Skip to main content
Log in

A phase field computational procedure for electromigration with specified contact angle and diffusional anisotropy

  • Original Paper
  • Published:
Computational Mechanics Aims and scope Submit manuscript

Abstract

Electromigration in electrical interconnects with high current density causes voids to form and grow near the cathode. These voids can possibly grow large enough to cause open circuit failure. The simulation of electromigration void nucleation and growth is challenging because of the complex interaction of electrical, thermal and mechanical fields that lead to the observed voids. In this paper, a phase-field approach is developed to study void growth due to electromigration taking into account diffusional anisotropy and specified contact angle. Consideration of anisotropy and contact angle is critical in applications including electronic assemblies that most commonly use Sn-based solder interconnects. In such systems, Sn exhibits significant anisotropy in its self-diffusivity, which assumes importance in small solder joints that are known to contain only a few grains. Furthermore, the contact angle at the void-bonding pad-solder triple junction is known to influence the void size and shape. To the best of the authors’ knowledge, only anisotropy in surface diffusivity has been considered in phase-field electromigration literature, whereas the contact angle at a triple junction has not been considered. The developed phase-field model incorporates effects of both contact angle and anisotropy in self-diffusivity. The model is used to simulate systems with different contact angles and dominant directions of diffusion. It is observed that diffusional anisotropy plays a more dominant role in failure, while the contact angle dictates evolution of the shape of voids.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  1. Bieler T, Jiang H, Lehman L, Kirkpatrick T, Cotts E (2006) Influence of Sn grain size and orientation on the thermomechanical response and reliability of Pb-free solder joints. In: Proceedings of the electronic components and technology conference, pp 1462–1467

  2. Chen C, Tong H, Tu K (2010) Electromigration and thermomigration in pb-free flip-chip solder joints. Annu Rev Mater Res 40:531–555

    Google Scholar 

  3. Zhang L, Ou S, Huang J, Tu KN, Gee S, Nguyen L (2006) Effect of current crowding on void propagation at the interface between intermetallic compound and solder in flip chip solder joints. Appl Phys Lett 88(1):012106

    Google Scholar 

  4. Basaran C, Lin M, Ye H (2003) A thermodynamic model for electrical current induced damage. Int J Solids Struct 40(26):7315–7327

    MATH  Google Scholar 

  5. Basaran C, Lin M (2008) Damage mechanics of electromigration induced failure. Mech Mater 40:66–79. https://doi.org/10.1016/j.mechmat.2007.06.006

    Article  Google Scholar 

  6. Rice JR, Drucker DC (1967) Energy C in stressed bodies due to void and crack growth. Int J Fract 3(1):19–27. https://doi.org/10.1007/BF00188642

    Article  Google Scholar 

  7. Lin M, Basaran C (2005) Electromigration induced stress analysis using fully coupled mechanical—diffusion equations with nonlinear material properties. Comput Mater Sci 34:82–98. https://doi.org/10.1016/j.commatsci.2004.10.007

    Article  Google Scholar 

  8. Liu Y, Zhang Y, Liang L (2010) Prediction of electromigration induced voids and time to failure for solder joint of a wafer level chip scale package. IEEE Trans Compon Packag Technol 33(3):544–552. https://doi.org/10.1109/TCAPT.2010.2042717

    Article  Google Scholar 

  9. Park M, Gibbons S, Arróyave R (2013) Phase-field simulations of intermetallic compound evolution in cu/sn solder joints under electromigration. Acta Mater 61(19):7142–7154

    Google Scholar 

  10. Udupa A, Sadasiva S, Subbarayan G (2016) A framework for studying dynamics and stability of diffusive-reactive interfaces with application to cu6sn5 intermetallic compound growth. Proc R Soc A Math Phys Eng Sci 472(2190):20160134

    MATH  Google Scholar 

  11. Kassner K, Misbah C (1999) A phase-field approach for stress-induced instabilities. EPL (Europhys Lett) 46(2):217

    Google Scholar 

  12. Hakim V, Karma A (2005) Crack path prediction in anisotropic brittle materials. Phys Rev Lett 95(23):235501

    Google Scholar 

  13. Miehe C, Welschinger F, Hofacker M (2010) Thermodynamically consistent phase field models of fracture: variational principles and multi-field FE implementations. Int J Numer Methods Eng 83(10):1273–1311

    MathSciNet  MATH  Google Scholar 

  14. Wang Y, Li J (2010) Phase field modeling of defects and deformation. Acta Mater 58(4):1212–1235

    Google Scholar 

  15. Caginalp G (1986) An analysis of a phase field model of a free boundary. Arch Ration Mech Anal 92(3):205–245

    MathSciNet  MATH  Google Scholar 

  16. Caginalp G, Fife P (1986) Phase-field methods for interfacial boundaries. Phys Rev B 33(11):7792

    Google Scholar 

  17. Penrose O, Fife PC (1990) Thermodynamically consistent models of phase-field type for the kinetic of phase transitions. Physica D Nonlinear Phenom 43(1):44–62

    MathSciNet  MATH  Google Scholar 

  18. Kobayashi R (1993) Modeling and numerical simulations of dendritic crystal growth. Physica D Nonlinear Phenom 63(3–4):410–423

    MATH  Google Scholar 

  19. Warren JA, Boettinger WJ (1995) Prediction of dendritic growth and microsegregation patterns in a binary alloy using the phase-field method. Acta Metall Mater 43(2):689–703

    Google Scholar 

  20. Wheeler AA, Boettinger WJ, McFadden GB (1992) Phase-field model for isothermal phase transitions in binary alloys. Phys Rev A 45(10):7424

    Google Scholar 

  21. Chen LQ, Khachaturyan A (1991) Computer simulation of structural transformations during precipitation of an ordered intermetallic phase. Acta Metall Mater 39(11):2533–2551

    Google Scholar 

  22. Wang Y, Chen LQ, Khachaturyan A (1993) Kinetics of strain-induced morphological transformation in cubic alloys with a miscibility gap. Acta Metall Mater 41(1):279–296

    Google Scholar 

  23. Chen LQ (2002) Phase-field models for microstructure evolution. Annu Rev Mater Res 32(1):113–140

    Google Scholar 

  24. Mahadevan M, Bradley RM (1999) Simulations and theory of electromigration-induced slit formation in unpassivated single-crystal metal lines. Phys Rev B 59(16):11037

    Google Scholar 

  25. Bhate DN, Bower AF, Kumar A (2002) A phase field model for failure in interconnect lines due to coupled diffusion mechanisms. J Mech Phys Solids 50(10):2057–2083

    MathSciNet  MATH  Google Scholar 

  26. Barrett JW, Nürnberg R, Styles V (2004) Finite element approximation of a phase field model for void electromigration. SIAM J Numer Anal 42(2):738–772

    MathSciNet  MATH  Google Scholar 

  27. Fensham PJ (1950) Self-diffusion in tin crystals. Aus J Chem 3(1):91–104

    Google Scholar 

  28. Lu M, Shih DY, Lauro P, Goldsmith C, Henderson DW (2008) Effect of Sn grain orientation on electromigration degradation mechanism in high Sn-based Pb-free solders. Appl Phys Lett 92(21):211909

    Google Scholar 

  29. Telang AU, Bieler TR (2005) The orientation imaging microscopy of lead-free Sn-Ag solder joints. JOM 57(6):44–49

    Google Scholar 

  30. Lane M, Liniger E, Lloyd J (2003) Relationship between interfacial adhesion and electromigration in cu metallization. J Appl Phys 93(3):1417–1421

    Google Scholar 

  31. Jacqmin D (2000) Contact-line Dynamics of a Diffuse Fluid Interface. Journal of Fluid Mechanics 402:57–88

    MathSciNet  MATH  Google Scholar 

  32. Khatavkar V, Anderson P, Meijer H (2007) Capillary spreading of a droplet in the partially wetting regime using a diffuse-interface model. J Fluid Mech 572:367–387

    MathSciNet  MATH  Google Scholar 

  33. Turco A, Alouges F, DeSimone A (2009) Wetting on rough surfaces and contact angle hysteresis: numerical experiments based on a phase field model. ESAIM Math Model Numer Anal 43(6):1027–1044

    MathSciNet  MATH  Google Scholar 

  34. Wheeler D, Warren JA, Boettinger WJ (2010) Modeling the early stages of reactive wetting. Phys Rev E 82(5):051601

    Google Scholar 

  35. Ding H, Spelt PD (2007) Wetting condition in diffuse interface simulations of contact line motion. Phys Rev E 75(4):046708

    Google Scholar 

  36. Lee HG, Kim J (2011) Accurate contact angle boundary conditions for the Cahn–Hilliard equations. Comput Fluids 44(1):178–186

    MATH  Google Scholar 

  37. Cahn J, Larché F (1983) An invariant formulation of multicomponent diffusion in crystals. Scr Metall 17(7):927–932

    Google Scholar 

  38. Larché F, Cahn JW (1985) Overview no. 41 the interactions of composition and stress in crystalline solids. Acta Metall 33(3):331–357

    Google Scholar 

  39. Balluffi RW, Seigle LL (1957) Growth of voids in metals during diffusion and creep. Acta Metall 5(8):449–454

    Google Scholar 

  40. Kirchheim R (1992) Stress and electromigration in al-lines of integrated circuits. Acta Metall Mater 40(2):309–323

    Google Scholar 

  41. Fischer F, Svoboda J (2008) Void growth due to vacancy supersaturation-a non-equilibrium thermodynamics study. Scr Mater 58(2):93–95

    Google Scholar 

  42. Cuitiño A, Ortiz M (1996) Ductile fracture by vacancy condensation in F.C.C. single crystals. Acta Mater 44(2):427–436. https://doi.org/10.1016/1359-6454(95)00220-0

    Article  Google Scholar 

  43. Hull D, Rimmer DE (1959) The growth of grain-boundary voids under stress. Philo Mag 4(42):673–687

    Google Scholar 

  44. Sadasiva S (2014) Simulations of diffusion driven phase evolution in heterogenous solids. Ph.D. thesis, Purdue University

  45. Balluffi RW, Allen S, Carter WC (2005) Kinetics of materials. Wiley, New York

    Google Scholar 

  46. Cahn JW, Elliott CM, Novick-Cohen A (1996) The Cahn–Hilliard equation with a concentration dependent mobility: motion by minus the Laplacian of the mean curvature. Eur J Appl Math 7(3):287–302

    MathSciNet  MATH  Google Scholar 

  47. Li X, Lowengrub J, Rätz A, Voigt A (2009) Solving PDEs in complex geometries: a diffuse domain approach. Commun Math Sci 7(1):81–107

    MathSciNet  MATH  Google Scholar 

  48. Eyre DJ (1998) An unconditionally stable one-step scheme for gradient systems. Unpublished article

  49. Wu X, Zwieten GJ, Zee KG (2014) Stabilized second-order convex splitting schemes for Cahn–Hilliard models with application to diffuse-interface tumor-growth models. Int J Numer Methods Biomed Eng 30(2):180–203

    MathSciNet  Google Scholar 

  50. Shen J, Yang X (2009) An efficient moving mesh spectral method for the phase field model of two-phase flows. J Comput Phys 228(8):2978–2992

    MathSciNet  MATH  Google Scholar 

  51. Kirk BS, Peterson JW, Stogner RH, Carey GF (2006) libMesh: a C++ library for parallel adaptive mesh refinement/coarsening simulations. Eng Comput 22(3–4):237–254. https://doi.org/10.1007/s00366-006-0049-3

    Article  Google Scholar 

  52. Stogner RH, Carey GF, Murray BT (2008) Approximation of Cahn–Hilliard diffuse interface models using parallel adaptive mesh refinement and coarsening with c1 elements. Int J Numer Methods Eng 76(5):636–661

    MATH  Google Scholar 

  53. Lloyd JR, Lane MW, Liniger EG (2002) Relationship between interfacial adhesion and electromigration in Cu metallization. In: Integrated reliability workshop final report, 2002. IEEE International, IEEE, pp 32–35

  54. Fife PC (1988) Dynamics of internal layers and diffusive interfaces, vol 53. SIAM, Philadelphia

    MATH  Google Scholar 

  55. Fife PC, Penrose O (1995) Interfacial dynamics for thermodynamically consistent phase-field models with nonconserved order parameter. Electron J Differ Equ 16:1–49

    MathSciNet  MATH  Google Scholar 

  56. Gurtin ME, Jabbour ME (2002) Interface evolution in three dimensions with curvature-dependent energy and surface diffusion: interface-controlled evolution, phase transitions, epitaxial growth of elastic films. Arch Ration Mech Anal 163(3):171–208

    MathSciNet  MATH  Google Scholar 

  57. Vaitheeswaran PK, Udupa A, Sadasiva S, Subbarayan G (2019) Interface balance laws, phase growth and nucleation conditions for multiphase solids with inhomogeneous surface stress. Contin Mech Thermodyn. https://doi.org/10.1007/s00161-019-00804-z

    Article  Google Scholar 

Download references

Acknowledgements

This study was partially supported by Intel Corporation and the Semiconductor Research Corporation under Task 1292.090.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to G. Subbarayan.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A: Formal asymptotic analysis

Appendix A: Formal asymptotic analysis

In general, it is necessary to show that the diffuse interface equations used in phase field modeling lead to the those given in Sect. 2 in the limit of the sharp interface. Such an exercise is carried out using the method of matched formal asymptotic analysis. A more detailed discussion of the technique of formal asymptotic analysis can be found in [54, 55].

1.1 Appendix A.1: Transformations to an interface coordinate system

The domain for the asymptotic analysis is shown in Fig. 13. In the outer region I, the Heaviside step function of Eq. (32), \(H(\mathbf {x})= 1\), while in region II, \(H(\mathbf {x})= 0\). In both region I and II, the Dirac \(\delta \) function of Eq. (33), \(\delta (\mathbf {x}) = 0\). At any instant, the interface \(\varGamma \) moves with a normal velocity \(v_n\mathbf {n}\), where \(\mathbf {n}\) is the unit normal to the interface at the origin of the coordinate system. Following [46], a moving coordinate system attached to the interface \(\varGamma \) based on the scaled distance \(\rho \) from the interface is used (Fig. 14); \(\rho \) is the normal signed distance from the interface, r, scaled by \(\frac{1}{\epsilon }\):

$$\begin{aligned} \rho = \frac{r}{\epsilon } \end{aligned}$$
(53)
Fig. 13
figure 13

Domain for the formal asymptotic analysis

In the moving inner coordinate system, any point \(\mathbf {x}\)  in space can be written as,

$$\begin{aligned} \mathbf {x}&= r\mathbf {n}+ \mathbf {P}\mathbf {x}\end{aligned}$$
(54)

where \(\mathbf {P}\) is the projection tensor defined as:

$$\begin{aligned} \mathbf {P}=\mathbf {I}-\mathbf {n}\otimes \mathbf {n}\end{aligned}$$
(55)

with \(\mathbf {I}\) being the identity tensor. Thus,

$$\begin{aligned} \frac{\partial \mathbf {x}}{\partial t}&= \frac{\partial r}{\partial t}\mathbf {n}+ \mathbf {P}\frac{\partial \mathbf {x}}{\partial t} \end{aligned}$$
(56)

Using the definitions given in [56] and described in detail in [57], the surface gradient of the scalar field \(u(\mathbf {x},t)\) is:

$$\begin{aligned} \nabla _{\varGamma }u=\mathbf {P}\nabla u \end{aligned}$$
(57)

Similarly, the surface gradient and the surface divergence of any vector field \(\mathbf {u}(\mathbf {x},t)\) is given by,

$$\begin{aligned}&\nabla _{\varGamma }\mathbf {u}=\mathbf {P}\nabla \mathbf {u} \end{aligned}$$
(58)
$$\begin{aligned}&\nabla _{\varGamma }\cdot \mathbf {u} =\mathrm {tr}\left( \nabla _{\varGamma }\mathbf {u}\right) =\mathbf {P}:\nabla \mathbf {u} \end{aligned}$$
(59)

Thus, using the definition of \(\mathbf {P}\), it follows from Eqs. (57) to (59) that

$$\begin{aligned}&\nabla u = \frac{{u}_{\rho }}{\epsilon }\mathbf {n}+\nabla _{\varGamma }u \end{aligned}$$
(60)
$$\begin{aligned}&\nabla \cdot \mathbf {u} =\mathbf {n}\cdot \nabla \mathbf {u} \cdot \mathbf {n}+\nabla _{\varGamma }\cdot \mathbf {u} \nonumber \\&=\frac{1}{\epsilon }\mathbf {u}_\rho \cdot \mathbf {n}+ \nabla _{\varGamma }\cdot \mathbf {u} \end{aligned}$$
(61)

where the first derivative of the fields with respect to \(\rho \) is denoted as \({u}_{\rho }=\frac{\partial u}{\partial \rho }\) and \(\mathbf {u}_\rho =\frac{\partial \mathbf {u}}{\partial \rho }\). In the above expression, we have also used \(\nabla \rho =\frac{1}{\epsilon }\nabla r = \frac{1}{\epsilon }\mathbf {n}\).

Fig. 14
figure 14

Transformation of coordinate system from a global coordinate system to an interface attached coordinate system

The Laplacian of a scalar field \(u(\mathbf {x})\) can next be derived as,

$$\begin{aligned} \nabla ^2u&= \nabla \cdot \nabla u = \nabla \cdot \left( \frac{{u}_{\rho }}{\epsilon }\mathbf {n}+\nabla _{\varGamma }u\right) \nonumber \\&= \nabla _{\varGamma }\cdot \left( \frac{{u}_{\rho }}{\epsilon }\mathbf {n}+\nabla _{\varGamma }u\right) +\frac{1}{\epsilon }\frac{\partial }{\partial \rho }\left( \frac{{u}_{\rho }}{\epsilon }\mathbf {n}+\nabla _{\varGamma }u\right) \cdot \mathbf {n} \nonumber \\ \Rightarrow \nabla ^2u&= \frac{1}{\epsilon ^2}{u}_{\rho \rho } + \frac{1}{\epsilon }{u}_{\rho } \nabla _{\varGamma }\cdot \mathbf {n} + \nabla _\varGamma ^2u \end{aligned}$$
(62)

where \({u}_{\rho \rho }=\frac{\partial ^2 u}{\partial \rho ^2}\). Using the curvature relation that \(\nabla _{\varGamma }\cdot \mathbf {n}=-\kappa \), we get the relation for the Laplacian as,

$$\begin{aligned} \nabla ^2u = \frac{1}{\epsilon ^2}{u}_{\rho \rho } - \frac{1}{\epsilon }\kappa {u}_{\rho } + \nabla _\varGamma ^2u \end{aligned}$$
(63)

The signed distance field moves along with the interface and hence \(\frac{d r}{dt}=0\). This gives,

$$\begin{aligned} \frac{d r}{dt}&= \frac{\partial r}{\partial t} + \mathbf {v}\cdot \nabla r = \frac{\partial r}{\partial t} + v_n = 0 \end{aligned}$$
(64)
$$\begin{aligned} \Rightarrow \frac{\partial r}{\partial t}&= -v_n \end{aligned}$$
(65)

Recognizing that, in the inner coordinate system, a scalar field is of the form \(u(\mathbf {x}(r,\mathbf {P}),t)\), the temporal partial derivative of a scalar field is given by,

$$\begin{aligned} \frac{\partial }{\partial t}u(\mathbf {x}(r,\mathbf {P}),t)&= \frac{\partial }{\partial t}u(r,\mathbf {P},t) + \frac{\partial u}{\partial \mathbf {x}}\cdot \frac{\partial \mathbf {x}}{\partial t} \nonumber \\&= {u}_{t} + \left( \frac{\partial r}{\partial t}\mathbf {n}+ \mathbf {P}\frac{\partial \mathbf {x}}{\partial t}\right) \cdot \nabla u \nonumber \\&= {u}_{t} + \frac{1}{\epsilon }\frac{\partial r}{\partial t}{u}_{\rho } + \frac{\partial \mathbf {x}}{\partial t}\cdot \nabla _{\varGamma }u \nonumber \\ \Rightarrow \frac{\partial }{\partial t}u(\mathbf {x}(r,\mathbf {P}),t)&= {u}_{t} - \frac{v_n}{\epsilon }{u}_{\rho } + \frac{\partial \mathbf {x}}{\partial t}\cdot \nabla _{\varGamma }u \end{aligned}$$
(66)

1.2 Appendix A.2: Asymptotic analysis of the modified Cahn–Hiliard equations

The asymptotic analysis for Eq. (39) is based on the analysis in [46]. The objective of this asymptotic analysis is to determine the constants \(c_1\)-\(c_3\) for the modified Cahn–Hiliard [Eq. (67)] equation given below.

$$\begin{aligned} \frac{\partial \phi }{\partial t}&= c_1 \left( \varOmega _\text {sol}\nabla \cdot \left[ \frac{M_\varGamma \delta (\phi ) }{\epsilon } \left( \nabla \mu _\text {sol}^\varGamma \right. \right. \right. \nonumber \\&\quad \left. \left. \left. +\, Z^*_\varGamma e \nabla \varphi _e +~Q^*_\varGamma \nabla T \right) \right] \right. \nonumber \\&\quad \left. +\, \frac{\delta (\phi )}{\epsilon } \lambda x_{\text {vac}} \left( \mu _\text {vac}- \mu _\text {sol}^\varGamma \right) \right) \end{aligned}$$
(67a)
$$\begin{aligned} \nonumber \\ \mu _\text {sol}^\varGamma&= c_2 \varOmega _\text {sol}\gamma \left( -\epsilon \nabla ^2 \phi + \frac{1}{\epsilon } f'(\phi ) \right) \nonumber \\&\quad +\, c_3 \varOmega _\text {sol}\delta (\phi ) W \end{aligned}$$
(67b)

The solutions \(\phi ,\mu _\text {sol}^\varGamma \) to the above equation can be expanded in the outer regions as,

$$\begin{aligned} {\hat{\mu }}(r,\epsilon )&= {\hat{\mu }}^0 + \epsilon {\hat{\mu }}^1+ \epsilon ^2 {\hat{\mu }}^2 + \cdots \end{aligned}$$
(68a)
$$\begin{aligned} {\hat{\phi }}(r,\epsilon )&= {\hat{\phi }}^0 + \epsilon {\hat{\phi }}^1 + \epsilon ^2 {\hat{\phi }}^2 + \cdots . \end{aligned}$$
(68b)

The solution in the inner region can be written as,

$$\begin{aligned} \mu (\rho ,\epsilon )&= \mu ^0 + \epsilon \mu ^1 +\epsilon ^2\mu ^2 + \ldots , \end{aligned}$$
(69a)
$$\begin{aligned} \phi (\rho ,\epsilon )&= \phi ^0 + \epsilon \phi ^1 + \epsilon ^2 \phi ^2 + \ldots . \end{aligned}$$
(69b)

The analysis of outer region is performed first. Substituting Eq. (68) into Eq. (67b), the leading order in the outer expansion is the \(\epsilon ^{-1}\) term. This leads to,

$$\begin{aligned} f'({\hat{\phi }}^0) = 0 \end{aligned}$$
(70)

Since the roots of \(f'({\hat{\phi }}^0)\) are at \(\pm 1\) (see Fig. 8), the solution to \({\hat{\phi }}\) in the outer regions converges to either \(+ 1\) or \(-1\). This allows the phases to be discriminated, with \(+1\) indicating the presence of material and \(-1\) the void. The only other term in the \(\epsilon ^0\) order in Eq. (67b) is the \(\varOmega _\text {sol}\delta ({\hat{\phi }}) W\) term, which is zero since \(\delta ({\hat{\phi }})=0\) in the outer region. This leads to the leading order term \({\hat{\mu }}^0 = 0\) in the outer region, and furthermore, Eq. (67a) can be reduced to \(0=0\) in all orders.

For the inner-expansion, Eq. (67) needs to be transformed into a co-ordinate system that is attached to the interface. Using Eqs. (63) and (66), the modified Cahn–Hilliard equation can be written as,

$$\begin{aligned} \begin{aligned}&{\phi }_{t} - \frac{1}{\epsilon } v_n{\phi }_{\rho } + \frac{\partial \mathbf {x}}{\partial t}\cdot \nabla _{\varGamma }\phi = \\&c_1\frac{\delta (\phi )}{\epsilon }\lambda x_\text {vac}\left( \mu _\text {vac}- \mu _\text {sol}^\varGamma \right) \\&\quad + c_1\varOmega _\text {sol}\left[ \frac{M \delta (\phi ) }{\epsilon }\left( \frac{1}{\epsilon ^2} {G}_{\rho \rho }- \frac{1}{\epsilon } \kappa { G}_{\rho } + \nabla _\varGamma ^2 G \right) \right. \\&\quad + \left. \frac{M\delta '(\phi )}{\epsilon } \left( \frac{{\phi }_{\rho }}{\epsilon } \mathbf {n}+\nabla _\varGamma \phi \right) . \left( \frac{ {G}_{\rho }}{\epsilon } \mathbf {n}+ \nabla _\varGamma G \right) \right] \end{aligned} \end{aligned}$$
(71a)
$$\begin{aligned} \nonumber \\ \begin{aligned} \mu _\text {sol}^\varGamma = c_2 \varOmega _\text {sol}\gamma \left( \frac{-1}{\epsilon }{\phi }_{\rho \rho } + \kappa {\phi }_{\rho } +\frac{1}{\epsilon }f'(\phi ) - \epsilon \nabla _\varGamma ^2 \phi \right)&\\ + c_3\varOmega _\text {sol}\delta (\phi ) W&\end{aligned} \end{aligned}$$
(71b)

where

$$\begin{aligned} G = \mu _\text {sol}^\varGamma + Z^*e \varphi _e+ Q^*T \end{aligned}$$
(72)

Substituting Eq. (69) into Eq. (71), the leading order terms in Eq. (71b) are at the \(\epsilon ^{-1}\) order and can be written as,

$$\begin{aligned} -\phi ^0_{\rho \rho } + f'(\phi ^0) = 0 \end{aligned}$$
(73)

The boundary conditions that Eq. (73) needs to satisfy the matching conditions with the outer regions I and II are, \(\lim _{\rho \rightarrow -\infty }\phi ^0 \rightarrow -1 \) and \(\lim _{\rho \rightarrow \infty }\phi ^0 \rightarrow 1\). With these boundary conditions, the solution to Eq. (73) is found to be,

$$\begin{aligned} \phi ^0 = \tanh \frac{\rho }{\sqrt{2}} \end{aligned}$$
(74)

This implies that the equilibrium profile of \(\phi \) is a hyperbolic tangent. Eq. (74) also explains the choice of the approximation to the Heaviside step function in Eq. (32), as that can be computed quite simply as \(H(\mathbf {x})= \frac{1+\phi (\mathbf {x})}{2}\). Further, Eq. (74) also provides a relation for the normal and curvature in terms of \(\phi ^0\) as,

$$\begin{aligned} \nabla \phi ^0&= \frac{1}{\sqrt{2}}\mathrm {sech}^2\frac{\rho }{\sqrt{2}}\nabla \rho = \frac{1}{\sqrt{2}\epsilon }\mathrm {sech}^2\frac{\rho }{\sqrt{2}}\mathbf {n}\end{aligned}$$
(75)
$$\begin{aligned} \nabla ^2\phi ^0&= \frac{-1}{\epsilon }\tanh \frac{\rho }{\sqrt{2}}\mathrm {sech}^2\frac{\rho }{\sqrt{2}}\nabla \cdot \mathbf {n} \nonumber \\&= \frac{-\kappa }{\epsilon }\tanh \frac{\rho }{\sqrt{2}}\mathrm {sech}^2\frac{\rho }{\sqrt{2}} \end{aligned}$$
(76)

where we have used the curvature-normal relation. Also, using the definition of surface gradient [Eq. (58)], we can now show that

$$\begin{aligned} \nabla _{\varGamma }\phi ^0&= \mathbf {P}\nabla \phi ^0 = \frac{1}{\sqrt{2}\epsilon }\mathrm {sech}^2\frac{\rho }{\sqrt{2}}\mathbf {P}\mathbf {n}= 0 \end{aligned}$$
(77)

Thus, at the \(\epsilon ^0\) order in Eq. (71b), the terms that remain are,

$$\begin{aligned} \mu ^0 = c_2 \gamma \varOmega _\text {sol}\kappa {\phi ^0_\rho }_{\,}+ c_3\varOmega _\text {sol}\delta (\phi ^0)W \end{aligned}$$
(78)

To make the left hand side of Eq. (78) integrable, both sides are multiplied by \(\phi ^0_\rho \). Integrating with respect to \(\rho \) over \(-\infty , \infty \), Eq. (78) can be shown to yield,

$$\begin{aligned} 2 \mu ^0 = \varOmega _\text {sol}\left( c_2 \frac{2\sqrt{2}}{3} \gamma \kappa + c_3 \frac{16}{15} W \right) \end{aligned}$$
(79)

Comparing with the sharp interface relation given in Eq. (27), we get

$$\begin{aligned} c_2=-\frac{3}{\sqrt{2}}, c_3=\frac{15}{8} \end{aligned}$$
(80)

In the above, the following integrals are used,

$$\begin{aligned} \int _{-\infty }^{\infty }&{\left( \tanh \frac{\rho }{\sqrt{2}}\right) }_{\rho }d\rho =2 \end{aligned}$$
(81a)
$$\begin{aligned} \int _{-\infty }^{\infty }&{\left( \tanh \frac{\rho }{\sqrt{2}}\right) }_{\rho }^2 d\rho = \frac{2\sqrt{2}}{3}\end{aligned}$$
(81b)
$$\begin{aligned} \int _{-\infty }^{\infty }&{\left( \tanh \frac{\rho }{\sqrt{2}}\right) }_{\rho }\left( 1- \phi ^2 \right) ^2 d\rho = \frac{16}{15} \end{aligned}$$
(81c)

An important assumption made in the above derivation is that the computed value of the strain energy is nearly constant over the interfacial region.

We observe now that Eq. (79) leads to \({\mu ^0_\rho } = 0\), and the results of Eqs. (74) and (80) can be used to analyze the solutions to Eq. (71a). Considering now the \(\epsilon ^{-2}\) terms in the inner region and noting that \(\mathbf {n}\cdot \nabla _{\varGamma }(\cdot )=0\) (surface gradient is orthogonal to normal \(\mathbf {n}\)) yields,

$$\begin{aligned}&{G^0_\rho }=0 \end{aligned}$$
(82)
$$\begin{aligned}&{G^0_{\rho \rho }}=0 \end{aligned}$$
(83)

The leading order in Eq. (71a) is thus \(\epsilon ^{-1}\). Now, considering terms of order \(\epsilon ^{-1}\), and using Eq. (77), we obtain the following,

$$\begin{aligned} \begin{aligned}&-v_n{\phi ^0_\rho } = c_1\varOmega _\text {sol}M\delta (\phi ^0) \nabla _\varGamma ^2 G \\&\quad +\, c_1 \delta (\phi ^0)\lambda x_\text {vac}(\mu _\text {vac}-\mu ^0) \end{aligned} \end{aligned}$$
(84)

Integrating the above from \(-\infty \) to \(\infty \) with respect to \(\rho \), substituting the terms of G and writing \(x_\text {vac}=\varOmega _\text {sol}c_\text {vac}\), the expression reduces to,

$$\begin{aligned} \begin{aligned} -2v_n= c_1\frac{4\sqrt{2}}{3}\varOmega _\text {sol}\Bigg [M \nabla _\varGamma ^2 \left( \mu ^0 +Z^*e\phi ^0 + Q^*T^0 \right) \Bigg .&\\ \Bigg .+\lambda c_\text {vac}(\mu _\text {vac}-\mu ^0) \Bigg ]&\end{aligned} \end{aligned}$$
(85)

On simplification, Eq. (85) therefore recovers the relation for the motion of the interface [Eq. (30)] if \(c_1=\frac{3}{2\sqrt{2}}\). Thus, the constants in Eq. (39) are determined.

1.3 Appendix A.3: Asymptotic analysis of the vacancy diffusion equation

Finally, the phase field version of the vacancy diffusion equation is studied. A crucial assumption made in this derivation here is that the pressure variation in the interfacial region is negligible. This is not strictly true, as only the displacements are constant over the interfacial region. However, in this paper, the value of the pressure and other stress/strain related quantities is averaged over the interfacial region and hence \({p_{\text {in}}}_{\rho } =0 \) is reasonable. The objective is to find a constant \(c_4\) such that the vacancy diffusion [Eq. (86)] reduces to the sharp interface boundary condition for the vacancies on the moving interface Eq. (22).

$$\begin{aligned} \begin{aligned}&\frac{\partial }{\partial t}\left( H(\mathbf {x})x_\text {vac}\right) = \\&\nabla \cdot \left[ H(\mathbf {x})\frac{\mathbf {D}_\text {vac}}{RT}{x_\text {vac}} \left( RT \frac{\nabla x_\text {vac}}{x_\text {vac}} + \beta \nabla p + Z^*e \nabla \varphi _e\right. \right. \\&\quad \left. \left. ~~~+ Q^*\nabla T \right) \right] + c_4\frac{\delta (\phi )}{\epsilon } \lambda x_\text {vac}\left( \mu _\text {vac}- \mu _\text {sol}^\varGamma \right) \end{aligned} \end{aligned}$$
(86)

In the outer domain, the interfacial terms drop out as \(\delta ({\hat{\phi }}) = 0\), and the standard bulk diffusion equation [Eq. (21)] is recovered. In the inner region, the vacancy mole fraction is expanded as

$$\begin{aligned} x_\text {vac}= x_\text {vac}^0 + \epsilon x_\text {vac}^1 + \cdots \end{aligned}$$
(87)

This lets the chemical potential to be expanded as,

$$\begin{aligned} \begin{aligned} \mu&= \mu + \epsilon \mu ^1 + \epsilon ^2\mu ^2 + \cdots \\&= RT\ln \frac{x_\text {vac}^0}{x_\text {vac}^\text {eq}} + \epsilon RT\frac{x^1_\text {vac}}{x_\text {vac}^0} + \cdots&\\&+ \beta ( p^0 +\epsilon p^1 + \cdots ) \end{aligned} \end{aligned}$$
(88)

Thus, we get in the leading order

$$\begin{aligned} \mu ^0 = RT \ln \frac{x_\text {vac}^0}{x_\text {vac}^\text {eq}} + \beta p^0 \end{aligned}$$
(89)

Transforming Eq. (86) into the inner co-ordinate system, the diffusion equation in the inner region can be written using Eq. (60) as,

$$\begin{aligned} \begin{aligned}&\frac{\partial }{\partial t}(H(\phi )x_\text {vac}) - \frac{v_n}{\epsilon } {(H (\phi )x_\text {vac})}_{\rho } \\&\quad +\frac{\partial \mathbf {x}}{\partial t}\cdot \nabla _{\varGamma }(H(\phi )x_\text {vac}) = \\&\quad -\frac{1}{\epsilon } \mathbf {n}.{\mathbf {j}_\text {vac}}_{\rho } -\nabla _{\varGamma }\cdot \mathbf {j}_\text {vac} \\&\quad + c_4\frac{\delta (\phi )}{\epsilon } \lambda x_\text {vac}\left( \mu _\text {vac}- \mu _\text {sol}^\varGamma \right) \end{aligned} \end{aligned}$$
(90)

where \(\mathbf {j}_\text {vac}\) is,

$$\begin{aligned} \begin{aligned} \mathbf {j}_\text {vac}=&-H (\phi )\frac{D_\text {vac}}{RT} x_\text {vac}\left( \nabla \mu _\text {vac}+ Z^*e \nabla \phi _e \right. \\&\left. + Q^*\nabla T \right) \end{aligned} \end{aligned}$$
(91)

Expanding Eq. (91) in the inner region and using Eq. (72), terms of order \(\epsilon ^{-1}\) can be grouped to get,

$$\begin{aligned} H(\phi )D_\text {vac}x_\text {vac}\left( {\mu ^0_\text {vac}}_{\rho } + G^0_\rho - {\mu ^0_\text {sol}}_{\rho }\right) = 0 \end{aligned}$$
(92)

From Eq. (79), we know that \(\mu _{\text {sol}_\rho }^0=0\). Next, using Eq. (82), and since \(H(\phi )D_\text {vac}x_\text {vac}\ge 0\) in the inner region, the above relation implies that \(\mu ^0_{\text {vac}_\rho }=0\) or \(\mu ^0_{\text {vac}} = {constant}\) in the inner region. Therefore, from Eq. (17)

$$\begin{aligned} {x_\text {vac}^0}_{\rho } = -\frac{x_\text {vac}^0}{RT} \beta {p^0_\rho } \end{aligned}$$
(93)

With the stress/strain quantities smoothed over the interfacial region, \({p^0_\rho } \approx 0\) and this leads to,

$$\begin{aligned} {x_\text {vac}^0}_{\rho } = 0 \end{aligned}$$
(94)

Terms of order \(\epsilon ^{-1}\) in Eq. (90) can now be grouped to get,

$$\begin{aligned} \begin{aligned}&-v_n{H(\phi )}_{\rho } x_\text {vac}- v_nH(\phi ){x_\text {vac}}_{\rho } = -\mathbf {n}.{\mathbf {j}_\text {vac}}_{\rho } \\&\quad + c_4 \delta (\phi ) \lambda x_\text {vac}\left( \mu _\text {vac}- \mu _\text {sol}^\varGamma \right) \end{aligned} \end{aligned}$$
(95)

Integrating the above equation from \( -\infty \) to \(\infty \), the relation can be rewritten as,

$$\begin{aligned} -v_nx_\text {vac}= -n.\mathbf {j}_\text {vac}+ c_4\frac{4\sqrt{2}}{3} \lambda x_\text {vac}\left( \mu _\text {vac}- \mu _\text {sol}^\varGamma \right) \end{aligned}$$
(96)

This recovers the boundary condition for the vacancies on the moving interface Eq. (22) for \(c_4=-\frac{3}{4\sqrt{2}}\). This determines the required constant in Eq. (40).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sadasiva, S., Vaitheeswaran, P. & Subbarayan, G. A phase field computational procedure for electromigration with specified contact angle and diffusional anisotropy. Comput Mech 66, 373–390 (2020). https://doi.org/10.1007/s00466-020-01855-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00466-020-01855-9

Keywords

Navigation