PDF Computational Methods for Electromagnetic and Optical Systems Second Edition Banerjee download
PDF Computational Methods for Electromagnetic and Optical Systems Second Edition Banerjee download
com
https://ebookultra.com/download/computational-
methods-for-electromagnetic-and-optical-systems-
second-edition-banerjee/
https://ebookultra.com/download/computational-methods-for-electric-
power-systems-2nd-edition-mariesa-l-crow/
ebookultra.com
https://ebookultra.com/download/computational-methods-for-electric-
power-systems-3rd-edition-mariesa-l-crow-author/
ebookultra.com
https://ebookultra.com/download/advanced-modeling-in-computational-
electromagnetic-compatibility-1st-edition-dragan-poljak/
ebookultra.com
https://ebookultra.com/download/computational-modeling-methods-for-
neuroscientists-1st-edition-erik-de-schutter/
ebookultra.com
Computational methods for mass spectrometry proteomics 1st
Edition Ingvar Eidhammer
https://ebookultra.com/download/computational-methods-for-mass-
spectrometry-proteomics-1st-edition-ingvar-eidhammer/
ebookultra.com
https://ebookultra.com/download/computational-methods-for-two-phase-
flow-and-particle-transport-wen-ho-lee/
ebookultra.com
Computational Methods for Electromagnetic and Optical
Systems Second Edition Banerjee Digital Instant
Download
Author(s): Banerjee, Partha P.; Jarem, John M
ISBN(s): 9781439891285, 1439891281
Edition: 2nd ed
File Details: PDF, 55.22 MB
Year: 2014
Language: english
Computational Methods
for Electromagnetic
and Optical Systems
SECOnd EditiOn
OPTICAL SCIENCE AND ENGINEERING
Founding Editor
Brian J. Thompson
University of Rochester
Rochester, New York
RECENTLY PUBLISHED
Computational Methods for Electromagnetic and Optical Systems, Second Edition, John M. Jarem,
and Partha P. Banerjee
Optical Methods of Measurement: Wholefield Techniques, Second Edition, Rajpal S. Sirohi
Optoelectronics: Infrared-Visible-Ultraviolet Devices and Applications, Second Edition,
edited by Dave Birtalan and William Nunley
Photoacoustic Imaging and Spectroscopy, edited by Lihong V. Wang
Polarimetric Radar Imaging: From Basics to Applications, Jong-Sen Lee and Eric Pottier
Near-Earth Laser Communications, edited by Hamid Hemmati
Laser Safety: Tools and Training, edited by Ken Barat
Slow Light: Science and Applications, edited by Jacob B. Khurgin and Rodney S. Tucker
Dynamic Laser Speckle and Applications, edited by Hector J. Rabal and Roberto A. Braga Jr.
Biochemical Applications of Nonlinear Optical Spectroscopy, edited by Vladislav Yakovlev
Tunable Laser Applications, Second Edition, edited by F. J. Duarte
Optical and Photonic MEMS Devices: Design, Fabrication and Control, edited by Ai-Qun Liu
The Nature of Light: What Is a Photon?, edited by Chandrasekhar Roychoudhuri, A. F. Kracklauer,
and Katherine Creath
Introduction to Nonimaging Optics, Julio Chaves
Introduction to Organic Electronic and Optoelectronic Materials and Devices, edited by
Sam-Shajing Sun and Larry R. Dalton
Fiber Optic Sensors, Second Edition, edited by Shizhuo Yin, Paul B. Ruffin, and Francis T. S. Yu
Terahertz Spectroscopy: Principles and Applications, edited by Susan L. Dexheimer
Photonic Signal Processing: Techniques and Applications, Le Nguyen Binh
Smart CMOS Image Sensors and Applications, Jun Ohta
Organic Field-Effect Transistors, Zhenan Bao and Jason Locklin
Coarse Wavelength Division Multiplexing: Technologies and Applications, edited by Hans Joerg
Thiele and Marcus Nebeling
Microlithography: Science and Technology, Second Edition, edited by Kazuaki Suzuki and
Bruce W. Smith
Physical Properties and Data of Optical Materials, Moriaki Wakaki, Keiei Kudo, and Takehisa
Shibuya
Microwave Photonics, edited by Chi H. Lee
Photonics: Principles and Practices, Abdul Al-Azzawi
Lens Design, Fourth Edition, Milton Laikin
Gas Lasers, edited by Masamori Endo and Robert F. Walker
Optical Waveguides: From Theory to Applied Technologies, edited by Maria L. Calvo and
Vasudevan Lakshiminarayanan
Computational Methods
for Electromagnetic
and Optical Systems
SECOnd EditiOn
John M. Jarem
Partha P. Banerjee
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to
publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials
or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material repro-
duced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any
form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming,
and recording, or in any information storage or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://www.copy-
right.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400.
CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that have been
granted a photocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identifica-
tion and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
vii
viii Contents
Chapter 9 Bipolar Coordinate RCWA Computational Examples and Case Studies................. 341
9.1 Introduction.................................................................................................... 341
9.2 Case Study I: Fourier Series Expansion, Eigenvalue and Eigenfunction
Analysis, and Transfer Matrix Analysis......................................................... 343
9.3 Case Study II: Comparison of KPE BA, BC Validation Methods,
and SV Methods for Relatively Small Diameter Scattering Objects ............ 353
9.4 Case Study III: Comparison of BA, BC, and SV Methods for Gradually,
Stepped-Up, Index Profile Scattering Objects................................................ 356
9.5 Case Study IV: Comparison of BA, BC, and SV Methods for
Mismatched, Index Profile, Scattering Objects.............................................. 368
9.6 Case Study V: Comparison of BA, BC, and SV Methods for Gradually,
Stepped-Up, Index Scattering Objects with High Index Core....................... 377
Contents xi
xiii
xiv Preface
Partha P. Banerjee is a professor of electro-optics and electrical and computer engineering (ECE)
at the University of Dayton, where he was chair of the ECE department from 2000 to 2005. Prior
to that, he was with the ECE department at the University of Alabama in Huntsville from 1991 to
2000 and at Syracuse University from 1984 to 2001. He received his BTech from the Indian Institute
of Technology in 1979, and his MS and PhD from the University of Iowa in 1980 and 1983, respec-
tively. His research interests include optical processing, nonlinear optics, photorefractive materials,
and acousto-optics. He has authored/coauthored four books, several book chapters, over 100 refer-
eed journal articles, and many conference papers. He is fellow of the Optical Society of America
(OSA) and of the Society of Photo-optical Instrumentation Engineers (SPIE), and is a senior mem-
ber of the Institute of Electrical and Electronics Engineers (IEEE). He received the National Science
Foundation (NSF) Presidential Young Investigator award in 1987.
xv
1 Mathematical Preliminaries
1.1 INTRODUCTION
Popular t-shirts advertising Maxwell’s equations do not go beyond merely stating them. In this
book, we enter into a little more depth and solve these equations for analyzing various electromag-
netic (EM) and optical problems, e.g., diffraction gratings, radiation and scattering from dielectric
objects, photorefractive materials, metamaterials, and chiral materials. The emphasis on finding the
solutions in our text concerns the use of Fourier and state variable analyses, and beam propagation
methods. In this chapter, we briefly review some mathematical techniques pertinent to the analyses
presented in later chapters.
x0 + 2 π /K
⎛ 2π ⎞
I=
∫x0
exp( jnKx ) exp( − jmKx )dx ≡ ⎜ ⎟ δ m,n
⎝ K⎠
(1.1)
⎧1, m = n
δ m,n = ⎨ (1.2)
⎩0, m ≠ n
Using this, a function f (x) can be expanded in a Fourier series over an interval (x0, x0 + 2π/K) as
f ( x) = ∑ F exp( jnKx)
n = −∞
n (1.3)
Multiplying Equation 1.3 by exp(−jmKx), integrating over the interval (x0, x0 + 2π/K), interchanging
the summation and the integral, and using Equations 1.1 and 1.2, we get
x0 + 2π /K ∞ x0 + 2 π /K
∫
x0
f ( x ) exp( − jmKx )dx = ∑F ∫
n = −∞
n
x0
exp( jnKx ) exp( − jmKx )dx
∞
⎡ 2π ⎤ ⎛ 2π ⎞
= ∑ F ⎢⎣⎛⎜⎝ K ⎞⎟⎠ δ
n = −∞
n m,n ⎥ = ⎜⎝ K ⎟⎠ Fm
⎦
1
2 Computational Methods for Electromagnetic and Optical Systems
Now, replacing m by n,
x0 + 2 π /K
⎛ K⎞
Fn = ⎜ ⎟
⎝ 2π ⎠ ∫
x0
f ( x ) exp( − jnKx )dx (1.4)
Note that if a function fe(x) is defined as fe(x) = f(x) exp(−jαx), where α is a constant, then over the
interval (x0, x0 + 2π/K), it can be written as
∞
fe ( x ) = ∑ F exp(− jk
n = −∞
n xn x ), k xn = α − nK (1.5)
∞ ∞
f ( x) = ∑
n = −∞
Fn exp( jnKx ), g( x ) = ∑ G exp( jnKx)
n = −∞
n (1.6)
over the same interval are multiplied, the product function h(x) has a Fourier series expansion
∞
h( x ) = ∑ H exp( jnKx)
n = −∞
n (1.7)
over the same interval. We can find the Fourier coefficients of h(x) in the following way:
∞ ∞
h( x ) = f ( x )g( x ) = ∑
n = −∞
Fn exp( jnKx ) ∑G
m = −∞
m exp( jmKx)
∞ ∞ ∞ ∞
= ∑∑
n = −∞ m = −∞
FnGm exp( j (n + m)Kx ) = ∑∑F
l = −∞ m = −∞
l −m Gm exp( jlKx )
≡ ∑ H exp( jlKx)
l = −∞
l (1.8)
The limits on l are −∞ to +∞ since l = m + n, and m and n each have limits −∞ to +∞. Hence the
Fourier coefficients Hl of h(x) can be expressed as
Hl = ∑F
m = −∞
l −m Gm (1.9)
Equation 1.7 is sometimes referred to as the Laurent rule [1]. To be more precise, Equations 1.8 and
1.9 should be understood in the following sense [1]:
N L M
⎛ ⎞
h( x ) = lim
N →∞ ∑
l=−N
Hl exp( jlKx ) = lim
L →∞ ∑ ⎜
l=−L ⎝
lim
M →∞
m= − M
∑
Fl − mGm ⎟ exp( jlKx )
⎠
(1.10)
Mathematical Preliminaries 3
The above equation, in the way it is written, emphasizes two important points. First, the two limits
L and M are independent of each other, and the inner limit has to be taken first. Second, the upper
and lower bounds in each sum should tend to infinity simultaneously [1].
fˆ (ω) =
∫ f (t ) exp(− jωt )dt
−∞
(1.11)
The one-dimensional spatial Fourier transform of a square-integrable function f(x) is given as [2]
∞
f (k x ) =
∫ f ( x) exp( jk x)dx
−∞
x (1.13)
∞
1
f ( x) =
2π ∫ f (k ) exp(− jk x)dx
−∞
x x (1.14)
The definitions for the forward and backward transforms are consistent with the engineering con-
vention for a traveling wave, as explained in [2]. If f (x) denotes a phasor EM field quantity, multipli-
cation by exp jωt gives a collection or spectrum of forward traveling plane waves.
The two-dimensional extensions of Equations 1.13 and 1.14 are
∞ ∞
f (k x , k y ) =
∫ ∫ f ( x, y) exp( jk x + jk y)dx dy
−∞ −∞
x y (1.15)
∞ ∞
1
f ( x, y ) =
(2 π )2 ∫ ∫ f (k , k ) exp(− jk x − jk y)dx dy
−∞ −∞
x y x y (1.16)
In many EM applications, the function f (x, y) represents the transverse profile of an EM wave at a
plane z. Hence in (1.15) and (1.16), f (x, y) and f̃ (k x, k y) have z as a parameter. For instance (1.16)
becomes
∞ ∞
1
f ( x, y, z ) =
(2 π )2 ∫ ∫ f (k , k ; z) exp(− jk x − jk y)dx dy
−∞ −∞
x y x y (1.17)
4 Computational Methods for Electromagnetic and Optical Systems
The usefulness of this transform lies in the fact that when substituted into Maxwell’s equations,
one can reduce the set of three-dimensional PDEs to a set of one-dimensional differential equa-
tions (ODEs) for the spectral amplitudes f̃ (k x, k y; z).
∞ 2π
g(kρ , φʹ ) =
∫ ∫ g(ρ) exp{ jk ρ cos(φ − φʹ)}ρ dρ dφ
0 0
ρ (1.19)
It turns out that the result of the integration w.r.t. ϕ is independent of ϕ′ since [3]
2π
1
J 0 (kρρ) =
2π ∫ exp{ jk ρ cos(φ − φʹ)}dφ
0
ρ (1.20)
∞
1
g(kρ ) =
2π ∫
g(ρ)J 0 (kρρ)ρ dρ
0
(1.21)
Equation 1.21 is called the Fourier–Bessel transform or the Hankel transform of the circularly sym-
metric function g(ρ) [4]. The inverse transform is given by
∞
1
g(ρ) =
2π ∫
g(kρ )J 0 (kρρ)kρ dkρ
0
(1.22)
f p ( nΛ ) = ∑ f (nΛ + rN Λ)
r = −∞
(1.23)
Mathematical Preliminaries 5
The discrete function f (nΛ) may be formed by the discrete values of a continuous function f (x) evalu-
ated at the points x = nΛ.
The discrete Fourier Transform (DFT) of fp(nΛ) is defined as
N −1
2π
f˜p (mK ) = ∑ f (nΛ) exp( jmnK Λ),
n=0
p K=
NΛ
(1.24)
N −1
1
f p ( nΛ ) =
N ∑ f˜ (mK ) exp(− jmnK Λ)
n=0
p (1.25)
For properties of the DFT, e.g., linearity, symmetry, and periodicity, as well as relationship to the
z-transform, the Fourier transform, and the Fourier series, the readers are referred to any standard
book on digital signal processing [5].
For the purposes of this book, the DFT is a way of numerically approximating the continuous
Fourier transform of a function. The DFT is of interest because it can be efficiently and rapidly eval-
uated by using standard Fast Fourier Transform (FFT) packages. The direct connection between
the continuous Fourier transform and the DFT is given in the following. For a function f (x) and its
continuous Fourier transform ƒ̃(k x ),
1 mK < π
f˜p (mK ) ≅ f˜(mK ), (1.26)
Λ Λ
In Equation 1.26, ƒ˜p(mK) is defined, as in Equation 1.24, to be the DFT of fp(nΛ). The equality holds
for the fictitious case when the function is both space and spatial frequency limited.
Ax = qx (1.27)
These column vectors are called the eigenvectors of the system. The values of q which satisfy (1.27)
are known as the eigenvalues, the characteristic values, or the latent roots of the matrix A. Equation
1.27 may be written as a linear set of equations as
where I is the identity matrix. The result of Equation 1.29 is an nth-order polynomial called the
characteristic equation or eigenvalue equation. The equation is given by
The roots of this equation are the eigenvalues of the matrix A . When the roots are all unequal to
one another, the roots or eigenvalues are called distinct. When the eigenvalue occurs m times, the
eigenvalue is a repeated value of order m. When the root has a real and nonzero imaginary part, the
roots occur in complex conjugate pairs. In factored form, Equation 1.30 may be written as
The coefficients of the eigenvalue equation may be found directly from the matrix A. For
instance, setting q to zero in (1.31), we find
The coefficient a1 may be found by expanding the factored characteristic equation and comparing
the polynomial coefficients of the resulting equation. For example, if n = 2,
and thus
a1 = −(q1 + q2 ) (1.35)
If the determinant is expanded, we also find that the determinant is the negative sum of diagonal
coefficients, that is,
k
Let Tk = Tr ( A ). Then a useful formula for the coefficient an of the characteristic equation is
a1 = −T1
1
a2 = − (a1T1 + T2 )
2
1
a3 = − (a2T1 + a1T2 + T3 ) (1.39)
3
1
an = − (an −1T1 + an − 2T2 + ⋅⋅⋅ + a1Tn −1 + Tn )
n
For the case when the roots of P(q) are distinct, a nontrivial vector x i can be found for each root
which satisfies
(qi I − A) xi = 0 (1.40)
The matrix formed of the columns of –x i is called the modal matrix M. The name modal matrix
comes from control theory where a dynamical system can be decomposed into dynamic modes of
operation. For EM diffraction grating problems and also for EM problems that use k-space tech-
niques, the EM field solutions associated with a state variable analysis may be decoupled into spatial
mode solutions. These modes are analogous to the dynamical modes of operation encountered in
control systems.
If the eigenvalues are distinct, which is mainly the case under consideration in this text, the
modal matrix is nonsingular and therefore its inverse exists. Letting M be the modal matrix, we
may write
MQ = AM (1.41)
where Q is a diagonal matrix holding the eigenvalues qi on the diagonal. It can be shown that the
inverse of M exists, hence, Equation 1.41 gives
−1
Q = M AM (1.42)
If Q is squared, we have
2 −1 −1 −1 2
Q = ( M A M )( M A M ) = ( M A M ) (1.43)
−1
and if we further pre- and post-multiply by M and M , respectively, we have
2 2 −1
A = MQ M (1.44)
p p −1
A = MQ M (1.45)
8 Computational Methods for Electromagnetic and Optical Systems
p
where Q is the diagonal matrix formed by raising each eigenvalue qi to the pth power. A matrix
polynomial N ( A) can be conveniently evaluated as
−1
N ( A) = MN (Q) M (1.46)
where linear combinations of powers of A as given by Equation 1.45 have been used. N (Q) is the
diagonal matrix formed by placing in each diagonal matrix entry the polynomial N(qi). Thus the
modal matrix provides a convenient way to quickly and accurately evaluate powers and polynomials
of the matrix A.
In this text, we will be greatly concerned with calculating the exponential function of the
matrix A. The exponential function of the matrix A, namely exp( A) , is defined as [7]
⎛ 1⎞ 1
exp( A) = I + A + ⎜ ⎟ ( A)2 + ⋅⋅⋅ + ( A)k + (1.47)
⎝ 2⎠ k!
which is the same infinite series expansion as that used to define the exponential function exp(a).
We now review two important aids that help in the solution and evaluation of exponential matrix
and in fact any function of the matrix A. These are called the Cayley–Hamilton theorem and the
Cayley–Hamilton technique. These aids will be presented only for the cases of matrices with dis-
tinct eigenvalues.
The first theorem to be reviewed is the Cayley–Hamilton theorem. If we have a polynomial
N(q) = qn + c1qn−1 + · · · + cn−1 q + cn, then using Equation 1.46 we have
⎡ N (q1 ) 0 0 ⎤
N ( A) = M ⎢⎢ 0 N (q2 ) ⎥ M −1 (1.48)
⎥
⎢⎣ 0 N (q3 )⎥⎦
where M is the modal matrix. If the polynomial N(q) is chosen to the characteristic equation, that
is, N(q) = P(q), then N(qi ) = P(qi ) = 0; i = 1, 2, …, n and thus
⎡0 0 0 ⎤
⎢0 0 0 ⎥
P( A) = M ⎢ ⎥ M −1 = 0 (1.49)
⎢0 0 0 ⎥
⎢ ⎥
⎣ ⋅⋅⋅⎦
N (q ) R (q )
= Q(q ) + (1.50)
P (q ) P (q )
N (q ) = Q(q ) P (q ) + R (q ) (1.51)
If q is an eigenvalue or root of P(q), then P(q) = 0 and N(q) = R(q). If we substitute A for q in Equation
1.51, we have
N ( A) = Q( A)P( A) + R( A) (1.52)
N ( A) = R( A) (1.53)
Thus a higher order polynomial matrix may be represented and evaluated using an n−1 polynomial
expression.
Consider next the case where the matrix function is a general analytic function over a region
of interest, for example, F ( A) = exp ( A) . In this case, F(q) may be expanded in an infinite power
series over the analytic region of interest. As in the case when F(q) was a polynomial, F(q) may be
written as
F (q ) = Q(q ) P (q ) + R (q ) (1.54)
Let q = q1, q2, …, qn be the distinct roots of P(qi) = 0, i = 1,…, n. We have, after evaluating
Equation 1.54,
F (q1 ) = R(q1 )
F (q2 ) = R(q2 )
(1.56)
F ( qn ) = R ( qn )
This defines a set of n × n linear equations from which the coefficients αi, i = 1,…, n may be
determined. At this point, we would like to show that the function Q(q) is analytic. To do this,
we write Q(q) as
F (q ) − R (q )
Q(q ) = (1.57)
P (q )
In this expression, we note that over the region of interest the numerator and denominator of Equation
1.57 have the same zeros. Since in Equation 1.57 all functions F(q), Q(q), P(q), and R(q) are analytic
over the range where F(q) is, we may replace q by the matrix A. We therefore have
F ( A) = Q( A)P( A) + R( A) (1.58)
10 Computational Methods for Electromagnetic and Optical Systems
F ( A) = R( A) (1.59)
Thus we have shown that the analytic matrix function F ( A) may be evaluated by using a polynomial
matrix expression of order n − 1 as given by R ( A) in Equation 1.55.
PROBLEMS
1.1 Derive the wave equation for the electric and magnetic fields starting from Maxwell’s equations
in a homogeneous, isotropic, source-free region (see Chapter 3). How does this change if the
material is anisotropic?
1.2 Find from first principles the Fourier series coefficients for a periodic square wave s(x) of unit
amplitude and 50% duty cycle. Now find the Fourier series coefficients of s2(x): (a) from first
principles and (b) using the Laurent rule. Plot s2(x) versus x by employing the Fourier series
coefficients you found using (b). Use 5, 10, and 100 Fourier coefficients. Describe the general
trend(s).
1.3 Find the two-dimensional Fourier transform of a rectangle (rect) function of unit height and width
a in each dimension.
1.4 Show that the two-dimensional Fourier transform of a Gaussian function of width w is another
Gaussian function. Functions like this are called self-Fourier transformable. Find its width
in the spatial frequency domain. Can you think of any other functions that are self-Fourier
transformable?
1.5 Find the Hankel transform of (a) a circular function defined as g(ρ) = circ(ρ/ρ0), which has
a value of 1 within a circle of radius ρ0 and is 0 otherwise; (b) a Gaussian function given by
g(ρ) = exp−(ρ/ρ0)2.
1.6 Find the DFT of a square wave function using a software of your choice. Comment on the
nature of the spectrum numerically computed as the width of the square wave changes.
⎛ 1 20 0⎞
1.7 Find sin –A, where –A is a matrix given by ⎜ −1 7 1 ⎟ , using the Cayley–Hamilton theorem [7].
⎜ ⎟
⎝3 0 −2⎠
REFERENCES
1. L. Li, Use of Fourier series in the analysis of discontinuous periodic structures, J. Opt. Soc. Am. A 15,
1808–1816, 1996.
2. P.P. Banerjee and T.-C. Poon, Principles of Applied Optics, Irwin, New York, 1991.
3. M. Abramowitz and I. Stegun, Handbook of Mathematical Functions, Dover, New York, 1965.
4. I. Sneddon, The Use of Integral Transforms, McGraw-Hill, New York, 1972.
5. A. Antoniou, Digital Filters: Analysis and Design, McGraw-Hill, New York, 1979.
6. P.M. Deruso, R.J. Roy, and C.M. Close, State Variables for Engineers, John Wiley & Sons, Inc.,
New York, 1967.
7. L.A. Pipes and L.R. Harvill, Advanced Mathematics for Engineers and Scientists, McGraw-Hill,
New York, 1970.
2 Scalar EM Beam Propagation
in Inhomogeneous Media
2.1 INTRODUCTION
In the previous chapter, we reviewed some of the mathematical preliminaries that will be useful
later on in the text. In this chapter, we discuss some of the basic concepts of scalar wave propaga-
tion, and discuss an important numerical method, called the beam propagation method (BPM), to
study propagation in linear media and in media with induced nonlinearities. Furthermore, we also
discuss propagation through induced gratings, both transmission and reflection type, in order to
assess energy coupling between participating waves. Finally, we introduce readers to an important
characterization method, called the z-scan method, which is often used to determine the focal length
of an induced lens.
∂2 E
− v 2∇2 E = 0 (2.1)
∂t 2
and substitute
E ( x, y, z, t ) = Re{Ee ( x, y, z ) exp[ j (ω 0t − k0 z)]} (2.2)
and we will use one or the other notation according to convenience. Substituting Equation 2.2 into
Equation 2.1 and assuming that Ee is a slowly varying function of z (the direction of propagation) in
the sense that ∂ 2 Ee / ∂z 2 ∂Ee / ∂z << k0, we obtain the paraxial wave equation [1]
∂ Ee
2 jk0 = ∇2⊥ Ee (2.4)
∂z
where ∇2⊥ denotes the transverse Laplacian. Equation 2.4 describes the propagation of the envelope
Ee(x, y, z) starting from the initial profile Ee│z= 0 = Ee0(x, y).
Equation 2.4 can be solved readily using Fourier transform techniques. Assuming Ee to be
Fourier transformable, we can employ the definition of the Fourier transform as in Equation 1.15
11
12 Computational Methods for Electromagnetic and Optical Systems
and its properties to transform Equation 2.4 into the ordinary differential equation (ODE)
dEe j
= (k x2 + k y2 )Ee (2.6)
dz 2k0
⎧ j (k 2 + k y2 )z ⎫
Ee (k x , k y ; z) = Ee 0 (k x , k y ) exp ⎨ x ⎬ (2.7)
⎩ 2k0 ⎭
where Ẽe0(k x, k y) is the Fourier transform of Ee0(x, y). We can interpret Equation 2.7 in the following
way: Consider a linear system with an input spectrum of Ẽe0(k x, k y) at z = 0 where the output spec-
trum at z is given by Ẽe(k x, k y; z). The spatial frequency response of the system, which we will call
the paraxial transfer function for propagation is then given by
Ee ⎧ j (k x2 + k y2 )z ⎫
≡ H ( k , k ; z ) = exp ⎨ ⎬ (2.8)
Ee 0
x y
⎩ 2k0 ⎭
As we will show later, in the split-step BPM, we model propagational diffraction by means of the
transfer function for propagation derived above. For more exact calculations, the nonparaxial trans-
fer function can be used. This may be derived starting from the nonparaxial wave equation, but will
not be presented here for the sake of simplicity.
Incidentally, the inverse Fourier transform of the transfer function for propagation yields the
impulse response for propagation. Starting from the paraxial transfer function for propagation
which resembles a complex Gaussian, the inverse Fourier transform is a complex Gaussian as well,
and has the form
jk0 ⎡ − j ( x 2 + y 2 )k0 ⎤
h( x, y, z ) = exp ⎢ ⎥ (2.9)
2πz ⎣ 2z ⎦
This, when convolved with the initial beam profile, yields the profile of the diffracted beam in the
spatial domain directly. This convolution integral is in fact the Fresnel diffraction formula.
∂Ee 1
= ∇2⊥ Ee − j Δnk0 Ee (2.10)
∂z 2 jk0
Scalar EM Beam Propagation in Inhomogeneous Media 13
The quantity Δn is the change in the refractive index over the ambient refractive index n 0 = c/v,
where c is the velocity of light in vacuum. Equation 2.10 is a modification of Equation 2.4 and
can be derived from the scalar wave equation when the propagation constant or equivalently the
velocity of the wave is a function of (x, y, z) explicitly as in gratings or fibers, or implicitly such as
through the intensity dependent refractive index. An alternative, though heuristic, way to justify
the presence of the additional term on the RHS of Equation 2.10 is to note that in the absence of
diffraction (the first term on the RHS), the solution of the equation is of the form Ee ∝ exp[−jΔnk0z],
which explicitly shows the additional phase change due to propagation in the perturbed refractive
index.
The paraxial propagation equation (2.10) is a partial differential equation (PDE) that does not
always lend itself to analytical solutions, except for some very special cases involving special spatial
variations of Δn, or when as in nonlinear optics, one looks for particular soliton solution of the result-
ing nonlinear PDE using exact integration or inverse scattering methods. Numerical approaches are
often sought for to analyze beam (and pulse) propagation in complex systems such as optical fibers,
volume diffraction gratings, Kerr and photorefractive (PR) media, etc. A large number of numerical
methods can be used for this purpose. The pseudospectral methods are often favored over finite dif-
ference methods due to their speed advantage. The split-step BPM is an example of a pseudospectral
method.
To understand the philosophy behind the BPM, it is useful to rewrite Equation 2.10 in the
form [2,3]
∂Ee
= ( Dˆ + Sˆ )Ee (2.11)
∂z
where D̂, Ŝ are a linear differential operator and a space dependent or nonlinear operator respectively
(see, for instance, the structure of Equation 2.10). Thus, in general, the solution of Equation 2.11 can
be symbolically written as
⎛ ⎛ 1⎞ ⎞
exp( Dˆ Δz) exp(SˆΔz ) = exp ⎜ Dˆ Δz + SˆΔz + ⎜ ⎟ [ Dˆ , Sˆ ]( Δz)2 + ⎟ (2.13)
⎝ ⎝ 2 ⎠ ⎠
according to the Baker–Hausdorff formula [2], where [D̂, Ŝ] represents the commutation of D̂, Ŝ.
Thus, up to second order in Δz,
which implies that in Equation 2.13 the diffraction and the inhomogeneous operators can be treated
independent of each other.
The action of the first operator on the RHS of Equation 2.14 is better understood in the spectral
domain. Note that this is the propagation operator that takes into account the effect of diffraction
between planes z and z + Δz. Propagation is readily handled in the spectral or spatial frequency
domain using the transfer function for propagation written in Equation 2.8 with z replaced by Δz.
The second operator describes the effect of propagation in the absence of diffraction and in the
presence of medium inhomogeneities, either intrinsic or induced, and is incorporated in the spatial
14 Computational Methods for Electromagnetic and Optical Systems
Initial profile
Ee (x, y; z = 0)
~
Ee (kx , ky , z) = x,y {Ee (x, y, z)}
~ ~
Ee (kx , ky ; z + Δz) = Ee0 (kx , ky , z)exp{ j(k x2 + k 2y )Δz/2k0}
–1 ~
E΄e (x, y, z + Δz) = x,y {Ee (kx , ky ; z + Δz)}
ˆ
Ee (x, y, z + Δz) = E΄e (x, y, z + Δz)exp{SΔz}
domain. A schematic block diagram of the BPM method in its simplest form is shown in Figure
2.1. There are other modifications to the simple scheme, viz., the symmetrized split-step Fourier
method, and the leap-frog techniques; these are discussed in detail elsewhere [2].
⎛ π⎞ ⎡ (k 2 + k y2) ⎤
Ee (k x , k y ; 0) = E0 ⎜ ⎟ exp ⎢ − x ⎥ (2.16)
⎝ α⎠ ⎣ 4α ⎦
Scalar EM Beam Propagation in Inhomogeneous Media 15
Ee (k x , k y ; z) = Ee (k x , k y ; 0) H (k x , k y ; z)
⎛ π⎞ ⎡ (k 2 + k y2) ⎤ ⎡ j (k x2 + k y2 )z ⎤
= E0 ⎜ ⎟ exp ⎢ − x ⎥ exp ⎢ ⎥
⎝ α⎠ ⎣ 4α ⎦ ⎣ 2k0 ⎦
⎛ π⎞ ⎡ j (k 2 + k y2 )q ⎤
≡ E0 ⎜ ⎟ exp ⎢ x ⎥ (2.17)
⎝ α⎠ ⎣ 2k0 ⎦
jk0 ⎛ π⎞ ⎡ jk ( x 2 + y 2 ) ⎤
Ee ( x, y; z) = E0 ⎜ ⎟ exp ⎢ − 0 ⎥ (2.18)
2πq ⎝ α ⎠ ⎣ 2q ⎦
jk0
q = z+ (2.19)
2α
If the initial Gaussian beam has waist w0 and plane wavefronts, α = 1/w02 and q = z + jzR , where
zR = k0 w02 /2 is commonly referred to as the Rayleigh range of the Gaussian beam. Upon simplify-
ing Equation 2.18 using Equation 2.19, we get the standard expression for the diffracted Gaussian
beam profile as
w0 ⎡ ( x 2 + y2 ) ⎤ ⎡ jk0 ( x 2 + y 2 ) ⎤ − jφ( z )
Ee ( x, y; z) = E0 exp ⎢ − 2 ⎥ exp ⎢ − ⎥e (2.20)
w( z ) ⎣ w ( z) ⎦ ⎣ 2 R( z ) ⎦
where
⎡ ⎛ z ⎞2⎤ ( z 2 + z R2 ) ⎛ z⎞
2 2
w ( z) = w ⎢1 + ⎜ ⎟ ⎥ , R( z) =
0 , φ( z ) = − tan −1 ⎜ ⎟ (2.21)
⎢⎣ ⎝ z R ⎠ ⎥⎦ z ⎝ zR ⎠
Incidentally, if a Gaussian beam with initially plane wavefronts passes through a (thin) lens of focal
length f, the latter introduces a quadratic phase curvature which can be accommodated by multiply-
ing the initial Gaussian with a complex exponential exp[ jk0(x2 + y2)/2f ]. In this case, the scalar optical
field immediately behind the lens (z = 0+) can be written in the form as in Equation 2.15 above, but
where α is redefined as
1 jk ⎛ 1 ⎞ ⎛ 1 − jz R ⎞
α= − 0 = ⎜ 2⎟⎜ (2.22)
w0 2 f ⎝ w0 ⎠ ⎝ f ⎟⎠
2
16 Computational Methods for Electromagnetic and Optical Systems
1
0.8 1
0.8
0.6
0.6
0.4
0.4
0.2
0.2
0
80 0
60 80 80
60 60 80
40 40 60
20 40 40
20 20 20
(a) 0 0 (b) 0 0
FIGURE 2.2 Diffraction of a Gaussian beam during free-space propagation: (a) profile at z = 0 (plane
wavefronts assumed) and (b) profile at z = z R, where z R is the Rayleigh range of the original Gaussian beam.
The paraxial propagation of a Gaussian beam through an optical system involving translation and
lensing can be equivalently studied in terms of the variation of the q-parameter. For instance, due
to propagation in air through a distance z alone, the change in the q-parameter can be readily found
from Equation 2.17 as
On the other hand, if a Gaussian beam is incident on a lens of focal length f, the change in the q-parameter
can be readily found from Equation 2.18 and the phase transformation by the lens given above as
1 1 1
= − (2.23b)
qnew qold f
Equation 2.23a and b will be later used to analyze propagation of Gaussian beams through media
which have a distributed lensing effect, such as a nonlinear medium.
We now demonstrate the application of BPM to Gaussian beam propagation through free space.
In this case, the inhomogeneous operator is zero, and propagation from a plane z = 0 to arbitrary
z can indeed be performed in one step in this case. However, in the example we provide, we use the
split-step method to convince readers that the result is identical to what one would get if the propaga-
tion was covered in one step. In Figure 2.2, we show the profile of a diffracted Gaussian beam after
propagation through free space, and the results agree with the physical intuition of increased width
and decreased on-axis amplitude during propagation as well as the analytical results in Equations
2.20 and 2.21.
n = n0 + n( 2 ) ( x 2 + y 2 ) (2.24)
where n0 denotes the intrinsic refractive index of the medium, and n(2) is a measure of the gradation
in the refractive index.
In this case, the operator Ŝ becomes
Sˆ = − jk0 n( 2 ) ( x 2 + y 2 ) (2.25)
Scalar EM Beam Propagation in Inhomogeneous Media 17
60 60
50 50
40 40
30 30
20 20
10 10
10 20 30 40 50 60 10 20 30 40 50 60
60
50
40
30
20
10
10 20 30 40 50 60
FIGURE 2.3 Contour plots showing periodic focusing of an initial Gaussian profile.
Propagation of a Gaussian beam in a medium with a graded index profile is shown in Figure 2.3.
The contour plots show the initial (Gaussian) beam profile, the beam profile where the initial
Gaussian attains its minimum waist during propagation before returning back to its original shape
again, due to periodic focusing by the graded index distribution. Note that there exists a specific
eigenmode (a Gaussian of a specific width, related to the refractive index gradient) for which the
beam propagates through the material without a change in shape as a result of a balance between
the diffraction of the beam and the guiding due to the parabolic gradient index profile. The con-
tour plot of such a beam is shown in Figure 2.4. Analytical expressions for the Gaussian beam
60
50
40
30
20
10
10 20 30 40 50 60
during propagation through a graded index medium and for the eigenmode can be derived using
the q-parameter of the Gaussian beam and the ABCD laws of q-transformation, but will not be
pursued here for the sake of brevity.
where C is an interaction constant (for details, see Korpel [4]) and s(x, z, t) is the real sound ampli-
tude given by,
1
s( x, z, t ) = [ Se ( x, z ) exp( − jKx ) exp( jΩt ) + c.c.] (2.27)
2
where Se is the complex amplitude of the sound field that interacts with the light beam and is travel-
ing in the x direction, and c.c. denotes the complex conjugate.
The quantities K and Ω are the propagation constant and the angular frequency, respectively, of
the sound field. Following Refs. [4,5], a snapshot of the sound field is used at t = 0, so that using
2.26 and 2.27,
⎛ 1⎞
= 1 − ⎜ ⎟ jk0 ΔzC [ Se ( x, z ) exp( − jKx ) + Se * ( x, z ) exp(+ jKx )] (2.28)
⎝ 2⎠
In the modified split-step technique, we take the Fourier transform of the above operator operating
on the optical field Ee(x, z), taking care to note from the property of Fourier transforms that
The main propagation loop of the algorithm is modified from Figure 2.1 and is shown in Figure
2.5. The boxes marked “Shift ±K” are used to facilitate the operation shown in 229 in the spatial
frequency domain.
Figure 2.6 shows problem geometry of a Gaussian beam incident nominally at Bragg angle on
a sound column of width z = L. The simulated evolution of the Gaussian beam is shown in Figure
2.7. The peak phase delay α of the light traveling through the acoustooptic cell is taken equal
to π, and the Klein–Cook parameter Q = K 2 L /k 0 = 13.1. We would like to point out that the same
answers could be derived by using the transfer function for acoustooptic interaction, as given in
Refs. [4,6].
Scalar EM Beam Propagation in Inhomogeneous Media 19
Initial profile
Ee (x, z = 0)
~
Ee (kx , z) = x {Ee (x, z)}
~ ~
Ee (kx , z + Δz) = Ee0 (kx , z) exp{ jk x2 Δz/2k0}
–1 ~
E΄e (x, z + Δz) = x {Ee (kx ; z + Δz)}
x x
Shift-K Shift-K
–1
x
FIGURE 2.5 Flow diagram for the modified split-step technique to analyze acoustooptic interaction.
Acoustic grating
Incident 0 order
Gaussian
beam
–1 order
Z=0 Z=L
FIGURE 2.6 Geometry of acoustooptic interaction with a Gaussian beam at nominal Bragg incidence.
∂ Ee 2
(2.30)
2 jk0 = ∇2⊥ Ee + 2n2 k02 Ee Ee
∂z
where n2 is the nonlinear refractive index coefficient defined by the functional dependence of the
total refractive index n on the intensity [7]:
20 Computational Methods for Electromagnetic and Optical Systems
Zeroth order
First order
z=0
In φ
te
rac
tio
n φB
len
gt
h –φB Z
z=L
FIGURE 2.7 Simulation plot of the intensity of the angular spectrum of the total field at different positions
along interaction length. (From Venzke, C. et al., Appl. Opt., 31, 656, 1992. With permission.)
n = n0 + δn Ee ( )= n +n
2
0 2 Ee
2
(2.31)
In writing (2.30), we have taken the linear refractive index n0 equal to unity for the sake of simplic-
ity. For a medium with n2 > 0, one can observe self-focusing of a Gaussian beam traveling through a
medium, while self-defocusing is observed for a medium with n2 < 0. The physical reasoning behind
self-focusing is as follows. The Gaussian beam induces a positive lens in the nonlinear material for
n2 > 0 due to the fact that where the beam intensity is high (e.g., on-axis), the induced refractive index
is higher as well, amounting to larger slowing down of the wavefronts. The wavefronts are therefore
bent similar to the action of a positive lens, resulting in initial focusing of the beam. This process
continues till the beam width is small enough for the diffraction effects to take over, leading to an
increase in the beam width. The converse is true for the case of n2 < 0. In this case, the beam spreads
more than in the linear diffraction limited case. A stable nonspreading solution in one transverse
dimension can be analytically found from the NLS equation for n2 > 0 and has the form
1/ 2
⎛ 8κ ⎞ x
Ee ( x; z) = ⎜ sech (2.32)
⎝ n2 k0 ⎟⎠ (2κk0 )1/ 2
where κ is a free parameter. The phase of Ee in the nonspreading solution is linearly proportional to
the propagation distance z.
As discussed above, self-focusing results in increase in the on-axis intensity and the narrowing
of the beam width. For powers above a certain critical power Pc [7,8], the beam may theoretically
collapse with an intensity so high that it can either cause breakdown in the material, triggering
some other physical effects, such as saturation of the index of refraction or failure of the assump-
tions about slowly varying amplitude and paraxial approximation. Zakharov and Shabat [9] have
pointed out that if the nonlinearity is strong enough, this results in higher index of refraction toward
the center of the beam and on-axis rays undergo total internal reflection and are thereby trapped.
As shown above, nonlinearity can balance diffraction of a beam in one dimension, resulting in the
formation of first order spatial solitons, see Equation 2.32. Also, if the nonlinear effect is higher
Scalar EM Beam Propagation in Inhomogeneous Media 21
than diffraction, periodic focusing occurs, or may result in higher order solitons. This may not be
the case in two or three dimensions where spatial collapse may occur.
In normalized form, and assuming n2 > 0, the NLS equation in Equation 2.30 can be rewritten as
∂ue
j + ∇2⊥ue + ue2 ue = 0 (2.33)
∂z
The NLS equation as written in Equation 2.33 can also be modified to model pulse propagation
through a nonlinear fiber and in the presence of group velocity dispersion. This is possible due to the
fact that the interchange x → t in the NLS equation (2.33) with a suitable coefficient in front of the
second order derivative term, signifying (anomalous) material dispersion and subsequent renormaliza-
tion transforms the equation to one that can model the propagation of pulses in time τ along a fiber [2]:
∂ue 1 ∂ 2ue 2
j + 2
+ γ ʹ ue ue = 0 (2.35)
∂z 2 ∂τ
Analogous to Equation 2.32, the first order soliton solution of (2.35) can be expressed as
(
ue (τ; z) = K sech K γ ʹ τ ) (2.36)
where K is a free parameter. This profile is called a temporal soliton and can be regarded as a non-
(
linear eigenmode of the NLS system. The propagation of an initial profile ue (τ; z = 0) = sech γ ʹ τ )
with γ ′ = 1 using the BPM to model the NLS equation as in Equation 2.35 is shown in Figure 2.8.
The linear term in Equation 2.35 can be handled in the temporal frequency domain by using Fourier
transforms, similar to the case of propagational diffraction. Note that the nonlinear operator is
1
~ |2
0.5
|u e
0
3
2 10
1 0
z
0 –10 τ
FIGURE 2.8 Evolution of a first order 1-D soliton. (From Nehmetallah, G. and Banerjee, P.P., Nonlinear
Optics and Applications, H.A. Abdeldayem and D.O. Frazier, eds., Research Signpost, Trivandrum, India,
2007. With permission.)
22 Computational Methods for Electromagnetic and Optical Systems
|ue| 0
–2
2
10
1
0
z τ
0 –10
FIGURE 2.9 Evolution of a second order 1-D soliton. (From Nehmetallah, G. and Banerjee, P.P., Nonlinear
Optics and Applications, H.A. Abdeldayem and D.O. Frazier, eds., Research Signpost, Trivandrum, India,
2007. With permission.)
exp{Sˆ Δz} = exp − j γ ʹ ue Δz . As expected, the “pulse” remains unchanged with propagation.
2
A similar result is obtained if we program the propagation of an initial beam profile and use
Equation 2.30 in one transverse dimension (x). The result is a spatial soliton.
The second order soliton input ũe(τ, 0) = 2 sech(τ) and its evolution in time is shown in Figure 2.9.
The split-step technique has also been applied to analyze propagation of profiles in two transverse
dimensions [11], and also to analyze propagation of optical fields that are pulsed in time and have a
spatial profile in the transverse dimension [12].
The NLS equation for beams using the paraxial approximation has been previously derived in
Section 2.6.1 as
∂ue
j + ∇2⊥ue + ue2 ue = 0 (2.37)
∂z
During the last stages of self-focusing, the assumptions about slowly varying amplitude and the
paraxial approximation may not be valid for large focusing angles. It has been proposed that there is
no singularity if one accounts for nonparaxiality [13]. The paraxial and nonparaxial NLS equations
which are classically used to model the self-focusing phenomenon can be written in the general
operator form:
∂ue ∂ 2u
− j ε 2e = jLr ue + jN nl (ue ) ue , r ∈ℜ D , z ≥ 0; ue (r, z = 0) = ue0 (r ) (2.38)
∂z ∂z
where D is the transverse dimension in space. The parameter ε = (λ/4πr0)2, where r0 is the initial beam
2σ
radius is often referred to as the nonparaxiality parameter. Also, Lr ue = ∇⊥2 ue, N nl (ue ) = ue . For most
2
cases, σ = 1 and hence N nl (ue ) = ue . The nonlinear operator Nnl(ũe) may be also conveniently modi-
fied to reflect any saturation in the change of refractive index:
δn ue( ) = n u
2
2 e
2 4
+ n4 ue , n4 < 0 (2.39)
Scalar EM Beam Propagation in Inhomogeneous Media 23
5
Mode 1
4 Mode 2
Mode 3
3
2
~a
1
–1
–2
0 2 4 6 8 10
~r
FIGURE 2.10 Townes soliton profiles for the first three “modes.”
Incidentally, for the paraxial case, two transverse dimensions and assuming radial symmetry, and
σ = 1, a special solution of (2.38), called the Townes soliton [7], can be found as the solution of the
normalized differential equation
Typical solutions for three different initial conditions, called modes, are plotted in Figure 2.10. These
modes, named ã(r̃ ), are characterized by the fact that the solutions tend to zero for r̃ → ∞. The solu-
tions of Equation 2.40 are very sensitive to initial conditions: for other initial conditions, the solutions
do not converge to zero as r → ∞.
The basic concept behind adaptive numerical algorithms is rather simple: imagine that a Gaussian
beam is spreading during linear propagation due to diffraction. If a numerical solution is being com-
puted, it is clear that after some distance of propagation, the size of the beam will become comparable to
the total transverse grid size, which will result in errors in the numerical computation, such as aliasing if
fast Fourier transform (FFT) techniques are used. The problem can be alleviated if the transverse profile
is re-sampled using a coarser grid size. The step size for propagation can also be increased, based on
the presumption that no sudden changes in the beam would occur at distances larger than the Rayleigh
range. The converse should be true for beams that are focusing: the transverse grid size should be made
finer and the propagational step size smaller.
In the case of the NLS equation, for the case D = 2, and assuming radial symmetry, McLaughlin
et al. [14] predicted similarity solutions of the form
1 ⎛ r ⎞
ue (r , z ) ∝ F (2.41)
( zr − z )1/ 2 ⎜⎝ ( zr − z )1/ 2 ⎟⎠
where F is an arbitrary function. We can use this knowledge to adaptively vary the longitudinal
stepping Δz and the transverse grid size for the beam.
A few more technical details must be mentioned before commencing the discussion on our adaptive
numerical techniques. Weinstein [15] has shown that if the initial beam power P0 is less than the lower
2
bound for the critical power for blowup Pclb = (λ 24πn 0 n2 ) N c, where N c =
∫R r dr = 1.86225, there
24 Computational Methods for Electromagnetic and Optical Systems
is no collapse of the paraxial NLS equation. Also, for an arbitrary initial profile ũe0 ≠ R, Fibich [13]
proved that there is an upper bound for the critical power Pcub = (λ 24πn 0 n2 ) G(ue 0 ) where
⎛ 2 u 2 dr ∇u 2 dr ⎞
G (ue 0) =
⎝ ∫
e0 e0
∫
⎠
⎛ u 4 dr ⎞
⎝ e0 ∫ ⎠
for which blowup will occur for this initial profile if it has sufficient high power. There are few
known integral invariants for the paraxial NLS equation above when σD ≥ 2. These invariants
are based on the symmetry of the NLS equation under gauge, space, and time transformations,
and may be derived from a Lagrangian density of Equation 2.37. Three of these invariants are the
Hamiltonian and the variance, defined as [14,16]
2
N=
∫ u e dr (2.42a)
⎛ 2 1 4⎞
⎝ ∫
H = ⎜ ∇ue − ue ⎟ dr
2 ⎠
(2.42b)
1 d2 2 2 D−2 4
8 dz 2 ∫r ue dr = H −
4 ∫ u
e dr (2.42c)
where the term |∇ũe|2 results from diffraction, and |ũe|4 from the nonlinear effect.
Note that from Equation 2.42c blowup occurs only if D ≥ 2. The “variance” is given as
2 2
∫
V ( z) = r ue dr = 4 Hz 2 + (dV (0) /dz )z + V (0) for D = 2 [17]. Thus, for H < 0, the function V(z)
vanishes at a distance zr = [V0 /−4H0]1/2 > 0. A sufficient condition for blowup is H < 0, i.e., when
the nonlinear effect is stronger than diffraction, the beam self-focuses and collapse occurs at a
distance zc ≤ zr.
− r 2 / 2 r02 1/ 2
For an input Gaussian of the form ue 0 = ce , zr = (1 ( pN c 2 − 1)) , zc = 0.317(p − 1)−0.6346 for
r0 = 1, and zc = 0.1585(p − 1) −0.6346 for r0 = 1/ 2 , where p = N0/Nc and N 0 = c 2r0D [18]. Note that the
ub
condition H = 0, which leads to Pc , leads to an overestimate of the actual critical power.
We now outline our numerical adaptive spectral technique called the adaptive split-step fast
Hankel transform (AFHTSS) used to track the solution of the NLS equation for σ = 1, D = 2 and vari-
able ε. Our scheme is based on the combination of the standard split-step fast Fourier transform
(SSFFT) and the Hankel transform, which exploits the cylindrical symmetry of the problem. This
enhances the computation time and precision appreciably. In addition, we also use the concepts
from the similarity solution developed by McLaughlin et al. [14] and apply them to our split-step
spectral method mentioned above, so that the grid transverse spatial range and the longitudinal
spatial step are adaptively updated. As its name indicates, we use the Hankel transform (see Chapter 1)
instead of the usual Fourier transform, relying on algorithms already developed in the literature [19].
In cylindrical coordinates, Equation 2.38 becomes
∂ue ∂ 2u ⎛ ∂2 1 ∂ ⎞ 2
− j ε 2e = j ⎜ 2 + ⎟ ue + j ue ue ; ue (r , z = 0) = ue0 (r ) = c exp − ar 2 (2.43)
∂z ∂z ⎝ ∂r r ∂r ⎠
Scalar EM Beam Propagation in Inhomogeneous Media 25
We use the definition of the lth order Hankel Transform pair [19,20]:
∞ ∞
∫
Ψ(ρ, z ) = 2π ψ (r , z )J l (2πrρ)rdr ,
0
∫
ψ (r , z ) = 2π Ψ(ρ, z )J l (2πrρ)ρdρ
0
(2.44a)
⎛ ∂2 1 ∂ ⎞
j⎜ 2 +
⎝ ∂r r ∂r ⎟⎠
ψ (r , z ) → − j 4π ρ Ψ(ρ, z)
HT 2 2
(2.44b)
The AFHTSS algorithm in Figure 2.11 resembles the symmetrized Fourier split-step technique,
where we change the longitudinal spatial step Δz ∝ (1/c(z1)2 − 1/c(z2)2) ≈ 1/c(z1)2 when c(z2) ≫ c(z1),
adaptively using McLaughlin’s similarity formula (2.41) and the grid spatial range Δrmax ∝ 1/c(z) in
order to track the varying amplitude of the focusing beam when ε = 0.
There are several numerical approaches for implementing the Hankel transform [19,21]. The
importance of Siegman’s method [19] resides in the fact that, depending on the parameters, one can
employ a non-uniform sampling that is denser near the focusing region, which has advantages over
uniform sampling. Yu et al.’s method [21] is based on the expansion of the function and its transform
by a zero order Bessel series that can be written as
N N
ˆ (m) =
Ψ ∑n =1
Cmn ψˆ (n), ψˆ (n) = ∑C
m =1
nm
ˆ (m)
Ψ (2.45a)
2 ⎛ jn jm ⎞ −1
Cmn = J0 J1 ( jn ) J1−1 ( jm ) (2.45b)
S ⎜⎝ S ⎟⎠
The jn’s are the positive roots of the zero order Bessel function J 0, J1 is the first order Bessel
function, and R1, R2 are the spatial and transform ranges respectively, with S = 2R1R2. For rn ∈ r ≥ R1
and ρm ∈ ρ ≥ R2, we have ψ(rn ) = Ψ(ρm ) = 0, where rn = jn /2πR2 and ρm = jm / 2πR1.
We now show sample simulation results using the AFHTSS method that uses Yu et al.’s Hankel
transform technique [21], as well as a novel adaptive version of a split-step fast fourier transform
2
(AFFTSS) technique. For the test function ψ 0 = 4e − r / 2 , self-focusing and collapse is expected
at zc = 0.1487 for ε = 0. Note that zr = 0.288, based on the study above, which is obviously overes-
timated. Figure 2.12 shows the maximal focusing as a function of grid size h = Δr, which proves
the convergence of our method to the numerical focusing point when Δr decreases by varying the
S parameter defined above [22]. Although the trend is similar, this is a considerable improvement
over the convergence test results in Fibich and Ilan [23]. Figure 2.13 shows the growth of the on-
axis intensity using AFHTSS technique for the test function above [22]. Also, in this figure, we take
S = 2πR1R2 = 2π × 2000 (corresponding to approximately 4000 cylindrical samples), which permits
computation for the paraxial case till z = zc = 0.14817 corresponding to the expected zc = 0.1487 writ-
ten above. Figure 2.14 shows the corresponding AFFTSS technique for 1024 × 1024 grid where we
get z = zc = 0.1581 [22]. This implies that the critical distance zc is a little bit over estimated and the
maximum intensity reached is also less than the AFHTSS method. It is instructive to note that
with increase in sampling points, the AFFTSS approaches the results from AFHTSS, but on the
26 Computational Methods for Electromagnetic and Optical Systems
Initial profile
2 2
ψ0 = ce–r /2r0
K=K+1
j 1/2 Δz
1– 1 – 16επ2ρ2
Δz Ψ(ρ, z)e – 2ε 2
Ψ ρ, z + =
2
Δz Δz
ψ r, z + = IFHT Ψ ρ, z +
2 2
j 2 1/2 Δz
1– 1 + 4ε|ψ|
Δz – 2ε
ψ r, z + 3 Δz = ψ r, z + e 2
2 2
3Δz 3Δz
Ψ ρ, z + = FHT Ψ r, z +
2 2
j 2 2 1/2 Δz
1– 1 – 16επ ρ
3Δz – 2ε 2
Ψ(ρ, z + 2Δz) = Ψ ρ, z + e
2
1 1
Δz and Δrmax
c(z)2 c(z)
FIGURE 2.11 The AFHTSS algorithm, a symmetrized version of the split-step FFT using Hankel transform
instead, and using adaptive longitudinal stepping and transverse grid management. FHT: fast Hankel trans-
form, IFHT: inverse fast Hankel transform.
Scalar EM Beam Propagation in Inhomogeneous Media 27
150
ψ = 4 exp(–r 2)
ψ = 4e(–r 2/2)
100
ψ2max/ψ2max0
50
0
0 0.005 0.01 0.015 0.02 0.025
Grid size: Δr = h
FIGURE 2.12 Maximal focusing as a function of grid size. (From Banerjee, P.P. et al., Opt. Commun., 249,
293, 2005. With permission.)
(8) ε = 0 (5)
(7)
103
(4)
(6)
(3)
102 (2)
(1)
101
0 0.05 0.1 0.15 0.2 0.25
z
2
FIGURE 2.13 On-axis intensity of ψ 0 = 4e − r / 2 as a function of propagation for fixed values of ε rang-
ing from 10 −2 to 10 −8 where z = zc = 0.1481 for ε = 10 −8, and for an adaptive ε varying as ε = (λ/4πr)2, using
AFHTSS with S = 2πR1R2 = 2π × 2000 (4000 cylindrical samples). (From Banerjee, P.P. et al., Opt. Commun.,
249, 293, 2005. With permission.)
expense of calculation time. Also we note that changing the number of samples in both techniques
does not affect the value of the critical distance drastically, but the maximum intensity reached
at that point will be less or more depending on the number of samples, which is in agreement
with the convergence test mentioned above. We stress that in all computations using AFHTSS and
AFFTSS, energy is always conserved. Consequently, AFHTSS permits us to track peak intensi-
ties higher, faster, and more accurate than what is achievable by AFFTSS. Also, by using adaptive
28 Computational Methods for Electromagnetic and Optical Systems
(8) ε = 0 (5)
(7)
103
(4)
(6)
(3)
102 (2)
(1)
101
0 0.05 0.1 0.15 0.2 0.25
z
2
FIGURE 2.14 On-axis intensity of ψ 0 = 4e − r / 2 as a function of propagation for fixed values of ε rang-
ing from 10 −2 to 10 −8 where z = zc = 0.1581 for ε = 10 −8, and for an adaptive ε varying as ε = (λ/4πr)2, using
AFFTSS with 10242 samples. (From Banerjee, P.P. et al., Opt. Commun., 249, 293, 2005. With permission.)
non-paraxiality parameter ε in the scalar nonparaxial equation, we obtain results similar to those of
the more complex vector method [24] and superior to those when ε is constant [13].
Finally, we compare computation speeds of the AFHTSS and AFFTSS. It can be shown that for
Siegman’s method, the number of computations is proportional to 4N log2 2N + 2N, compared with
2N 2 log2 N computations for the two dimensional FFT as in AFFTSS. Although at first glance, the
number of computations in Yu et al.’s method is proportional to N2, we can make the number of com-
putations comparable to Siegman’s method by a priori computing and storing the zeros of the Bessel
function. The advantages of Yu et al.’s method over Siegman’s are the accuracy for the sampled points
and a simple retrieval expression. For more comparison between Yu et al.’s and Siegman’s method, we
refer readers to table 1 in Yu et al. [21]. Also, we note that the use of the adaptive variation of the longi-
tudinal propagation stepping size Δz and the transverse spatial sampling size according to 1/c2(z) and
1/c(z) allow us to track on-axis amplitudes, for the paraxial case, up to two orders of magnitude more
than what is achievable without the adaptive algorithm, for both the AFHTSS and AFFTSS methods.
Without the adaptive variation, the numerical methods become unstable, and we witness oscillatory
focusing and defocusing of the beam from numerical instability. Typical run times on a Pentium IV
2.4 GHz processor with a 2 GB RAM are around 1 min for AFHTSS when S = 2πR1R2 = 2π × 2000,
10 min for AFFTSS when the mesh size is N 2 = (210)2. Note that the Hankel transform based method,
which exploits the cylindrical symmetry, is one dimensional, and therefore is expected to be faster
than other two dimensional FFT based numerical methods.
Kalumniator, 400.
Kaput legis, 463, n. 6.
Κήρυκες, 153, n. 3.
King, auspices of, 103;
presidency of contio, 140;
of comitia calata, 154;
curiata, 173 ff.;
right to address people, 145, 173;
as legislator, 177 f.;
irresponsible, 180;
powers of, 181;
jurisdiction, 182;
election, 182-4, 189 f.;
declares war, 175 f., 181, 230.
Klebs, on reformed comitia centuriata, 223, 225.
Knights, see Equites.
Kornemann, on lex Scantinia, 357, n. 13.
Our website is not just a platform for buying books, but a bridge
connecting readers to the timeless values of culture and wisdom. With
an elegant, user-friendly interface and an intelligent search system,
we are committed to providing a quick and convenient shopping
experience. Additionally, our special promotions and home delivery
services ensure that you save time and fully enjoy the joy of reading.
ebookultra.com