for Journal of Structural Geology Manuscript Draft

0 downloads 0 Views 4MB Size Report
Page 1 ... Nevertheless, I have a few suggestions the authors may want to consider that I ..... One major problem in microstructural studies is that we normally only have a ... investigated deformation related features such as dynamic recrystallization .... In this case, the number of equations to be solved is not limited; hence.
Elsevier Editorial System(tm) for Journal of Structural Geology Manuscript Draft Manuscript Number: SG-D-17-00310R1 Title: A review of numerical modelling of the dynamics of microstructural development in rocks and ice: Past, present and future Article Type: SI:JSG - 40th Anniversary Keywords: review, evolution of microstructures, numerical modelling, dynamic recrystallization, process interaction Corresponding Author: Dr. Sandra Piazolo, Ph.D. Corresponding Author's Institution: University of Leeds First Author: Sandra Piazolo, Ph.D. Order of Authors: Sandra Piazolo, Ph.D.; Paul D Bons; Albert Griera; Maria-Gema Llorens; Enrique Gomez-Rivas; Daniel Koehn; John Wheeler; Robyn Gardner; Jose R. A. Godinho; Lynn Evans; Ricardo A Lebensohn; Mark W Jessell Manuscript Region of Origin: Abstract: This review provides an overview of the emergence and current status of numerical modelling of microstructures, a powerful tool for predicting the dynamic behaviour of rocks and ice at the microscale with consequence for the evolution of these materials at a larger scale. We emphasize the general philosophy behind such numerical models and their application to important geological phenomena such as dynamic recrystallization and strain localization. We focus in particular on the dynamics that emerge when multiple processes, which may either be enhancing or competing with each other, are simultaneously active. Here, the ability to track the evolving microstructure is a particular advantage of numerical modelling. We highlight advances through time and provide glimpses into future opportunities and challenges.

Cover Letter

Revision of Ms. Ref. No.: SG-D-17-00310 Title: A review of numerical modelling of the dynamics of microstructural development in rocks and ice: Past, present and future Journal of Structural Geology Dear Prof. Passchier, Dear Cees,

Thank you for handling our contribution and the opportunity to revise our manuscript: A review of numerical modelling of the dynamics of microstructural development in rocks and ice: Past, present and future. We have now carried out the revisions in the light of the useful and constructive comments by the reviewer. We took great care in addressing all the comments. Reviewer #1 suggested to add a table with the models cited and processes modelled into the main text. We hope that you agree with the reviewer and us that this table that we now supply will be very useful to the community and therefore warrants to be placed in the main body of the manuscript. Below we have outlined our responses and reasoning behind those. We hope you agree with us that the manuscript is now acceptable for publication in JSG.

Kind regards, Sandra Piazolo (corresponding author, on behalf of all authors)

Comments to review: Reviewers' comments are given in black font below, with our replies and changes made shown in red font. A copy of the revised manuscript with all changes tracked is also included in this resubmission.

Reviewer #1: This review paper on the topic of microstructural numerical modeling is a valuable addition to the JSG volume. Modeling is an extremely useful component of research, and one that has probably been neglected in microstructural work more than in other fields. As such, any pathway - such as this paper - to expand the reach of modeling will serve the community well. This manuscript does a nice job of presenting the historical development and range of applications in current modeling approaches. The vignettes, such as diffusion creep, stylolite formation, and dissolution, provide readers with good fuel for consideration in their own research programs. The introduction is particularly compelling and well structured. This manuscript could be published as is. The graphics and writing are clear and raise no concerns about the validity of the work. We are glad to see the positive overall assessment of our manuscript Nevertheless, I have a few suggestions the authors may want to consider that I think would increase the impact of this contribution and be more accessible to a wider audience of new and old modeling

practitioners. 1. I would find a table of the different models cited, links to access the code, if available, and the different processes included in each of the models a useful reference. In fact we as the authors did start with such a compilation when initially writing this contribution but realized that this would significantly increase the length of this short review if it were included in the printed version, hence decided against such a table. We now, however, as suggested by the reviewer, provide the table, but restrict ourselves to the cited models stating this explicitly. We hope you agree that this table is valuable enough to be in the main text. 2. The discussion on lines 50-55 seems to stop a bit short of the reality of our current understanding of microstructural processes. I think some additional thoughts about how non-linear feedbacks and potentially chaotic systems operate in the microstructural domain would strengthen the argument for modeling. We have now included a couple of extra sentences discussing non-linear feedbacks and their potential importance. 3. For several sections, particularly "2.2. Multiprocess modelling - approaches", and "2.3. The numerical representation of a microstructure", a more thorough description of the pros and cons of different approaches (e.g., operator splitting, different ways to represent the microstructure) would be welcome. We agree that this would be desirable, however due to the space limitations a thorough discussion is not possible. However, we have now added several sentences to identify the main pros and cons of the different approaches. 4. In a similar vein, overall, the text leans much more in the descriptive direction (making it very readable), than in the technical. I suggest considering adding some more technical discussion about the challenges of the modeling and the areas that we can consider robust. For the audience this is intended for, we would like to keep this contribution mainly on the descriptive side, however we take the point that some of the technical challenges and achievements need to be highlighted more. Hence we added for each section another sentence specifically about technical issues/ advances and robust approaches. 5. Another topic that could be expanded builds off Lines 100-103, which mention the role of the microscale stress and strain fields. The non-evolutionary models mentioned in section 5.5 contribute to that, but this is an area that is ripe for more investigation (e.g., Powell, R., Evans, K. A., Green, E. C. R. and White, R. W., On equilibrium in non-hydrostatic metamorphic systems. J Metamorph Geol. Accepted Author Manuscript. doi:10.1111/jmg.12298; Jessell et al., 2005). We agree that this area ripe for more investigation, especially in the dynamic field. However, Powell et al. 2018, as well as being critical of co-author Wheeler, do not provide a way in which to understand

diffusion creep as coupled to other chemical processes. So instead we cite Wheeler 2018, who does. We have now added “…local variations in stress and strain play a significant role in the evolution of microstructures and material behaviour, with their effects on chemistry being particularly intriguing (Wheeler 2018)”. Wheeler, J., 2018. The effects of stress on reactions in the Earth: sometimes rather mean, usually normal, always important. Journal of Metamorphic Geology 36, 439-461. (https://doi.org/10.1111/jmg.12299) 6. Section 6 (line 518) provides a nice look back, and I would welcome more forward looking thoughts on the most pressing problems or the problems that have the best potential for near-term solutions. This group of authors has a unique and exceptional perspective on the field, and this seems a good opportunity to share with the community more details of what they see as the future. In section 5 we already discuss some of the future challenges, hence in section 6, for sake of brevity, we only highlight what we considered the main challenge – linking chemistry and deformation in a robust manner

These above suggestions are merely suggestions - the authors should feel free to take them or leave them. I have a few minor additional comments: Line 117: Perhaps begin the paragraph with the sentence that starts on line 121, to provide more of a segue. Done, we have restructured the paragraph accordingly. Line 297-298: "In this case dissolution and/or precipitation are driven by changes in elastic and surface energies as well as normal stress at the interface." This probably needs some citation, as I wouldn't say that there is universal agreement about the drivers of dissolution and precipitation (e.g., Kristiansen, Kai, Markus Valtiner, George W. Greene, James R. Boles, and Jacob N. Israelachvili. "Pressure solutionThe importance of the electrochemical surface potentials." Geochimica et Cosmochimica Acta 75, no. 22 (2011): 6882-6892.) This leads us to a change in Section 3.2 the general description of diffusion along stressed boundaries to: “The mechanism involves diffusion driven by gradients in normal stress along boundaries, and in simple models no other driving forces need to be included”. This is then contrasted to the mechanisms discussed in the next section 3.3 “In this case dissolution and/or precipitation are driven not just by normal stress gradients at the interface (c.f. section 3.2) but also by changes in elastic and surface energies. In addition to these driving forces, the detailed electrochemistry of interfaces has a kinetic effect on rates (Gratier et al. 2013) ”.

We decided to cite Gratier et al. here, as their work is clearly relevant and in line with results from coauthor J. Wheeler (Sheldon et al. 2003, & 2004) Sheldon, H.A., Wheeler, J., Worden, R.H., Cheadle, M.J., 2003. An analysis of the roles of stress, temperature and pH in chemical compaction of sandstones. Journal of Sedimentary Research 73, 64-71. Sheldon, H.A., Wheeler, J., Worden, R.H., Cheadle, M.J., 2004. An analysis of the roles of stress, temperature and pH in chemical compaction of sandstones: Reply. Journal of Sedimentary Research 74, 449-450. Line 488: The main description paper for the elastic code used by Naus-Thijssen is: Vel, S.S., Cook, A., Johnson, S.E., and Gerbi, C., 2016, Computational Homogenization and Micromechanical Analysis of Textured Polycrystalline Materials, Computer Methods in Applied Mechanics and Engineering, v. 310, p. 749-779, doi: 10.1016/j.cma.2016.07.037. Apologies, we have now added this reference.

Highlights (for review)

Highlights:     

Review of numerical modelling of dynamic interaction of processes at the microscale Assessment of the effect of competing processes allows in-depth understanding Modelling provides insights into link between processes and material properties Modelling of the interplay between physical and chemical processes is needed Future developments promise enhanced understanding and predictive powers

*Manuscript Click here to view linked References

1

A review of numerical modelling of the dynamics of microstructural development in

2

rocks and ice: Past, present and future

3 4

S. Piazolo1, P. D. Bons2, A. Griera3, M.-G. Llorens2, E. Gomez-Rivas4,5, D. Koehn6, J.

5

Wheeler7, R. Gardner8, J. R. A. Godinho9,, L. Evans8,10, R. A. Lebensohn11, M. W.

6

Jessell12

7

1

School of Earth and Environment, University of Leeds, Leeds LS2 9JT, United Kingdom

8

2

Department of Geosciences, Eberhard Karls University Tübingen, Wilhelmstr. 56, 72074

9

Tübingen, Germany

10

3

11

(Cerdanyola del Vallès), Spain

12

4

13

080280 Barcelona, Spain

14

5

15

United Kingdom

16

6

17

Glasgow G12 8QQ, UK

18

7

19

Herdman Building, Liverpool University, Liverpool L69 3GP, U.K.

20

8

21

Systems/GEMOC, Department of Earth and Planetary Sciences, Macquarie University, NSW

22

2109, Australia

23

9

24

M13 9PL, Manchester, UK.

Departament de Geologia, Universitat Autònoma de Barcelona, 08193 Bellaterra

Departament de Mineralogia, Petrologia i Geologia Aplicada, Universitat de Barcelona,

School of Geosciences, King’s College, University of Aberdeen, Aberdeen AB24 3UE,

School of Geographical and Earth Sciences, Gregory Building, University of Glasgow,

Dept. Earth, Ocean and Ecological Sciences, School of Environmental Sciences, Jane

Australian Research Council Centre of Excellence for Core to Crust Fluid

Henry Moseley X-ray Imaging Facility, School of Materials, The University of Manchester,

25

10

26

Vic 3800, Australia

27

11

Los Alamos National Laboratory, MST8, MS G755, Los Alamos, NM, 87545, USA

28

12

Centre for Exploration Targeting, School of Earth Sciences, The University of Western

29

Australia, 35 Stirling Hwy, Crawley, 6009, Australia

School of Earth, Atmosphere and Environmental Sciences, Monash University, Clayton,

30 31

Corresponding author: Sandra Piazolo ([email protected])

32 33

Abstract

34

This review provides an overview of the emergence and current status of numerical

35

modelling of microstructures, a powerful tool for predicting the dynamic behaviour of rocks

36

and ice at the microscale with consequence for the evolution of these materials at a larger

37

scale. We emphasize the general philosophy behind such numerical models and their

38

application to important geological phenomena such as dynamic recrystallization and strain

39

localization. We focus in particular on the dynamics that emerge when multiple processes,

40

which may either be enhancing or competing with each other, are simultaneously active.

41

Here, the ability to track the evolving microstructure is a particular advantage of numerical

42

modelling. We highlight advances through time and provide glimpses into future

43

opportunities and challenges.

44 45

Keywords:

46

recrystallization, process interaction

review,

47 48

1.

Introduction

evolution

of

microstructures,

numerical

modelling,

dynamic

49

Structural geology studies how rocks deform under applied stress or strain. These studies are

50

applied from the sub-grain scale to that of mountain belts and tectonic plates. To know what

51

has happened to a volume of rock millions of years ago, or to predict what would happen to a

52

rock volume under certain (future) conditions, we need to know the link between (i) the

53

boundary conditions (e.g. applied stress or strain, pressure-temperature conditions), (ii) the

54

intrinsic material properties (e.g. Young's modulus, crystal symmetry, slip systems) and (iii)

55

the processes that may be activated (e.g. dislocation creep, pressure solution, mineral phase

56

changes). These three together determine the evolution of the state of the rock (e.g.

57

developing cleavage, folds, lineation, stylolites) and its bulk material properties (e.g.

58

viscosity). The microstructure is thus a state variable of a rock, describing the state it

59

achieved as a result of the interplay between various processes and boundary conditions.

60

Importantly, the microstructure is not a passive log of events, but plays an active and central

61

role in the evolution of a rock throughout its history (Gottstein, 2004). In this context non-

62

linear feedback between a specific microstructural configuration and local changes in

63

intrinsic and extrinsic parameters may be important but difficult to predict. Hence, the

64

microstructure is the link that couples the material properties, boundary conditions and

65

processes that together control the behaviour and evolution of a rock. Consequently, the

66

analysis and correct interpretation of microstructures is crucial in gaining an understanding of

67

how rocks, including ice, deform on Earth and other planets. Here we define microstructure

68

as the full spatial, compositional and orientational arrangement of all entities in a rock,

69

typically on the scale of a thin section to hand specimen (roughly µm to cm) (Hobbs et al.,

70

1976). These entities include minerals, grain and subgrain boundaries, crystal lattice

71

orientations, and chemical composition from the nano- to micro-scale.

72

While the rock's temperature or elastic strain are ephemeral, the microstructure may be

73

preserved for the geologist to interpret millions to billions of years later. Therefore,

74

microstructures are one of the prime forensic tools to unravel the history of a rock that allows

75

us to deduce the succession of strain rates, stresses, diagenetic and metamorphic conditions

76

and the rock's rheology during deformation (e.g. Hobbs et al., 1976; Vernon, 2004; Passchier

77

and Trouw, 2005).

78 79

One major problem in microstructural studies is that we normally only have a static record as

80

post-mortem "images" (grain arrangement in thin section, chemical and orientation

81

characteristics derived using electron microscopy, 3D datasets from computer tomography,

82

neutron diffraction and synchrotron beam analysis, etc.). To interpret microstructures, we

83

need to know how they form and change. Means (1977) wrote: "a more valuable kind of

84

kinematic investigation is that in which the time sequence of incremental strains or

85

incremental displacements, the kinematic history, is correlated with progressive structural

86

change. This promises to reveal more about the physics of rock deformation and to provide

87

more sensitive structural methods for reconstructing tectonic history". Such investigations

88

first became possible using in-situ analogue experiments that were introduced to geology by

89

influential scientists such as Cloos (1955) and Ramberg (1981), and at the microstructural

90

scale by Means (1977). In in-situ analogue experiments by Means (1977), inspired by those

91

of McCrone and Chen (1949), a thin-section sized sample of crystalline analogue material is

92

sandwiched and deformed between glass plates to observe the changing microstructure under

93

a microscope. These experiments were followed by a large number of studies that

94

investigated deformation related features such as dynamic recrystallization and crystal

95

plasticity (e.g. Means, 1977; 1980; Means and Ree, 1988; Park and Means, 1996; Urai et al.,

96

1980; Urai and Humphreys, 1981; Wilson, 1986; Jessell, 1986; Ree, 1991; Bons and Jessell,

97

1999; Ree and Park, 1997; Herwegh and Handy, 1996, 1998; Nama et al., 1999; Wilson et

98

al., 2014), melt or fluid-bearing microstructures (e.g. Urai, 1983; Rosenberg and Handy,

99

2000, 2001;

Rosenberg, 2001; Schenk and Urai, 2004, 2005; Walte et al., 2005), the

100

development of vein and fringe microstructures (e.g. Hilgers et al., 1997; Koehn et al., 2003)

101

and the formation of -clasts (ten Brink et al., 1995). Many of these experiments made clear

102

that when a polycrystalline material deforms, multiple processes act upon the microstructure

103

and govern its dynamic behaviour throughout its deformation and post-deformation history

104

(Fig. 1a; see also movies at http://www.tectonique.net/MeansCD/). In addition, pre-existing

105

heterogeneities and/or those developing during deformation can have a significant effect on

106

the dynamics of the system since local variations in stress and strain play a significant role in

107

the evolution of microstructures and material behaviour, with their effects on chemistry being

108

particularly intriguing (Wheeler, 2018).

109

Although changes in microstructural characteristics (e.g. grain size) can be quantified in such

110

experiments, the underlying principles of the active processes still need to be deduced.

111

Therefore, an additional tool is needed. Here numerical models simulating microstructural

112

development based on several concurrent processes become important. Such numerical

113

simulations share the advantage of in-situ experiments that the full microstructural

114

development can be traced in form of time-series, with the opportunity to systematically

115

study the effect of different processes and/or pre-existing heterogeneities on the

116

microstructural development. In addition, numerical simulations are neither constrained by

117

time nor specific boundary conditions. Once calibrated against laboratory experiments (e.g.

118

Piazolo et al., 2004) or analytical solutions, they can be applied to a wide range of conditions

119

or materials including those not attainable or suitable for laboratory experiments. In many

120

cases, the developed microphysical behaviour may be applied to problems on the continental

121

scale.

122

In this contribution, we aim to give an overview of the concepts behind numerical modelling

123

of microstructures, i.e. microdynamic modelling, and the achievements that have been made

124

in this field. This is followed by a selection of examples of the current state of the art. Finally,

125

future possibilities and directions are briefly discussed, including some work in progress. We

126

focus on those studies that (i) are applied to geological materials, (ii) involve several

127

processes and (iii) allow prediction and visualization of the development of the

128

microstructure. Hence, we exclude studies where microstructures are not spatially mapped

129

along with models in which grains are modelled as if embedded in a medium with averaged

130

properties. Table 1 summarizes the microdynamic numerical models cited, the numerical

131

method used and processes included in each model.

132 133

2. Numerical simulation of microstructures

134

2. 1. Emergence and philosophy

135

Soon after the first in-situ experiments, computers had advanced to a stage allowing the first

136

numerical simulations of microstructural development, with a full "image" of the

137

microstructure, to appear in geology (Etchecopar, 1977; Jessell, 1988a,b; Jessell and Lister,

138

1990). When using such models we accept the fact that the scientific description of the

139

phenomena studied does not fully capture reality. This is not a shortcoming but a strength, as

140

general system behaviour and interrelationships can be studied systematically and

141

transferred to geological questions. In a numerical model the behaviour of a system emerges

142

from the piece-meal enumeration of the behaviour of each of its component parts. As a

143

consequence, numerical models are powerful and necessary when the bulk behaviour is

144

influenced by the local interaction of the components. In contrast, global, averaged or mean

145

field solutions cannot fully encompass the effect of processes that interact with each other

146

on the small scale. In nature, patterns emerge (e.g. orientation and arrangement of high-

147

strain zones, foliation, fractures) which may be used to deduce conditions of their formation.

148

Numerical models are powerful tools to investigate this link between patterns observed in

149

nature and processes that are responsible for their development.

150 151

2.2. Multiprocess modelling – approaches

152

As shown convincingly by the aforementioned experiments, several processes act

153

concurrently on a microstructure and control the dynamics of the microstructural evolution

154

and material properties (e.g. Means, 1977, 1980). While there are a large number of models

155

that focus on one process alone, there are fewer systems that aim to model the effect of

156

multiple interacting processes. Here, we focus on the latter. Solving all necessary equations

157

simultaneously is only practical when the number of interacting processes is small.

158

Alternatively, the set of equations, hence processes, needs to be reduced to a manageable

159

number necessitating a user driven selection which may result in the omission of important

160

feedbacks. Operator splitting provides an alternative approach, whereby each process acts

161

sequentially on a microstructure that was incrementally changed by all operating processes in

162

the previous step. In this case, the number of equations to be solved is not limited; hence

163

there is no need to artificially reduce the number of modelled processes. Clearly, length scale

164

and time steps have to be considered carefully for this approach to be valid. The operator

165

splitting approach is fundamental to the numerical platform Elle (Jessell et al., 2001, Bons et

166

al., 2008, Piazolo et al., 2010; www.elle.ws) which is, within the realm of microdynamic

167

modelling, the most used numerical platform in geology. However, operator splitting has also

168

been employed in other numerical schemes (e.g. Cross et al., 2015).

169 170

2.3. The numerical representation of a microstructure

171

It is necessary to describe numerically the microstructure so that in the model the processes

172

can act upon it spatially. Up to now, for rocks and ice, microstructural models have been

173

mostly restricted to two-dimensions. Two basic approaches can be taken: (i) a lattice data

174

structure or (ii) an element data structure. In the lattice data structure, the microstructure is

175

mapped onto an irregular or regular lattice, like the pixels in a digital image or gridded

176

analysis points obtained during chemical mapping and orientation mapping of geological

177

samples. Potts, cellular automata, lattice-spring, particle-in-cell, phase field and

178

micromechanical Fast Fourier Transform (FFT) based models utilize the lattice data structure

179

(Fig. 2; see also chapter 2 of Bons et al., 2008 for a detailed review of different numerical

180

methods). Numerically the lattice data structure is easy to use and calculations are

181

straightforward. In the element data structure, the microstructure is described by discrete

182

elements which can be points, line segments, or polygons (Fig. 2). Finite-element models and

183

boundary models typically use elements to describe a system. Element data structure is well

184

suited for processes acting upon a surface, e.g. grain boundary migration, but is numerically

185

more “expensive” due to necessary topology checks. Both elements and lattice points can

186

have, in addition to their location in xy space, specific properties or store specific state

187

variables, such as mineral phase, chemical composition, crystallographic orientation, and

188

stress and strain (rate) values, respectively. The numerical platform Elle (Jessell et al., 2001)

189

combines these two approaches: a microstructure is represented by boundary nodes

190

connected by straight segments (Fig.2). A polygon enclosed by boundary nodes represents a

191

grain or subgrain. Property variations within the individual polygons are defined at interior

192

points where information is recorded and tracked. Processes acting on the microstructure can

193

move boundaries and/or interior points and/or change the properties at those points, e.g.

194

composition, crystallographic orientation or dislocation density. The system has now matured

195

to a stage that it can robustly be used for a large variety of models relevant to geology.

196 197

3. Examples of modelling several coupled processes

198

3.1.

199

3.1.1. A historical perspective

200

Dynamic recrystallization is the response of a crystalline aggregate to lower its free energy by

201

formation and movement of sub- and grain boundaries (Means, 1983). It occurs by a number

202

of concurrent processes that act upon the polycrystalline material. For decades, geologists

203

have used microstructural characteristics developed during dynamic recrystallization to infer

204

dominance of processes and linked boundary conditions (e.g. Urai et al., 1986; Hirth and

205

Tullis, 1992; Stipp et al., 2002). Following the first attempts of such a model (Etchecopar,

206

1977), Jessell (1988a,b) and Jessell and Lister (1990) developed a numerical model

207

incorporating rotation of the crystal lattice and dynamic recrystallization for quartz. The data

208

structure used was a 100x100 hexagonal lattice structure where each lattice point represented

209

a grain or subgrain (Fig. 3a inset). Calculation of the crystal lattice rotation in Jessell and

210

Lister (1990) uses the Taylor-Bishop-Hill calculation method (Taylor, 1938; Bishop and Hill,

211

1951a, 1951b; Lister and Paterson, 1979) and the critical resolved shear stress of the different

212

slip systems in quartz. This work reproduced the general microstructures seen in a mylonite.

213

A major step forward was that the simulations showed that dynamic recrystallization can

214

significantly modify crystallographic preferred orientation (CPO) development. Piazolo et al.

215

(2002) advanced Jessell’s model by utilizing the more robust Elle data structure to describe

216

the local deformation and the dynamics of the grain boundary network (Jessell et al., 2001).

217

The model now included a finite element solution for incompressible, linear or nonlinear

218

viscous flow (BASIL; Barr and Houseman 1992, 1996) to compute the local velocity field,

219

the front tracking method for the motion of grain boundaries by moving nodes and segments

220

(e.g. Bons and Urai, 1992), the Taylor-Bishop-Hill formulation (Lister and Paterson, 1979)

221

and additional features such as crystal lattice rotation, formation of subgrains,

222

recrystallization by nucleation, grain boundary migration, recovery, work hardening and

Dynamic recrystallization

223

tracking of dislocation densities. This list of additional features shows the advance made in

224

approximately a decade in reproducing natural microstructures (Fig. 3b).

225

So far, the rheological anisotropy of minerals deforming by dislocation creep was not taken

226

into account in the calculation of the stress-strain rate field and lattice rotation. About another

227

decade after Piazolo et al. (2002), the viscoplastic implementation of the full-field crystal

228

plasticity micromechanical code based on Fast Fourier Transforms (VPFFT; Lebensohn,

229

2001) was introduced in geological microstructural modelling (Griera et al., 2011, 2013,

230

2015). The VPFFT is a spectral method that specifically assumes that deformation is

231

achieved by dislocation glide on crystallographic slip planes (each with their own critical

232

resolved shear stress) whose orientations are mapped on a regular grid (for details see Griera

233

et al. (2013), Montagnat et al. (2011, 2014) and Llorens et al. (2016a)). The simulations of

234

Griera et al. (2011, 2013) and Ran et al. (2018) illustrate that heterogeneous stress within and

235

between grains that emerge from inhomogeneous slip has a significant effect on the rotation

236

rate of porphyroclasts and -blasts and the strain (rate) field around these objects.

237

More recently, Llorens et al. (2016a; 2016b; 2017) and Steinbach et al. (2016; 2017) coupled

238

dynamic recrystallization (DRX) by grain boundary migration (GBM), intracrystalline

239

recovery and polygonisation with VPFFT crystal plasticity models, and applied them to study

240

the deformation of polar ice. Llorens et al. (2016a) presents a comprehensive description of

241

the method. They found that DRX produces large and equidimensional grains, but only

242

marginally affects the development of the c-axes CPO. However, DRX can alter the activity

243

of slip systems and does modify the distribution of a-axes. In simple shear, the strong

244

intrinsic anisotropic of ice crystals is transferred from the crystal to the polycrystal scale,

245

leading to strain localisation bands that can be masked by GBM (Llorens et al., 2016b).

246

Llorens et al. (2017) compared the dynamics of pure polar ice polycrystalline aggregates in

247

pure and simple shear deformation. It was found that, due to the vorticity of deformation, it is

248

expected that ice is effectively weaker in the lower parts of ice sheets (where simple shear

249

dominates) than in the upper parts (where ice is mostly deformed coaxially). The method was

250

also applied to reproduce the development of tilted-lattice (kink) bands found in ice cores

251

(Jansen et al., 2016) and crenulation cleavage during folding (Ran et al., 2018) as a result of

252

mechanical anisotropy. Steinbach et al. (2016) included air inclusions as a second phase to

253

simulate DRX in porous firn and found that DRX can occur despite the low strain and stress in

254

firn.

255

Apart from polar ice, the VPFFT-Elle model has also been used to analyse subgrain rotation

256

recrystallization of halite polycrystals in simple shear (Gomez-Rivas et al., 2017) by coupling

257

VPFFT with the Elle routines that simulate intracrystalline recovery (Borthwick et al, 2013)

258

and subgrain rotation. They found that recovery does not affect CPOs, but strongly decreases

259

grain size reduction. These authors also evaluated the use of mean subgrain misorientations

260

as a strain gauge.

261

3.1.2. Testing the effect of different combinations of recrystallization processes

262

We show results from three combinations of processes that occur during dynamic

263

recrystallization. These processes act on the same initial 10x10 cm aggregate of ice Ih

264

crystals (Fig. 4a). The numerical approach in these simulations is based on the VPFFT

265

micromechanical model in combination with several Elle processes (Llorens et al., 2016a,

266

2016b, 2017; Steinbach et al., 2016; 2017). The deformation-induced lattice rotation and the

267

estimation of geometrically necessary dislocation densities calculated from the stress and

268

velocity field provided by the VPFFT algorithm are used to simulate recrystallization by

269

intra-crystalline recovery, grain boundary migration and nucleation. Grain boundary

270

migration (GBM) is simulated based on the algorithm by Becker et al. (2008). Here, a front-

271

tracking approach is used in which the movement of boundaries is mapped through time.

272

Recovery reduces the intra-granular stored energy in a deformed crystal (Borthwick et al.,

273

2013). Nucleation creates new grain boundaries with areas of misorientation values above a

274

pre-defined threshold (Piazolo et al., 2002; Llorens et al., 2017; Steinbach et al., 2017).

275

Dextral simple shear deformation is applied in strain increments of ∆ = 0.04. Three different

276

combinations of recrystallization processes are tested: (1) GBM only, (2) GBM and recovery

277

or (3) GBM, recovery and nucleation (Fig. 4a). Simulations including VPFFT viscoplastic

278

deformation show a similar evolution of c-axis orientation, regardless of the dynamic

279

recrystallization processes included (Fig. 4a). The lack of influence of nucleation on the CPO

280

is due to the fact that new grains are modelled to have lattice orientations close to those of

281

their parent grains. However, inclusion of the nucleation process result in grain size reduction

282

(Fig. 4b). Results illustrate how different combinations of microdynamic processes affect the

283

microstructural characteristics (e.g. CPO, grain size, shape and orientation) in different ways

284

where no single microstructural parameter can grasp the full dynamics of polycrystalline

285

deformation.

286

3.2. Diffusion creep

287

Diffusion creep is a physico-chemical process that operates in large regions of the Earth – in

288

the upper Earth as “pressure solution” where diffusion is enhanced along fluid films and

289

elsewhere at low strain rates, maybe in large parts of the lower mantle (Karato, 1992). The

290

mechanism involves diffusion driven by gradients in normal stress along boundaries. In

291

simple models no other driving forces need to be included. It is implicit in our understanding

292

of this mechanism that grain boundary sliding occurs in conjunction with diffusion. Ford et

293

al. (2002) devised a numerical model to understand the evolution of grain shapes and CPO in

294

grain boundary diffusion creep. The mathematical framework allows for the consideration of

295

diffusion and sliding together. Operator splitting is not required and the evolution at each

296

timestep is based on a single matrix inversion operation. The model couples the

297

microstructure to an evolving system of local stresses (Fig. 5). The model predicts that grain

298

shapes become somewhat elongate, in accordance with experiments on calcite (Schmid et al.,

299

1987) and observations on an albite-bearing mylonite (Jiang et al., 2000) (Fig. 5). It also

300

predicts that grain rotations are not chaotic and that CPO may be present to high strains

301

(Wheeler, 2009), as seen in, for example, experiments on olivine-orthopyroxene (Sundberg

302

and Cooper, 2008) and predicted by numerical simulations of Bons and den Brok (1999). A

303

second phase, if insoluble, can be included and such a model was used by Berton et al. (2006,

304

2011) to prove the hypothesis that a different grain boundary diffusion coefficient along the

305

two-phase boundaries could explain fibre growth in pressure shadows. Not all model

306

predictions are in agreement with other investigations – for example some materials

307

deformed by diffusion creep e.g. olivine show quite equant grains (Karato et al., 1986). In

308

such experiments grain growth occurs but this process was not included in the numerical

309

model, hence results differ. In a later section we address this issue, but this example

310

illustrates that, in common with other numerical models, the applicability reflects the

311

processes included.

312

3.3. Stress driven dissolution, growth and dynamic roughening

313

Another example where stress leads to micro- and macrostructures is the roughening of grain

314

boundaries in the presence of a fluid. In this case dissolution and/or mineral growth are

315

driven not just by normal stress gradients at the interface (c.f. section 3.2) but also by changes

316

in elastic and surface energies. In addition to these driving forces, the detailed

317

electrochemistry of interfaces has a kinetic effect on rates (Gratier et al., 2013). In the

318

example presented here, the process is modelled in Elle by coupling a lattice spring code that

319

calculates the strain with a background fluid that dissolves lattice elements. This process can

320

produce transient patterns of interface geometry at the grain boundaries (Koehn et. al., 2006).

321

In addition, the process itself produces rough interfaces that can be seen on the larger scale as

322

stylolites (Koehn et al., 2007; Ebner et al., 2009). The scaling behaviour of stylolites is

323

important for stress inversion and compaction on the basin scale (Koehn et al., 2012; 2016)

324

and the shape of stylolites is partly rooted in the microstructure of the host-rock (Fig. 6).

325

Slower dissolving material on various scales, from small grains to fossils and layers, can

326

influence the dissolution process and thus the development of patterns, whereas deformation

327

leads to changes in long-range interactions (Koehn et al., 2016). Similarly, pinning behaviour

328

in the presence of a fluid is seen in grain growth where grain boundary pinning results in

329

restriction of grain growth. Pinning particles can also be moved around so that growing

330

grains become clean. Such a process can lead to layered rocks, for example, zebra dolomites

331

(Kelka et al., 2015).

332

4. Numerical simulations of microdynamic processes: Examples of ongoing work

333

4.1.

334

Even now, most of the numerical models that incorporated more than one process are

335

restricted to deformation processes alone. Although in diffusion creep there is an intrinsic

336

chemical aspect, the model described in 3.2 involves just a single phase of fixed composition.

337

There are some models that incorporate the growth of new mineral phases and investigate the

338

rheological effect of such growth (e.g. Groome et al., 2006; Smith et al., 2015), however the

339

location of new phase growth and/or growth rate is predefined. While Park et al. (2004)

340

showed convincingly the effect of microstructure on the diffusion pathways and chemical

341

patterns in garnet and biotite, there are now many opportunities available in a numerical

342

system such as Elle (Jessell et al., 2001) to couple local chemistry and microstructural

343

evolution. Below we present preliminary results from two current projects linking chemical

344

changes and a dynamically evolving grain network.

345

4.1.1. Grain boundary diffusion creep and grain growth

346

Here we present the first results using a model combining, using operator splitting,

347

deformation by diffusion creep (Ford et al., 2002; Wheeler, 2009) and surface energy driven

Linking chemistry and microstructural evolution

348

grain boundary migration (GBM), i.e. grain growth (modified after Becker et al., 2008). Four

349

different scenarios are shown: diffusion creep only, grain boundary migration only and two

350

combinations with medium and high grain boundary migration GBM rates (Fig. 7a). The

351

starting material is an almost regular hexagonal grain mesh with equant grains in which triple

352

junctions have had small random perturbations imposed (Fig. 7a). The perturbations are

353

required to avoid mathematical problems that arise in perfectly regular hexagonal networks

354

(Wheeler, 2010). Topology checks such as neighbour switching routines are performed after

355

each step in all simulations, ensuring that no topology problems arise. The microstructural

356

development is markedly different if grain growth is coupled with diffusion creep (Fig. 7a)

357

where with increasing GBM rate the aspect ratios are generally reduced relative to that of

358

comparable experiments modelling diffusion creep only. This is provisionally in accord with

359

the hypothesis that there will be a ‘steady state’ grain elongation developed which is a

360

function of the relative magnitudes of strain rate and grain growth kinetic parameters

361

(Wheeler, 2009) but more simulations are required (with less regular starting

362

microstructures). In contrast, for grain growth only, the microstructure does not change as the

363

initial configuration consists of stable near 120° triple junctions (Fig. 7a). Grain numbers (i.e.

364

grain size) do not change in any of the simulations.

365

4.1.2. Trace element partitioning between fluid and solid coupled with recrystallization

366

Field and experimental studies illustrate the importance of the time and length scales on the

367

evolution of grain networks during trace element diffusion (e.g. Ashley et al., 2014 and

368

references therein). Therefore, understanding the influence of recrystallization on trace

369

element distribution is essential for correctly interpreting patterns of trace element

370

distribution (e.g. Wark and Watson, 2006). However, to our knowledge there are currently no

371

numerical approaches able to simulate diffusion coupled with microstructure evolution.

372

Recently, a new Elle finite difference process that solves Fick’s second law to simulate trace

373

element diffusion in a polycrystalline medium has been developed. The polycrystal is defined

374

by n-phases where grain boundaries and grains are differentiated (Fig. 7b). The trace element

375

partitioning coefficient between different phases is also considered. The approach uses an

376

element data structure representing grain boundaries and a lattice data structure of a regular

377

grid of unconnected lattice point to track diffusion along the grain boundaries and the

378

physical and chemical properties within grains, respectively. To ensure elemental exchange

379

between grain boundaries and grain interiors, a grain interior-grain boundary partition

380

coefficient is implemented that takes the proximity to grain boundaries into account. This

381

allows the simulation of coupled bulk and grain boundary diffusion. This approach can be

382

fully coupled with other processes of the numerical platform Elle and therefore allows

383

simulating simultaneous diffusion, deformation and static or dynamic recrystallization. The

384

use of this approach is demonstrated through the modelling of diffusion of a tracer with

385

fractionation during static grain growth of a single-phase aggregate (Fig. 7b) (after that

386

presented by Jessell et al., 2001). Patterns of chemical concentration qualitatively resemble

387

those of Ti-distribution observed in recrystallized quartz in shear zones and can help to

388

understand the redistribution of Ti in quartz during dynamic recrystallization (e.g. Grujic et

389

al., 2011), among many other geological problems.

390

4.2.

391

We restricted this review to two-dimensional numerical approaches, as very few models have

392

been published that are in three dimensions and also specific to geological microstructures.

393

An exception is the phase field approach by Ankit et al. (2015) who investigate the growth of

394

crystals in an open fracture. However, within the material science community numerical

395

models exist that investigate the behaviour of a microstructure in response to single processes

396

in three dimensions (e.g. three dimensional crystal plasticity modelling platform DAMASK

397

(Roters et al., 2012; Eisenlohr et al., 2013), which includes the option of solving the

Modelling in three dimensions

398

micromechanical problem using a 3D implementation of the FFT-based formulation

399

(presently implemented in Elle in its 2-D version) and grain growth (e.g. Krill and Chen,

400

2002; Kim et al., 2006). With the technical advances made within the geological and material

401

science community it is now in principle possible to extend currently available 3D models to

402

geological questions and/or multi-process scenarios.

403

4.2.1. Dissolution of reactive surfaces in three dimensions

404

Mineral dissolution controls important processes in geoscience such as serpentinization (e.g.

405

Seyfried et al., 2007), retrogression and replacement reactions (Putnis and Austrheim, 2012),

406

deformation by dissolution-precipitation creep (e.g. Rutter, 1976) and also affects processes

407

relevant to society such as the stability of spent nuclear fuel and mine waste. Recent work has

408

shown that dissolution rates are linked to the evolving surface structure, and thus are time-

409

dependent (Godinho et al., 2012). Different surfaces have different structures in three

410

dimensions and the presence of etch pits play a major role in dissolution (Pluemper et al.,

411

2012). Hence, for accurate representation of dissolution behaviour, modelling in three

412

dimensions is important. Here, we present a numerical model that simulates the dissolution

413

process as a potential tool to quantify the links between dissolution rates, reactive surface

414

area and topography over periods of time beyond reasonable for a laboratory experiment. The

415

program uses empirical equations that relate the dissolution rate of a point of the surface with

416

its crystallographic orientation (Godinho et al., 2012) to simulate changes of topography

417

during dissolution, which ultimately results in the variation of the overall dissolution rate

418

(Godinho et al., 2012, 2014). The initial surface is composed of a group of nodes with a xy

419

position and a set height (z). For each lattice point a local surface orientation is calculated

420

from the inclination of the segment node and its neighbours. This orientation together with

421

the crystallographic orientation of the grain the surface belongs to is then used to calculate a

422

dissolution rate (surface reactivity). Based on this the displacement of the node using the

423

equations published in Godinho et al. (2012) is calculated. The model allows the graphical

424

display of the three-dimensional topographic development of the surface, tracking of the

425

variation of the surface area and calculation of the overall dissolution rate at each stage of the

426

simulation. Results obtained with the 3D simulation at three consecutive stages of dissolution

427

(150 hrs, 180 hrs, 276 hrs dissolution duration) show that numerical results are in accordance

428

to experimental results (Fig. 7c).

429 430

5. Numerical simulations of microstructures: Possibilities and challenges

431

5.1. Linking laboratory and natural data with numerical models

432

Over the last two decades numerical capabilities have advanced markedly and models have

433

come of age. Consequently, a large range of models and process combinations is now

434

available that can be utilized to gain further insight into the link between processes, material

435

properties and boundary conditions. This also includes polymineralic systems making the

436

models more appropriate to use to investigate processes occurring within polyphase rocks

437

(e.g. Roessiger et al., 2014; Steinbach et al, 2016). New avenues of effective verification

438

against laboratory and natural data are now opening up, due to the development of analytical

439

tools allowing rapid complete microstructural and microchemical analysis providing dataset

440

similar to those possible in numerical models (e.g. Steinbach et al, 2017, Piazolo et al., 2016);

441

many of the analytical data sets have a lattice data structure.

442

5.2. Linking chemistry and microstructural evolution

443

Advances in numerical methods, theoretical treatment of thermodynamic data, analytical

444

tools and theoretical and experimental insights into the coupling between chemical and

445

physical process that emerged over the last decade, call for a major effort in this area of

446

research. Examples of areas of opportunities include investigation of the (1) significance of

447

local stress versus bulk stress on mineral reactions and reaction rates (e.g. Wheeler, 2014), (2)

448

characteristics of replacement microstructures and their potential significance for

449

microstructural interpretation (e.g. Putnis, 2009, Spruzeniece et al., 2017), (3) mobility of

450

trace elements enhanced by deformation that change significantly the local elemental

451

distributions (e.g. Reddy et al., 2007, Piazolo et al., 2012, 2016) and (4) coupling between

452

reactive fluid-solid systems and hydrodynamics (Kelka et al., 2017). Increasing computer

453

processing speeds will aid the running of such models. However, the technical challenges

454

include the harmonisation of fundamentally different numerical approaches (crystal plasticity

455

versus diffusion creep, for example) and the resolution of fundamental mathematical

456

problems, e.g. the current lack of an internally consistent model for multiphase diffusion

457

creep, highlighted by Ford and Wheeler (2004).

458

5.3. Linking brittle and ductile deformation: elasto-viscoplastic behaviour

459

When examining the rock record, it is clear that in many cases, brittle and ductile behaviour

460

often occurs at the same time within a rock (e.g. Bell and Etheridge, 1973, Hobbs et al.,

461

1976). With the potential significance of grain-scale brittle behaviour now measurable on

462

seismic signals (e.g. Fagereng et al., 2014) there is an increased need in developing numerical

463

techniques that allow us to model the dynamic link between brittle and ductile behaviour.

464

Such elasto-viscoplastic behaviour combines the elastic reversible fast deformation with a

465

viscous time dependent flow and a plastic behaviour. These behaviours can be included in

466

continuous or discontinuous models. For example, in the numerical platform Elle, a lattice

467

structure is used to deform the model elastically up to a critical stress where bonds fracture

468

(plastic behaviour) and the particles themselves deform as a function of stress and time

469

(viscous behaviour). In this case the viscous behaviour conserves the volume whereas shear

470

forces and differential stresses converge to zero (Sachau and Koehn, 2010, 2012; Arslan et

471

al., 2012; Koehn and Sachau, 2014). Alternatively, linking the elasto-visocplastic FFT based

472

(EVPFFT) (Lebensohn et al., 2012) with models such as Elle would allow calculation of the

473

Cauchy stresses that are the local driving force for grain scale damage processes.

474

One of the major challenges is the large range of time-scales in these processes, with

475

fracturing and fluid flow on fast to intermediate and viscous deformation and potentially

476

reactions on very long time scales. This complexity either requires the assumption that some

477

processes are instantaneous or it requires an “up-scaling in time” or non-linear time scales in

478

models.

479 480

5.4. Expansion of capability to three dimensions

481

With the advent of supercomputers and new numerical approaches now is the time to develop

482

techniques to investigate microstructural development in three dimensions. This is of

483

particular importance if material transport such as aqueous fluid and/or melt flow is to be

484

considered. This would enable, for example, modelling of strain fringes, shear veins and en-

485

echelon tension gashes. However, these require models that link brittle and ductile behaviour

486

along with modelling in three dimensions.

487

One of the biggest problems faced with three-dimensional models of microstructures are the

488

three-dimensional topology changes that are common in dynamic microstructural

489

development (e.g. Fig. 1-7). This is a major problem if an element data structure with

490

segments, i.e. grain boundaries is used. However, the phase field approach does not work

491

with such distinct boundaries, and is therefore well suited for 3D problems (e.g. Ankit et al.

492

2014). In addition, a three dimensional network of unconnected nodes, in which there is no

493

physical movement of boundaries but only changes in the properties of the 3D nodes (voxels)

494

(cf. Fig. 7c) may be a way forward (e.g. Sachau and Koehn, 2012).

495 496

5.5. Link between geophysical signals and microstructure

497

At a time where there is an ever-increasing amount of geophysical data being collected,

498

numerical simulations that are used to test the link between microstructural development and

499

geophysical signal will become increasingly important. Cyprych et al. (2017) showed that not

500

only crystallographic preferred orientation but also the spatial distribution of phases i.e. the

501

microstructure, has a major impact on seismic anisotropy. Therefore, a direct link between

502

the microdynamic models such as Elle allowing tracking of microstructural changes through

503

time and space and geophysical signal generation offers a wealth of new opportunities

504

including interpretation of strong reflectors in the lower crust and mantle. Current efforts by

505

Johnson, Gerbi and co-workers (Naus-Thijsen et al., 2011; Cook et al., 2013; Vel et al., 2016)

506

are in line with this direction.

507 508

5.6. Application of numerical models to polymineralic rock deformation and ice-related

509

questions

510

The dynamic behaviour of the Earth is strongly influenced by the deformation of

511

polymineralic rocks. At the same time Earth’s polar ice caps and glacial ice which often

512

include ice and a second phase (e.g. air inclusions, dust, entrained bedrock) is of major

513

importance to society, especially in view of changing climate (e.g. Petit et al., 1999; EPICA,

514

2004). Application of the current numerical capabilities to polycrystalline ice and

515

polymineralic rocks is therefore urgently needed. Over the last years, there has been an

516

increased effort in this direction (Roessiger et al., 2011, 2014; Piazolo et al., 2015; Llorens et

517

al., 2016a, 2016b, 2017; Jansen et al., 2016; Steinbach et al. 2016, 2017), which promises to

518

continue. Here, the development of the link between elemental mobility and microstructural

519

development is of major importance, as only with such models the can chemical signals of,

520

for example, ice cores be correctly interpreted.

521

522

5.7. Upscaling: Utilizing operator splitting and utilities developed for microdynamic

523

systems to larger-scale problems

524

One of the strengths of the numerical approach taken by the microstructural community has

525

been the close link between different processes and the ability of the models to take into

526

account the local differences in properties such as stress, strain and chemistry. The technique

527

of operator splitting has proven extremely powerful. Furthermore, the ability to model

528

anisotropic material behaviour utilizing for example VPFFT viscoplastic deformation

529

formulations (Lebensohn, 2001), has enabled realistic and dynamic models. Upscaling this

530

approach to investigate problems at a large scale e.g. folding (Llorens et al., 2013a, 2013b;

531

Bons et al., 2016; Ran et al., 2018) and shear deformation (Gardner et al., 2017) have shown

532

to be very beneficial. There is great scope to expand further on this in view of fluid flow,

533

mineralization and fault formation.

534 535

6. Numerical simulations of microstructures: Lessons learnt and future challenges

536

Numerical simulations of microstructural development have caught our imagination over the

537

last three decades. They have markedly advanced our ability to explain phenomena and

538

patterns we observe in nature and experiments by allowing us to test the link between

539

boundary conditions, material properties, processes, and microstructural development.

540

Importantly, models, especially those that couple several process and/or investigate pre-

541

existing heterogeneities can train the geologist to think of the dynamics of the system rather

542

than a linear development. For example, different patterns of strain localization observed in

543

nature can be explained by differences in the relative rates of interacting processes (e.g.

544

Jessell et al., 2005; Gardner et al., 2017). At the same time, specific indicative microstructural

545

parameters can be developed to help interpret natural microstructures (e.g. Piazolo et al.,

546

2002; Gomez-Rivas et al., 2017, Llorens et al., 2017; Steinbach et al., 2017).

547

However, including chemistry coupled to other processes remains a particular challenge.

548

Whilst, for example, trace-element diffusion can be enacted in parallel with other processes

549

(section 4.1.2), chemical transport of major elements and diffusion creep cannot yet be fully

550

integrated with many other processes. Indeed, when multiphase systems are considered, there

551

are unsolved problems with diffusion creep modelling even in the absence of other processes

552

(Ford and Wheeler, 2004). This challenge is closely linked to our current inability to

553

confidently model grain boundary sliding.

554

Nevertheless, the studies we describe have shown that numerical models are extremely

555

powerful in providing benchmark results to investigate what kind of microstructure may

556

develop under certain conditions. These models are sophisticated mind experiments that are

557

firmly based on physical and chemical laws for which the theory is well known individually

558

but their interaction is difficult to predict analytically.

559 560

Acknowledgements:

561

We would like to thank Win Means for opening up a whole new perspective on

562

microstructures with his inspirational in-situ experiments. The authors thank the DFG, ARC,

563

ESF, NSF, EU through Marie Curie Fellowship to SP, the Government of Catalonia's

564

Secretariat for Universities and Research for a Beatriu de Pinós fellowship to EGR (2016 BP

565

00208), and NERC (NERC grant NE/M000060/1) for support of the numerical endeavours.

566

The authors thank C. Gerbi for his helpful and constructive review as well as C. Passchier for

567

editorial handling of the manuscript.

568 569 570

References

571 572

Ankit, K., Urai, J. L., Nestler, B. 2015. Microstructural evolution in bitaxial crack‐ seal veins: A phase-field study. Journal of Geophysical Research 120, 3096-3118.

573

Arslan, A., Koehn, D., Passchier, C.W., Sachau, T. 2012. The transition from single layer to

574

foliation boudinage: A dynamic modelling approach. Journal of Structural Geology 42,

575

118-126.

576

Ashley, K.T., Carlson, W.D., Law, R.D., Tracy, R.J. 2014. Ti resetting in quartz during

577

dynamic recrystallization: mechanisms and significance. American Mineralogist 99, 2025-

578

2030.

579 580 581 582 583

Barr, T.D., Houseman, G.A. 1992. Distribution of deformation around a fault in a non-linear ductile medium. Geophys Res Lett 19, 1145-1148. Barr, T.D., Houseman, G.A. 1996. Deformation fields around a fault embedded in a nonlinear ductile medium. Geophys J Int 125, 473-490. Becker, J.K., Bons, P.D., Jessell, M.W. 2008. A new front-tracking method to model

584

anisotropic grain and phase boundary motion in rocks. Computers & Geosciences 34, 201-

585

212.

586 587 588 589 590

Bell, T. H., Etheridge, M. A. 1973. Microstructure of mylonites and their descriptive terminology. Lithos 6, 337-348. Berton, J. R., Durney, D. W., Wheeler, J., Ford, J. M., 2006. Diffusion-creep modelling of fibrous pressure-shadows. Tectonophysics 425, 191-205. Berton, J. R., Durney, D. W., Wheeler, J., 2011. Diffusion-creep modelling of fibrous

591

pressure-shadows II: influence of inclusion size and interface roughness. In: Deformation

592

Mechanisms, Rheology & Tectonics: Microstructures, Mechanics & Anisotropy (eds

593

Prior, D. J., Rutter, E. H. & Tatham, D.), Geological Society Special Publications, 360,

594

.319-328, .

595 596 597 598 599 600 601 602 603 604 605 606 607

Bishop J.F.W., Hill R. 1951a. A theory of the plastic distortion of a polycrystalline aggregate under combined stresses. Philosophical Magazine 42, 414–427. Bishop, J.F.W., Hill, R. 1951b A Theoretical Derivation of the Plastic Properties of a Polycrystalline Face-Centred Metal. Philosophical Magazine 42, 1298-1307. Bons, P.D., den Brok, S.W.J. 2000. Crystallographic preferred orientation development by dissolution-precipitation creep. Journal of Structural Geology 22, 1713-1722. Bons, P.D., Jessell, M.W. 1999. Micro-shear zones in experimentally deformed octachloropropane. Journal of Structural Geology 21, 323-334. Bons, P.D., Urai J.L. 1992. Syndeformational grain growth: microstructures & kinetics. Journal of Structural Geology 14, 1101-1109. Bons, P.D., Koehn, D., Jessell, M.W. (Eds) (2008) Microdynamic Simulation. Lecture Notes in Earth Sciences 106, Springer, Berlin. 405 pp. ISBN 978-3-540-44793-1. Bons, P.D., Jansen, D., Mundel, F., Bauer, C.C., Binder, T., Eisen, O., Jessell, M.W.,

608

Llorens, M.G., Steinbach, F., Steinhage, D. Weikusat, I., 2016. Converging flow and

609

anisotropy cause large-scale folding in Greenland's ice sheet. Nature communications 7,

610

11427.

611

Borthwick, V., Piazolo, S., Evans, L., Griera, A., Bons, P.D. 2013. What happens to

612

deformed rocks after deformation? A refined model for recovery based on numerical

613

simulations. In: "Deformation Structures and Processes within the Continental Crust",

614

Special Volume 394, Geol. Soc. London. doi: 10.1144/SP394.11

615 616 617

Cloos, E. 1955. Experimental analysis of fracture patterns. Geological Society of America Bulletin 66, 241-256. Cook, A., Vel., S., Johnson, S.E., Gerbi, C., Song, W.J., 2013, Elastic and Seismic Properties

618

(ESP) Toolbox (beta version), http://umaine.edu/mecheng/faculty-andstaff/

619

senthil-vel/software/ESP_Toolbox/

620

Cross, A. J., Ellis, S., Prior, D. J. 2015. A phenomenological numerical approach for

621

investigating grain size evolution in ductiley deforming rocks. Journal of Structural

622

Geology 76, 22-34.

623

Cyprych, D., Piazolo, S., Almqvist, B.S.G. 2017. Seismic anisotropy from compositional

624

banding in granulites from the deep magmatic arc of Fiordland, New Zealand, Earth and

625

Planetary Science Letters 477, 156-167.

626

Ebner, M., Koehn, D., Toussaint, R, Renard, F. 2009. The influence of rock heterogeneity on

627

the scaling properties of simulated and natural stylolites. Journal of Structural Geology 31,

628

72-82.

629

Eisenlohr, P., Diehl, M., Lebensohn, R.A., Roters, F., 2013. A spectral method solution to

630

crystal elasto-viscoplasticity at finite strains. International Journal of Plasticity 46, 37-53.

631

EPICA Community members (2004) Eight glacial cycles from an Antarctic ice core. Nature

632 633 634 635 636 637 638 639 640

429, 623-628. Etchecopar, A. 1977. A plane kinematic model of progressive deformation in a polycrystalline aggregate. Tectonophysics 39, 121-139. Fagereng, Å., Hillary, G. W. B., Diener, J. F. A. 2014. Brittle-viscous deformation, slow slip and tremor. Geophysical Research Letters 41, 4159-4167. Ford, J. M., Wheeler, J., 2004. Modelling interface diffusion creep in two-phase materials. Acta Materialia 52, 2365-2376. Ford, J. M., Wheeler, J., Movchan, A. B., 2002. Computer simulation of grain boundary creep. Acta Materialia 50, 3941-3955.

641

Ford, J.M., Ford, N.J., Wheeler, J. 2004. Simulation of grain boundary diffusion creep:

642

analysis of some new numerical techniques: Proc R Soc, London 460, 2395-2413

643

Gardner, R., Piazolo, S., Evans, L., Daczko, N. 2017. Patterns of strain localization in

644

heterogeneous, polycrystalline rocks - a numerical perspective, Earth and Planetary

645

Science Letters, 463, 253-265.

646

Gratier, J.P., Dysthe, D.K., Renard, F., 2013. The Role of Pressure Solution Creep in the

647

Ductility of the Earth's Upper Crust, in: Dmowska, R. (Ed.), Advances in Geophysics, Vol

648

54. Elsevier, Amsterdam, pp. 47-179.

649 650 651

Griera, A., Bons, P.D., Jessell, M.W., Lebensohn, R.A., Evans, L., Gomez-Rivas, E. 2011. Strain localization and porphyroclast rotation. Geology 39, 275-278. Griera, A., Llorens, M.-G., Gomez-Rivas, E., Bons, P.D., Jessell, M.W., Evans, L.A.,

652

Lebensohn, R. 2013. Numerical modelling of porphyroclast and porphyroblast rotation in

653

anisotropic rocks. Tectonpophysics 587, 4-29,

654

Griera, A., Gomez-Rivas, E., Llorens, M.-G., Bons, P.D., Lebensohn, R. 2015.

655

Microstructural evolution of polycrystalline materials in simple shear: insights from full-

656

field numerical simulations. Geotectonic Research 97, 37–39.

657 658 659

Gottstein, G. 2004. Integral materials modelling. Modelling and Simulation in Materials Science and Engineering, 12, I. Godinho, J.R., Piazolo, S., Evans, L., 2014. Simulation of surface dynamics during

660

dissolution as a function of the surface orientation: Implications for non-constant

661

dissolution rates. Earth and Planetary Science Letters 408, 163-170.

662 663 664

Godinho, J.R.A., Piazolo, S., Evins, L.Z. 2012. Effect of surface orientation on dissolution rates and topography of CaF 2, Geochimica et Cosmochimica Acta 86, 392-403. Gomez-Rivas, E., Griera, A., Llorens, M-.G., Bons, P.D., Lebensohn, R.A., and Piazolo, S.

665

(2017). Subgrain rotation recrystallization during shearing: insights from full-field

666

numerical simulations of halite polycrystals. Journal of Geophysical Research, in press.

667

Grujic, D., Stipp, M., Wooden, J.L., 2011. Thermometry of quartz mylonites: Importance of

668

dynamic recrystallization on Ti‐ in‐ quartz reequilibration. Geochemistry, Geophysics,

669

Geosystems, 12.

670

Groome, W. G., Johnson, S. E., Koons, P. O. 2006. The effects of porphyroblast growth on

671

the effective viscosity of metapelitic rocks: implications for the strength of the middle

672

crust. Journal of Metamorphic Geology 24: 389–407.

673

Herwegh, M., Handy, M. R. 1998. The evolution of high-temperature mylonitic microfabrics:

674

Evidence from simple shearing of a quartz analogue (norcamphor). J. Struct. Geol. 18,

675

689-710.

676

Herwegh, M., Handy M. R.. 1998 The origin of shape preferred orientations in mylonite:

677

inferences from in-situ experiments on polycrystalline norcamphor. Journal of Structural

678

Geology 20, 681-694.

679

Hilgers, C., Urai, J.L., Post, A.D., Bons, P.D. 1997. Fibrous vein microstructures:

680

Experimental and numerical simulation. Aardkundige Mededelingen 8, 107-109.

681

Hirth, G., Tullis, J. 1992. Dislocation creep regimes in quartz aggregates. Journal of

682 683 684

Structural Geology 14, 145 - 159. Hobbs B.E., Means W.O., Williams P.F. 1976. An Outline of Structural Geology. John Wiley and Sons, 571 pp.

685

Jansen, D., Llorens, M.-G., Westhoff, J., Steinbach, F., Kipfstuhl, S., Bons, P.D., Griera, A.,

686

Weikusat, I. 2016. Small-scale disturbances in the stratigraphy of the NEEM ice core:

687

observations and numerical model simulations. The Cryosphere 10, 359-370.

688

Jiang, Z., Prior, D. J., Wheeler, J., 2000. Albite crystallographic preferred orientation and

689

grain misorientation distribution in a low-grade mylonite: implications for granular flow.

690

Journal of Structural Geology 22, 1663-1674.

691 692 693 694 695 696 697 698

Jessell, M.W. 1986. Grain boundary migration and fabric development in experimentally deformed octachloropropane. Journal of Structural Geology 8, 527-542. Jessell, M. W. 1988a. Simulation of fabric development in recrystallizing aggregates—I. Description of the model. Journal of Structural Geology 10, 771-778. Jessell, M. W. 1988b. Simulation of fabric development in recrystallizing aggregates—II. Example model runs. Journal of Structural Geology 10, 779-793. Jessell, M. W., Lister, G. S. 1990. A simulation of the temperature dependence of quartz fabrics. Geological Society, London, Special Publications 54, 353-362.

699

Jessell, M., Bons, P.D., Evans, L., Barr, T., Stüwe, K. 2001. Elle: the numerical simulation of

700

metamorphic and deformation microstructures. Computers And Geosciences 27, 17-30,

701

http://dx.doi.org/10.1016/S0098-3004(00)00061-3

702

Jessell, M.W., Siebert, E., Bons, P.D., Evans, L., Piazolo, S. 2005. A new type of numerical

703

experiment on the spatial and temporal patterns of localization of deformation in a

704

material with a coupling of grain size and rheology. Earth and Planetary Science Letters

705

239, 309-326.

706

Karato, S. I., Paterson, M. S., Fitz Gerald, J. D., 1986. Rheology of synthetic olivine

707

aggregates: influence of grain size and water. Journal of Geophysical Research 91, 1851-

708

8176.

709 710 711 712 713 714

Karato, S.I., Li, P., 1992. Diffusion creep in perovskite: implications for the rheology of the lower mantle. Science 255, 1238. Kelka, U., Koehn, D., Beaudoin, N. 2015. Zebra pattern in rocks as a function of grain growth affected by second-phase particles. Frontiers in Physics 3, 74. Kelka, U., Veveakis, M., Koehn, D., Beaudoin, N., 2017. Zebra rocks: compaction waves create ore deposits. Scientific Reports 7, 14260.

715

Kim, S. G., Kim, D. I., Kim, W. T., Park, Y. B. 2006. Computer simulations of two-

716

dimensional and three-dimensional ideal grain growth. Physical Review E 74, 061605.

717

Koehn, D., Bons, P.D., Passchier, C.W. 2003. Development of antitaxial strain fringes during

718

non-coaxial deformation: an experimental study. Journal of Structural Geology 25, 263-

719

275.

720

Koehn, D., Renard, F., Toussaint, R., Passchier, C. 2007. Growth of stylolite teeth patterns

721

depending on normal stress and finite compaction. Earth and Planetary Science Letters,

722

257, 582-595.

723 724 725

Koehn, D., Ebner, M., Renard, F., Toussaint, R., Passchier, W. 2012. Modelling of stylolite geometries and stress scaling. Earth and Planetary Science Letters 341, 104-113. Koehn, D., Sachau, T. 2014. Two-dimensional numerical modeling of fracturing and shear

726

band development in glacier fronts. Journal of Structural Geology 61, 133-142.

727

(doi:10.1016/j.jsg.2012.11.002).

728

Koehn, D., Rood, M.P., Beaudoin, N. , Chung, P., Bons, P.D., Gomez-Rivas, E. (2016) A

729

new stylolite classification scheme to estimate compaction and local permeability

730

variations. Sedimentary Geology 346, 60-71.

731 732 733 734 735

Krill Iii, C. E., Chen, L. Q. 2002. Computer simulation of 3-D grain growth using a phasefield model. Acta Materialia 50, 3059-3075. Lebensohn, R. A. 2001. N-site modeling of a 3D viscoplastic polycrystal using fast Fourier transform. Acta Materialia 49, 2723-2737. Lebensohn, R.A., Kanjarla, A.K., Eisenlohr, P., 2012. An elasto-viscoplastic formulation

736

based on fast Fourier transforms for the prediction of micromechanical fields in

737

polycrystalline materials. International Journal of Plasticity 32, 59-69.

738 739

Lister GS, Paterson MS 1979. The simulation of fabric development during plastic deformation and its application to quartzite: fabric transitions. J Struct Geol 1, 99-115.

740

Llorens, G.-M., Griera, A., Weikusat, I., Bons, P.D., Roessiger, J., Lebensohn, R.A. 2016a.

741

Dynamic recrystallisation of ice aggregates during co-axial viscoplastic deformation: a

742

numerical approach. Journal of Glaciology 62, 359-377.

743

Llorens, G.-M., Griera, A., Bons, P.D., Lebensohn, R.A., Evans, L.A., Jansen, D., Weikusat,

744

I. 2016b. Full-field predictions of ice dynamic recrystallisation under simple shear

745

conditions. Earth and Planetary Science Letters 450, 233-242.

746

Llorens, M. -G., Griera, A., Steinbach, F., Bons, P. D., Gomez-Rivas, E., Jansen, D.,

747

Roessiger, J., Lebensohn, R. A. and Weikusat, I. 2017. Dynamic recrystallization during

748

deformation of polycrystalline ice: insights from numerical simulations. Phil. Trans. R.

749

Soc. A 375, 20150346.

750 751 752 753

Llorens, M.G., Bons, P.D., Griera, A., Gomez-Rivas, E. and Evans, L.A., 2013a. Single layer folding in simple shear. Journal of Structural Geology 50, 209-220. Llorens, M-.G., Bons, P.D., Griera, A. Gomez-Rivas, E. 2013b. When do folds unfold during progressive shear?. Geology 41, 563-566

754

McCrone WC, Chen PT .1949. Grain growth in octochloropropane. J Appl Phys 20, 230-231.

755

Means, W.D. 1977. A deformation experiment in transmitted light. Earth Planet. Sci. Letts.,

756 757 758 759 760

35, 169-179. Means W.D. 1980. High temperature simple shearing fabrics: a new experimental approach. J. Struct. Geol. 2, 197-202. Means W.D. 1983. Microstructure and micromotion in recrystallization flow of octachloropropane: a first look. Geol. Rundsch. 72, 511-528.

761

Means, W. D., Ree, J. H. 1988. Seven types of subgrain boundaries in

762

octachloropropane. Journal of Structural Geology 10, 765-770.

763

Montagnat, M., Blackford, J.R., Piazolo, S., Arnaud, L., Lebensohn, R.A., 2011.

764

Measurements and full-field predictions of deformation heterogeneities in ice. Earth and

765

Planetary Science Letters 305, 153-160.

766

Montagnat, M., Castelnau, O., Bons, P.D., Faria, S.H., Gagliardini, O., Gillet-Chaulet, F.,

767

Grennerat, F., Griera, A., Lebensohn, R.A., Moulinec, H., Roessiger, J., Suquet, P. 2014.

768

Multiscale modeling of ice deformation behavior. Journal of Structural Geology 61, 78-

769

108.

770

Nama, T.N., Otoh, S., Masuda, T. , 1999. In-situ annealing experiments of octachloropropane

771

as a rock analogue: kinetics and energetics of grain growth. Tectonophysics 304, 57-70.

772

Naus-Thijssen, F.M., Goupee, A.J., Johnson, S.E., Vel, S.S., Gerbi, C., 2011. The influence

773

of crenulation cleavage development on the bulk elastic and seismic properties of

774

phyllosilicate-rich rocks. Earth and Planetary Science Letters 311, 212-224.

775 776 777 778 779 780 781

Park Y, Means WD. 1996. Direct observation of deformation processes in crystal mushes. J Struct Geol 18, 847-858. Park, Y., Park, D., Evans, L., Ree, J. H. 2004. An Elle-based 2-D model for cation exchange reaction between garnet and biotite. Journal of the Virtual Explorer 15. Passchier, C.W., Trouw, R.A.J., 2005. Microtectonics (2nd enlarged edition). Springer Verlag, Berlin. Petit, J.R., J. Jouzel, D. Raynaud, N.L. Barkov, J.-M. Barnola, I. Basile, M. Bender, J.

782

Chappellaz, M. Davis, G. Delaygue, M. Delmotte, V. M. Kotlyakov, M. Legrand, V.Y.

783

Lipenkov, C. Lorius, L.Pépin, C. Ritz, E.Saltzman, M. Stievenard 1999, Climate and

784

atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature

785

399, 429-436.

786

Piazolo, S., Austrheim, H., Whitehouse, M., 2012. Brittle-ductile microfabrics in naturally

787

deformed zircon: Deformation mechanisms and consequences for U-Pb dating. American

788

Mineralogist 97, 1544-1563.

789

Piazolo, S., Bons, P.D., Jessell, M.W., Evans, L., Passchier, C.W. 2002. Dominance of

790

microstructural processes and their effect on microstructural development: insights from

791

numerical modelling of dynamic recrystallization. Geol. Soc, London, Spec. Publ. 200,

792

149-170.

793

Piazolo, S., Jessell, M. J., Prior, D. J., Bons, P. D. 2004. The integration of experimental in-

794

situ EBSD observations and numerical simulations: a novel technique of microstructural

795

process analysis. Journal of Microscopy, 213, 273-284.

796

Piazolo, S., Jessell, M.W., Bons, P.D., Evans, L., Becker, J.K. 2010. Numerical simulations

797

of microstructures using the Elle platform: A modern research and teaching tool. Journal

798

of the Geological Society of India 75, 110-127.

799

Piazolo, S., La Fontaine, A., Trimby, P., Harley, S., Yang, L., Armstrong, R. and Cairney,

800

J.M., 2016. Deformation-induced trace element redistribution in zircon revealed using

801

atom probe tomography. Nature communications, 7. Piazolo, S., Kaminsky, F.V., Trimby,

802

P., Evans, L., Luzin, V., 2016. Carbonado revisited: Insights from neutron diffraction,

803

high resolution orientation mapping and numerical simulations. Lithos 265, 244-256.

804

Piazolo, S., Montagnat, M., Grennerat, F., Moulinec, H., Wheeler, J., 2015. Effect of local

805

stress heterogeneities on dislocation fields: examples from transient creep in

806

polycrystalline ice. Acta Materialia 90, 303-309.

807 808 809 810

Plümper, O., Røyne, A., Magrasó, A., Jamtveit, B. 2012. The interface-scale mechanism of reaction-induced fracturing during serpentinization. Geology 40, 1103-1106. Putnis, A. 2009. Mineral replacement reactions. Reviews in mineralogy and geochemistry 70, 87-124.

811 812 813 814 815

Putnis, A., Austrheim, H. 2010. Fluid‐ induced processes: metasomatism and metamorphism. Geofluids 10, 254-269. Ramberg, H., 1981. Gravity, deformation, and the earth's crust: In theory, experiments, and geological application. Academic press. Ran, H., de Riese, T., Llorens, M.-G., Finch, M., Evans, L.A., Gomez-Rivas, E., Griera, A.,

816

Jessell, M.W., Lebensohn, R.A., Piazolo, S., and Bons, P.D. 2018. Time for anisotropy:

817

The significance of mechanical anisotropy for the development of deformation structures.

818

Journal of Structural Geology, in press.

819 820

Ree, J. H., Park, Y., 1997. Static recovery and recrystallization microstructures in sheared octachloropropane. Journal of Structural Geology 19, 1521–1526.

821

Reddy, S. M., Timms, N. E., Pantleon, W., Trimby, P. 2007. Quantitative characterization of

822

plastic deformation of zircon and geological implications. Contributions to Mineralogy

823

and Petrology 153, 625-645.

824

Roessiger, J., Bons, P.D., Griera, A., Jessell, M.W., Evans, L. Montagnat, M., Kipfstuhl, S.,

825

Faria, S.H., Weikusat, I. 2011. Competition between grain growth and grain size reduction

826

in polar ice. Journal of Glaciology 57 942-948.

827 828 829 830 831

Roessiger, J., Bons, P.D., Faria, S.H. 2014. Influence of bubbles on grain growth in ice. Journal of Structural Geology 61, 123-132. Rosenberg, C. L. 2001. Deformation of partially molten granite: a review and comparison of experimental and natural case studies. International Journal of Earth Sciences 90, 60-76. Rosenberg, C.L., Handy, M.R. 2000. Syntectonic melt pathways during simple shearing of a

832

partially molten rock analogue (Norcamphor‐ Benzamide). J. Geophys. Res. 105, 3135-

833

3149.

834

Rosenberg, C.L., Handy, M.R. 2001. Mechanisms and orientation of melt segregation paths

835

during pure shearing of a partially molten rock analog (norcamphor–benzamide). J. Struct.

836

Geol. 23, 1917-1932.

837

Roters, F., P. Eisenlohr, C. Kords, D. D. Tjahjanto, M. Diehl, D. Raabe. 2012. DAMASK:

838

the Düsseldorf Advanced MAterial Simulation Kit for studying crystal plasticity using an

839

FE based or a spectral numerical solver. Procedia IUTAM 3, 3-10.

840 841

Rutter, E.H., 1976. The kinetics of rock deformation by pressure solution. Philosophical Transactions of the Royal Society, London, 203-219.

842

Sachau, T., Koehn, D. 2010. Faulting of the lithosphere during extension and related rift-

843

flank uplift: a numerical study. International Journal of Earth Sciences 99, 1619-1632.

844

Sachau, T., Koehn, D. (2012) Melange: a viscoelastic lattice-particle model applicable to the

845

lithosphere. Geochemistry, Geophysics, Geosystems 13, 1-26.

846

Schenk, O., Urai, J. L., 2004. Microstructural evolution and grain boundary structure during

847

static recrystallization in synthetic polycrystals of sodium chloride containing saturated

848

brine. Contributions to Mineralogy and Petrology 146, 671–682.

849

Schenk, O., Urai, J.L. 2005. The migration of fluid-filled grain boundaries in recrystallizing

850

synthetic bischofite: first results of in-situ high-pressure, high-temperature deformation

851

experiments in transmitted light. J. Metamorph. Geol. 23, 695-709.

852 853 854

Schmid, S.M., Panozzo, R., Bauer, S., 1987. Simple shear experiments on calcite rocks: rheology and microfabric. Journal of structural Geology 9, 747-778. Seyfried, W.E., Foustoukos, D.I., Fu, Q., 2007. Redox evolution and mass transfer during

855

serpentinization: An experimental and theoretical study at 200° C, 500bar with

856

implications for ultramafic-hosted hydrothermal systems at Mid-Ocean

857

Ridges. Geochimica et Cosmochimica Acta 71, 3872-3886.

858

Smith, J.R., Piazolo, S., Daczko, N.R., Evans, L., 2015. The effect of pre‐ tectonic reaction

859

and annealing extent on behaviour during subsequent deformation: insights from paired

860

shear zones in the lower crust of Fiordland, New Zealand. Journal of Metamorphic

861

Geology 33, 557-577.

862

Spruzeniece, L., Piazolo, S., Maynard-Casely, H.E., 2017. Deformation-resembling

863

microstructure created by fluid-mediated dissolution-precipitation reactions, Nature

864

Communications 8.

865

Steinbach, F., Bons, P.D., Griera, A., Jansen, D., Llorens Verde, M.G., Roessiger, J.,

866

Weikusat, I., 2016. Strain localization and dynamic recrystallization in the ice–air

867

aggregate: a numerical study. The Cryosphere 10, 3071-3089.

868

Steinbach F., Kuiper, E-J. N., Eichler, J., Bons P.D., Drury M. R., Griera, A., Pennock, G.M.,

869

Weikusat I., 2017. The Relevance of Grain Dissection for Grain Size Reduction in Polar

870

Ice: Insights from Numerical Models and Ice Core Microstructure Analysis. Frontiers in

871

Earth Science 5, 66.

872

Stipp, M., Stünitz, H., Heilbronner, R., Schmid, S.M. 2002. The eastern Tonale fault zone: a

873

'natural laboratory' for crystal platic deformation of quartz over a temperature range from

874

250 to 750 °C. J. Struct. Geol. 24, 1861-1884.

875

Sundberg, M. & Cooper, R. F., 2008. Crystallographic preferred orientation produced by

876

diffusional creep of harzburgite: Effects of chemical interactions among phases during

877

plastic flow. Journal of Geophysical Research-Solid Earth 113.

878 879

ten Brink CE, Passchier CW. 1995. Modelling of mantle porphyroclasts using nonNewtonian rock analogue materials. J Struct Geol 17, 131-146

880

Taylor GI. 1938. Plastic strain in metals. Journal of the Institute of Metals 62, 307–324.

881

Urai, J.L., 1983. Water assisted dynamic recrystallization and weakening in polycrystalline

882

bischofite. Tectonophysics 96, 125-157.

883 884 885

Urai, J. L., Humphreys, F. J., 1981. The development of shear zones in polycrystalline camphor. Tectonophysics, 78, 677–685. Urai, J. L., Humphreys, F. J., Burrows, S. E., 1980. In-situ studies of the deformation and

886

dynamic recrystallization of rhombohedral camphor. Journal of Materials Science 15,

887

1231–1240.

888

Urai, J. L., Means, W. D., Lister, G. S. 1986. Dynamic recrystallization of minerals. In:

889

Mineral and Rock Deformation: Laboratory Studies (edited by Hobbs, B. E. & Heard, H.

890

C.). Am. Geophys, Un. Geophys. Monogr. 36,161-199.

891

Vel, S.S., Cook, A., Johnson, S.E., Gerbi, C., 2016. Computational Homogenization and

892

Micromechanical Analysis of Textured Polycrystalline Materials, Computer Methods in

893

Applied Mechanics and Engineering 310, 749-779.

894 895 896

Vernon, R.H., 2004. A practical guide to rock microstructures. Cambridge University Press, Cambridge. Walte, N.P., Bons, P.D., Passchier, C.W. 2005. Deformation of melt-bearing systems -

897

insight from in situ grain-scale analogue experiments. Journal of Structural Geology 27,

898

1666-1679.

899 900 901 902

Wark, D.A., Watson, E.B., 2006. TitaniQ: a titanium-in-quartz geothermometer. Contributions to Mineralogy and Petrology 152, 743-754. Wheeler, J., 2009. The preservation of seismic anisotropy in the Earth’s mantle during diffusion creep. Geophysical Journal International 178, 1723-1732.

903

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

904

relation to grain rotation, grain boundary sliding and superplasticity. Philosophical

905

Magazine 90, 2841-2864.

906

Wheeler, J. 2014. Dramatic effects of stress on metamorphic reactions. Geology 42, 647-

907

650.Wheeler, J., 2018. The effects of stress on reactions in the Earth: sometimes rather

908

mean, usually normal, always important. Journal of Metamorphic Geology 36, 439-461.

909

Wilson, C. J. L. 1986. Deformation induced recrystallization of ice: the application of in situ

910

experiments. Mineral and rock deformation: laboratory studies: The Paterson volume,

911

213-232.

912

Wilson, C. J., Peternell, M., Piazolo, S., Luzin, V. 2014. Microstructure and fabric

913

development in ice: lessons learned from in situ experiments and implications for

914

understanding rock evolution. Journal of Structural Geology 61, 50-77.

915

916

Table Captions

917

Table 1 List of numerical models of microstructural (mm to dm) development identifying

918

processes modelled, numerical method used and providing relevant references. This list is

919

restricted to geological applications and those referenced in the main text. Note processes are

920

categorized as "S" for static (material points do not move) and "D" for dynamic (material

921

points move). Furthermore, unless stated otherwise models are two-dimensional.

922

Abbreviations: TBH – Taylor Bishop Hill calculation method for crystal lattice rotation,

923

VPFFT – Viscoplastic Fast Fourier Transform based model, EVPFFT - Elasto-viscoplastic

924

Fast Fourier Transform based model, FEM Finite Element, reXX – recrystallization.

925 926

Figure Captions

927

Figure 1 Microstructural development during in-situ deformation of the rock analogue

928

octochloropropane within a circular shear zone (Bons and Jessell, 1999); dashed red lines

929

indicate the shear distribution between the two steps shown. Experiments run with top to the

930

right shear at an average strain rate of 4.6·10-4 s-1 where the strain rate near edge of the shear

931

zone ( 1.2·10-3 s-1) is 10 x higher than in the top half of the image ( 1.2·10-4 s-1). (a) t1 at a

932

bulk shear strain of 40; (b) t1+16 min.. Note multiple concurrent processes: grain boundary

933

migration, leading to dissection of grains (locations 1 & 2), subgrain rotation (black arrow),

934

nucleation (white arrow). The different shear rates lead to a different balance of

935

recrystallization processes and differences in microstructures. At the low shear rate grains are

936

equant, have straight sub-grain boundaries and basal planes at an angle to the NS and EW

937

polarisers. The high-strain rate zone shows serrate grain boundaries, an oblique grain-shape

938

foliation, basal grains approximately parallel to the shear-zone boundary, as well as shear

939

localisation on grain boundaries (location 3), indicative of grain boundary sliding/shearing.

940

Such "micro-shear zones" may now have been detected in polar ice sheets as well (Weikusat

941

et al., 2009).

942

Figure 2 Numerical representation of a microstructure. (a) Micrograph of quartzite; (b)

943

numerical representation combining an element data structure with nodes (black circles),

944

segments (black lines) and polygons (enclosed area) and a lattice data structure with

945

unconnected lattice points (open circles). This structure is used in the numerical platform Elle

946

(see text for details).

947

Figure 3 Numerical modelling of dynamic recrystallization – a historical perspective;

948

mineral modelled is quartz; simple shear (see text for details); (a) numerical microstructure

949

after = 1; different grey scales signify different crystallographic orientations; top inset shows

950

data structure of hexagonal lattice points (modified after Jessell and Lister, 1990); (b)

951

numerical microstructure after = 1; colours show crystallographic orientation, grain

952

boundaries are red, subgrain boundaries black; for data structure see inset (modified after

953

Piazolo et al., 2002).

954

Figure 4 Numerical modelling of dynamic recrystallization – testing the effect of process

955

combination on microstructural development; model parameters: mineral modelled - ice; time

956

step - 20 years, simple shear; ∆ - 0.04 per time step. FFT and GBM signify fast fourier

957

transform formulation for crystal plasticity and grain boundary migration, respectively; (a)

958

initial microstructure and results after=2.4 for different process combinations. Results are

959

shown as grain network with orientation related colour coding according to crystal

960

orientations relative to the shortening direction y (see legend) and in pole figures. In the latter

961

the colour bar indicates the multiples of uniform distribution; (b) grain area distribution

962

normalized to the initial average grain area for all models shown.

963

Figure 5 Results from diffusion creep modelling in pure shear; 2D microstructure as in the

964

starting frame of Fig 1b of Wheeler (2009). Shown is the oblique view of microstructure with

965

patterns of normal stress shown along grain boundaries in the 3rd dimension as “fences”. Red

966

arrows show stretching direction, hence stresses are tensile (shown as negative) on

967

boundaries at a high angle to stretching direction are tensile. Fences are colour coded

968

according to dissolution rate with blue low and red high. When there is no relative grain

969

rotation the fences have a single colour and the stress is parabolic. When there is relative

970

grain rotation the fences vary in colour and the stress is a cubic function of position.

971

Figure 6 Dynamic development of stylolite roughness in numerical simulations; (a) time

972

series (left to right) with the stylolite nucleating in the middle of a slow dissolving layer

973

(layer in green and stylolite in black colour). Once the stylolite has dissolve the layer on one

974

side the layer starts to pin and teeth develop. (b) Variation of the pinning strength of the layer

975

in three different simulations showing a strong dependency. The compaction (movement of

976

upper and lower walls) is shown in quite arrows. L in the picture on the right hand side is the

977

initial position of the layer and P shown as black arrows indicates the pinning of the layer

978

upwards and downwards during dissolution.

979

Figure 7 Example of new development in microdynamic numerical modelling; preliminary

980

results (see text for details). (a) Coupling of diffusion creep and surface energy driven grain

981

boundary migration (GBM); (left) starting microstructure; (middle) microstructure at stretch

982

2; (right) graph showing average aspect ratio versus stretch; note that the microstructure after

983

a significant period of exclusive grain boundary migration is the same as the starting

984

microstructure as no movement occurs as all triple junctions are 120°. Number of grains

985

stays constant for all simulations. (b) Evolution of chemical concentration of an arbitrary

986

element during surface energy driven GBM. The material is a single-phase polycrystalline

987

aggregate with different initial chemical content. This example assumes very low bulk

988

diffusion (Dbulk =1e-20 m2/s) and fast grain boundary diffusion (Dboundary =1e-8 m2/s), hence

989

grain boundary diffusion dominates. Colour code indicates chemical concentration and white

990

lines represent grain boundaries. (c) Dissolution of reactive surfaces in 3D using the example

991

of fluorite dissolution. Microstructures after three dissolution periods are show (150hrs, 180

992

hrs, 276 hrs); surfaces correspond to the {111} plane at the start of experiment/simulation;

993

colours identify different depths where blue signifies low and red high; images are 60 μm

994

width. Top panel row shows numerical results. Lower panel shows experimental results using

995

confocal microscopy images of a grain of a sintered CaF2 pellet at the same three dissolution

996

times as the numerical models. Note the formation of etch pits with similar triangular shape

997

and the faster/enhanced dissolution of the grain boundaries in both experiment and numerical

998

simulations.

*Manuscript mark-up changes Click here to view linked References

1

A review of numerical modelling of the dynamics of microstructural development in

2

rocks and ice: Past, present and future

3 4

S. Piazolo1, P. D. Bons2, A. Griera3, M.-G. Llorens2, E. Gomez-Rivas4,5, D.

5

Koehn5Koehn6, J. Wheeler6Wheeler7, R. Gardner7Gardner8, J. R. A.

6

Godinho8Godinho9,, L. Evans7,9Evans8,10, R. A. Lebensohn10Lebensohn11, M. W.

7

Jessell11Jessell12

8

1

School of Earth and Environment, University of Leeds, Leeds LS2 9JT, United Kingdom

9

2

Department of Geosciences, Eberhard Karls University Tübingen, Wilhelmstr. 56, 72074

Formatted: Space Before: 0 pt, After: 0 pt Formatted: Superscript

10

Tübingen, Germany

11

3

12

(Cerdanyola del Vallès), Spain

13

44

14

080280 Barcelona, Spain

15

5

16

United Kingdom

17

56

18

Glasgow G12 8QQ, UK

19

67

20

Herdman Building, Liverpool University, Liverpool L69 3GP, U.K.

21

78

22

Systems/GEMOC, Department of Earth and Planetary Sciences, Macquarie University, NSW

23

2109, Australia

24

89

25

Manchester, M13 9PL, Manchester, UK.

Formatted: Space Before: 0 pt

Departament de Geologia, Universitat Autònoma de Barcelona, 08193 Bellaterra

Departament de Mineralogia, Petrologia i Geologia Aplicada, Universitat de Barcelona,

School of Geosciences, King’s College, University of Aberdeen, Aberdeen AB24 3UE,

School of Geographical and Earth Sciences, Gregory Building, University of Glasgow,

Dept. Earth, Ocean and Ecological Sciences, School of Environmental Sciences, Jane

Australian Research Council Centre of Excellence for Core to Crust Fluid

Henry Moseley X-ray Imaging Facility, School of Materials, The University of

Formatted: Space Before: 0 pt

26

910

27

Vic 3800, Australia

28

1011

Los Alamos National Laboratory, MST8, MS G755, Los Alamos, NM, 87545, USA

29

1112

Centre for Exploration Targeting, School of Earth Sciences, The University of Western

30

Australia, 35 Stirling Hwy, Crawley, 6009, Australia

School of Earth, Atmosphere and Environmental Sciences, Monash University, Clayton,

31 32

Corresponding author: Sandra Piazolo ([email protected])

33 34

Abstract

35

This review provides an overview of the emergence and current status of numerical

36

modelling of microstructures, a powerful tool for predicting the dynamic behaviour of rocks

37

and ice at the microscale with consequence for the evolution of these materials at a larger

38

scale. We emphasize the general philosophy behind such numerical models and their

39

application to important geological phenomena such as dynamic recrystallization and strain

40

localization. We focus in particular on the dynamics that emerge when multiple processes,

41

which may either be enhancing or competing with each other, are simultaneously active.

42

Here, the ability to track the evolving microstructure is a particular advantage of numerical

43

modelling. We highlight advances through time and provide glimpses into future

44

opportunities and challenges.

45 46

Keywords:

47

recrystallization, process interaction

review,

48 49

1.

Introduction

evolution

of

microstructures,

numerical

modelling,

dynamic

50

Structural geology studies how rocks deform under applied stress or strain. These studies are

51

applied from the sub-grain scale to that of mountain belts and tectonic plates. To know what

52

has happened to a volume of rock millions of years ago, or to predict what would happen to a

53

rock volume under certain (future) conditions, we need to know the link between (i) the

54

boundary conditions (e.g. applied stress or strain, pressure-temperature conditions), (ii) the

55

intrinsic material properties (e.g. Young's modulus, crystal symmetry, slip systems) and (iii)

56

the processes that may be activated (e.g. dislocation creep, pressure solution, mineral phase

57

changes). These three together determine the evolution of the state of the rock (e.g.

58

developing cleavage, folds, lineation, stylolites) and its bulk material properties (e.g.

59

viscosity). The microstructure is thus a state variable of a rock, describing the state it

60

achieved as a result of the interplay between various processes and boundary conditions.

61

Importantly, the microstructure is not a passive log of events, but plays an active and central

62

role in the evolution of a rock throughout its history (Gottstein, 2004). In this context non-

63

linear feedback between a specific microstructural configuration and local changes in

64

intrinsic and extrinsic parameters may be important but difficult to predict. Hence, the

65

microstructure is the link that couples the material properties, boundary conditions and

66

processes that together control the behaviour and evolution of a rock. Consequently, the

67

analysis and correct interpretation of microstructures is crucial in gaining an understanding of

68

how rocks, including ice, deform on Earth and other planets. Here we define microstructure

69

as the full spatial, compositional and orientational arrangement of all entities in a rock,

70

typically on the scale of a thin section to hand specimen (roughly µm to cm) (Hobbs et al.,

71

1976). These entities include minerals, grain and subgrain boundaries, crystal lattice

72

orientations, and chemical composition from the nano- to micro-scale.

73

While the rock's temperature or elastic strain are ephemeral, the microstructure may be

74

preserved for the geologist to interpret millions to billions of years later. Therefore,

Formatted: Space After: 0 pt

75

microstructures are one of the prime forensic tools to unravel the history of a rock that allows

76

us to deduce the succession of strain rates, stresses, diagenetic and metamorphic conditions

77

and the rock's rheology during deformation (e.g. Hobbs et al., 1976; Vernon, 2004; Passchier

78

and Trouw, 2005).

79 80

One major problem in microstructural studies is that we normally only have a static record as

81

post-mortem "images" (grain arrangement in thin section, chemical and orientation

82

characteristics derived using electron microscopy, 3D datasets from computer tomography,

83

neutron diffraction and synchrotron beam analysis, etc.). To interpret microstructures, we

84

need to know how they form and change. Means (1977) wrote: "a more valuable kind of

85

kinematic investigation is that in which the time sequence of incremental strains or

86

incremental displacements, the kinematic history, is correlated with progressive structural

87

change. This promises to reveal more about the physics of rock deformation and to provide

88

more sensitive structural methods for reconstructing tectonic history". Such investigations

89

first became possible using in-situ analogue experiments that were introduced to geology by

90

influential scientists such as Cloos (1955) and Ramberg (1981), and at the microstructural

91

scale by Means (1977). In in-situ analogue experiments by Means (1977), inspired by those

92

of McCrone and Chen (1949), a thin-section sized sample of crystalline analogue material is

93

sandwiched and deformed between glass plates to observe the changing microstructure under

94

a microscope. These experiments were followed by a large number of studies that

95

investigated deformation related features such as dynamic recrystallization and crystal

96

plasticity (e.g. Means, 1977; 1980; Means and Ree, 1988; Park and Means, 1996; Urai et al..,

97

1980; Urai and Humphreys, 1981; Wilson, 1986; Jessell, 1986; Ree, 1991; Bons and Jessell,

98

1999; Ree and Park, 1997; Herwegh and Handy, 1996, 1998; Nama et al., 1999; Wilson et

99

al., 2014) and), melt or fluid-bearing microstructures (e.g. Urai, 1983; Rosenberg and Handy,

100

2000, 2001; Rosenberg, 2001; Schenk and Urai, 2004, 2005; Walte et al., 2005) and), the

101

development of vein and fringe microstructures (e.g. Hilgers et al., 1997,; Koehn et al..,

102

2003) and the formation of -clasts (ten Brink et al., 1995). Many of these experiments made

103

clear that when a polycrystalline material deforms, multiple processes act upon the

104

microstructure and govern its dynamic behaviour throughout its deformation and post-

105

deformation history (Fig. 1a; see also movies at http://www.tectonique.net/MeansCD/). In

106

addition, pre-existing heterogeneities and/or those developing during deformation can have a

107

significant effect on the dynamics of the system since local variations in stress and strain play

108

a significant role in the evolution of microstructures and material behaviour. , with their

109

effects on chemistry being particularly intriguing (Wheeler, 2018).

110

Although changes in microstructural characteristics (e.g. grain size) can be quantified in such

111

experiments, the underlying principles of the active processes still need to be deduced.

112

Therefore, an additional tool is needed. Here numerical models simulating microstructural

113

development based on several concurrent processes become important. Such numerical

114

simulations share the advantage of in-situ experiments that the full microstructural

115

development can be traced in form of time-series, with the opportunity to systematically

116

study the effect of different processes and/or pre-existing heterogeneities on the

117

microstructural development. In addition, numerical simulations are neither constrained by

118

time nor specific boundary conditions. Once calibrated against laboratory experiments (e.g.

119

Piazolo et al.., 2004) or analytical solutions, they can be applied to a wide range of conditions

120

or materials including those not attainable or suitable for laboratory experiments. In many

121

cases, the developed microphysical behaviour may be applied to problems on the continental

122

scale.

123

In this contribution, we focus on those worksIn this contribution, we that (i) are applied to

124

geological materials, (ii) involve several processes and (iii) allow prediction and visualization

Formatted: Font: Times New Roman, 12 pt Formatted: Font: Times New Roman, 12 pt

125

of the development of the microstructure. Hence, we exclude studies where microstructures

126

are not spatially mapped along with models in which grains are modelled as if embedded in a

127

medium with averaged properties. We aim to give an overview of the concepts behind

128

numerical modelling of microstructures, i.e. microdynamic modelling, and the achievements

129

that have been made in this field. This is followed by a selection of examples of the current

130

state of the art. Finally, future possibilities and directions are briefly discussed, including

131

some work in progress. We focus on those studies that (i) are applied to geological materials,

132

(ii) involve several processes and (iii) allow prediction and visualization of the development

133

of the microstructure. Hence, we exclude studies where microstructures are not spatially

134

mapped along with models in which grains are modelled as if embedded in a medium with

135

averaged properties. Table 1 summarizes the microdynamic numerical models cited, the

136

numerical method used and processes included in each model.

137 138

2. Numerical simulation of microstructures

139

2. 1. Emergence and philosophy

140

Soon after the first in-situ experiments, computers had advanced to a stage allowing the first

141

numerical simulations of microstructural development, with a full "image" of the

142

microstructure, to appear in geology (Etchecopar, 1977; Jessell, 1988a,b; Jessell and Lister,

143

1990). When using such models we accept the fact that the scientific description of the

144

phenomena studied does not fully capture reality. This is not a shortcoming but a strength, as

145

general system behaviour and interrelationships can be studied systematically and

146

transferred to geological questions. In a numerical model the behaviour of a system emerges

147

from the piece-meal enumeration of the behaviour of each of its component parts. As a

148

consequence, numerical models are powerful and necessary when the bulk behaviour is

149

influenced by the local interaction of the components. In contrast, global, averaged or mean

Formatted: English (Australia)

150

field solutions cannot fully encompass the effect of processes that interact with each other

151

on the small scale. In nature, patterns emerge (e.g. orientation and arrangement of high-

152

strain zones, foliation, fractures) which may be used to deduce conditions of their formation.

153

Numerical models are powerful tools to investigate this link between patterns observed in

154

nature and processes that are responsible for their development.

155 156

2.2. Multiprocess modelling – approaches

157

As shown convincingly by the aforementioned experiments, several processes act

158

concurrently on a microstructure and control the dynamics of the microstructural evolution

159

and material properties (e.g. Means, 1977, 1980). While there are a large number of models

160

that focus on one process alone, there are fewer systems that aim to model the effect of

161

multiple interacting processes. Here, we focus on the latter. Solving all necessary equations

162

simultaneously is only practical when the number of interacting processes is small.

163

Alternatively, the set of equations, hence processes, needs to be reduced to a manageable

164

number necessitating a user driven selection which may result in the omission of important

165

feedbacks. Operator splitting provides an alternative approach, whereby each process acts

166

sequentially on a microstructure that was incrementally changed by all operating processes in

167

the previous step. In this case, the number of equations to be solved is not limited; hence

168

there is no need to artificially reduce the number of modelled processes. Clearly, length scale

169

and time steps have to be considered carefully for this approach to be valid. The operator

170

splitting approach is fundamental to the numerical platform Elle (Jessell et al., 2001, Bons et

171

al., 2008, Piazolo et al., 2010; www.elle.ws) which is, within the realm of microdynamic

172

modelling, the most used numerical platform in geology. However, operator splitting has also

173

been employed in other numerical schemes (e.g. Cross et al., 2015).

174

175

2.3. The numerical representation of a microstructure

176

It is necessary to describe numerically the microstructure so that in the model the processes

177

can act upon it spatially. Up to now, for rocks and ice, microstructural models have been

178

mostly restricted to two-dimensions. Two basic approaches can be taken: (i) a lattice data

179

structure or (ii) an element data structure. In the lattice data structure, the microstructure is

180

mapped onto an irregular or regular lattice, like the pixels in a digital image or gridded

181

analysis points obtained during chemical mapping and orientation mapping of geological

182

samples. Potts, cellular automata, lattice-spring, particle-in-cell, phase field and

183

micromechanical Fast Fourier Transform (FFT) based models utilize the lattice data structure

184

(Fig. 2; see also chapter 2 of Bons et al., 2008 for a detailed review of different numerical

185

methods). Numerically the lattice data structure is easy to use and calculations are

186

straightforward. In the element data structure, the microstructure is described by discrete

187

elements which can be points, line segments, or polygons (Fig. 2). Finite-element models and

188

boundary models typically use elements to describe a system. Element data structure is well

189

suited for processes acting upon a surface, e.g. grain boundary migration, but is numerically

190

more “expensive” due to necessary topology checks. Both elements and lattice points can

191

have, in addition to their location in xy space, specific properties or store specific state

192

variables, such as mineral phase, chemical composition, crystallographic orientation, and

193

stress and strain (rate) values, respectively. The numerical platform Elle (Jessell et al.., 2001)

194

combines these two approaches: a microstructure is represented by boundary nodes

195

connected by straight segments (Fig.2). A polygon enclosed by boundary nodes represents a

196

grain or subgrain. Property variations within the individual polygons are defined at interior

197

points where information is recorded and tracked. Processes acting on the microstructure can

198

move boundaries and/or interior points and/or change the properties at those points, e.g.

199

composition, crystallographic orientation or dislocation density. The system has now matured

200

to a stage that it can robustly be used for a large variety of models relevant to geology.

201 202

3. Examples of modelling several coupled processes

Formatted: Space After: 0 pt

203

3.1.

Formatted: Indent: Left: 0", Hanging: 0.39", Space After: 0 pt

204

3.1.1. A historical perspective

205

Dynamic recrystallization is the response of a crystalline aggregate to lower its free energy by

206

formation and movement of sub- and grain boundaries (Means, 1983). It occurs by a number

207

of concurrent processes that act upon the polycrystalline material. For decades, geologists

208

have used microstructural characteristics developed during dynamic recrystallization to infer

209

dominance of processes and linked boundary conditions (e.g. Urai et al., 1986; Hirth and

210

Tullis, 1992; Stipp et al., 2002). Following the first attempts of such a model (Etchecopar,

211

1977), Jessell (1988a,b) and Jessell and Lister (1990) developed a numerical model

212

incorporating rotation of the crystal lattice and dynamic recrystallization for quartz. The data

213

structure used was a 100x100 hexagonal lattice structure where each lattice point represented

214

a grain or subgrain (Fig. 3a inset). Calculation of the crystal lattice rotation in Jessell and

215

Lister (1990) uses the Taylor-Bishop-Hill calculation method (Taylor, 1938; Bishop and Hill,

216

1951a, 1951b; Lister and Paterson, 1979) and the critical resolved shear stress of the different

217

slip systems in quartz. This work reproduced the general microstructures seen in a mylonite.

218

A major step forward was that the simulations showed that dynamic recrystallization can

219

significantly modify crystallographic preferred orientation (CPO) development. Piazolo et al.

220

(2002) advanced Jessell’s model by utilizing the more robust Elle data structure to describe

221

the local deformation and the dynamics of the grain boundary network (Jessell et al., 2001).

222

The model now included a finite element solution for incompressible, linear or nonlinear

223

viscous flow (BASIL; Barr and Houseman 1992, 1996) to compute the local velocity field,

Dynamic recrystallization

Formatted: Space After: 0 pt

224

the front tracking method for the motion of grain boundaries by moving nodes and segments

225

(e.g. Bons and Urai, 1992), the Taylor-Bishop-Hill formulation (Lister and Paterson, 1979)

226

and additional features such as crystal lattice rotation, formation of subgrains,

227

recrystallization by nucleation, grain boundary migration, recovery, work hardening and

228

tracking of dislocation densities. This list of additional features shows the advance made in

229

approximately a decade in reproducing natural microstructures (Fig. 3b).

230

So far, the rheological anisotropy of minerals deforming by dislocation creep was not taken

231

into account in the calculation of the stress-strain rate field and lattice rotation. About another

232

decade after Piazolo et al. (2002), the viscoplastic implementation of the full-field crystal

233

plasticity micromechanical code based on Fast Fourier Transforms (VPFFT; Lebensohn,

234

2001) was introduced in geological microstructural modelling (Griera et al., 2011, 2013,

235

2015). The VPFFT is a spectral method that specifically assumes that deformation is

236

achieved by dislocation glide on crystallographic slip planes (each with their own critical

237

resolved shear stress) whose orientations are mapped on a regular grid (for details see Griera

238

et al. (2013), Montagnat et al. (2011, 2014) and Llorens et al. (2016a)). The simulations of

239

Griera et al. (2011, 2013) and Ran et al. (2018) illustrate that heterogeneous stress within and

240

between grains that emerge from inhomogeneous slip has a significant effect on the rotation

241

rate of porphyroblastporphyroclasts and -clastblasts and the strain (rate) field around these

242

objects.

243

More recently, Llorens et al. (2016a; 2016b; 2017) and Steinbach et al. (2016; 2017) coupled

244

dynamic recrystallization (DRX) by grain boundary migration (GBM), intracrystalline

245

recovery and polygonisation with VPFFT crystal plasticity models, and applied them to study

246

the deformation of polar ice. Llorens et al. (2016a) presents a comprehensive description of

247

the method. They found that dynamic recrystallizationDRX produces large and

248

equidimensional grains, but only marginally affects the development of the c-axes CPO.

Formatted: Space Before: 0 pt, After: 0 pt

249

However, DRX can alter the activity of slip systems and does modify the distribution of a-

250

axes. In simple shear, the strong intrinsic anisotropic of ice crystals is transferred from the

251

crystal to the polycrystal scale, leading to strain localisation bands that can be masked by

252

GBM (Llorens et al., 2016b). Llorens et al. (2017) compared the dynamics of pure polar ice

253

polycrystalline aggregates in pure and simple shear deformation. It was found that, due to the

254

vorticity of deformation, it is expected that ice is effectively weaker in the lower parts of the

255

ice sheets (where simple shear dominates) than in the upper parts (where ice is mostly

256

deformed coaxially). In Jansen et al. (2016) thisThe method was also applied to reproduce the

257

development of tilted-lattice (kink) bands found in ice cores (Jansen et al., 2016) and

258

crenulation cleavage during folding (Ran et al., 2018) as a result of mechanical anisotropy.

259

Steinbach et al. (2016) included air inclusions as a second phase to simulate DRX in porous

260

firn and found that DRX can occur despite the low strain and stress in firn.

261

Apart from polar ice, the VPFFT-Elle model has also been used to analyse subgrain rotation

262

recrystallization of halite polycrystals in simple shear (Gomez-Rivas et al., 2017) by coupling

263

VPFFT with the Elle routines that simulate intracrystalline recovery (Borthwick et al, 2013)

264

and subgrain rotation. They found that recovery does not affect CPOs, but strongly decreases

265

grain size reduction. These authors also evaluated the use of mean subgrain misorientations

266

as a strain gauge.

267

3.1.2. Testing the effect of different combinations of recrystallization processes

268

We show results from three combinations of processes that occur during dynamic

269

recrystallization. These processes act on the same initial 10x10 cm aggregate of ice Ih

270

crystals (Fig. 4a). The numerical approach in these simulations is based on the VPFFT

271

micromechanical model in combination with several Elle processes (Llorens et al., 2016a,

272

2016b, 2017; Steinbach et al., 2016; 2017). The deformation-induced lattice rotation and the

273

estimation of geometrically necessary dislocation densities calculated from the stress and

274

velocity field provided by the VPFFT algorithm are used to simulate recrystallization by

275

intra-crystalline recovery, grain boundary migration and nucleation. Grain boundary

276

migration (GBM) is simulated using a front-tracking approach based on the algorithm by

277

Becker et al. (2008). Here, a front-tracking approach is used in which the movement of

278

boundaries is mapped through time. Recovery reduces the intra-granular stored energy in a

279

deformed crystal (Borthwick et al., 2013). Nucleation creates new grain boundaries with

280

areas of misorientation values above a pre-defined threshold (Piazolo et al.., 2002; Llorens et

281

al., 2017; Steinbach et al., 2017). Dextral simple shear deformation is applied in strain

282

increments of ∆ = 0.04. Three different combinations of recrystallization processes are

283

tested; namely: (1) GBM only, (2) GBM and recovery or (3) GBM, recovery and nucleation

284

(Fig. 4a). Simulations including VPFFT viscoplastic deformation show a similar evolution of

285

c-axis orientation, regardless of the dynamic recrystallization processes included (Fig. 4a).

286

The lack of influence of nucleation on the CPO is due to the fact that new grains are modelled

287

to have lattice orientations close to those of their parent grains. However, inclusion of the

288

nucleation process result in grain size reduction (Fig. 4b). Results illustrate how different

289

combinations of microdynamic processes affect the microstructural characteristics (e.g. CPO,

290

grain size, shape and orientation) in different ways where no single microstructural parameter

291

can grasp the full dynamics of polycrystalline deformation.

292

3.2. Diffusion creep

Formatted: Indent: Left: 0", Hanging: 0.3", Space After: 0 pt

293

Diffusion creep is a physiophysico-chemical process that operates in large regions of the

Formatted: Space After: 0 pt

294

Earth – in the upper Earth as “pressure solution” where diffusion is enhanced along fluid

295

films and elsewhere at low strain rates, maybe in large parts of the lower mantle (Karato,

296

1992). The mechanism involves diffusion driven by gradients in normal stress along

297

boundaries. In simple models no other driving forces need to be included. It is implicit in our

298

understanding of this mechanism that grain boundary sliding occurs in conjunction with

299

diffusion. Ford et al. (2002) devised a numerical model to understand the evolution of grain

300

shapes and CPO in grain boundary diffusion creep. The mathematical framework allows for

301

the consideration of diffusion and sliding together. Operator splitting is not required and the

302

evolution at each timestep is based on a single matrix inversion operation. The model couples

303

the microstructure to an evolving system of local stresses (Fig. 5). The model predicts that

304

grain shapes become somewhat elongate, in accordance with experiments on calcite (Schmid

305

et al., 1987) and observations on an albite-bearing mylonite (Jiang et al., 2000) (Fig. 5). It

306

also predicts that grain rotations are not chaotic and that CPO may be present to high strains

307

(Wheeler, 2009), as seen in, for example, experiments on olivine-orthopyroxene (Sundberg

308

and Cooper, 2008).) and predicted by numerical simulations of Bons and den Brok (1999). A

309

second phase, if insoluble, can be included and such a model was used by Berton et al. (2006,

310

2011) to prove the hypothesis that a different grain boundary diffusion coefficient along the

311

two-phase boundaries could explain fibre growth in pressure shadows. Not all model

312

predictions are in agreement with other investigations – for example some materials

313

deformed by diffusion creep e.g. olivine show quite equant grains (Karato et al., 1986). In

314

such experiments grain growth occurs but this process was not included in the numerical

315

model, hence results differ. In a later section we address this issue, but this example

316

illustrates that, in common with other numerical models, the applicability reflects the

317

processes included.

318 319

3.3. Stress driven dissolution, growth and dynamic roughening

Formatted: Indent: Left: 0", Hanging: 0.3", Space After: 0 pt

320

Another example where stress leads to micro- and macrostructures is the roughening of grain

Formatted: Space After: 0 pt

321

boundaries in the presence of a fluid. In this case dissolution and/or precipitationmineral

322

growth are driven not just by normal stress gradients at the interface (c.f. section 3.2) but also

323

by changes in elastic and surface energies as well as normal stress at the interface.. In

324

thisaddition to these driving forces, the detailed electrochemistry of interfaces has a kinetic

325

effect on rates (Gratier et al., 2013). In the example presented here, the process is modelled in

326

Elle by coupling a lattice spring code that calculates the strain with a background fluid that

327

dissolves lattice elements. This process can produce transient patterns of interface geometry

328

at the grain boundaries (Koehn et. al., 2006). In addition, the process itself produces rough

329

interfaces that can be seen on the larger scale as stylolites (Koehn et al., 2007; Ebner et al.,

330

2009). The scaling behaviour of stylolites is important for stress inversion and compaction on

331

the basin scale (Koehn et al., 2012; 2016) and the shape of stylolites is partly rooted in the

332

microstructure of the host-rock (Fig. 6). Slower dissolving material on various scales, from

333

small grains to fossils and layers, can influence the dissolution process and thus the

334

development of patterns, whereas deformation leads to changes in long-range interactions

335

(Koehn et al., 2016). SimilarSimilarly, pinning behaviour in the presence of a fluid is seen in

336

grain growth where grain boundary pinning results in restriction of grain growth. Pinning

337

particles can also be moved around so that growing grains become clean. Such a process can

338

lead to layered rocks, for example, the Zebrazebra dolomites (Kelka et al., 2015).

339

4. Numerical simulations of microdynamic processes: Examples of ongoing work

340

4.1.

341

Even now, most of the numerical models that incorporated more than one process are

342

restricted to deformation processes alone. Although in diffusion creep there is an intrinsic

343

chemical aspect, the model described in 3.2 involves just a single phase of fixed composition.

344

There are some models that incorporate the growth of new mineral phases and investigate the

345

rheological effect of such growth (e.g. Groome et al., 2006; Smith et al., 2015), however the

346

location of new phase growth and/or growth rate is predefined. While Park et al. (2004)

347

showed convincingly the effect of microstructure on the diffusion pathways and chemical

348

patterns in garnet and biotite, there are now many opportunities available in a numerical

Linking chemistry and microstructural evolution

Formatted: Indent: Left: 0", First line: 0", Space After: 0 pt Formatted: Space After: 0 pt

Formatted: Font: Not Italic, No underline

349

system such as Elle (Jessell et al., 2001) to couple local chemistry and microstructural

350

evolution. Below we present preliminary results from two current projects linking chemical

351

changes and a dynamically evolving grain network.

352

4.1.1. Grain boundary diffusion creep and grain growth

353

Here we present the first results using a model combining, using operator splitting,

354

deformation by diffusion creep (Ford et al., 2002,; Wheeler, 2009) and surface energy driven

355

grain boundary migration (GBM), i.e. grain growth (modified after Becker et al., 2008). Four

356

different scenarios are shown: diffusion creep only, grain boundary migration only and two

357

combinations with medium and high grain boundary migration GBM rates (Fig. 7a). The

358

starting material is an almost regular hexagonal grain mesh with equant grains in which triple

359

junctions have had small random perturbations imposed (Fig. 7a). The perturbations are

360

required to avoid mathematical problems that arise in perfectly regular hexagonal networks

361

(Wheeler, 2010). Topology checks such as neighbour switching routines are performed after

362

each step in all simulations, ensuring that no topology problems arise. The microstructural

363

development is markedly different if grain growth is coupled with diffusion creep (Fig. 7a)

364

where with increasing GBM rate the aspect ratios are generally reduced relative to that of

365

comparable experiments modelling diffusion creep only. This is provisionally in accord with

366

the hypothesis that there will be a ‘steady state’ grain elongation developed which is a

367

function of the relative magnitudes of strain rate and grain growth kinetic parameters

368

(Wheeler, 2009) but more simulations are required (with less regular starting

369

microstructures). In contrast, for grain growth only, the microstructure does not change as the

370

initial configuration consists of stable near 120° triple junctions (Fig. 7a). Grain numbers (i.e.

371

grain size) do not change in any of the simulations.

372 373

4.1.2. Trace element partitioning between fluid and solid coupled with recrystallization

374

Field and experimental studies illustrate the importance of the time and length scales on the

375

evolution of grain networks during trace element diffusion (e.g. Ashley et al., 2014 and

376

references therein). Therefore, understanding the influence of recrystallization on trace

377

element distribution is essential for correctly interpreting patterns of trace element

378

distribution (e.g. Wark and Watson, 2006). However, to our knowledge there are currently no

379

numerical approaches able to simulate diffusion coupled with microstructure evolution.

380

Recently, a new Elle finite difference process that solves Fick’s second law to simulate trace

381

element diffusion in a polycrystalline medium has been developed. The polycrystal is defined

382

by n-phases where grain boundaries and grains are differentiated (Fig. 7b). The trace element

383

partitioning coefficient between different phases is also considered. The approach uses an

384

element data structure representing grain boundaries and a lattice data structure of a regular

385

grid of unconnected lattice point to track diffusion along the grain boundaries and the

386

physical and chemical properties within grains, respectively. To ensure elemental exchange

387

between grain boundaries and grain interiors, a grain interior-grain boundary partition

388

coefficient is implemented that takes the proximity to grain boundaries into account. This

389

allows the simulation of coupled bulk and grain boundary diffusion. This approach can be

390

fully coupled with other processes of the numerical platform Elle and therefore allows

391

simulating simultaneous diffusion, deformation and static or dynamic recrystallization. The

392

use of this approach is demonstrated through the modelling of diffusion of a tracer with

393

fractionation during static grain growth of a single-phase aggregate (Fig. 7b) (after that

394

presented by Jessell et al., 2001). Patterns of chemical concentration qualitatively resemble

395

those of Ti-distribution observed in recrystallized quartz in shear zones and can help to

396

understand the redistribution of Ti in quartz during dynamic recrystallization (e.g. Grujic et

397

al. 2011).., 2011), among many other geological problems.

398

4.2.

399

We restricted this review to two -dimensional numerical approaches, as very few models

400

have been published that are in three dimensions and also specific to geological

401

microstructures. An exception is the phase field approach by Ankit et al. (2015) who

402

investigate the growth of crystals in an open fracture. However, within the material science

403

community numerical models exist that investigate the behaviour of a microstructure in

404

response to single processes in three dimensions (e.g. three dimensional crystal plasticity

405

modelling platform DAMASK (Roters et al., 2012; Eisenlohr et al., 2013), which includes the

406

option of solving the micromechanical problem using a 3D implementation of the FFT-based

407

formulation (presently implemented in Elle in its 2-D version) and grain growth (e.g. Krill

408

and Chen, 2002; Kim et al., 2006; Zöllner, 2011). With the technical advances made within

409

the geological and material science community it is now in principle possible to extend

410

currently available 3D models to geological questions and/or multi-process scenarios.

Modelling in three dimensions

Formatted: Indent: Left: 0", First line: 0", Space After: 0 pt

Formatted: Space After: 0 pt

411 412

4.2.1. Dissolution of reactive surfaces in three dimensions

413

Mineral dissolution controls important processes in geoscience such as serpentinization (e.g.

414

Seyfried et al., 2007), retrogression and replacement reactions (Putnis and Austrheim, 2012),

415

deformation by dissolution-precipitation creep (e.g. Rutter, 1976) and also affects processes

416

relevant to society such as the stability of spent nuclear fuel and mine waste. Recent work has

417

shown that dissolution rates are linked to the evolving surface structure, and thus are time-

418

dependent (Godinho et al., 2012). Different surfaces have different structures in three

419

dimensions and the presence of etch pits play a major role in dissolution (Pluemper et al..,

420

2012). Hence, for accurate representation of dissolution behaviour, modelling in three

421

dimensions is important. Here, we present a numerical model that simulates the dissolution

422

process as a potential tool to quantify the links between dissolution rates, reactive surface

423

area and topography over periods of time beyond reasonable for a laboratory experiment. The

424

program uses empirical equations that relate the dissolution rate of a point of the surface with

425

its crystallographic orientation (Godinho et al., 2012) to simulate changes of topography

426

during dissolution, which ultimately results in the variation of the overall dissolution rate

427

(Godinho et al., 2012, Godinho et al., 2014). The initial surface is composed of a group of

428

nodes with a xy position and a set height (z). For each lattice point a local surface orientation

429

is calculated from the inclination of the segment node and its neighbours. This orientation

430

together with the crystallographic orientation of the grain the surface belongs to is then used

431

to calculate a dissolution rate (surface reactivity). Based on this the displacement of the node

432

using the equations published in Godinho et al. (2012) is calculated. The model allows the

433

graphical display of the three-dimensional topographic development of the surface, tracking

434

of the variation of the surface area and calculation of the overall dissolution rate at each stage

435

of the simulation. Results obtained with the 3D simulation at three consecutive stages of

436

dissolution (150 hrs, 180 hrs, 276 hrs dissolution duration) show that numerical results are in

437

accordance to experimental results (Fig. 7c).

438 439

5. Numerical simulations of microstructures: Possibilities and challenges

440

5.1. Linking laboratory and natural data with numerical models

441

Over the last two decades numerical capabilities have advanced markedly and models have

442

come of age. Consequently, a large range of models and process combinations is now

443

available that can be utilized to gain further insight into the link between processes, material

444

properties and boundary conditions. This also includes polymineralic systems making the

445

models more appropriate to use to investigate processes occurring within

446

polymineralicpolyphase rocks (e.g. Roessiger et al., 2014; Steinbach et al, 2016). New

447

avenues of effective verification against laboratory and natural data are now opening up, due

448

to the development of analytical tools allowing rapid complete microstructural and

449

microchemical analysis providing dataset similar to those possible in numerical models; (e.g.

450

Steinbach et al, 2017, Piazolo et al., 2016); many of the analytical data sets have a lattice data

451

structure.

452 453

5.2. Linking chemistry and microstructural evolution

454

Advances in numerical methods, theoretical treatment of thermodynamic data, analytical

455

tools and theoretical and experimental insights into the coupling between chemical and

456

physical process that emerged over the last decade, call for a major effort in this area of

457

research. Examples of areas of opportunities include investigation of the (1) significance of

458

local stress versus bulk stress on mineral reactions and reaction rates (e.g. Wheeler, 2014;

459

Hobbs and Ord, 2014), (2) characteristics of replacement microstructures and their potential

460

significance for microstructural interpretation (e.g. Putnis, 2009, SpuzienceSpruzeniece et al.,

461

2017), (3) mobility of trace elements enhanced by deformation that change significantly the

462

local elemental distributions (e.g. Reddy et al., 2007, Piazolo et al., 2012, 2016) and (4)

463

coupling between reactive fluid-solid systems and hydrodynamics (Kelka et al., 2017).

464

Increasing computer processing speeds will aid the running of such models. However, the

465

technical challenges include the harmonisation of fundamentally different numerical

466

approaches (crystal plasticity versus diffusion creep, for example) and the resolution of

467

fundamental mathematical problems, e.g. the current lack of an internally consistent model

468

for multiphase diffusion creep, highlighted by Ford and Wheeler (2004).

469

5.3. Linking brittle and ductile deformation: elasto-viscoplastic behaviour

470

When examining the rock record, it is clear that in many cases, brittle and ductile behaviour

471

often occurs at the same time within a rock (e.g. Bell and Etheridge, 1973, Hobbs et al.,

472

1976). With the potential significance of grain-scale brittle behaviour now measurable on

Formatted: Space After: 0 pt

Formatted: No underline, Font color: Auto, English (Australia)

473

seismic signals (e.g. Fagereng et al., 2014) there is an increased need in developing numerical

474

techniques that allow us to model the dynamic link between brittle and ductile behaviour.

475

Such elasto-viscoplastic behaviour combines the elastic reversible fast deformation with a

476

viscous time dependent flow and a plastic behaviour. These behaviours can be included in

477

continuous or discontinuous models. For example, in the numerical platform Elle, a lattice

478

structure is used to deform the model elastically up to a critical stress where bonds fracture

479

(plastic behaviour) and the particles themselves deform as a function of stress and time

480

(viscous behaviour). In this case the viscous behaviour conserves the volume whereas shear

481

forces and differential stresses converge to zero (Sachau and Koehn, 2010, 2012; Arslan et

482

al., 2012; Koehn and Sachau, 2014). Alternatively, linking the elasto-visocplastic FFT based

483

(EVPFFT) (Lebensohn et al., 2012) with models such as Elle would allow calculation of the

484

Cauchy stresses that are the local driving force for grain scale damage processes.

485

One of the major challenges is the large range of time-scales in these processes, with

486

fracturing and fluid flow on fast to intermediate and viscous deformation and potentially

487

reactions on very long time scales. This complexity either requires the assumption that some

488

processes are instantaneous or it requires an “up-scaling in time” or non-linear time scales in

489

models.

490 491

5.4. Expansion of capability to three dimensions

492

With the advent of supercomputers and new numerical approaches now is the time to develop

493

techniques to investigate microstructural development in three dimensions. This is of

494

particular importance if material transport such as aqueous fluid and/or melt flow is to be

495

considered. This would enable, for example, modelling of strain fringes, shear veins and en-

496

echelon extensiontension gashes. However, these require models that link brittle and ductile

497

behaviour along with modelling in three dimensions.

498

One of the biggest problems faced with three-dimensional models of microstructures are the

499

three-dimensional topology changes that are common in dynamic microstructural

500

development (e.g. Fig. 1-7). This is a major problem if an element data structure with

501

segments, i.e. grain boundaries is used. However, the phase field approach does not work

502

with such distinct boundaries, and is therefore it is well suited for 3D problems (e.g. Ankit et

503

al. 2014). In addition, a three dimensional network of unconnected nodes, in which there is

504

no physical movement of boundaries but only changes in the properties of the 3D nodes

505

(voxels) (cf. Fig. 7c) may be a way forward (e.g. Sachau and Koehn, 2012).

506 507

5.5. Link between geophysical signals and microstructure

508

At a time where there is an ever-increasing amount of geophysical data being collected,

509

numerical simulations that are used to test the link between microstructural development and

510

geophysical signal will become increasingly important. CrypchCyprych et al. (2017) showed

511

that not only crystallographic preferred orientation but also the spatial distribution of phases

512

i.e. the microstructure, has a major impact on seismic anisotropy. Therefore, a direct link

513

between the microdynamic models such as Elle allowing tracking of microstructural changes

514

through time and space and geophysical signal generation offers a wealth of new

515

opportunities including interpretation of strong reflectors in the lower crust and mantle.

516

Current efforts by Johnson, Gerbi and co-workers (Cook et al., 2013; Naus-Thijsen et al.,

517

2011; Cook et al., 2013; Vel et al., 2016) are in line with this direction.

518 519

5.6. Application of numerical models to polymineralic rock deformation and ice-related

520

questions

521

The dynamic behaviour of the Earth is strongly influenced by the deformation of

522

polymineralic rocks. At the same time Earth’s polar ice caps and glacial ice which often

523

include ice and a second phase (e.g. air inclusions, dust, entrained bedrock) is of major

524

importance to society, especially in view of changing climate (e.g. Petit et al., 1999; EPICA,

525

2004). Application of the current numerical capabilities to polycrystalline ice and

526

polymineralic rocks is therefore urgently needed. Over the last years, there has been an

527

increased effort in this direction (Roessiger et al., 2011, 2014; Piazolo et al., 2015; Llorens et

528

al., 2016a, 2016b, 2017; Jansen et al., 2016; Steinbach et al. 2016, 2017; Roessiger et al.,

529

2011, 2014), which promises to continue. Here, the development of the link between

530

elemental mobility and microstructural development is of major importance, as only with

531

such models the can chemical signals of, for example, ice cores be correctly interpreted.

532 533

5.7. Upscaling: Utilizing operator splitting and utilities developed for microdynamic

534

systems to larger -scale problems

535

One of the strengths of the numerical approach taken by the microstructural community has

536

been the close link between different processes and the ability of the models to take into

537

account the local differences in properties such as stress, strain and chemistry. The technique

538

of operator splitting has proven extremely powerful. Furthermore, the ability to model

539

anisotropic material behaviour utilizing for example VPFFT viscoplastic deformation

540

formulations (Lebensohn, 2001), has enabled realistic and dynamic models. Upscaling this

541

approach to investigate problems at a large scale e.g. folding (Llorens et al., 2013a, 2013b;

542

Bons et al.., 2016; Ran et al., 2018) and shear deformation (Gardner et al., 2017) have shown

543

to be very beneficial. There is great scope to expand further on this in view of large scale

544

shear deformation, fluid flow, mineralization and fault formation.

545 546

6. Numerical simulations of microstructures: Lessons learnt and future challenges

Formatted: Space After: 0 pt

547

Numerical simulations of microstructural development have caught our imagination over the

548

last three decades. They have markedly advanced our ability to explain phenomena and

549

patterns we observe in nature and experiments by allowing us to test the link between

550

boundary conditions, material properties, processes, and microstructural development.

551

Importantly, models, especially those that couple several process and/or investigate pre-

552

existing heterogeneities can train the geologist to think of the dynamics of the system rather

553

than a linear development. For example, different patterns of strain localization observed in

554

nature can be explained by differences in the relative rates of interacting processes (e.g.

555

Jessell et al.., 2005; Gardner et al.., 2017). At the same time, specific indicative

556

microstructural parameters can be developed to help interpret natural microstructures (e.g.

557

Piazolo et al., 2002; Gomez-Rivas et al., 2017, Llorens et al., 2017; Steinbach et al., 2017).

558

Such

559

However, including chemistry coupled to other processes remains a particular challenge.

560

Whilst, for example, trace-element diffusion can be enacted in parallel with other processes

561

(section 4.1.2), chemical transport of major elements and diffusion creep cannot yet be fully

562

integrated with many other processes. Indeed, when multiphase systems are considered, there

563

are unsolved problems with diffusion creep modelling even in the absence of other processes

564

(Ford and Wheeler, 2004). This challenge is closely linked to our current inability to

565

confidently model grain boundary sliding.

566

Nevertheless, the studies we describe have shown that numerical models are extremely

Formatted: No underline, Font color: Auto, English (Australia)

567

powerful in providing benchmark results to investigate what kind of microstructure may

Formatted: No underline, Font color: Auto, English (Australia)

568

develop under certain conditions. These models are sophisticated mind experiments that are

569

firmly based on physical and chemical laws for which the theory is well known individually

570

but their interaction is difficult to predict analytically.

571

572

Acknowledgements:

573

We would like to thank Win Means for opening up a whole new perspective on

574

microstructures with his inspirational in-situ experiments. The authors thank the DFG, ARC,

575

ESF, NSF, EU through Marie Curie Fellowship to SP and NERC for support of the numerical

576

endeavours., the Government of Catalonia's Secretariat for Universities and Research for a

577

Beatriu de Pinós fellowship to EGR (2016 BP 00208), and NERC (NERC grant

578

NE/M000060/1) for support of the numerical endeavours. The authors thank C. Gerbi for his

579

helpful and constructive review as well as C. Passchier for editorial handling of the

580

manuscript. Formatted: Font: Bold

581 582 583

References

584

Ankit, K., Urai, J. L., Nestler, B. 2015. Microstructural evolution in bitaxial crack‐ seal

585

veins: A phase‐ -field study. Journal of Geophysical Research 120, 3096-3118.

586

Arslan, A., Koehn, D., Passchier, C.W., Sachau, T. 2012. The transition from single layer to

587

foliation boudinage: A dynamic modelling approach. Journal of Structural Geology 42,

588

118-126.

589

Ashley, K.T., Carlson, W.D., Law, R.D., Tracy, R.J. 2014. Ti resetting in quartz during

590

dynamic recrystallization: mechanisms and significance. American Mineralogist 99, 2025-

591

2030.

592 593 594 595

Barr, T.D., Houseman, G.A. 1992. Distribution of deformation around a fault in a non-linear ductile medium. Geophys Res Lett 19, 1145-1148. Barr, T.D., Houseman, G.A. 1996. Deformation fields around a fault embedded in a nonlinear ductile medium. Geophys J Int 125, 473-490.

596

Becker, J.K., Bons, P.D., Jessell, M.W. 2008. A new front-tracking method to model

597

anisotropic grain and phase boundary motion in rocks. Computers & Geosciences 34, 201-

598

212.

599 600 601 602 603

Bell, T. H., Etheridge, M. A. 1973. Microstructure of mylonites and their descriptive terminology. Lithos 6, 337-348. Berton, J. R., Durney, D. W., Wheeler, J., Ford, J. M., 2006. Diffusion-creep modelling of fibrous pressure-shadows. Tectonophysics 425, 191-205. Berton, J. R., Durney, D. W., Wheeler, J., 2011. Diffusion-creep modelling of fibrous

604

pressure-shadows II: influence of inclusion size and interface roughness. In: Deformation

605

Mechanisms, Rheology & Tectonics: Microstructures, Mechanics & Anisotropy (eds

606

Prior, D. J., Rutter, E. H. & Tatham, D.), 319-328, The Geological Society Special

607

Publications, 360, .319-328, .

608 609 610 611 612 613 614 615 616 617 618 619

Bishop J.F.W., Hill R. 1951a. A theory of the plastic distortion of a polycrystalline aggregate under combined stresses. Philosophical Magazine 42, 414–427. Bishop, J.F.W., Hill, R. 1951b A Theoretical Derivation of the Plastic Properties of a Polycrystalline Face-Centred Metal. Philosophical Magazine 42, 1298-1307. Bons, P.D., den Brok, S.W.J. 2000. Crystallographic preferred orientation development by dissolution-precipitation creep. Journal of Structural Geology 22, 1713-1722. Bons, P.D., Jessell, M.W. 1999. Micro-shear zones in experimentally deformed octachloropropane. Journal of Structural Geology 21, 323-334. Bons, P.D., Urai J.L. 1992. Syndeformational grain growth: microstructures & kinetics. Journal of Structural Geology 14, 1101-1109. Bons, P.D., Koehn, D., Jessell, M.W. (Eds) (2008) Microdynamic Simulation. Lecture Notes in Earth Sciences 106, Springer, Berlin. 405 pp. ISBN 978-3-540-44793-1.

Formatted: Font: Not Italic, Pattern: Clear

620

Bons, P.D., Jansen, D., Mundel, F., Bauer, C.C., Binder, T., Eisen, O., Jessell, M.W.,

621

Llorens, M.G., Steinbach, F., Steinhage, D. Weikusat, I., 2016. Converging flow and

622

anisotropy cause large-scale folding in Greenland's ice sheet. Nature communications 7,

623

11427.

624

Borthwick, V., Piazolo, S., Evans, L., Griera, A., Bons, P.D. 2013. What happens to

625

deformed rocks after deformation? A refined model for recovery based on numerical

626

simulations. In: "Deformation Structures and Processes within the Continental Crust",

627

Special Volume 394, Geol. Soc. London. doi: 10.1144/SP394.11

628 629 630

Cloos, E. 1955. Experimental analysis of fracture patterns. Geological Society of America Bulletin 66, 241-256. Cook, A., Vel., S., Johnson, S.E., Gerbi, C., Song, W.J., 2013, Elastic and Seismic Properties

631

(ESP) Toolbox (beta version), http://umaine.edu/mecheng/faculty-andstaff/

632

senthil-vel/software/ESP_Toolbox/

633

Cross, A. J., Ellis, S., Prior, D. J. 2015. A phenomenological numerical approach for

634

investigating grain size evolution in ductiley deforming rocks. Journal of Structural

635

Geology 76, 22-34.

636

Cyprych, D., Piazolo, S., Almqvist, B.S.G. 2017. Seismic anisotropy from compositional

637

banding in granulites from the deep magmatic arc of Fiordland, New Zealand, Earth and

638

Planetary Science Letters 477, 156-167.

639

Ebner, M., Koehn, D., Toussaint, R, Renard, F. 2009. The influence of rock heterogeneity on

640

the scaling properties of simulated and natural stylolites. Journal of Structural Geology 31,

641

72-82.

642 643

Eisenlohr, P., Diehl, M., Lebensohn, R.A., Roters, F., 2013. A spectral method solution to crystal elasto-viscoplasticity at finite strains. International Journal of Plasticity 46, 37-53.

644 645 646 647 648 649 650 651 652 653

EPICA Community members (2004) Eight glacial cycles from an Antarctic ice core. Nature 429, 623-628. Etchecopar, A. 1977. A plane kinematic model of progressive deformation in a polycrystalline aggregate. Tectonophysics 39, 121-139. Fagereng, Å., Hillary, G. W. B., Diener, J. F. A. 2014. Brittle-viscous deformation, slow slip and tremor. Geophysical Research Letters 41, 4159-4167. Ford, J. M., Wheeler, J., 2004. Modelling interface diffusion creep in two-phase materials. Acta Materialia 52, 2365-2376. Ford, J. M., Wheeler, J., Movchan, A. B., 2002. Computer simulation of grain boundary creep. Acta Materialia 50, 3941-3955.

654

Ford, J.M., Ford, N.J., Wheeler, J. 2004. Simulation of grain boundary diffusion creep:

655

analysis of some new numerical techniques: Proc R Soc, London 460, 2395-2413

656

Gardner, R., Piazolo, S., Evans, L., Daczko, N. 2017. Patterns of strain localization in

657

heterogeneous, polycrystalline rocks - a numerical perspective, Earth and Planetary

658

Science Letters, 463, 253-265.

659

Gratier, J.P., Dysthe, D.K., Renard, F., 2013. The Role of Pressure Solution Creep in the

660

Ductility of the Earth's Upper Crust, in: Dmowska, R. (Ed.), Advances in Geophysics, Vol

661

54. Elsevier, Amsterdam, pp. 47-179.

662 663 664

Griera, A., Bons, P.D., Jessell, M.W., Lebensohn, R.A., Evans, L., Gomez-Rivas, E. 2011. Strain localization and porphyroclast rotation. Geology 39, 275-278. Griera, A., Llorens, M.-G., Gomez-Rivas, E., Bons, P.D., Jessell, M.W., Evans, L.A.,

665

Lebensohn, R. 2013. Numerical modelling of porphyroclast and porphyroblast rotation in

666

anisotropic rocks. Tectonpophysics 587, 4-29,

667

Griera, A., Gomez-Rivas, E., Llorens, M.-G., Bons, P.D., Lebensohn, R. 2015.

668

Microstructural evolution of polycrystalline materials in simple shear: insights from full-

669

field numerical simulations. Geotectonic Research 97, 37–39.

670 671 672

Gottstein, G. 2004. Integral materials modelling. Modelling and Simulation in Materials Science and Engineering, 12, I. Godinho, J.R., Piazolo, S., Evans, L., 2014. Simulation of surface dynamics during

673

dissolution as a function of the surface orientation: Implications for non-constant

674

dissolution rates. Earth and Planetary Science Letters 408, 163-170.

675 676 677

Godinho, J.R.A., Piazolo, S., Evins, L.Z. 2012. Effect of surface orientation on dissolution rates and topography of CaF 2, Geochimica et Cosmochimica Acta 86, 392-403. Gomez-Rivas, E., Griera, A., Llorens, M-.G., Bons, P.D., Lebensohn, R.A., and Piazolo, S.

678

(2017). Subgrain rotation recrystallization during shearing: insights from full-field

679

numerical simulations of halite polycrystals. Journal of Geophysical Research, in press.

680

Grujic, D., Stipp, M., Wooden, J.L., 2011. Thermometry of quartz mylonites: Importance of

681

dynamic recrystallization on Ti‐ in‐ quartz reequilibration. Geochemistry, Geophysics,

682

Geosystems, 12.

683

Groome, W. G., Johnson, S. E., Koons, P. O. 2006. The effects of porphyroblast growth on

684

the effective viscosity of metapelitic rocks: implications for the strength of the middle

685

crust. Journal of Metamorphic Geology 24: 389–407.

686

Herwegh, M., Handy, M. R. 1998. The evolution of high-temperature mylonitic

687

microfabrics: Evidence from simple shearing of a quartz analogue (norcamphor). J. Struct.

688

Geol. 18, 689-710.

689

Herwegh, M., Handy M. R.. 1998 The origin of shape preferred orientations in mylonite:

690

inferences from in-situ experiments on polycrystalline norcamphor. Journal of Structural

691

Geology 20, 681-694.

692

Hilgers, C., Urai, J.L., Post, A.D., Bons, P.D. 1997. Fibrous vein microstructures:

693

Experimental and numerical simulation. Aardkundige Mededelingen 8, 107-109.

694

Hirth, G., Tullis, J. 1992. Dislocation creep regimes in quartz aggregates. Journal of

695 696 697 698 699

Structural Geology 14, 145 - 159. Hobbs B.E., Means W.O., Williams P.F. 1976. An Outline of Structural Geology. John Wiley and Sons, 571 pp. Hobbs, B. E., & Ord, A. (2014). Structural geology: the mechanics of deforming metamorphic rocks. Elsevier.

700

Jansen, D., Llorens, M.-G., Westhoff, J., Steinbach, F., Kipfstuhl, S., Bons, P.D., Griera, A.,

701

Weikusat, I. 2016. Small-scale disturbances in the stratigraphy of the NEEM ice core:

702

observations and numerical model simulations. The Cryosphere 10, 359-370.

703

Jiang, Z., Prior, D. J., Wheeler, J., 2000. Albite crystallographic preferred orientation and

704

grain misorientation distribution in a low-grade mylonite: implications for granular flow.

705

Journal of Structural Geology 22, 1663-1674.

706 707 708 709 710 711 712 713

Jessell, M.W. 1986. Grain boundary migration and fabric development in experimentally deformed octachloropropane. Journal of Structural Geology 8, 527-542. Jessell, M. W. 1988a. Simulation of fabric development in recrystallizing aggregates—I. Description of the model. Journal of Structural Geology 10, 771-778. Jessell, M. W. 1988b. Simulation of fabric development in recrystallizing aggregates—II. Example model runs. Journal of Structural Geology 10, 779-793. Jessell, M. W., Lister, G. S. 1990. A simulation of the temperature dependence of quartz fabrics. Geological Society, London, Special Publications 54, 353-362.

714

Jessell, M., Bons, P.D., Evans, L., Barr, T., Stüwe, K. 2001. Elle: the numerical simulation of

715

metamorphic and deformation microstructures. Computers And Geosciences 27, 17-30,

716

http://dx.doi.org/10.1016/S0098-3004(00)00061-3

717

Jessell, M.W., Siebert, E., Bons, P.D., Evans, L., Piazolo, S. 2005. A new type of numerical

718

experiment on the spatial and temporal patterns of localization of deformation in a

719

material with a coupling of grain size and rheology. Earth and Planetary Science Letters

720

239, 309-326.

721

Karato, S. I., Paterson, M. S., Fitz Gerald, J. D., 1986. Rheology of synthetic olivine

722

aggregates: influence of grain size and water. Journal of Geophysical Research 91, 1851-

723

8176.

724 725 726 727 728 729 730

Karato, S.I., Li, P., 1992. Diffusion creep in perovskite: implications for the rheology of the lower mantle. Science 255, 1238. Kelka, U., Koehn, D., Beaudoin, N. 2015. Zebra pattern in rocks as a function of grain growth affected by second-phase particles. Frontiers in Physics 3, 74. Kelka, U., Veveakis, M., Koehn, D., Beaudoin, N., 2017. Zebra rocks: compaction waves create ore deposits. Scientific Reports 7, 14260. Kim, S. G., Kim, D. I., Kim, W. T., Park, Y. B. 2006. Computer simulations of two-

731

dimensional and three-dimensional ideal grain growth. Physical Review E 74, 061605.

732

Koehn, D., Bons, P.D., Passchier, C.W. 2003. Development of antitaxial strain fringes during

733

non-coaxial deformation: an experimental study. Journal of Structural Geology 25, 263-

734

275.

735

Koehn, D., Renard, F., Toussaint, R., Passchier, C. 2007. Growth of stylolite teeth patterns

736

depending on normal stress and finite compaction. Earth and Planetary Science Letters,

737

257, 582-595.

738 739

Koehn, D., Ebner, M., Renard, F., Toussaint, R., Passchier, W. 2012. Modelling of stylolite geometries and stress scaling. Earth and Planetary Science Letters 341, 104-113.

740

Koehn, D., Sachau, T. 2014. Two-dimensional numerical modeling of fracturing and shear

741

band development in glacier fronts. Journal of Structural Geology 61, 133-142.

742

(doi:10.1016/j.jsg.2012.11.002).

743

Koehn, D., Rood, M.P., Beaudoin, N. , Chung, P., Bons, P.D., Gomez-Rivas, E. (2016) A

744

new stylolite classification scheme to estimate compaction and local permeability

745

variations. Sedimentary Geology 346, 60-71.

746 747 748 749 750

Krill Iii, C. E., Chen, L. Q. 2002. Computer simulation of 3-D grain growth using a phasefield model. Acta Materialia 50, 3059-3075. Lebensohn, R. A. 2001. N-site modeling of a 3D viscoplastic polycrystal using fast Fourier transform. Acta Materialia 49, 2723-2737. Lebensohn, R.A., Kanjarla, A.K., Eisenlohr, P., 2012. An elasto-viscoplastic formulation

751

based on fast Fourier transforms for the prediction of micromechanical fields in

752

polycrystalline materials. International Journal of Plasticity 32, 59-69.

753

Lister GS, Paterson MS 1979. The simulation of fabric development during plastic

754

deformation and its application to quartzite: fabric transitions. J Struct Geol 1, 99-115.

755

Llorens, G.-M., Griera, A., Weikusat, I., Bons, P.D., Roessiger, J., Lebensohn, R.A. 2016a.

756

Dynamic recrystallisation of ice aggregates during co-axial viscoplastic deformation: a

757

numerical approach. Journal of Glaciology 62, 359-377.

758

Llorens, G.-M., Griera, A., Bons, P.D., Lebensohn, R.A., Evans, L.A., Jansen, D., Weikusat,

759

I. 2016a2016b. Full-field predictions of ice dynamic recrystallisation under simple shear

760

conditions. Earth and Planetary Science Letters 450, 233-242.

761

Llorens, M. -G., Griera, A., Steinbach, F., Bons, P. D., Gomez-Rivas, E., Jansen, D.,

762

Roessiger, J., Lebensohn, R. A. and Weikusat, I. 2017. Dynamic recrystallization during

763

deformation of polycrystalline ice: insights from numerical simulations. Phil. Trans. R.

764

Soc. A 375, 20150346.

765 766 767 768

Llorens, M.G., Bons, P.D., Griera, A., Gomez-Rivas, E. and Evans, L.A., 20132013a. Single layer folding in simple shear. Journal of Structural Geology 50, 209-220. Llorens, M-.G., Bons, P.D., Griera, A. Gomez-Rivas, E. 20132013b. When do folds unfold during progressive shear?. Geology 41, 563-566

769

McCrone WC, Chen PT .1949. Grain growth in octochloropropane. J Appl Phys 20, 230-231.

770

Means, W.D. 1977. A deformation experiment in transmitted light. Earth Planet. Sci. Letts.,

771 772 773 774 775

35, 169-179. Means W.D. 1980. High temperature simple shearing fabrics: a new experimental approach. J. Struct. Geol. 2, 197-202. Means W.D. 1983. Microstructure and micromotion in recrystallization flow of octachloropropane: a first look. Geol. Rundsch. 72, 511-528.

776

Means, W. D., Ree, J. H. 1988. Seven types of subgrain boundaries in

777

octachloropropane. Journal of Structural Geology 10, 765-770.

778

Montagnat, M., Blackford, J.R., Piazolo, S., Arnaud, L., Lebensohn, R.A., 2011.

779

Measurements and full-field predictions of deformation heterogeneities in ice. Earth and

780

Planetary Science Letters 305, 153-160.

781

Montagnat, M., Castelnau, O., Bons, P.D., Faria, S.H., Gagliardini, O., Gillet-Chaulet, F.,

782

Grennerat, F., Griera, A., Lebensohn, R.A., Moulinec, H., Roessiger, J., Suquet, P. 2014.

783

Multiscale modeling of ice deformation behavior. Journal of Structural Geology 61, 78-

784

108.

785

Nama, T.N., Otoh, S., Masuda, T. , 1999. In-situ annealing experiments of octachloropropane

786

as a rock analogue: kinetics and energetics of grain growth. Tectonophysics 304, 57-70.

787

Naus-Thijssen, F.M., Goupee, A.J., Johnson, S.E., Vel, S.S., Gerbi, C., 2011. The influence

788

of crenulation cleavage development on the bulk elastic and seismic properties of

789

phyllosilicate-rich rocks. Earth and Planetary Science Letters 311, 212-224.

790 791 792 793 794 795 796

Park Y, Means WD. 1996. Direct observation of deformation processes in crystal mushes. J Struct Geol 18, 847-858. Park, Y., Park, D., Evans, L., Ree, J. H. 2004. An Elle-based 2-D model for cation exchange reaction between garnet and biotite. Journal of the Virtual Explorer 15. Passchier, C.W., Trouw, R.A.J., 2005. Microtectonics (2nd enlarged edition). Springer Verlag, Berlin. Petit, J.R., J. Jouzel, D. Raynaud, N.L. Barkov, J.-M. Barnola, I. Basile, M. Bender, J.

797

Chappellaz, M. Davis, G. Delaygue, M. Delmotte, V. M. Kotlyakov, M. Legrand, V.Y.

798

Lipenkov, C. Lorius, L.Pépin, C. Ritz, E.Saltzman, M. Stievenard 1999, Climate and

799

atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature

800

399, 429-436.

801

Piazolo, S., Austrheim, H., Whitehouse, M., 2012. Brittle-ductile microfabrics in naturally

802

deformed zircon: Deformation mechanisms and consequences for U-Pb dating. American

803

Mineralogist 97, 1544-1563.

804

Piazolo, S., Bons, P.D., Jessell, M.W., Evans, L., Passchier, C.W. 2002. Dominance of

805

microstructural processes and their effect on microstructural development: insights from

806

numerical modelling of dynamic recrystallization. Geol. Soc, London, Spec. Publ. 200,

807

149-170.

808

Piazolo, S., Jessell, M. J., Prior, D. J., Bons, P. D. 2004. The integration of experimental in-

809

situ EBSD observations and numerical simulations: a novel technique of microstructural

810

process analysis. Journal of Microscopy, 213, 273-284.

811

Piazolo, S., Jessell, M.W., Bons, P.D., Evans, L., Becker, J.K. 2010. Numerical simulations

812

of microstructures using the Elle platform: A modern research and teaching tool. Journal

813

of the Geological Society of India 75, 110-127.

814

Piazolo, S., La Fontaine, A., Trimby, P., Harley, S., Yang, L., Armstrong, R. and Cairney,

815

J.M., 2016. Deformation-induced trace element redistribution in zircon revealed using

816

atom probe tomography. Nature communications, 7. Piazolo, S., Kaminsky, F.V., Trimby,

817

P., Evans, L., Luzin, V., 2016. Carbonado revisited: Insights from neutron diffraction,

818

high resolution orientation mapping and numerical simulations. Lithos 265, 244-256.

819

Piazolo, S., Montagnat, M., Grennerat, F., Moulinec, H., Wheeler, J., 2015. Effect of local

820

stress heterogeneities on dislocation fields: examples from transient creep in

821

polycrystalline ice. Acta Materialia 90, 303-309.

822 823 824 825 826 827 828 829 830

Plümper, O., Røyne, A., Magrasó, A., Jamtveit, B. 2012. The interface-scale mechanism of reaction-induced fracturing during serpentinization. Geology 40, 1103-1106. Putnis, A. 2009. Mineral replacement reactions. Reviews in mineralogy and geochemistry 70, 87-124. Putnis, A., Austrheim, H. 2010. Fluid‐ induced processes: metasomatism and metamorphism. Geofluids 10, 254-269. Ramberg, H., 1981. Gravity, deformation, and the earth's crust: In theory, experiments, and geological application. Academic press. Ran, H., de Riese, T., Llorens, M.-G., Finch, M., Evans, L.A., Gomez-Rivas, E., Griera, A.,

831

Jessell, M.W., Lebensohn, R.A., Piazolo, S., and Bons, P.D. 2018. Time for anisotropy:

832

The significance of mechanical anisotropy for the development of deformation structures.

833

Journal of Structural Geology, in press.

834 835

Ree, J. H., Park, Y., 1997. Static recovery and recrystallization microstructures in sheared octachloropropane. Journal of Structural Geology 19, 1521–1526.

836

Reddy, S. M., Timms, N. E., Pantleon, W., Trimby, P. 2007. Quantitative characterization of

837

plastic deformation of zircon and geological implications. Contributions to Mineralogy

838

and Petrology 153, 625-645.

Formatted: English (United States), Pattern: Clear

839

Roessiger, J., Bons, P.D., Griera, A., Jessell, M.W., Evans, L. Montagnat, M., Kipfstuhl, S.,

840

Faria, S.H., Weikusat, I. 2011. Competition between grain growth and grain size reduction

841

in polar ice. Journal of Glaciology 57 942-948.

842 843 844 845 846

Roessiger, J., Bons, P.D., Faria, S.H. 2014. Influence of bubbles on grain growth in ice. Journal of Structural Geology 61, 123-132. Rosenberg, C. L. 2001. Deformation of partially molten granite: a review and comparison of experimental and natural case studies. International Journal of Earth Sciences 90, 60-76. Rosenberg, C.L., Handy, M.R. 2000. Syntectonic melt pathways during simple shearing of a

847

partially molten rock analogue (Norcamphor‐ Benzamide). J. Geophys. Res. 105, 3135-

848

3149.

849

Rosenberg, C.L., Handy, M.R. 2001. Mechanisms and orientation of melt segregation paths

850

during pure shearing of a partially molten rock analog (norcamphor–benzamide). J. Struct.

851

Geol. 23, 1917-1932.

852

Roters, F., P. Eisenlohr, C. Kords, D. D. Tjahjanto, M. Diehl, D. Raabe. 2012. DAMASK:

853

the Düsseldorf Advanced MAterial Simulation Kit for studying crystal plasticity using an

854

FE based or a spectral numerical solver. Procedia IUTAM 3, 3-10.

855 856

Rutter, E.H., 1976. The kinetics of rock deformation by pressure solution. Philosophical Transactions of the Royal Society, London, 203-219.

857

Sachau, T., Koehn, D. 2010. Faulting of the lithosphere during extension and related rift-

858

flank uplift: a numerical study. International Journal of Earth Sciences 99, 1619-1632.

859

Sachau, T., Koehn, D. (2012) Melange: a viscoelastic lattice-particle model applicable to the

860

lithosphere. Geochemistry, Geophysics, Geosystems 13, 1-26.

861

Schenk, O., Urai, J. L., 2004. Microstructural evolution and grain boundary structure during

862

static recrystallization in synthetic polycrystals of sodium chloride containing saturated

863

brine. Contributions to Mineralogy and Petrology 146, 671–682.

864

Schenk, O., Urai, J.L. 2005. The migration of fluid-filled grain boundaries in recrystallizing

865

synthetic bischofite: first results of in-situ high-pressure, high-temperature deformation

866

experiments in transmitted light. J. Metamorph. Geol. 23, 695-709.

867 868 869

Schmid, S.M., Panozzo, R., Bauer, S., 1987. Simple shear experiments on calcite rocks: rheology and microfabric. Journal of structural Geology 9, 747-778. Seyfried, W.E., Foustoukos, D.I., Fu, Q., 2007. Redox evolution and mass transfer during

870

serpentinization: An experimental and theoretical study at 200° C, 500bar with

871

implications for ultramafic-hosted hydrothermal systems at Mid-Ocean

872

Ridges. Geochimica et Cosmochimica Acta 71, 3872-3886.

873

Smith, J.R., Piazolo, S., Daczko, N.R., Evans, L., 2015. The effect of pre‐ tectonic reaction

874

and annealing extent on behaviour during subsequent deformation: insights from paired

875

shear zones in the lower crust of Fiordland, New Zealand. Journal of Metamorphic

876

Geology 33, 557-577.

877

Spruzeniece, L., Piazolo, S., Maynard-Casely, H.E., 2017. Deformation-resembling

878

microstructure created by fluid-mediated dissolution-precipitation reactions, Nature

879

Communications 8.

880

Steinbach, F., Bons, P.D., Griera, A., Jansen, D., Llorens Verde, M.G., Roessiger, J.,

881

Weikusat, I., 2016. Strain localization and dynamic recrystallization in the ice–air

882

aggregate: a numerical study. The Cryosphere 10, 3071-3089.

883

Steinbach F., Kuiper, E-J. N., Eichler, J., Bons P.D., Drury M. R., Griera, A., Pennock, G.M.,

884

Weikusat I., 2017. The Relevance of Grain Dissection for Grain Size Reduction in Polar

885

Ice: Insights from Numerical Models and Ice Core Microstructure Analysis. Frontiers in

886

Earth Science 5, 66.

887

Stipp, M., Stünitz, H., Heilbronner, R., Schmid, S.M. 2002. The eastern Tonale fault zone: a

888

'natural laboratory' for crystal platic deformation of quartz over a temperature range from

889

250 to 750 °C. J. Struct. Geol. 24, 1861-1884.

890

Sundberg, M. & Cooper, R. F., 2008. Crystallographic preferred orientation produced by

891

diffusional creep of harzburgite: Effects of chemical interactions among phases during

892

plastic flow. Journal of Geophysical Research-Solid Earth 113.

893 894

ten Brink CE, Passchier CW. 1995. Modelling of mantle porphyroclasts using nonNewtonian rock analogue materials. J Struct Geol 17, 131-146

895

Taylor GI. 1938. Plastic strain in metals. Journal of the Institute of Metals 62, 307–324.

896

Urai, J.L., 1983. Water assisted dynamic recrystallization and weakening in polycrystalline

897 898 899 900

bischofite. Tectonophysics 96, 125-157. Urai, J. L., Humphreys, F. J., 1981. The development of shear zones in polycrystalline camphor. Tectonophysics, 78, 677–685. Urai, J. L., Humphreys, F. J., Burrows, S. E., 1980. In-situ studies of the deformation and

901

dynamic recrystallization of rhombohedral camphor. Journal of Materials Science 15,

902

1231–1240.

903

Urai, J. L., Means, W. D., Lister, G. S. 1986. Dynamic recrystallization of minerals. In:

904

Mineral and Rock Deformation: Laboratory Studies (edited by Hobbs, B. E. & Heard, H.

905

C.). Am. Geophys, Un. Geophys. Monogr. 36,161-199.

906

Vel, S.S., Cook, A., Johnson, S.E., Gerbi, C., 2016. Computational Homogenization and

907

Micromechanical Analysis of Textured Polycrystalline Materials, Computer Methods in

908

Applied Mechanics and Engineering 310, 749-779.

909 910

Vernon, R.H., 2004. A practical guide to rock microstructures. Cambridge University Press, Cambridge.

911

Walte, N.P., Bons, P.D., Passchier, C.W. 2005. Deformation of melt-bearing systems -

912

insight from in situ grain-scale analogue experiments. Journal of Structural Geology 27,

913

1666-1679.

914 915 916 917

Wark, D.A., Watson, E.B., 2006. TitaniQ: a titanium-in-quartz geothermometer. Contributions to Mineralogy and Petrology 152, 743-754. Wheeler, J., 2009. The preservation of seismic anisotropy in the Earth’s mantle during diffusion creep. Geophysical Journal International 178, 1723-1732.

918

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

919

relation to grain rotation, grain boundary sliding and superplasticity. Philosophical

920

Magazine 90, 2841-2864.

921

Wheeler, J. 2014. Dramatic effects of stress on metamorphic reactions. Geology 42, 647-650.

922

Wheeler, J., 2018. The effects of stress on reactions in the Earth: sometimes rather mean,

923

usually normal, always important. Journal of Metamorphic Geology 36, 439-461.

924

Wilson, C. J. L. 1986. Deformation induced recrystallization of ice: the application of in situ

925

experiments. Mineral and rock deformation: laboratory studies: The Paterson volume,

926

213-232.

927

Wilson, C. J., Peternell, M., Piazolo, S., Luzin, V. 2014. Microstructure and fabric

928

development in ice: lessons learned from in situ experiments and implications for

929

understanding rock evolution. Journal of Structural Geology 61, 50-77.

930 931 932 933

Zöllner, D. 2011. A Potts model for junction limited grain growth. Computational Materials Science 50, 2712-2719.

Formatted: English (United Kingdom)

Formatted: English (United Kingdom)

934

935

Table Captions

936

Table 1 List of numerical models of microstructural (mm to dm) development identifying

937

processes modelled, numerical method used and providing relevant references. This list is

938

restricted to geological applications and those referenced in the main text. Note processes are

939

categorized as "S" for static (material points do not move) and "D" for dynamic (material

940

points move). Furthermore, unless stated otherwise models are two-dimensional.

941

Abbreviations: TBH – Taylor Bishop Hill calculation method for crystal lattice rotation,

942

VPFFT – Viscoplastic Fast Fourier Transform based model, EVPFFT - Elasto-viscoplastic

943

Fast Fourier Transform based model, FEM Finite Element, reXX – recrystallization.

944 945

Figure Captions

946

Figure 1 Microstructural development during in-situ deformation of the rock analogue

947

deformationoctochloropropane within a circular shear zone (Bons and Jessell, 1999); dashed

948

red lines indicate the shear distribution between the two steps shown; octachloropropane,.

949

Experiments run with top to the right shear, at an average strain rate isof 4.6·10-4 s-1 where

950

the strain rate near edge of the shear zone ( 1.2·10-3 s-1) is 10 x higher than in the top half of

951

the image ( 1.2·10-4 s-1). (a) at t1 at a bulk shear strain of 40; (b) t1+16 min; note.. Note

952

multiple concurrent processes: grain boundary migration, leading to dissection of grains

953

(locations 1 & 2), subgrain rotation (black arrow), nucleation (white arrow). The different

954

shear rates lead to a different balance of recrystallization processes and differences in

955

microstructures. At the low shear rate grains are equant, have straight sub-grain boundaries

956

and basal planes at an angle to the NS and EW polarisers. The high-strain rate zone shows

957

serrate grain boundaries, an oblique grain-shape foliation, basal grains approximately parallel

958

to the shear-zone boundary, as well as shear localisation on grain boundaries (location 3),

Formatted: Space After: 0 pt

959

indicative of grain boundary sliding/shearing. Such "micro-shear zones" may now have been

960

detected in polar ice sheets as well (Weikusat et al., 2009).

961

Figure 2 Numerical representation of a microstructure;. (a) micrographMicrograph of

962

quartzite; (b) numerical representation combining an element data structure with nodes (black

963

circles), segments (black lines) and polygons (enclosed area) and a lattice data structure with

964

unconnected lattice points (open circles). This structure is used in the numerical platform Elle

965

(see text for details).

966

Figure 3 Numerical modelling of dynamic recrystallization – a historical perspective;

967

mineral modelled is quartz; simple shear (see text for details); (a) numerical microstructure

968

after = 1; different grey scales signify different crystallographic orientations; top inset shows

969

data structure of hexagonal lattice points (modified after Jessell and Lister, 1990); (b)

970

numerical microstructure after = 1; colours show crystallographic orientation, grain

971

boundaries are red, subgrain boundaries black; for data structure see inset (modified after

972

Piazolo et al., 2002).

973

Figure 4 Numerical modelling of dynamic recrystallization – testing the effect of process

974

combination on microstructural development; model parameters: mineral modelled is- ice;

975

intrinsic grain boundary mobility M0 = 1·10-10 m2kg-1s-1 (see Llorens et al., 2017 for

976

details); time step =- 20 years, simple shear; ∆ =- 0.04 per time step. FFT and GBM signify

977

fast fourier transform formulation for crystal plasticity and grain boundary migration,

978

respectively; (a) initial microstructure and results after=2.4 for different process

979

combinations. Results are shown as grain network with orientation related colour coding

980

according to crystal orientations relative to the shortening direction y (see legend) and in pole

981

figures, where. In the latter the colour bar indicates the multiples of uniform distribution; see

982

text for further details; (a) initial microstructure (left) and numerical results and results

Formatted: Font: Not Bold, English (United Kingdom)

Formatted: Font: Not Bold, English (United States)

983

after=2.4; (b) grain area distribution normalized to the initial average grain area for all

984

models shown.

985

Figure 5 Results from diffusion creep modelling in pure shear; 2D microstructure as in the

986

starting frame of Fig 1b of Wheeler (2009);). Shown is the oblique view of microstructure

987

with patterns of normal stress shown along grain boundaries in the 3rd dimension as “fences”;

988

red”. Red arrows show stretching direction, hence stresses are tensile (shown as negative) on

989

boundaries at a high angle to stretching direction are tensile. Fences are colour coded

990

according to dissolution rate with blue low and red high. When there is no relative grain

991

rotation the fences have a single colour and the stress is parabolic. When there is relative

992

grain rotation the fences vary in colour and the stress is a cubic function of position.

993

Figure 6 Dynamic development of stylolite roughness in numerical simulations; (a) time

994

series (left to right) with the stylolite nucleating in the middle of a slow dissolving layer

995

(layer in green and stylolite in black colour). Once the stylolite has dissolve the layer on one

996

side the layer starts to pin and teeth develop. (b) Variation of the pinning strength of the layer

997

in three different simulations showing a strong dependency. The compaction (movement of

998

upper and lower walls) is shown in quite arrows. L in the picture on the right hand side is the

999

initial position of the layer and P shown as black arrows indicates the pinning of the layer

1000

upwards and downwards during dissolution.

1001

Figure 7 Example of new development in microdynamic numerical modelling; preliminary

1002

results (see text for details). (a) Coupling of iffusiondiffusion creep and surface energy driven

1003

grain boundary migration (GBM); (left) starting microstructure; (middle) microstructure at

1004

stretch 2; (right) graph showing average aspect ratio versus stretch; note that the

1005

microstructure after a significant period of exclusivelyexclusive grain boundary migration is

1006

the same as the starting microstructure as no movement occurs as all triple junctions are 120°;

1007

number°. Number of grains stays constant for all simulations. (b) Evolution of chemical

1008

concentration of an arbitrary element during surface energy driven GBM. The material is a

1009

single-phase polycrystalline aggregate with different initial chemical content. This example

1010

assumes very low bulk diffusion (Dbulk =1e-20 m2/s) and fast grain boundary diffusion

1011

(Dboundary =1e-8 m2/s), hence grain boundary diffusion dominates; colour. Colour code

1012

indicates chemical concentration; and white lines displayingrepresent grain boundaries. (c)

1013

Dissolution of reactive surfaces in 3D; material using the example of fluorite; microstructures

1014

dissolution. Microstructures after three dissolution durationsperiods are show (150hrs, 180

1015

hrs, 276 hrs); surfaces correspond to the {111} plane at the start of experiment/simulation;

1016

colours identify different depths where blue signifies low and red high; images are 60 μm

1017

width; note the formation of etch pits with similar triangular shape and the faster/enhanced

1018

dissolution of the grain boundaries in both experiment and numerical simulation; (top. Top

1019

panel row) shows numerical results; (lower. Lower panel row)shows experimental results

1020

using confocal microscopy images of a grain of a sintered CaF2 pellet at the same three

1021

dissolution times.

1022 1023

as the numerical models. Note the formation of etch pits with similar triangular shape and

1024

the faster/enhanced dissolution of the grain boundaries in both experiment and numerical

1025

simulations.

Formatted: Space After: 0 pt, Don't adjust space between Latin and Asian text, Don't adjust space between Asian text and numbers

Formatted: English (Australia)

a

t1

b

t1+16 min

strain rate increase

Figure1

Figure 2

unconnected polygon node

fsp

qtz

a

b

segment

node

Figure 3

grain

a

γ=1

b

γ=1

Figure 4

initial

FFT+GBM

γ = 2.4 {0001}

(a) c

a

{1120}

FFT+GBM+recovery +nucleation

FFT+GBM+recovery

{1010}

γ = 2.4 {0001}

{1120}

{1010}

γ = 2.4 {0001}

{1120}

b

8 {0001} {1120} {1010}

(b) 30

log (area/initial average grain area)

y {0001}

{1120} {1010}

frequency

0

FFT+GBM FFT+GBM+recovery

20

FFT+GBM+recovery +nucleation 10

0

-1.5

-1

-0.5

0

0.5

1

1.5

{1010}

Figure 5

0.04 0.02

Stress

0 -0.02 -0.04 -0.06

5

-0.08

4 3 0

2 1

1 2

0 3

-1 4

-2 5 -3

6 -4

Figure 6

t1

nucleation t 2 compaction

roughening in time compaction

t3 t4 teeth growth compaction

p L p

layer dissolved

small parts pinning increase in pinning strength

layer pinning

Figure 7

a

diffusion creep average aspect ratio

3

starting geometry

en masse neighbour switch diffusion creep

diffusion creep + low GBM rate 2

diffusion creep + low GBM rate diffusion creep + high GBM rate

diffusion creep + high GBM rate 1 1

2

3 Stretch

4

b

progression in time 10

c etch pits

150 hrs

enhanced dissolution

180 hrs

chemical concentration

100

grain boundary

276 hrs

1.62 micron

enhanced dissolution relative height etch pits - 1.54 micron

Table 1

(micro) structure/process General papers – review, overview, state of the art Single processes Crystal lattice rotation for single phase

Numerical Method

References Jessell et al., 2001; Jessell and Bons, 2002; Bons et al., 2008; Piazolo et al., 2010

D D D

slip, rotation, translation TBH VPFFT

D

VPFFT – 3D

Etchecopar, 1977 Lister and Paterson, 1979 Lebensohn et al., 2001; Montagnat et al. ,2011, 2014 Roters et al., 2012; Eisenlohr et al., 2013

Grain rotations for two phases

D

VPFFT

Elasto-viscoplastic behaviour

D D

Diffusion creep for single phase

D

EVPFFT lattice spring model + viscous deformation front tracking

Diffusion creep for two phases

D

front tracking

Dissolution and/or precipitation

D

lattice spring model

Dissolution at reactive surfaces

D

front tracking front tracking - 3D

Grain growth – isotropic surface energy – isotropic surface energy

S SD

Roessiger et al. 2011, 2014 Bons & Urai, 1992

S S SD S SD

front tracking front tracking (triple points only) front tracking front tracking Potts – 2 & 3D front tracking phase field – 3D

D S D

front tracking front tracking FEM – thin sheet model

Hilgers et al. ,1997; Koehn et al., 2003 Becker et al., 2008; Roessiger et al., 2011 Barr and Houseman, 1992, 1996

S D

finite difference VPFFT

Park et al., 2004 Griera et al., 2011, 2015; Ran et al., 2018

S

Potts-like

Borthwick et al., 2013

D

FEM + front tracking

Groome et al., 2006; Smith et al., 2015

D

TBH+Potts-like

Jessell, 1988a, b; Jessell & Lister, 1990

S D

front tracking TBH+FEM+front tracking

Roessiger et al., 2011 Piazolo et al., 2002

D

VPFFT+ front tracking

D

FEM+front tracking

Jansen et al., 2016; Llorens et al., 2016a, b, 2017; Steinbach et al., 2016, 2017; Gomez-Rivas et al., 2017; Piazolo et al., this contribution Jessell et al., 2005; Gardner et al., 2017

D D S

FEM front tracking front tracking

Cross et al., 2015 Piazolo et al. this contribution Piazolo et al. this contribution

– anisotropic surface energy – growth with pinning – isotropic grain growth – two phases Growth of crystals into a crack or in strain fringes Strain-induced grain boundary migration Single phase ductile deformation without reXX + crystallography Grain boundary diffusion Strain/stress localisation Combination of Processes Intracrystalline recovery and subgrain rotation Ductile deformation and grain growth polyphase Dynamic reXX: - crystal plastic deformation, grain boundary migration, rotation recrystallization - grain growth and polygonisation - crystal plastic deformation, grain boundary migration, rotational recrystallization, nucleation, recovery - crystal plastic deformation, grain boundary migration, rotational recrystallization, recovery Strain/stress localization - linked to grain size variations & rheology - linked to grain size variations & rheology Diffusion and grain growth Trace element diffusion and reXX

Griera et al., 2011, 2013, 2015; Ran et al., 2018 Lebensohn et al., 2012 Sachau and Koehn, 2010, 2012; Arslan et al., 2012; Koehn and Sachau, 2014 Ford et al., 2002, 2004; Wheeler, 2009, 2010 Berton et al., 2006, 2011; Ford and Wheeler, 2004 Koehn et al., 2006, 2007, 2012, 2016; Ebner et al., 2009, Godinho et al., 2014 Piazolo et al. this contribution

Becker et al., 2008; Piazolo et al., 2016 Kelka et al., 2015 Kim et al. 2006; Krill and Chen, 2002 Roessiger et al., 2014 Ankit et al., 2015