Skip to main content
Log in

Convergence of an ADI splitting for Maxwell’s equations

  • Published:
Numerische Mathematik Aims and scope Submit manuscript

Abstract

The convergence of an alternating direction implicit method for Maxwell’s equations on product domains is investigated. Unlike the classical Yee scheme and most other integrators proposed in the literature, this method is both unconditionally stable and computationally cheap. We prove second-order convergence of the time-discretization in the framework of operator semigroup theory. In contrast to formal considerations based on Taylor expansions, our convergence analysis respects the unboundedness of the involved differential operators. The proofs are based on results concerning the regularity of the Cauchy problems, which then allow to apply an abstract convergence proof by Hansen and Ostermann (Numer Math 108:557–570, 2008).

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  1. Amrouche, C., Bernardi, C., Dauge, M., Girault, V.: Vector potentials in three-dimensional non-smooth domains. Math. Methods Appl. Sci. 21(9), 823–864 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  2. Chen, W., Li, X., Liang, D.: Energy-conserved splitting FDTD methods for Maxwell’s equations. Numer. Math. 108(3), 445–485 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  3. Costabel, M., Dauge, M.: Singularities of electromagnetic fields in polyhedral domains. Arch. Ration. Mech. Anal. 151(3), 221–276 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  4. Dautray, R., Lions, J.-L.: Mathematical analysis and numerical methods for science and technology. In: Spectral Theory and Applications. With the Collaboration of Michel Artola and Michel Cessenat, Vol. 3, 2nd edn. Springer, Berlin (2000a)

  5. Dautray, R., Lions, J.-L.: Mathematical analysis and numerical methods for science and technology. In: Evolution Problems. I. With the Collaboration of Michel Artola, Michel Cessenat and Hélène Lanchon, Vol. 5, 2nd edn. Springer, Berlin (2000b)

  6. Engel, K.-J., Nagel, R.: One-Parameter Semigroups for Linear Evolution Equations, Volume 194 of Graduate Texts in Mathematics. Springer, Berlin (2000)

    Google Scholar 

  7. Faragó, I., Horváth, R., Schilders, W.H.A.: Investigation of numerical time-integrations of Maxwell’s equations using the staggered grid spatial discretization. Int. J. Numer. Model. Electron. Netw. Dev. Fields 18(2), 149–169 (2005)

    Article  MATH  Google Scholar 

  8. Gao, L., Zhang, B., Liang, D.: The splitting finite-difference time-domain methods for Maxwell’s equations in two dimensions. J. Comput. Appl. Math. 205(1), 207–230 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  9. Garcia, S.G., Lee, Tae-Woo, Hagness, S.C.: On the accuracy of the ADI-FDTD method. IEEE Antennas Wirel. Propag. Lett. 1(1), 31–34 (2002)

    Article  Google Scholar 

  10. Gonzáles García, S., Rubio Bretones, A., Gómez Martín, R., Hagness, S.C.: Accurate implementation of current sources in the ADI-FDTD scheme. IEEE Antennas Wirel. Propag. Lett. 3(1), 141–144 (2004)

  11. Gonzáles García, S., Godoy Rubio, R., Rubio Bretones, A., Gómez Martín, R.: On the dispersion relation of ADI-FDTD. IEEE Microw. Wirel. Compon. Lett. 16(6), 354–356 (2006)

    Article  Google Scholar 

  12. Hairer, E., Lubich, Ch., Wanner, G.: Geometric Numerical Integration, Structure-Preserving Algorithms for Ordinary Differential Equations, Volume 31 of Springer Series in Computational Mathematics. Springer, Berlin, Heidelberg (2006)

    Google Scholar 

  13. Hansen, E., Ostermann, A.: Dimension splitting for evolution equations. Numer. Math. 108, 557–570 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  14. Lee, J., Fornberg, B.: A split step approach for the 3-D Maxwell’s equations. J. Comput. Appl. Math. 158(2), 485–505 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  15. Lee, J., Fornberg, B.: Some unconditionally stable time stepping methods for the 3D Maxwell’s equations. J. Comput. Appl. Math. 166(2), 497–523 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  16. Leis, R.: Initial-Boundary Value Problems in Mathematical Physics. B. G. Teubner, Stuttgart (1986)

    Book  MATH  Google Scholar 

  17. Namiki, T.: A new FDTD algorithm based on alternating-direction implicit method. IEEE Trans. Microw. Theory Tech. 47(10), 2003–2007 (1999)

    Article  Google Scholar 

  18. Namiki, T.: 3-D ADI-FDTD method-unconditionally stable time-domain algorithm for solving full vector Maxwell’s equations. IEEE Trans. Microw. Theory Tech. 48(10), 1743–1748 (Oct 2000)

  19. Rauch, J.: Partial Differential Equations, Volume 128 of Graduate Texts in Mathematics. Springer, New York (1991)

    Google Scholar 

  20. Reed, M., Simon, B.: Methods of modern mathematical physics I. Functional Analysis, 2nd edn. Academic Press, Inc., Harcourt Brace Jovanovich, Publishers, New York (1980)

  21. Sz-Nagy, B.: Spektraldarstellung Linearer Transformationen des Hilbertschen Raumes. Berichtigter Nachdruck, Volume 39 of Ergebnisse der Mathematik und ihrer Grenzgebiete. Springer, Berlin, New York (1967) (reprint of the 1942 original edition edition)

  22. Taflove, A., Hagness, S.C.: Computational Electrodynamics: The Finite-Difference Time-Domain Method, 3rd edn. Artech House Publishers, Norwood (2005)

    Google Scholar 

  23. Tucsnak, M., Weiss, G.: Observation and control for operator semigroups. Birkhäuser, Basel (2009)

    Book  MATH  Google Scholar 

  24. Verwer, J.G.: Component splitting for semi-discrete Maxwell equations. BIT 51(2), 427–445 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  25. Verwer, J.G.: Composition methods, Maxwell’s equations, and source terms. SIAM J. Numer. Anal. 50(2), 439–457 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  26. Verwer, J.G., Botchev, M.A.: Unconditionally stable integration of Maxwell’s equations. Linear Algebra Appl. 431(3–4), 300–317 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  27. Yee, K.: Numerical solution of initial boundary value problems involving Maxwell’s equations in isotropic media. IEEE Trans. Antennas Propag. 14(3), 302–307 (1966)

    Article  MATH  Google Scholar 

  28. Zheng, F., Chen, Z., Zhang, J.: Toward the development of a three-dimensional unconditionally stable finite-difference time-domain method. IEEE Trans. Microw. Theory Tech. 48(9), 1550–1558 (2000)

    Article  Google Scholar 

Download references

Acknowledgments

We thank the referees for useful comments which in particular led to a simplification of the proof of Theorem 4.2.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Roland Schnaubelt.

Appendix

Appendix

We now present two proofs of Lemmas 3.6 and 3.7 we omitted in Sect. 3.

Proof of Lemma 3.6

Lax–Milgram provides us with a unique \(v\in H^1_\Gamma (Q)\) satisfying (21). To show the asserted regularity of \(v\), we consider the operators \(A_j=-\partial _j^2\) on \(L^2(Q)\) whose domain consists of those \(w\in L^2(Q)\) such that \(\partial _j^2 w\in L^2(Q)\), \(w=0\) on \(\Gamma _j^+\) or on \(\Gamma _j^-\) if \(\Gamma _j^+\subseteq \Gamma \) or if \(\Gamma _j^-\subseteq \Gamma \), respectively, and \(\partial _j w=0\) on \(\Gamma _j^+\) or on \(\Gamma _j^-\) if \(\Gamma _j^+\subseteq \Gamma '\) or if \(\Gamma _j^-\subseteq \Gamma '\), respectively. Here and below we have \(j=1,2,3\). For \(u\in D(A_j)\) and \(v\in D_j\), an integration by parts shows

$$\begin{aligned} \int _Q (uv +A_ju \, v)\,\mathrm {d}x=\int _Q (uv+\partial _j u\,\partial _j v)\,\mathrm {d}x=:a(u,v), \end{aligned}$$

where \(a\) is a symmetric, continuous and coercive bilinear form. It is routine to check that \(A_j\) is the self adjoint operator induced by \(a\). It is clear that \(A_j\) is positive. In particular, \(D_j\) is the domain of \(A_j^\frac{1}{2}\) and hence \(\partial _j A_j^{-\frac{1}{2}}\) is bounded on \(L^2(Q)\).

To see that the resolvents of \(A_i\) and \(A_j\) commute, we observe that the resolvent of, say, \(A_1\) is given by \(((\lambda I+ A_1)^{-1}f)(x,y,z)= (R_1(\lambda )f(\cdot , y,z))(x)\) for \(\lambda >0\), for almost every \((x,y,z)\in Q\) and the resolvent \(R_1(\lambda )\) of the negative second derivative on \(L^2(a_1^-,a_1^+)\) with the boundary conditions of \(A_1\). Analogous facts hold for \(A_2\) and \(A_3\). If \(f\) is the product of \(f_k\in L^2(a_k^-,a_k^+)\) for \(k=1,2,3\), then \((\lambda I+ A_i)^{-1}(\lambda I+ A_j)^{-1}f=(\lambda I+ A_j)^{-1}(\lambda I+ A_i)^{-1}f\). Since the span of such functions is dense in \(L^2(Q)\), the resolvents commute.

As explained in Sects. III.4, VII.2 and X.1 of [21], we thus have a joint functional calculus with respect to \(A_1\), \(A_2\) and \(A_3\) for Borel measurable functions \(\phi :{\mathbb {R}}^3_+\rightarrow {\mathbb {R}}\). The operator \(\phi (A_1,A_2,A_3)\) is bounded if \(\phi \) is bounded, and for \(h(\lambda )=1+\lambda _1+\lambda _2 +\lambda _3\) we have \(h(A_1,A_2,A_3)=I+A_1+A_2+A_3=:I+A\) on the domain \(D(A):=D(A_1)\cap D(A_2) \cap D(A_3)\). Set \(\rho =1/h\). Then \(\rho (A_1,A_2,A_3)\) is bounded and it is the inverse of \(I+A\), so that \(A\) is closed. Using the bounded functions \(h_{i,j}(\lambda )=\lambda _i^\frac{1}{2}\,\lambda _j^\frac{1}{2} \rho (\lambda )\), we see that the operator \(h_{ij}(A_1,A_2,A_3)=A_i^\frac{1}{2}\,A_j^\frac{1}{2}\, (I+A)^{-1}\) is bounded for all \(i,j\in \{1,2,3\}\). This means that \(D(A)\hookrightarrow H^2(Q)\) implying \(D(A)=D\) and the equivalence of graph norm of \(\Delta \) and the \(H^2\)-norm on \(D\). It is then clear that \(v=(I+A)^{-1}f\) is the required weak solution.\(\square \)

Proof of Lemma 3.7

1) Throughout, let \((\mathbf {E},\mathbf {H})\in D(M_0^2)\). It is known that a map \(u\in H(\mathrm {rot})\cap H(\mathrm {div})\) belongs to \(H^1(Q)^3\) if \(u\times \nu =0\) or \(u\cdot \nu =0\) holds on \(\partial Q\). Moreover, the \(H^1\) norm of \(u\) is then dominated by \(\Vert u\Vert _{L^2}+\Vert \mathrm{div }u\Vert _{L^2} +\Vert \mathrm{rot }u\Vert _{L^2}\), see, e.g., Theorem 2.17 in [1]. Note that the Eqs. (14) and (16) still hold on \(Q\). In particular \(\mathrm{div }\mathbf {E}\) and \(\mathrm{div }\mathbf {H}\) belong to \(L^2(Q)^3\). We thus have \(\mathbf {E},\mathbf {H}\in H^1(Q)^3\) and \(\Vert (\mathbf {E},\mathbf {H})\Vert _{H^1}\le c\, (\Vert (\mathbf {E},\mathbf {H})\Vert _X+\Vert M_0(\mathbf {E},\mathbf {H})\Vert _X)\). The asserted zero-order traces for \(\mathbf {E}\) and \(\mathbf {H}\) now are a direct consequence of the boundary conditions \(\mathbf {E}\times \nu =0\) and \(\mathbf {H}\cdot \nu =0\), respectively.

Since \(\mathbf {E},\mathbf {H}\in H^1(Q)^3\hookrightarrow L^6(Q)^3\) and \(M^2 (\mathbf {E},\mathbf {H})\in X\), Eq. (16) and the assumptions on \(\varepsilon \) and \(\mu \) imply that \(\Delta E_j, \Delta H_j\in L^2(Q)\). A standard localization argument then yields \(E_j,H_j\in H^2_{\text {loc}}(Q)^3\) for \(j=1,2,3.\) In addition, the \(X\)-norm of \((\Delta \mathbf {E}, \Delta \mathbf {H})\) is bounded by that of \(M_0^2(\mathbf {E},\mathbf {H})\) and \((\mathbf {E},\mathbf {H})\). We next establish the properties of the traces of \(\mathbf {E}\) and \(\mathbf {H}\) needed to derive \(\mathbf {E},\mathbf {H}\in H^2(Q)^3\) from Lemma 3.6.

2) We first consider \(E_1\). We will actually show that \(\varepsilon E_1\) belongs to \(H^2(Q)\) by applying Lemma 3.6 to \(\varepsilon E_1\). Because of

$$\begin{aligned} \partial _{kl}E_1 = \frac{1}{\varepsilon } \,\partial _{kl}(\varepsilon E_1)- \frac{\partial _k \varepsilon }{\varepsilon }\, \partial _l E_1 - \frac{\partial _l \varepsilon }{\varepsilon }\, \partial _k E_1 -\frac{\partial _{kl} \varepsilon }{\varepsilon }\,E_1, \end{aligned}$$
(38)

it will then follow that \(E_1\in H^2(Q)\) employing \(E_1\in H^1(Q)\) and the assumed regularity of \(\varepsilon \). At the present stage, from (38), \(\Delta E_1\in L^2 (Q)\) and \(E_1\in H^2_{\text {loc}}(Q)^3\) we can already infer that \(f:=(I-\Delta )(\varepsilon E_1)\in L^2(Q)\) and \(\varepsilon E_1\in H^2_{\mathrm {loc}}(Q)\). Part 1) shows that \(\varepsilon E_1=0\) on the faces \(\Gamma :=\Gamma _2^-\cup \Gamma _2^+\cup \Gamma _3^-\cup \Gamma _3^+\). Fix a function \(\psi \in H^1(Q)\) with \(\partial _2\psi ,\partial _3\psi \in H^1(Q)\) and having support in \([a_1^-,a_1^+]\times [a_2^-+\eta ,a_2^+-\eta ] \times [a_3^-+\eta ,a_3^+-\eta ]\) for some small \(\eta =\eta (\psi )>0\). A given \(\varphi \in H^1_\Gamma (Q)\) can be approximated in \(H^1(Q)\) by such \(\psi \) employing cutoff and mollification in the \((x_2,x_3)\) directions. For each sufficiently small \(\kappa >0\), we set

$$\begin{aligned} Q_\kappa&=(a_1^-+\kappa ,a_1^+-\kappa )\times (a_2^-+\kappa ,a_2^+-\kappa ) \times (a_3^-+\kappa ,a_3^+-\kappa ). \end{aligned}$$

We take \(\kappa \in (0,\eta (\psi ))\) and denote by \(\Gamma _1^\pm (\kappa )\) the open faces of \(Q_\kappa \) containing points of the form \((a_1^\pm \pm \kappa , x_2,x_3)\). Integrating by parts and using \(\mathrm{div }(\varepsilon \mathbf {E})=0\) as well as \(\partial _j(\varepsilon E_j)\in H^1_{\mathrm {loc}}(Q)\) for \(j=1,2,3\), we conclude that

$$\begin{aligned}&\int _Q \nabla (\varepsilon E_1)\cdot \nabla \psi \,\mathrm {d}x +\int _Q \varepsilon E_1\psi \,\mathrm {d}x = \lim _{\kappa \rightarrow 0}\int _{Q_\kappa } \big (\varepsilon E_1\psi + \nabla (\varepsilon E_1)\cdot \nabla \psi \big ) \,\mathrm {d}x \nonumber \\&\quad \quad = \lim _{\kappa \rightarrow 0}\left[ \int _{Q_\kappa } (I-\Delta ) (\varepsilon E_1)\psi \,\mathrm {d}x +\int _{\partial Q_\kappa } \psi \,\nabla (\varepsilon E_1)\cdot \nu \,\mathrm {d}\sigma \right] \nonumber \\&\quad \quad = \int _{Q} f\psi \,\mathrm {d}x \pm \lim _{\kappa \rightarrow 0} \int _{\Gamma _1^\pm (\kappa )} \psi \,\partial _1 (\varepsilon E_1) \,\mathrm {d}(x_2,x_3)\nonumber \\&\quad \quad =\int _{Q} f\psi \,\mathrm {d}x \mp \lim _{\kappa \rightarrow 0} \int _{\Gamma _1^\pm (\kappa )} \psi (\partial _2 (\varepsilon E_2)+\partial _3(\varepsilon E_3)) \,\mathrm {d}(x_2,x_3)\nonumber \\&\quad \quad = \int _{Q} f\psi \,\mathrm {d}x \pm \lim _{\kappa \rightarrow 0} \int _{\Gamma _1^\pm (\kappa ) } \big ( \varepsilon E_2 \partial _2 \psi + \varepsilon E_3 \partial _3\psi \big ) \,\mathrm {d}(x_2,x_3)\nonumber \\&\quad \quad = \int _Q f\psi \,\mathrm {d}x. \end{aligned}$$
(39)

We have used that \(\psi \) vanishes near \(\Gamma \) for the penultimate equation and that \(\varepsilon E_j, \partial _j \psi \in H^1(Q)^3\) and \(\varepsilon E_j=0\) on \(\Gamma _1^\pm \) for \(j=2,3\) in the last identity, see part 1). By approximation, Eq. (39) then holds for all \(\psi \in H^1_\Gamma (Q)\), and hence Lemma 3.6 yields \(\varepsilon E_1\in H^2(Q)\) so that \(E_1\in H^2(Q)\) as explained above. In the same way, one sees that \(E_2,E_3\in H^2(Q)\). Moreover, \(\Vert E_j\Vert _{H^2}\) is bounded by \(c\,(\Vert E_j\Vert _{L^2} + \Vert \Delta E_j\Vert _{L^2})\) due to Lemma 3.6 and hence by \(c\,(\Vert (\mathbf {E},\mathbf {H})\Vert _X+ \Vert M^2_0(\mathbf {E},\mathbf {H})\Vert _X)\) in view of step 1).

We denote by \(\gamma _i\) the trace operator to \(\Gamma _i^\pm \), where \(i,j,k\in \{1,2,3\}\). Since \(E_k\in H^2(Q)\), one can approximate \(E_k\) in \(H^2(Q)\) by \(v_n\in C^2(\overline{Q})\). Clearly, \(\gamma _i \partial _j v_n= \partial _j \gamma _i v_n\) and thus \(\gamma _i\partial _j E_k =\partial _j \gamma _i E_k\). As a result, the asserted first order boundary conditions of \(\mathbf {E}\) follow from the already established 0-order boundary conditions of \(\mathbf {E}\).

3) Next, we consider \(H_1\) and set \(g:=(I-\Delta ) H_1\in L^2(Q)\). Here we have less Dirichlet boundary conditions, namely \(H_j=0\) on \(\Gamma _j^\pm \) for \(j=1,2,3\). To deal with the Neumann conditions, we first note that

$$\begin{aligned}&\mathrm{rot }(\varepsilon ^{-1}\mathrm{rot }\mathbf {H})\in L^2(Q)^3, \qquad \varepsilon ^{-1}\mathrm{rot }\mathbf {H}\times \nu =0 \text {on}\partial Q,\\&\mathrm{div }(\varepsilon ^{-1}\mathrm{rot }\mathbf {H})= \nabla \varepsilon ^{-1}\cdot \mathrm{rot }\mathbf {H}\in L^2(Q). \end{aligned}$$

Hence, \(\varepsilon ^{-1}\mathrm{rot }\mathbf {H}\) belongs to \(H^1(Q)^3\) which yields \(\mathrm{rot }\mathbf {H}\in H^1(Q)^3\). It also follows that \(\mathrm{rot }\mathbf {H}\times \nu =0\) on \(\partial Q\). In particular, the first component of \(\mathrm{rot }\mathbf {H}\) vanishes on \(\Gamma _2^\pm \cup \Gamma _3^\pm \).

We set \(\tilde{\Gamma }= \Gamma _1 ^-\cup \Gamma _1^+\) and define the faces \(\Gamma _j^\pm (\kappa )\) of \(Q_\kappa \) in the \(j\)th direction for \(j=2,3\), cf. step 2). We take functions \(\psi \in H^1(Q)\) with \(\partial _1\psi \in H^1(Q)\) and having support in \([a_1^-+\eta ,a_1^+-\eta ]\times [a_2^-,a_2^+] \times [a_3^-,a_3^+]\) for some \(\eta >0\). We choose \(\kappa \in (0,\eta )\) so that \(\psi \) vanishes around \(\Gamma _1^\pm (\kappa )\). As above, we deduce

$$\begin{aligned} \int _{Q}&\nabla H_1\cdot \nabla \psi \,\mathrm {d}x +\int _Q H_1\psi \,\mathrm {d}x = \lim _{\kappa \rightarrow 0} \int _{Q_\kappa } \big (H_1\psi +\nabla H_1\cdot \nabla \psi \big )\,\mathrm {d}x\\&= \lim _{\kappa \rightarrow 0}\Big [\int _{Q_\kappa } \psi \,(I-\Delta ) H_1\,\mathrm {d}x + \int _{\partial Q_\kappa } \psi \,\nu \cdot \nabla H_1\,\mathrm {d}\sigma \Big ]\\&=\int _{Q} \psi \,(I-\Delta ) H_1\,\mathrm {d}x + \lim _{\kappa \rightarrow 0} \int _{\partial Q_\kappa } \big [\psi \,\nu \cdot \nabla H_1 -(\mathrm{rot }\mathbf {H}\times \nu )\cdot (\psi ,0,0)\big ] \,\mathrm {d}\sigma \\&=\int _{Q} g\psi \,\mathrm {d}x + \lim _{\kappa \rightarrow 0} \int _{\partial Q_\kappa } \psi \,\nu \cdot \partial _1 \mathbf {H}\,\mathrm {d}\sigma \\&=\int _{Q} g\psi \,\mathrm {d}x \pm \lim _{\kappa \rightarrow 0}\Big [ \int _{\Gamma _2^\pm (\kappa )} \psi \, \partial _1 H_2 \,\mathrm {d}\sigma + \int _{\Gamma _3^\pm (\kappa )} \psi \, \partial _1 H_3 \,\mathrm {d}\sigma \Big ]\\&=\int _{Q} g\psi \,\mathrm {d}x \mp \lim _{\kappa \rightarrow 0}\Big [ \int _{\Gamma _2^\pm (\kappa )} H_2\, \partial _1 \psi \,\mathrm {d}\sigma + \int _{\Gamma _3^\pm (\kappa )} H_3\, \partial _1 \psi \,\mathrm {d}\sigma \Big ]\\&= \int _{Q} g\psi \,\mathrm {d}x. \end{aligned}$$

The remaining assertions now follow as in step 2.\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hochbruck, M., Jahnke, T. & Schnaubelt, R. Convergence of an ADI splitting for Maxwell’s equations. Numer. Math. 129, 535–561 (2015). https://doi.org/10.1007/s00211-014-0642-0

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00211-014-0642-0

Mathematics Subject Classification (2010)

Navigation