From Continuum Modeling, to Geometric Simulation, to Mean-Field

1 downloads 0 Views 1MB Size Report
composer J.S. Bach concluded each of his church Cantatas, and with that example I use ...... jected surface energy for a unit of length along an axis provided by the planar ...... height. This approach result shows that (a) the stability boundary for faceted interfaces .... We consider surfaces formed by a single crystal with cubic.
NORTHWESTERN UNIVERSITY

Evolving Faceted Surfaces: From Continuum Modeling, to Geometric Simulation, to Mean-Field Theory

A DISSERTATION SUBMITTED TO THE GRADUATE SCHOOL IN PARTIAL FULFILLMENT OF THE REQUIREMENTS for the degree DOCTOR OF PHILOSOPHY

Field of Applied Mathematics

By Scott A. Norris

EVANSTON, ILLINOIS December 2006

2

c Copyright by Scott A. Norris 2006

All Rights Reserved

3

ABSTRACT

Evolving Faceted Surfaces: From Continuum Modeling, to Geometric Simulation, to Mean-Field Theory

Scott A. Norris

We first consider the directional solidification, in two dimensions, of a dilute binary alloy having a large anisotropy of surface energy, where the sample is pulled in a high-energy direction such that the planar state is thermodynamically prohibited. Analyses including reduction of dynamics, matched asymptotic analysis, and energy minimization are used to show that the interface assumes a faceted profile with small wavelength. Questions on stability and other dynamic behavior lead to the derivation of a facet-velocity law. This shows the that faceted steady solutions are stable in the absence of constitutional supercooling, while in its presence, coarsening replaces cell formation as the mechanism of instability. We next proceed to introduce a computational-geometry tool which, given a facetvelocity law, performs large-scale simulations of fully-faceted coarsening surfaces, first in the special case with only three allowed facet orientations (threefold symmetry), and then for arbitrary surfaces. Topological events including coarsening are comprehensively

4

considered, and are treated explicitly by our method using both a priori knowledge of event outcomes and a novel graph-rewriting algorithm. While careful attention must be paid to both non-unique topological events and the imposition of a discrete timestepping scheme, the resulting method allows rapid simulation of large surfaces and easy extraction of statistical data. Example statistics are provided for the threefold case based on simulations totaling one million facets. Finally, a mean-field theory is developed for the scale-invariant length distributions observed during the coarsening of one-dimensional faceted surfaces. This theory closely follows the LSW theory of Ostwald ripening in two-phase systems, but the mechanism of coarsening in faceted surfaces requires the derivation of additional terms to model the coalescence of facets. The model is solved by the exponential distribution, but agreement with experiment is limited by the assumption that neighboring facet lengths are uncorrelated. However, the method concisely describes the essential processes operating in the scaling state, illuminates a clear path for future refinement, and offers a generic framework for the investigation of faceted surfaces evolving under arbitrary dynamics.

5

Acknowledgements Soli Deo Gloria – To God alone be glory. With this great cry of the Reformation, the composer J.S. Bach concluded each of his church Cantatas, and with that example I use it to conclude this thesis. While fully accepting that evolution is the only well-reasoned theory describing the mechanism of our origins, I still believe, on the basis of Christ’s resurrection, that God is ultimately responsible in some way for everything that exists. Thus, the credit for anything of worth in this work belongs ultimately and solely to Him. I could thank God for many things here, but will limit myself to two. First, I am grateful for Northwestern’s Graduate Christian Fellowship, where I found friends who love me, colleagues who have shared my struggles, and a continual struggling conversation about how to honor God amidst academic pursuit. Second, as I have struggled to justify the pursuit of mathematics in a world with so much loneliness and suffering, I thank God for giving me a vision of how my profession fits into the larger whole of the person He wants me to become. My quest at this institution, and indeed my life as a whole, would be much emptier without the presence of His truth and love. Mom and Dad, thank you for your encouragement, love, and for teaching me all of things growing up that have allowed me to succeed on my own. I am so grateful for all the time and energy you have poured into me, and I hope I can do half as well with my own children. Keith and Carla, thank you for being great in-laws – I’ll always look forward to spending time with you guys. Finally, to my dear wife Tara, thank you for

6

your constant love and support – for celebrating with me in my successes, encouraging me in my failures, and assuring me above all that you’d keep loving me no matter what happened. It has been a long road, but sharing it with you has given me strength and hope along the way. I have been blessed to make some very good friends during my time here, and so I thank: Michael and Bolu, for two fantastic years as roommates, much encouragement, and the times I borrowed your cars; Matt, Edy, and Young Cheol, for a truly special year of friendship and fellowship in small group, and many great eating expeditions; Kyle, for many enjoyable lunches, and for teaching me to almost love the Mac; Gogi, for more lunches and for laughing with me about the ups and downs of grad life; and Peter, for all the wonderful conversations, reflections, and video game binges. You guys have truly blessed my life, and it was you that made me sad to leave Chicago. A good adviser makes graduate school easier, and I’ve been lucky to have two. Steve, a huge thanks for your flexibility in allowing me to telecommute, and for always being available for guidance and advice despite your busy schedule. Most of all, thank you for your example of quiet humility and patience despite all you’ve achieved. Stephen, thank you for our many conversations of all kinds, and for the inspiration of your sheer love of doing mathematics. Also, thank you especially for sharing your life with me as well as your ideas. To both of you, thanks for leading me to a topic of research that has turned out to be a lot of fun to pursue. No matter the extent to which research is a part of the rest of my life, I will look back fondly on this work. Finally, despite rumors to the contrary propagated by PhD Comics, graduate students must eat more than Ramen, and therefore I thank NASA for supporting me during the

7

majority of my time here. Though I have to wonder if my work will ever directly aid the exploration of space, it is fun to be associated with a program that brought such wonder to my childhood, and an honor to know they thought my work promising enough to fund. In addition, I thank Northwestern and IGERT for support during my first and second years, respectively, and NSF, via Stephen, for that extra bit of commuting money third and fourth years.

8

Table of Contents ABSTRACT

3

Acknowledgements

5

List of Tables

11

List of Figures

12

Chapter 1. Introduction

14

Chapter 2. Faceted Interfaces in Directional Solidification

20

2.1. Introduction

20

2.2. Background: Governing equations, basic state, and linear stability

27

2.3. Statics. Non-planar interface due to strong anisotropy, φ = 0

39

2.4. Energetics. Optimal Wavelength, Comparison with Planar State

51

2.5. Dynamics. Stability, Wavelength Trapping, and Coarsening

56

2.6. Conclusions and Comments

62

Chapter 3. Large-Scale Simulations of Coarsening Faceted Surfaces

65

3.1. Introduction

65

3.2. Faceted Surfaces: Description, Kinematics, and Dynamics

68

3.3. Topological Events

71

3.4. Demonstration

76

9

3.5. Conclusions Chapter 4. The Kinematics of Faceted Surfaces with Arbitary Symmetry

82 84

4.1. Introduction

84

4.2. Data structures and simple motion: a 3D cellular network

87

4.3. Topological Events

90

4.4. Discretization and Performance of Topological Events

110

4.5. Demonstration and Discussion

117

4.6. Conclusions

122

Chapter 5. A Mean-Field Theory for Coarsening Faceted Surfaces

124

5.1. Introduction

124

5.2. Example Dynamics and Problem Formulation

127

5.3. Application: Our chosen facet dynamics

133

5.4. Solution and Comparison with Numerical-Experimental Data

134

5.5. Conclusions

136

References

140

Appendix A. Appendices for Chapter 1

149

A.1. Justification of the Quasi-steady state

149

A.2. Homogenized Linear Stability Analysis

151

Appendix B. Appendices to Chapter 3

154

B.1. Elaboration on Far-Field Reconnection

154

B.2. Non-Uniqueness

157

10

Appendix C. Appendix to Chapter 4 C.1. Numerical Simulation

163 163

11

List of Tables 3.1

Relevant coarsening phenomena.

78

12

List of Figures 2.1

Summary of linear stability results

36

2.2

Surface-energy minimizing slopes

40

2.3

Analysis of outer solutions

49

2.4

Sample faceted profiles

50

2.5

Representative behaviors of dynamic faceted interfaces

61

3.1

Facet Merge event

72

3.2

Merging Facet Pinch event

73

3.3

Vanishing Facet events

75

3.4

Example coarsening sequence

77

3.5

One-point geometric distributions

79

3.6

Correlational distributions

81

4.1

Diagram of neighbor relations

88

4.2

Normal diagrams for Vanishing Edge events

93

4.3

Neighbor Switch event

94

4.4

Irregular Neighbor Switch event

95

4.5

Facet Join event

96

13

4.6

Signatures of Constricted Facet events

98

4.7

Facet Pierce event

100

4.8

Irregular Facet Pierce event

101

4.9

Facet Pinch event

103

4.10

Joining Facet Pinch event

104

4.11

Example Vanishing Facet Event

106

4.12

Facets Vanishing in a group

108

4.13

Facets Vanishing as a step

109

4.14

Coarsening sequence with threefold symmetry

119

4.15

Coarsening sequence with fourfold symmetry

120

4.16

Coarsening sequence with sixfold symmetry

121

5.1

Survey of coarsening behavior

128

5.2

Representative facet configuration

129

5.3

Comparison between theory and experiment

138

5.4

Correlations in neighboring facet lengths

139

B.1

Method of listing binary trees

161

B.2

Saddle versions of Vanishing Edge events

162

14

CHAPTER 1

Introduction This thesis, in the most general sense, is about what happens to materials with a large anisotropy of surface energy when placed in dynamic situations. Thus, we begin with an overview of the definition, causes, and effects of large surface energy anisotropy. A unique property of crystals (in contrast to amorphous solids like glass) is that their material properties usually depend on orientation. This is due, in turn, to the packing symmetries of the internal crystal lattice on which individual atoms reside. When such materials expose a planar interface, the structure of that interface looks different at atomic scales depending on its orientation relative to the crystal lattice. These different surface configurations cause the material to react differently with its surrounding environment. Thus, the surface energy γ (and many other properties) depends on the surface orientation relative to the lattice. This dependence is modeled by making these quantities functions of an angle θ, which describes the deviation of the interface normal from some reference orientation associated with the lattice. While many different functions describing anisotropy may be considered, it will suffice for what follows to consider the particular form γ(θ) = γ0 [1 + α4 cos(4θ)]. This surface energy for crystals in two dimensions exhibits fourfold symmetry, with mean magnitude γ0 and relative anisotropy varying through the parameter α4 .

15

The earliest studies of anisotropy focused on the equilibrium problem, which seeks the surface-energy minimizing shape in two dimensions of a particle in equilibrium with a surrounding melt, given the form of γ(θ). This problem was solved in 1901 by Wulff [1], who devised a geometric method of constructing the correct shape, now known as the Wulff construction. For the example surface energy γ considered here, increasing the value of the anisotropy coefficient α4 from zero causes an initially round particle to change shape, eventually developing missing orientations and corners at the critical value of α4 = 1/15. A related problem inquires about the equilibrium shape of an initially planar interface exposed to its melt at the high-energy orientation θ = 0. Here, Herring [2] also used geometric considerations to show that the surface remains planar for small α4 , but at the identical critical value of α4 = 1/15, this interface becomes thermodynamically unstable and is replaced by a completely faceted “sawtooth” interface. Herring also showed that these two equilibrium results are linked, as it is precisely the missing orientations of the former problem that are thermodynamically unstable in the latter. Furthermore, the shared critical value is no accident – high-energy orientations will be unstable in any context if the surface stiffness γ + γθθ of that orientation is negative [3]. This thermodynamic instability is a key concept in the following chapters, and may be generally remembered as follows: “an interface orientation is thermodynamically unstable if it can reduce its surface energy by assuming a faceted sawtooth form.” The modeling of thermodynamically-faceting surfaces in dynamic situations was initiated by Mullins [4] in 1961 and Cabrera [5] in 1963; these authors noted independently the similarities between the process of surface faceting and the then-fledgling modeling

16

of evolving phase boundaries. However, even as understanding of phase boundaries progressed, including the seminal work of Cahn and Hilliard [6], the problem of faceting dynamics lay dormant for nearly thirty years, until it was picked up again and studied from two perspectives – purely continuum thermodynamic models [7, 8, 9], and continuum approximations of discrete step-flow models [10, 11, 12]. In each of these works, initially planar interfaces are shown to undergo spinodal decomposition into faceted hill-and-valley structures, just as one would expect from knowledge of the equilibrium problem. However, after this process is completed, coarsening phenomena are observed, where small facets shrink to zero length and vanish, causing a corresponding continuous increase in the average length of those that remain. Of even more interest is that these coarsening faceted surfaces are often observed to obey dynamic scaling, in which the surface looks the same at all sizes if first scaled by the average length. In such a state, the statistical geometric properties of the surface remain constant even as the average lengthscale grows, and this constant state can be considered to concisely summarize all of the salient information present in the evolving system. This progression – from faceting, to coarsening, to a dynamically scaling state – represents a typical surface behavior in models of many systems, and is the pattern on which this thesis will focus. In this historical context, then, we begin in Chapter 1 by applying mathematical methods recently fruitful in other faceting contexts to the problem of the directional solidification of a dilute binary alloy – a system which, though possessing a rich history of its own, has been mostly neglected in the context of faceting. While the inclusion of a solute field makes this system is slightly more complex than the pure-material systems

17

considered previously, the mathematical progression follows a similar course, with similar results. Starting from a full free-boundary problem, thermodynamic instability of an initially planar surface with negative surface stiffness leads to faceting. This occurs initially on a small wavelength which, in turn, allows a simplification to a partial differential equation that governs just the interface position. Matched asymptotic analysis of this equation reveals a family of steady faceted solutions, and energetic considerations specify an optimal member. Turning to the unsteady problem, we follow Watson [13] to obtain a key theoretical simplification – the further reduction of surface dynamics to a facet-velocity law which describes the normal velocity of each facet as a function of its geometric configuration. This law reveals that, below the critical pulling speed causing supercooling, the steady faceted profiles are stable. However, above that critical speed, coarsening occurs, which (a) replaces the formation of cells as the principal mechanism of instability, and (b) appears at pulling speeds smaller than those at which cells would otherwise appear. Chapters 2 and 3 begin by observing that facet velocity laws of the type found in Chapter 1 exist for many systems. Because they so efficiently describe the evolving faceted surface, they allow correspondingly efficient computational methods; and since the statistical study of dynamic scaling necessitates the simulation of large surfaces, the development of a tool to exploit such theoretical simplification is urgently needed. While easily implemented for 1+1D surfaces z = h(x, t), such tools are more difficult to construct for the 2+1D surfaces z = h(x, y, t) commonly seen in crystal-growing experiments. This is due to the occurrence of various topological events allowing surface re-organization and coarsening. Existing simulation attempts must choose between speed [14, 15] and robust

18

topological handling [16, 17]; here we present a method that achieves both. The method is outlined in Chapter 2 for threefold-symmetric surfaces, where it is used to study the dynamic scaling exhibited by Watson’s annealing dynamics [13]. The method is then more fully described and generalized to arbitrary symmetry in Chapter 3, where the same phenomenon is illustrated by the dynamics derived in Chapter 1. In addition to speed and topological accuracy, a further advantage of the method over previous approaches is easy access to geometric data presented by the surface. This is achieved by choosing a data structure which mirrors the natural structure of the surface, and demonstrated by the exhibition of numerous statistical measures of surface geometry present in the dynamically scaling states of the systems studied. Finally, since dynamic scaling pushes complex evolving surfaces into a state which can be effectively characterized by just a few statistics, it is natural to seek simplified models which replicate this behavior. Chapter 4 illustrates just such a model, which exploits the dynamics derived in Chapter 1 to describe the evolution of the facet length distribution to the scale-invariant steady state found using the tool in Chapters 2 and 3. The model recalls the famous Lifshitz, Slyozov, and Wagner theory [18, 19, 20] of Ostwald ripening in a two-phase system; however, due to the geometric consequences of coarsening in faceted systems, it also includes a coagulation term reminiscent of that in the work of Von Smoluchowski [21] and Schumann [22]. While certain simplifying assumptions keep the resulting model from quantitatively matching numerically-collected data, it qualitatively illustrates the essential forces at work in the dynamically scaling state, while suggesting more general models that would increase its predictive value.

19

In summary, this thesis traces the particular faceting system of directional solidification through a series of generally-applicable mathematical treatments. These include the analytical reduction of a free-boundary problem to a facet-velocity law describing surface evolution (Chapter 1), the numerical study of coarsening through an efficient computational tool (Chapters 2-3), and the development of a mean-field theory that attempts to describe the resulting dynamically scaling state (Chapter 4). These three tools – continuum modeling, geometric simulation, and mean-field analysis – naturally parallel the three stages of faceted-surface evolution – faceting, coarsening, and dynamic scaling. They thus together form a framework for the comprehensive study and comparison of faceting in many contexts.

20

CHAPTER 2

Faceted Interfaces in Directional Solidification 2.1. Introduction In the process of directional solidification, a sample liquid is pulled through a temperature gradient produced by the presence of a heating element and a cooling element, set at temperatures above and below the freezing point of the liquid, respectively. At a position in between these elements, the liquid freezes, forming a liquid/solid interface. If the material being solidified is a binary alloy, then solute is rejected at the interface, and must diffuse away into the bulk liquid, creating a solute gradient directed oppositely to the thermal gradient. At high enough pulling speeds, the concentration gradient steepens sufficiently to create an instability, replacing normally planar interface morphologies with more complicated cellular or dendritic structures [23]. These phenomena may be easily observed when transparent organic alloys are solidified within a narrow channel between two glass planes, an environment known as a Hele-Shaw cell [24]. Because complex behavior may thus be easily observed and measured, and because the nearly-two-dimensional nature of the Hele-Shaw cell leads to analytical simplicity, this particular procedure has been popular for the comparison of theory and experiment. Combined with the fact that the instability described above is similar to important hydrodynamic instabilities [25],

21

the opportunities afforded by directional solidification have made it one of the most commonly studied forms of crystal growth, and placed it among the classical problems of mathematical physics.

2.1.1. History 1: Directional Solidification The first qualitative explanation of the morphological instability described above was given in 1953 by Rutter and Chalmers [26]; these ideas were quantified the same year by Tiller et. al. [27]. These authors argued that at high-enough pulling speeds, the steepening solute gradient at the interface eventually leads to a layer of liquid that is supercooled for its chemical composition. In the presence of this consitutional supercooling, the planar interface was predicted to be unconditionally unstable. This idea was generalized in 1964 to include the effect of surace energy, when Mullins and Sekerka [23] performed a linear stability analysis showing that surface energy could stabilize the interface against small-wavelength perturbations even in the presence of constitutional supercooling1. This accurately predicts the phenomenon of absolute stability, where the front restabilizes at very high solidification rates as the effective surface energy regains dominance over the solute gradient. Their analysis also predicted the critical wavelength at which instability would occur, allowing for careful comparison with experimental results. In 1970, Wollkind and Segel [28] extended this analysis into the weakly nonlinear regime. For pulling velocities near the critical velocity, they derived ordinary differential equations (Landau equations) describing the post-instability amplitude of cellular 1If

thermal conductivity is greater in the solid than in the liquid (common for metals), they also showed that instability is possible without constitutional supercooling in the liquid. However, the central idea is that a solute gradient large enough to overcome the thermal gradient is necessary for instability; thus, the idea of supercooling as a necessary condition is still instructive.

22

solutions having the critical wavelength predicted by Mullins and Sekerka. This analysis predicted the conditions under which the instability is sub-critical or super-critical. In limits where the critical wavelength is large compared to the solute boundary layer thickness, the result of weakly nonlinear analysis is not an ODE governing the amplitude of a single cellular mode, but rather a PDE governing the interface evolution as a whole. In 1983 Sivashinsky [29] performed the first such analysis of solidification, obtaining, in the limit of small segregation coefficient, an equation that describes subcritical bifurcations. In 1988, Brattkus and Davis [30] studied the near-absolute-stability limit, obtaining an equation describing supercritical bifurcation. These cases were then generalized in 1990 into a single framework by Riley and Davis [31], who derived, in an intermediate limit, equations able to capture the change from subcritical to supercritical bifurcation. Beyond the Mullins-Sekerka instability, cellular interfaces are generically exhibited, and as pulling speeds increase, these grow in amplitude, until a secondary instability causes the formation of dendritic structures [32, 33]. In these regimes, numerical simulation is a primary tool of investigation, and research has focused in several areas. In the cellular regime, detailed examination of cell shape were carried out in [34, 35, 36], while questions of wavelength selection have been investigated by looking for steady solutions using an integral formulation [37, 38]. Finally, in the dendritic regime, phase-field methods have been developed to study the shape of solidifying structures in three dimensions [39, 40]. The previous review covers only the behavior of the simplest possible solidification model, including only the effects of solute diffusion and rejection and surface energy; a

23

more comprehensive review including many generalizations may be found in [41]. However, as we have seen, even this minimal model exhibits a wide variety of complex behavior. The important general theme to extract from the above summary is the competition between a destabilizing solute gradient and a stabilizing surface energy. The progression from planar, to cellular to dendritic solutions as pulling speed increases is driven by the corresponding increase in solute gradients at the interfaces. However, the effective surface energy is also dependent on pulling speed, and as it overtakes solute gradients in relative strength it causes a reverse progression from dendritic, to cellular, and finally back to planar states in the absolute stability limit. Since in what follows we consider the effect of modifications to surface energy, it is important to keep this basic scenario firmly in mind.

2.1.2. History 2: Anisotropy The generalization to the above model of primary interest here is that we allow the surface energy γ to be anisotropic. As discussed in the Introduction, we let γ depend on the surface orientation θ, which measures the angle between the surface normal and a reference orientation associated with the bulk crystal lattice. The anisotropy of γ may be classified as either “small” or “large,” depending on whether or not the surface stiffness γ + γθθ [42] is strictly positive or not. Large anisotropy, in which we are most interested, has long been studied in equilibrium problems, where geometric considerations reveal that cornering and faceting are generically present on energy-minimizing interfaces [1, 2, 3]. However, the study of large anisotropy in dynamic systems is problematic because, in the Gibbs-Thompson-like equations describing interface motion, the surface stiffness

24

appears multiplying the curvature, which is ordinarily the highest derivative present in the equation. Thus, for orientations with negative surface stiffness, the Gibbs-Thompson equation is ill-posed mathematically, and some way to regularize the problem is needed. The way forward is provided adding relevant physics in the form of a small corner energy, which penalizes rapid changes in orientation [43, 42, 2]. Modeling this energy results in the inclusion of higher-order derivative terms, which supply the needed mathematical regularization [44, 45]. The strategy just described has been applied to enable the study of large anisotropy in a variety of dynamic settings [7, 8, 9, 10, 11, 12]. The generic behavior revealed by these studies is that initially planar surfaces rapidly decompose into a faceted sawtooth, or “hill-and-valley” configuration with very small wavelength. Having done so, such faceted surfaces then proceed to increase that wavelength via coarsening, where small facets shrink and vanish, causing in increase in the average length of those that remain.

2.1.3. History 3: Directional Solidification and Anisotropy A natural step at this point is to inquire what happens when large anisotropy, with its thermodynamic instability leading to faceting, is added to the above model of solidification, where solute gradients drive a morphological instability. However, despite the fascinating behavior of faceted surfaces caused by large anisotropy in other dynamic contexts, most work in directional solidification has considered only the case of small anisotropy2. In that case, because all orientations are thermodynamically stable, no new instabilities appear, 2We

note that [46, 47] considers large-anisotropy solidification, but in a low-energy direction which is thermodynamically stable. Also, [48] considers faceted cellular solidification above the supercooling speed, but simply starts with a faceted array, and does not consider how such an array came to exist.

25

and so the effect on, for example, linear stability [49], amplitude equations [50], and longwave reductions [51] is only modulatory. Such a treatment is valuable – part of the appeal of the transparent organics used in Hele-Shaw cells is that they, like many metals, indeed possess only small anisotropies [52]. However, most non-metals have stronger anisotropies, and when directionally solidified in thermodynamically unstable directions, exhibit just the faceting behaviors predicted in other dynamic contexts (see, for example, [53, 54]). Additionally, the solidification of these materials in such orientations is interesting because of the possibility that the planar state, about which all subsequent analysis is typically based, may cease to occur at all. Besides admitting qualitatively new system behaviors, this possibility may require the development of novel analytical methods to describe those behaviors. Developing such methods and exploring this behavior is the aim of this work.

2.1.4. Kinetics Finally, since directional solidification is a dynamic process, we also include in our model the oft-neglected effect of attachment kinetics, which is also anisotropic. This property effectively represents the mobility µ(θ) of a moving interface. Differences in mobility correspond to differences in the amount of supercooling necessary to maintain a given interface speed, with the end result that the inverse mobility µ−1 (θ) also appears in the Gibbs-Thompson equation. Anisotropy of attachment kinetics has no small/large distinction as does that of surface energy; however, it can generate behaviors such as traveling waves [49, 55], and can additionally cause faceting itself in some circumstances.

26

Because the relative importance of kinetic anisotropy and surface energy anisotropy is unknown a priori, we include this effect in our analysis.

2.1.5. Summary The aim here, then, is to study the behavior of a binary alloy with large surface-energy anisotropy that is directionally solidified in a high-energy orientation (with negative surface stiffness). We are most interested in understanding the relationship between the morphological instability expected due to solute gradients with the thermodynamic instability expected due to negative surface stiffness. Specific questions we aim to answer include the following. Does faceting occur as in other dynamic contexts? How does the presence of solute gradients affect this behavior? Does a faceted steady state replace the usual planar state below supercooling? If so, what are the characteristics of this state, and what is the resulting effect on the solidified microstructure? What happens to such an interface when the pulling speed is increased past its supercooling critical value? Does the presence of this supercooled liquid layer cause further destabilization? Does coarsening occur? Under what circumstances? Finally, since solidification is a dynamic process, does anisotropy of attachment kinetics play a significant role in any of the above considerations? To briefly summarize our results, linear stability analysis reveals that, for large enough anisotropy, the high-energy planar state is indeed thermodynamically unstable for all concentrations and pulling speeds. The search for the anticipated faceted steady solutions reveals that concentration does not affect their shape, and leads to singularly perturbed equation describing the steady interface. Matched asymptotic analysis of this equation

27

indeed yields a family of nearly faceted hill-and-valley structures. Members of this family may be compared by considering an appropriate free energy, the minimization of which yields an optimal wavelength that scales with the small corner energy. Questions concerning interface dynamics lead to the derivation of a steepest-descent surface evolution, and ultimately, an effective facet dynamics. This dynamics reveals that faceted interfaces are stable below the supercooling speed, where coarsening is prohibited as a mechanism for dynamic wavelength adjustment. In contrast, above the supercooling speed, but below the usual Mullins-Sekerka limit, a reversal of the effective thermal gradient allows coarsening, which replaces cellular growth as the mechanism of instability . Interestingly, at no stage of this analysis does anisotropy of attachment kinetics play a significant role; however, the presence of kinetic effects, often neglected elsewhere, is pivotal in deriving the facet dynamics which so simplifies the later analysis.

2.2. Background: Governing equations, basic state, and linear stability Governing equations. We consider the directional solidification of a dilute binary alloy with anisotropic surface energy. We use the Frozen Temperature Approximation (FTA) [25], in which one neglects latent-heat generation, assumes equal thermal conductivities in the solid and liquid3, and assumes that thermal diffusion is much faster than solute diffusion, with the result that the temperature is linear over characteristic solutediffusion lengths. We also assume a one-sided model, which neglects diffusion in the solid phase. Let x be the co-ordinate lateral to the initially planar surface, z be normal to that surface, and shift to a frame of reference moving with the mean interface position. In this 3Note

that this assumption precludes the presence of instability below the constitutional supercooling velocity, a possibility footnoted above and discussed in [23].

28

two-dimensional co-ordinate system, the temperature T (z), concentration C(x, z, t), and interface position h(x, t) are then described by the following equations: (2.1a)

T = T0 + GT z

all z

(2.1b)

Ct = V Cz + D∇2 C

z > h(x)

(2.1c)

C → C∞

z→∞

(2.1d)

Cn = −[C(1 − k)]V¯n /D

z = h(x)

(2.1e)

T I = Tm + mC +

 Tm  (˜ γ + γ˜θθ )K − ν(Kss + K3 ) − µ ˜−1 V¯n Lv

z = h(x).

Here, GT is the (imposed) temperature gradient, V is the speed at which the sample is pulled through the temperature gradient, and C∞ is the original concentration of solute in the bulk sample. These three parameters are the extrinsic parameters subject to experimental control. The intrinsic parameters are more numerous. In the bulk, there is a free parameter T0 which anchors the z-coordinate, the melting point Tm of the pure solvent, the volumetric latent heat of fusion Lv , the diffusion coefficient D of the solute in the liquid, and the (negative) liquidus slope m describing freezing-point depression. At the interface, there is the local curvature K, the second derivative Kss of curvature with respect to arc length, a modified normal velocity V¯n (described below), the normal derivative Cn of concentration, the interface temperature T I , the segregation coefficient k describing solute rejection (the ratio, at an interface, of the concentration on the solid side to that on the liquid side), a small parameter ν which is the magnitude of the corner energy (typically near atomic scales), the anisotropic surface energy γ˜ (θ), and the

29

anisotropic mobility µ ˜(θ) describing attachment kinetics. Note the use of tilde to denote a few dimensional quantities; this is for notational convenience later. Here θ denotes the angle formed by the surface normal with the z-axis, so that hx = − tan θ. Also note that V¯n describes the normal velocity in the (old) stationary coordinate system. We’ll use Vn to describe normal velocity in the (new) moving coordinate; the two are related via the expression V + ht = V cos θ + Vn . V¯n = p 1 + h2x

(2.2)

To describe anisotropic surface energy and surface mobility, we use the identical smooth, sinusoidal functions (2.3)

γ˜ (θ) = γ0 γ(θ),

γ(θ) = 1 + α4 cos[4(θ − φ)]

(2.4)

µ ˜(θ) = µ0 µ(θ),

µ(θ) = 1 + β4 cos[4(θ − φ)],

where γ0 and µ0 are dimensional constants giving the average magnitudes of γ˜ and µ ˜, and γ and µ are nondimensional functions describing the variation in θ of the same. These forms have fourfold symmetry, and allow the direction of solidification to vary from the high-energy orientation by an angle φ. Thus, φ = π/4 describes solidification in a lowenergy direction (see [46]), while φ = 0 describes a high-energy direction. Equation (2.3) yields a surface stiffness of γ + γθθ = 1 − 15α4 cos[4(θ − φ)], so that the cases of “small” and “large” anisotropy are characterized respectively by α4 < 1/15 and α4 > 1/15.

30

Remark.

It will be noted that surface energy and mobility have identical form, and

thus, orientations with lowest (highest) surface energy are also those with the lowest (highest) mobility. This is due, like anisotropy itself, to atomic packing patterns. Orientations with low surface energy are typically atomically smooth, requiring energetically expensive atomic layer nucleations for growth, and therefore exhibiting low mobility. By contrast, orientations with high surface energy are typically atomically rough, offering a constant supply of neighbor-rich “holes” for growth, and therefore exhibiting high mobility. A recent overview of the derivation of these equations, and the associated assumptions, may be found in Chapter 3 of [41]. The related equations there have been modified, however, by the appropriate addition to Eqn. (2.1e) of regularization terms described in [44, 45], to allow the possibility of “large” anisotropy. Scaling and Non-Dimensionalization.

For steady, x-independent solutions of

Eqns. (2.1), one finds that 

V 1−k exp(− z) 1+ k D

(2.5a)

C = C∞

(2.5b)

T0 = Tm +

(2.5c)

h = 0.

mC∞ −µ ˜−1 (0)V k



We pause to note several things. First, the point z = 0 was placed to ensure h = 0, which specifies the constant T0 . Second, the positive value Tm − T0 is the total freezing point depression due to attachment kinetics and the presence of solute, which causes a displacement of the interface from the T = Tm isotherm – this displacement will be important later when we consider non-planar interfaces. Third, the concentration profile reacts to

31

this displacement by simply shifting with the interface – its shape is unchanged. Resuming our argument, the form of the steady solution (2.5) informs the following scalings and dimensionless variables: D ¯ [¯ x, z¯, h] V D t = 2 t¯ V  C∞  1 − (1 − k)C¯ C= k mC∞ D T = Tm + −µ ˜−1 (0)V + GT T¯ k V

(2.6a)

[x, z, h] =

(2.6b) (2.6c) (2.6d)

T¯ = z¯.

(2.6e)

Here temperature, fixed by the FTA, is scaled so as to replace it by the variable z¯. Also, the dimensionless concentration is scaled such that C = 0 at z = 0, which requires that it scale with the negative physical concentration. After performing these scalings, and eliminating all references to the fixed temperature T through (2.6e), we arrive at the following non-dimensional equations describing the evolution of solute concentration and interface shape (where bars have been dropped): (2.7a)

C t = C z + ∇2 C

for z > h(x)

(2.7b)

C →1

as z → ∞

(2.7c)

Cz = hx Cx + [1 − (1 − k)C](1 + ht )

on z = h(x)

(2.7d)

 C = M−1 h − Γˆ s(θ)K + ν¯ Kss + K3 + µ ¯−1 A(hx , ht )

on z = h(x)

32

where (2.8)

sˆ(θ; α4 , φ) = γ + γθθ

has been introduced as a shorthand for the surface stiffness, and (2.9)

(1 + ht ) −1 A(hx , ht ) = p µ (hx ) − µ−1 (0) 2 1 + hx

represents the effect of attachment kinetics on concentration at the interface. The dimensionless constants (M−1 , Γ, ν¯, µ ¯) are respectively proportional to temperature gradient, surface energy, corner energy, and mobility, and are given by (2.10a)

M−1 =

(2.10b)

Γ=

(2.10c)

ν¯ =

(2.10d)

µ ¯−1 =

k D GT V |m(1 − k)|C∞

V Tm k γ0 D Lv |m(1 − k)|C∞

V 3 Tm k ν D 3 Lv |m(1 − k)|C∞ k V µ−1 0 . |m(1 − k)|C∞

Of particular interest is the parameter M−1 , the reciprocal of the morphological number, which can be written in the simpler form GT /|m|GC , where GC = (1 − k)C∞ V /kD is the solute gradient at the interface. Thus if M−1 < 1, the gradient of freezing point depression is larger than the gradient of temperature, which precisely defines the existence of constitutional supercooling. In addition, we also note that the parameter ν¯ is typically very small, and it will be important in later asymptotic analysis.

33

Basic State and linear stability analysis. The non-dimensional version of the basic state (2.5) has the form: (2.11)

h0 = 0,

C0 = 1 − exp(−z).

To determine the linear stability of this state, we introduce and study the evolution of a ˆ C] ˆ as follows, small disturbance [h, ˆ h = h0 + h,

(2.12)

ˆ C = C0 + C,

ˆ |C|] ˆ C] ˆ ≪ [|h0 |, |C0 |]. Linearizing in [h, ˆ we find that the disturbance obeys the where [|h|, equations (2.13a)

Cˆt = Cˆz + Cˆzz + Cˆxx

for z > 0

(2.13b)

Cˆ → 0

as z → ∞

(2.13c) (2.13d)

ˆ = −(1 − k)[Cˆ + h] ˆ +h ˆt Cˆz − h ˆ = M−1 h ˆ − sˆ0 Γh ˆ xx + ν¯h ˆ xxxx + µ ˆ x + AT h ˆ t) Cˆ + h ¯−1 (AX h

on z = 0 on z = 0,

where Taylor series have been used to allow all boundary terms to be evaluated at z = 0. Here AX and AT describe the derivatives of A(hx , ht ) with respect to hx and ht , respectively, evaluated at hx = ht = 0. Also, it was first observed by Coriell and Sekerka [49] that only the surface stiffness of the planar interface sˆ0 is present; (2.14)

sˆ0 (α4 , φ) = sˆ(0; α4 , φ) = (1 − 15α4 ) cos(4φ).

34

This term multiplies the nondimensional surface energy Γ; thus, anisotropy here simply creates an effective surface energy parameter sˆ0 Γ compared to the isotropic case (where ˆ = [c(z), h] exp(σt) exp(iax) ˆ h] sˆ ≡ 1). From here, the introduction of normal modes [C, leads, upon considering Eqns. (2.13a,b), to the selection of a decaying exponential solution for c(z). Application of the boundary conditions (2.13c,d) then results a homogeneous system of algebraic equations, the solution of which requires the following implicit dispersion relation for σ: (2.15) σ+k i M−1 = 1 − sˆ0 Γa2 − ν¯a4 + µ . ¯−1 (iAX a + AT σ) + µ ¯−1 p 0 σ− h 1 2 + σ) − (1 − k) (1 + 1 + 4(a 2

We now investigate this relation in four cases of primary interest – small and large

surface energy anisotropy for high- and low-energy pulling orientations. We note that, for each of these cases, anisotropic attachment kinetic terms vanish. For both high-energy (φ = 0) and low-energy (φ = π/4) pulling directions, the term AX = 0, as can be seen by considering the symmetries in equation (2.4). At intermediate pulling orientations, this term causes traveling solutions [49], but that has already been investigated and will not be considered further in this paper. To study neutral stability boundaries, we set σ = 0 in expression (2.15), and note that, since σ = AX = 0, kinetics plays no role in neutral stability for our choice of φ. Before proceeding, however, we make an important observation. The resulting neutral stability curves would live in (M−1 , Γ, ν¯)-space, and measure the relative effects of morphological number (i.e. constitutional supercooling) and surface energy. These are, indeed, the theoretical effects at work, but they are awkward parameters to work with experimentally

35

– the directly controllable parameters of interest are pulling velocity and sample solute concentration. Additionally, we see that ν¯ is actually dependent on these parameters, and so makes a poor description of corner energy. For these reasons, “natural parameters” also introduced which scale independently on pulling speed, concentration, and corner energy [56]. The relationships between these sets of parameters are given by the transformations (2.16)

V=

r

Γ M−1

C=

r

1 M−1 Γ

N =

M−1 ν¯ . Γ2

With these parameter issues in mind, we now present neutral stability curves associated with each of the four cases described above in both of the discussed parameter spaces. Figure 2.1a shows curves in the experimental (log(C), log(V))-space with fixed N , while Figure 2.1b gives results in the theoretical (Γ, M−1 )-space with fixed N (not fixed ν¯). In each case, we fix k = 0.5, N = 10−4. Before considering the effect of anisotropy, it is instructive to consider the stability boundaries present in the isotropic problem (α4 = 0, sˆ0 = 1), shown in blue in both figures. In the theoretical space, the curve actually terminates on the axes; these terminal points correspond to the linear asymptotes observed in the experimental space. Theoretically, the interface is stable if M−1 > 1 or Γ > 1/k; the former is called the constitutional supercooling boundary, while the latter is called the absolute stability boundary. These correspond experimentally to the lower asymptote (shown) and the upper asymptote (not shown). The succinct summary of behavior is that some amount of constitutional supercooling M−1 < 1 is required for instability, but a large enough surface energy Γ can stabilize no matter the supercooling; thus the name absolute stability [23].

36

6

1.5 sˆ0 < 0

sˆ0 > 0

sˆ0 < 0

4

1 M−1

log(V)

2

0

sˆ0 > 0 −2

0.5

−4

−6 −4

−2

0

2 log(C)

4

6

8

0

0

0.5

1

1.5 Γ

2

2.5

3

Figure 2.1. Summary of linear stability results (color). Neutral stability curves are shown in both (C, V)-space (a), and (Γ, M−1 )-space (b). Blue shows isotropic stability boundary. Red shows solidification in high-energy direction (φ = 0). Green shows solidification√in low-energy direction (φ = π/4). Dashed lines represent α4 =√1/15 − N (small anisotropy), while dotted lines represent α4 = 1/15 + N (large anisotropy). The black line shows the constitutional supercooling boundary. The effect of φ. Turning to the effect of anisotropy, we begin by considering the effect of pulling direction. Recall that the only effect of anisotropy is the presence of the constant sˆ0 multiplying the non-dimensional surface energy Γ. For φ = π/4, sˆ0 > 1, and thus anisotropy enhances stabilizing effect of surface energy. However, for φ = π/4, sˆ0 > 1, which diminishes the surface energy stabilization. Thus, the φ = π/4 curves lie inside the isotropic curves, while the φ = 0 curves lie outside of them.

4

The effect of α4 . We next consider the effect of the anisotropy strength, expressed in the constant α4 . We see that the effect of α4 turns out to be φ-dependent. For both φ, two values of α4 are considered, one just below the “strong anisotropy” value of 1/15, and one just above. For the φ = π/4 case, crossing this boundary has no significant effect (ˆ s0 4The

effect in (Γ, M−1 )-space is simply to stretch or compress the neutral stability curve along the Γ-axis as compared to the isotropic case. The corresponding effect in (C, V)-space is a shift of the top branch.

37

simply crosses 2 – a quantitative change), while for the φ = 0 case, the effect is dramatic (ˆ s0 crosses 0, which is a qualitative change). The effect of sˆ0 . The above behavior is put into a single, simple framework if we only consider sˆ0 . Whenever sˆ0 > 0, surface energy is stabilizing. Thus, constitutional supercooling is required for instability, causing the three positive-ˆ s0 curves to remain anchored to the constitutional supercooling boundary, while the absolute stability boundary moves (ˆ s0 is, after all, simply a surface energy modifier). However, when sˆ0 < 0, surface energy becomes a destabilizing agent. Supercooling is no longer necessary for instability, as the negative surface stiffness encourages bending; thus, the neutral stability curve de-anchors from the supercooling boundary. Indeed, only the presence of corner energy allows a stable region at all. The de-anchored curves in Figure 2.1 have a very small negative value of sˆ0 , but as sˆ0 continues to decrease, the curves begin a singular migration upward (leftward) in the √ theoretical (experimental) parameter space. Finally, for sˆ0 < −2 N , the stable region disappears entirely, signaling universal instability. Mathematically, this can be most easily seen by converting to “natural” parameters, and finding the critical concentration Cc at which instability occurs, (2.17)

Cc = min a

1 V

[1 + sˆ0 (aV)2 + N (aV)4 ] . k 1 − 1 1+√1+4a 2 −(1−k) ] 2[

√ It turns out that, for sˆ0 < −2 N , Cc attains a negative value, implying instability for all positive (physical) C. The form of (2.17) also allows us to describe the asymptotic behavior of the neutral stability curve. Since the top and bottom branches represent large- and small-V limits, respectively, and since C is largely a function of aV, large V

38 √ implies small a and vice-versa. If we let sˆ0 = −2 N + ǫ, this allows an asymptotic approximation of the singular migration of Cc , showing that (2.18a) (2.18b)

√ log(Ctop ) ∼ − log(V) + log(ǫ) − log( N ) log(Cbottom ) ∼

log(V) + log(ǫ) − log(4N /k).

As ǫ → 0, the curve migrates singularly leftward, vanishing at ǫ = 0. Finally, it is instructive to examine the change that occurs in the dimensional critical wavelength of instability ac as sˆ0 crosses zero. This wavelength, which is O(D) for much p s0 |/N ) of the visible part of the curves in Figure 2.1 for sˆ0 > 0, becomes instead O( |ˆ

for sˆ0 < 0. Thus, accompanying the transition to negative surface stiffness is a transition of the instability wavelength from diffusional scales to corner scales. This occurs over the same values as the migration of the NSC, and indicates that surface energy becomes the primary cause of instability. √ Summary. For large enough anisotropy sˆ0 < −2 N , the typical conditional morphological instability on diffusional scales is replaced by universal thermodynamic instability on capillary scales. This result is generally unsurprising; the thermodynamic instability of planar surfaces with negative surface stiffness was is a result that was obtained by Herring [57]. However, that result is modified here by the presence of the regularization term. In particular, besides stabilizing the planar solution for not-too-negative surface stiffness, the corner energy coefficient provides a scale for the instability that occurs; no such scale exists for Herring’s instability [58]. The obvious remaining question, then, is what happens to the interface after instability?

39

2.3. Statics. Non-planar interface due to strong anisotropy, φ = 0 We now proceed to look for non-planar solutions present in the range of universal linear √ √ instability, where sˆ0 < −2 N . Specifically, we let α4 > 1/15 + 2 N while pulling in the thermodynamically unstable direction φ = 0. We begin by listing three expectations based on heuristic reasoning; these will then be detailed. Expectation 1: Shape.

We have already stated Herring’s result that, in equi-

librium, planar surfaces with large anisotropy of surface energy are thermodynamically unstable to lower-energy faceted surfaces. To explain this briefly, we consider the projected surface energy for a unit of length along an axis provided by the planar surface. This energy is described by (2.19)

E=

Z

γ(θ)ds =

Z

γ(q)

p

1 + q 2 dx,

where here and in what follows q = hx . In the presence of large anisotropy, this functional is minimized by any5 faceted hill-and-valley structure with slopes of q ∗ (α4 ), where q ∗ minimizes the integrand of (2.19). For the surface energy given in Eqn. (2.3), performing this minimization yields an implicit relation for the value of q ∗ : (2.20)

α4 =

(1 + q 2 )2 . 15 − 10q 2 − q 4

The value q ∗ (α4 ) will be called the optimal slope, and is plotted in Figure 2.2. The same result can also be obtained using the convexification argument found in [59], where q ∗ is found by locating double tangent points on the polar plot of inverse surface energy. Now, 5Note

that all hill-and-valley structures have the same projected surface energy. Thus, while surface energy anisotropy provides an energetically favored slope, it is not sufficient, by itself, to provide a scale[58].

40

1

q*

0.5

0

−0.5

−1

0

0.1

0.2

0.3

0.4

0.5 α4

0.6

0.7

0.8

0.9

1

Figure 2.2. Surface-energy-minimizing slope for various values of α4 . directional solidification includes many more effects than just surface energy anisotropy, including, importantly, solute rejection and diffusion and anisotropy of attachment kinetics. However, we have seen in the previous section that the thermodynamic instability due to surface energy is preserved, and indeed dominant, in the linear theory. Thus, we expect to see shapes similar to those found in the planar equilibrium problem – corners separating facets with energetically favorable slopes near q ∗ . Expectation 2: Size.

Below the constitutional supercooling speed, the unstable

planar state exists within a positive effective thermal gradient (c.f. [60]). We expect the interface to reduce surface energy by faceting, and the resulting change to the concentration field will result in a nonlinear interaction between the two. However, because of the thermal gradient, large interface deformations would result in large thermal energetic penalties. Thus, we expect an interface with small amplitude. If the facet slopes are not small, then this corresponds to a small wavelength λ as well, an expectation which will be useful throughout the analysis that follows. Expectation 3: Displacement. Equation (2.5) above shows how the planar state is displaced from the T = Tm isotherm by the presence of attachment kinetics. Because

41

this effect depends on orientation and the normal velocity of the surface, any faceted interface with slope q ∗ 6= 0 will be displaced by a mean amount that is different from the displacement of the planar interface. Turning to the solute concentration, we recall that any such kinetic displacement does not affect the shape of the concentration field, which is merely displaced along with the interface. Formulation. In light of our expectations, then, we look for a small, faceted nonplanar interface h, having slopes near q ∗ and with mean displacement Z from the z = 0 isotherm; this is accompanied by a corresponding small correction to the concentration profile C. Just as in the above linear stability analysis, we write (2.21)

ˆ h = Z + h,

ˆ C = C0 (z − Z) + C,

ˆ |C|] ˆ ≪ [|h0 |, |C0|]. In contrast to the linear theory, however, we do not assume where [|h|, that the corrections are infinitesimal. Therefore, we expand C(h) and Cz (h) about z = Z in the boundary conditions there, but we do not yet discard any nonlinear derivative

42

ˆ and Cˆ are then terms. The resulting semi-linearized equations governing h (2.22a) Cˆt = Cˆz + Cˆzz + Cˆxx

for z > Z

(2.22b) Cˆ → 0

as z → ∞

(2.22c) ˆ x Cˆx + k h ˆ − (1 − k)Cˆ + h ˆt Cˆz = h

on z = Z

(2.22d)  ˆ − Γˆ ˆ x, h ˆ t) Cˆ = M−1 Z + (M−1 − 1)h s(θ)K + νˆ Kss + K3 + µ ¯−1 A(h

on z = Z.

Looking for steady solutions (∂/∂t → 0), we first consider the interface h to be given, and solve the system of equations (2.22a-c) governing the concentration C. This will yield a steady solution for C which depends on the (fixed, but unknown) interface h. We then evaluate the solution for C at Z, and insert this value into the left-hand side of the Gibbs-Thompson condition (2.22d) to obtain a nonlinear equation describing h (c.f. [46]). Now, the many nonlinear terms retained in equations (2.22) will prevent analytical solutions at several points in the following analysis. This problem could be remedied by considering a small-slope limit of these equations (c.f. [61]). Such a situation can be created by choosing value of α4 that is only slightly above the critical value of 1/15, introducing into the problem a small parameter ε = 15α4 − 1. The optimal slope q ∗ can p √ be shown in this case to equal q ∗ = 3ε/8, suggesting a scaling x → εx; if performed, such a scaling would eliminate all the difficult nonlinearities. However, such a careful

43

selection of α4 is not very general, and so rather than applying this limit a priori, we prefer to call upon it when analytical difficulty requires, and retain nonlinearities where they can be handled. A postiore estimates well validate this approach.

2.3.1. Solution for C ˆ x Cˆx in the boundary condition (2.22c) Equations (2.22a-c) have a nonlinearity – the term h describing solute balance. Recalling the small-slope approximation idea just discussed, let us neglect it for a moment, and return to it later. Having done so, we perform a Fourier transform in the x-coordinate, after which exponentials in z satisfying (2.22a) are easily found. Application of the (transformed) boundary conditions (2.22b,c) (neglecting the ˆ x Cˆx term) at z = 0 then yields the solution: h (2.23)

ˆ z) = C(x,

Z



−∞

−k

¯ h(κ) exp[−p(κ)(z − Z)] exp(iκx) dκ p(κ) − (1 − k)

where (2.24) (2.25)

1 ¯ h(κ) = 2π p(κ) =

Z

1+



ˆ h(x) exp(−iκx)dx

−∞



1 + 4κ2 ; 2

here bars indicate Fourier transforms (in the variable κ), and hats still indicate the smallamplitude variables described above. We now need to obtain a form for the value of (2.23) at z = Z, for insertion into Eqn. ˆ ¯ (2.22d). Assuming periodic profiles h(x) with wavenumber κ0 , then h(κ) is just a series of ˆ Z) δ-spikes occurring at wavenumbers nκ0 . We can thus expand the expression for C(x,

44

as (2.26)

ˆ 0) = −k C(x,

∞ X

¯ ρ(nκ0 )h(nκ 0 ) cos(nκ0 x),

n=1

where (2.27)

1 ρ(κ) = p(κ) − (1 − k)

  1 ≈ for κ ≫ 1 . κ

ˆ We now see that this is roughly just h(x), but scaled by a factor ρ(κ0 ), and “rounded,” with higher-frequency modes damped. We now use the first application of our expectation of small wavelength. Given the asymptotic form ρ(κ) ∼ 1/κ, and the fact that the ˆ ∼ O(λ2 ). It turns out, then, that the wavelength satisfies λ = 2π/κ0 , we see that Cˆ ∼ λh correction to the concentration caused by a small non-planar interface is very small, and may usually be neglected in equations describing the interface h itself. ˆ x Cˆx term. While we cannot solve the equations We now re-consider the neglected h on the concentration field with this term included, we can use a scaling argument to show that the correction to the concentration field is still small. With λ again the small ˆ x now be O(1). Also, let the value of Cˆ at the interface wavelength, let the surface slope h be of unknown magnitude C. Now, we expect Cˆ at the interface to vary on the same ˆ itself, suggesting that Cx ∼ O(C/λ). Finally, following the argument above, scale as h equations (2.22a,b) can still be solved using Fourier transforms to show that the scale of Cˆz at the interface is also O(C/λ). Replacing each term in equation (2.22c) with its appropriate scale, we obtain the following dimensional form: (2.28)

C C ∼ + λ + C. λ λ

45

This “equation” is only for the purpose of identifying dominant terms, and is not an equality (specifically, we cannot cancel the two C/λ terms). We observe first that, when ˆ x Cˆx term is of the same C/λ size as the Cˆz term, which O(1) slopes are allowed, the h means (a) it cannot be neglected, and (b) probably has a qualitative effect on the shape of the concentration field. However, looking for a balance of terms in equation (2.28), we see that since C ≪ C/λ, we must balance C/λ and λ. This shows that again C ∼ λ2 . We ˆ remains O(λh), ˆ and thus, is still thus conclude that the magnitude of Cˆ evaluated at h ˆ itself. small enough to neglect in equations involving h

2.3.2. Solution for h Dropping the bars on h, a nonlinear equation governing the interface h is obtained by inserting the just-derived value of C at the interface into the Gibbs-Thompson Equation (2.22d). Since Equation (2.22d) contains terms h, and since we just saw that C ∼ O(λ2 ) ∼ O(λh), we conclude that this term is small enough to neglect. This leads to the equation: (2.29)

O(λ2 ) = M−1 Z + Geff h − Γˆ s(θ)K + δ 2 (Kss + K3 ) + µ ¯−1 A0 (q)

where Geff = (M−1 − 1) is an effective thermal gradient, δ =



ν¯ is a small parameter

associated with corner energy, and A0 (q) = A(hx , 0). The interface h is thus described by a singularly perturbed nonlinear equation, which, like equations describing faceting in other contexts, is similar to the Cahn-Hilliard equation describing phase separation [6]. The form of such equations suggests a matched asymptotic analysis; we follow [62] by looking first for an inner solution describing corners, which will then provide boundary conditions for an outer equation describing facets. This approach will yield a family of

46

composite solutions with varying wavelength, the comparison of which is the topic of Section 2.4. 2.3.2.1. Inner Scale δ. Since the primary effect of strong anisotropy is expected to be the presence of corners in solutions [1], we begin by looking at the inner scale where we expect to find them. Using the small corner-energy parameter δ, an inner equation for √ ¯ To leading order, this gives x, h]. Eqn. (2.29) is found with the scaling [x, h] → (δ/ Γ)[¯ the equation (2.30)

Kss + K3 = sˆ(θ)K.

Because this equation is strongly nonlinear, we limit ourselves to consideration of the small-slope form. It will be seen later that only the existence of an inner solution is necessary for further analysis, and the precise form of that solution is of little interest. The small-slope form of (2.30) is [61] (2.31)

hxxxx = (8h2x − 1)hxx .

This equation, while still nonlinear, may be directly integrated as follows (letting q = hx ):

(2.32a) (2.32b) (2.32c) (2.32d)

qxxx = (8q 2 − 1)qx 8 qxx = q 3 − q + A 3 1 2 2 4 1 2 q = q − q + Aq + B 2 x 3 2 p qx = 2W(q).

47

In the final step, W(q) is a double-welled potential for the slope q, and is simply the smallslope surface stiffness integrated twice. Solving this equation on (−∞, ∞) and requiring a bounded solution, we find that we must choose A and B according to the bitangent construction, which lets both wells of W rest exactly on the q = 0 axis. This procedure gives W(q) a final form of (2/3)(q 2 − 3/8)2 , and admits an exact solution for q: (2.33)

r

q=±

3 x tanh( √ ). 8 2

This clearly represents a corner in h, and matches precisely the two energetically favored slopes in the small-slope limit as described above. The inner, corner solution, in turn, provides boundary conditions for the outer, facet equation – any outer solution must connect to two corners, and must therefore twice achieve an optimal slope ±q ∗ . Remark.

While the corner solution (2.33) is valid only in the small-slope regime,

in general we require only the existence of some similar solution to provide boundary conditions on the outer equations. Since the bitangent construction described here can be viewed as the small-slope limit of the Wulff construction [61], this requirement is ensured. 2.3.2.2. Outer Scale. An outer equation describing facets is provided, as usual, by simply neglecting the corner term because it is multiplied by δ 2 : (2.34)

Geff h = Γˆ s(θ)K − [M−1 Z + µ ¯−1 A0 (q)].

Suitable solutions of Eqn. (2.34) must match our inner solutions (corners); i.e., they must twice attain one of the preferred slopes q ∗ (α4 ). In addition, they must have a mean height of zero, since mean displacement is described by the so-far-undetermined Z. To see what solutions may exist satisfying these requirements, we numerically examine all solutions of

48

(2.34) by restating it as a dynamical system in (h, q) (2.35a) (2.35b)

h˙ = q   Geff h + M−1 Z + µ ¯−1 A0 (q) (1 + q 2 )3/2 , q˙ = Γˆ s(θ)

where the dot represents differentiation in x. Now, to determine Z, we will first assume that Z = 0. This will produce a solution h with some nonzero mean height H. Then, the correct zero-mean solution is obtained by letting Z = (Geff /M−1)H. A representative resulting phase plane is given in Figure 2.3a for Geff = Γ = µ ¯ −1 = 1, α4 = β4 = 0.5, q ∗ (α4 ) = 0.8908; only the upper half is shown since the system is invariant under the transformation q → −q, x → −x. There, we see several interesting families of solutions; however, only those enclosed in the red triangular regions meet the boundary conditions just described. These solutions are displaced from the q-axis by the A0 (q) term, because the displacement due to attachment kinetics of the faceted state is different from that of the planar state. Notably, if µ ¯−1 = 0, then the system would be invariant under the transformation q → −q, h → −h, and thus, symmetric about h = 0; the corrective term Z would not be needed. Now, let these solutions be parametrized by L, the total solution length in x; then for each parameter set (Geff , Γ, µ ¯−1 , α4 , β4 ) there exists an implicit relation qmin (L) describing the minimum slope attained by each family member, and, following the above argument, a relation Z(L) describing the mean displacement from zero. These functions are shownin Figs. (2.3b,c) for the parameter set chosen above.

49

1

0.8

q

0.6

0.4

0.2

0 −1.5

−1

−0.5

0

0.5 h

1

1

1.5

2

2.5

1.2

0.9 1.1

0.8



qmin

0.7 1

0.6 0.5

0.9

0.4

0

1

2

3 L

4

5

0.8

0

1

2

3

4

5

L

Figure 2.3. (a) The phase plane for outer equation solutions with positive slopes (color). Admissible solutions live within the red triangular regions. Black solid (dashed) lines represent degenerate stable (unstable) regions. (b) The function qmin (L). (c) The function Z(L).

2.3.2.3. Composite Solution. Piecing together appropriate inner and outer solutions (corners and facets), we plot in Figure 2.4 some example composite solutions [h, q](x) for different L (note that L is the length of a single facet, and so the the total wavelength is 2L). These solutions are not precisely piecewise-planar (faceted), as would be expected in the equilibrium problem. This is caused primarily by the presence of the thermal

50

gradient; Figure 2.4 shows how this gradient “pressures” the facets to bend in such a way as to reduce the area of the solid which projects into the melt, and melt into the solid. However, this effect is only strongly present at large wavelengths. 1 1

q

h

2 0

0 −1

−1 0

2

4

6

8

10

1.4

0

2

4

6

8

10

0

0.2

0.4

0.6

0.8

1

0

0.02

0.04

0.06

0.08

0.1

1 q

h

1.2 0

1 −1 0.2

0.4

0.6

0.8

1

1.14

1

1.12

0

q

h

0

1.1 −1 0

0.02

0.04

0.06 x

0.08

0.1

x

Figure 2.4. Sample height and slope profiles for various values of L, for the parameters given in the text. Note that the interface h(x) is very nearly planar even for O(1) wavelengths. As the wavelength decreases, it becomes even more so – see figure (2.3b).

Since we have been anticipating a small-wavelength solution, we now inspect the smallL limit, which reveals two important facts. First, it is seen in Figure 2.3b, and can be shown analytically, that for small L, qmin (L) = q ∗ + O(L2 ). Since the slope of any solution h(x; L) lies on [qmin (L), q ∗ ], we have the important consequence that small wavelength solutions are nearly linear. Second, if solutions are nearly linear, then A0 (q) ≈ A0 (q ∗ ) everywhere (not of course, in the corners, but its value is irrelevant there). In this limit,

51

then, the interface displacement is decoupled from the interface shape, and tends to the value (2.36)

Z = −¯ µ−1 A0 (q ∗ )M,

which is exactly the value at zero of the curve in Figure 2.3c. Meanwhile, the interface shape tends to that which would occur if kinetics were neglected. Summary. We have found a family of small, nearly-faceted steady interfaces which replace the traditional planar state due to strong surface energy anisotropy. Accompanying these interfaces is a correction to the concentration profile C which is smaller still, to the point that the final effect on solid microstructure is negligible. Due to anisotropic attachment kinetics, these solutions are displaced from the isotherm occupied by the planar state. Questions of wavelength selection and stability will now be addressed in Sections 2.4 and 2.5.

2.4. Energetics. Optimal Wavelength, Comparison with Planar State In Section 2.3 we used matched asymptotic methods to find a family of nearly-faceted solutions, parametrized by their wavelength λ, that satisfy equation (2.29) describing the steady interface. However, our matched asymptotic approach told us nothing about which wavelengths λ, if any, are preferred, nor the mechanisms of such a preference. To inquire about wavelength selection, we show that, in spite of the fact that the system is not in equilibrium and is placed in a thermal gradient, the surface can still be characterized by a free energy E(λ), which we then minimize to obtain an optimal λ. In the context of faceted surfaces, this approach has been used previously by Voorhees et. al. [60], who

52

considered Eqn. (2.29) without corner energies. In addition, this approach has been used on equations essentially identical to ours in the context of elastic bars (see [63, 64, 65, 66]. Here, we repeat the important points. We begin by re-stating, for convenience, the equation describing h: (2.37)

O(λ2 ) = Geff (h + Z) − Γˆ s(θ)K + δ 2 (Kss + K3 ) + µ ¯ −1 A0 (hx ).

Now, we saw in Section 2.3.2.3 that, in the small-wavelength limit, solutions are nearly perfectly faceted, allowing us to replace the function A0 (q) with the constant A∗0 = A0 (q ∗ ), thus specifying the displacement Z. We then find that (2.38)

O(λ2 ) = Geff h − Γˆ s(θ)K + δ 2 (Kss + K3 ) =

δE , δh

where (2.39)

E=

Z

R

1 Geff h2 dx + 2

 Z  1 2 2 Γγ(θ) + δ K ds. 2 R

Thus, in the small-wavelength limit, the right-hand side of Eqn. (2.37) is the variational derivative of a free energy functional given in Eqn. (2.39). The terms in this energy represent, respectively, an effective thermal energy penalty ET due to supercooling, a surface energy Es , and a corner energy Ec . Any solution to h(x), therefore, minimizes this energy functional at least locally; the preferred solution minimizes it absolutely. Since q is everywhere near q ∗ (faceting), the anisotropic surface energy Es is nearly constant over a solution period, and large-scale solution characteristics are determined by a competition between ET and Ec . Then, since corner energy δ 2 is expected to be small,

53

while the effective thermal gradient Geff may be O(1) below the supercooling boundary, we expect an expensive supercooling penalty (projection into the melt) to be reduced by frequent, inexpensive cornering, resulting in a small wavelength. This provides an energetic rationale predicting for small wavelength which complements the geometric reasoning we used earlier. To obtain a precise value for the optimal wavelength, we calculate the average energy hEi (L) in terms of wavelength, and then minimize it. Since we expect small wavelength solutions, and since even not-too-large solutions are almost perfectly linear, we simplify our work by assuming perfectly faceted solutions with |hx | = q ∗ . The average surface energy Es is then constant everywhere except in the corners, and we need only to consider a balance between the supercooling energy on the facets, and the corner energies at the corners. For a single solution period of wavelength λ, there are two facets and two corners, giving total energies of (2.40a)

λ/4

ET = 2

Z

Ec ≈ 2

Z



−λ/4

(2.40b)

−∞

1 Geff h2 dx 2 1 2 2 δ q dx 2 x

Gq ∗ 2 3 = λ , 96 Z q∗ p 2W(q) dq = Iδ, ≈δ −q ∗

where I replaces the integral in (2.40b). The former quantity is exact and trivially calculated, while the latter is actually a small-slope approximation of the corner energy that substitutes h2xx for K2 , dx for ds, and uses result (2.32d) to set up a change of variables from x to q. Additionally, surface energy is higher in the corners, and so an exact calculation of the total surplus energy at the corners would also integrate γ(q) − γ(q ∗ ) over the corner. However, both of these neglected contributions can be seen to be O(δ),

54

and thus merely change the form of I, and not its scale. Dividing by λ to get average energies and differentiating in λ, we find that the minimum average energy satisfies 0 = Geff q ∗ 2 λ3 − 96Iδ.

(2.41)

The energy-minimizing wavelength is then (2.42)

λ=2



12Iδ Geff q ∗2

1/3

= O(δ 1/3 ) = O(¯ ν 1/6 ).

Remark 1 (Optimal Wavelength near Supercooling).

It will be noticed that this

wavelength blows up as the supercooling boundary Geff = 0 is approached. However, as Geff approaches O(λ), we see that Geff h is O(λ2 ). Recalling from Section 2.3.1 that the value of C at the interface is also O(λh) = O(λ2 ), this means that the concentration correction due to the faceted interface is no longer small enough to neglect in the GibbsThompson equation, and thus the analysis leading to Eqn. (2.42) is no longer valid. Had we specifically included the approximation C(x, 0) ≈ −χλh(x) in the Gibbs-Thompson equation, we would have found that Geff → Geff + χλ, adding a term χλ4 to equation (2.41). As Geff → 0, this new term would balance the corner energy term, giving (2.43)

λ=2



2Iδ χq ∗2

1/4

= O(δ 1/4 ) = O(¯ ν 1/8 ).

Thus, while the concentration correction can be neglected for most purposes, it does serve to limit the size of λ, protecting against blowup as the supercooling boundary is approached.

55

Remark 2 (Comparison with the Energy of the Planar State).

We also note that, in

reducing surface energy by faceting, a surface increases supercooling and corner energies. Thus, faceting in this situation will be expected only if the total energy of the optimallyfaceted state is less than the surface energy of the planar interface. The condition for this is, after some manipulation, p Geff q ∗ 2 3 λ − (E0 − Es 1 + q ∗ 2 )λ < −Iδ. 96

(2.44)

For the optimal solution wavelength just found, this translates to γ(0) − γ(q)

(2.45)

p

1+

q2

>



Geff ν¯q ∗ 2 I 2 144

1/3

.

Since ν¯ ≪ 1, this condition is met for realistic solidification environments. However, the relationship (2.44) will render some wavelengths in the family of possible wavelengths unacceptable. Summary.

Our assumption of small wavelength, argued heuristically on geometric

grounds, has been shown to be verified on the basis of energetic arguments. Sufficiently below the supercooling boundary, the optimal wavelength scales as δ 1/3 , or equivalently ν 1/6 [64, 65], where we recall that ν is the dimensional corner energy, having a scale on atomic lengths. However, we note that it remains to be seen if this optimal wavelength is, in fact, achieved. To speak to this question, we investigate the dynamics of a moving interface in Section 2.5.

56

2.5. Dynamics. Stability, Wavelength Trapping, and Coarsening Several of the remaining questions we wish to address concern dynamics; the evolution of perturbations to our solution, the ability of the interface to change its wavelength, and so on. To address these questions, we derive an evolution equation for h valid in the smallwavelength limit. This equation may be obtained by using Laplace transforms on the fully time-dependent problem (2.7), a process performed in Appendix A.1. However, for the sake of clarity here, we merely state that we can justifiably make a quasi-static assumption on the concentration field to neglect the Ct term in Eqn. (2.7a) (thus, the concentration quickly equilibrates to a slowly-moving interface [49, 67, 68]). Then, following the same analysis as is found in Section 2.3, we find a time-dependent value of C at the interface. However, upon insertion into the Gibbs-Thompson relation, it may still be neglected due to its small size. Thus, the appropriate evolution equation may be directly read from the right-hand side of the Gibbs-Thompson equation; in the small-wavelength limit, that equation is (2.46)

  δE µ−1 (q) p ht + A0 (q) + µ ¯M−1 Z = −¯ µ , 2 δh 1+q

with E defined in Eqn. (2.39).

Equation (2.46) turns out to be similar to a family of equations used to describe faceted crystal growth in other contexts [69, 7, 8, 10, 9], and, like them, is a modification of the Cahn-Hilliard equation describing phase separation. Of special importance to us in these works is the observation that initially smooth surfaces evolving under Cahn-Hilliard-type equations first rapidly decompose into faceted sawtooth surfaces, each facet of which then slowly evolves as a unit. Since numerical simulations of (2.46) reveal a similar behavior,

57

we conclude that solution dynamics may be summarized by finding the facet-velocity law which governs facet evolution [62]. To obtain this law, we first exploit the fact that small-wavelength solutions form nearperfect facets to replace instances of the variable q with the constant q ∗ as discussed above in Section 2.3.2.3.

6

Then, the Z and A0 (q ∗ ) terms cancel, and we can write

(2.47)

i δE h p , ht = − µ ¯µ(q ∗ ) 1 + q ∗ 2 δh

which describes a steepest descent of the value of the energy E. We are now in a position to follow Watson in [13], who considered faceted surfaces in three dimensions governed by equations of precisely this form. Given a variational form like (2.47), and assuming that individual facets each move as a unit on a slow timescale, he showed that facets evolve according to the relation (2.48)



dh dt



i

=−

1 ∂E , Ai ∂hi

where [dh/dt]i is the vertical velocity of the entire facet, Ai is its projected area, and dE/dhi is the rate of change of the energy E resulting from vertical facet translation. This result is independent of the form of the energy E, so we may apply it to our problem. A quick calculation reveals that, for the energy (2.39), the value of dE/dhi is Li Geff hm ,

6The

variable q terms may not, of course, be so replaced in the corners. However, by writing out the variational derivative in Eqn. (2.46) and making the space-time scaling [x, t] → [x, t]/δ, it can be seen that corner evolution satisfies an equation with a much faster time scale than that which describes facet motion. Thus, we make another quasi-static assumption that the corners remain in equilibrium with the facets that form them, and proceed to consider only facet behavior.

58

with Li the width of the facet, and hm = hhi

(2.49)

its mean height. Then, dividing by Li (instead of Ai ), we arrive at the facet-velocity law specific to our energy functional: (2.50)



dh dt



i

h i p = −Geff µ ¯µ(q ∗ ) 1 + q ∗2 hm

Remark. An argument might validly be raised at this point that the near-perfect facet assumption is only valid for static surfaces, not evolving ones. However, if we make the substitutions h → h + hm and ht → ht + V in equation (2.47), we see that the overall facet velocity is governed by the discrete equation V = −AGeff hm which, since all hm are small, has a characteristic slow time scale t/hm . The facet shape, on the other hand, is governed again by equation (2.47), or at least the outer equation (2.34), which evolves on the “regular” time scale t (also c.f. the previous footnote). Thus, a quasi-static assumption on the shape of moving facets is justifiable, and shows that moving facets have shapes which are identical to the static facet shapes found above.

2.5.1. Stability From the above discussion, we infer that small-amplitude disturbances to a periodic faceted profile will quickly settle into a disturbed faceted profile. Thus, stability may be determined simply by examining the evolution of such profiles. Equation (2.50) shows that, as long as the effective thermal gradient Geff > 0, facets are driven to the z = 0 isotherm. Since a sawtooth surface with all facets centered on z = 0 is easily seen to

59

be necessarily periodic, we see that initially non-periodic surfaces are driven to become periodic ones, which are stable. If Geff < 0, the above argument no longer holds, facets are driven away from the z = 0, and periodic surfaces are unstable. The instability criteria for such surfaces is thus precisely Geff < 0, or M−1 > 1. Remark 1.

Interestingly, the instability requirement M−1 > 1 is just the super-

cooling criteria first hypothesized by Rutter and Chalmers and Tiller et. al. for the isotropic problem, but which is only valid there in the absence of surface energy. The difference between this behavior and the classical Mullins-Sekerka instability is illustrated in Figure 2.1a where, for any bulk concentration C, a gap exists between the supercooling boundary and the Mullins-Sekerka boundary. Inside this gap, isotropic surfaces are stable, while faceted surfaces are not. Thus, for surfaces with strong surface energy anisotropy, surface energy serves only to cause faceting – its usual role as a stabilizing agent is absent. Remark 2. It will be noted that Equation (2.50), and thus the stability result just given, are only valid for small-wavelength surfaces. Thus, one might inquire about the stability of a periodic, small-wavelength faceted surface to disturbances that are large compared to the surface wavelength. Such an analysis is carried out in Appendix A.2, where it is shown that for large-wavelength disturbances, an effective homogenized surface energy does stabilize against supercooling. These disturbances thus first become unstable at pulling speeds higher than the critical speed causing supercooling. Since small-wavelength disturbances are already unstable at that point, large-wavelength disturbances need not be considered.

60

2.5.2. Below supercooling: Wavelength Selection While an energetically optimal solution wavelength was identified in Section 2.4, it is not clear if, or how, a non-optimal solution would change its wavelength to become optimal. Because surface energy keeps the surface faceted, the only way to increase wavelength is through coarsening. However, coarsening of a periodic interface requires some kind of facet motion, and a consideration of the free energy (2.39) reveals that any facet motion on a periodic surface increases supercooling energy, and is therefore prohibited under the dynamics (2.50). Energetically, smaller-than-optimal solutions are trapped in local energy wells created by the thermal gradient. (Mechanisms of decreasing wavelength, such as “facet shattering” [70] or “tip splitting” [71, 48], are not considered here.) However, knowing that any one of a range of wavelengths is stable, we may still conclude that the optimal wavelength may never be reached, and we are encouraged in this direction by the fact that elastic materials modeled with similar equations [63, 64, 65, 66] are experimentally observed to exhibit hysteresis in their equilibrium-pattern wavelength [72].

2.5.3. Above supercooling: Coarsening Finally, when the solidification speed is increased beyond the constitutional supercooling boundary, the sign of Geff changes, and the above-mentioned energy wells are replaced by energy hills. Now, facet motion away from the z = 0 isotherm actually decreases supercooling energy (since the thermal gradient is negative). As facets accelerate away from z = 0, the surface coarsens as boundaries between facets meet and annihilate. An example of this behavior is given in Figure 2.5, where it is also contrasted with belowsupercooling behavior. The coarsening process continues until the typical wavelength

61

is no longer small. At such point, the facet-dynamics model loses its validity, and the full free-boundary problem must be considered. Consequently, the wavelength of a final steady or unsteady state is selected not by near-instability analysis of competing cellular modes, but by nonlinear dynamic interactions between the fully-faceted surface and its associated concentration field. (See, for example, [48] for several facet-dynamics models proposed to describe this late-time regime.) 8

(a)

7 6

t

5 4 3 2 1 1

1.2

1.4

1.6 x

1.8

2

2.2

(b)

7 6

t

5 4 3 2 1 16

17

18

19

20

21

22

23

x

Figure 2.5. Representative solution behavior below and above supercooling. The evolution of corners is shown; peaks are red, and valleys are blue. (a) Below the supercooling velocity, coarsening is prohibited, and non-periodic interfaces are driven toward periodicity. (b) However, above the supercooling velocity, coarsening replaces cellular growth as the mechanism of instability.

62

2.6. Conclusions and Comments When materials with large anisotropy are solidified in high-energy orientations, negative surface stiffness renders the planar interface unstable for all solidification environments. Instead, the interface assumes one of a family of small, faceted profiles. Consideration of attachment kinetics reveals a displacement of the faceted interface relative to the planar one during solidification, whereas an associated correction to the concentration field results in a very slight variation in the composition of the final product. While geometric considerations predict a small interface wavelength, the matched asymptotic methods used to derive interface shapes do not reveal a wavelength selection mechanism. Instead, the minimization of an appropriate surface free energy reveals that the optimal solution wavelength scales as ν 1/6 [64, 65], where ν is a very small corner energy parameter. Questions of dynamics lead, in the small-wavelength limit, to the derivation of a facet velocity law, which specifies the vertical velocity of each facet as a function of its mean height. This approach result shows that (a) the stability boundary for faceted interfaces is precisely the M−1 = 1 supercooling boundary; (b) wavelength change to reach the optimal wavelength is inhibited below supercooling, leading to the prevalence of varied, non-optimal solutions; and (c) above supercooling, coarsening replaces the usual cellular growth as the mechanism of instability. Perhaps the most interesting feature of the contrast between small and large anisotropy is that the role of surface energy changes fundamentally. For isotropy and small anisotropy, surface energy stabilizes against an instability driven by constitutional supercooling. The presence of anisotropy in this regime simply quantitatively modifies a pre-existing morphological instability. However, for large anisotropy, surface energy becomes destabilizing,

63

and drives its own, thermodynamic faceting instability before supercooling is reached. In addition, the sole effect of surface energy is to drive the interface toward its optimal slope via faceting. In the absence of its further stabilizing influence, the stability of the faceted surface depends solely on the presence or absence of supercooling; interestingly, this is exactly the qualitative instability criteria originally hypothesized by Rutter and Chalmers and Tiller et. al. for the isotropic problem. Thus, where surface energy and supercooling are competing effects for small or zero anisotropy, they are divorced for large anisotropy. An additional interesting qualitative change is the mechanism by which solutions evolve after becoming unstable. For materials with small or zero anisotropy, the latestage behavior of linearly-identified instabilities are well-described by the usual weakly nonlinear cellular solutions, where a small band of admissible solution wavelengths is apparent from onset, and destabilization to nearby wavelengths within that band can be considered. In the large-anisotropy regime, by contrast, destabilization of periodic faceted interfaces occurs when supercooling simply causes the interface to begin coarsening under the appropriate facet dynamics. Here, no intrinsic wavelength is apparent at onset, and the final wavelength will be selected only eventually, through nonlinear interactions between interface shape and concentration profile. While this work was motivated by the addition of the particular effect of large anisotropy to the long-standing problem of directional solidification, its broader significance is best seen by viewing it as a sample study of large-anisotropy surfaces in dynamic contexts. In such systems, faceting is the generic outcome regardless of other environmental conditions. This renders traditional analytical methods of limited value – linear stability

64

analysis simply returns the expected universal instability of the planar state, while further destabilization occurs through coarsening rather than cellular instability. In the place of these traditional methods, it was the derivation of the facet velocity law in Section 2.5 that allows real advancement of understanding. This approach would be of use in general in the study of faceting interfaces.

65

CHAPTER 3

Large-Scale Simulations of Coarsening Faceted Surfaces 3.1. Introduction When a crystalline material is cut along an energetically unfavorable direction and then allowed to evolve by some mechanism, surface faceting may result. In this phenomenon, an initially flat surface will decompose into a faceted, pyramidal hill-and-valley configuration. Faceting is caused by the strong crystalline anisotropy of surface energy – a faceted interface, while exposing more surface area than a flat one, may have a lower surface energy if the facets have low-energy orientations. As a faceted surface continues to evolve, it may also exhibit coarsening, whereby small facets continually vanish, and the average length scale L of those that remain increases with time. Of primary interest is whether such systems exhibit dynamic scaling, whereby the surface approaches a constant statistical state which is preserved even as the length scale increases. This intriguing phenomenon, observed in many coarsening systems, is studied because the system may be described at all stages of evolution by a single set of statistical distributions. The detailed statistical study of coarsening and dynamic scaling requires a method of rapidly simulating large faceted surfaces, which may be developed as follows. Assuming the orientation of each facet is prescribed and fixed, then individual facet motion is constrained to translation along its normal. Therefore, the evolution of a completely faceted surface is concisely expressed by a discrete collection of individual normal facet velocities

66

(nucleation of new facets is not treated here). If the facet velocity law providing this collection at each time can be determined, then the computational complexity of evolving the surface can be reduced to that of a system of ordinary differential equations; an evolving surface which possesses such a known law is termed a Piecewise-Affine Dynamic Surface (PADS) [13]. Such PADS are, in fact, known. In what are now known as van der Drift models [73, 74], the facets of diamond grown under vapor deposition advance according to a fixed velocity which depends only on orientation. Other configurational rules have more recently been proposed or derived for systems as varied as the evolving faceted interface between two elastic solids [75], the thermal annealing of a faceted crystal with its melt [13], and several models of solidification [71, 48, 76, 62]. Because of the variety of systems described by facet velocity laws, and the need for large simulations to investigate coarsening and dynamic scaling phenomena, there is a strong incentive for the development of a computational geometry tool to investigate evolving faceted surfaces. Several such geometric methods have been considered in the past. For 1+1D surfaces z = h(x, t) evolving in time t, many examples exist. Pfeiffer et. al. [71] and Shangguan and Hunt [48] included, to our knowledge, the first such simulations in their proposals of facet dynamics describing the solidification of pure silicon and binary alloys, respectively. Later, Wild et. al. [77], Dammers and Radelaar [78], and Paritosh et. al. [79] all used the same approach to study the evolution of diamond films under the previouslyknown van der Drift evolution. Additionally, what are essentially 1+1D geometric surface simulations are found in two simulations of the convective Cahn-Hilliard equation [76, 62] – these authors actually develop explicit expressions for corner evolution. Finally, whereas each of the above simulations were specifically implemented for the particular dynamics

67

being studied, Zhang and Adams [80, 81] have recently released a more general software package which allows the selection of a variety of facet behaviors. While their simplicity makes them efficient, the primary complication of direct geometric methods is the need to manually detect and resolve topological events. These changes in the neighbor relations between facets occur when, as the surface evolves, facets merge, split, or vanish. Trivial in 1+1D, difficulty associated with topology increases with the number of dimensions considered. For 2+1D surfaces z = h(x, y, t), the only known geometric simulations of faceted surfaces are due to Thijssen [14] and Barrat et. al. [15], who studied the law for diamond films; a similar method was also applied to spiral-mode growth of thin films [82]. These authors did not explicitly address topology, instead allowing diamond grains to interpenetrate, and describing the actual surface as the envelope of these grains. Indeed, the increased topological difficulty associated with high-dimensional geometric methods is an oft-cited motivation for the development of “topology free” methods such as phase-field [83, 84] and level-set [16, 17] methods. However, these methods sacrifice speed and ease of access to geometrical data. Furthermore, the appropriate resolution of potentially non-unique events as described in [14] requires explicit intervention, which is made difficult by topology free methods, and negates much of their benefit. In light of these concerns, we have chosen an explicit resolution approach, and we find that, for certain symmetries at least, the topological complexity has been overstated. In particular, for 2+1D surfaces possessing only three facet normals (threefold symmetry), only three kinds of topological event are possible, each of which recalls similar events observed in the related 2D work of Roosen and Taylor et. al. [70, 85, 86]1. With 1The

purely 2D work of these authors represents an important intermediate case between 1+1D and 2+1D. Their method captured the kinematics of evolving completely faceted crystal domains in the

68

the abundance of systems for which facet-velocity rules are known, and the speed and accessibility advantages offered by geometric methods, the challenges posed by topology are worth tackling. The aim here, then, is to present a general-purpose geometric method, which implements topological events, for the simulation of coarsening, threefold-symmetric faceted surfaces in 2+1 dimensions. We shall begin with a description of the faceted surface and associated data structure, and a discussion of the kinematics and dynamics which govern its evolution. We then present illustrations and discussions of each class of topological event, including detection and resolution procedures. Finally, we demonstrate our method by simulating a total of one million facets under a sample dynamics describing thermal annealing. The efficiency of this method allows such large collections of facets to be simulated rapidly, while the geometrical nature of the network allows the easy collection of a rich variety of statistical data.

3.2. Faceted Surfaces: Description, Kinematics, and Dynamics In this Section we present the basic elements of our method. We first give a mathematical description of completely faceted surfaces, and present a three-component structure used to simulate them. We then discuss the kinematics of such surfaces; i.e., how facets move and how their motion drives the evolution of other surface elements. Last, we consider the imposition of a dynamics on the system, and we apply a sample dynamics associated with thermal annealing. plane, and was used to simulate dynamics associated with growth due to diffusion fields, attachment kinetics, and surface diffusion. Of especial importance to us is that topological events were handled explicitly, and indeed, each of the three events observed there exhibits aspects of a corresponding event considered here.

69

3.2.1. Description We consider evolving fully-faceted surfaces z = h(x, y, t) consisting of planar facets {Fi} with prescribed normals {ni }. We consider surfaces formed by a single crystal with cubic symmetry, and restrict our attention facets possessing one of three orthogonal normals, given in spherical polar co-ordinates by (3.1)

ni ∈ (1, 2πi/3, α) ,

√ i = {0, 1, 2} , α = sin−1 (1/ 2)

These normals represent the ([100],[010],[001]) orientations of a cubic crystal viewed from the [111] direction (with a different choice of α, different symmetric orientations could be modeled). The facets F are bounded by and meet at straight edges, which in turn which in turn meet at triple-junctions (while any number of edges greater than three could theoretically meet at a junction, we assume that all junctions are formed by exactly three edges). This facet-edge-junction structure is reminiscent of two-dimensional cellular networks [87, 88, 89, 90, 91]. Several other purely two-dimensional coarsening physical systems – such as soap froths [92, 93] and polycrystalline grains [94, 95] – are also cellular in nature, and have been effectively simulated using three-component models.

3.2.2. Kinematics Since the orientation of each facet Fi is fixed and constrained by (3.1), its motion is completely described by a displacement in the normal direction, parametrized by a local distance parameter. The surface kinematics Vn are thereby captured by specifying the instantaneous normal velocity Vi of each facet. The motion of edges and junctions, being

70

merely the intersections between the two and three facets that comprise them, are then uniquely determined by the motion of those facets. In practice, we only use facet velocities indirectly, as a means to calculate junction velocities. When junctions are moved correctly, edges (connections between two junctions) and thus facets (collections of edges) are necessarily moved correctly as well.

3.2.3. Dynamics With the kinematics of fully-faceted surfaces now set, we may consider imposing a surface dynamics associated with some physical problem. To do so, we choose a facet-velocity law which specifies Vi , yielding a piecewise-affine dynamic surface (PADS) [13]. (To connect with mathematical theory, we note that this step amounts to imposing a vector field on the manifold of fully-faceted surfaces just defined.) For concreteness, we consider a PADS associated with the annealing of a faceted surface. It has recently been shown [13] that, because the equations describing this system are variational in nature, a principle of maximal dissipation may be applied, which shows that the surface evolves so as to always maximally reduce its energy, which to leading order is stored in the edges between facets. This approach allows the matched asymptotic extraction of the facet velocity law, which is expressed by the equation (3.2)

Vi = −

1 ∂P . A(Fi ) ∂ni

Here A(Fi ) represents the area of facet Fi , while P per unit displacement of Fi along its normal.

∂P ∂ni

is the rate of change in total perimeter

71

3.3. Topological Events In presenting our data structure in Section 3.2.1, we described a set of neighbor relations inherent to the surface. Surface elements of each class (facet, edge, junction) neighbor members from each of the other classes. Together, this set of neighbor relations comprises the topological state of the surface. As the surface evolves under a prescribed dynamics (see Sec. 3.2.3), these relations may change as facets merge, split, or vanish. Such changes to the topological state of the surface are called topological events. On an actual faceted surface, these happen naturally as the system evolves; however, in an approach like ours they must be performed manually. Each event is found by looking for an appropriate “trigger” condition on the surface; once detected, the proper resolution follows from geometrical considerations. In this Section, we consider the occurrence, detection and resolution of topological events, limiting our attention to those occurring under the symmetry and dynamics already described.

3.3.1. Facet Merge When an edge shrinks to zero length on an evolving surface with cubic symmetry, two facets of like orientation meet, and merge to form a larger facet. This event is called a Facet Merge. To see why this occurs, consider that each edge is composed of two facets of which it is the intersection (its composite facets), and stretches between two facets at which it terminates (its terminal facets). Because we only allow three distinct facet orientations, the terminal facets of any edge necessarily have the same orientation. Thus, when an edge shrinks to zero length in isolation (i.e. none of its neighbor facets are vanishing), its terminal facets meet exactly. Having the same orientation, they merge into

72

a single facet, and the total number of facets on the surface is reduced by one. Thus, the facet merge is one possible mechanism of coarsening for a PADS with cubic symmetry. Figure 3.1 depicts a representative facet merge.

Figure 3.1. Example facet merge event, viewed from above. Arrows represent gradients of the facets on which they appear. Dotted lines indicate past edges no longer present.

Detection and Resolution. Edges in the data structure are directed, having an initial and terminal junction. Since facets, and hence edges, have fixed orientation, each edge thus has a unique orientation. The only way a tangent may change is to reverse direction when an edge shrinks to zero, as just described. If such an “edge flip” occurs, it indicate that a facet merge ought to have occurred during the preceding timestep. This serves as our “trigger” condition. Having found the flipped edge, the like-oriented facets are first adjusted to equal height in preparation for merging. Further resolution is then essentially an exercise in labeling. All edges and faces which touch the shrinking edge are first identified; then surface elements are created/deleted, and neighbor relations reassigned as appropriate to effect the change shown in Figure 3.1. Complete numerical details will be published elsewhere.

73

3.3.2. Merging Facet Pinch In addition to merging together, facets may also be split apart if they are pinched by non-adjacent neighbors. Under cubic symmetry, a facet may only be so pinched by two of its neighbors with identical orientations. While the pinched facet splits in two, the impinging neighbors (having the same orientation) merge to form a larger facet; the total number of facets is thus conserved. This event is called a Merging Facet Pinch. It is not a coarsening event, but rather a re-organization which allows further coarsening to occur. In Figure 3.2, we see an example of this event.

Figure 3.2. Example merging facet pinch event, viewed from above. Arrows represent normals of the facets on which they appear. Dotted lines represent past edges no longer present.

Detection and Resolution. To describe the trigger which indicates a merging facet pinch, we note that the edges forming the boundary of a facet form a polygon in the plane, A merging facet pinch is detected when, after a timestep, this polygon is found to be self-intersecting. Such a polygon represents a geometric inconsistency, as surface facets are necessarily simply-connected. This indicates that the facet in question was pinched during the previous timestep, and a merging facet pinch should have occurred. Having identified the impinging facets, we first ensure that they are of equal height to

74

allow proper merging. Then, as in the facet merge, further resolution of the merging facet pinch is mostly an exercise in element labeling. Using neighbor relations inherent to the network, we can collect all the affected edges and junctions. Elements are then created and deleted as necessary to perform the change illustrated in Figure 3.2. Again, numerical details will be presented elsewhere.

3.3.3. Removal of Vanishing Facets The final class of topological events involves one or more contiguous facets shrinking to zero area. When this occurs, these facets must be removed, and the surrounding neighbor facets re-connected appropriately. The total number of facets clearly decreases, making these events an additional mechanism of coarsening. A variety of these events are possible kinematically; however, after many tests under both dynamics with random initial conditions, we observe only three kinds: a “step removal,” a “ridge removal,” and a “box removal.” We list these in Figure 3.3 for illustrative purposes. The appearance and relative incidence of each configuration is a consequence of the dynamics. Detection. In practice, we do not actually wait for a zero area facet to occur. Instead, we seek to eliminate small facets whenever their area decreases below a certain threshold. The gathering of two- and three-facet groups is accomplished by maintaining a second, more liberal threshold. Whenever a facet is found to have decreased below the first threshold, its neighbors are recursively examined to find those smaller than the second threshold. This procedure is not foolproof, but since the area of vanishing facets tends to zero, all facet areas must lie beneath the second threshold for at least some finite time

75

(a)

(b)

(c)

Figure 3.3. Three kinds of vanishing facet events. (a) One facet vanishing alone – a “step removal.” (b) Two facets vanishing simultaneously – a “ridge removal.” (c) Three facets vanishing simultaneously – a “box removal.” Arrows represent normals of facets on which they appear.

before the event. Therefore, if the procedure applied to a flagged facet does not yield a recognized configuration, we abort and do nothing, to try again during a future timestep. Resolution. After deleting the small facet or collection of facets, a “hole” is left in the network. This must be repaired by reconnecting the surrounding facets, which we call the “far field.” For the step removal, height averaging is again necessary before merging the neighboring large facets together. For the ridge removal, the far field is O(4), and

76

the reconnection structure is a set of two points and a line, which must terminate at the two (necessarily) identically-oriented far field faces. Finally, for the cube removal, the far field is O(3), and a single point must result. In each case, the location of created points is easily calculated from the positions of the surrounding facets.

3.4. Demonstration We now apply our method to the annealing dynamics described above in Equation (3.2). We begin in Figure 3.4 with a sequence of images from a relatively small test run. There the reader may locate regions of the surface near to each of the topological events described above. In the subsections which follow, we then present statistical data averaged from 40 runs of 25,000 facets each. We first summarize the coarsening behavior of annealing surfaces, including mechanism, power law, and convergence toward the dynamically scaling “steady state”. Next, we describe aspects of that state through some easily-gathered morphometric data describing distributions of relative geometric quantities. Finally, we consider some topological and correlational statistics, which illustrate, respectively, some properties of facets based on number of sides, and the degree to which neighboring facets have similar geometric properties.

3.4.1. Rates and Mechanisms of coarsening, Convergence to dynamic scaling We begin our statistical data with results concerning the rate and mechanism of coarsening, as well as the convergence to the scale invariant state (SIS – actual SIS data are found below). First, coarsening is achieved primarily through the step-removal mechanism shown in Figure 3.3a. As the system approaches the SIS, it exhibits the power-law

77

Figure 3.4. A top-down view of a small coarsening faceted surface. Spatial scale is constant, but irrelevant; time increases down the column. coarsening seen in many dynamically scaling systems, with the characteristic morphological length scale LM growing as t 1/3 . This has been observed in the past, and has

78

recently been explained by observing that the dynamics (3.2) are invariant under the scaling x → αx, t → α3 t [13]. Finally, we observe that the system converges to a scale-invariant state (SIS) at a rate of E ∼ t−2/3 , with E the 2-norm of the difference between any given statistical distribution and its scale-invariant form. However, the mechanism by which convergence to scale invariance occurs is coarsening, and so the convergence rate implicitly depends on the coarsening rate. For this reason, we favor expressing E in terms of the fraction of remaining facets N /N0 rather than time. This defines a function (3.3)

N E(N ) ∼ N0 

p

for



 N ≪1 N0

which describes the coarsening efficiency of the dynamics being studied, where the efficiency exponent p reflects the attractive strength of the scale-invariant state. This representation allows the transparent comparison of coarsening phenomena in systems with different coarsening rates; in particular, systems with p > 1 are expected to verifiably achieve scale-invariance before running out of facets, while systems with p < 1 are not. The dynamics studied here exhibit the intermediate value p = 1.

Coarsening Mechanism Step removal Coarsening Rate t 1/3 Convergence to SIS t−2/3 Efficiency Exponent p=1 Table 3.1. Relevant coarsening phenomena.

79

3.4.2. Some 1D distributions In Figure 3.5 we present a series of distributions of scaled, dimensionless geometric quantities. That is, for a dimensional quantity q, we present a distribution of the dimensionless q/ hqi, where hqi denotes the system-wide average of that quantity. For a system in a state of dynamic scaling, all such distributions are constant in time. In the following figures, data have been time-averaged over the dynamic scaling regime to minimize noise.

0.8

0.5

(b)

(a)

0.45

0.7

0.4 0.6

number of facets

0.35 0.5 0.3

ρ

0.25 0.2

0.4

0.3

0.15 0.2 0.1 0.1 0.05 0 0

2

4

6

8

10

12

14

16

18

0 0

20

0.5

1

1.5

2

number of sides

2.5

3

3.5

4

4.5

5

L /

0.8

1

(c)

(d)

0.9

0.7 0.8 0.6 0.7 0.5 0.6

ρ

ρ

0.4

0.5 0.4

0.3 0.3 0.2 0.2 0.1 0.1 0 0

0.5

1

1.5

2

2.5

A /

3

3.5

4

4.5

5

0 0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

P /



Figure 3.5. (a) Distribution of facets by number of sides n. (b) Distribution of dimensionless edge lengths ρ(L/L), with contributions from concave and convex edges. (c) Distribution of dimensionless facet areas ρ(A/A), with contributions from 2n-sided facets. (d) Distribution of dimensionless facet perimeters ρ(P/P), with contributions from 2n-sided facets.

80

Figure 3.5a shows the probability of a facet having a given number of sides. Under cubic symmetry, all facets under threefold symmetry have an even number of edges (because only three facet orientations are available, and facets of like orientation cannot touch, the neighbors of a facet with one orientation must facet alternate between the other two orientations). In Figure 3.5b we show the distribution of edge lengths. The total distribution is in dotted black, while the solid blue and green lines show the contributions from convex and concave edges, respectively. The equality of these reflects the underlying up-down symmetry of the dynamics. Finally, in Figures 3.5c,d we display distributions of facet area and perimeter, respectively, relative to their global averages. These are broken down into contributions from 2n-sided facets, illustrating the level of detail that may be extracted using our method.

3.4.3. Topological results and neighbor relations In our last set of data, we consider two scale-invariant topological properties, which describe average geometrical quantities as functions of the number of sides; and two correlational properties, which are two-point distributions associated with neighbor pairs. Figure 3.6a shows that average facet area grows linearly with the number of sides per cell, a relationship known as Lewis’s law [96]. In Figure 3.6b, we show that the average number of sides of the neighbors of n-sided cells, mn , obeys Aboav’s law [97, 98, 99]: (3.4)

mn = (6 − a) +

6a + µ2 , n

with µ2 the second moment about the mean of the distribution of sides per cell, and a a fitting parameter (we find µ2 = 7.07 and a = 5.345). These relationships are commonly

81

8

3.5

(b)

(a) 3

7.5

2.5

A /


7

mn 2 6.5 1.5

6 1

5.5

0.5 4

6

8

10

12

14

4

16

6

8

10

number of sides

(c)

14

16

18

(d)

4.5

4

4

3.5

3.5

3

3

P /



A /


4.5

12

number of sides

2.5

2.5

2

2

1.5

1.5

1

1 0.5

0.5 0.5

1

1.5

2

2.5

3

3.5

4

4.5

0.5

1

1.5

2

A /


2.5

3

3.5

4

4.5

P /



Figure 3.6. (a) Lewis’s law, showing the average area of 2n-sided facets. Also shown is the average perimeter of the same. (b) The Aboav-Weaire law, showing the average number of sides of the neighbors of 2n − sided facets. (c) A two-variable distribution of the relative areas of neighboring facets. (d) A two-variable distribution of the relative perimeters of neighboring facets. observed in evolving 2D cellular network problems such as soap froth evolution and grain growth. Finally, in Figures 3.6c,d, we give a pair of distributions measuring the probability of two neighboring facets having a given pair of areas and perimeters, respectively. These data, together with the data in Figures 3.5c,d, can be used to determine whether the behaviors of neighboring facets are correlated. Understanding the extent of such correlations will, in turn, inform the future pursuit of mean-field theories.

82

3.5. Conclusions We have presented a computational geometry tool for the simulation of coarsening faceted surfaces in 2+1 dimensions. Such surfaces are expressed geometrically as a 3D cellular network consisting of facets, edges, junctions, and the connections between them. Kinematic relationships between facet displacement and junction/edge motion have been discussed, and an example dynamics has been imposed, resulting in a Piecewise-Affine Dynamic Surface, or PADS. We considered a faceted surface with cubic symmetry, corresponding to the growth of a cubic crystal in the [111] direction. For this symmetry group, we identified and discussed the three classes of topological events: • a facet merge in which two facets merge to form a larger facet • a merging facet pinch in which one facet splits in two, and two others merge • a facet removal in which one of more vanishing facets are removed. The detection and implementation of these topological events is the main contribution of our tool. The primary benefits of our approach are its speed and easy access to a variety of geometrical data, which are both highlighted through our demonstration on the facet dynamics (3.2) associated with thermal annealing. Because the method is intrinsically geometrical, we can easily extract statistics describing distributions of geometric quantities, as well as those describing correlation among quantities and neighbors. On the other hand, because the method efficiently handles topology, we can quickly measure not only the dynamically scaling state itself, but also the convergence toward that state, as described by the introduced quantity of coarsening efficiency.

83

It is our hope that this tool, generalized to arbitrary symmetry groups, will be widely useful to anyone investigating fully-faceted surface evolution. The speed and design of this tool will allow the rapid simulation of large faceted surfaces, and the consequent collection of geometric statistics suitable for the comparison of different physical causes of coarsening. Finally, neighbor relations inherent in the data structure will allow the search for correlations in high-order statistics, the presence or absence of which should help to inform the future pursuit of mean-field theories for coarsening faceted surfaces.

84

CHAPTER 4

The Kinematics of Faceted Surfaces with Arbitary Symmetry 4.1. Introduction In many crystal-growing procedures of interest, a nano-scale faceted surface appears and proceeds to evolve, often exhibiting coarsening and even dynamic scaling, whereby characteristic statistics describing the surface remain constant even as the characteristic lengthscale increases through the vanishing of small facets. For many evolving faceted surfaces, a facet velocity law can be observed [74, 73], assumed [71, 48], or derived [75, 62, 13] which specifies the normal velocity of each facet, often in configurational form which depends on the geometry of the facet. In this way, the dynamics of a continuous, two-dimensional surface can be concisely represented by a discrete collection of such velocities, and overall computational complexity reduced to that of a system of ODE’s; the resulting system is known as a Piecewise-Affine Dynamic Surface, or PADS. Such theoretical simplification, in turn, enables the large-scale numerical simulations necessary for the statistical investigation of coarsening and dynamic scaling. The numerics involved in the direct geometric simulation of an arbitrary PADS is straightforward for one-dimensional surfaces, requiring nothing beyond traditional ODE

85

techniques except simple geometric translation between facet displacement and edge displacement, and a small surface correction associated with each coarsening event. Consequently, such simulations accompany many of the above facet treatments of facet dynamics, and have also been independently repeated elsewhere [77, 100, 79, 80, 81]. However, in two dimensions, the corrections due to coarsening events are much more involved, and any code must be able to deal with a family of non-coarsening topological events that alter the neighbor relations between nearby facets. Consequently, the fewer simulation attempts use either fast but poentially imprecise envelope methods [14, 15, 82], or more robust but slower phase-field [83, 84] or level-set [16, 17] methods to avoid explicitly performing topological changes. Besides the speed/accuracy trade-off exhibited by these approaches, both methods obscure the natural geometric simplicity of the native surface, complicating the extraction of detailed surface statistics which, after all, motivates large simulations in the first place. Additionally, as will be seen, the presence of non-unique topological events requires explicit intervention regardless of topological scheme, which negates much of the advantage of a “hands-free” treatment. In the previous chapter, we introduced a direct-simulation method which explicitly performs topological events along the way, thus preserving both simulation speed and topological accuracy. In addition, by representing the surface as a collection of facets, edges, and junctions, plus the neighbor relations between them, the method mirrors the natural geometry of the surface being modeled, which allows easy extraction of geometric statistics. There, however, the restricted case of threefold symmetry was chosen for ease of topological implementation; under this symmetry, a limited number of topological events were observed, and both vanishing facets and non-vanishing surface rearrangements could

86

be handled explicitly using a priori knowledge of the before and after surface states. While many surfaces exhibit threefold symmetry, making the method useful even in this special case, it could not handle other common crystal symmetries, notably fourfold and sixfold. In this chapter, then, we generalize the previous model to allow the simulation of surfaces with arbitrary symmetry groups. We begin in Section 4.2 with a brief summary of the basic method, including surface representation, facet kinematics, and the application of a dynamics. Next, in Section 4.3, we provide a careful enumeration of topological events which may occur on surfaces of arbitrary symmetry; this includes discussion of the Far-Field Reconnection algorithm, by which network holes left by vanishing facets may be consistently repaired without knowledge of the post-event state. Then, we provide in Section 4.4 a careful consideration of the consequences of using (necessarily discrete) timesteps during the simulation of a surface whose evolution equations change qualitatively between steps (at topological events); the issues that arise are discussed in the context of three sample strategies. The completed method is illustrated from three-, four, and six-fold symmetric surfaces in Section 4.5; these exhibit all of the topological events likely to be encountered on a real surface, and demonstrate that the method is robust enough to generically simulate faceted surfaces of any symmetry class for which a facetvelocity law is uniquely specified. Finally, in addition to detailing the FFR algorithm, the appendix includes a discussion of kinematically non-unique topological events, where two resolutions are possible, and highlights the need to refer to the dynamics or even first principles to decide how the surface should evolve in those cases.

87

4.2. Data structures and simple motion: a 3D cellular network 4.2.1. Characterization

We consider the evolution of a single-valued, fully-faceted surface z = h(x, y, t); this definition explicitly forbids overhangs and inclusions. We assume that the surface bounds a single crystal which exists on exactly one lattice; thus, we are not treating surfaces with multiple grains. The surface is piecewise-affine, consisting of facets {Fi } with fixed normals {ni }. These are bounded by and meet at edges E which are necessarily straight line segments; edges in turn meet at triple-junctions J . This three-component structure is reminiscent of two-dimensional cellular networks [101, 87, 89, 90, 91] and indeed, while we consider three dimensional surfaces, the projection of the edge set onto the plane z = 0 is a 2D cellular network. This structure and the neighbor relations inherent within it suggest a doubly-linked object-oriented data structure, consisting of: (1) a set of junctions, each having a location, pointing to three edges and three facets; (2) a set of edges, each having a tangent, pointing to two junctions and two facets; and (3) a set of facets, each having a normal, pointing to m edges and m junctions. These objects and the associated neighbor relations are illustrated in Figure 4.2.1; this structure is the natural structure of the surface, and uniquely and exactly describes it. We now consider each element in more detail. 4.2.1.1. Junctions. A junction is a point in space formed where edges (and hence, facets) intersect. The order n of a junction is simply the number of edges which meet there. While junctions of any order n ≥ 3 are possible, we restrict ourselves here to the case of order 3 junctions or “triple junctions.” This greatly simplifies analysis and

88

(a)

(b)

(c)

Figure 4.1. Neighbor relations for each kind of surface element. (a) A junction neighbors three edges and three faces. (b) An edge neighbors two junctions and two faces. (c) A face neighbors m junctions and m edges (here m = 5).

code, as triple junctions are uniquely positioned by the three facets meeting there. The intrinsic geometric information carried by a junction is its location. Junctions are stored in a Junction class, which contains this location, as well as pointers to the three edges and three facets which meet there. 4.2.1.2. Edges. An edge is a line segment formed by the intersection of exactly two facets, and bounded by exactly two junctions. The intrinsic geometrical quantity of an edge is its orientation, which is fixed since facets have fixed normals. Edges are stored in the Edge class, which records the tangent, as well as pointers to the two neighboring facets and two bounding junctions. At creation, edges are “directed”: one junction is arbitrarily deemed the origin, and the other the terminus, establishing a tangent. This has two important consequences. First, if we imagine walking along the edge in the tangent direction, then one neighboring facet may be labeled “left”, and the other “right.” This information allows us to distinguish between convex and concave edges, and also to determine the clockwise direction around a given facet, which is necessary for effective navigation of the network, as well as the

89

proper calculation of boundary integrals on facets. Second, the tangent allows us to detect when an edge “flips” (see [70]); this will be discussed in more detail in Section 4.3.1. 4.2.1.3. Faces. A facet is a simply-connected planar polygonal region in space, which is bounded by an equal number of edges and junctions. The intrinsic geometric information carried by a facet is its normal, which is fixed. Our surface definition z = h(x, t) requires that the normal of each facet is constrained to be on the hemispherical shell of unitlength vectors with positive z component. The imposition of a particular symmetry on the crystal may further restrict available normals, but no such restriction is here assumed. Facets are stored in a Facet class, which contains the normal, as well as a list of bounding edges and junctions, sorted in counter-clockwise order.

4.2.2. Kinematics The intrinsic geometric means of characterizing surface evolution is by specifying the normal velocity of each point on the surface. A piecewise-affine surface is composed of a collection of planar, fixed-normal facets, whose motion is limited to displacement along the normal. Therefore, the kinematics Vn of the entire surface may be expressed by a discrete set of individual facet velocities Vi . As edges and junctions are merely intersections between two and three facets, respectively, their motion is uniquely specified by the motion of the facets that neighbor them. In particular, if p is the location in space of a triple junction, then the velocity of that a triple junction may be calculated through the expression (4.1)

dp = A−1 v, dt

90

where the rows of A and entries of v are the unit normals and normal velocity, respectively, of the three facets intersecting to form p. In practice, facet velocities are only used indirectly to calculate junction velocities – if junctions are moved correctly, edges (connections between two junctions) and thus faces (collections of edges) are necessarily moved correctly as well.

4.2.3. Dynamics All that remains now is to select a particular dynamics; that is, to specify an expression for the normal velocity Vi of each facet. Having chosen one, we follow [13] and refer to the resulting evolving structure as a Piecewise-Affine Dynamic Surface (PADS). Example dynamics describing many different physical situations were listed in the introduction, and the exact dynamics is not of special concern here (although we will select one for demonstration later). It is worth noting here, however, that most of the dynamics proposed to date are configurational, depending on properties of the facet such as area, perimeter, number of junctions, or mean height. Thus, sudden changes in the geometric properties of a facet can lead to sudden changes in its velocity, an issue which will be explored in more detail in Section 4.4.

4.3. Topological Events We have just discussed how elements of each class (facet, edge, vertex) neighbor members from each of the other classes. Taken together, the set of all of these neighbor relations comprises the topological state of the surface. It is a complete record of every neighbor relationship on the surface, and is unique for a given surface. As the system

91

evolves, these neighbor-relations may change as facets exchange neighbors, join together, split apart, or vanish. Each of these cases is an example of a topological event, and represents a change to the topological state of the surface (topological events are a defining feature of evolving cellular networks – again see [101, 87, 89, 90, 91]). To maintain an accurate representation of the surface, a direct geometric method like that described here must manually perform topological events as necessary. Because actual surface evolution is fairly trivial, this is the main difficulty of our method. A natural first question to ask at this point is “how many topological events are possible?” To begin answering this question, we point out that on a physical surface, topological events occur automatically, and by geometric necessity. If a detected event signals the need to change neighbor relationships at some location on the surface, we may therefore infer that failing to change them would produce a cellular network with “wrong” relationships, that do not correspond to a physical surface. We call such erroneous configurations geometrically inconsistent; examples include primarily edge networks that intersect when viewed from above, since these correspond to overhangs and inclusions, which are prohibited. Since topological events serve to avoid possible geometric inconsistencies, we may discover what events are possible by considering how inconsistencies may occur. This is most easily accomplished by considering each surface element in turn. We first consider junctions, which are simply a location in space. A junction can, in the course of surface evolution, leave the periodic domain, in which case it is wrapped to the other side. However, this is only a bookkeeping operation, and does not represent a real topological event. Turning to edges, we note that edges possess a directed length. As already hinted in section 4.2.1.2, this length could become negative if the edge were to

92

“flip” [70] A flipped edge has no geometrical meaning on a single-valued surface, and so we introduce a class of Vanishing Edge events which occur when edges reach zero length. Finally, we consider facets. Since a facet has fixed orientation, its changing properties are loosely its shape and size. Specifically, a facet is a simply connected planar region with positive area. These two defining properties of facets lead, through consideration of their potential violation, to two additional classes of topological event: Facet Constriction events which prevent the formation of self-intersecting facets, and Vanishing Facet events which remove facets from the network when they reach zero area.

4.3.1. Vanishing Edges An adjacent point-point event occurs when an edge shrinks to zero length, and its junctions meet. To consider what might happen to the faceted surface when this occurs, we first label the immediate surroundings of an edge. Each edge is composed of two faces of which it is the intersection, its composite faces, and stretches between two faces at which it terminates, its terminal faces. In addition, we will also use the term emanating edges to refer to those edges immediately neighboring the shrinking edge. Now, consider the hemispherical shell of available facet normals (Section 4.2.1.3). The (necessarily distinct) normals of the composite faces specify a great circle about this hemisphere, which divides it into two parts. The normals of the terminal faces cannot lie on this boundary, and unless they are identical (a special case), they form a second great circle around the hemisphere. While terminal normals may not lie on the composite great circle, the reverse is not true, and this fact effectively divides Vanishing Edge events into three sub-classes.

93

(a)

(b)

(c)

Figure 4.2. Normal Diagrams for different types of Vanishing Edge events. Blue dots represent the normals of composite faces, while red dots represent the normals of terminal faces. Dotted lines represent great circles between two points. (a) Terminal great circle touches neither composite point. (b) Terminal great circle touches one composite point. (c) Terminal normals occupy the same point. Great circle undefined. Figure 4.3.1 illustrates this idea, and gives an example of each of the three possible cases. If the terminal great circle touches neither composite point, then the well-studied Neighbor Switch occurs. If the terminal great circle touches one composite point, then an Irregular Neighbor Switch results. Finally, if the terminal normals occupy the same point, then no great circle is defined – the terminal facets have he same normal, and when the edge between them shrinks to zero, they join into a single facet: a Facet Join. 4.3.1.1. Neighbor Switch. On a general surface, the most common Vanishing Edge event is the neighbor switch, which is frequently encountered in other evolving cellular networks. In this event, neither composite normal touches the terminal great circle, so any three of the normals involved form a linearly independent set – this property is the defining feature of the neighbor switch. When an edge with this configuration shrinks to zero length, the surrounding facets simply exchange neighbors. Figure 4.3.1.1 gives an example of this event. Resolution. The neighbor switch is performed by the NS_repairman class. To resolve this event, it simply deletes the old edge, and creates a new edge. The composite faces

94

Figure 4.3. Example of a Neighbor Switch Event. Arrows represent gradients of regions in which they appear. which formed the old edge become terminal faces of the new edge, and cease to neighbor each other. Conversely, the terminal faces of the old edge become the composite faces of the new edge, and thus become neighbors. This symmetric exchange in neighbor relations is the cause of the name Neighbor Switch, which comes from the grain-growth literature – the less-descriptive name “T1 process” in often used in the soap froth literature. In addition to replacing the vanishing edge, the junctions on either side of this edge are replaced. Each new junction is formed by the intersection of the deleted edge’s (formerly non-adjacent) terminal faces with one of its composite faces. Comments. Readers familiar with other cellular-network literature will note that the example Neighbor Switch in Figure 4.3.1.1 lacks the typical “X” shape. This is due to the constrained nature of facet normals, and hence, edge orientations. Additionally, we note that the neighbor switch is a reversible event; in fact it is its own reversal. Finally, a certain sub-class of neighbor switches posessing “saddle” structure are non-unique, as was observed by Thijssen [14]. For a discussion of this non-uniqueness and its consequences, see Appendix B.2. 4.3.1.2. Irregular Neighbor Switch. When the normal of one of the composite faces lies on the great circle formed by the terminal normals, the neighbor switch cannot occur. Here, the terminal faces cannot form a new junction with the offending composite face

95

because the three normals involved are not independent. Instead, when an edge with this configuration shrinks to zero, two closely related events are possible, depending on the configuration of the nearby edges. These events are collectively called Irregular Neighbor Switches, with two varieties called a “gap opener” and “gap closer” that are exact opposites. These are illustrated in figure 4.3.1.2.

Figure 4.4. An example of the Irregular Neighbor Switch event. From left to right is the non-unique “gap opener.” From right to left is the unique “gap closer.” Arrows represent gradients of regions in which they appear, while circles indicate flat facets with zero gradient (vertical normal). Resolution. The irregular neighbor switch is performed by the INS_repairman class. Because one composite normal lies on the terminal great circle, exactly two of the emanating edges are parallel in R3 . The gap opener occurs when these edges emanate from the shrinking edge in opposite directions, while the gap closer occurs when the edges emanate in the same direction. To resolve the gap opener, we select one of the parallel emanating edges to be split apart (see below). The gap will go here, filled by the terminal face that touches the other parallel edge, and will extend all the way to the far end of the split edge, where a new edge is introduced to link the two edges resulting from the split edge. This is all illustrated in Figure 4.3.1.2. To resolve the gap closer, simply reverse the steps.

96

Comments. Several comments on this pair of events are in order. First, while the gap closer is uniquely resolved, the gap-opener is an inherently non-unique event, as either of the parallel edges could be the one split (we will discuss this further in section B.2). Second, both resolution options have the potentially dissatisfying property of being non-local in effect, because the collision of two junctions causes an entire edge to split apart. What is perhaps more likely is the nucleation of a new, tiny facet at the moment the junctions collide; however, we have excluded that possibility from consideration here. Finally, while common experimentally-encountered surfaces usually have either high symmetry (only a few facet orientations) or no symmetry (as many orientations as facets), the irregular neighbor switch with its three coplanar orientations requires what may be called “intermediate symmetry,” where orientations are limited, but many are available. Because it poses resolution difficulties, and because it is not encountered in any surfaces we wish to study, we have not yet actually implemented this event. 4.3.1.3. Facet Join. Finally, we consider the special case where the terminal normals are identical. When such an edge shrinks to zero length, the terminal faces meet exactly. Having the same orientation, they then join to form a larger face. Figure 4.3.1.3 depicts a representative facet join event.

Figure 4.5. An example facet join procedure. Arrows represent gradients of regions in which they appear.

97

Resolution. Facet Joins are performed by the FJoin_Repairman class. To perform a facet join, a new face is created to replace the joining faces, and all edges and junctions that neighbored the old faces are re-assigned to this new face. Next, the vanishing edge and its two junctions are deleted, leaving the four emanating edges to be considered. These are most logically grouped into the (necessarily parallel) pairs of edges bordering, respectively, the left and right composite faces of the vanished edge. In the example event shown in Figure 4.3.1.3, these two pairs look different: one pair meets side-to-side, while the other pair meets end-to-end. Computationally, however, this makes no difference; each pair is replaced by a single edge connecting their remaining non-deleted junctions. This behavior is generic for all face joins. Comments. We note that the face join is, strictly speaking, non-reversible (though see Section 4.3.2.1). The exact opposite of the face join would be a facet which spontaneously “shatters,” as described in [70]; this behavior is certainly worth studying, but is not currently implemented. Second, although this is a “special case” in general, for highsymmetry crystal surfaces it may be very common – indeed, for the case of a cubic crystal with only three available facet orientations considered in Chapter 2, Facet Joins are the only Vanishing Edge event exhibited. Finally, we note that this event is the only Vanishing Edge event which does not conserve the number of facets. It is, in fact, one mechanism by which coarsening may occur, and may be the dominant mechanism for high-symmetry surfaces.

98

4.3.2. Facet Constrictions The second class of topological event occurs whenever a facet ceases to be simply-connected, and results in that facet being split into two new facets. Remembering that the edges of a facet trace out a polygon in the plane, we observe that the non-simply connected polygon, if allowed to continue evolving, would become self-intersecting, which clearly has no geometrical interpretation. So, how may an evolving polygon become self-intersecting? Since the boundary consists of edges and junctions, there are three possible modes: (a) two non-adjacent junctions meet, (b) a junction meets an edge, or (c) two edges meet. Each case has a distinct “signature,” illustrated in Figure 4.3.2, which can be used to tell them apart.

(a)

(b)

(c1)

(c2)

Figure 4.6. Signatures of Constricted Facet events. (a) Non-Adjacent Junction-Junction collision signature. (b) Junction-Edge collision signature. (c1),(c2) Asymmetric and Symmetric Edge-Edge collision signatures.

The Junction-Junction collision shown in Figure 4.3.2a represents the formation of a perfect O(6) junction. While theoretically interesting, such events are not considered here; we hypothesize that, given random initial data, two junctions not connected by an edge will never exactly meet. Furthermore, by considering Figure 4.3.2, it can be seen that all Junction-Junction collisions, if perturbed as we hypothesize, result in either junction-edge or edge-edge collisions, and can therefore be resolved accordingly.

99

Junction-Edge collisions occur when a facet is pinched into two pieces by three of its neighbors, depicted in Figure 4.3.2b. There, two adjacent neighbors of the facet, forming a wedge, meet a third neighbor and pierce it. Two separate events are possible in this class. In most cases, the wedge simply splits the central facet into two parts, in an event called a Facet Pierce. However, if the normals of the wedge facets and the normal of the central facet lie on the same great circle, then, as the central facet is split, the opposing facet opens up a gap in the wedge: an Irregular Facet Pierce. Edge-edge collisions occur when a facet is pinched by four neighbors, shown in Figure 4.3.2c. In these events, two non-adjacent, exactly parallel edges meet, which requires that the normals of the impinging facets be coplanar with the normal of the pinched facet. Again, two variations are possible. If the impinging faces have different normals, the event is called a Facet Pinch. However, if they have the same normal, they join even as they pinch the facet in question, in a process called a Joining Facet Pinch. In addition, each event may occur in either symmetrical or asymmetrical flavors, which are shown in Figure 4.3.2c1,c2 respectively. The meeting of two edges requires the involvement of two junctions; these lie on the same edge for the symmetrical case, and on different edges for the asymmetrical case, as seen in the figure. 4.3.2.1. Facet Pierce. The first self-intersection we will study is the simplest; the facet pierce. It is a point-line event as described above; that is, a facet is split when a triangular wedge formed by two adjacent neighboring facets intersects the edge formed with a third, opposing neighbor. The facet pierce is functionally the opposite of a facet join, and is illustrated in Figure 4.3.2.1.

100

Figure 4.7. Example of a facet pierce procedure. At the moment the event occurs, an O(5) junction is formed, which immediately breaks in one of three ways, depending on the dynamics. Arrows represent gradients of the regions in which they appear.

Resolution. Each Facet Pierce is performed by the FJoin_Repairman class. Given the constricted facet, as well as the junction and edge which meet, it can label all of the surrounding facet elements and deterministically reconnect them correctly. First, two new facets are created to replace the constricted facet. The junctions and edges that bordered the old facet can be reassigned to these based on the labels created initially. The colliding junction and edge are deleted, to be replaced by three new junctions and two new edges. The locations of the former and neighbor relationships of each can be determined by considering Figure 4.3.2.1 and using the labels. Comments. First, technically, at the moment of the event, an O(5) junction forms, which as shown in Figure 4.3.2.1 may proceed to break in one of three ways. This does not, however, constitute a non-uniqueness; rather, the dynamics governing the surface

101

evolution at the moment of topological change specify which exit pathway is chosen. Second, while Thijssen [14] rightly objected to this resolution for the case of separate grains, we find it satisfactory for the case of a single crystal considered here. 4.3.2.2. Irregular Facet Pierce. A special modification of the Facet Pierce just described occurs when the normal of the opposing facet shares a great circle with the normals of the facets forming the wedge. This event is called an Irregular Facet Pierce. Recall that three new junctions were created during the facet pierce. However here, since the two newly created facets have identical normals, and the remaining three have normals which are not independent, those junctions cannot be created. Instead, as the wedge facets meet the opposing facet, one of two things happen – either the center edge of the wedge is split apart by the opposing facet (a “wedge split”), or the opposing facet is split apart by the wedge (a “wedge extension”). We see an illustration of each possibility in Figure 4.3.2.2.

Figure 4.8. Example of an irregular facet pierce procedure. Arrows indicate gradients; circles flat planes with no gradient. Resolution.

The Irregular Facet Pierce is performed by the IFP_Repairman class,

which at instantiation is given the constricted facet, as well as the junction and edge which

102

meet. This event is repaired quite similarly to the regular facet pierce, with modifications. As is done there, two new facets are created to replace the constricted facet, and junctions and edges bordering the old facet are reassigned to the new ones. The resolution differs in how to replace the colliding junction and edge. If the “wedge split”resolution is chosen, then the middle edge of the wedge and its far junction are also deleted – these are replaced by two parallel edges and junctions. Finally, an edge is formed which links them and borders the facet on the far side of the deleted edge. If the “wedge extension” resolution is chosen, not only is the constricted facet split apart, but so is the one opposite the edge split by the wedge. One must first determine which edge of this second split facet the extended wedge will intersect. Having done so, that facet is deleted, to be replaced by two new facets. The extension is formed by adding two edges parallel to the middle edge of the wedge, and the edge it intersects is split in two. Two new edges and three junctions must be created to link the extension with the edge it intersects. Finally, all edges and junctions bordering the deleted facet, plus those created to form the extension, are re-assigned appropriately to the new facets. Figure 4.3.2.2 is especially helpful here. Comments.

The event clearly recalls the “gap opener” described above. It shares

with that event three coplanar surface normals, and as a result, two possible resolutions. Additionally, while the two options here are qualitatively different compared to the symmetric options of the gap opener, they are additionally both non-local effects due to a local cause. Again, perhaps the best resolution is to nucleate a new facet, which we do not yet consider. Finally, both events require “intermediate symmetry,” and for the same reasons discussed above, we have not implemented this event.

103

4.3.2.3. Facet Pinch. We now turn to consider the case of Edge-Edge events, the first of which is called a Face Pinch. Here the normals of the pinching facets are not identical, and so junctions can be created as needed – an illustration of this event is shown in Figure 4.3.2.3. This event is philosophically similar to the face split described above. In each case, a facet is split into two by non-joining neighbors; the difference is just whether the procedure is “sharp” or “blunt”; i.e., caused by parallel edges or a junction and an edge.

Figure 4.9. Example of a symmetric face pinch. Arrows represent gradients of regions in which they appear.

Resolution. Each Facet Pinch is performed by the FPinch_Repairman class. Because of the similarities between the facet pierce and facet pinch, the associated Repairman classes behave similarly as well. Here, the Repairman class constructor takes the constricted facet and the two colliding edges. With this information, it can label all of the surrounding facet elements and deterministically achieve the change shown in Figure 4.3.2.3. As with the Facet Pierce, two new facets are created to replace the constricted facet, and the junctions and edges that bordered the old facet are reassigned as required. The colliding edges are deleted, as are the associated junctions discussed above. Five edges and four junctions are created to complete the reconnection, as shown in Figure 4.3.2.3.

104

Comments. This event, like the Irregular Neighbor Switch and Irregular Facet Pierce, requires a surface with “intermediate symmetry.” While it is uniquely resolved and poses no great difficulty of implementation, we have not yet implemented it for this reason. 4.3.2.4. Joining Facet Pinch. Finally, a special modification of the face-pinch occurs when the impinging facets have identical normals. The constricted is split in exactly the same way as in a face pinch; however, since the two facets doing the “pinching” are identically oriented, they join together to form a larger facet. We see an illustration of this situation in Figure 4.3.2.4.

Figure 4.10. Example of an asymmetric face swap. Arrows represent gradients of regions in which they appear.

Resolution. Each Joining Facet Pinch is performed by the JFPinch_Repairman class, which operates similarly to the Repairman classes associated with the Facet Pierce and Facet Pinch. This class is again instantiated with the constricted facet and the two meeting edges, which allows the necessary labeling. Again, two new facets are created to replace the constricted facet, but in this case the two facets which meet must join, and so another new facet must be created to replace them – necessary junction and edge reassignments are again easily carried out. Finally, rather than deleting the edges which meet and the associated junctions involved, the meeting edges are simply re-connected as shown in Figure 4.3.2.4.

105

Comments. Note that the final configuration is similar to the original configuration; in fact, with suitable facet motion, the surface could return to its original configuration via another face swap; the event is thus self-reversible in a sense. Also, since both a facet pinch and a facet join occur simultaneously, the total numbers of each surface element remain unchanged during this event.

4.3.3. Vanishing Facets The final class of topological event occurs when a facet shrinks to zero area and is removed. However, as has been noted numerous times previously in the context of cellular networks, very small facets can result in stiff dynamics that are difficult to numerically simulate accurately. For this reason, we follow previous authors by pre-emptively removing facets with areas below some small threshold (but see Section 4.4.1). This process is summarized in Figure 4.3.3. There, we see a single small flat facet vanishing into a pentagonal well (4.3.3a). Being smaller than the allowed threshold, it is removed, leaving a “hole” in the network (4.3.3b). The facets and edges bordering this hole we call the far field, and they need to be reconnected correctly to patch the hole. The correct reconnection for this particular well is shown in Figure 4.3.3c. The principal difficulty in this process occurs during the reconnection step (Figure 4.3.3c). Here, we are assigning new neighbor relationships to the far-field facets, which also involves the creation of new edges and junctions to form boundaries between them. In other cellular-network problems, these neighbor relationships (and hence the reconnection) is usually chosen randomly, under the reasoning that any error introduced is small

106

(a)

(b)

(c)

Figure 4.11. (a) a single flat facet vanishing into a pentagonal well. (b) removing the facet leaves a far field outside the dotted circle, which must be reconnected. (c) the unique reconnection shown inside the dotted circle. Arrows indicate gradients. enough to neglect and quickly corrected1. However, because the faceted-surface network represents a piecewise-planar geometrical surface, we are not free to choose randomly. Since each facet in the far field has a normal and a local height, neighbor relationships determine junction locations and thus edge placement. However, the final reconnection must be geometrically consistent – all facets must be simply connected, and thus no edges may intersect. If we were to randomly choose our neighbor relationships, the resulting reconnection would likely fail this test, and would thus represent a non-physical “surface.” To guarantee a geometrically consistent reconnection, we must search through all virtual reconnections until we find one that does not result in any self-intersecting facets. Several questions immediately arise: 1: How can we effectively characterize a “reconnection”? 2: How many virtual reconnections are there to search? 3: How can we efficiently list all these choices? 4: Can we be sure a good reconnection exists? 1See,

however, [102, 103], where the effects of this random choice in soap froths is investigated and found to be significant. A deterministic method of re-connection is proposed, based on the assumption that a cell loses sides as it shrinks until it has only three.

107

5: Is this reconnection unique?

For our method to be effective, all but the last of these questions must be answered satisfactorily. The detailed answers to (1-3) are found in the appendix, but we will summarize them here. The edges and junctions created during an O(n) reconnection may be effectively characterized as a binary tree with n − 2 nodes. The number of m−noded binary trees is given by the Catalan number Cm =

2n! . n!(n+1)!

Finally, these trees may be

efficiently listed using a greedy recursive algorithm in O(Cm ) time. For the fourth question regarding existence, we argue heuristically that a facet reaching zero area proves the existence of its own reconnection, since a surface with a zero-area facet is functionally the same as the surface with that facet removed. We then assume the existence of that same reconnection for some window of time before the facet reaches zero. A fuller proof would appeal to manifold theory. Finally, the fifth question regarding uniqueness is addressed in Section B.2. Having established these facts, we have a robust method for reconnecting an arbitrary far field of facets. Before considering some special cases of this method, let us summarize the general process so far: Whenever facets smaller than a threshold area are detected, we:

a: remove them, leaving a hole in the mesh. b: list all virtual reconnections (VR’s) as n-node binary trees. c: use associated neighbor relationships to find edge locations. d: test each VR until one with no intersecting edges is found.

108

We note that this approach represents a comprehensive reconnection method for any cellular network problem. Though it is necessary for the faceted surface problem, it may be useful in any situation where a verifiably optimal reconnection is sought. 4.3.3.1. Special Case: Facets Disappearing in Groups. It is possible for groups of facets to shrink together, in such a way that they cannot be removed sequentially. For an example, consider the configurations in Figure 4.3.3.1. In such a case, it is necessary to identify and collect a contiguous group of small facets for simultaneous deletion – we call this a near field. Any facet neighboring the near field is assigned to the far field, which may be reconnected as described previously after the near field is deleted.

Figure 4.12. Example of a group of disappearing facets. Reconstruction shown in dotted lines. To gather the near-field facets, we maintain a second, more liberal threshold. Whenever a face shrinks below the first threshold, as described above, its neighbors are recursively examined to collect those smaller than the second threshold. This method is rather simplistic, and, in cases of oddly-shaped pyramids, may not return the entire near field. This, in turn, will result in an incorrect far field, which will most likely be nonreconnectable. However, a group of facets vanishing together eventually all head to zero area, and for some window of time before they would physically vanish, all are small

109

enough to be detected in this way. Thus, we allow the code to “skip over” small facet combinations that it cannot remove successfully, and try again during a future timestep. 4.3.3.2. Special Case: Facets Disappearing as Steps. It is also possible, on highsymmetry crystal surfaces, that the small facet or group of facets forms a “step” between two much larger facets of identical orientation, but different height. Figure 4.3.3.2 illustrates this situation, in which the near field is bounded by exactly four facets, two of which have identical orientations. In such a case, the final fate of the surface is that the small facets vanish as the large facets join together. The method described above contains no provision for joining far-field facets during reconnection, and so there is no way to reconnect the far field produced in this case.

Figure 4.13. Example of a step removal. Left: A chain of small facets separates two large facets of identical orientation. Right: The small facets have been removed, and the large facets joined. Having identified a near field as forming a step, one solution is to delete the small facets, then move the two large faces to the same height and join them. This results in two pairs of unconnected edges, which are each deleted and replaced with an appropriate single edge. Since facet groups forming steps are, in fact, bordered by four facets generically, a separate Repairman class could be written to handle this case. However, the small

110

adjustment to the positions of the large facets can lead to subtle problems, as will be seen in Section 4.4. Therefore, a more robust if less elegant approach is to simply add one of the large parallel facets to the (step-forming) near field; a good choice is the one with fewer edges. Since the far field surrounding this modified near field requires no joins, it can be repaired using the FFR method. Again, failures are possible as described in the above section, but resolution is always possible near enough to the time the event would physically occur.

4.4. Discretization and Performance of Topological Events We have now discussed the general kinematics of a PADS, and surveyed all topological events which may occur as the surface evolves. Before our treatment is complete, however, we must consider with care the application of a time-stepping scheme. The accurate performance of topological events under such a scheme is problematic because, while events on a continuously evolving surface happen at precise times (Ei at ti ), any time-stepping method invariably skips over these times. This has three consequences, concerning detection, consistency, and accuracy. After discussing them briefly, we will present three possible timestepping methods which illustrate them in more detail. Detection.

Because timestepping will always skip over moments of topological

change, we must abandon hope of simply finding topological events ready to perform. Instead, we must either look ahead before each timestep and anticipate when events will occur (a predictive method); or step before looking, and then by examining the network infer where events should have occurred (a corrective method). Class A events can be

111

easily be detected either way, while class B events are easier to correct, and class C are easier to predict. Consistency.

Once the occurrence of an event has been detected by either means,

it must be performed in a way that preserves geometric consistency – i.e., the network always corresponds to a physical surface s = h(x). For example, two joining facets can only be mechanically fused if they exhibit the same local height. If, in addition to the occurrence of an event, a detection scheme can determine the exact time at which it occurred, then one strategy is to move the network to the precise event time, at which resolution is trivial. However, one may wish to attempt resolutions at other times, and the geometric consequences of doing so must be weighed. Accuracy Finally, we must consider the possibility of error that is produced during topological change. This error is most easily understood if we view the evolving surface in its abstract form as a highly nonlinear system of ODE’s. The (usually configurational) evolution function is moderated by the topological state; thus, topological events can represent sudden, qualitative changes in the evolution function. A naive time-stepping scheme which steps over these without appropriate measures will produce large localized errors at moments of topological change.

4.4.1. Method 1. Predict Events, Travel Exactly to Each Event Assume that, at all times, we accurately predict the time and location of the next topological event2. Then, a straightforward timestepping strategy which avoids consistency 2An

example of this approach may be found in the early soap froth simulations of [92], where edges and cells shrinking to zero are anticipated. A similar predictive approach could be developed for facets which become non-simply connected, by anticipating the possibility of junctions crossing edges.

112

and accuracy concerns is to continually calculate the time of the next topological event (accurate to the the order of the time-stepping method), and then step from event to event. Under this approach, time is divided into slices with constant equations of motion, guaranteeing that that the system always evolves under the correct equations, and accurately representing the continuously evolving surface. In addition, high-order single-step methods such as Runga-Kutta methods may be used to obtain high accuracy. Though neatly eliminating consistency and accuracy concerns, this method has a serious disadvantage. The frequency of topological events scales with the system size, and since we can never step farther than the next event, we effectively make the timestep dependent on system size – ∆t ∼ O(N −1 ). Since moving the system through a single timestep is itself an O(N) operation, then advancing the system through any O(1) period of time takes O(N 2 ) time. While acceptable for the detailed study of a small surface, it is obviously undesirable for the statistical study of large surfaces. This is chiefly because, consistency concerns aside, it makes little sense to halt the entire surface at every single topological event, when each of these involves only a few facets. Thus, our next method has as its chief objective the use of timesteps which are independent of system size.

4.4.2. Method 2. Use Fixed timestep – Late Correction of Observed Events A second strategy is to take fixed timesteps, use a corrective method of topological detection, and attempt to perform topological corrections late. Since timestep is independent of system size, many events will now occur per timestep, the size of which is chosen to produce a fixed small percentage of facets undergoing topological change each step. While this approach theoretically eliminates the O(N 2 ) contribution to running time, it

113

introduces hurdles to event detection, as well as geometrically consistent and accurate resolution. Detection. We just stated that, in this corrective detection scheme, more than one event occurs per timestep. Whether or not this is a problem depends on the Domain of Influence of each event, defined to be the set of network elements that event affects. If these sets contain no common elements, then the associated events occur too far apart in space to affect each other – they are independent. Consequently, a detection routine can hand them in arbitrary order to the repair routines, there to be confidently performed in isolation. However, occasionally two or more domains of influence overlap. In this situation, called a Discrete Compound Event, the associated topological events are no longer independent, and a detection routine can no longer ensure a priori their correct, consistent resolution when handed off. Even worse, the very signatures used to identify separate events may be obscured in the resulting “tangle,” such that the routine does not even recognize what has happened. Given the variety of event signatures described in Section 4.3, and the many combinations in which they might occur, creating a complete list of all DCE’s would be prohibitive if not impossible. Instead, we reason that, on a random surface, no two events will ever occur at exactly the same moment (It is possible to artificially construct faceted surfaces such two or more events must occur simultaneously – we do not consider this case). Thus, if we simply refine our timestep when necessary, formerly overlapping events can be sorted out, and detected in sequence. A robust strategy for handling compound events is thus to (a) retrace the problematic timestep, (b) refine it into smaller slices, and (c) repeat steps (a) and (b) recursively, until only single events are detected.

114

Consistency. Since the surface is allowed to evolve unrepaired past numerous topological events per timestep, surface regions near these events will be geometrically inconsistent after the step. To say the same thing, facets involved in the bypassed events will have incorrect neighbor relationships. However, we have already classified all possible events, so having identified which event occurred, and which facets were involved, we know a priori what the correct neighbor relationships should be after the event. This knowledge, along with knowledge of the position of each facet involved, allows us to reconstruct the consistent surface that should have emerged during the event3. Unfortunately, not all events can be consistently corrected at a late time in this way. In particular, Facet Joins and Joining Facet Pinches involve the joining of two facets that meet each other at a single local height. Since this condition exists for only a single instant, such events cannot be performed in a geometrically consistent way at any time other than the “correct” one. To accommodate this requirement while preserving a topology-independent timestep, we are forced to manually adjust the height of the joining facets before the event is performed. Besides the error induced by this strategy (discussed next), this need illustrates a second problem that can arise. In a RepairInduced Inconsistency, the very act of performing one event, because it is done late, triggers a second event that was not detected originally. An example is when the justdescribed height adjustment required for the delayed repair of a facet join triggers, say, a neighbor-switching event. Since this newly-triggered event was not originally detected, the system is left in an inconsistent state after all repairs are made. An ad-hoc strategy 3This

strategy is similar to the Far-Field Reconnection algorithm described above, except that except that the correct neighbor relationships are already known. However, FFR is a general algorithm for finding correct relationships between neighbors. Thus, many of the above topological events described above may be performed “lazily,” by simply identifying the involved facets, and applying FFR.

115

to find such RII’s is infeasible for the same reason as is a complete listing of all possible Discrete Compound Events (indeed, an RII may produce a DCE, which rules out a simple multiple-rechecking strategy). Thus, a similar retrace/refine/repeat strategy is required, with the added requirement that all events performed prior to detecting the RII must first be undone. Accuracy. Finally, as alluded above, repairing topological events after they occur can introduce large isolated errors. This can be due to the “fudging” required for the delayed repair of facet joins and swaps, but more generally is caused by facets involved in (uncorrected) topological events having been evolved under the wrong equations of motion for part of the relevant timestep. Consider an event Ei : tj < ti < tj+1 , with domain of influence Di . Since topological events likely correspond to a change in the surface’s evolution equation, the facets in Di are moved using the wrong equations for the time interval [ti , tj+1]. Since the equations guiding Di are wrong by as much as O(1) for a time of order O(∆t), facets in Di may accumulate O(∆t) location errors during the timestep in which the event occurs. Since the quantity of topological events does not depend on ∆t, the method retains first order accuracy globally. However, this error introduces a barrier to achieving higher-order accuracy later on.

4.4.3. Method 3. Localized Adaptive Replay The previous method, alas, contains one subtle problem that keeps it from being a true O(N) method. This problem is that the frequency of DCEs and RIIs, though small, still scales with the system size, and these necessitate timestep refinement. So although the late method does not have to explicitly step according to the O(1/N) time between

116

topological events, yet to accurately detect and resolve those events it is still implicitly driven by the refinement strategy to step along a time associated with DCEs and RIIs. While this characteristic time is longer than that between individual topological events, and does not greatly slow the simulation of tens of thousands of facets, it still results in a method that is formally O(N 2 ), which becomes prohibitive when considering systems of millions of facets. Thus, we now sketch a third method, not yet implemented, which eliminates this effect all together. In addition, the method allows us to perform topological events in a way which confers all the accuracy benefits of the first, predictive method. We first re-state that, on any given timestep, most facets are not involved in any topological changes. While it was therefore obviously wasteful to move the entire surface from event to event in the first method, it is also conceptually wasteful to perform a global retrace/refine/repeat step to DCEs and RIIs in the second method. Instead, after every timestep, we should identify for each DCE/RII the Topological Subdomain containing all facets involved in the event. The few facets within these subdomains would be retraced/refined/repeated as required, while the rest of the (unaffected) facets would be left undisturbed in their post-timestep state. Since operating on a given, constant number of facets takes O(1) time, and since the number of events per timestep scales only like O(N), we see that a single timestep and all associated corrections – including DCEs and RIIs – can now be performed in O(N) time, with a final state that is guaranteed to be consistent. This produces a true O(N) method. In addition, this “Localized Replay” strategy has an accuracy benefit. Regular, recognized topological events also have easily identifiable topological subdomains. If the facets within these domains are retraced, then

117

the predictive detection mechanism of the first method can be applied within the domain to eliminate consistency and accuracy problems associated with late removal. One difficulty remains, however. Facets involved in topological events may, under configurational facet-velocity laws, exhibit abrupt changes in velocity a result of the event. During the remaining segment of timestep, these facets may “break out” of the subdomain initially created to contain them, and begin interacting with facets outside of it. Thus, we would need a mechanism to detect this, and start over with a larger subdomain if it occurs. Finally, if subdomains can change size, then there is the possibility that two nearby subdomains will come to overlap as the algorithm progresses. Therefore, we must include the ability to merge them if necessary, start over with the new, larger sub-domain, and repeat adaptively until everything can be sorted. This adaptivity ensures the robustness of the method, as highlighted by the method’s formal name of Adaptive Localized Replay. The reader may note that the pattern of adaptive repetition is similar to that used to resolve DCEs and RIIs above, and worry that another, even smaller O(N 2 ) effect lurks in the shadows. However, in both of the previous methods, such effects were due to the global response to a local problem. Since this latter method is designed to be localized, there is no longer any mechanism to generate such effects.

4.5. Demonstration and Discussion We demonstrate our method using the sample dynamics derived in Chapter 1, associated with the directional solidification of a strongly anisotropic dilute binary alloy. When a sample is solidified at a pulling velocity which is greater than some critical value, solute gradients caused by solute rejection at the interface create a solute gradient which opposes

118

and overcomes the thermal gradient, resulting in a negative effective thermal gradient. In this environment, facets move away from the freezing isotherm at a rate proportional to their mean distance from the isotherm, as given by the dynamics (4.2)

Vi = hhii .

In Figures 4.5, 4.15, and 4.16, this dynamics is applied to surfaces with common three-, four-, and six-fold symmetries to illustrate the flexibility of our topology-handling approach. A series of snapshots from the coarsening surface are presented, in which surface configurations near to many of the topological events described above may be observed. (However, neither the Irregular Neighbor Switch, Irregular Facet Pierce, nor Facet Pinch occur because no three facet normals are coplanar in these symmetries; indeed, these events are not expected to occur on most physical surfaces, and were included for theoretical completeness.)

119

Figure 4.14. A top-down view of a small coarsening faceted surface with threefold symmetry. Spatial scale is constant, but irrelevant; time increases down the column.

120

Figure 4.15. A top-down view of a small coarsening faceted surface with fourfold symmetry. Spatial scale is constant, but irrelevant; time increases down the column.

121

Figure 4.16. A top-down view of a small coarsening faceted surface with sixfold symmetry. Spatial scale is constant, but irrelevant; time increases down the column.

122

About half of computational time is spent looking for topological changes, which is significant but not prohibitive. With appropriate timestep choice, using even the stillinefficient timestepping method 2 above, a surface of 25, 000 facets may be simulated to a 99 percent coarsened state in about an hour on currently available workstations. With the implementation of method 3 above, this time should be cut in half, and since method 3 is truly O(N), a single million-facet simulation should take about a day. Looking further ahead, since facet velocity calculations and topology checks require only local information, the method should be easily parallelizable, making possible even larger speed gains.

4.6. Conclusions We have presented a complete method for the simulation of fully-faceted interfaces of a single bulk crystal, with arbitary symmetry, where an effective facet velocity law is known. The surface, which is reminiscent of two-dimensional cellular networks, is encoded numerically in a geometric three-component structure consisting of facets, edges, and junctions, and the neighbor relationships between them. Consistent surface evolution specified by the facet velocity law is accomplished via a simple relationship between facet motion and junction motion. Although requiring the explicit handling of topological events, the method is efficient, using the natural structure of the surface, and accessible, allowing easy extraction of geometrical data. This combination makes it ideal for the statistical study of extremely large surfaces necessary for the investigation of dynamic scaling phenomena. A comprehensive listing of all topological events has been presented. These allow single-crystal surface with arbitrary symmetry (or no symmetry at all) to be simulated.

123

Events are classified into three categories, representing three ways that surface elements can become geometrically inconsistent. These are: edges which approach zero length, facets which become constricted, and facets which approach zero area. Resolution strategies for the former two classes can be determined a priori, while repairing surface “holes” left by vanishing facets requires a novel Far-Field Reconnection algorithm, which iteratively searches through all virtual reconnections to find one which produces a consistent surface. Finally, intrinsic non-uniqueness of several events is discussed; since ours is a purely kinematic method, decisions regarding resolution of these events must be made ahead of time through consideration of the dynamics or other physics. In addition, a detailed discussion of the issues associated with a discrete time-stepping scheme has been presented. The core issue is that topological events, which occur at discrete times throughout surface evolution, invariably fall between timesteps, with consequences for the detection of events, as well as their geometrically consistent and numerically accurate resolution. Since topological change corresponds (under configurational facet velocity laws at least) to qualitative changes in the local evolution function, some way to reach these in-between times must be introduced, while recognizing that only a few facets are involved in topological change during each timestep. A comparison of three approaches showed that the optimal solution is one of Localized Adaptive Replay, where large timesteps are taken to improve speed, but local surface subdomains associated with topological change are reverted, and then replayed in a way that re-visits events with the necessary precision as necessary. While further work remains to implement this approach, the method as presented is capable of comparing million-facet datasets via averaged runs.

124

CHAPTER 5

A Mean-Field Theory for Coarsening Faceted Surfaces 5.1. Introduction In many examples of faceted surface evolution, a facet velocity law giving the normal velocity of each facet can be observed, assumed, or derived. Examples of such dynamic laws describe growth of polycrystalline diamond films from the vapor [74, 73], evolution of faceted boundaries between two elastic solids [75], the evaporation/condensation mechanism of thermal annealing [13], and various solidification systems [71, 48, 76, 62]. Such velocity laws are typically configurational, depending on surface properties of the facet such as area, perimeter, orientation, or position, and reduce the computational complexity of evolving a continuous surface to the level of a finite-dimensional system of ordinary differential equations. This theoretical simplification enables and invites large numerical simulations for the study of statistical behavior. This has been done frequently for onedimensional surfaces [71, 48, 76, 62, 77, 100, 79, 80, 81], while less frequently for two-dimensional surfaces due to the necessity of handling complicated topological events [13, 14, 15, 16, 17]. Such inquiries reveal that many of the systems listed above exhibit coarsening – the continual vanishing of small facets and the increase in the average length of those that remain. Notably, these systems also display dynamic scaling, in which common geometric surface properties approach a constant statistical state, which is preserved even as the length scale increases.

125

The dynamical scaling behavior of coarsening faceted surfaces recalls the process of Ostwald ripening [104], in which small solid-phase grains in a liquid matrix dissolve, while larger grains accrete the resulting solute and grow. As this proceeds, the distribution of relative particle sizes tends to a constant state1. In fact, general 2D faceted surfaces fall conceptually into the same class of phase-ordering systems, except that the a vector order parameter reflecting surface normal replaces the scalar order parameter reflecting phase [105]; other systems exhibiting similar behavior include coarsening cellular networks describing soap froths and polycrystalline films [87, 88, 89, 90, 91], and films growing via spiral defect [82]. Each of these generalizations is characterized by a network of evolving boundaries which separate domains of possibly differing composition, and exhibit coarsening and convergence toward scale-invariant steady states. Since dynamic scaling pushes complex systems into a state which can be effectively characterized by just a few statistics, it is natural to seek simplified models which replicate this behavior. The canonical example of this approach is the celebrated theory of Lifshitz, Slyozov, and Wagner describing Ostwald ripening [18, 19, 20]. Generically, such an approach selects a distribution of some quantity, and includes just enough of the total system behavior to specify the effective behavior of that quantity – for example, the original LSW theory first identifies the average behavior of particles as a function of size, and uses that result to identify a continuity equation describing distribution evolution. Ideas of this kind have been applied to several of the higher-order cellular systems introduced above, notably froths [106, 107] and spiral-growth films [82]. To the extent that 1Indeed,

it was observed some time ago that facets of alternating orientations on a one-dimensional surface are analogous to alternating phases of a separating two-phase alloy [4, 5], and the Cahn-Hilliard equation [6] which models phase separation has been used, in modified form, to describe several different kinds of faceted surface evolution [8, 9].

126

such approaches mirror experimental data, they can yield valuable physical insight which cannot be gained by considering single particles, nor even by direct numerical simulation of larger ensembles. However, to date no similar attempt has been made for evolving faceted interfaces. Given the wide variety of examples of purely faceted motion, the membership of this problem class in the wider class of phase ordering systems, and the past success in applying mean-field analyses to these systems, it is somewhat surprising that no such attempt has been made to describe the mean-field evolution of faceted surfaces. In this chapter, therefore, we take a first step in that direction by introducing a framework for describing the distribution of facet lengths in 1D faceted surface evolution. Our approach closely resembles the LSW theory of Ostwald ripening, in that a facet-velocity law allows the effective behavior of facets by length, and thus the specification of a continuity equation governing the evolution of the length distribution. However, our model differs in that facets do not vanish in isolation as do grains in Ostwald ripening – instead each vanishing facet causes its two immediate neighbors to join together. This process of merging is not treated in LSW theory, and requires the introduction of a convolution integral reminiscent of equations due to Smoluchowski [21] and Schumann [22] describing coagulation. We apply our method to one particular facet dynamics, associated with the directional solidification of faceting binary alloys. However, the method is general and can be applied to any dynamics where effective facet behavior is accessible.

127

5.2. Example Dynamics and Problem Formulation To exhibit our method, we consider the dynamics derived in Chapter 1. During the directional solidification of a strongly anisotropic binary alloy, small-wavelength faceted surfaces develop. If the alloy is solidified above a critical velocity, a layer of supercooled liquid is created at the interface, which drives a coarsening instability governed by the facet dynamics (5.1)



dh dt



i

= hhii .

Figure 5.1a displays representative surface evolution during coarsening. There, the locations of corners are plotted over time. We see that this system exhibits binary coarsening, whereby a single facet shrinks to zero length, causing two corners meet and annihilate. As coarsening proceeds, the average facet length increases (Figure 5.1b), and a scale-invariant distribution of relative facet lengths is reached (Figure 5.1c). To describe the evolution of a scale-invariant length distribution such as that shown in Figure 5.1c, we will derive equations governing the evolution of the facet distribution ρ(x), where the value of ρ at x represents the number of facets with length L = x. This distribution can be used to obtain the total number of facets N(t), the average facet length L(t), and the (constant) total surface area A, via the relations (5.2)

N=

Z



ρ(x)dx

0

(5.3)

A=

Z



xρ(x)dx

0

(5.4)

L = A/N.

128

(a)

8

t

6 4 2 82

84

86

88

90

92

94

x 3

1 (b)

2.5

(c) 0.8 0.6

1.5



log

2

0.4

1

0.2

0.5 0

0

1

2 t

3

4

0

0

1

2

3

4

5

L /

Figure 5.1. Survey of coarsening behavior. (a) A representative example of the kink/anti-kink evolution (red/blue). (b) Facet lengthscale growth with time. (c) Scale-invariant distribution of relative facet lengths (the tail is gaussian).

Additionally, we will refer in what follows to the normalized probability distribution of facet lengths P (x), and the probability distribution of relative facet lengths Pˆ (ˆ x), given respectively by (5.5)

P (x) = P (x)/N

(5.6)

Pˆ (ˆ x) = LP (x/L),

with xˆ = x/L.

129

To simplify the calculations, we assume that all facets have slopes of ±1, and neighboring facet lengths are uncorrelated. Additionally, in what follows, it will be helpful to consider Figure 5.2, which shows a representative facet F , and its two neighbors F (−) and F (+) .

h − L/2 + L(+) h + L/2 h − L/2 h + L/2 − L(−)

L(−)

L

L(+)

Figure 5.2. A diagram illustrating a representative facet and its two neighbors. Here h is an arbitrary reference height that occurs at the midpoint of the center facet.

5.2.1. Flux Law We begin by observing that, because all slopes are fixed with alternating values ±1, then the rate of length change for any single facet is independent of its own vertical velocity, and is instead completely determined by the vertical velocity of its immediate neighbors. This general, geometric property can be obtained by inspection of Figure 5.2, and may be written (5.7)

  1 dhi−1 dLi i dhi+1 . = (−1) − dt 2 dt dt

130

where odd (even) facets have negative (positive) slopes. Assume for now that a consideration of the facet dynamics (5.1), the existing facet distribution ρ, and Equation (5.7) allow the derivation of an effective mean facet behavior (5.8)

φ(L) =



dL dt



(L),

which gives the average rate of length change as a function of facet length. Then ρφ gives the total flux of facets in the space of facet lengths parameterized by x. This allows us to use the divergence theorem to write a simple continuity equation for ρ(x), as follows: (5.9)

∂ ∂ρ + [ρ(x)φ(x)] = 0. ∂t ∂x

5.2.2. Coarsening Terms Since equation (5.9) aims to describe a coarsening faceted surface, it must accurately address the primary feature of coarsening – the shrinking and vanishing of facets over time. To see if it does, we consider Figure 5.2 again, and imagine that the facet F goes to length 0. We see that three things occur: first, the facet F itself vanishes; second, the neighbors of F also vanish; and third, a new facet is created which is the merging of the neighbors of F . If these processes are not captured by Eqn. (5.9), then we must add terms to it so that it does. The first process of facet vanishing is indeed captured by the continuity equation (5.9). In our framework, vanishing facets shrink to zero length and flow through the domain boundary at x = 0. Equation (5.9) naturally exhibits this behavior, and allows the easy extraction of the rate R at which coarsening occurs. This is simply the flux at

131

the origin, given by (5.10)

R = −ρ(0)φ(0).

The second process, neighbor loss, is not captured by (5.9). To include it, we model it as a sink S, which eliminates facets in a way that is probabilistically accurate. Recalling the assumption that adjacent facet lengths are not correlated, we can assume that Li−1 and Li+1 are each described by the (normalized) distribution P (x). Thus, the appropriate sink is (5.11)

S(x) = −2P (x).

The final process of facet creation is also not captured by (5.9). We model it, in contrast to neighbor loss, as a source Ψ, which creates facets of length Li−1 + Li+1 . Since these variables are independent, and each described by P , then the sum Li−1 + Li+1 is described by the joint probability function (5.12)

Ψ(x) = P2 (x) =

Z

0

x

P (s)P (x − s) ds,

obtained by integrating a two-point (probability) distribution P (x)P (y) along lines of constant x + y. The net modifications required by coarsening may now be summed into a single term (5.13)

C(x) ≡ S(x) + Ψ(x) = −2P (x) +

Z

x 0

P (s)P (x − s) ds.

132

Since C describes the additional effects of facets coarsening (i.e. leaving the domain), it must be multiplied by R, the rate at which this occurs, and then added to Eqn. (5.9). This gives the result (5.14)

∂ ∂ρ = − (ρφ) + RC(x). ∂t ∂x

5.2.3. Assuming a Scaling Viewpoint

Equation (5.14), with C defined as in (5.13), now correctly describes the distribution, by length, of facets on a coarsening surface. In particular, it models the two primary features of coarsening – the decrease in the number N of facets over time, and the corresponding increase in the average length L of those that remain. Indeed, by performing the same integrals used to obtain N and L above to the entire equation (5.14), we can estabilish that (5.15a) (5.15b)

∂N = −2R ∂t Z ∞ ∂L = P φdx. ∂t 0

These rates of change in N and L accurately describe the relative change in the number of facets and average length scale of any initial faceted surface. However, the ultimate aim of this chapter is to obtain a description of the (normalized) distribution of relative facet lengths Pˆ (ˆ x), which reaches a steady state during dynamic scaling. Since Pˆ is defined in terms of P , which in turn is defined in terms of ρ, and since since we know that ρ evolves

133

according to Equation (5.14), we can use several iterations of the chain rule to show that (5.16)

i ∂ hˆ ∂ Pˆ ˆ x) + R ˆ P (ˆ x)φ(ˆ =− ∂t ∂ xˆ

"Z

0



# ˆ ∂ P ) , Pˆ (s)Pˆ (ˆ x − s)ds + 2(Pˆ + xˆ ∂ xˆ

ˆ x) = φ(Lˆ ˆ ˆ = −Pˆ (0)φ(0). where φ(ˆ x)/L, and R

5.3. Application: Our chosen facet dynamics The framework just derived was completely general, describing the coarsening of any faceted surface, and indeed any binary system exhibiting binary coarsening. We here apply that general framework to the specific facet dynamics (5.1), by deriving the appropriate effective flux function φ(x). Using the dynamics itself, the general kinematic form of φ given in Equation (5.7), and recalling the diagram in Figure 5.2, we perform the following calculation: (5.17a) (5.17b)

=

(5.17c)

=

(5.17d)

=

(5.17e)

=

(5.17f) (5.17g)

 dh(−) dh(+) − dt dt 1 (−) (+)  h − h 2   1 1 1 (−) (+) [2h + L − L ] − [2h − L + L ] 2 2 2   1 1 L − [L(−) + L(+) ] 2 2   Z 1 ∞ 1 L− xP2 (x) dx 2 2 0

dL 1 = dt 2



1 (x − L(t)) 2 1 ˆ x − 1) . φ(hatx) = (ˆ 2 φ(x) =

134

In step (5.17e), since both L(−) and L(+) obey the probability distribution P , the sum L(−) and L(+) is again modelled by the joint probability function P2 . The effective contribution to φ is then obtained by performing a weighted integral of possible sums x multiplied by their relative prevelance P2 . Since the total mass of P2 equals unity, that integration describes the center of mass of P2 , which by considering the form of Eqn. (5.12) can be shown to equal 2L(t). This result, in turn, informs the final nondimensionalized result in step (5.17g). We see from this that facets smaller than the average shrink, while facets larger than the average grow. Broadly speaking, this is how coarsening works, so our approximation at least has the right form. ˆ we now write the complete evolution equation for the distribution Having calculated φ, of relative facet lengths under the specific facet dynamics (5.1). Dropping all hats, we have (5.18)

1 ∂ ∂P =− [(x − 1)P ] + R ∂t 2 ∂x

Z

0

x

 ∂ P (s)P (x − s) ds + 2 (xP (x)) . ∂x

5.4. Solution and Comparison with Numerical-Experimental Data Solution of Equation (5.18) is currently performed numerically, by letting arbitrary initial conditions relax to a steady state; our numerical method is given in Appendix C.1. This state is unsurprisingly independent of the initial condition chosen, but surprisingly simple in form – it is simply the exponential distribution P (x) = exp(−x), which is easily shown (after the fact) to satisfy Equation (5.18). We now proceed to compare the characteristics of this predicted steady state with those of the actual steady state found by direct simulation of the dynamics (5.1); our main results are shown in Figure 5.3.

135

We begin in Figure 5.3a by comparing the distribution P itself. The predicted exponential distribution is shown in blue; comparison with the green actual distribution reveals qualitative but not quantitative agreement. In particular, while the tail of the predicted distribution is (obviously) exponential, the tail of the actual distribution is gaussian. As a consequence, our mean-field steady state exhibits far too great an incidence of extremely long facets. Seeking the cause of this discrepancy, we next test the accuracy of our effective flux function φ(x). Figure 5.3b shows the contour plot of the distribution of length/velocity pairs ρ(x, φ). Finding the mean velocity for each length gives the statistical φ (dashes), which turns out to compare favorably with the predicted φ (solid). Both are linear with form φ = α(x − 1), and while the actual slope of 0.39 differs from the predicted slope of 0.5, they can be made to agree by simply scaling time, since every term in Equation (5.18) contains either φ or φ(0). So this approximation seems to be valid. Finally, we examine the coarsening terms: the sink S(x) in Figure 5.3c and the source Ψ(x) in Figure 5.3d. These are functionals of P , and so we are not surprised that the predicted values (blue) are different from the values calculated from the actual steady state (green). However, for both S and Ψ, we assumed that neighboring facet lengths were uncorrelated. If we instead calculate statistically the S and Ψ generated by vanishing facets (that is, the neighbors of vanishing facets), we get the curves in red, which are different not only from the predicted quantities, but also from the quantites we would have gotten from using the actual distribution P and assuming no correlations. This suggests that the ultimate culprit is the assumption that neighboring facet lengths are uncorrelated. Going back to the simulation, we now measure the correlation of the

136

lengths of nearby facets as a function of neighbor distance, in Figure 5.4. There we see a small but significant correlation for at least the first two neighbors. This produces the discrepancy between the green and red curves in Figures 5.3c,d above, and may be responsible for the discrepancy in the tail as well. This result is not surprising, as the main weakness of the original LSW theory which inspires our approach was also a failure to address correlations; later generalizations which corrected this deficiency agreed well with experimental data [108].

5.5. Conclusions We have presented a mean-field theory for the evolution of length distributions associated with coarsening faceted surfaces. In the spirit of LSW theory, a facet-velocity law governing surface evolution is used to establish a characteristic length-change law; this in turn leads to a simple continuity equation governing the evolution of the facet length distribution ρ(x). However, because the vanishing of any facet forces the joining of its two neighbors, this equation must be modified by the addition of appropriate terms describing coarsening, including a convolution term recalling models of coagulation. Our model therefore serves, apart from the direct application to facet dynamics, as a study in the union of these two mechanisms of steady statistical behavior. The scale-invariant distribution is tracked by studying the evolution of the normalized probability distribution of relative facet lengths Pˆ (ˆ x), which preserves both zeroeth and first moments. The resulting equation is solved by the exponential distribution, and numerical simulation reveals that any initial condition converges to this solution. This result unfortunately does not agree quantitatively with the more gaussian distribution

137

obtained by sampling a large surface simulated directly under the facet dynamics. Further investigation reveals that the likely culprit is the assumption that neighboring facet lengths are uncorrelated. Indeed, a similar assumption plauged the original LSW theory, and relaxing that assumption resulted in much better agreement with experiment. However, even with this deficiency, the model captures the essential feature of the dynamically scaling state – an effective facet behavior law which grows large facets and shrinks small ones, moderated by competing terms describing coarsening and continuous change of viewpoint, which respectively redistribute probability density toward infinity and zero, respectively. While later improvements to our model addressing neighbor correlation will undoubtedly increase its predictive capabilities, these same forces will still balance in the steady state. The model as presented thus serves as a qualitative explanation of the essential features of the scaling state, as well as a guide to further reasearch efforts.

138

1 (a) 0.8



0.6 0.4 0.2 0

0

1

2

3

4 L /

5

6

7

8

3 (b) 2

φ

1

0

−1

0

0.5

1

1.5

2

2.5 L /

3

3.5

4

4.5

5

0 0

(c)

−4 log |Ψ|

−4 log |S|

(d)

−2

−2

−6 −8

−6 −8

−10

−10

−12 0

2

4 L /

6

−12

0

2

4 L /

6

8

Figure 5.3. (a) Comparison between the theoretically-predicted (blue) and statistically-gathered (green) steady states. The former exhibits exponential decay, while the latter is gaussian. (b) Contour of the statistical distribution of length/velocity pairs ρ(x, φ). For each length, mean statistical velocity is plotted as dotted line, while the predicted velocity φ(x) is a solid line. (c,d) Comparison of log(−S/2), log(Ψ) as obtained by various means: from the predicted steady state (blue), from the actual steady state (green), and from measuring the neighbors of vanishing facets in the actual steady state (red).

139

correlation coefficient

1.5

1

0.5

0

−0.5

0

2

4

6

8 10 12 neighbor distance

14

16

18

20

Figure 5.4. Statistically-sampled correlation of facet lengths as a function of neighbor distance.

140

References

[1] G. Wulff, Zur frage der Ge-. schwindigkeit des wachstums und der auflsung von. krystallflchen, Z. Kristallogr. 34 (1901) 449, For a recent review, see M. Wortis, in Chemistry and Physics of Solid Surfaces, Edited by R. Vanselow and R. Howe (Springer, Berlin, 1988), vol. 7. [2] C. Herring, Some theorems on the free energies of crystal surfaces, Physical Review 82 (1) (1951) 87–93. [3] A. Chernov, The spiral growth of crystals, Soviet Physics - Uspekhi 4 (1961) 116– 148. [4] W. Mullins, Theory of linear facet growth during thermal etching, Philosophical Magazine 6 (71) (1961) 1313–1341. [5] N. Cabrera, On stability of structure of crystal surfaces, in: Symposium on Properties of Surfaces, ASTM Materials Science Series – 4, American Society for Testing and Materials, American Society for Testing and Materials, 1916 Race St., Philadelphia, PA, 1963, pp. 24–31. [6] J. Cahn, J. Hilliard, Free energy of a nonuniform system. I. Interfacial free energy, J. Chem Phys 28 (1958) 258. [7] J. Stewart, N. Goldenfeld, Spinodal decomposition of a crystal surface, Physical Review A 46 (10) (1992) 6505–6512. [8] F. Liu, H. Metiu, Dynamics of phase separation of crystal surfaces, Physical Review B 48 (9) (1993) 5808–5818. [9] A. Golovin, S. Davis, A. Nepomnyashchy, A convective Cahn-Hilliard model for the formation of facets and corners in crystal growth, Physica D 122 (1998) 202–230. [10] M. Siegert, M. Plischke, Slope selection and coarsening in molecular beam epitaxy, Physical Review Letters 73 (11) (1994) 1517–1522.

141

[11] M. Seigert, Coarsening dynamics of crystalline thin films, Physical Review Letters 81 (25) (1998) 5481–5484. ˇ [12] P. Smilauer, M. Rost, J. Krug, Fast coarsening in unstable epitaxy with desorption, Physical Review E 59 (6) (1999) R6263–R6266. [13] S. Watson, S. Norris, Scaling theory and morphometrics for a coarsening multiscale surface, via a principle of maximal dissipation, Physical Review Letters 96 (2006) 176103. [14] J. Thijssen, Simulations of polycrystalline growth in 2+1 dimensions, Physical Review B 51 (3) (1995) 1985–1988. [15] S. Barrat, P. Pigeat, E. Bauer-Grosse, Three-dimensional simulation of CVD diamond film growth, Diamond and Related Materials 5 (1996) 276–280. [16] G. Russo, P. Smereka, A level-set method for the evolution of faceted crystals, SIAM Journal of Scientific Computing 21 (6) (2000) 2073–2095. [17] P. Smereka, X. Li, G. Russo, D. Srolovitz, Simulation of faceted film growth in three dimensions: Microstructure, morphology, and texture, Acta Materialia 53 (2005) 1191–1204. [18] I. Lifshitz, V. Slezov, Kinetics of diffusive decomposition of supersaturated solid solutions, Soviet Physics JETP 38 (1959) 331–339. [19] I. Lifshitz, V. Slyozov, The kinetics of pecipitation from supersaturated solid solutions, J. Phys. Chem. Solids 19 (1961) 35–50. [20] C. Wagner, Theorie der alterung von niederschl¨agen durch uml¨osen, Zeitschrift f¨ ur Elektrochemie 65 (1961) 581–591. [21] M. von Smoluchowski, Drei vortr¨age u ¨ ber diffusion, brownsche molekularbewegung und koagulation von kolloidteilchen, Physikalische Zeitschrift 17 (1916) 557–571. [22] T. Schumann, Theoretical aspects of the size distribution of fog particles, Quarterly Journal of the Royal Meteorological Society 66 (1940) 195–207. [23] W. Mullins, R. Sekerka, Stability of a planar interface during solidification of a dilute binary alloy, Journal of Applied Physics 35 (2) (1964) 444–451. [24] R. Trivedi, Interdendritic spacing: Part II. A comparison of theory and experiment, Metallurgical Transactions A 15 (6) (1984) 977–982.

142

[25] J. Langer, Instabilities and pattern formation in crystal growth, Reviews of Motern Physics 52 (1) (1980) 1–28. [26] J. Rutter, B. Chalmers, A prismatic substructure formed during solidification of metals, Canadian Journal of Physics 31 (1953) 15. [27] W. Tiller, J. Rutter, K. Jackson, B. Chalmers, The redistribution of solute atoms during the solidification of metals, Acta Metallurgica 1 (1953) 428–437. [28] D. Wollkind, L. Segel, A nonlinear stability analysis of the freezing of a dilute binary alloy, Philos. Trans. Roy. Soc. London A 268 (1970) 351–380. [29] G. Sivashinsky, On cellular instability in the solidification of a dilute binary alloy, Physica D 8 (1983) 243–248. [30] K. Brattkus, S. Davis, Cellular growth near absolute stability, Physical Review B 38 (16) (1988) 11452–11460. [31] D. Riley, S. Davis, Long-wave morphological instabilities in the directional solidification of a dilute binary mixture, SIAM Journal of Applied Mathematics 50 (2) (1990) 420–436, connects [29] and [30]. [32] S. Coriell, G. McFadden, Morphological stability, in: D. Hurle (Ed.), Handbook of Crystal Growth, Vol. 1b, Elsevier Science Publishers, 1993, Ch. 12, pp. 785–857. [33] B. Billia, R. Trivedi, Pattern formation in crystal growth, in: D. Hurle (Ed.), Handbook of Crystal Growth, Vol. 1b, Elsevier Science Publishers, 1993, Ch. 14, pp. 899–1073. [34] G. McFadden, S. Coriell, Nonplanar interface morphologies during unidirectional solidification of a binary alloy, Physica D 12 (1984) 253–261. [35] L. Ungar, M. Bennett, R. Brown, Cellular interface morphologies in directional solidification. III. The effects of heat transfer and solid diffusivity, Physical Review B 31 (9) (1985) 5923–5930. [36] L. Ungar, R. Brown, Cellular interface morphologies in directional solidification. IV. The formation of deep cells, Physical Review B 31 (9) (1985) 5931–5940. [37] A. Karma, Wavelength selection in directional solidification, Physical Review Letters 57 (1986) 858–861.

143

[38] D. Kessler, H. Levine, steady-state cellular growth during directional solidification, Physical Review A 39 (1988) 3041–3052. [39] W. Boettinger, J.A.Warren, C. Beckermann, A. Karma, Phase-field simulation of solidification, Annual Review of Materials Science 32 (2002) 163–194. [40] B. Echebarria, R. Folch, A. Karma, M. Plapp, Quantitative phase-field model of alloy solidification, Physical Review E 70 (2004) 061604. [41] S. Davis, Theory of Solidification, Cambridge University Press, 2001. [42] C. Herring, Surface tension as a motivation for sintering, in: W. Kingston (Ed.), The physics of powder metallurgy, McGraw-Hill, 1951, p. Chap. 8, proceedings of a symposium held at Bayside, L. I., New York, August 24-26, 1949. [43] J. Gibbs, On the equilibrium of heterogeneous substances, Vol. 1, Longmans, Green, and Co., 1928, Ch. 3, p. 219. [44] S. Angenent, M. Gurtin, Multiphase thermomechanics with interfacial structure. 2. Evolution of an isothermal interface, Archive for Rational Mechanics and Analysis 108 (1989) 323–391. [45] A. Di Carlo, M. Gurtin, P. Podio-Guidugli, A regularized equation for anisotropic motion-by-curvature, SIAM Journal of Applied Mathematics 52 (4) (1992) 1111– 1119. [46] R. Bowley, B. Caroli, C. Caroli, F. Graner, P. Nozi‘eres, B. Roulet, On directional solidification of a faceted crystal, Journal de Physique (France) 50 (1989) 1377–1391. [47] B. Caroli, C. Caroli, B. Roulet, Directional solidification of a faceted crystal. II. phase dynamics of crenellated front patterns, Journal de Physique (France) 50 (1989) 3075–3087. [48] D. Shangguan, J. Hunt, Dynamical study of the pattern formation of faceted cellular array growth, Journal of Crystal Growth 96 (1989) 856–870. [49] S. Coriell, R. Sekerka, The effect of the anisotropy of surface tension and interface kinetics on morphological stability, Journal of Crystal Growth 34 (1976) 157–163. [50] G. McFadden, S. Coriell, R. Sekerka, The effect of surface tension anisotropy on cellular morphologies, Journal of Crystal Growth 91 (1988) 180–198.

144

[51] R. Hoyle, G. McFadden, S. Davis, Pattern selection with anisotropy during directional solidification, Phil. Trans. Roy. Soc. Lond. A 354 (1996) 2915–2949. [52] K. Jackson, J. Hunt, Transparent compounds that freeze like metals, Acta Metallurgica 13 (1965) 1212–1215. [53] F. Melo, P. Oswald, Destabilization of a faceted smectic-A – smectic-B interface, Physical Review Letters 64 (12) (1990) 1381–1384. [54] D. Shangguan, J. Hunt, In situ observation of faceted cellular array growth, Metallurgical Transactions A 22A (1991) 941–945. [55] G. Young, S. Davis, K. Brattkus, Anisotropic interface kinetics and tilted cells in unidirectional solidification, Journal of Crystal Growth 83 (1987) 560–571. [56] G. Merchant, S. Davis, Morphological instability in rapid directional solidification, Acta Metallurgica et Materialia 38 (1990) 2683–2693. [57] C. Herring, The use of classical macroscopic concepts in surface energy problems, in: R. Gomer, C. Smith (Eds.), Structure and Properties of Solid Surfaces, University of Chicago Press, 1953, p. 5. [58] W. Mullins, Solid surface morphologies governed by capillarity, in: W. Robertson, N. Gjostein (Eds.), Metal Surfaces: Structure, Energetics, and Kinetics, American Society for Metals, Metals Park, OH, 1963, Ch. 2, pp. 17 – 66. [59] F. Frank, The geometrical thermodynamics of surfaces, in: W. Robertson, N. Gjostein (Eds.), Metal Surfaces: Structure, Energetics, and Kinetics, American Society for Metals, Metals Park, OH, 1963, Ch. 1, pp. 1–16. [60] P. Voorhees, S. Coriell, R. Sekerka, G. McFadden, The effect of anisotropic crystalmelt surface tension on grain boundary groove morphology, Journal of Crystal Growth 67 (1984) 425–440. [61] S. J. Watson, Crystal growth, coarsening and the convective Cahn-Hilliard equation, in: P. Colli, C. Verdi, A. Visintin (Eds.), Free Boundary Problems: Theory and Applications, International Series of Numerical Mathematics, Vol. 147, Birkhaeuser, 2003. [62] S. Watson, F. Otto, B. Rubinstein, S. Davis, Coarsening dynamics of the convective Cahn-Hilliard equation, Physica D 178 (3-4) (2003) 127–148. [63] J. Ericksen, Equilibrium of bars, Journal of Elasticity 5 (1975) 191–202.

145

[64] S. M¨ uller, Minimizing sequences for nonconvex functionals, phase transitions and singular perturbations, in: Kirchg¨assner (Ed.), Problems involving change of type, Vol. 359 of Lecture Notes in Physics, Springer-Verlag, Berlin, 1990, Ch. 1, pp. 31–44. [65] S. Mller, Singular perturbation as a selection criterion for periodic minimizing sequences, Calculus of Variations 1 (1993) 169–204. [66] L. Truskinovsky, G. Zanzotto, Ericksen’s bar revisited: Energy wiggles, J. Mech. Phys. Solids 44 (8) (1996) 1371–1408. [67] R. Sekerka, A time-dependent theory of stability of a planar interface during dilute binary alloy solidification, in: H. Peiser (Ed.), Crystal Growth, Pergamon, Oxford, 1967, pp. 691–702. [68] R. Sekerka, Application of the time-dependent theory of interface stability to an isothermal phase transformation, Journal of Physics and Chemistry of Solids 28 (1967) 983–994. [69] J. Villain, Continuum models of crystal growth from atomic beams with and without desorption, Journal de Physique I (France) 1 (1991) 19–42. [70] A. Roosen, J. Taylor, Modeling crystal growth in a diffusion field using fully faceted interfaces, Journal of Computational Physics 114 (1) (1994) 113–128. [71] L. Pfeiffer, S. Paine, G. Gilmer, W. van Saarloos, K. West, Pattern formation resulting from faceted growth in zone-melted thin films, Physical Review Letters 54 (17) (1985) 1944–1947. [72] S. Fu, I. M¨ uller, H. Xu, The interior of the pseudoelastic hysteresis, Mater. Res. Soc. Symp. Proc. 246 (1992) 39–42. [73] A. van der Drift, Evolutionary selection, a principle governing growth orientation in vapor-deposited layers, Philips Research Reports 22 (1967) 267–288. [74] A. Kolmogorov, To the ”geometric selection” of crystals, Dokl. Acad. Nauk. USSR 65 (1940) 681–684. [75] M. Gurtin, P. Voorhees, On the effects of elastic stress on the motion of fully faceted interfaces, Acta Materialia 46 (6) (1998) 2103–2112. [76] C. Emmot, A. Bray, Coarsening dynamics of a one-dimensional driven Cahn-Hilliard system, Physical Review E 54 (5) (1996) 4568–4575.

146

[77] C. Wild, N. Herres, P. Koidl, Texture formation in polycrystalline diamond films, Journal of Applied Physics 68 (3) (1990) 973–978. [78] A. Dammers, S. Radelaar, 2-dimensional computer modeling of polycrystalline growth, Textures and Microstructures 14 (1991) 757–762. [79] Paritosh, D. Srolovitz, C. Battaile, J. Butler, Simulation of faceted film growth in two dimensions: Microstructure, morphology, and texture, Acta Materialia 47 (7) (1999) 2269–2281. [80] J. Zhang, J. Adams, FACET: a novel model of simulation and visualization of polycrystalline thin film growth, Modelling Simul. Mater. Sci. Eng 10 (2002) 381– 401. [81] J. Zhang, J. Adams, Modeling and visualization of polycrystalline thin film growth, Computational Materials Science 31 (3-4) (2004) 317–328. [82] T. Schulze, R. Kohn, A geometric model for coarsening during spiral-mode growth of thin films, Physica D 132 (1999) 520–542. [83] J. Taylor, J. Cahn, Diffuse interfaces with sharp comers and facets: Phase field models with strongly anisotropic surfaces, Physica D 112 (1998) 381–411. [84] J. Eggleston, G. McFadden, P. Voorhees, A phase-field model for highly anisotropic interfacial energy, Physica D 150 (1-2) (2001) 91–103. [85] W. Carter, A. Roosen, J. Cahn, J. Taylor, Shape evolution by surface diffusion and surface attachment limited kinetics on completely faceted surfaces, Acta Metallurgica et Materialia 43 (12) (1995) 4309–4323. [86] A. Roosen, W. Carter, Simulations of microstructural evolution: Anisotropic growth and coarsening, Physica A 261 (1998) 232–247. [87] J. Stavans, The evolution of cellular structures, Reports on Progress in Physics 56 (6) (1993) 733–789. [88] V. Fradkov, D. Udler, Two-dimensional normal grain growth: Topological aspects, Advances in Physics 43 (6) (1994) 739–789. [89] H. Frost, C. Thompson, Computer simulation of grain growth, Current Opinion in Solid State and Materials Science 1 (3) (1996) 361–368.

147

[90] C. Thompson, Grain growth and evolution of other cellular structures, Solid State Physics 55 (2001) 269–314. [91] G. Schliecker, Structure and dynamics of cellular systems, Advances in Physics 51 (5) (2002) 1319–1378. [92] D. Weaire, J. Kermode, Computer simulation of a two-dimensional soap froth I. Method and motivation, Philosophical Magazine B 48 (3) (1983) 245–259. [93] D. Weaire, J. Kermode, Computer simulation of a two-dimensional soap froth II. Analysis of results, Philosophical Magazine B 50 (3) (1984) 379–395. [94] M. Anderson, D. Srolovitz, G. Grest, P. Sahni, Computer-simulation of grain-growth .1. Kinetics, Acta Metallurgica et Materialia 32 (5) (1984) 783–791. [95] D. Srolovitz, M. Anderson, P. Sahni, G. Grest, Computer-simulation of grain-growth .2. Grain-size distribution, topology, and local dynamics, Acta Metallurgica et Materialia 32 (5) (1984) 793–802. [96] F. Lewis, The correlation between cell division and the shapes and sizes of prismatic cells in the epidermis of cucumis, Anatomical Record 38 (3) (1928) 341–376. [97] D. Aboav, The arrangement of grains in a polycrystal, Metallography 3 (1970) 383–390. [98] D. Weaire, Some remarks on the arrangement of grains in a polycrystal, Metallography 7 (1974) 157–160. [99] D. Aboav, Arrangement of cells in a net, Metallography 13 (1) (1980) 43–58. [100] J. Thijssen, H. Knops, A. Dammers, Dynamic scaling in polycrystalline growth, Physical Review B 45 (15) (1992) 8650–8656. [101] J. Glazier, D. Weaire, The kinetics of cellular-patterns, Journal of PhysicsCondensed Matter 4 (8) (1992) 1867–1894. [102] B. Levitan, E. Domany, Topological model of soap froth evolution with deterministic T2 processes, Europhysics Letters 32 (7) (1995) 543–548. [103] B. Levitan, E. Domany, Dynamical features in coarsening soap froth: Topological approach, International Journal of Modern Physics B 10 (28) (1996) 3765–3805. [104] W. Ostwald, Z. Phys. Chem. 34 (1900) 495.

148

[105] A. Bray, Theory of phase-ordering kinetics, Advances in Physics 43 (1994) 357–459. [106] C. Beenakker, Evolution of two-dimensional soap-film networks, Physical Review Letters 57 (19) (1986) 2454–2457. [107] H. Flyvbjerg, Model for coarsening froths and foams, Physical Review E 47 (6) (1993) 4037–4054. [108] M. Marder, Correlations and ostwald ripening, Physical Review A 36 (2) (1987) 858–874. [109] M. Miksis, L. Ting, Effective equations for multiphase flows – waves in a bubbly liquid, in: Advances in Applied Mechanics Vol. 28, Vol. 28 of Advances in Applied Mathematics, Academic Press, 1992, pp. 141–260.

149

APPENDIX A

Appendices for Chapter 1 A.1. Justification of the Quasi-steady state The quasi-steady approximation is often justified by identifying a small Peclet number, which measures the ratio of typical structural lengths to diffusional lengths. No Peclet number is directly obtained from the non-dimensionalization performed above; however, a typical definition of the Peclet number looks like (A.1)

P =

VL , D

where V , L, and D are characteristic velocities, lengths, and the diffusion coefficient. With this definition, a consideration of the facet-velocity (2.50) shows that, for nonperiodic evolving faceted surfaces, V ∼ λ, L ∼ λ, and D ∼ 1. This gives an O(λ2 ) Peclet number, in agreement with the original quasi-steady assumption. However, we can more rigorously arrive at the steepest-descent form (2.46), at least in the small-slope sense. Beginning again with Equations (2.7), but neglecting the hx Cx term, we now keep all the time derivatives, and use Laplace and Fourier transforms to solve the equations. In Fourier-Laplace space, we obtain the following solution: (A.2) (A.3)

LF[ht + kh] = −LF [GR ] m(κ, s) − (1 − k) p 1 m(κ, s) = (1 + 1 + 4(κ2 + s)) 2

150

where LF indicates the Fourier-Laplace transform, GR represents the right hand side of the Gibbs-Thompson equation, and κ and s are the Fourier and Laplace variables, respectively. Now, the right-hand side can be directly inverted, which simply returns GR . The question is what to do about the left-hand side. The effect of the quasi-steady approximation Ct → 0 is to replace m(κ, s) with m(κ, 0), but we’ll attempt to directly perform the inverse Laplace transform to see if this is justified. We begin with the convolution theorem for Laplace transforms, which says that L

(A.4)

−1

[F (s)G(s)] =

Z

0

t

f (τ )g(t − τ )dτ

and we let (A.5)

F (s) = L[kh + ht ],

G(s) =

1 . m(κ, s) − (1 − k)

Our task is thus to find the inverse Laplace transform of G(s), and examine the resulting integral. While this expression exhibits no easily-invertible form, we can approximate the Bromwich integration to leading order for small t, and arrive at the leading-order approximation (in κ) of (A.6)

1 exp(−κ2 t) √ . g(t) ≈ √ π t

(We want the small t behavior because of the form of Eqn. (A.4) – the values function g(τ ) with small τ multiply the values of h and ht with τ near t, which are expected to matter the most.) Then, in the small-wavelength limit κ → ∞, we can extract the leading

151

order behavior of the convolution as (A.7a) (A.7b)

−1

L F (s)G(s) ≈ [kh + ht ] ≈

Z

0



2 exp[−(κ2 + 1/4)t] √ √ dτ π t

1 [kh + ht ] κ0

which is exactly what we would get from the the quasi-steady approximation in the same limit (see the argument leading to Eqn. (2.26)).

A.2. Homogenized Linear Stability Analysis In Section 2.5.1, we developed a stability criteria for small-wavelength periodic faceted surfaces. However, that criteria was only valid for small-wavelength disturbances – namely, perturbations to the facet heights of the periodic surface. Here, we consider the stability of the micro-faceted solution to perturbations with wavelength much larger than the solution wavelength. We will show that long-wavelength disturbances are categorically less destabilizing than the small disturbances considered in the main text. We again consider disturbed solutions of the form (A.8)

˜ h = h0 + h,

˜ C = C0 + C,

but where [h0 , C0 ] now include the faceted corrections derived in the text. We then insert forms (A.8) into the original non-dimensional governing equations (2.7). The result of this is a system of equations similar to those describing the linearization about the planar state (2.13). However, since [h0 , C0 ] are no longer planar, there are some additional terms

152

in the boundary conditions at z = h having variable coefficients: (A.9a)

Cˆt = Cˆz + Cˆzz + Cˆxx

for z > 0

(A.9b)

Cˆ → 0

as z → ∞

(A.9c) (A.9d)

h i ˆ − (1 − k)Cˆ + h ˆ t + C 0h ˆ x + h0 Cˆx + h0 C 0 h ˆ Cˆz = k h x x x xz Cˆ = Geff h

on z = 0 on z = 0

h i ˆ ˆ − Γ SX hx + SXX hxx

h i ˆ x + KXX h ˆ xx + KXXX h ˆ xxx + KXXXX h ˆ xxxx , + ν¯ KX h where S = sˆ(hx )K, K = (Kss + K3 ), subscripts indicate derivatives with respect to the proper derivative of h (i.e. SX indicates the derivative of S with respect to hx ), and each quantity is evaluated at h0 . Now, for disturbances [h¯n , C¯n ] with wavelength much greater than λ (the wavelength of [h¯0 , C¯0 ]), we may use the technique of homogenization. In this approach, we “average out” the effects of the (relatively) rapidly varying coefficients above. Following [109], we define a fast-space variable ξ = x/λ (where λ is the small wavelength of the steady solution). Then we let (A.10)

¯ C] ¯ C](x, ¯ = [h, ¯ [h, ξ, t),

(A.11)

¯ C] ¯ = [h,

∞ X

where ξ =

x λ

λn [h¯n , C¯n ]

0

and collect like powers of λ. With the assumption that all φ¯ are bounded and four-times ¯ 0 through h ¯ 3 are differentiable, the large orders λ−4 through λ−1 serve to establish that h ξ-independent, as are C¯0 and C¯1 . Then we take the average of the order λ0 equations,

153

and find as a solvability condition, equations (A.9), but with all coefficients averaged over one period. Now, considerations of the form of h, as well as symmetries in S and K, reveal that many of the variable coefficients vanish (h0x , Cx0 , SX , KX , KXXX ). Of those that remain, 0 hh0x Cxz i is small compared to k, hKXX i is small compared to hSXX i, and hKXXXX i merely

modifies ν¯; this leaves SXX for our consideration. Recalling the definition of S, and comparing Eqns. (2.13) and (A.9), we see that this term is an effective, period-averaged surface stiffness which replaces the surface stiffness of the planar state sˆ0 . Critically, this turns out to be positive for all values of the anisotropy coefficient α4 , and thus all solution slopes q ∗ . Thus, whereas the planar state had a negative surface stiffness leading to universal instability, the nonplanar solutions under consideration have a positive effective homogenized surface stiffness. The homogenized linear stability analysis then is identical to the original Mullins-Sekerka stability analysis for positive surface stiffness, leading to a Mullins-Sekerka-like region of instability similar to that in Figure 2.1. However, this region of instability lies entirely within the supercooled region M−1 > 1, indicating that facetwise instability and coarsening as described in Section 2.5.3 will always occur at lower pulling velocities than the long-wave instability discussed here. Therefore, this analysis is needed only for completeness, and serves functionally only to rule out behaviors not already considered in the main text.

154

APPENDIX B

Appendices to Chapter 3 B.1. Elaboration on Far-Field Reconnection Our method of removing facets and facet groups requires patching a “hole” in the network left by the deleted facets. This requires selecting a geometrically consistent reconnection from a list of potential, or virtual reconnections. As outlined in the text, this involves searching through a complete list of virtual reconnections and testing each for geometric consistency. In this Appendix, we address in more detail questions (1-3) posed in Section 4.3.3 regarding the details of this method. For convenience, we repeat them here:

1: How can we effectively characterize a “reconnection”? 2: How many potential reconnections are there to search? 3: How can we efficiently list all potential choices?

We show here that an effective means of answer these questions is to think of reconnections as extended binary trees. This characterization enables us to easily count potential reconnections, distinguish them through naming, and suggests an algorithm for efficiently listing them for testing. An exhaustive illustration of the process is given for the case of an O(5) far field in Figure B.1. It will be useful to refer to that diagram during the following discussion.

155

B.1.1. Characterization: Binary Trees Patching network holes always involves finding unknown neighbor relations between a given number of adjacent facets – that is, no facets are ever created, only edges and junctions. These are always connected into a single graph. In fact, the edges and junctions created during reconnection (the “reconnection set”) form a binary tree1. In Figure B.1, the trees associated with each possible reconnection are shown in thick blue lines. The far-field edges touching the reconnection set shown in gray represent the completion of this tree. That is, they take the interior tree, and add leaves to it so that every node of the interior tree is a triple-node. Each virtual reconnection corresponds to a unique interior tree and completed tree in this manner.

B.1.2. Enumeration: The Catalan Number The counting of binary trees is, fortunately, a solved problem of graph theory. Given n nodes, they may be arranged in Cn distinct binary trees, where Cn is the nth Catalan Number ; (B.1)

Cn =

2n! . n!(n + 1)!

Now, re-connecting an O(n) far field requires the creation of n − 3 edges and n − 2 nodes; this may be visually confirmed for the case n = 5 in Figure B.1. This creates an n − 2 noded binary tree, and so an O(n) far field has Cn−2 virtual reconnections to search. 1A

binary tree is a graph consisting of edges and nodes such that: (1) each node connects to 1,2 or 3 other nodes, (2) no cycles exist. Condition (1) is met because we consider only triple junctions, while condition (2) is met because the creation of new facets is excluded.

156

B.1.3. Naming If take the completed binary tree and arbitrarily select a root node, then from each virtual interior tree we can generate a unique sequence of letters which identify it and encode its construction. This can be formed in one of two ways. The first way is to specify the tree by a set of recursive function calls. Each branch point has left and right branches, each of which may terminate in either a leaf, or another branch point. The middle column of Figure B.1 gives such a function for each tree shown in the left column. The second way is to walk around the tree in a counter-clockwise manner, recording each branch point or leaf as it is encountered. Either method produces a series of letters that uniquely identify the tree. Since the beginning ’LB’ and terminal ’LL’ are guaranteed, we may use only an abbreviated version consisting of n − 3 of each letter. B.1.4. Listing The problem is now reduced to generating all possible letter combinations. We can recursively build these combinations letter by letter using a greedy algorithm which chooses ’B’ over ’L’ if possible. This approach is subject to three restrictions which must be true of a “legal” word. At each step, (a) L ≤ B + 1, (b) B ≤ n − 3, (c) L ≤ n − 3. The function we use is sufficiently short that we simply reproduce it here: list_trees(n) { rec_list_trees(n, 0, 0, ’’) ; }

rec_list_trees(n, leaves, branches, word) {

157

max = n-3 ; if (branches < max) rec_list_trees(n, leaves, branches+1, word+’B’) ; if (leaves < max and leaves

Suggest Documents