Discrete-to-continuum limits of planar disclinations
Abstract
In materials science, wedge disclinations are defects caused by angular mismatches in the crystallographic lattice. To describe such disclinations, we introduce an atomistic model in planar domains. This model is given by a nearest-neighbor-type energy for the atomic bonds with an additional term to penalize change in volume. We enforce the appearance of disclinations by means of a special boundary condition.
Our main result is the discrete-to-continuum limit of this energy as the lattice size tends to zero. Our proof method is relaxation of the energy. The main mathematical novelty of our proof is a density theorem for the special boundary condition. In addition to our limit theorem, we construct examples of planar disclinations as solutions to numerical minimization of the model and show that classical results for wedge disclinations are recovered by our analysis.
1 Introduction
This paper is devoted to the mathematical analysis of a discrete model that describes frustrations in atomistic lattices induced by rotational mismatches. Such configurations are called wedge disclinations, which are angular defects. Disclinations are observed in solids in situations where the rotational symmetry is violated at the level of the crystal lattice. Figure 1 illustrates classical, simple examples of disclinations.
Historically, the existence of disclinations was predicted by Volterra alongside dislocations (translational defects) in a celebrated paper [Vol07]. However, it was not until the late 1960s that disclinations saw a systematical investigation both from an experimental and theoretical perspective. First examples of disclinations over planar lattices have been discovered in superconductors and reported in [TE68] and [ET67]. While a dislocation is a singularity of the deformation field which may be described by a lattice-valued vector, called Burgers vector, a disclination, as stated in [AEST68] (see also [Nab67]), is characterized by a closure failure of rotation … for a closed circuit round the disclination centre. A continuum theory for disclinations in linearized elasticity has been first derived by de Wit in [dW70] based on the idea of compatible elasticity and later elaborated by the same author in a series of articles [dW73a], [dW73b], [dW73c]. A comprehensive theory for disclinations (alongside dislocations) in non-linear elasticity has been developed by Zubov [Zub97]. We refer to [ZA18] for more recent developments on unified approaches to treat dislocations and disclinations as well as for a review of mechanical models of disclinations.
Our work is inspired by a number of experimental observations which have revealed formation of disclinations in metallic structures under a variety of mechanical loading and forces, geometrical regimes and kinematical constraints. Here, we elaborate on two such observations.
In austenite-to-martensite transformations, disclinations may emerge during the formation of rotated (and constant-strain) shear-bands [Bha03]. Such transformations are purely elastic and of the type solid-to-solid. They appear in a class of metals; in particular, in shape-memory alloys. Upon symmetry-break (typically triggered and driven by a negative temperature gradient), austenite, the high-symmetry and highly homogeneous crystal phase, turns into martensite, an anisotropic crystal phase with lower symmetry. The crystallographic lattice accommodates this phase change by forming a complicated microstructure composed of a mixture of thin plates and needle-shaped regions exhibiting differently rotated copied of martensitic phases. In a zero stress-microstructure, martensite can be described by a piecewise constant deformation gradient. Constant-strain regions, that is, regions of constant crystal orientation, are separated by sharp planar interfaces according to kinematical compatibility. This compatibility is called a rank-1 connection of the corresponding deformation tensors. However, such configurations are ideal and typically atomistic non-idealities such as dislocations and disclinations appear in large numbers. Outstanding examples of disclinations in martensite are represented by ”nested” star-shaped geometries observed in . Experiments are described in [KK91, MA80]; modeling work as well as numerical and exact constructions are described in [PL13] and [CPL14] respectively, and mathematical theories are developed in [CDPR+20]. Significantly more complicated microstructures rich in defects are described in [ILTH17] for Ti-Nb-Al-based alloys. Here, martensite nucleates and evolves in the form of thin plates embedded in an austenitic lattice. The evolution of these plates is complex due to plates colliding against surrounding structures, undergoing further branching into additional martensitic sub-plates or being reflected after hitting a grain boundary. Such evolution results in self-similar patterns resembling fractal structure which are rich in disclinations and dislocations [ILTH17]. The available models of such microstructures are essentially based on statistical analysis [IHM13] and probability [BCH15, CH], and thus the consistency with atomistic models remains elusive. This is where we aim to contribute.
The second experimental observation is the discovery of a superior (and, yet to date, largely unexplained) structural reinforcement mechanism which has recently spurred scientific interest on the experimental as well as on the theoretical investigation of kink formation in certain classes of metal alloys. Formation of kinks consisting of approximately constant-strain bands accompanied by high rotational stretches of the lattice are observed in classes of laminate ”mille-feuille” structures under uniaxial compression [HOI+16, HMH+16]. In one of their most typical morphologies, bands manifest themselves in the form of planar and sharp ridge-shaped regions which appear at various length-scales and are accompanied by localized plastic stresses and formation of disclinations [LN15]. Kinks of various morphologies have been described in [Ina19] where constructions of piecewise affine deformation maps based on the rank-1 connection rule and incorporating angular mismatches are presented. While [Ina19]’s analysis sheds light on the kinematics of the disclination-kinking mechanism for various morphologies of kinks, there is no available model based on atomistic descriptions which describes the energy of such disclination-kinking mechanisms.
A common aspect on both martensitic microstructures and kink formation is that planar regions of approximately constant strains and constant-orientation lattices need to rotate in order to preserve the continuity of the deformation field across their common border. As a result, these materials necessarily develop angular lattice misfits which are striking examples of wedge disclinations. This motivates the modeling assumption of this paper that wedge disclinations are caused by large (non-linear) rotational stretches in planar-confined geometries.
Our aim is to take the first step in the direction of a comprehensive variational theory that is suitable to simultaneously treat microscale and localized defects and to predict their effect on large (non-linear) elastic and plastic deformations including kinks and shear-bands. By designing a simple, nearest-neighbor-type interaction mechanism, we construct a model which we apply to describe a single disclination in a planar lattice and which is at the same time potentially adaptable to describe multi-disclination systems and to incorporate other lattice defects such as voids, dislocations and grain boundaries.
By pursuing this aim, we also fill a gap in the literature on atomistic modeling of planar lattice defects and the limit thereof as the lattice spacing tends to . To identify this gap, we review related literature. First, in the framework of linearized elasticity, the formal asymptotic expansions in [SN88] provide a continuum model for lattice defects. Second, in [LPS13] two triangular lattices with different lattice spacings are attached together, which forces dislocations to form at the interface. The main result is a continuum limit of this model as the lattice spacing tends to 0. Third, based on the discrete calculus of lattices constructed in [AO05], the stress in a periodic crystal induced by parallel screw dislocations is computed in [Pon07, HO14, HO15, EOS16, BBO19]. These results provide quantitative estimates between the displacement in the lattice and the displacement computed from linear elasticity in the continuum counter part. In [BHO19] these techniques are extended to capture cracks. All these works deal with either localized or small deformation to the underlying two dimensional lattice, which is unfit for disclinations (see, e.g., the relatively large deformations in Figure 1).
For large deformations, the recent work in [KM18] presents and anlyzes an atomistic model of a stretchable hexagonal lattice defined over a smooth manifold, in which the main result is the continuum limit in the form of -convergence. Since the approach in [KM18] is designed to describe the energetics of highly distorted membranes, we follow a similar approach. However, the choice in [KM18] that the number of bonds per atom is always makes the result not applicable to disclinations.
As in [KM18], we assign an energy to a deformed lattice, where is the lattice spacing. Two examples of such deformed lattices are illustrated by the dotted nodes in Figure 1. Given a deformed lattice, the energy penalizes stretching and compression of the atomic bonds and changes in volume of the triangles formed by three neighboring atoms. To enforce the appearance of a - or -type disclination, we require the deformed lattice to satisfy a special boundary condition such that or rotated copies of it fit together such as in Figure 1.
Our main result, Theorem 3.1, is the exact derivation of , the macroscale energy obtained in the limit as , that is,
in a suitable topology.
From the mathematical perspective, the interest of our analysis lies in the treatment of the special boundary condition. Conceptually, this boundary conditions requires us to incorporate a pointwise constraint in the space of traces on -Sobolev vector maps which are characterized by means of a nonlocal norm. Our main contribution on this is Proposition 4.4.
The paper is organized as follows. In Section 2 we present the discrete mechanical model and the mathematical setting required for the related variational analysis. In Section 3 we state and prove our main result, Theorem 3.1, on the -convergence of the lattice energy . In this proof, we postpone the proofs of technical lemmas to Section 4, which includes the proof of Proposition 4.4 on density in the space of admissible lattice displacements. In Section 5 we explore the physical implications of our -convergence analysis in terms of energy and stress states of the minimizers of the continuum model. Finally in Section 6 we present several numerical realizations of both positive and negative disclinations.
2 The lattice energy
Here we define the atomic lattice energy as briefly introduced in the introduction. We start with the kinematics. Inspired by Figure 1, we consider a two-dimensional model. This corresponds conceptually to the mid-section of a 3-dimensional body. Furthermore, we impose rotational symmetry so that we may confine the domain to a single wedge in Figure 1 indicated by the black dots. The reference domain is given by the equilateral open triangle of size 1 with boundary as depicted in Figure 2. We take the reference positions of the atoms as a triangular lattice , where is such that it fits on top of as in Figure 2, i.e., the lattice spacing is such that and is positioned such that each closed line segment fits on top of the atoms in one of the lattice directions. We denote by
(1) |
the set of the six outward-pointing bonds in from any lattice point, where is the counter-clockwise rotation matrix by angle . For later use, we sometimes interpret as a planar graph , where is the set of all vertices , is the set of all edges between neighboring vertices , and is the set of all open triangles with sides given by the three edges .
The set of admissible displacements is given by
(2) |
where is the angle associated with a - or -type disclination (see Figure 1). Since is a finite set, the map can be identified with a vector in . The boundary condition in (2) is such that the rotated copy of fits seamlessly to . Adding more rotated copies, we obtain a deformed lattice with rotational symmetry such as that in Figure 1, which has a - or -type disclination at the origin.
Next we define the lattice energy on as inspired by [KM18]. We start with a formal description. Given , we set the lattice energy formally as
(3) |
where the potential
penalizes atomic bonds which are not of length (the parameter value corresponds to linear elasticity), the weight function
(4) |
counts the outer edges as half111we model edges as part of the volume of the medium around them., and the potential penalizes change in volume of the triangles (here, is the volume of , and is the volume of the triangle after applying the displacement ), especially if the volume of gets inverted under . Note that while the identity map minimizes , it is not in because of the boundary condition. Hence, the boundary condition enforces mechanical frustration. For and , Figure 1 illustrates the deformed lattice of a local minimizer in of which does not contain negative change in volume. In this paper we always assume unless specified otherwise.
Since the term involving is not derived from physical principles, we elaborate on this modelling choice. It is well-known that when , lattice energies with only nearest-neighbor interactions such as do not penalize folding or any other negative change in volume, and as a result, the related continuum energy may not penalize compression. While our boundary condition in (2) is not standard, we show in Section 6 by means of numerical simulations that folded patterns appear as local minimizers of when .
To penalize folding, we pose the following minimal requirements on :
-
()
;
-
()
, and .
Condition () is a continuity estimate, which implies both local Lipschitz continuity and -growth. Condition () ensures that any change in volume is penalized. A simple example of is . These conditions are more general than those in [KM18, Section 3.3].
The term in related to has an inconvenient form. We fix this by changing variables. The result is a rigorously defined energy functional , which we will use in the remainder of the paper.
We change variables by describing the state as the deformation of a displacement defined on . Given , we set as the piecewise linear extension of to , i.e.,
We note that
(5) |
where is any lattice bond of unit direction (see (1)), and the indices depend on . In particular, is constant on each ,
is the relative change in volume of under , and the sign of the determinant determines whether the volume of is inverted under . From these observations, we define the second term in (3) rigorously by
(6) |
Next we rewrite and in terms of . This yields
Using (5) and (6), we rewrite (3) as
(7) |
where
(8) |
In the computation above, the weight function turns into the factor due to the fact that each edge in the interior of borders two triangles in . Eq. (7) motivates us to define
(9) |
Remark 2.1 (Properties of ).
We note that and that any satisfies . While does not contain a subspace of constants, it does contain a subspace of linear maps . A possible choice for is the one that satisfies and .
3 Continuum limit
Having introduced the lattice energy on the triangular lattice, we are now in a position to discuss the continuum limit as . To keep track of the asymptotic behavior of minima and minimizers of we characterize the continuum model with -convergence [DM93].
Let . The domain of the continuum energy is
Observe that is linear (and, in particular, convex), and non-empty since . Moreover, if , there holds for any . Thanks to the properties of traces (see Lemma 4.6 below), we have that is strongly closed and, therefore, weakly closed as well thanks to convexity.
The continuum energy functional is given by
(10) |
where is as in (8) and is the quasiconvex envelop of defined by
(11) |
where being quasiconvex means that is Borel measurable, locally bounded and satisfies
for any bounded open set , any and any .
Theorem 3.1.
For and , (see (9)) -converges as to in the strong topology.
We prove Theorem 3.1 in Section 3.2. Since the -dependence of appears only in its domain , proving a -limit result reduces to proving a relaxation result, i.e., finding the lower semi-continuous envelope of
in the right functional framework. There is a large literature on such relaxation problems; in Section 3.1 we cite the relevant classical theory. While this theory gives a useful roadmap for proving Theorem 3.1, it does not capture Theorem 3.1 because of the periodic boundary condition in . Therefore, in Section 3.2 we give the skeleton of the proof of Theorem 3.1 based on the classical theory, and identify the missing steps as technical lemmas which we prove in Section 4.
3.1 Classical relaxation result on
We review some classical relaxation results as preparation for proving Theorem 3.1. All theorem references below in this section refer to Dacorogna’s book [Dac08]. Another relevant reference is [AF84].
We recall that is a bounded Lipschitz domain. Here and in what follows we adopt the Frobenius norm for matrices. Let satisfy
-
()
-growth. ;
-
()
Continuity estimate. .
Note that () includes a uniform bound from below, and that () provides a local Lipschitz estimate. We set
Here and henceforth, we remove the range from the notation of the function space if there is no danger for confusion.
As preparation, we cite Theorem 6.9 for the alternative characterization of as the quasiconvexification of given by
(12) |
where is a subset of with (i.e., has zero two-dimensional volume).
Since satisfies (), Theorem 9.1 implies that for all there exists a sequence such that
and
(13) |
By the two properties of , it follows from Theorem 6.9 and Theorem 5.3(iv) that is continuous. Then, Theorem 1.13 implies that is sequentially weakly lower semicontinuous in . In particular, for any converging weakly in to some , we have that
(14) |
3.2 Proof of Theorem 3.1
The proof below relies on the following three statements which we make precise and prove in Section 4:
The proof of Theorem 3.1 follows from matching a lower bound with an upper bound, which is the standard method for computing -limits [DM93]. To identify the set where the limit functional is finite, we first investigate the (equi-)compactness of minimizing sequences.
Proof of Theorem 3.1.
Compactness. By the lower bound in (),
for some constants independent of and , and thus any finite-energy sequence is bounded in . By the Poincaré Inequality (Lemma 4.1), we then infer that is bounded in , and thus strongly convergent (along a subsequence) in .
-liminf. Since we can focus on finite-energy sequences in , the compactness statement implies that in as . Since , which is a closed subspace of , we also have . Hence, by applying (14), we obtain the required -liminf estimate.
4 Technical steps in the proof of Theorem 3.1
Here we state rigorously and prove the three statements mentioned at the start of Section 3.2. We start with the Poincaré Inequality.
Lemma 4.1 (Poincaré Inequality on ).
There exists such that for all it holds that
Proof.
We follow a standard proof by contradiction. Assuming that there exists such that
we obtain by compactness that converges along a subsequence (not relabelled) to strongly in and weakly in . Hence, and . Since is connected, for some constant and non-zero vector . However, for a.e. ,
Hence, , which completes the proof. ∎
4.1 Properties of
Proof.
If , then () is obvious since . Then, since by (), the lower bound in () is immediate. The upper bound follows by () from
() follows by () from
∎
For later use (although not necessary for the proof of Theorem 3.1) we elaborate on the rigidity properties of the energy density . The rigidity estimate is a direct consequence of Assumptions (W1), (W2). Lengthy computations are postponed to the Appendix.
We start with introducing the singular value decomposition
(16) |
In (16), are the ordered singular values of , that is, the square roots of the eigenvalues of the matrix . We also recall the definition of the distance function
Lemma 4.3.
For defined in (8), there exists such that for all
(17) |
Proof.
Let be the singular decomposition as in (16). We split two cases; and . We start with the first case. From the definition of we get
(18) |
where is the unit vector, is the rotation matrix, and is fixed by . We first consider the case . Thanks to Lemma A.1 we have
Recalling the well-known relation
and noting that implies , we obtain (17). For the case , applying Jensen’s inequality in (18) yields
Then, (17) follows from the argument above.
We continue with the second case, . By in Section 3.1, there exist independent of such that . We separate two cases:
- 1.
-
2.
If , we note from the triangle inequality
that
for some independent of .
∎
4.2 Density of in
Proposition 4.4 (Density of in ).
For all there exists parametrized by such that as .
Before giving the proof of Proposition 4.4 at the end of this section, we first outline the idea of the proof, and then establish some technical lemmas.
In order to explain the idea of the proof, we first recall two classical density results in the following lemma. To state it, we define .
Lemma 4.5 (Density of ).
For any with either or there exists parametrized by such that as .
Proof.
This is a standard result in numerical analysis; see e.g. [ET99, Chap. X, Prop. 2.1, 2.6 and 2.9]. We give the details of the proof to show how our boundary condition fits in.
For , it is not restrictive by density to assume that . Then, setting with piecewise affine continuation, it is obvious that and that uniformly as .
For , we set again with piecewise affine continuation, and define . Since , we have , and thus . For the gradient, we split , where the disjoint sets and are such that and has volume that vanishes as . Then,
First, since , we have by the argument above that , and thus as . Second, since , , and thus uniformly in . Hence, by taking small enough with respect to and , we conclude that as . ∎
Thanks to Lemma 4.5, the proof of Proposition 4.4 narrows down to constructing a decomposition where
(19) |
Indeed, if such a decomposition exists, then Lemma 4.5 provides approximations in of and , and can simply be approximated by . The difficulty in constructing such a decomposition is in finding a for which is sufficiently close to . Approximation by convolution does not work directly since has to satisfy the boundary condition in . Instead, we use the Trace Theorem to approximate in an appropriate function space on . This approximation is based on convolution, but care is needed because of the boundary condition, the corners of and the fact that the norm of the function space on is nonlocal.
Next, we prepare for proving Proposition 4.4 by citing a Trace Theorem (Lemma 4.6) and proving a density result on (Lemma 4.9). To avoid technical difficulties with the corners in , we first transform to the unit disc . With this aim, let be a related transformation (see Figure 3) such that
-
•
is bi-Lipschitz;
-
•
;
-
•
for , where are given, in terms of the polar angle coordinate, by
-
•
if , then
Above and in the following, we will often identify the unit circle with the periodic interval . We also adopt the convention that a subscript in a function space indicates the boundary condition. We note that there are constants such that for all
Hence, it is sufficient to construct the decomposition of after transforming it to .
To cite the Trace Theorem, we fix and recall the usual norm of the fractional Sobolev space :
Lemma 4.6 (Trace [Gag57, Thm. 1.I]).
There exists a such that for all
Conversely, there exists a such that for all there exists a with such that
In order to prove a density result in , we first recall two technical lemmas:
Lemma 4.7 ([DNPV12], Lem. 5.2).
If , then the even extension satisfies
for some universal constant .
Lemma 4.8 (Continuity of the translation operator).
The translation operator is continuous in the strong topology, that is, for all ,
Proof.
The proof is standard; we give a sketch. The statement is obvious for . For general , it suffices to approximate it by , and note that
∎
Lemma 4.9 (Approximation on ).
is dense in .
Proof.
Let . We split with such that
Note that with , and that it is not directly clear that . We will construct approximating sequences of and respectively such that . By construction of , is a suitable approximating sequence, where is the usual mollifier.
To construct , we first show that . Since by construction , it is sufficient to show that for some . We start with . For any and any parameter , let
Clearly, for any . Since for a.e. and , we also have . Since is the even extension of , we obtain from Lemma 4.7 that . Hence, for the odd extension , we find from the linear relation
that . Finally,
The proof of is analogous.
Finally we construct the approximating sequences . Care is needed to ensure that satisfies the boundary condition. With this aim, we first approximate and by the translations and with ; see Lemma 4.8. Note that and . Then, we define for some such that , and note that . This completes the proof. ∎
We are ready to prove Proposition 4.4.
Proof of Proposition 4.4.
Step 1: decomposition of . Let and be given. Set and . Then, by Lemma 4.6 we have . By Lemma 4.9 we find with . By Lemma 4.6, there exists with and
for some independent of .
Next, we take as the harmonic extension of . Then, translating back to , we set for . Taking , we observe that satisfy (19) for .
Step 2: Construction of . By Lemma 4.5 we find sequences parametrized by such that and
for . Hence, setting and taking small enough with respect to , we obtain
for some independent of , or . Since is arbitrary, we conclude that as . ∎
5 Physical interpretation of Theorem 3.1
We show that, despite the relaxation process implied by -convergence, the equilibrium solution of the model are necessarily stressed. This is a consequence of the rotational boundary conditions incorporated into and of the finite penalization to folding imposed by .
Proposition 5.1.
(20) |
The proof of Proposition 5.1 follows the lines of [KM18, Sec. 5]. For the readers’ convenience, we display the main steps.
Proof.
First, we show
(21) |
for some . From Lemma 4.3, we have for every . Now, thanks to (12), for every and every smooth open with , there exists a map (in fact even ) such that
We now invoke Rigidity Theorem 3.1 [FJM02] (which applies for , see [CS06, Sec. 2.4]) yielding the existence of a constant and a matrix such that
(22) |
Now, interpreting as a convex function on , we obtain
(23) |
for all . By applying (23) to (22) we have
because the integral of vanishes as has zero boundary datum. Putting our estimates together and minimising over , we get
and the desired result in (21) follows by sending .
6 Numerical computations
In this section we explore numerically several energy wells of and inspect whether the obtained minimizers satisfies . We do this for the simplified setting given by , which does not penalize folded patterns. Fortunately, this simplified setting turns out to be stable enough to reproduce configurations with (see, e.g., Figure 1) as local minima. Yet, local minimizers are of limited use to our theoretical result Theorem 3.1, because -convergence only guarantees the convergence of global minimizers in the limit . Instead, we show by direct inspection of the energy values at the computed local minimizers that there is a good agreement with available continuum models for disclinations for decreasing values of . We compute local minimizers by employing Newton’s method for different initial conditions. In all computations below, Newton’s method converges quadratically.
6.1 The limit
Here we verify and quantify some of the computations in [SN88], in which disclinations satisfying are obtained. The only difference with [SN88] is in the definition of in (4); in [SN88] is taken instead.
For both and we chose the initial condition such that the local minimizer of satisfies . With this aim, for we took as initial condition one of the linear deformations in with (see Remark 2.1). For the subsequent values , we constructed the initial condition from the minimizer obtained for the previous value of by linear interpolation. The resulting local minimizers turn out to satisfy . For , is illustrated in Figure 1. The energy values are plotted in Figure 4. The behavior of as is qualitatively similar to the computations of [SN88] which are based on a model in linearized elasticity.
Next, we test the convergence of as . Indeed, since need not be global minimizers, the -convergence result in Theorem 3.1 does not guarantee convergence. To test the convergence, we impose the power law ansatz
(24) |
where the constants need to be fitted from the data. Without computing and , we obtain the exponent of the power law numerically from
(25) |
Table 1 lists the values of . Since the values of are positive and increasing with , seems to converge as implying a power law behavior.
1.552 | 1.276 | |
1.712 | 1.488 | |
1.812 | 1.595 | |
1.866 | 1.645 | |
1.898 | 1.671 | |
1.918 | 1.686 |
6.2 Folded configurations
In the next simulations we explore other local minimizers of the energy by starting Newton’s method from other initial conditions. In particular, since we take , there is no penalisation on , and thus local minimizers with folded patterns may occur.
With this aim, we set and fold the reference lattice as illustrated in Figure 5. This folding procedure is done such that the following property used in the boundary condition in (2)
is conserved during the folding process. After folding, we construct the initial condition for Newton’s method by deforming the folded reference domain by the linear map defined in Remark 2.1.
Figures 7 and 8 illustrate the local minimizers obtained for each folded pattern shown in Figure 5 for and respectively. Figure 6 shows the related energy values. It is clear that the folded local minimizers have lower energy. In fact, it seems that the energy values decrease in an affine manner with the number of folds, and that, after linear extrapolation, energy would be obtained around 4 folds.
The observation that has folded patterns as local minima with low energy is not unexpected from the expression of the continuum energy . Indeed, implies that the minimization problem of lacks rigidity; Lemma 4.3 and Proposition 5.1 are not valid anymore. In fact, if we substitute into (8), we obtain (see Lemma A.2) , which implies that is a global minimizer of . This is consistent with the observation in Figure 6, where the energy values decay with the number of folds.
7 Future perspectives
We develop a continuum model (the energy in (10)) for the energy of a disclination in a planar geometry. We construct rigorously from an atomistic model (Theorem 3.1), and show that any minimizer of corresponds to a stressed configuration of the medium (Proposition 5.1). We intend this first rigorous study as a stepping stone towards a general variational theory capable of describing both morphology and energetics of disclinations and their effect on the macrosocopic properties of a material. By analyzing the interaction of mismatches and distortions on the hexagonal lattice, we take a first step towards the investigation of the interaction of defects with the lattice kinematics, thus opening a possible path to modeling more complex systems such as austenite-martensite microstructures in Shape-Memory Alloys and kink formations as mentioned in the introduction.
Acknowledgements
P.C. is supported by JSPS Grant-in-Aid for Young Scientists (B) 16K21213 and by JSPS Innovative Area Grant 19H05131. P.C. holds an honorary appointment at La Trobe University and is a member of GNAMPA. The authors acknowledge the Research Institute for Mathematical Sciences, an International Joint Usage and Research Center located in Kyoto University, where part of the work contained in this paper was carried out.
Appendix A Appendix
Lemma A.1.
For all and all ,
(26) |
Proof.
We separate 2 cases: and . In the first case, we start with bounding the right-hand side of (26) from above. If , then clearly . If , then from the observation that the line segment between the points and on the circle is below the arc of between the same two points, we obtain that . Hence,
This together with the bound for we get
(27) |
We continue by bounding the left-hand side in (26) from below. Writing
we note from the facts that is -periodic and that there are at least two values for for which . For these values of , we have by and that
and thus
(28) |
Estimating the convex function from below by its tangent at yields
Plugging this estimate into (28), we observe that the resulting lower bound equals the upper bound in (27). This completes the proof in the case .
In the second case, , we follow a similar procedure. We claim that . Then, instead of (27), we get
(29) |
To prove this claim, we note from that . To get an upper bound for , we obtain from that
Since the right-hand side is concave in , we can bound it from above by its tangent at , this yields , and thus . The claim follows.
For the left-hand side of (26), similar to the previous case, there are at least two values for for which . For these values of , we have by , and that
and thus
(30) |
Using that , we obtain
Plugging this estimate into (30), we observe that the resulting lower bound is larger than the upper bound in (29). ∎
Lemma A.2.
Proof.
Since for all , it follows . We are left to show the reverse inequality. To do this, we introduce , the rank-1 convex envelope of (see [Dac08, Sec. 6.4]) which is characterized by the following formula ([Dac08, Thm. 6.10])
(31) |
where is the hierarchical compatibility constraint defined in [Dac08, Definition 5.14]. Recall also that for , then (see [Dac08, Sec. 6.1]). Hence, it remains to show that .
To show that , we take in (31) , and consider the family
and observe that
-
1)
;
-
2)
;
-
3)
;
-
4)
, where and .
By [Dac08, Example 5.15], Properties imply that satisfy condition . Then, it is then easy to see from 1) that constitute an admissible candidate in the minimization problem (31). Finally, condition implies for
Collecting all the results above, we have
∎
References
- [AEST68] K. Anthony, U. Essmann, A. Seeger, and H. Trauble. Disclinations and the Cosserat-Continuum with Incompatible Rotations, volume Mechanics of Generalized Continua, Proceedings of the IUTAM-Symposium on The Generalized Cosserat Continuum and the Continuum Theory of Dislocations with Applications, Freudenstadt and Stuttgart (Germany) 1967, pages 355–358. Springer-Verlag Berlin Heidelberg, 1968.
- [AF84] E. Acerbi and N. Fusco. Semicontinuity problems in the calculus of variations. Archive for Rational Mechanics and Analysis, 86(2):125–145, 1984.
- [AO05] M. P. Ariza and M. Ortiz. Discrete crystal elasticity and discrete dislocations in crystals. Archive for Rational Mechanics and Analysis, 178(2):149–226, 2005.
- [BBO19] J. Braun, M. Buze, and C. Ortner. The effect of crystal symmetries on the locality of screw dislocation cores. SIAM Journal on Mathematical Analysis, 51(2):1108–1136, 2019.
- [BCH15] J. Ball, P. Cesana, and P. Hambly. A probabilistic model for martensitic avalanches. MATEC Web of Conferences, 33, 2015.
- [Bha03] K. Bhattacharya. Microstructure of martensite: why it forms and how it gives rise to the shape-memory effect. Oxford University Press, 2003.
- [BHO19] M. Buze, T. Hudson, and C. Ortner. Analysis of an atomistic model for anti-plane fracture. Mathematical Models and Methods in Applied Sciences, 29(13):2469–2521, 2019.
- [BJ87] J. Ball and R. James. Fine phase mixtures as minimizers of eenergies. Arch. Ration. Mech. Anal., 100:13–52, 1987.
- [CDPR+20] P. Cesana, F. Della Porta, A. Rueland, C. Zillinger, and B. Zwicknagl. Exact constructions in the (non-linear) planar theory of elasticity: From elastic crystals to nematic elastomers. Archive for Rational Mechanics and Analysis, 237(1):383–445, 2020.
- [CH] P. Cesana and P. Hambly. A probabilistic model for interfaces in a martensitic phase transition. Preprint.
- [CPL14] P. Cesana, M. Porta, and T. Lookman. Asymptotic analysis of hierarchical martensitic microstructure. Journal of the Mechanics and Physics of Solids, 72, 2014.
- [CS06] S. Conti and B. Schweizer. Rigidity and gamma convergence for solid‐solid phase transitions with SO(2) invariance. Commun. Pure Appl. Math., 59(06):830–868, 2006.
- [Dac08] B. Dacorogna. Direct Methods in the Calculus of Variations. Springer, Heidelberg, 2nd edition, 2008.
- [DM93] G. Dal Maso. An Introduction to -Convergence. Birkhäuser Verlag, Boston, 1993.
- [DNPV12] E. Di Nezza, G. Palatucci, and E. Valdinoci. Hitchhiker’s guide to the fractional Sobolev spaces. Bulletin des Sciences Mathématiques, 136(5):521–573, 2012.
- [dW70] R. de Wit. Linear theory of static disclinations. In J. A. Simmons, R. de Wit, and R. Bullough (eds.) Fundamental Aspects of Dislocation Theory, volume 317, pages 651–673. Nat. Bur. Stand. (US), Spec. Publ., 1970.
- [dW73a] R. de Wit. Theory of disclinations: II. continuous and discrete disclinations in anisotropic elasticity. Journal of Research of the National Bureau of Standards - A, Physics and Chemistry, 77A(1), 1973.
- [dW73b] R. de Wit. Theory of disclinations: III. continuous and discrete disclinations in isotropic elasticity. Journal of Research of the National Bureau of Standards - A, Physics and Chemistry, 73A(3), 1973.
- [dW73c] R. de Wit. Theory of disclinations: IV. straight disclinations. Journal of Research of the National Bureau of Standards - A Physics and Chemistry, 77A(5), 1973.
- [EOS16] V. Ehrlacher, C. Ortner, and A. V. Shapeev. Analysis of boundary conditions for crystal defect atomistic simulations. Archive for Rational Mechanics and Analysis, 222(3):1217–1268, 2016.
- [ET67] D. Essmann and H. Träuble. The direct observation of individual flux lines in type ii superconductors. Phys. Letters, 24A(526), 1967.
- [ET99] I. Ekeland and R. Temam. Convex Analysis and Variational Problems, volume 28. SIAM, 1999.
- [FJM02] G. Friesecke, R. James, and S. Müller. A theorem on geometric rigidity and the derivation of non linear plate theory from three dimensional elasticity. Commun. Pure Appl. Math., 55:1461—1506, 2002.
- [Gag57] E. Gagliardo. Caratterizzazioni delle tracce sulla frontiera relative ad alcune classi di funzioni in n variabili. Rend. Sem. Mat. Univ. Padova, 27(405):284–305, 1957.
- [HMH+16] K. Hagihara, T. Mayama, M. Honnami, M. Yamasaki, H. Izuno, T. Okamoto, T. Ohashi, T. Nakano, and Y. Kawamura. Orientation dependence of the deformation kink band formation behavior in Zn single crystal. International Journal of Plasticity, 77:174–191, 2016.
- [HO14] T. Hudson and C. Ortner. Existence and stability of a screw dislocation under anti-plane deformation. Archive for Rational Mechanics and Analysis, 213(3):887–929, 2014.
- [HO15] T. Hudson and C. Ortner. Analysis of stable screw dislocation configurations in an antiplane lattice model. SIAM Journal on Mathematical Analysis, 47(1):291–320, 2015.
- [HOI+16] K. Hagihara, T. Okamoto, H. Izuno, M. Yamasaki, M. Matsushita, T. Nakano, and Y. Kawamura. Plastic deformation behavior of -type synchronized LPSO phase in a Mg-Zn-Y system. Acta Materialia, 109:90–102, 2016.
- [IHM13] T. Inamura, H. Hosoda, and S. Miyazaki. Incompatibility and preferred morphology in the self-accommodation microstructure of -titanium shape memory alloy. Philosophical Magazine, 93(6):618–634, 2013.
- [ILTH17] T. Inamura, M. Li, M. Tahara, and H. Hosoda. Formation process of the incompatible martensite microstructure in a beta-titanium shape memory alloy. Acta Materialia, 124:351–359, 2017.
- [Ina19] T. Inamura. Geometry of kink microstructure analysed by rank-1 connection. Acta Materialia, 173:270–280, 2019.
- [KK91] Y. Kitano and K. Kifune. HREM study of disclinations in MgCd ordered alloy. Ultramicroscopy, 39:279–286, 1991.
- [KM18] R. Kupferman and C. Maor. Variational convergence of discrete geometrically-incompatible elastic models. Calculus of Variations and Partial Differential Equations, 57, 2018.
- [LN15] X.W. Lei and A. Nakatani. A deformation mechanism for ridge-shaped kink structure in layered solids. Journal of Applied Mechanics, 82:071016, 1–6, 2015.
- [LPS13] G. Lazzaroni, M. Palombaro, and A. Schlömerkemper. Dislocations in nanowire heterostructures: from discrete to continuum. Proceedings in Applied Mathematics and Mechanics, 13(1):541–544, 2013.
- [MA80] C. Manolikas and S. Amelinckx. Phase transitions in ferroelastic lead orthovanadate as observed by means of electron microscopy and electron diffraction. Phys. Stat. Sol., 60(2):607–617, 1980.
- [Nab67] F.R.N. Nabarro. Theory of crystal dislocations. International Series of Monographs on Physics. Oxford: Clarendon Press, 1967.
- [PL13] M. Porta and T. Lookman. Heterogeneity and phase transformation in materials: Energy minimization, iterative methods and geometric nonlinearity. Acta Materialia, 61(14):5311–5340, 2013.
- [Pon07] M. Ponsiglione. Elastic energy stored in a crystal induced by screw dislocations: from discrete to continuous. SIAM Journal on Mathematical Analysis, 39(2):449–469, 2007.
- [SN88] S. Seung, H. and R. Nelson, D. Defects in flexible membranes with crystalline order. Physical Review A, 38:1005, 1988.
- [TE68] H. Träuble and D. Essmann. Fehler im flussliniengitter von supraleitern zweiter art. Phys. Stat. Sol., 25(373), 1968.
- [Vol07] V. Volterra. Sur l’équilibre des corps élastiques multiplement connexes. Annales scientifiques de l’École Normale Supérieure, 24:401–517, 1907.
- [ZA18] C. Zhang and A. Acharya. On the relevance of generalized disclinations in defect mechanics. Journal of the Mechanics and Physics of Solids, 119:188–223, 2018.
- [Zub97] L. M. Zubov. Nonlinear Theory of Dislocations and Disclinations in Elastic Bodies. Lecture Notes in Physics Monographs. Springer, 1997.