Skip to main content

Introduction to Yasuura’s Method of Modal Expansion with Application to Grating Problems

  • Chapter
  • First Online:
The Generalized Multipole Technique for Light Scattering

Abstract

In this chapter we introduce the theory of the Yasuura’s method based on modal expansion and explain the methods of numerical computation in detail for several grating problems. After a sample problem we discuss the methods for solving two types of problems that require additional knowledge and steps, that is, scattering by a dielectric cylinder and diffraction by a grating. Some numerical results are shown to give an evidence of an experimental rule for the number of linear equations in formulating the least-squares problem that determines the modal coefficients. After confirming the rule we show a couple of examples of practical interest, i.e., scattering by a relatively deep metal grating, plasmon surface waves on a metal grating placed in conical mounting, scattering by a metal surface modulated in two directions, and scattering by periodically located dielectric spheres. To provide supplementary explanations of particular problems, four appendices are given; H-wave scattering from a cylinder, the normal equation and related topics, conical diffraction by a dielectric grating, and comparison of modal functions and the algorithm of the smoothing procedures.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 109.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 139.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 139.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    The reason why we set a limit “2-D” is that the SP, in the present form, is available only in 2-D problems. This is because we employ an indefinite integral to realize a low-pass spatial filter.

  2. 2.

    It is also termed Transverse-Electric (TE) wave, which means the electric field is orthogonal to the xy-plane. While in the H-wave (or TM-wave) the magnetic field has the z-component alone.

  3. 3.

    The component is called a leading field if it gives other nonzero components as in (8.3). Note that the derivation of \(\mathbf{H}^\mathrm{s}\) by (8.3) is a proper procedure because the sequence of our approximate solutions converges to the true solution uniformly in wider sense in the exterior region \(\mathrm{S}_\mathrm{e}\) as we will see later.

  4. 4.

    We hope the readers consult a treatise on Functional Analysis, e.g. [14], in case of need.

  5. 5.

    If C is a circle centered at the origin, it is apparent that the sets of boundary values and normal derivatives are both complete because the members of each set are nothing other than the Fourier bases . Even in case if C is not a circle, the sets are still complete because of Example 2: let L be a circle centered at the origin and take the Fourier bases for \(f_m(t)\) in (8.8), then we get a set of separated solutions.

  6. 6.

    This is not a strong exception because we can modify the contour L (and hence \(\mathrm{D}_\mathrm{i}\)) slightly to avoid the coincidence. Example 2 is a key theorem of generation of complete sets, which has been proven by Yasuura and Itakura [39] as an analogy of Runge’s (or Runge-Walsh’s) theorem known in Theory of Complex Functions.

  7. 7.

    Unfortunately, separated solutions are not very efficient in a problem where C is strongly modulated from a circle (or, in general, a coordinate surface of the system of coordinates employed). As an example, we show a comparison between types of modal functions: the separated solutions (8.7) and monopole fields (8.9) in Appendix 4.

  8. 8.

    This dependence is natural because the boundary values of modal functions, in general, do not form an orthogonal set in H. This type of summation is usually called a flexible summation. Note that the approximate solution is defined in a finite summation of modal functions. By considering a sequence of finite-sum solutions, we can avoid the constraint of the convergence area of an infinite series solution. Yasuura’s original papers [38,39,40] has been written from this point of view. Reference [9] includes an interpretation of the difference between series and sequence solutions.

  9. 9.

    \(G(\mathbf{r},\mathbf{r}')\) is a total electric field observed at \(\mathbf{r}'\,(\not = \mathbf{r})\) when a unit line source is placed at \(\mathbf{r}\) in Fig. 8.1. Note that employment of the Green function satisfying (8.12) is for convenience and is not essential: The whole theory has been established in [38,39,40], where the free-space Green function alone was used.

  10. 10.

    This kind of convergence is called uniform convergence in wider sense in \(\mathrm{S}_\mathrm{e}\).

  11. 11.

    Until the middle of 80s we employed normal equations (NE) in solving LSP 1. Now we solve the problem using the method in Sect. 8.2.3.2. We state the reason why we stopped using the NE and attach some comments in Appendix 2.

  12. 12.

    On the other hand, there is a possibility to make possible use of the weighting function accompanying the variable transformation. For example, a Schwarz–Christoffel-type transformation works to remove the singularity of Green’s function in a problem of an edged cross section [23].

  13. 13.

    It is more reasonable to ask “How many linear equations do we need?” This is because (i) we get two equations at one sampling point in a 2-media problem (see Sect. 8.2.4); and (ii) we should understand (8.31) as a relation between the numbers of equations J and unknowns M.

  14. 14.

    In addition, the solution by a QRD program, usually, is not inferior in accuracy to one by an SVD program. This may be because of the greater computational complexity of the SVD.

  15. 15.

    The machine epsilon, EPS, is the minimum positive number that satisfies \(1 + \mathrm{EPS} > 1\) in the floating-point system employed.

  16. 16.

    To get the latter we set \(\mathbf{u}_\nu \times (\mathbf{H}_\mathrm{e} - \mathbf{H}_\mathrm{i}) = 0\). Insertion of \(\mathbf{H}_\mathrm{e} = (\mathrm{i}/\omega \mu )\nabla E_\mathrm{e}\times \mathbf{u}_z\) etc. finds the desired relation. Here, \(E_\mathrm{e} = F + \varPsi _\mathrm{e}\) stands for the total electric field in \(\mathrm{S}_\mathrm{e}\).

  17. 17.

    We cannot include the derivation of equations from (8.45) through (8.49) because it takes much space. Interested readers can find the details in [35, 43,44,45]. The paper by Petit and Cadilhac [33] is also helpful.

  18. 18.

    The use of intrinsic impedance is also possible and is widely employed. That is: find the coefficients by minimization of \(|\text{ error } \text{ in } \text{ E }|^2 + Z_0^2 |\text{ error } \text{ in } \text{ H }|^2\). Here, \(Z_0\) is the intrinsic impedance of vacuum or surrounding material. We use this formulation in Appendix 3.

  19. 19.

    If the compensation by \(\gamma \) is not necessary, we can set \(p=1/ \mathbf{f\,}^\dag \mathbf{f}\) and \(q=1/ \mathbf{g}^\dag \mathbf{g}\) or use the intrinsic impedance.

  20. 20.

    s stands for senkrecht (German) , which means the electric field is perpendicular to the plane of incidence, the plane spanned by \(\mathbf{u}_Y\) (grating normal) and the incident wavevector.

  21. 21.

    The superscript E denotes that the coefficients concern the E-wave. Later we will also use the superscripts H, TE, and TM in accordance with polarizations.

  22. 22.

    If we employ the SP, this correspondence is essentially important because we need periodicity of the functions defined on the boundary. In using Yasuura’s method without the SP, we can say the following points: (1) If we get the solution through the NE, this modification is not necessary because it is done automatically in calculating the inner products; (2) While if we employ the QRD or SVD: (2.i) The modification may accelerate the convergence of the solutions because the target function and the modal functions are periodically continuous after modification; (2.ii) And, a quadrature by parts (or rectangular-rule) approximation is equivalent to a trapezoidal-rule in numerical integrations.

  23. 23.

    Note that the bias setting (we used \({-}1\) here) has an effect on the accuracy of numerical computation when the grating is deep.

  24. 24.

    Although the use of normalization by wavelength (i.e., \(kd = 2\pi d/\lambda \) etc.) is convenient in handling a problem with a PC obstacle, we employ real length here.

  25. 25.

    \(\rho _m\) is referred to as the (reflection) efficiency of the mth order.

  26. 26.

    When the number of truncation is small (e.g., \(N \le 10\)), we sometimes observe a phenomenon that the condition number continues to decrease slightly beyond \(J=2M\) due to tiny increment of \(\sigma _\mathrm{min}\).

  27. 27.

    We should notice, however, that the accuracy of an H-wave solution is lower than that of an E-wave solution by one or two digits. This is observed generally; and was Yasuura’s motivation of introducing the SP. His idea came from the fact that a Neumann problem for an electrostatic potential is equivalent to a Dirichlet problem for a stream function. The prototype of the SP, hence, was called an algorithm using the stream function in a wave field.

  28. 28.

    The efficiencies are given by \(\rho _{m}^\mathrm{TE}=\left( \gamma _{1m}/\gamma _{10} \right) |A_{1m}^\mathrm{TE}|^2\) and \(\rho _{m}^\mathrm{TM}=\left( \gamma _{1m}/\gamma _{10} \right) |A_{1m}^\mathrm{TM}|^2\) where \(\gamma _{1m}\) is the propagation constant in the Z-direction of the mth-order propagating mode \(\left( \mathrm{Re} \left( \gamma _{1m} \right) \ge 0 \right) \) concerning the upper region \(\text{ V }_1\), and \(A_{1m}^\mathrm{TE}\) and \(A_{1m}^\mathrm{TM}\) are the expansion coefficients of the approximate solutions defined in (8.120) of Appendix 3.

  29. 29.

    \(\rho ^\mathrm{Total}\) is a summation of \(\rho _m\) over the propagating orders.

  30. 30.

    \(1-\rho ^\mathrm{Total}\) represents the ratio of the absorbed light power by a metal grating to the incident light power.

  31. 31.

    The vector \(\mathbf{m}_{mn}^\mathrm{e,h}(\mathbf{r}_q)\) is perpendicular to the \(r_q\) axis, whereas \(\mathbf{n}_{mn}^\mathrm{e,h}(\mathbf{r}_q)\) has an \(r_q\) component. That is, the superscript TE (TM) in (8.89) means transverse electric (transverse magnetic) with respect to \(r_q\).

  32. 32.

    If \(f \in \varPhi _N\), then \(E_N = 0\). This, however, cannot occur in practice: For example, even in the case of scattering from a circular cylinder made of a PC, we need an infinite series to represent a rigorous solution because the boundary value has a form \(\exp [-\mathrm{i}ka\cos (\theta _s - \iota )]\). In addition, note that \(\varPhi _N\) is closed.

  33. 33.

    We get (8.104) by setting \((\varphi _m, \varPsi _N-f) = 0\) \((m=0, \pm 1,\ldots , \pm N)\); or from \(\partial E_N/\partial \overline{A}_m = 0\).

  34. 34.

    Assume a PC surface-relief grating with a TE-wave incidence, for simplicity, and imagine the surface current induced. It apparently has a Z-oriented ingredient, which excites a TM-wave component.

  35. 35.

    \(\mathbf{e}^\mathrm{TE}\) is perpendicular to the plane of incidence; the fact that the magnetic field accompanying \(\mathbf{e}^\mathrm{TM}\) is orthogonal to the plane can be seen by manipulation.

  36. 36.

    We can use monopole fields also in the grating problems discussed in Sect. 8.2.5. A countably infinite set of monopoles located periodically in x—i.e., the location is given by \((x_1 + \ell d, y_1)\) \((0< x_1 \le d;\; y_1 < \eta (x_1);\; \ell = 0, \pm 1, \pm 2, \ldots )\)—radiates a plane wave [4, 36] satisfying (GD1) and (GD2). If we let the monopoles be accompanied by phase factors \(\exp (\mathrm{i}\ell kd\sin \theta )\), the plane wave meets the periodicity (GD3). Increasing the number of monopoles in the first strip region to M, i.e., \((x, y) = (x_1, y_1), (x_2, y_2), (x_3, y_3), \ldots , (x_M, y_M)\), and repeating the same procedure, we have a set of M plane waves, which is the desired set of modal functions [28, 37].

  37. 37.

    Although the employment of polyphase wave functions is effective because of the periodicity, we do not use them for simplicity.

  38. 38.

    According to the result of numerical computation, an optimum d was in the rage [0.85, 0.90] when the total number of poles was between 40 and 120. If we increased (or decreased) the number of poles, the optimum d approached 0.90 (or 0.85). Note, however, that the trends were observed in solving a particular problem with specific computational parameters and are no more than reference data.

  39. 39.

    The result of sample calculation has shown that the use of \(|\text{ BC }|^\alpha \) (\(\alpha > 1\)) instead of BC (i.e., further emphasis of the convex part in locating poles) gives better solutions.

  40. 40.

    We have applied the rule \(J=2M\) and have omitted J.

  41. 41.

    The relation is referred to as the optical theorem, which implies energy conservation.

  42. 42.

    We have employed the monopole fields and have seen their effectiveness [30]. It is worth noting that inclusion of a few dipoles located near the convex part of L in addition to the monopoles improves the efficiency greatly. This might be related to Cadilhac-Petit’s opinion [7] in locating the poles near an internal focus.

  43. 43.

    We got the elements under the assumption that the length of C is 1. This is convenient in mathematical analysis and does not affect applications to obstacles made of a lossless material including PC. In dealing with a lossy material, in particular a metal in light frequency, the normalization should be accompanied by a law of similitude in time-dependent EM field [19] and, hence, the use of actual length might be appropriate.

References

  1. M. Bass (ed.), Handbook of Optics; Volume II — Devices, Measurements, and Properties, 2nd edn. (McGraw-Hill, 1995)

    Google Scholar 

  2. R.H.T. Bates, Analytic constraints on electromagnetic field computations. IEEE Trans. Microw. Theory Tech. MTT-23(8), 605–623 (1975)

    Google Scholar 

  3. R.H.T. Bates, J.R. James, I.N.L. Gallett, R.F. Millar, An overview of point matching. Radio Electron. Eng. 43(3), 193–200 (1973)

    Article  Google Scholar 

  4. A. Boag, Y. Leviatan, A. Boag, Analysis of two-dimensional electromagnetic scattering from a periodic grating of cylinders using a hybrid current method. Radio Sci. 23(4), 612–624 (1988)

    Article  ADS  Google Scholar 

  5. G.P. Bryan-Brown, J.R. Sambles, M.C. Hutley, Polarization conversion through the excitation of surface plasmons on a metallic grating. J. Modern Opt. 37(7), 1227–1232 (1990)

    Article  ADS  Google Scholar 

  6. A.P. Calderón, The multipole expansion of radiation fields. J. Ration. Mech. Anal. (J. Math. Mech.) 3, 523–537 (1954)

    Google Scholar 

  7. M. Cadilhac, R. Petit, On the diffraction problem in electromagnetic theory: a discussion based on concepts of functional analysis including an example of practical application, in Huygens’ Principle 1690–1990: Theory and Applications, Studies in Mathematical Physics, ed. by H. Blok, et al. (Elsevier, Amsterdam, 1992)

    Google Scholar 

  8. G. Hass, L. Hardley, Optical properties of metal, in American Institute of Physics Handbook, ed. by D.E. Gray, 2nd ed. (McGraw-Hill, 1963), pp. 6–107

    Google Scholar 

  9. J.P. Hugonin, R. Petit, M. Cadilhac, Plane-wave expansions used to describe the field diffracted by a grating. J. Opt. Soc. Am. 71(5), 593–598 (1981)

    Article  ADS  Google Scholar 

  10. H. Ikuno, K. Yasuura, Numerical calculation of the scattered field from a periodic deformed cylinder using the smoothing process on the mode-matching method. Radio Sci. 13(6), 937–946 (1978)

    Article  ADS  Google Scholar 

  11. H. Ikuno, M. Gondoh, M. Nishimoto, Numerical analysis of electromagnetic wave scattering from an indented body of revolution. Trans. IEICE Electron. E74-C(9), 2855–2863 (1991)

    Google Scholar 

  12. T. Inagaki, J.P. Goudonnet, J.W. Little, E.T. Arakawa, Photoacoustic study of plasmon-resonance absorption in a bigrating. J. Opt. Soc. Am. B 2(3), 433–439 (1985)

    Article  ADS  Google Scholar 

  13. M. Kawano, H. Ikuno, M. Nishimoto, Numerical analysis of 3-D scattering problems using the Yasuura method. Trans. IEICE Electron. E79-C(10), 1358–1363 (1996)

    Google Scholar 

  14. A.N. Kolmogorov, S.V. Fomin, Elements of the Theory of Functions and Functional Analysis (Dover, New York, 1999)

    Google Scholar 

  15. C.L. Lawson, R.J. Hanson, Solving Least Squares Problems (Prentice-Hall, New Jersey, 1974)

    Google Scholar 

  16. T. Matsuda, D. Zhou, Y. Okuno, Numerical analysis of plasmon-resonance absorption in a bisinusoidal metal grating. J. Opt. Soc. Am. A 19(4), 695–701 (2002)

    Google Scholar 

  17. A. Matsushima, Y. Momoka, M. Ohtsu, Y. Okuno, Efficient numerical approach to electromagnetic scattering from three-dimensional periodic array of dielectric spheres using sequential accumulation. Progr. Electromagn. Res. 69, 305–322 (2007)

    Article  Google Scholar 

  18. R.F. Millar, Rayleigh hypothesis and a related least-squares solution to scattering problems for periodic surfaces and other scatterers. Radio Sci. 8(8–9), 785–796 (1973)

    Article  ADS  MathSciNet  Google Scholar 

  19. H. Nakano, Frequency-independent antennas: spirals and log-periodics, in Modern Antenna Handbook, ed. by C.A. Balanis (Wiley, New Jersey, 2008), pp. 263–323

    Google Scholar 

  20. Y. Nakata, M. Koshiba, M. Suzuki, Finite-element analysis of plane wave diffraction from dielectric gratings. Trans. IEICE Jpn. J69-C(12), 1503–1511 (1986)

    Google Scholar 

  21. M. Neviér, The homogeneous problems, in Electromagnetic Theory of Gratings, ed. by R. Petit (Springer, Berlin, 1980), pp. 123–157

    Google Scholar 

  22. M. Ohtsu, Y. Okuno, A. Matsushima, T. Suyama, A Combination of up- and down-going Floquet modal functions used to describe the field inside grooves of a deep grating. Progr. Electromagn. Res. 64, 293–316 (2006)

    Article  Google Scholar 

  23. Y. Okuno, A numerical method for solving edge-type scattering problems. Radio Sci. 22(6), 941–946 (1987)

    Article  ADS  Google Scholar 

  24. Y. Okuno, The mode-matching method, in Analysis Methods in Electromagnetic Wave Problems, ed. by E. Yamashita (Artech House, 1990), pp. 107–138

    Google Scholar 

  25. Y. Okuno, An introduction to the Yasuura method, in Analytical and Numerical Methods in Electromagnetic Wave Theory, ed. by M. Hashimoto, M. Idemen, O.A. Tretyakov (Science House, 1993), pp. 515–565

    Google Scholar 

  26. Y. Okuno, H. Ikuno, Completeness of the boundary values of equivalent sources. Mem. Fac. Eng. Kumamoto Univ. 38(1), 1–8 (1993)

    ADS  Google Scholar 

  27. Y. Okuno, H. Ikuno, Yasuura’s method, its relation to the fictitious source methods, and its advancements in the solution of 2D problems, in Generalized Multipole Techniques for Electromagnetic and Light Scattering, ed. T. Wriedt (Elsevier, Amsterdam, 1999)

    Google Scholar 

  28. Y. Okuno, T. Matsuda, T. Kuroki, Diffraction efficiency of a grating with deep grooves, in Proceedings of the 1995 Sino-Japanese Joint Meeting on Optical Fiber Science and Electromagnetic Theory (OFSET’95), vol. 1 (Tianjin, China, 1995), pp. 106–111

    Google Scholar 

  29. Y. Okuno, T. Suyama, R. Hu, S. He, T. Matsuda, Excitation of surface plasmons on a metal grating and its application to an index sensor. Trans. IEICE Electron. E90-C(7), 1507–1514 (2007)

    Google Scholar 

  30. Y. Okuno, H. Yamaguchi, The idea of equivalent sources in the Yasuura method, in Proceedings 1992 International Symposium on Antennas Propagat (ISAP’92), vol. 1E3-2 (Sapporo, Japan, 1992)

    Google Scholar 

  31. Y. Okuno, K. Yasuura, Numerical algorithm based on the mode-matching method with a singular-smoothing procedure for analysing edge-type scattering problems. IEEE Trans. Antennas Propagat. 30(4), 580–587 (1982)

    Article  ADS  MATH  Google Scholar 

  32. R. Petit (ed.), Electromagnetic Theory of Gratings (Springer, Berlin, 1980)

    Google Scholar 

  33. R. Petit, M. Cadilhac, Electromagnetic theory of gratings: some advances and some comments on the use of the operator formalism. J. Opt. Soc. Am. A 7(9), 1666–1674 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  34. H. Raether, Surafce plasmon and roughness, in Surface Polaritons — Electromagnetic Waves at Surfaces and Interfaces, ed. by V.M. Agranovich, D.L. Mills (North Holland, 1982), pp. 331–403

    Google Scholar 

  35. M. Tomita, K. Yasuura, The Rayleigh expansion theorem for the boundary value problem in two media. Kyushu Univ. Tech. Rep. 52(2), 142–154 (1979)

    Google Scholar 

  36. J.R. Wait, Reflection from a wire grid parallel to a conducting plane. Can. J. Phys. 32, 571–579 (1954)

    Article  ADS  MATH  Google Scholar 

  37. X. Xu, B.W. Chen, R. Gong, M. Zheng, Use of auxiliary source fields in Yasuura’s method, in Proceedings of the 2017 IEEE International Conference on Computational Electromagnetics (ICCEM2017), vol. 2C1.2 (Kumamoto, Japan, 2017)

    Google Scholar 

  38. K. Yasuura, T. Itakura, Approximation method for wave functions (I). Kyushu Univ. Tech. Rep. 38(1), 72–77 (1965)

    Google Scholar 

  39. K. Yasuura, T. Itakura, Complete set of wave functions – approximation method for wave functions (II). Kyushu Univ. Tech. Rep. 38(4), 378–385 (1966)

    Google Scholar 

  40. K. Yasuura, T. Itakura, Approximation algorithm by complete set of wave functions – approximation method for wave functions (III). Kyushu Univ. Tech. Rep. 39(1), 51–56 (1966)

    Google Scholar 

  41. K. Yasuura, H. Ikuno, Smoothing process on the mode-matching method for solving two-dimensional scattering problems. Mem. Fac. Eng. Kyushu Univ. 37(4), 175–192 (1977)

    Google Scholar 

  42. K. Yasuura, Y. Okuno, Singular-smoothing procedure on Fourier analysis. Mem. Fac. Eng. Kyushu Univ. 41(2), 123–141 (1981)

    Google Scholar 

  43. K. Yasuura, M. Tomita, Convergency of approximate wave functions on the boundary – the case of inner domain. Kyushu Univ. Tech. Rep. 52(1), 79–86 (1979)

    Google Scholar 

  44. K. Yasuura, M. Tomita, Convergency of approximate wave functions on the boundary – the case of outer domain. Kyushu Univ. Tech. Rep. 52(1), 87–93 (1979)

    Google Scholar 

  45. K. Yasuura, M. Tomita, Numerical analysis of plane wave scattering from dielectric cylinders. Trans. IECE Jpn. 62-B(2), 132–139 (1979)

    Google Scholar 

  46. K.A. Zaki, A.R. Neureuther, Scattering from a perfectly conducting surface with a sinusoidal height profile: TE polarization. IEEE Trans. Antennas Propagat. AP-19(2), 208–214 (1971)

    Google Scholar 

Download references

Acknowledgements

The authors thank Mr. BenWen Chen and Mr. Rui Gong, Centre for Optical and Electromagnetic Research, South China Academy of Advanced Optoelectronics, South China Normal University for preparing the figures in Sect. 8.3.1 including numerical computations.

One of the authors (A.M.) wish to express his thanks to Japan Society for Promotion of Science (JSPS) for partial support to the work in Sects. 8.3.2 and 8.3.5 under Grant Number JP15K06023 (KAKENHI).

Another one of the authors (Y.O.) is grateful to Prof. S. He, COER-SCNU, COER-ZJU, and JORSEP-KTH for his continuous help and encouragement.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Akira Matsushima .

Editor information

Editors and Affiliations

Appendices

Appendix 1: H-Wave Scattering by a PC Cylinder

Let us consider a problem where an H-wave (TM-wave) is incident to the obstacle shown in Fig. 8.1. That is, the incident wave is polarized in the xy-plane so that the incident magnetic field has only a z-component

$$\begin{aligned} \mathbf{H}^\mathrm{i}(\mathbf{r}) = \mathbf{u}_z F(\mathbf{r}) = \mathbf{u}_z \exp [-\mathrm{i}kr \cos (\theta -\iota )]. \end{aligned}$$
(8.92)

The scattered magnetic field has only a z-component

$$\begin{aligned} \mathbf{H}^\mathrm{s}(\mathbf{r}) = \mathbf{u}_z \varPsi (\mathbf{r}) \end{aligned}$$
(8.93)

which is a leading field of the problem. Thus, we have

Problem 1’: H-wave, PC. Find \(\varPsi (\mathbf{r})\) that satisfies:  

(N1):

The 2-D Helmholtz equation in \(\mathrm{S}_\mathrm{e}\);

(N2):

The 2-D radiation condition at infinity;

(N3):

The boundary condition

$$\begin{aligned} \partial _\nu \varPsi (s) = g(s) \equiv -\partial _\nu F(s) \quad (s \in \text{ C }). \end{aligned}$$
(8.94)

Here, \(\partial _\nu \) denotes a normal derivative at s. Equation (8.94) is called Neumann’s or the second-kind boundary condition.

 

Employing the Green’s (or Neumann’s) function of this boundary-value problem satisfying a homogeneous boundary condition

$$\begin{aligned} \partial _\nu N(\mathbf{r}, s) = 0\quad (\mathbf{r}\in \mathrm{S}; s \in \mathrm{C}), \end{aligned}$$
(8.95)

we get a formal representation similar to (8.13)

$$\begin{aligned} \varPsi _N(\mathbf{r}) - \varPsi (\mathbf{r}) = -\int \limits _{s=0}^C N(\mathbf{r},s) \left[ \partial _\nu \varPsi _N(s) - g(s)\right] \,ds\quad (\mathbf{r}\in \mathrm{S}). \end{aligned}$$
(8.96)

Here, \(\varPsi _N\) denotes an approximate solution defined by

$$\begin{aligned} \varPsi _N(\mathbf{r}) = \sum _{m=-N}^N B_m(M)\varphi _m(\mathbf{r}). \end{aligned}$$
(8.97)

After a discussion similar to that in Sect. 8.2.2.3, we have a least-squares problem for the H-wave problem:

LSM 1’: H-wave, PC. Find the coefficients \(B_m(M)\) \((m=0, \pm 1, \ldots , \pm N)\) that minimize the mean-squares boundary residual

$$\begin{aligned} E_N = \frac{\left\| \partial _\nu \varPsi _N - g \right\| ^2}{\Vert g \Vert ^2} = \frac{1}{\Vert g \Vert ^2} \left\| \sum _{m=-N}^N B_m(M)\partial _\nu \varphi _m - g \right\| ^2. \end{aligned}$$
(8.98)

We can solve this problem on a computer following the procedure in Sect. 8.2.3. Approximations to other nonzero components can be found by

$$\begin{aligned} \mathbf{E}_N^\mathrm{s}(\mathbf{r}) = \frac{1}{\mathrm{i}\omega \varepsilon _0}\nabla \varPsi _N(\mathbf{r}) \times \mathbf{u}_z. \end{aligned}$$
(8.99)

It is worth noting that in an H-wave scattering from a dielectric obstacle, the boundary condition (8.42) should be altered slightly. Let \(\mathbf{H}^\mathrm{s}(\mathbf{r})=\mathbf{u}_z\varPsi _\mathrm{e}(\mathbf{r})\), and \(\mathbf{H}^\mathrm{t}(\mathbf{r})=\mathbf{u}_z\varPsi _{\,\mathrm i}(\mathbf{r})\), then we have

$$\begin{aligned} \left\{ \begin{array}{l} \varPsi _\mathrm{e}(s)-\varPsi _{\,\mathrm i}(s)=f(s)\equiv -F(s)\\ \partial _\nu \varPsi _\mathrm{e} - n^{-2} \partial _\nu \varPsi _{\,\mathrm i}(s) = g(s) \equiv -\partial _\nu F(s), \end{array}\right. \end{aligned}$$
(8.100)

where the second line means the electric-field continuity and \(n^2 = \varepsilon /\varepsilon _0\).

Appendix 2: Solution of LSP 1 by a Normal Equation and Related Topics

Although we do not use a normal equation in numerical analysis, we look over the solution method by the equation because it is an important theoretical tool in working with a least-squares problem. Let us define an inner product between two functions in \(\mathbf{H}= L^2(0, C)\) by

$$\begin{aligned} (f,g)=\int \limits _{s=0}^C \overline{f(s)}g(s)\,ds, \end{aligned}$$
(8.101)

then we find that \(\Vert f \Vert = \sqrt{(f,f)}\). Employing these relations, we modify (8.22) to obtain

$$\begin{aligned} E_N = \!\! \sum _{m=-N}^N \sum _{n=-N}^N \overline{A_m}(\varphi _m,\varphi _n) A_n - \!\! \sum _{m=-N}^N \overline{A_m} (\varphi _m,f) - \!\! \sum _{n=-N}^N (f,\varphi _n)A_n + \Vert f \Vert ^2. \end{aligned}$$
(8.102)

The predictable M is not shown.

Now we define a subspace of \(\mathbf{H}\), \(\varPhi _N\), spanned by the boundary values of a finite number of modal functions \(\{\varphi _0(s),\varphi _{\pm 1}, \ldots ,\) \(\varphi _{\pm N}\}\). An element of \(\varPhi _N\) can be represented as

$$\begin{aligned} \varPsi _N(s) = \sum _{n=-N}^N A_n\,\varphi _n(s). \end{aligned}$$
(8.103)

Apparently, there is a minimum value of \(E_N\), which is a squared distance between f(s) and a point in \(\varPhi _N\).Footnote 32 The minimum is achieved when (8.103) agrees with the foot of a perpendicular line from f(s) to the surface of \(\varPhi _N\). The necessary and sufficient condition for this is that: The \(A_m\) coefficients are the solutions of the set of linear equations

$$\begin{aligned} \sum _{n=-N}^N (\varphi _m,\varphi _n)\,A_n = (\varphi _m, f)\quad (m=0,\pm 1,\ldots ,\pm N). \end{aligned}$$
(8.104)

This is referred to as the normal equation (NE) of LSP 1 and is a formal solution to the problem.Footnote 33

Next, let us consider the minimization from a computational point of view. That is, we try to find the \(A_m\) coefficients using the sampled values of boundary functions; and the functions are represented by J-dimensional complex-valued vectors \(\mathbf{f}\), \(\varvec{\varphi }_m\), and \(\varvec{\varPsi }_N\) as in Sect. 8.2.3. This leads us to DLSP 1. We know the orthogonal decompositions are useful tools for solving the problem. However, setting them aside, we here consider a NE based on DLSP 1. Because the Jacobian matrix \(\varPhi \) is \(J \times M\) \((J > M)\), the set of linear equations

$$\begin{aligned} \varPhi \mathbf{A} = \mathbf{f} \end{aligned}$$
(8.105)

is over-determined and does not have a usual solution. However, if we multiply (8.105) by \(\varPhi ^\dag \) from the left, we have

$$\begin{aligned} \text{ H } \mathbf{A} = \mathbf{b}, \end{aligned}$$
(8.106)

where

$$\begin{aligned} \text{ H } = \varPhi ^\dag \varPhi \end{aligned}$$
(8.107)

is an \(M \times M\) positive-definite Hermitian matrix provided \(\varPhi \) is full rank. And,

$$\begin{aligned} \mathbf{b} = \varPhi ^\dag \mathbf{f} \end{aligned}$$
(8.108)

is an M-dimensional right-hand side. Usually, (8.106) is referred to as the NE of DLSP 1 and has been employed as a standard method of solution for a long time.

Obviously, (8.106) is an approximation of (8.104). For example, an (mn)th element of the coefficient matrix of (8.104) can be represented as

$$\begin{aligned} (\varphi _m,\varphi _n) = \int \limits _{s=0}^C\overline{\varphi _m(s)}\varphi _n(s)\, ds \simeq \frac{C}{J}\sum _{j=1}^J\overline{\varphi _m(s_j)}\varphi _n(s_j) = \frac{C}{J}\,\varvec{\varphi }_m^\dag \varvec{\varphi }_n. \end{aligned}$$
(8.109)

The right-hand side of (8.109) is the (mn)th entry of H multiplied by the line element C / J. Hence, (8.104) and (8.106) are essentially the same thing, and they have common weak points in numerical computations. Widely-accepted key observations are:

  • The NE is rigorous, in principle, and can be employed in theoretical considerations;

  • The NE combined with Gaussian elimination (diagonal pivoting assumed) is equivalent to the (modified) Schmidt QRD except for the next two items;

  • The NE may lose information in constructing \(\mathrm{H} = \varPhi ^\dag \varPhi \), and this process is time consuming usually;

  • The NE is dominated by the condition number of H that is square of the original condition number: \(\text{ cond }(\text{ H }) = [\text{ cond }(\varPhi )]^2\).

The last item means (8.104) and (8.106) are more sensitive to computational errors than LSP 1 and DLSP 1. Therefore, the NE’s are more difficult to solve on a computer than the original least-squares problems. We, hence, do not recommend the use of (8.104) or (8.106). Even if we are working in the case where the inner products in (8.104) can be calculated analytically, we should not employ (8.104) because of the last item.

Before closing this Appendix, we would like to state a couple of comments on (8.105). Apparently, J cannot be less than M because (8.105) is indeterminate for \(J < M\). If we set \(J = M\), we have a point-matching method (PMM) or a collocation method. The method is known to be effective if the contour C coincides with a part of a coordinate curve of a system of coordinates in which Helmholtz’s equation is separable; and that the modal functions are the separated solutions in that system. Convergence of the PMM solution is related to the validity of the Rayleigh hypothesis [2, 3, 18].

In Yasuura’s method we usually set \(J = 2M\) as we see in Sect. 8.3.1. That is, we employ 2M linear equations to determine M unknown coefficients. This may be understood as a small device or improvement of the PMM. However, this produces good results such as proof of convergence, wide range of application, and so on with little increase of computational complexity as a reasonable cost.

Appendix 3: Conical Diffraction by a Grating

In Sect. 8.2.5 we dealt with diffraction by a grating, where all the field components were functions of two variables (X and Y) and two independent cases of polarization [E-wave (TE, s) and H-wave (TM, p)] existed. In addition, the directions of propagating diffraction-orders were parallel to the plane of incidence. These were possible because: (1) the grating surface was uniform in Z; and (2) the plane of incidence was in parallel to the direction of periodicity \(\mathbf{u}_X\). Here, we concisely examine the problem of a lossless dielectric grating in which the second condition is not satisfied, i.e., the plane of incidence makes a nonzero angle \(\phi \) with the positive X-direction as shown in Fig. 8.14a. We will see that

  • The field components are functions of X, Y, and Z, but the dependence on Y—the direction of uniformity—is limited;

  • The two cases of polarization are not independent, i.e., both TE and TM diffracted waves exist for TE (or TM) incidenceFootnote 34;

  • The direction of propagating orders lie on the surface of a cone whose vertex agrees with the coordinate origin O; the direction of the zeroth mode is on the plane of incidence at the same time.

Because of the third characteristic, this arrangement (\(\phi \not = 0\)) is called conical mounting and the term conical diffraction is used. In this connection, the arrangement in Sect. 8.2.5 is termed planar mounting.

Let the incident wave be

$$\begin{aligned} \left[ \begin{array}{c} \mathbf{E}^\mathrm{i} \\ \mathbf{H}^\mathrm{i} \end{array} \right] (\mathbf{r}) = \left[ \begin{array}{c}{} \mathbf{e}^\mathrm{i}\\ \mathbf{h}^\mathrm{i} \end{array}\right] \exp (-\mathrm{i}\mathbf{k}^\mathrm{i} \cdot \mathbf{r}). \end{aligned}$$
(8.110)

Here, \(\mathbf{e}^\mathrm{i}\) and \(\mathbf{h}^\mathrm{i}\) are electric- and magnetic-field amplitude, which are related by

$$\begin{aligned} \mathbf{h}^\mathrm{i} = \frac{1}{\omega \mu _0}\, \mathbf{k}^\mathrm{i} \times \mathbf{e}^\mathrm{i} \end{aligned}$$
(8.111)

and \(\mathbf{k}^\mathrm{i}\) is the incident wavevector defined by

$$\begin{aligned} \mathbf{k}^\mathrm{i} = (n_1 k\sin \theta \cos \phi , n_1 k\sin \theta \sin \phi , -n_1 k\cos \theta ) \equiv (\alpha , \beta , -\gamma ) \end{aligned}$$
(8.112)

with \(\theta \) being the polar angle between the wavevector \(\mathbf{k}^\mathrm{i}\) and the grating normal \(\mathbf{u}_Z\). We decompose the incident wave into a TE(s)- and a TM(p)-component, where TE (or TM) means that the electric (or magnetic) field of the relevant incident wave is perpendicular to the plane of incidence. To do this, we define two unit vectors that span a plane orthogonal to the incident wavevector

$$\begin{aligned} \mathbf{e}^\mathrm{TE}=(\sin \phi ,-\cos \phi ,0), \quad \mathbf{e}^\mathrm{TM}=(\cos \theta \cos \phi ,\cos \theta \sin \phi ,\sin \theta ). \end{aligned}$$
(8.113)

They give the directions of the incident electric fields that are in the TE- and TM-polarization.Footnote 35 Thus the decomposition is

$$\begin{aligned} \mathbf{e}^\mathrm{i} = \mathbf{e}^\mathrm{TE}\cos \delta + \mathbf{e}^\mathrm{TM}\sin \delta , \end{aligned}$$
(8.114)

where \(\delta \) is a polarization angle shown in Fig. 8.14b. \(\delta = 0\) and \(\pi /2\) mean TE- and TM-incidence. Hence, an incident wave has three angular parameters: \(\phi \), \(\theta \), and \(\delta \).

We consider the problem to seek the diffracted electric and magnetic field in the semi-infinite regions \(\text{ V }_1\) and \(\text{ V }_2\) over and below the grating surface \(\text{ S }_\mathrm{G}\).

Problem 4 conical, dielectric grating. Find the solutions that satisfy the following requirements:  

(CD1):

The Helmholtz equation in \(\text{ V }_1\) and \(\text{ V }_2\);

(CD2):

Radiation conditions in the positive and negative Z-direction;

(CD3):

A periodicity condition that: the relation \(f(X+d,Y,Z) = e^{\mathrm{i}\alpha d}f(X,Y,Z)\) holds for any component of the diffracted wave, and the phase constant in Y is \(\beta \);

(CD4):

The total tangential component of electric and magnetic field must be continuous across the grating surface \(\text{ S }_\mathrm{G}\).

  Dealing with a problem of conical diffraction, we should keep in mind the unique nature of the problem. First, because every field component has a common phase constant \(\beta \) in Y, it is sufficient to match the boundary condition on a cross section between the grating surface and a plane \(Y = \text{ const }\). The conically-mounted gratings, hence, belong to the class of quasi-3-D structures. Second, because the TE- and TM-wave are not independent, we always need both TE and TM vector modal functions in constructing approximate solutions.

We define the modal functions satisfying (CD1)–(CD3) by

$$\begin{aligned} \left\{ \begin{array}{l} \varvec{\varphi }_{\ell m}^\mathrm{TE}(\mathbf{r}) = \mathbf{e}_{\ell m}^\mathrm{TE} \exp (-\mathrm{i}\mathbf{k}_{\ell m} \cdot \mathbf{r}), \quad \varvec{\varphi }_{\ell m}^\mathrm{TM}(\mathbf{r}) = \mathbf{e}_{\ell m}^\mathrm{TM} \exp (-\mathrm{i}\mathbf{k}_{\ell m} \cdot \mathbf{r}) \\ \quad \quad \quad (\ell = 1, 2;\; m = 0, \pm 1, \pm 2, \ldots ). \end{array} \right. \end{aligned}$$
(8.115)

Here,

$$\begin{aligned} \mathbf{e}_{\ell m}^\mathrm{TE} = \frac{ \mathbf{k}_{\ell m} \times \mathbf{u}_Z}{|\mathbf{k}_{\ell m} \times \mathbf{u}_Z|}, \quad \mathbf{e}_{\mathrm{p}m}^\mathrm{TM} = \frac{ \mathbf{e}_{\ell m}^\mathrm{TE} \times \mathbf{k}_{\ell m}}{|\mathbf{e}_{\ell m}^\mathrm{TE} \times \mathbf{k}_{\ell m}|} \quad (\ell = 1, 2), \end{aligned}$$
(8.116)
$$\begin{aligned} \mathbf{k}_{1m} = (\alpha _m, \beta , \gamma _{1m}),\quad \mathbf{k}_{2m} = (\alpha _m, \beta , -\gamma _{2m}), \end{aligned}$$
(8.117)

and

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \alpha _m = \alpha + \frac{2m\pi }{d}, \quad \gamma _{\ell m} = \sqrt{(n_\ell k)^2 - (\alpha _m^2 + \beta ^2)} \\ \quad \quad \quad (\text{ Re }\,\gamma _{\ell m} \ge 0,\; \text{ Im }\,\gamma _{\ell m} \le 0). \end{array} \right. \end{aligned}$$
(8.118)

Note that the functions in (8.115) are for constructing electric fields. For the magnetic fields we get

$$\begin{aligned} \varvec{\psi }_{\ell m}^\mathrm{q}(\mathbf{r}) = \frac{1}{\omega \mu _0}\, \mathbf{k}_{\ell m} \times \varvec{\varphi }_{\ell m}^\mathrm{q}(\mathbf{r}) \quad (\ell = 1, 2;\; \mathrm{q}= \mathrm{TE, TM}) \end{aligned}$$
(8.119)

through Maxwell’s equations. Finite linear combinations of the modal functions define approximate solutions:

$$\begin{aligned} \left[ \begin{array}{c} \mathbf{E}_{\ell N} \\ \mathbf{H}_{\ell N} \end{array} \right] (\mathbf{r}) =\sum _{m=-N}^N A_{\ell m}^\mathrm{TE} \left[ \begin{array}{c} \varvec{\varphi }_{\ell m}^\mathrm{TE}\\ \varvec{\psi }_{\ell m}^\mathrm{TE}\end{array}\right] (\mathbf{r}) + \sum _{m=-N}^N A_{\ell m}^\mathrm{TM} \left[ \begin{array}{c} \varvec{\varphi }_{\ell m}^\mathrm{TM}\\ \varvec{\psi }_{\ell m}^\mathrm{TM} \end{array}\right] (\mathbf{r}) \quad (\ell = 1, 2) \end{aligned}$$
(8.120)

Here, the number of modal functions M is neglected.

The unknown coefficients in (8.120) should be determined in order that the solutions satisfy the boundary condition (CD4) approximately in the mean-squares sense. For this purpose we first consider the cross section C between the grating surface \(\mathrm{S}_\mathrm{G}\) and a plane \(Y = 0\). This is the same thing as the periodic curve C in Sect. 8.2.5. In a similar way to one in Sect. 8.2.5, we define the primary period \(\mathrm{S}_1\), whose boundary \(\mathrm{C}_1\) (\(\subset \mathrm{C}\)), the function space \(\mathbf{H}\) consisting of all the square integrable functions on \(\mathrm{C}_1\), and the norm \(\Vert f \Vert \) of a function f(s). Then, we can state the least-squares problem that determines the unknown coefficients:

LSP 4: conical, dielectric grating. Find the coefficients \(A_{\ell m}^\mathrm{TM}\) and \(A_{\ell m}^\mathrm{TM}\) \((\ell = 1, 2;\; m=0, \pm 1, \ldots , \pm N)\) that minimize the mean-square error

$$\begin{aligned} E_N = \left\| \varvec{\nu }\times \left( \tilde{\mathbf{E}}_{1N} + \tilde{\mathbf{E}}^\mathrm{i} - \tilde{\mathbf{E}}_{2N} \right) \right\| ^2 + Z_0^2 \left\| \varvec{\nu }\times \left( \tilde{\mathbf{H}}_{1N} + \tilde{\mathbf{H}}^\mathrm{i} - \tilde{\mathbf{H}}_{2N} \right) \right\| ^2. \end{aligned}$$
(8.121)

Here, \(Z_0\) denotes the intrinsic impedance of vacuum and \(\tilde{\mathbf{E}}^\mathrm{i}\) etc. mean periodic functions with respect to x defined in the same way as one in (8.69)–(8.71). The method of discretization and the solution method are found in Sect. 8.2.4.

Appendix 4: Comparison of Modal Functions and Algorithm of the SP

Here we show some results of effectiveness comparison between three kinds of modal functions in solving a sample problemFootnote 36: E-wave scattering from a PC cylinder whose cross section is given byFootnote 37

$$\begin{aligned} \text{ C }: r_s = a(1+0.2 \cos 3\theta _s). \end{aligned}$$
(8.122)

Let us normalize every quantity having dimension of length by the total length of C. And, we assume the incident wave comes along the x-axis from the negative x-direction (i.e., \(\iota =0\)).

The modal functions considered here are: (a) the separated solutions, which we defined by (8.7) in Sect. 8.2.2; (b) monopole fields defined by (8.9); (c) monopole fields whose poles are located densely near the convex part of C. Because the separated solutions are known widely, we explain the monopole fields below:  

(b) Equally spaced poles.:

Let L be a similar curve to C with the ratio of similitude \(d\ (0< d < 1)\).Footnote 38 We arrange M poles on L at regular intervals. Then, the distance between two poles is L / M where L is the length of L.

(c) Concentration of poles near the convex parts of L.:

(i) First, we draw the similar curve L. (ii) Next,we calculate the curvature \(\kappa (t)\) of L as a function of \(t\;(\in \text{ L })\), and add some positive bias c in order that the biased curvature (BC) be no less than 0: \(\tilde{\kappa }(t) = \kappa (t) + c\ (\ge 0)\). (iii) Thirdly, we define a probability density function by normalizing the BC.Footnote 39

$$\begin{aligned} f(t) = \frac{\displaystyle \int \limits _0^t \tilde{\kappa }(t')\, dt'}{\displaystyle \int \limits _0^L \tilde{\kappa }(t')\, dt'}. \end{aligned}$$
(8.123)

Thus we get the number of poles between \(t_1\) and \(t_2\) by

$$\begin{aligned} n(t_1, t_2) = M\int \limits _{t_1}^{t_2} f(t)\, dt. \end{aligned}$$
(8.124)

 

We have solved the problem using the method explained in Sects. 8.2.2 and 8.2.3. We used three kinds of modal functions (a), (b), and (c); and tried at two frequencies: \(ka = 10\) and 30. The parameter d was set to be 0.87. To see the accuracy of a solution we calculated two kinds of errors: the normalized mean-square error \(E_M(\text{ m })\) and the error on energy balance (or on the optical theorem) \(e_M(\text{ m })\). The former is the same thing as one defined in (8.22) and (8.30)Footnote 40 except that the subscript shows the total number of modal functions. The latter shows the deviation from a proportional relation between the forward scattering amplitude and total cross section.Footnote 41 The argument m shows the type of modal functions: m \(=\) sep, esp, and pcc, which mean (a) separated solutions, (b) equally-spaced poles, and (c) poles concentrated near the convex parts.  

Results at \(ka=10\).:

Because the obstacle size is handy, the \(E_M\) errors fall off rapidly: \(E_{45}(\mathrm{sep})\), \(E_{35}(\mathrm{esp})\), and \(E_{31}(\mathrm{pcc})\) are below 1%. As for the \(e_M\) errors of the solutions, the situation is different. The solutions with esp or pcc modal functions converge rapidly as \(e_M(\mathrm{esp})\) and \(e_M(\mathrm{pcc})\) are below 1% at \(M \simeq 30\). On the other hand, \(e_{31}(\mathrm{sep})\) is about 10%. Increasing M to 70, we have: \(e_{70}(\mathrm{esp})=9\times 10^{-5}\)%; \(e_{70}(\mathrm{pcc})=1\times 10^{-5}\)%; and \(e_{71}(\mathrm{sep})=4\)%.

Results at \(ka=30\).:

The advantage of the monopole fields is clear in this range of frequency. Setting \(M\simeq 100\), we have \(E_{101}(\mathrm{sep})=4\)%, \(E_{100}(\mathrm{esp})=2\times 10^{-1}\)%, \(E_{100}(\mathrm{pcc})=2\times 10^{-3}\)%, \(e_{101}(\mathrm{sep})=7\)%, \(e_{100}(\mathrm{esp})=2\times 10^{-1}\)%, and \(e_{100}(\mathrm{pcc})=5\times 10^{-3} \)%. The pcc modal functions seems to be the best choice in solving the problem. In fact, we can find an accurate solution with a \(10^{-5}\)% \(e_M\) error by setting \(\text{ m } = \text{ pcc }\) and \(M=120\).

 

These results mean that the potential of a combination of separated solutions is not so strong in describing scattered fields from obstacles deformed strongly from a circle. We have two ways to cope with this issue: (i) employment of a set of modal functions other than the separated solutionsFootnote 42; and (ii) employment of the SP.

The Algorithm of the SP

Here we include a guidance how to apply the SP in the boundary-matching process based on Yasuura’s method of modal expansion for convenience. We start from DLSP 1, i.e., minimization of the numerator of (8.30), \(\Vert \varPhi \mathbf{A}- \mathbf{f} \Vert ^2\). Instead of minimizing it directly, we force a constraint

$$\begin{aligned} (\mathbf{1}, \varPhi \mathbf{A}- \mathbf{f}) = 0 \end{aligned}$$
(8.125)

on the M-dimensional solution vector A, where the parentheses mean an inner product and \(\mathbf{1}=[1\ 1\ \cdots \ 1]^\mathrm{T}\) is a J-dimensional constant vector.

An operator of the smoothing procedure (in a discretized form) is a \(J\times J\) matrix given by

$$\begin{aligned} \mathrm{K}^{(p)} = \left[ K^{(p)}_{j\ell } \right] , \end{aligned}$$
(8.126)

where p means the order of the SP. The explicit forms of the matrix elements for \(p=1, 2\), and 3 areFootnote 43:

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle K^{(1)}_{j\ell } = u(j-\ell ) - \frac{j-\ell }{J} - \frac{1}{2}, \\ \displaystyle K^{(2)}_{j\ell } = -\frac{1}{2} \left[ \frac{(j-\ell )^2}{J^2} - \frac{|j-\ell |}{J} + \frac{1}{6} \right] , \\ \displaystyle K^{(3)}_{j\ell } = \frac{1}{6} \left[ \frac{|j-\ell |^3}{J^3} - \frac{3(j-\ell )^2}{2J^2} + \frac{|j-\ell |}{2J} + \frac{1}{6} \right] . \end{array} \right. \end{aligned}$$
(8.127)

Thus we can state a method of solution with the SP as follows:

DLSP 3: E-wave, PC, SP. Find the solution vector A that minimizes the discretized mean-square error

$$\begin{aligned} E_{MJ}= \frac{\Vert \mathrm{K}^{(p)}(\varPhi \mathbf{A}- \mathbf{f}) \Vert ^2}{\Vert \mathrm{K}^{(p)}{} \mathbf{f\,} \Vert ^2} \end{aligned}$$
(8.128)

under the constraint (8.125).

Two ways are possible to solve this conditioned least-squares problem: (i) employment of Lagrange’s multiplier; and (ii) elimination of a modal coefficient by using the constraint. Although (i) is a standard way in handling a constraint, we take (ii) because our constraint is simple and can eliminate one of the M unknowns to deduce a least-squares problem with \(M-1\) unknowns.

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG, part of Springer Nature

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Matsushima, A., Matsuda, T., Okuno, Y. (2018). Introduction to Yasuura’s Method of Modal Expansion with Application to Grating Problems. In: Wriedt, T., Eremin, Y. (eds) The Generalized Multipole Technique for Light Scattering. Springer Series on Atomic, Optical, and Plasma Physics, vol 99. Springer, Cham. https://doi.org/10.1007/978-3-319-74890-0_8

Download citation

Publish with us

Policies and ethics