Skip to main content
Log in

A Stabilized Semi-Implicit Euler Gauge-Invariant Method for the Time-Dependent Ginzburg–Landau Equations

  • Published:
Journal of Scientific Computing Aims and scope Submit manuscript

Abstract

In this paper, we propose and analyze a stabilized semi-implicit Euler gauge-invariant method for numerical solution of the time-dependent Ginzburg–Landau (TDGL) equations in the two-dimensional space. The proposed method uses the well-known gauge-invariant finite difference approximations with staggered variables in a rectangular mesh, and a stabilized semi-implicit Euler discretization for time integration. The resulted fully discrete system leads to two decoupled linear systems at each time step, thus can be efficiently solved. We prove that the proposed method unconditionally preserves the point-wise boundedness of the solution and is also energy-stable. Moreover, the proposed method under the zero-electric potential gauge is shown to be equivalent to a mass-lumped version of the lowest order rectangular Nédélec edge element approximation and the Lorentz gauge scheme to a mass-lumped mixed finite element method. These indicate the method is also effective in solving the TDGL problems in non-convex domains although the solutions are often of low-regularity in such situation. Various numerical experiments are also presented to demonstrate effectiveness and robustness of the proposed method.

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
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22

Similar content being viewed by others

References

  1. Alstrom, T., Sorensen, M., Pedersen, N., Madsen, S.: Magnetic flux lines in complex geometry type-II superconductors studied by the time dependent Ginzburg–Landau equation. Acta Appl. Math. 115, 63–74 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  2. Arnold, D., Falk, R., Winther, R.: Finite element exterior calculus: from Hodge theory to numerical stability. Bull. Am. Math. Soc. (N.S.) 47, 281–354 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  3. Buscaglia, G., Bolech, C., Lopez, A.: On the numerical solution of the time-dependent Ginzburg–Landau equations in multiply connected domains. In: Berger, J., Rubinstein, J. (eds.) Connectivity and superconductivity. LNP m62, pp. 200–214. Springer, Berlin (2000)

    Chapter  Google Scholar 

  4. Carey, G., Knezevic, D.: Multiscale and hysteresis effects in vortex pattern simulations for Ginzburg–Landau problems. Int. J. Numer. Meth. Eng. 81, 866–891 (2010)

    MathSciNet  MATH  Google Scholar 

  5. Chapman, S.J., Howison, S.D., Ockendon, J.R.: Macroscopic models for superconductivity. SIAM Rev. 34, 529–560 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  6. Chen, Z.: Mixed finite element methods for a dynamical Ginzburg–Landau model in superconductivity. Numer. Math. 76, 323–353 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chen, Z., Hoffmann, K.: Numerical studies of a non-stationary Ginzburg–Landau model for superconductivity. Adv. Math. Sci. Appl. 5, 363–389 (1995)

    MathSciNet  MATH  Google Scholar 

  8. Christiansen, S., Halvorsen, T.: A Gauge invariant discretization on simplicial grids of the Schrödinger eigenvalue problem in an electromagnetic field. SIAM J. Numer. Anal. 49, 331–345 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  9. Coskun, E., Kwong, M.: Simulating vortex motion in superconducting films with the time-dependent Ginzburg–Landau equations. Nonlinearity 10, 579–593 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  10. Du, Q.: Finite element methods for the time-dependent Ginzburg–Landau model of superconductivity. Comput. Math. Appl. 27, 119–133 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  11. Du, Q.: Discrete gauge-invariant approximations of a time dependent Ginzburg–Landau model of superconductivity. Math. Comput. 67, 965–986 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  12. Du, Q., Gunzburger, M.D., Peterson, J.S.: Analysis and approximation of the Ginzburg–Landau model of superconductivity. SIAM Rev. 34, 54–81 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  13. Du, Q., Ju, L., Li, X., Qiao, Z.: Stabilized linear semi-implicit schemes for the nonlocal Cahn–Hilliard equation. J. Comput. Phys. 363, 39–54 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  14. Du, Q., Nicolaides, R., Wu, X.: Analysis and convergence of a covolume approximation of the Ginzburg–Landau model of superconductivity. SIAM J. Numer. Anal. 35, 1049–1072 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  15. Frahm, H., Ullah, S., Dorsey, A.: Flux dynamics and the growth of the superconducting phase. Phys. Rev. Lett. 66, 3067–3070 (1991)

    Article  Google Scholar 

  16. Gao, H., Li, B., Sun, W.: Optimal error estimates of linearized Crank–Nicolson Galerkin FEMs for the time-dependent Ginzburg–Landau equations in superconductivity. SIAM J. Numer. Anal. 52, 1183–1202 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  17. Gao, H., Sun, W.: An efficient fully linearized semi-implicit Galerkin-mixed FEM for the dynamical Ginzburg–Landau equations of superconductivity. J. Comput. Phys. 294, 329–345 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  18. Gao, H., Sun, W.: A new mixed formulation and efficient numerical solution of Ginzburg–Landau equations under the temporal gauge. SIAM J. Sci. Comput. 38, A1339–A1357 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  19. Gao, H., Sun, W.: Analysis of linearized Galerkin—mixed FEMs for the time-dependent Ginzburg–Landau equations of superconductivity. Adv. Comput. Math. 44, 923–949 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  20. Gor’kov, L.P., Eliashberg, G.M.: Generalization of the Ginzburg–Landau equations for non-stationary problems in the case of alloys with paramagnetic impurities. Sov. Phys. JETP 27, 328–334 (1968)

    Google Scholar 

  21. Gropp, W., Kaper, H., Leaf, G., Levine, D., Palumbo, M., Vinokur, V.: Numerical simulation of vortex dynamics in type-II superconductors. J. Comput. Phys. 123, 254–266 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  22. Gunter, D., Kaper, H., Leaf, G.: Implicit integration of the time-dependent Ginzburg–Landau equations of superconductivity. SIAM J. Sci. Comput. 23, 1943–1958 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  23. Kato, R., Enomoto, Y., Maekawa, S.: Computer simulations of dynamics of flux lines in type-II superconductors. Phys. Rev. B 47, 8016–8024 (1993)

    Article  Google Scholar 

  24. Kaper, H., Kwong, M.: Vortex configurations in type-II superconducting films. J. Comput. Phys. 119, 120–131 (1995)

    Article  MATH  Google Scholar 

  25. Kim, S., Burkardt, J., Gunzburger, M., Peterson, J., Hu, C.: Effects of sample geometry on the dynamics and configurations of vortices in mesoscopic superconductors. Phys. Rev. B 76, 024509 (2007)

    Article  Google Scholar 

  26. Li, B.: Convergence of a decoupled mixed FEM for the dynamic Ginzburg-Landau equations in nonsmooth domains with incompatible initial data. Calcolo 54, 1441–1480 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  27. Li, B., Zhang, Z.: Mathematical and numerical analysis of time-dependent Ginzburg–Landau equations in non-convex polygons based on Hodge decomposition. Math. Comput. 86, 1579–1608 (2017)

    Article  MATH  Google Scholar 

  28. Li, B., Zhang, Z.: A new approach for numerical simulation of the time-dependent Ginzburg–Landau equations. J. Comput. Phys. 303, 238–250 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  29. Liao, H., Zhou, S., Shi, X.: Simulating the time-dependent Ginzburg–Landau equations for type-II superconductors by finite-difference method. Chin. Phys. 13, 0737-09 (2004)

    Google Scholar 

  30. Monk, P.: Finite Element Methods for Maxwell’s Equations. Oxford University Press, New York (2003)

    Book  MATH  Google Scholar 

  31. Mu, M.: A linearized Crank–Nicolson–Galerkin method for the Ginzburg–Landau model. SIAM J. Sci. Comput. 18, 1028–1039 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  32. Mu, M., Huang, Y.: An alternating Crank–Nicolson method for decoupling the Ginzburg–Landau equations. SIAM J. Numer. Anal. 35, 1740–1761 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  33. Raza, N., Sial, S., Siddiqi, S.: Approximating time evolution related to Ginzburg–Landau functionals via Sobolev gradient methods in a finite-element setting. J. Comput. Phys. 229, 1621–1625 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  34. Richardson, W., Pardhanani, A., Carey, G., Ardelea, A.: Numerical effects in the simulation of Ginzburg–Landau models for superconductivity. Int. J. Numer. Meth. Eng. 59, 1251–1272 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  35. Stosic, Darko, Stosic, Dusan, Ludermir, T., Stosic, B., Milosevic, M.: GPU-advanced 3D electromagnetic simulations of superconductors in the Ginzburg–Landau formalism. J. Comput. Phys. 322, 183–198 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  36. Thomée, V.: Galerkin Finite Element Methods for Parabolic Problems. Springer, Berlin (1997)

    Book  MATH  Google Scholar 

  37. Tinkham, M.: Introduction to Superconductivity. McGraw-Hill, New York (1975)

    Google Scholar 

  38. Xu, C., Tang, T.: Stability analysis of large time-stepping methods for epitaxial growth models. SIAM J. Numer. Anal. 44, 1759–1779 (2006)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Huadong Gao.

Additional information

Publisher's Note

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

H. Gao’s research is partially supported by National Natural Science Foundation of China under Grant Number 11871234 and Fundamental Research Funds for the Central Universities (HUST) under Grant Number 2017KFYXJJ089. L. Ju’s research is partially supported by US National Science Foundation under Grant Number DMS-1818438.

Appendices

Appendix

An Example with Explicit Coefficient Matrices

In this appendix, we provide an example on a \(2\times 2\) mesh to show the equivalence of the gauge invariant finite difference approximation and the mass-lumped mixed finite element method. To this end, we shall only consider the Lorentz gauge scheme for \({\mathbf {A}}\) (2.34) and the mixed FEM (4.8)–(4.9). For simplicity, we take \(\Omega = (0,1)^2\) and set \(h= \frac{1}{2}\). There are 9 degrees of freedom (DOFs) for \(\Phi \) and 12 DOFs for \({\mathbf {A}}\), see Fig. 23.

Fig. 23
figure 23

An illustration of labeled global DOFs for \({\mathbf {A}}\) and \(\Phi \) on a \(2 \times 2\) mesh

The Lorentz gauge scheme (2.33)–(2.34) can be written as

$$\begin{aligned} \frac{1}{\Delta t} \overrightarrow{A}^{n} - \nabla _h \, D_h \overrightarrow{A}^{n} + {\mathbf {K}}_1 \overrightarrow{A}^{n} = \text {RHS}, \end{aligned}$$
(A.1)

where \(\nabla _h\) is a \(12 \times 9\) gradient matrix and \(D_h\) is a \(9 \times 12\) divergence matrix defined respectively by

$$\begin{aligned} {\nabla }_h&= \frac{1}{h} \left[ \begin{array}{*{9}{r}} -1&{} \quad 1&{} \quad 0&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0&{} \quad -1&{} \quad 1&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -1&{} \quad 0&{} \quad 0&{} \quad 1&{} \quad 0&{} \quad 0&{} \quad &{} \quad &{} \quad \\ 0&{} \quad -1&{} \quad 0&{} \quad 0&{} \quad 1&{} \quad 0&{} \quad &{} \quad &{} \quad \\ 0&{} \quad 0&{} \quad -1&{} \quad 0&{} \quad 0&{} \quad 1&{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad -1&{} \quad 1&{} \quad 0&{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad 0&{} \quad -1&{} \quad 1&{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad -1&{} \quad 0&{} \quad 0&{} \quad 1&{} \quad 0&{} \quad 0\\ &{} \quad &{} \quad &{} \quad 0&{} \quad -1&{} \quad 0&{} \quad 0&{} \quad 1&{} \quad 0\\ &{} \quad &{} \quad &{} \quad 0&{} \quad 0&{} \quad -1&{} \quad 0&{} \quad 0&{} \quad 1\\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -1&{} \quad 1&{} \quad 0\\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0&{} \quad -1&{} \quad 1\\ \end{array} \right] ,\\ D_h&= \frac{1}{h} \left[ \begin{array}{rrrrrrrrrrrr} 2 &{} \quad 0 &{} \quad 2 &{} \quad 0 &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -1 &{} \quad 1 &{} \quad 0 &{} \quad 2 &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad -2 &{} \quad 0 &{} \quad 0 &{} \quad 2 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad -1 &{} \quad 0 &{} \quad 0 &{} \quad 2 &{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad 0 &{} \quad &{} \quad \\ &{} \quad &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad -1 &{} \quad 1 &{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad &{} \quad \\ &{} \quad &{} \quad 0 &{} \quad 0 &{} \quad -1&{} \quad 0 &{} \quad -2 &{} \quad 0 &{} \quad 0 &{} \quad 1 &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -2 &{} \quad 0 &{} \quad 0 &{} \quad 2 &{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad -2 &{} \quad 0 &{} \quad -1 &{} \quad 1 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad 0&{} \quad -2 &{} \quad 0 &{} \quad -2 \\ \end{array} \right] , \end{aligned}$$

and \({\mathbf {K}}_1\) is the coefficient matrix defined by

$$\begin{aligned} {\mathbf {K}}_1 = \frac{1}{h^2} \left[ { \begin{array}{*{12}{r}} 2&{} \quad 0&{} \quad -2&{} \quad 2&{} \quad 0&{} \quad -2&{} \quad 0&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0&{} \quad 2&{} \quad 0&{} \quad -2&{} \quad 2&{} \quad 0&{} \quad -2&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -2&{} \quad 0&{} \quad 2&{} \quad -2&{} \quad 0&{} \quad 2&{} \quad 0&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 1&{} \quad -1&{} \quad -1&{} \quad 2&{} \quad -1&{} \quad -1&{} \quad 1&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0&{} \quad 2&{} \quad 0&{} \quad -2&{} \quad 2&{} \quad 0&{} \quad -2&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -1&{} \quad 0&{} \quad 1&{} \quad -1&{} \quad 0&{} \quad 2&{} \quad 0&{} \quad -1&{} \quad 1&{} \quad 0&{} \quad -1&{} \quad 0 \\ 0&{} \quad -1&{} \quad 0&{} \quad 1&{} \quad -1&{} \quad 0&{} \quad 2&{} \quad 0&{} \quad -1&{} \quad 1&{} \quad 0&{} \quad -1 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -2&{} \quad 0&{} \quad 2&{} \quad -2&{} \quad 0&{} \quad 2&{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 1&{} \quad -1&{} \quad -1&{} \quad 2&{} \quad -1&{} \quad -1&{} \quad 1 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0&{} \quad 2&{} \quad 0&{} \quad -2&{} \quad 2&{} \quad 0&{} \quad -2 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -2&{} \quad 0&{} \quad 2&{} \quad -2&{} \quad 0&{} \quad 2&{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0&{} \quad -2&{} \quad 0&{} \quad 2&{} \quad -2&{} \quad 0&{} \quad 2 \end{array} } \right] . \end{aligned}$$

Let us recall that the mixed FEM introduces \(\Phi = -\mathrm {div}\, {\mathbf {A}}\) as an extra variable. Based on the lowest order bilinear element space \(S_h\) and lowest order rectangular Nédélec edge element space \(V_h\), we look for \((\Phi _h^n, {\mathbf {A}}_h^n) \, \in \, S_h \times V_h\) such that

$$\begin{aligned}&(\Phi _h^n, \omega _h) - ({\mathbf {A}}_h^n, \nabla \omega _h) = 0, \quad \forall \,\omega _h \in S_h , \end{aligned}$$
(A.2)
$$\begin{aligned}&\frac{1}{\Delta t}({\mathbf {A}}_h^{n}, {\mathbf {v}}_h) + (\nabla \Phi _h^n, {\mathbf {v}}_h) + (\mathrm {curl}{\mathbf {A}}_h^n, \mathrm {curl}{\mathbf {v}}_h ) = \text {RHS}\,, \quad \forall \, {\mathbf {v}}_h \in V_h, \end{aligned}$$
(A.3)

which further can be represented by the following matrix equations

$$\begin{aligned}&{\mathbf {M}}_{\text {bilinear}} \overrightarrow{\Phi }^{n} - {\mathbf {D}}_h\overrightarrow{A}^{n} = 0, \end{aligned}$$
(A.4)
$$\begin{aligned}&\frac{1}{\Delta t} \mathbf {\varvec{M}} \overrightarrow{A}^{n} + {\mathbf {D}}_h^{T} \overrightarrow{\Phi }^{n} + {\mathbf {K}}_2\overrightarrow{A}^{n} = \text {RHS}\,. \end{aligned}$$
(A.5)

On the \(2 \times 2\) mesh, the above mass matrix \({\mathbf {M}}_{\text {bilinear}}\) generated by the lowest order bilinear Lagrange element space and the matrix \(\mathbf {\varvec{D}}_h\) are defined by

$$\begin{aligned} {\mathbf {M}}_{\text {bilinear}}&=h^2 \left[ { \begin{array}{*{9}{r}} \frac{1}{9} &{} \quad \frac{1}{18}&{} \quad 0 &{} \quad \frac{1}{18}&{} \quad \frac{1}{36}&{} \quad 0 &{} \quad &{} \quad &{} \quad \\ \frac{1}{18}&{} \quad \frac{2}{9} &{} \quad \frac{1}{18}&{} \quad \frac{1}{36}&{} \quad \frac{1}{9}&{} \quad \frac{1}{36} &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad \frac{1}{18}&{} \quad \frac{1}{9} &{} \quad 0 &{} \quad \frac{1}{36}&{} \quad \frac{1}{18}&{} \quad &{} \quad &{} \quad \\ \frac{1}{18} &{} \quad \frac{1}{36} &{} \quad 0 &{} \quad \frac{2}{9} &{} \quad \frac{1}{9} &{} \quad 0 &{} \quad \frac{1}{18}&{} \quad \frac{1}{36} &{} \quad 0 \\ \frac{1}{36} &{} \quad \frac{1}{9} &{} \quad \frac{1}{36} &{} \quad \frac{1}{9} &{} \quad \frac{4}{9} &{} \quad \frac{1}{9} &{} \quad \frac{1}{36}&{} \quad \frac{1}{9} &{} \quad \frac{1}{36} \\ 0 &{} \quad \frac{1}{36} &{} \quad \frac{1}{18} &{} \quad 0 &{} \quad \frac{1}{9} &{} \quad \frac{2}{9} &{} \quad 0 &{} \quad \frac{1}{36} &{} \quad \frac{1}{18} \\ &{} \quad &{} \quad &{} \quad \frac{1}{18}&{} \quad \frac{1}{36}&{} \quad 0 &{} \quad \frac{1}{9} &{} \quad \frac{1}{18}&{} \quad 0 \\ &{} \quad &{} \quad &{} \quad \frac{1}{36}&{} \quad \frac{1}{9}&{} \quad \frac{1}{36}&{} \quad \frac{1}{18} &{} \quad \frac{2}{9} &{} \quad \frac{1}{18} \\ &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{36}&{} \quad \frac{1}{18}&{} \quad 0 &{} \quad \frac{1}{18}&{} \quad \frac{1}{9} \end{array} } \right] ,\\ \mathbf {\varvec{D}}_h&= \left[ { \begin{array}{*{12}{r}} -\frac{1}{3} &{} \quad 0 &{} \quad -\frac{1}{3} &{} \quad -\frac{1}{6} &{} \quad 0 &{} \quad -\frac{1}{6}&{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ \frac{1}{3} &{} \quad -\frac{1}{3} &{} \quad -\frac{1}{6} &{} \quad -\frac{2}{3} &{} \quad -\frac{1}{6}&{} \quad \frac{1}{6}&{} \quad -\frac{1}{6}&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad \frac{1}{3} &{} \quad 0 &{} \quad -\frac{1}{6} &{} \quad -\frac{1}{3}&{} \quad 0 &{} \quad \frac{1}{6} &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -\frac{1}{6} &{} \quad 0 &{} \quad \frac{1}{3} &{} \quad \frac{1}{6} &{} \quad 0 &{} \quad -\frac{2}{3}&{} \quad 0 &{} \quad -\frac{1}{3}&{} \quad -\frac{1}{6} &{} \quad 0 &{} \quad -\frac{1}{6} &{} \quad 0 \\ \frac{1}{6} &{} \quad -\frac{1}{6} &{} \quad \frac{1}{6} &{} \quad \frac{2}{3} &{} \quad \frac{1}{6}&{} \quad \frac{2}{3}&{} \quad -\frac{2}{3}&{} \quad -\frac{1}{6}&{} \quad -\frac{2}{3} &{} \quad -\frac{1}{6} &{} \quad \frac{1}{6} &{} \quad -\frac{1}{6} \\ 0 &{} \quad \frac{1}{6} &{} \quad 0 &{} \quad \frac{1}{6} &{} \quad \frac{1}{3}&{} \quad 0 &{} \quad \frac{2}{3}&{} \quad 0 &{} \quad -\frac{1}{6} &{} \quad -\frac{1}{3} &{} \quad 0 &{} \quad \frac{1}{6} \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -\frac{1}{6}&{} \quad 0 &{} \quad \frac{1}{3}&{} \quad \frac{1}{6} &{} \quad 0 &{} \quad -\frac{1}{3} &{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{6}&{} \quad -\frac{1}{6}&{} \quad \frac{1}{6}&{} \quad \frac{2}{3} &{} \quad \frac{1}{6} &{} \quad \frac{1}{3} &{} \quad -\frac{1}{3} \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{6}&{} \quad 0 &{} \quad \frac{1}{6} &{} \quad \frac{1}{3} &{} \quad 0 &{} \quad \frac{1}{3} \\ \end{array} } \right] , \end{aligned}$$

and the mass matrix \({\mathbf {M}}\) and stiffness matrix \({\mathbf {K}}_2\) generated by the lowest order rectangular Nédélec edge element space are defined respectively by

$$\begin{aligned} {\mathbf {M}}&= h^2 \left[ { \begin{array}{*{12}{r}} \frac{1}{3}&{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{6} &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad \frac{1}{3}&{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{6} &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad \frac{1}{3}&{} \quad \frac{1}{6}&{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad \frac{1}{6}&{} \quad \frac{2}{3}&{} \quad \frac{1}{6}&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad 0 &{} \quad \frac{1}{6}&{} \quad \frac{1}{3}&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ \frac{1}{6}&{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad \frac{2}{3} &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{6}&{} \quad 0 \\ 0 &{} \quad \frac{1}{6}&{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{2}{3} &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{6} \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{3} &{} \quad \frac{1}{6}&{} \quad 0 &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{6} &{} \quad \frac{2}{3}&{} \quad \frac{1}{6} &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{6}&{} \quad \frac{1}{3} &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{6} &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{3} &{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{6} &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad \frac{1}{3} \end{array} } \right] ,\\ {\mathbf {K}}_2&= \left[ { \begin{array}{*{12}{r}} 1 &{} \quad 0 &{} \quad -1 &{} \quad 1 &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad 1 &{} \quad 0 &{} \quad -1&{} \quad 1 &{} \quad 0 &{} \quad -1 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -1 &{} \quad 0 &{} \quad 1 &{} \quad -1&{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 1 &{} \quad -1 &{} \quad -1 &{} \quad 2&{} \quad -1&{} \quad -1 &{} \quad 1 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad 1 &{} \quad 0 &{} \quad -1&{} \quad 1&{} \quad 0 &{} \quad -1 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ -1 &{} \quad 0 &{} \quad 1 &{} \quad -1&{} \quad 0&{} \quad 2 &{} \quad 0 &{} \quad -1 &{} \quad 1 &{} \quad 0 &{} \quad -1 &{} \quad 0 \\ 0 &{} \quad -1 &{} \quad 0 &{} \quad 1&{} \quad -1&{} \quad 0 &{} \quad 2 &{} \quad 0 &{} \quad -1&{} \quad 1 &{} \quad 0 &{} \quad -1 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -1 &{} \quad 0 &{} \quad 1 &{} \quad -1 &{} \quad 0 &{} \quad 1 &{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 1 &{} \quad -1 &{} \quad -1 &{} \quad 2 &{} \quad -1 &{} \quad -1 &{} \quad 1 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad -1 &{} \quad 1 &{} \quad 0 &{} \quad -1 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad -1 &{} \quad 0 &{} \quad 1 &{} \quad -1 &{} \quad 0 &{} \quad 1 &{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad 1 &{} \quad -1 &{} \quad 0 &{} \quad 1 \end{array} } \right] . \end{aligned}$$

If the mass-lumping is used, then it holds

$$\begin{aligned}&(\Phi _h^n, \omega _h)_h - ({\mathbf {A}}_h^n, \nabla \omega _h)_h = 0, \end{aligned}$$
(A.6)
$$\begin{aligned}&\frac{1}{\Delta t}({\mathbf {A}}_h^{n}, {\mathbf {v}}_h)_h + (\nabla \Phi _h^n, {\mathbf {v}}_h)_h + (\mathrm {curl}{\mathbf {A}}_h^n, \mathrm {curl}{\mathbf {v}}_h) = \text {RHS}\, , \end{aligned}$$
(A.7)

and we can obtain the following matrix equations

$$\begin{aligned}&\widetilde{{\mathbf {M}}}_{\text {bilinear}} \overrightarrow{\Phi }^{n} -\widetilde{{\mathbf {D}}}_h \overrightarrow{A}^{n} = 0 \,, \end{aligned}$$
(A.8)
$$\begin{aligned}&\frac{1}{\Delta t} \widetilde{{\mathbf {M}}} \overrightarrow{A}^{n} +\widetilde{{\mathbf {D}}}_h^{T}\overrightarrow{\Phi }^{n} +{\mathbf {K}}_2 \overrightarrow{A}^{n} = \text {RHS}\,, \end{aligned}$$
(A.9)

where \(\widetilde{{\mathbf {M}}}_{\text {bilinear}}\) and \(\widetilde{\mathbf {\varvec{M}}}\) are diagonal matrices as

$$\begin{aligned}&\widetilde{{\mathbf {M}}}_{\text {bilinear}} = \text {diag}\left( \left[ \begin{array}{*{9}{r}} \frac{1}{4}&\frac{1}{2}&\frac{1}{4}&\frac{1}{2}&1&\frac{1}{2}&\frac{1}{4}&\frac{1}{2}&\frac{1}{4} \end{array} \right] ^T\right) , \nonumber \\&\widetilde{\mathbf {\varvec{M}}} = \text {diag}\left( \left[ \begin{array}{*{12}{r}} \frac{1}{2}&\frac{1}{2}&\frac{1}{2}&1&\frac{1}{2}&1&1&\frac{1}{2}&1&\frac{1}{2}&\frac{1}{2}&\frac{1}{2} \end{array} \right] ^T\right) , \end{aligned}$$

and

$$\begin{aligned} {\widetilde{{\mathbf {D}}}_h = h \left[ { \begin{array}{*{12}{r}} -\frac{1}{2} &{} \quad 0 &{} \quad -\frac{1}{2} &{} \quad 0 &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ \frac{1}{2} &{} \quad -\frac{1}{2} &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ 0 &{} \quad \frac{1}{2} &{} \quad 0 &{} \quad 0 &{} \quad -\frac{1}{2}&{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \\ &{} \quad &{} \quad \frac{1}{2} &{} \quad 0 &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad -\frac{1}{2}&{} \quad 0 &{} \quad 0 &{} \quad &{} \quad \\ &{} \quad &{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad 1 &{} \quad -1 &{} \quad 0 &{} \quad -1 &{} \quad 0 &{} \quad &{} \quad \\ &{} \quad &{} \quad 0 &{} \quad 0 &{} \quad \frac{1}{2}&{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad 0 &{} \quad -\frac{1}{2} &{} \quad &{} \quad \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad \frac{1}{2}&{} \quad 0 &{} \quad 0 &{} \quad -\frac{1}{2} &{} \quad 0 \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad 1 &{} \quad 0 &{} \quad \frac{1}{2} &{} \quad -\frac{1}{2} \\ &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad &{} \quad 0 &{} \quad 0&{} \quad \frac{1}{2} &{} \quad 0 &{} \quad \frac{1}{2} \\ \end{array} } \right] .} \end{aligned}$$

By taking matrices \(\widetilde{{\mathbf {M}}}_{\text {bilinear}}\), \(\widetilde{\mathbf {\varvec{M}}}\) and \(\widetilde{{\mathbf {D}}}_h\) into (A.9), we can deduce that

$$\begin{aligned}&\frac{1}{\Delta t} \overrightarrow{A}^{n} + \widetilde{{\mathbf {M}}}^{-1} \widetilde{{\mathbf {D}}}_h^{T} \widetilde{{\mathbf {M}}}_{\text {bilinear}}^{-1} \widetilde{{\mathbf {D}}}_h \overrightarrow{A}^{n} + \widetilde{{\mathbf {M}}}^{-1}{\mathbf {K}}_2\overrightarrow{A}^{n} = \text {RHS}\,. \end{aligned}$$

It is easy to verify the following matrix qualities

$$\begin{aligned} \nabla _h = \widetilde{{\mathbf {M}}}^{-1} \widetilde{{\mathbf {D}}}_h^{T}, \quad -D_h = \widetilde{{\mathbf {M}}}_{\text {bilinear}}^{-1}\widetilde{{\mathbf {D}}}_h, \quad {\mathbf {K}}_1 = \widetilde{{\mathbf {M}}}^{-1}{\mathbf {K}}_2, \end{aligned}$$

which confirm our observation. Finally, we shall point out that similar equivalences also hold for L-shape and multi-connected domains.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gao, H., Ju, L. & Xie, W. A Stabilized Semi-Implicit Euler Gauge-Invariant Method for the Time-Dependent Ginzburg–Landau Equations. J Sci Comput 80, 1083–1115 (2019). https://doi.org/10.1007/s10915-019-00968-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-019-00968-5

Keywords

Navigation