remarkRemark \newsiamremarkhypothesisHypothesis \newsiamthmclaimClaim \headersWell-posedness of Cahn-Hilliard model for surface diffusionNiu, Xiang,Yan
Well-posedness of a modified degenerate Cahn-Hilliard model for surface diffusion
Abstract
We study the well-posedness of a modified degenerate Cahn-Hilliard type model for surface diffusion. With degenerate phase-dependent diffusion mobility and additional stabilizing function, this model is able to give the correct sharp interface limit. We introduce a notion of weak solutions for the nonlinear model. The existence result is obtained by approximations of the proposed model with nondegenerate mobilities. We also employ this method to prove existence of weak solutions to a related model where the chemical potential contains a nonlocal term originated from self-climb of dislocations in crystalline materials.
keywords:
Phase field model, degenerate Cahn-Hilliard equation, surface diffusion, well-posedness, weak solutions35A01, 35G20, 35K25, 74N20, 82C26
1 Introduction
We consider the following modified degenerate Cahn-Hilliard type model
(1) | |||||
(2) |
When , (1)-(2) becomes Cahn-Hilliard (CH) equation with degenerate mobility. The degenerate Cahn-Hilliard equation has been widely studied as a diffuse-interface model for phase separation in binary system [2, 3, 4, 5, 7, 10]. Over the years, the interface motion in the sharp limit has caught a lot of attention for various choice of mobility and homogeneous free energy . When and being either the logarithmic free energy
with temperature or the double obstacle potential
Cahn, Elliott, and Novick-Cohen [19] showed via asymptotic expansions that the sharp-interface limit in the time scale is interface motion by surface diffusion. Sharp interface limits for different time scales were discussed in [5] for highly disparate diffusion mobility and smooth double well . In particular, the system evolves in time scale according to the combination of a one-sided modified Mullins–Sekerka problem in the phase with nonzero constant mobility and a nonlinear diffusion process that solves a quasi-stationary porous medium equation in the phase with small mobility. A later work by the same authors [6] derived sharp interface limit for time scale with diffusion mobility and smooth double well potential , noting the effect of the diffusion field on the interface motion as a jump of fluxes. The analysis was done on the (unphysical) solution branch with on some region. For and , Lee, Münch and Süli [14] considered the physical branch of solution where everywhere and showed that there is an additional nonlinear bulk diffusion term appearing to leading order of the sharp interface limit. Further study in [15] indicates that the leading order sharp-interface motion depends sensitively on the choice of mobility.
The existence of weak solutions for degenerate Cahn-Hilliard equation was proved by Elliotte and Garcke [10] (see [24] for 1D case). Their results include the case and being the logarithmic free energy. Dai and Du [7] introduced a different notion of weak solutions for degenerate Cahn-Hilliard equation with mobility and smooth double well potentials; they showed that their model accommodates the Gibbs-Thomson effect, which was not by the method in [10].
There is a critical issue in modeling surface diffusion by the degenerate Cahn-Hilliard model [11, 21], due to the presence of incompatibility in the asymptotic matching between the outer and inner expansions. Rätz, Ribalta, and Voigt (RRV) [21] fixed this incompatibility by introducing a singular factor in front of the chemical potential to force it to vanish in the far field. Their model essentially consists of equations (1)-(2) without the term on the left side of (1), and other terms for modeling heteroepitaxial growth of thin films. The RRV model with the stabilizing function has been validated by numerical simulations [21] and asymptotic analyses [11, 21]. It has been successfully generalized to many applications, e.g., growth of nanoscale membranes [1], dewetting of ultrathin films [17], and grain boundary formation in nanoporous metals [9]. Recently, a phase field model for dislocation self-climb by vacancy pipe diffusion was developed based on degenerate Cahn-Hilliard model with such stabilizing function [20]. However, to the best of our knowledge, well-posedness of these degenerate Cahn-Hilliard models with singular factor that give the correct sharp interface limit for surface diffusion has not been established in the literature.
In this paper, in order to prove the well-posedness of the RRV type Cahn-Hilliard model with correct sharp interface limit for surface diffusion, we propose a modified degenerated Cahn-Hilliard model as given in (1)-(2), and discuss its well-posedness and sharp interface limit. In particular, we have modified the original RRV model so that the equation can be written in the form of gradient flow of the total energy.
Our first result is a sharp interface limit equation for (1) and (2) via formal asymptotic analysis. We obtain the following sharp interface equation
(3) |
as . Here , are constants whose exact forms are derived in section 2. This validates this equation as a diffuse-interface model for surface diffusion.
Our main result concerns the well-posedness of the initial value problem of (1)-(2). For this purpose, we set and consider the following problem in a periodic setting when .
(4) | |||||
(5) |
Here for , for some constant and satisfies the following assumptions.
-
(i)
and there exist constants , such that for all and some , the following growth assumptions hold.
(6) (7) (8)
We see that the classical double well potential satisfies (6)-(8) with .
Our existence proof is obtained via approximations of the proposed model (4)-(5) with positive mobilities. Given any , we define
(9) |
with
(10) |
Our first step is to find a sufficiently regular solution for (4)-(5) with mobility and stablizing function together with a smooth potential .
Theorem 1.1.
Proof of theorem 1.1 is based on Galerkin approximations. Due to the presence of the stablizing function , it is not obvious how to pass to the limit in the nonlinear term of the Galerkin approximations. Our key observation in this step is strong convergence of (up to a subsequence) in where which allows us to pass to the limit in the nonlinear term.
To obtain the weak solution to (4), we consider the limit of for a sequence . The key challenge is how to pass to the limit in both sides of (11). In the degenerate Cahn-Hilliard case, the estimates for the positive mobility approximations yield a uniform bound for and it is straighforward to pass to the limit on the left hand side in the approximating equations. Moreover, the bound on , together with bound on yields strong convergence of in . By this and the weak convergence of in , Dai and Du [7] showed (up to a subsequence) that weakly in where is the weak limit of . The main task left is to show and the limit equation becomes a weak form Cahn-Hilliard equation. They [7] proved that this is almost true in the set where . Their main idea is the following. For small numbers monotonically decreasing to , they consider the limit in a subset of where approximate solutions converges uniformly and . By decomposing where mobility is bounded from below uniformly in and controlled above in by suitable multiples of , they obtain the weak form equation for the limit function by passing to the limit of on then letting goes to . Under further regularity assumptions on , they obtained the explicit expression for in the weak form of the equation.
In this paper, we adapt their idea to our model. There are two main difficulties. The first obtacle is the bound estimate on blows up when goes to zero and we can not pass to the limit on the left hand side of (11); secondly, due to the presence of the stablizing function on the right hand side, it is more complicated to derive an explicit expression of the weak limit of in terms of on the right hand side of the limit equation. To overcome the first difficulty, we derive an alternative form of (11) by multiplying to both sides (valid due to regularity of , c.f. section 3.4 and equation (85)). From this, we obtain uniform estimates on which enables us to pass to the limit on the left hand side of the alternate equation (85). To find limit form on the right hand side of (85), we need convergence of in . Due to the lack of control on , such convergence can not be derived directly using Aubin-Lions Lemma [7]. Instead, we apply Aubin-Lions lemma to and derive convergence of (consequently on ) from convergence of through characterization of compact sets [22] in . We then follow the idea in [7] to pass to the limit on the right hand side of (85). Finally, we identify an explicit expression of the weak limit of in terms of the weak limit under additional integrability assumptions on derivatives of .
Theorem 1.2.
For any and , there exists a function satisfying
-
i)
, where ,
-
ii)
for and .
-
iii)
for all ,
which solves (4)-(5) in the following weak sense
-
a)
There exists a set with and a function satisfying such that
(13) for all with . Here is the set where are nondegenerate and is the characteristic function of set .
-
b)
Assume For any open set on which and for some , we have
(14)
Moreover, the following energy inequality holds for all
(15) | |||||
Adding an additional term to the chemical potential in (2), we can apply the same method to derive existence of weak solutions of the modified model (see section 5 for further details). Such nonlocal model originates from the phase field model for self-climb of dislocation loops [20].
The paper is organized as follows. We shall derive sharp interface limit for (1) and (2) through formal asymptotic expansions in section 2. Section 3 is devoted to the proof of Theorem 1.1 and Theorem1.2 is proved in section 4. Similar existence theorems for the modified model with an additional nonlocal term added to the chemical potential are presented in section 5.
2 Sharp interface limit via asymptotic expansions
In this section, we perform a formal asymptotic analysis to obtain the sharp interface limit of the proposed phase field model (1)- (2)as .
2.1 Outer expansions
We first perform expansion in the region far from the dislocations. Assume the expansion for is
(16) |
Correspondingly, we have
We also expand chemical potential as
(17) |
and rewrite equation (1) as
(18) |
Set
Plugging the expansions into (18) and (2) and matching the coefficients of powers in both equations, the of (18) and (2) yields
(19) | |||
(20) |
Since
then or satisfies equations (19)-(20). In particular, such choice of implies .
(23) | |||
(24) |
In summary, the or in the outer region.
2.2 Inner expansions
For the small inner regions near the dislocations, we introduce local coordinates near the dislocations. Considering a dislocation parameterized by arc length parameter . We denote a point on the dislocation by with tangent unit vector and inward normal vector . A point near the dislocation is expressed as
where is the signed distance from point to the dislocation. Since the gradients fields are of order , we introduce and use coordinates in the inner region. Under this setting, we write and equation (1) can be written as
(27) | |||
(28) |
Assume takes the same form expansion as (17) and the following expansions hold for within dislocation core region:
(29) |
Here we assume the leading order solution , which describes the dislocation core profile, remains the same at all points on the dislocation at any time.
Set
the leading order for equation (27) and (28) is , which yields
(30) | |||
(31) |
Substituting into (30), we can rewrite (30) as
Integrating this equation, we have
(32) |
Since in the outer region, we must have and as . Therefore . Dividing (32) by and integrating, we have . Since is independent of and is in the outer region, we must have . Thus
(33) |
Solution to (33) subject to far field condition and can be found numerically (see [20] for example). In particular, for all .
Next, the equation of (27) and (28) yields, using , that
(34) | |||
(35) |
When , we have . Sustituting into (34) and integrating, we have
(36) |
Matching with the outer solutions ( as ), we conclude that . Dividing (36) by and integrating, we have . Thus (35) can be written as
(37) |
where is a linear operator whose kernal is span. (37) is sovlable iff the right hand side is perpendicular to the kernal of , i.e.
From this, we conclude
where positive constants and are given by
Therefore
(38) |
Letting , (27) can be written as
(39) | |||||
Using , , the order equation of (39) reduces to
Integrating with respect to , we have . Matching with outer solutions, we must have . Thus which gives .
Next we look at the equation of (39). Using , and , we have
Integrating this equation with respect to and matching with outer solutions yields
(40) |
where we used the fact that is independent of and
By (38), we have , substitute this into (40), we obtain the sharp interface limit equation
(41) |
Remark 2.1.
Notice here the outer and inner expansions are similar to the expansions in [20]. We wrote out all details here for readers’ convenience.
3 Weak solution for phase field model with positive mobilities
In this section we prove existence of weak solutions for phase field model with positive mobilities. Let be the set of nonnegative integers and with . We pick an orthonormal basis for as
Observe is also orthogonal in for any . Here and throughout the paper, we denote .
3.1 Galerkin approximations
Define
where satisfy
(42) | |||||
(43) | |||||
(44) |
(42)-(44) is an initial value problem for a system of ordinary equations for . Since right hand side of (42) is continuous in , the system has a local solution.
Define energy functional
Direct calculation yields
integration over gives the following energy identity.
Here and throughout the paper, represents a generic constant possibly depending only on , , but not on . Since is bounded region, by growth assumption assumption (6) and Poincare’s inequality, the energy identity (3.1) implies with
(46) |
and
(47) |
By (46), the coefficients are bounded in time, thus the system (42)-(44) has a global solution. In addition, by Sobolev embedding theorem and growth assumption (7) on , we have
for any with
(48) | |||
(49) |
3.2 Convergence of
Given and any , let be the orthogonal projection of onto span. Then
Since
we have
Therefore
(50) |
For , since , by Sobolev embedding theorem and Aubin-Lions Lemma (see [22] and Remark 3.1) , the following embeddings are compact :
and
From this and the boundedness of and , we can find a subsequence, and such that as ,
(51) | |||||
(52) | |||||
(53) | |||||
(54) |
for . In addition
3.3 Weak solution
By (47) and the lower bound on , we have
(61) |
By (43), (46) and (48), we have
(61),(3.3) and Poincare’s inequality yield
Thus there exists a and a subsequence of , not relabeled, such that
(63) |
Therefore by (55), (63) and Sobolev embedding theorem, we have
(64) |
for any . Combining (56), (63)and (64), we have
(65) |
for any . By (47), we can improve this convergence to
(66) |
By (43), we have
Integrating with respect to from to , we have on ,
Passing to the limit in the equation above, by (53), (57) and (64), we have
(67) |
On the other hand,
Since strongly in , by (51),(58) and (64), as , (3.3) yields
(69) |
(70) |
By (46), weakly in , thus (70) implies
(71) |
It follows from (71), (72), (73) and generalized dominated convergence theorem (see Remark 3.2) that
(74) |
Let
by (74), we can extract a subsequence of , not relabeled, such that a.e. in (0,T). By Egorov’s theorem, for any given , there exists with such that converges to uniformly on .
Given , for any , there exists with such that
(75) |
Multiplying (42) by and integrating in time yield
Since , by (63) and (64), we have
(77) |
To prove convergence on , observe
To pass to the limit in , we write
By (47), (51), (73) and (75), we can bound by
For , we have
Since converges to uniformly on , and , letting in yields . Letting , we conclude as . Passing to the limit in (3.3), we have
(79) | |||||
Fix , given any , its Fourier series converges strongly to in . Hence
where
by(63), (64) and strong convergence of to in . We can bound by
Consequently (79) and (3.3) imply
(81) |
for all with . Moreover, since in , we see that by (52).
Remark 3.2.
(Generalized dominated convergence theorem) Assume is measurable. strongly in for and , : are measurable functions satisfying
with , then in .
3.4 Regularity of
We now consider the regularity of . Given any , ). Integrating (43) from to , by (58),(64) and (71), we have
for all . Given any , its Fouirier series strongly converges to in , therefore
(82) |
Recall and for any , regularity theory implies . Hence
(83) |
Since growth assumption on implies , pick , we have
Therefore with
Hence , combined with for any , we have and
(84) |
Regularity of implies . A simple interpolation shows for any . Given any with and , we have for any . From this, we can pick as a test function in (81), we have
(85) |
for any with .
3.5 Energy Inequality
4 Phase field model with degenerate mobility
In this section, we prove theorem 1.2. Fix initial data . We pick a montone decreasing positive sequence with . By theorem 1.1 and (85), for each , there exists
with weak derivative
where , , such that and for all ,
(86) | |||||
(87) |
Moreover, for all with , the following holds:
(88) |
Here we write , , for simplicity of notations.
4.1 Convergence of and equation for the limit function
Noticing the bound in (46) and (47) only depends on , we can find a constant , independent of such that
(89) | |||
(90) |
Growth condition on , and Sobolev embedding theorem give
for any . By (88), for any with ,
Let
(92) |
Thus (4.1) yields with and
(93) |
Moreover, by growth assumption on and estimates on , we have
(94) |
for . By (89), (90), (93)-(94) and Remark 3.1 we can find a subsequence, not relabeled, a function , a function and a function such that as ,
(95) | |||
(96) | |||
(97) | |||
(98) | |||
(99) | |||
(100) |
where . By (99) and (105) from Remark 4.1, we have
Thus given any , there exists such that for all and all ,
Given any , let . Consider the interval having and as end points. Denote this interval by . We consider three cases.
Case I: .
In this case, for any and by (92)
Case II: and .
In this case, we have
and
Case III: and
In this case, we have
Thus
Pick and fix . Let
Then
Taking maximum on the left side, we have for all , any ,
Thus
In addition, for any , (89) implies that for , we have
Therefore we conclude from Remark 4.1 that
(101) |
Similarly. we can prove
(102) |
Growth condition on and (101), (102) yield
Hence converges to a.e. in and . Passing to the limit in (88), we have
(103) |
for any with .
Remark 4.1.
(Compactness in Theorem 1 in [22]) Assume is a Banach space and . is relatively compact in for , or in for if and only if
(104) |
(105) |
Here for is defined on .
4.2 Weak convergence of
We now look for relation between and . Following the idea in [7], we decompose as follows. Let be a positive sequence monotonically decreasing to . By (96) and Egorov’s theorem, for every , there exists satisfying such that
(106) |
We can pick
(107) |
Define
Then
(108) |
Let and . Then and each can be split into two parts:
(109) |
For any with , we have
(110) | |||||
The left hand side of (110) converges to . We analyze the three terms on the right hand side separately. To estimate the first term on the right hand side of (110), noticing and
we have
By uniform convergence of to in , we introduce subsequence such that uniformly in and there exists such that for all ,
(111) |
Thus the third term on the right hand side of (110) can be estimated by
For the second term, we see that
Therefore is bounded in and we can extract a further subsequence, not relabeled, which converges weakly to some . Since is an increasig sequence of sets with , we have a.e. in . By setting outside , we can extend to a function defined in . Therefore for a.e. , there exists a limit of as . Let , we see that for a.e and for all .
By a standard diagonal argument, we can extract a subsequnce such that
(112) |
By strong convergence of to in for , we obtain
weakly in for and all . Recall weakly in , we have in for all . Hence in and consequently
weakly in for .
4.3 Relation between and
The desired relation between and is
(115) | |||||
Given the known regularity and degeneracy of , the right hand side of (115) might not be defined as a function. We can, however, under the additional assumption and suitable assumptions on integrability of , find an explicit expression of in terms of (115) in suitable subset of .
Claim I: If and for some , the interior of , denoted by , is not empty, then
and
Proof of the claim I. Since , we can have a subsequence, not relabeled such that, converges weakly to in . Since
(116) |
The right hand side of (116) weakly converges to in . Hence
On the other hand, using and as test functions in (82) yield
Passing to the limit, by (102), growth assumptions on and (116), we have
Therefore
Since , we can differentiate (116) and get
(117) |
and
(118) |
on . Thus
(119) |
Since
we have, for any ,
i.e.
Passing to the limit in (117), we obtain, in the sense of distribution, that
Since , we have , hence
(120) |
Since uniformly in , we have
Since uniformly in , we have
Passing to the limit in (118), we have
on . Noticing the value of on doesn’t matter since it does not appear on the right hand side of (113).
Claim II: For any open set in which for some and , we have
(121) |
in .
To prove this, since
(122) |
and
(123) |
The right hand side of (122) converges weakly to in for . Hence
The right hand side of (123) converges weakly to
in for each and therefore
in . The definition of can be extended to by our integrability assumption on . Define
Then is open and is defined by (121) on . Since , on and
we can take the value of to be zero outside , and it won’t affect the integral on the right side of (13).
Lastly the energy inequality (15) follows by taking limit in the energy inequality for .
Remark 4.2.
In Cahn-Hilliard case, there is convergence of on , and relation between and can be derived directly. Here we only have convergence of on . In order to obtain convergence of , we need convergence on for suitable , this is where we used the additional assumption .
5 A Modified phase field model for self-climb of prismatic dislocation loops
Dislocations are line defects in crystals [12, 23]. A phase field model [20] was derived based on the pipe diffusion model for self-climb of prismatic dislocation loops [18, 19] that describes the conservative climb of dislocation loops observed in experiments of irradiated materials [13, 12, 8]. In this section, we study the wellposedness of the following modified phase field model for self-climb of prismatic dislocation loops:
(124) | |||||
(125) |
Where , for , satisfy same assumptions (6)-(7) as those for Eqs. (1)-(2). Here is the total climb force with
where is the applied climb force, and
(126) |
represents the climb force generated by all the dislocations. Here is a bounded domain, is the shear modulus, is the Poisson ratio, and . In this model, we assume that the prismatic dislocation loops lie and evlove by self-climb in the plane and all dislocation loops have the same Burgers vector .
The chemical potential comes from variations of the classical Cahn-Hilliard energy and the elastic energy due to dislocations, i.e.
(127) |
where
(128) | |||
(129) |
are classical Cahn-Hilliard energy and elastic energy, respectively. Under periodic boundary conditions, the climb force generated by the dislocations can be expressed as
(130) |
Here is a fractional operator defined by
for . In the analysis below, without loss of generality, we set the coefficient of the climb force .
System (124)-(125) is a modified version of the phase field model introduced in [20], which does not have the term on the left side of (124). Putting an extra factor in front of the nonlocal climb force , the asymptotic analysis in [20] showed that the proposed phase field model yields accurate dislocation self-climb velocity in the sharp interface limit. Moreover, numerical simulations in [20] showed excellent agreement with experimental observations and discrete dislocation dynamics simulation results. Now we prove the wellposedness of the modified model (124)-(125). There is an extra nonlocal term in this model compared with the model considered in previous sections.
Define
where satisfy
(131) | |||||
(132) | |||||
(133) |
(131)-(133) is an initial value problem for a system of ordinary equations for . Since right hand side of (131) is continuous in , the system has a local solution.
Define energy functional
Direct calculation yields
integration over gives the following energy identity for any
where represents a generic constant possibly depending only on , , but not on . Since is bounded region, by growth assumption assumption (6) and Poincare’s inequality, the energy identity (5) implies with
(135) |
and
(136) |
Repeat the argument in Section 3 and Section 4, replacing energy functional by when necessary, we can prove the following existence theorem for (124)-(125) with nondegenerate and degenerate mobilities respectively.
Theorem 5.1.
Theorem 5.2.
For any and , there exists a function satisfying
-
i)
, where ,
-
ii)
for and .
-
iii)
for all ,
which solves (4)-(5) in the following weak sense
-
a)
There exists a set with and a function satisfying such that
(139) for all with . Here is the set where are nondegenerate and is the characteristic function of set .
-
b)
Assume For any open set on which and for some , we have
(140) a.e in .
Moreover, the following energy inequality holds for all
(141) | |||||
ACKNOWLEDGEMENTS X.H. Niu’s research is supported by National Natural Science Foundation of China under the grant number 11801214 and the Natural Science Foundation of Fujian Province of China under the grant number 2021J011193. Y. Xiang’s research is supported by the Hong Kong Research Grants Council General Research Fund 16307319. X. Yan’s research is supported by a Research Excellence Grant and CLAS Dean’s Summer Research Grant from University of Connecticut.
References
- [1] M. Albani, L. Ghisalberti, R. Bergamaschini, M. Friedl, M. Salvalaglio, A. Voigt, F. Montalenti, G. Tütüncüoglu, A. F. i Morral, and L. Miglio, Growth kinetics and morphological analysis of homoepitaxial gaas fins by theory and experiment, Phys. Rev. Mater., 2 (2018), p. 093404.
- [2] J. W. Cahn and J. E. Hilliard, Free energy of a nonuniform system. i. interfacial free energy, The Journal of Chemical Physics, 28 (1958), pp. 258–267.
- [3] , Spinodal decomposition: A reprise, Acta Metallurgica, 19 (1971), pp. 151–161.
- [4] J. W. Cahn and J. Taylor, Overview no. 113 surface motion by surface diffusion, Acta Metallurgica et Materialia, 42 (1994), pp. 1045–1063.
- [5] S. Dai and Q. Du, Motion of interfaces governed by the cahn–hilliard equation with highly disparate diffusion mobility, SIAM Journal on Applied Mathematics, 72 (2012), pp. 1818–1841.
- [6] , Coarsening mechanism for systems governed by the cahn–hilliard equation with degenerate diffusion mobility, Multiscale Modeling & Simulation, 12 (2014), pp. 1870–1889.
- [7] , Weak solutions for the Cahn-Hilliard equation with degenerate mobility, Arch. Ration. Mech. Anal., 219 (2016), pp. 1161–1184.
- [8] S. Dudarev, Density functional theory models for radiation damage, Annu. Rev. Mater. Res., 43 (2013), p. 35–61.
- [9] K. L. M. Elder, W. B. Andrews, M. Ziehmer, N. Mameka, C. Kirchlechner, A. Davydok, J.-S. Micha, A. F. Chadwick, E. T. Lilleodden, K. Thornton, and P. W. Voorhees, Grain boundary formation through particle detachment during coarsening of nanoporous metals, Proc. Natl. Acad. Sci., 118 (2021), p. e2104132118.
- [10] C. M. Elliott and H. Garcke, On the cahn–hilliard equation with degenerate mobility, SIAM Journal on Mathematical Analysis, 27 (1996), pp. 404–423.
- [11] C. Gugenberger, R. Spatschek, and K. Kassner, Comparison of phase-field models for surface diffusion., Physical review. E, Statistical, nonlinear, and soft matter physics, 78 1 Pt 2 (2008), p. 016703.
- [12] J. Hirth and J. Lothe, Theory of Dislocations, McGraw-Hill, New York, 1982.
- [13] F. Kroupa and P. B. Price, Conservative climb of a dislocation loop due to its interaction with an edge dislocation, Philos. Mag., 6 (1961), pp. 243–247.
- [14] A. A. Lee, A. Münch, and E. Süli, Degenerate mobilities in phase field models are insufficient to capture surface diffusion, Applied Physics Letters, 107 (2015), p. 081603.
- [15] , Sharp-interface limits of the cahn-hilliard equation with degenerate mobility, SIAM J. Appl. Math., 76 (2016), pp. 433–456.
- [16] J.-L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod; Gauthier-Villars, Paris, 1969.
- [17] M. Naffouti, R. Backofen, M. Salvalaglio, T. Bottein, M. Lodari, A. Voigt, T. David, A. Benkouider, I. Fraj, L. Favre, A. Ronda, I. Berbezier, D. Grosso, M. Abbarchi, and M. Bollani, Complex dewetting scenarios of ultrathin silicon films for large-scale nanoarchitectures, Sci. Adv., 3 (2017), p. eaao1472.
- [18] X. Niu, Y. Gu, and Y. Xiang, Dislocation dynamics formulation for self-climb of dislocation loops by vacancy pipe diffusion, Int. J. Plast., 120 (2019), pp. 262 – 277.
- [19] X. Niu, T. Luo, J. Lu, and Y. Xiang, Dislocation climb models from atomistic scheme to dislocation dynamics, J. Mech. Phys. Solids, 99 (2017), pp. 242 – 258.
- [20] X. Niu, Y. Xiang, and X. Yan, Phase field model for self-climb of prismatic dislocation loops by vacancy pipe diffusion, Int. J. Plast., 141 (2021), p. 102977.
- [21] A. Rätz, A. Ribalta, and A. Voigt, Surface evolution of elastically stressed films under deposition by a diffuse interface model, J. Comput. Phys., 214 (2006), p. 187–208.
- [22] J. Simon, Compact sets in the space , Annali di Matematica Pura ed Applicata, 146 (1986), pp. 65–96.
- [23] Y. Xiang, Modeling dislocations at different scales, Commun. Comput. Phys., 1 (2006), pp. 383–424.
- [24] J. Yin, On the existence of nonnegative continuous solutions of the cahn-hilliard equation, Journal of Differential Equations, 97 (1992), pp. 310–327.