for example, in subducting plates and blocks of crust rotating about a vertical axis. In 167

these cases, a proper objective treatment of large transformations, and elastic rotations in particular, 168

Page 8: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

seems desirable. This effort should also involve stress advection (e.g., Truesdell, 1956), which is 169

irrelevant in the simple-shear case at hand. 170

171

3. Derivations: kinematics of 2D simple shear 172

We consider 2D plane-strain isothermal simple shear of a unit square, where x1 and x2 denote the 173

horizontal and vertical direction, respectively. The lower-left corner of the square is centered on the 174

origin of the coordinate system. Shear is parallel to x1. The boundary conditions are: 175

176

v1 x2 = 0( ) = v2 x2 = 0( ) = 0v1 x2 =1( ) = const. > 0v2 x2 =1( ) = 0∂v1∂x2

0 < x2 <1( ) = const. > 0

(3), 177

178

where vi denote the velocities in the respective directions. Thus, the velocity-gradient tensor Lij has 179

only one non-zero, constant entry: 180

181

L12 =∂v1∂x2

= !γ (4), 182

183

where is the constant shear-strain rate applied to the unit square. Hence, the deformation-rate 184

tensor Dij and the spin tensor Sij , the antisymmetric component of Lij , read: 185

186

γ

Page 9: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Dij =0!γ2

!γ2

0

⎢⎢⎢⎢

⎥⎥⎥⎥

Sij =12Lij − L

Tij( ) =

0!γ2

−!γ2

0

⎢⎢⎢⎢

⎥⎥⎥⎥

(5a, b). 187

188

Equation (5b) expresses the internal rotations inherent to simple shear. Conceptually, the simple-189

shear model embodies an idealized plane-strain shear zone with homogeneous strain distribution. In 190

nature, shear zones generally display a heterogeneous strain pattern (Ramsay, 1980, Ramsay and 191

Graham, 1970). However, this heterogeneity is a function of the scale of observation. One can 192

identify length scales, at which deformation is, to first order, homogeneous. Hence, one may 193

consider the following analysis as an examination of a homogeneously deforming representative 194

elementary volume within an ideal shear zone, across which variations in parameters such as 195

temperature, composition, viscosity, fluid content, etc., are negligible. 196

3.1 Derivations: Maxwell model for small strain (SS model) 197

The SS model is solved by integrating Eq. (1) after substituting simple-shear kinematics (Eqs. (3), 198

(5a)), trel (Eq. (2a)), W (Eq. (2b)), and dimensionless time k = trel−1t , which yields: 199

200

σ12 k( ) =WG 1− e−k( )σ11 =σ 22 = 0

(6a, b). 201

202

The total deformational power per unit volume is: 203

204

!ETOT t( ) = ʹσ ij :Dij =σ12 !γ (7). 205

206

Page 10: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Integrating Eq. (7) over time with the initial condition ETOT t = 0( ) = 0 gives the total energy per unit 207

volume: 208

209

ETOT k( ) =W 2G k + e−k( )−1⎡⎣

⎤⎦ (8). 210

211

To decompose the total energy into elastic and viscous fractions, we use the additivity of the 212

deformation rates to introduce the elastic and viscous fraction of strain rate (e.g., Kaus and Becker, 213

2007): 214

215

1=!γel!γ+!γ vis!γ= fel + fvis (9), 216

217

where fel and fvis denote the elastic (subscript el) and viscous (subscript vis) fraction of the strain 218

rate, respectively. At infinite time, the shear stress is purely viscous: σ ∞vis =σ12 k→∞( ) =η !γ =WG . 219

Thus, the viscous and elastic stress fractions can be defined as: 220

221

fvis =σ12 k( )σ vis

∞=1− e−k

fel =1− fvis = e−k

(10a, b). 222

223

Inserting Eqs. (6), (9) and (10) into Eq. (7) and integrating over time yield the stored and viscous 224

energy per unit volume, again for zero initial energies: 225

226

Eel = !γ felσ12 dk =W2G 1

2+12e−2k − e−k

⎝⎜

⎠⎟ =

σ122

2G0

k

Evis = !γ fvisσ12 dk =W2G 2e−k − 1

2e−2k + k − 3

2⎛

⎝⎜

⎠⎟

0

k

∫ (11a, b). 227

Page 11: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

228

3.2 Derivations: Maxwell model with the Jaumann stress rate (MJ model) 229

The elastic stress rate in Eq. (1) ( ʹ!σ ij ) is now replaced by the Jaumann stress rate (Jaumann, 1911): 230

231

!τ ij = !σ ij − Sikσ kj +σ ikSkj (12). 232

233

This stress rate is equivalent to a rigid-body rotation of the stress tensor neglecting second and 234

third-order terms (Chakrabarty, 2006). Since in the problem at hand σ11 = −σ 22 ⇒ tr σ ij( ) = 0 (e.g., 235

Durban, 1990), the dashes for deviatoric quantities are dropped in the following.Substituting the 236

spin (Eq. (5b)), Eq. (12) reads: 237

238

!τ ij =!σ11 − !γσ12 !σ12 + !γσ11!σ 21 + !γσ11 !σ 22 + !γσ12

⎣⎢⎢

⎦⎥⎥

(13). 239

240

Inserting Eq. (13) into the constitutive model (Eq. (1)) with the deformation-rate tensor for simple 241

shear (Eq. (5a)) and substituting trel yield the system of coupled differential equations: 242

243

η !γ = trel !σ12 − !γ !σ 22( )+σ120 = trel !σ 22 + !γσ12( )+σ 22

(14). 244

245

The solution of Eq. (14) in terms of W and dimensionless time k is (Appendix 1): 246

σ12 k( ) = WG1+W 2 e−k W sin Wk( )− cos Wk( )⎡⎣ ⎤⎦+1⎡⎣ ⎤⎦

σ 22 k( ) = WG1+W 2 e−k W cos Wk( )+ sin Wk( )⎡⎣ ⎤⎦−W⎡⎣ ⎤⎦

(15a, b). 247

248

Page 12: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

The first fraction in Eq. (15a), b is the steady-state viscous shear stress for the MJ model: 249

σ vis∞ =σ12 k→∞( ) =WG 1+W 2( )

−1. It differs from the steady-state viscous stress of the SS model 250

by the factor 1+W 2( )−1

. This factor highlights an important effect of the stress-rotation terms 251

included in the Jaumann stress rate: they decrease the maximum shear stress attained at steady state. 252

Using Eq. (7) and Eq. (15a), the total deformational energy per unit volume as function of k is: 253

254

ETOT k( ) =σ vis∞W k − e−k

W 2 +1W 2 −1( )cos Wk( )+ 2W sin Wk( )( )+W

2 −1W 2 +1

⎣⎢

⎦⎥(16). 255

256

Eq. (16) is also decomposed into elastic and viscous fractions. The elastic and viscous stress 257

fractions read: 258

259

(17a, b). 260

261

To conclude the energy decomposition, the time integrals of Eq. (11) need to be solved again after 262

substituting the appropriate expressions for the shear stress (Eq. (15a)) and elastic and viscous stress 263

fractions (Eq. (17)). The results are cumbersome because of the lengthy expressions resulting from 264

the integration of the trigonometric terms and are thus not shown here. Many mathematics software 265

packages can conduct this integration, and the interested user may find this more convenient. 266

3.3 Derivations: Maxwell model at finite strain (FT model) 267

The logarithmic finite-strain model requires slight modifications of Eqs. (1) and (12). The Cauchy 268

stress σ ij is replaced by the Kirchhoff stress Σij = Jσ ij where J is the determinant of the 269

deformation-gradient tensor. We use the Kirchhoff stress with more general applications in mind, 270

fel = e−k cos Wk( )−W sin Wk( )( )

fvis = e−k W sin Wk( )− cos Wk( )( )+1

Page 13: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

even though J = 1 for simple shear. The co-rotational Cauchy stress rate !τ ij (Eq. (12)) is replaced by 271

the co-rotational rate of the Kirchhoff stress ⌣Σij : 272

273

⌣Σij = "Σij −ΩikΣkj + ΣikΩkj (18), 274

275

which involves a logarithmic spin of rotation discovered by Xiao and colleagues (Xiao et al., 276

1997a, Xiao et al., 1997b): 277

278

Ωij =Wij +λA + λBλA − λB

+2

lnλA − lnλB

⎝⎜

⎠⎟Pik

A( )DklPljB( )

A,B=1,A≠B

n∑ (19), 279

280

where A and B are dummy integers denoting the principal directions, λA is the Ath eigenvalue of the 281

left Cauchy-Green strain tensor, PikA( ) is its Ath eigenprojection (tensor product of an eigenvector 282

with itself), and n is the number of distinct eigenvalues. The sum vanishes naturally if the 283

eigenvalues are all equal. The spin of Eq. (19) contains a correction term compared to that of the 284

Jaumann formulation, which accounts for the stretching of the eigenprojections, not only their 285

rotation. The logarithmic spin produces a smooth, stable response under simple shear, unlike the 286

Jaumann spin, which produces aberrant oscillations (Lehmann, 1972b, Dienes, 1979). Moreover, 287

since the co-rotational logarithmic rate of the Hencky strain tensor coincides with the deformation-288

rate tensor, the former constitutes an ideal energy conjugate of the Kirchhoff stress (Bruhns, 2003). 289

Therefore, the Maxwell model stated in Eq. (1) involves another modification in the FT 290

model, in addition to inserting the stress rate of Eq. (18): the time derivative of the Eulerian 291

logarithmic Hencky strain tensor replaces the stretching. The associated system of differential 292

equations (Eq. (A9)) and the numerical method for integrating them are detailed in appendix 2. The 293

particular qualities, which render logarithmic kinematics the mathematically most consistent tool 294

Page 14: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

for the constitutive behavior of geomaterials, are further discussed in great depth by Xiao et al. 295

(2006). 296

4. Results 297

We present the evolution of normalized shear stress σ12G−1 and the ratios of shear stress and 298

energies up to a shear strain of γ =10 for the intermediate- and high-W regimes. The former is 299

represented by the discrete set of solutions withW ∈ 0.1;0.2;...;0.9{ } , and the latter by the set 300

W ∈ 1;2;...;10{ } . 301

4.1 Shear stress evolution 302

The shear-stress curves start to diverge at γ ≈ 0.5 (Fig. 2). In general, shear stresses of the SS 303

model and those of the MJ model bracket shear stresses of the FT model. Their discrepancy 304

increases with increasing W. The spurious strain softening of the MJ model is significant for 0.5 ≤ 305

W ≤ 1 while stress oscillations are visible for all greater W. The SS and FT models exhibit 306

monotonously increasing shear stresses over the tested strain range. We continue with a detailed 307

comparison of relative shear stresses and energies, considering the intermediate- and high-W 308

regimes separately (Fig. 3). For the high-W regime, MJ results are omitted because their stress 309

oscillations are highly problematic (see Fig. 2b and appendix 3 for an in-depth analysis). 310

4.2 Intermediate-W regime 311

The SS model predicts higher stresses than the FT model (Fig. 3a). The maximum magnitude of this 312

discrepancy increases from ca. 1% at W = 0.1 and γ ≈ 0.5 to ca. 25% at W = 0.9 and γ ≈ 2.3 . At 313

γ ≤ 0.5 , the difference never exceeds 5%. The MJ model predicts lower stresses than the FT model 314

(Fig. 3a). The maximum discrepancy increases nonlinearly with shear strain and W and reaches 315

from <2% at W = 0.1 to ca. 42% W = 0.9 at γ =10 . The difference does not exceed 3% at γ ≤1 . 316

Since the total energy is the integral of the stress-strain curve, the evolutions of the relative 317

total energies resemble those of the relative shear stresses (Fig. 3b). The SS model computes larger 318

Page 15: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

total energies than the FT model, ranging from ca. 1.5% excess at W = 0.1 and γ ≈1.3 to ca. 20% at 319

W = 0.9 and γ ≈ 3.8 . At γ =1 , the difference is <7%. Again, the MJ model computes lower 320

energies than the FT model with a maximum difference at γ =10 , being about 1% at W = 0.1 up to 321

ca. 32% at W = 0.9. At γ = 2 , the difference is <5% in all cases. 322

In the evolution of stored energy, the SS model predicts larger elastic energy than the FT 323

model, ranging from ca. 2% at W = 0.1 and γ ≈ 0.5 to about 56% at W = 0.9 and γ ≈ 2.3 (Fig. 3c). 324

The curve shapes mirror those of the shear-stress evolution because elastic energy and shear stress 325

are related by Eel = 0.5σ122G−1 (Eq. (11a)). The MJ model gives smaller energies than the FT-model 326

energies with the maximum difference appearing at γ =10 for W ≤ 0.7, ranging from ca. 5% at W = 327

0.1 to ~ 95% at W = 0.7. At larger W, negative ratios are obtained because the MJ theory computes 328

negative stored energy there. 329

In terms of dissipated viscous energy (Fig. 3d), the SS model first gives smaller and then 330

larger values than the FT model. The difference of the smaller values does not exceed 10% (at W = 331

0.9 and γ = 0.7 ). The difference of the lager values reaches ca. 35% at the largest W and smallest γ. 332

The MJ model exhibits the opposite behavior with considerably larger magnitudes of discrepancy, 333

attaining an excess of up to ~ 260% and a deficiency of ca. 30% for the largest W. 334

4.3 High-W regime 335

The overall trends of shear-stress evolution remain similar to those of the intermediate-W regime. 336

The SS model gives shear stresses higher than the FT model with a maximum discrepancy of ca. 337

30% at W = 1 to ca. 350% at W = 10 (Fig. 3e). This trend is reflected in the total-energy comparison 338

(Fig. 3f), which reveals a maximum difference of ~ 25% at W = 1 up to ~ 280% at W = 10. A 339

significant proportion of this energy error is due to problems with the computation of elastic energy 340

(Fig. 3g). The SS model computes larger stored energy than the FT model by up to ca. 50% at W = 341

1 up to more than one order of magnitude at W = 10. Accordingly, the differences in viscous-energy 342

evolution are smaller (Fig. 3h). The largest negative difference ranges from ca. 10% at W = 1 to 343

Page 16: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

about 70% at W = 10; the largest positive difference covers a range of ~ 13% at W = 1 to ~ 65% at 344

W = 10. 345

5. Discussion 346

The three examined models for viscoelastic deformation differ markedly. The SS model ignores 347

large transformations, frame indifference, and self-consistency sensu Bruhns et al. (1999) a priori. 348

As a result, the FT and SS model differ most significantly during highly non-coaxial large 349

deformations in the intermediate- and high-W regimes. The MJ formulation performs worse in some 350

cases than the SS model in spite of its use of a geometrically correct objective stress rate. However, 351

it violates the self-consistency requirement, which incurs unphysical stress oscillations and negative 352

stored energy (Fig. 3c, see also appendix 3). This poses a serious problem for multi-physics 353

formulations with energy feedbacks that rely on accurately tracking cross-scale energy fluxes 354

through thermodynamic upscaling methods (Regenauer-Lieb et al., 2013, Regenauer-Lieb et al., 355

2014). The FT model provides an energetically consistent large-transformation model that satisfies 356

the objectivity and self-consistency requirements. We now proceed to discuss the applicability of 357

the examined Maxwell models in the light of our results. 358

With engineering applications in mind, our comparison reveals a range of W where all 359

discussed models deliver qualitatively similar results. At small strains and W < 0.1, the three tested 360

models are identical. Under these conditions, the SS model provides the simplest choice. At 361

intermediate W and γ <1 , the shear stresses of all three models are within ≤ +/- 10%. If this 362

accuracy is acceptable in this strain and W range, it does not matter practically, which model is 363

used. At γ >1 , the shear-stress deviations exceed 10% at W > 0.3 in the MJ model, and at W > 0.4 364

in the SS model. Nevertheless, the shear-stress overestimation of the SS model does not exceed 365

25% in the intermediate-W regime. One may perhaps envisage applications, where for the sake of 366

minimizing the computational burden, this overestimate is acceptable. The MJ model exhibits 367

spurious softening of the shear stresses at W ≥ 0.3 (Fig. 2a). This unphysical effect renders its use 368

problematic at and beyond this W. In the high-W regime, shear stresses in the SS model exceed 369

Page 17: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

those of the FT model by ca. 10% at γ ≈ 0.5 and keep on increasing quickly while the MJ model 370

oscillates unrealistically (Fig. 2b). As it is clear that the formulation of the SS model is ill suited for 371

large transformation, the rapid increase and high stress magnitude of the SS model in high-W 372

regime cannot be considered acceptable. Therefore, we consider the FT model the best choice in the 373

high-W regime. 374

However, if energy consistency matters, the FT model is also better suited for most of the 375

intermediate-W regime. As emphasized in a review by Hobbs et al. (2011), the time evolution of the 376

stored energy of a coupled geological system holds the key to understanding dissipation and thus 377

the formation of localized structures. This notion is deeply rooted in thermodynamic principles 378

(Ziegler, 1977, Coussy, 2004). Therefore, if one wishes to tackle non-coaxial deformation 379

problems, where the mechanical energy of the system is coupled to non-mechanical dissipation 380

processes such as metamorphic processes or heat transfer, energy consistency is important. The SS 381

model overestimates stored mechanical energy in simple shear by well over 10% at W > 0.3. 382

Therefore, we recommend, as a pragmatic rule of thumb, to use the SS model in the intermediate-W 383

regime only in cases where energy consistency is not required or W ≤ 0.3. 384

The MJ model exhibits unphysical negative stored energy at W > 0.7, which obviously 385

violates thermodynamic principles. In fact, all MJ models exhibit negative time gradients in stored 386

energy in the examined W range (appendix 3, Fig. A1). Because the loading at the boundaries of the 387

homogeneously deforming unit square is never removed and material properties remain constant, 388

elastic strain and stored energy should not decrease. Therefore, the MJ model violates the self-389

consistency requirement and also the principle of energy conservation. The associated shear-stress 390

decrease becomes noticeable at W = 0.3 (Fig. 2a). Hence, the MJ model should not be used at W ≥ 391

0.3. The use of the MJ model is thus restricted to the W interval ]0.1;0.3[. This rather small 392

parameter range and the fundamental conceptual shortcomings of the MJ model lead to the question 393

if there is any practical merit in using it. 394

Page 18: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Our analysis suggests that the FT model should be the tool of choice for highly non-coaxial 395

finite viscoelastic deformation problems in the intermediate- and high-W regimes. This notion holds 396

true if one considers the concepts of frame indifference and self-consistency sensu Bruhns et al. 397

(1999) as meaningful physical constraints. Nevertheless, even though the FT model is objective and 398

energetically consistent, it remains to be seen if it can be falsified in laboratory experiments. 399

Experimental tests of the model with rotary rheometry and analogue materials are subject of 400

ongoing work. In addition to its unique shear-stress response, the FT model predicts non-zero 401

normal stresses in simple shear, which can also be observed and tested experimentally (Freeman 402

and Weissenberg, 1948). 403

404

6. Conclusions 405

We introduced a large-strain model for Maxwell viscoelasticity with a logarithmic co-rotational 406

stress rate (the “FT model”). An analysis of homogeneous isothermal simple shear with the FT 407

model compared to a small-strain formulation (the “SS model”) and a model using the Jaumann 408

stress rate (the “MJ model”) leads to the following key conclusions: 409

410

(1) At W ≤ 0.1, all models yield essentially identical results. 411

(2) At larger W, the models show increasing differences for γ > 0.5 . The SS model overestimates 412

shear stresses compared to the FT model while the MJ model exhibits an oscillatory response 413

underestimating the FT model. 414

(3) The MJ model violates the self-consistency condition resulting in stress oscillations and should 415

be disregarded. It does not deliver truly elastic behavior. 416

(4) In the intermediate-W regime, the shear-stress overestimates of the SS model may constitute 417

acceptable errors if energy consistency is not important. If energy consistency is desirable, the SS 418

model should not be used at W ≥ 0.3. 419

Page 19: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

(5) In the high-W regime, stresses in the SS model become unacceptably large. The FT model 420

should be used in this domain. 421

422

The FT model constitutes a physically consistent Maxwell model for large non-coaxial 423

deformations, even at high Weissenberg numbers. It overcomes the conceptual limitations of the SS 424

model, which is limited to small transformations, not objective and not self-consistent. It also solves 425

the problem of the energetically aberrant oscillations of the MJ model. The FT model promises 426

particular usefulness for simulations of bending plates, rotating blocks of crust, and lithospheric 427

viscoelastic shear zones, especially those at the scale of hundreds of meters or smaller as well as 428

those involving strong heterogeneity of material properties due to, for example, layering or 429

inclusions. In such systems, high Weissenberg numbers and large rotations can be expected. 430

Acknowledgements 431

The authors gratefully acknowledge funding by the Australian Research Council via Discovery 432

project DP140103015. CS is grateful for support of the QUT High Performance Computing group. 433

We are very thankful for the thorough and most helpful reviews of Greg Houseman, an anonymous 434

reviewer, and the editor Jörg Renner. 435

436

References 437

Abaqus/Explicit, 2015. User’s Manual, Version 6.14, Dassault Systèmes. 438

Abaqus/Standard, 2009. User's Manual, Version 6.9, (Hibbit, Karlsson, and Sorensen Inc., 439

Pawtucket, Rhode Island, 2000). 440

Abels, A. & Bischoff, L., 1999. Clockwise block rotations in northern Chile: Indications for a large-441

scale domino mechanism during the middle-late Eocene, Geology, 27, 751-754. 442

Bailey, R.C., 2006. Large time step numerical modelling of the flow of Maxwell materials, 443

Geophysical Journal International, 164, 460-466. 444

Page 20: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Bass, J.D., 1995. Elasticity of Minerals, Glasses, and Melts. in Mineral physics and 445

crystallography: a handbook of physical constants, pp. 45-63, ed. Ahrens, T. J. American 446

Geophysical Union. 447

Beuchert, M.J. & Podladchikov, Y.Y., 2010. Viscoelastic mantle convection and lithospheric 448

stresses, Geophysical Journal International, 183, 35-63. 449

Biermeier, C. & Stüwe, K., 2003. Strain rates from snowball garnet, Journal of Metamorphic 450

Geology, 21, 253-268. 451

Branlund, J.M., Kameyama, M.C., Yuen, D.A. & Kaneda, Y., 2000. Effects of temperature-452

dependent thermal diffusivity on shear instability in a viscoelastic zone: implications for 453

faster ductile faulting and earthquakes in the spinel stability field, Earth and Planetary 454

Science Letters, 182, 171-185. 455

Bruhns, O., 2003. Objective rates in finite elastoplasticity. in IUTAM Symposium on Computational 456

Mechanics of Solid Materials at Large Strains, pp. 151-160, ed Miehe, C. Springer, 457

Dordrecht, Stuttgart, Germany. 458

Bruhns, O., Meyers, A. & Xiao, H., 2001a. Analytical perturbation solution for large simple shear 459

problems in elastic-perfect plasticity with the logarithmic stress rate, Acta Mechanica, 151, 460

31-45. 461

Bruhns, O., Xiao, H. & Meyers, A., 1999. Self-consistent Eulerian rate type elasto-plasticity models 462

based upon the logarithmic stress rate, International Journal of Plasticity, 15, 479-520. 463

Bruhns, O., Xiao, H. & Meyers, A., 2001b. Large simple shear and torsion problems in kinematic 464

hardening elasto-plasticity with logarithmic rate, International Journal of Solids and 465

Structures, 38, 8701-8722. 466

Chakrabarty, J., 2006. Theory of Plasticity, Third edn, Vol., pp. Pages, Elsevier Butterworth-467

Heinemann Oxford, UK. 468

Page 21: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Colak, U.O., 2004. Modeling of large simple shear using a viscoplastic overstress model and 469

classical plasticity model with different objective stress rates, Acta Mechanica, 167, 171-470

187. 471

Conrad, C.P. & Hager, B.H., 1999. Effects of plate bending and fault strength at subduction zones 472

on plate dynamics, J. Geophys. Res., 104, 17551-17571. 473

Coussy, O., 2004. Poromechanics, edn, Vol., pp. Pages, John Wiley & Sons. 474

Dealy, J.M., 2010. Weissenberg and Deborah Numbers - Their Definition and Use, Rheology 475

Bulletin, 79, 28. 476

Dienes, J.K., 1979. On the analysis of rotation and stress rate in deforming bodies, Acta Mechanica, 477

32, 217-232. 478

Durban, D., 1990. A comparative study of simple shear at finite strains of elastoplastic solids, The 479

Quarterly Journal of Mechanics and Applied Mathematics, 43, 449-465. 480

Freeman, S. & Weissenberg, K., 1948. Some new rheological phenomena, Nature, 162, 320-323. 481

Hobbs, B.E., Ord, A. & Regenauer-Lieb, K., 2011. The thermodynamics of deformed metamorphic 482

rocks: A review, Journal of Structural Geology, 33, 758-818. 483

Jaumann, G., 1911. Geschlossenes System physikalischer und chemischer Differential-Gesetze, 484

Sitzber. Akad. Wiss. Wien, Abt. IIa, 120, 385-530. 485

Johnson, S.E., 2009. Porphyroblast rotation and strain localization: Debate settled!, Geology, 37, 486

663-666. 487

Karato, S., 2008. Deformation of Earth Materials: An Introduction to the Rheology of Solid Earth, 488

edn, Vol., pp. Pages, Cambridge University Press. 489

Karrech, A., Poulet, T. & Regenauer-Lieb, K., 2012. Poromechanics of saturated media based on 490

the logarithmic finite strain, Mechanics of Materials, 51, 118-136. 491

Karrech, A., Regenauer-Lieb, K. & Poulet, T., 2011. Frame indifferent elastoplasticity of frictional 492

materials at finite strain, International Journal of Solids and Structures, 48, 397-407. 493

Page 22: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Kaus, B.J.P. & Becker, T.W., 2007. Effects of elasticity on the Rayleigh–Taylor instability: 494

implications for large-scale geodynamics, Geophysical Journal International, 168, 843-862. 495

Kaus, B.J.P. & Podladchikov, Y.Y., 2006. Initiation of localized shear zones in viscoelastoplastic 496

rocks, J. Geophys. Res., 111. 497

Korenaga, J., 2007. Effective thermal expansivity of Maxwellian oceanic lithosphere, Earth and 498

Planetary Science Letters, 257, 343-349. 499

Kusznir, N.J. & Bott, M.H.P., 1977. Stress concentration in the upper lithosphere caused by 500

underlying visco-elastic creep, Tectonophysics, 43, 247-256. 501

Larsen, C.F., Motyka, R.J., Freymueller, J.T., Echelmeyer, K.A. & Ivins, E.R., 2005. Rapid 502

viscoelastic uplift in southeast Alaska caused by post-Little Ice Age glacial retreat, Earth 503

and Planetary Science Letters, 237, 548-560. 504

Lehmann, T., 1972a. Anisotropische plastische Formänderung, Romanian J. Techn. Sci. Appl. 505

Mech., 17, 1077-1086. 506

Lehmann, T., 1972b. Einige Bemerkungen zu einer allgemeinen Klasse von Stoffgesetzen für große 507

elasto-plastische Formänderungen, Ing.-Arch., 41, 297-310. 508

Liu, X., Eckert, A. & Connolly, P., 2016. Stress evolution during 3D single-layer visco-elastic 509

buckle folding: Implications for the initiation of fractures, Tectonophysics, 679, 140-155. 510

Mancktelow, N.S., 1999. Finite-element modelling of single-layer folding in elasto-viscous 511

materials: the effect of initial perturbation geometry, Journal of Structural Geology, 21, 512

161-177. 513

Mancktelow, N.S., 2002. Finite-element modelling of shear zone development in viscoelastic 514

materials and its implications for localisation of partial melting, Journal of Structural 515

Geology, 24, 1045-1053. 516

Mandal, N. & Khan, D., 1991. Rotation, offset and separation of oblique-fracture (rhombic) 517

boudins: theory and experiments under layer-normal compression, Journal of Structural 518

Geology, 13, 349-356. 519

Page 23: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Meyers, A., Bruhns, O. & Xiao, H., 2005. Objective stress rates in repeated elastic deformation 520

cycles, PAMM, 5, 249-250. 521

Moresi, L., Dufour, F. & Mühlhaus, H.B., 2002. Mantle Convection Modeling with 522

Viscoelastic/Brittle Lithosphere: Numerical Methodology and Plate Tectonic Modeling, 523

Pure and Applied Geophysics, 159, 2335-2356. 524

Moss, W.C., 1984. On instabilities in large deformation simple shear loading, Computer Methods in 525

Applied Mechanics and Engineering, 46, 329-338. 526

Mühlhaus, H.-B. & Regenauer-Lieb, K., 2005. Towards a self-consistent plate mantle model that 527

includes elasticity: simple benchmarks and application to basic modes of convection, 528

Geophysical Journal International, 163, 788-800. 529

Passchier, C.W. & Simpson, C., 1986. Porphyroclast systems as kinematic indicators, Journal of 530

Structural Geology, 8, 831-843. 531

Peltier, W.R., 1976. Glacial-Isostatic Adjustment—II. The Inverse Problem, Geophysical Journal 532

International, 46, 669-705. 533

Pfiffner, O.A. & Ramsay, J.G., 1982. Constraints on geological strain rates: Arguments from finite 534

strain states of naturally deformed rocks, Journal of Geophysical Research: Solid Earth, 87, 535

311-321. 536

Ramsay, J.G., 1980. Shear zone geometry: A review, Journal of Structural Geology, 2, 83-99. 537

Ramsay, J.G. & Graham, R.H., 1970. Strain variation in shear belts, Canadian Journal of Earth 538

Sciences, 7, 786-813. 539

Ramsay, J.G. & Huber, M.I., 1987. The Techniques of Modern Structural Geology, Vol. 2: Folds 540

and Fractures, edn, Vol. 2, pp. Pages, Academic Press, London. 541

Regenauer-Lieb, K., Veveakis, M., Poulet, T., Wellmann, F., Karrech, A., Liu, J., Hauser, J., 542

Schrank, C., Gaede, O., Fusseis, F. & Trefry, M., 2014. Multiscale coupling and 543

multiphysics approaches in earth sciences: Applications, Journal of Coupled Systems and 544

Multiscale Dynamics, 1, 281-323. 545

Page 24: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Regenauer-Lieb, K., Veveakis, M., Poulet, T., Wellmann, F., Karrech, A., Liu, J., Hauser, J., 546

Schrank, C., Gaede, O. & Trefry, M., 2013. Multiscale coupling and multiphysics 547

approaches in earth sciences: Theory, Journal of Coupled Systems and Multiscale 548

Dynamics, 1, 49-73. 549

Robin, C.M.I. & Bailey, R.C., 2009. Simultaneous generation of Archean crust and subcratonic 550

roots by vertical tectonics, Geology, 37, 523-526. 551

Sone, H. & Uchide, T., 2016. Spatiotemporal evolution of a fault shear stress patch due to 552

viscoelastic interseismic fault zone rheology, Tectonophysics, 684, 63-75. 553

Tohver, E., Trindade, R.I.F., Solum, J.G., Hall, C.M., Riccomini, C. & Nogueira, A.C., 2010. 554

Closing the Clymene ocean and bending a Brasiliano belt: Evidence for the Cambrian 555

formation of Gondwana, southeast Amazon craton, Geology, 38, 267-270. 556

Truesdell, C., 1956. Hypo-Elastic Shear, Journal of Applied Physics, 27, 441-447. 557

Wheeler, J., 2010. Anisotropic rheology during grain boundary diffusion creep and its relation to 558

grain rotation, grain boundary sliding and superplasticity, Philosophical Magazine, 90, 559

2841-2864. 560

White, J., 1964. Dynamics of viscoelastic fluids, melt fracture, and the rheology of fiber spinning, 561

Journal of Applied Polymer Science, 8, 2339-2357. 562

White, J. & Mawer, C., 1992. Deep-crustal deformation textures along megathrusts from 563

Newfoundland and Ontario: implications for microstructural preservation, strain rates, and 564

strength of the lithosphere, Canadian Journal of Earth Sciences, 29, 328-337. 565

Xiao, H., Bruhns, O. & Meyers, A., 1997a. Hypo-Elasticity Model Based upon the Logarithmic 566

Stress Rate, Journal of Elasticity, 47, 51-68. 567

Xiao, H., Bruhns, O. & Meyers, A., 2006. Elastoplasticity beyond small deformations, Acta 568

Mechanica, 182, 31-111. 569

Xiao, H., Bruhns, O.T. & Meyers, A., 1997b. Logarithmic strain, logarithmic spin and logarithmic 570

rate, Acta Mechanica, 124, 89-105. 571

Page 25: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Xiao, H., Bruhns, O.T. & Meyers, A., 1998a. On objective corotational rates and their defining spin 572

tensors, International Journal of Solids and Structures, 35, 4001-4014. 573

Xiao, H., Bruhns, O.T. & Meyers, A., 1998b. Strain Rates and Material Spins, Journal of Elasticity, 574

52, 1-41. 575

Zhang, Y., Mancktelow, N.S., Hobbs, B.E., Ord, A. & Mühlhaus, H.B., 2000. Numerical modelling 576

of single-layer folding: clarification of an issue regarding the possible effect of computer 577

codes and the influence of initial irregularities, Journal of Structural Geology, 22, 1511-578

1522. 579

Ziegler, H., 1977. An introduction to thermomechanics, edn, Vol. 21, pp. Pages, North-Holland 580

Publishing Co., Amsterdam. 581

582

Figures 583

584

Fig. 1: Double-logarithmic contour plots of the decadic logarithm of W over those of viscosity and strain rate 585

for the (a) lower- and (b) upper-bound estimate of the shear modulus. The white dotted lines denote 586

parameter combinations with W = 0.1; the white dashed lines mark those with W = 1. Textboxes indicate the 587

models, which are valid in the respective parameter fields separated by those lines. 588

Page 26: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

589

Fig. 2: Plots of normalized shear stress over shear strain for the SS, MJ, and FT models in the (a) 590

intermediate-W regime, and (b) the high-W regime. Each plot title indicates the respective W and the 591

dimensionless time kmax reached at final γ . Note that the results for W = 0.1 are omitted because the curves 592

cannot be distinguished at this scale of observation. 593

Page 27: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

594Fig. 3: Top panel (light-grey canvas): Comparison of model results in the intermediate-W regime. (a) Shear-595

stress ratio, (b) total-energy ratio, (c) stored-energy ratio, and (d) viscous-energy ratio for the SS model 596

(solid lines) and the MJ model (dashed lines). Each sub-figure contains two plots: on the left, SS results are 597

divided by FT results; on the right, MJ results are divided by FT results. Curve symbols denote the respective 598

W, defined in the right panel of (d). Bottom panel (grey canvas): Comparison of the SS to the FT model in 599

the high-W regime. Results of the SS model are again divided by those of the FT model. The legend for the 600

respective value of W is shown in (e). 601

602

Page 28: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Appendix 603

1. Analytical solution of the MJ model 604

The MJ model for simple shear requires the solution of the following nonhomogeneous linear 605

system of ordinary differential equations: 606

607

η !γ = trel !σ12 − !γ !σ 22( )+σ120 = trel !σ 22 + !γσ12( )+σ 22

(A1). 608

609

For convenience, we insert the placeholders: 610

611

A = −trel−1

B = !γC =η !γtrel

−1

(A2 a, b, c), 612

613

which leads, after rearrangement, to: 614

615

!σ12 = Aσ12 + Bσ 22 +C!σ 22 = Aσ 22 − Bσ12

(A3 a, b). 616

617

The solution is: 618

619

σ12 t( ) = eAt sin Bt( )Q1 − cos Bt( )Q2⎡⎣ ⎤⎦−

ACA2 + B2

σ 22 t( ) = eAt sin Bt( )Q2 + cos Bt( )Q1⎡⎣ ⎤⎦−BC

A2 + B2

(A4 a, b). 620

621

From the initial conditions σ12 t = 0( ) =σ 22 t = 0( ) = 0 , the constants Q1 and Q2 are calculated: 622

Page 29: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

Q1 =BC

A2 + B2

Q2 = −AC

A2 + B2

(A5 a, b). 623

624

Re-inserting all substitutions yields the final result: 625

626

σ12 t( ) =η !γ

trel−1 + trel !γ

2

⎝⎜

⎠⎟ e

−ttrel !γ sin !γt( )− trel−1 cos !γt( )⎡⎣ ⎤⎦+ trel

−1⎡

⎣⎢⎢

⎦⎥⎥

σ 22 t( ) =η !γ

trel−1 + trel !γ

2

⎝⎜

⎠⎟ e

−ttrel !γ cos !γt( )+ trel−1 sin !γt( )⎡⎣ ⎤⎦− !γ

⎣⎢⎢

⎦⎥⎥

(A6 a, b). 627

628

To compute the total energy, we integrate Eq. (7) after substituting Eq. (A6a) and W: 629

630

ETOT t( ) = GW !γW 2 +1

t − GW 2

W 2 +1( )2 e

−ttrel W 2 −1( )cos !γt( )+ 2W sin !γt( )⎡⎣

⎤⎦+Q3 (A7). 631

632

The integration constant Q3 is again determined from the initial condition ETOT t = 0( ) = 0 : 633

634

Q3 =GW 2 W 2 −1( )

W 2 +1( )2 (A8). 635

636

2. Numerical solution of FT model 637

From a numerical point of view, Ωij is completely defined at the beginning of each time increment, 638

given its purely kinematic nature. Similarly, to calculate the quantity −ΩikΣkj + ΣikΩkj , it is 639

customary to use the Kirchhoff stress of the beginning of the time increment in a material 640

integration module, the choice we opt for here. The only remaining quantity needed to obtain the 641

Page 30: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

co-rotational rate of the Kirchhoff stress ⌣Σij is the time derivative !Σij . For the Maxwell model, the 642

time increment is: 643

644

Σij + trel !Σij = 2η !hij (A9), 645

646

where hij = 0.5lnbij is Hencky’s strain tensor, and bij is the left Cauchy-Green strain tensor (for 647

derivations of these tensors, see for example Karrech et al., 2011, Bruhns et al., 1999). This 648

equation is the modified version of the constitutive model in Eq. (1), involving the substitution of 649

the Kirchhoff stress, insertion of the Maxwell relaxation time, and replacement of the deformation-650

rate tensor with the time derivative of the Eulerian logarithmic Hencky strain tensor. We use the 651

central-difference scheme to approximate the stress and strain increments as: 652

653

!gij t = t0 +Δt2

⎝⎜

⎠⎟ =

ΔgijΔt

gij t = t0 +Δt2

⎝⎜

⎠⎟ = gij t = t0( )+

Δgij2

(A10), 654

655

where gij is some second-order tensor, gij t = t0( ) is its value at the beginning of the increment, and 656

Δgij is the change in gij over the time increment Δt . Hence, it can be seen that: 657

658

Δt2+ trel

⎝⎜

⎠⎟ΔΣij = 2ηΔhij −ΔtΣij t = t0( ) (A11). 659

660

Therefore, the Jacobian matrix describing the incremental constitutive behavior before co-rotational 661

correction is: 662

663

Page 31: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

ΔΣij

Δ !hkl=

ηΔt2+ trel

δikδ jl +δilδ jk( ) (A12), 664

665

where !hkl is the Hencky strain tensor in Voigt notation, and δij is the Kronecker delta. Moreover, 666

the stress stored at the end of the increment is: 667

668

Σij t = t0 +Δt( ) = ΔΣij −ΩikΣkj + ΣikΩkj (A13). 669

670

The Jacobian matrix and the Kirchhoff stress at the end of the time increment expressed by Eqs. 671

(A12) and (A13), respectively, are sufficient to describe the constitutive behavior in a finite-element 672

code. We implemented the above formulation into the user routine UMAT (Abaqus/Standard, 2009) 673

to solve the FT model numerically. 674

675

3. Analysis of oscillations in the MJ model 676

The analytical solution in Eq. (15) explains the spurious softening and stress oscillations in the MJ 677

model (Fig. 2). Let us consider the expression for the shear stress in the following. It contains the 678

trigonometric term T12 =W sin Wk( )− cos Wk( ) , which is the obvious cause of the spurious 679

oscillations. The term T12 has a half-period of πW −1 . Its amplitude magnitude is: A = W 2 +1 . 680

The latter expression shows why T12 has no noticeable influence on the shear-stress response at W ≤ 681

0.1. In addition, T12 is multiplied with the damping factor e-k. Therefore, the oscillations are damped 682

at sufficiently large k ≥ 2π (Fig. A1). In the range of ten relaxation times, models with W < 1 683

exhibit spurious softening, while those at larger W show oscillations (Fig. A1). The gravity of this 684

fundamental problem becomes perhaps clearer when one considers the evolution of stored energy in 685

the MJ model. Eq. (15a) implies immediately that the elastic energy also contains trigonometric 686

Page 32: A comparative study of Maxwell viscoelasticity at large strains … · Only models that use the logarithmic spin are self-consistent. 85 To analyse the relevance of the co-rotational

terms and therefore is subject to oscillations. This can be seen in Fig. 3c, which reveals unphysical 687

negative stored energy. Due to the kinematics of the examined simple-shear case, the stored energy 688

in the system must not decrease but approach a constant maximum value of Eelmax = 0.5 σ vis

∞( )2G−1 in 689

the infinite-time limit. In other words, because constant-velocity boundary conditions are applied to 690

a homogeneous body with constant material properties, it should not relax stored stresses. However, 691

the elastic-energy solution for the MJ model exhibits negative gradients. The critical dimensionless 692

time kcrit, after which the time derivative of the elastic energy first becomes negative, is plotted for a 693

number of W in the right panel of Fig. A1. Beyond kcrit the MJ model is dissipative and violates the 694

first law of thermodynamics. Since the principal of energy conservation is at the heart of continuum 695

mechanics, we propose to abandon the MJ model. 696

697

Fig. A1: Left panel: Contour plot of e−k W sin Wk( )− cos Wk( )⎡⎣ ⎤⎦ in W-k space. Right panel: Plot of kcrit 698

over W. The time kcrit is the first point in time, at which ∂Eel / ∂k = 0 . For discussion, see text. 699