Anda di halaman 1dari 40
OPTICAL ELECTRONICS AJOY GHATAK and K. THYAGARAJAN Physics Department, Indian Institute of Technology, New Dethi CAMBRIDGE UNIVERSITY PRESS Cambridge New York New Rochelle Melbourne Sydney Published by the Press Syndicate of the University of Cambridge ‘The Pitt Building. Trumpington Street, Cambridge CB2 IRP 32 East S7th Street, New York, NY 10022. USA 10 Stamford Road, Oakleigh Melbourne 3166, Australia © Cambridge University Press 1989 First published 1989 Printed in Great Britain at the University Press, Cambrid, British Library cataloguing in publication data Ghatak, A.K. (Ajoy Kumar), 1939 Optical electronics. 1. Electro-optics 1 Title H. Thyagarajan, K 535 Library of Congress cataloguing in publication data Ghatak, A-K. (Ajoy K.), 1939 Optical electronies/A.K. Ghatak and K. Thyagarajan. p.cm. Bibliography: p. Includes index. ISBN 0-521-30643~4. ISBN 0-521-31408-9 (pbk.) 1Optoelectronics. 1. Thyagarajan, K. IL. Title. TAI750.G48 1989 621.38'0414-del9 8810291 CIP ISBN 0 521 30643 4 hard covers ISBN 0 521 31408 9 paperback ™ Ld 1.2 13 14 1.5 1.6 2.3 24 2.5 2.6 2.7 28 3.1 3.2 3.3 3.4 3.5 Contents Maxwell’s equations and propagation of electromagnetic waves Introduction Maxwell's equations Plane waves in a dielectric The Poynting vector The complex notation Wave propagation in an absorbing medium Additional problems Reflection and refraction of electromagnetic waves Introduction Reflection and refraction at the interface of two homog- eneous nonabsorbing dielectrics Total internal reflection and evanescent waves Reflection and transmission by a film Extension to two films Interference filters Periodic media Reflection and transmission in the presence of absorbing media Additional problems Wave propagation in anisotropic media Introduction Double refraction Some polarization devices Plane waves in anisotropic media Wave refractive index Mes ASEH 27 28 42 46 54 7 57 60 62 63 63 63 69 74 78 vi 3.6 3.7 3.8 41 42 43 44 45 4.6 47 48 49 4.10 4.11 5.1 5.2 5.3 5.4 5.5 5.6 5.7 58 59 5.10 6.1 6.2 6.3 64 6.5 Contents Ray refractive index The ray velocity surface The index ellipsoid Problems Fraunhofer diffraction Introduction The diffraction formula Rectangular aperture The single slit diffraction pattern Circular aperture Directionality of laser beams Limit of resolution Resolving power of a microscope Annular aperture and apodization Fraunhofer diffraction by a set of identical apertures Resolving power of a prism Additional problems Fresnel diffraction Introduction The diffraction integral Uniform amplitude and phase distribution Diffraction of a Gaussian beam Intensity distribution near the paraxial image point of a converging lens Fresnel diffraction by a circular aperture Babinet’s principle Fresnel diffraction due to a circular disc Diffraction at a straight edge Fresnel diffraction by a long narrow slit Problems Spatial frequency filtering Introduction The Fourier transform and some of its important properties The Fourier transforming property of a thin lens Some elementary examples of the Fourier transforming property of a lens Applications of spatial frequency filtering 84 87 89 97 99 102 106 108 114 117 121 122 124 135 137 140 140 141 141 144 148 152 152 154 160 165 167 167 168 169 172 175 6.6 67 71 72 73 14 75 76 17 8.1 8.2 83 84 8.5 8.6 8.7 88 9.1 9.2 93 9.4 9.5 9.6 97 98 99 9.10 10 10.1 10.2 10.3 10.4 Contents Phase contrast microscope Image deblurring Holography Introduction The basic principle Coherence requirements Resolution Fourier transform holograms Volume holograms Some applications Lasers: I Introduction The Einstein coefficients Light amplification The threshold condition Laser rate equations Variation of laser power around threshold Optimum output coupling Line broadening mechanisms Additional problems Lasers: I Introduction Modes ofa rectangular cavity and the open planar resonator The quality factor The ultimate linewidth of the laser Mode selection Q-switching Mode locking in lasers Modes of a confocal resonator system General spherical resonator Higher order modes Some laser systems Introduction Ruby lasers Neodymium based lasers The He-Ne laser vii 177 178 181 181 182 188 188 190 194, 194 245 246 252 254 256 264 272 280 287 291 294 294 294 297 300 viii 10.5 10.6 10.7 10.8 10.9 It M1 11.2 11.3 11.4 11.5 11.6 11.7 11.8 11.9 RHR won Contents The argon ion laser The CO, laser Dye iasers Excimer lasers Semiconductor lasers Problems Electromagnetic analysis of the simplest optical waveguide Introduction Classification of modes for a planar waveguide TE modes in a symmetric step index planar waveguide TM modes in a symmetric step index planar waveguide The relative magnitude of the longitudinal components of the E and H fields Power associated with a mode Radiation modes Excitation of guided modes Maxwell's equations in inhomogeneous media: TE and TM modes in planar waveguides Additional problems Leaky modes in optical waveguides Introduction Quasi-modes in a planar structure Leakage of power from the core The matrix method for determining the propagation charac- teristics of planar structures which may be leaky or absorbing Calculation of bending loss in optical waveguides Optical fibre waveguides Introduction The optical fibre The numerical aperture Pulse dispersion in step index fibres Scalar wave equation and the modes of a fibre Modal analysis for a step index fibre Modal analysis of a parabolic index medium Pulse dispersion Multimode fibres with optimum profiles 302 302 304 308 308 314 31s 315 318 319 328 330 331 333 334 339 342 347 347 349 352 354 358 364 364 367 368 369 374 377 387 390 395 13.10 13.11 13.12 13.13 13.14 14.1 14.2 14.3 14.4 14.5 14.6 15 15.1 15.2 15.3 15.4 15.5 15.6 16 16.1 16.2 16.3 16.4 17 17.1 17.2 17.3 17.4 Contents First and second generation fibre optic communication systems Single mode fibres The Gaussian approximation Splice loss The vector modes Integrated optics Introduction Modes in an asymmetric planar waveguide Ray analysis of planar waveguides WKB analysis of inhomogeneous planar waveguides Strip waveguides Some guided wave devices Additional problems The electrooptic effect Introduction The electrooptic effect in KDP crystal The electrooptic effect in KDP crystals: The electrooptic effect in lithium niobate and lithium tantalate crystals General considerations on modulator design The index ellipsoid in the presence ofan external electric field Additional problems longitudinal mode ransverse mode The strain optic tensor Introduction The strain optic tensor Calculation of Ae for a longitudinal acoustic wave propagat- ing in an isotropic medium Calculation of Ae for a shear wave propagating along the z- direction in lithium niobate Problems Acoustooptic effect: Raman—Nath diffraction Introduction Raman Nath and Bragg regimes of diffraction A simple experimental set up to observe Raman-Nath diffraction Theory of Raman. Nath diffraction 400 403 410 411 416 421 421 422 432 433 437 441 455 461 461 462 473 475 477 492 498 502 502 502 504 505 506 508 508 508 510 Si 18 18.1 18.2 18.3 18.4 18.5 18.6 18.7 18.8 18.9 19 19.1 19.2 19.3 19.4 19.5 20 20.1 20.2 20.3 GUae> aAmm Contents Acoustooptic effect: Bragg diffraction Introduction Small Bragg angle diffraction Basic equations governing Bragg diffraction Coupled wave analysis for small Bragg angle diffraction Large Bragg angle diffraction Application to periodic media Transition to the Raman-Nath regime Vector approach to coupled wave equations Evaluation of é'Acé, Acoustooptic devices Introduction Raman-Nath acoustooptic modulator Bragg modulator Acoustooptic deflectors Acoustooptic spectrum analyser Nonlinear optics Introduction The self focussing phenomenon Second harmonic generation Appendices Wave equation and its solutions The index ellipsoid Density of modes Solution of the scalar wave equation for an infinite square law medium Leakage calculations of a packet of radiation modes WKB analysis of multimode fibres Coupled mode equations References and suggested reading Index 519 519 520 524 526 532 537 540 541 542 546 546 546 548 559 562 564 564 565 569 587 589 592 595 596 598 663 609 612 621 Preface Ever since the invention of the laser in 1960, there has been a renaissance in the field of optics and the field of optical electronics encompassing generation, modulation, transmission etc. of optical radiation has gained tremendous importance. With optics and optical electronics now finding applications in almost all branches of science and engineering, study of these subjects is becoming extremely important. The present book intended for senior undergraduate and first year graduate students is an attempt at a coherent presentation of the basic physical principles involved in the understanding of some of the important optoelectronic effects and devices. The book starts with the basic formulation of the study of propagation of electromagnetic waves, reflection and refraction and propagation through anisotropic media. This is followed by diffraction and its application in the study of spatial frequency filtering and holography. Basic physics behind laser operation is treated next with a brief discussion on different laser types. The next four chapters deal with the subject of optical waveguides including fibre and integrated optics which are already revolutionizing the field of information transmission. The next five chapters deal with three very important effects which are used in many opto-electronic devices namely the electrooptic, acoustooptic and nonlinear optical effects. The various concepts in the book have been derived from first principles and hence it can also be used for self study. A large number of solved and unsolved problems have been scattered throughout the book. This should particularly help the reader to a better appreciation of the concepts developed and also to get a feel for the numbers involved. Some of the problems are intended to extend the range of understanding beyond what is derived in the book. The writing of the book started in 1979 and during the past eight years portions of the book have been used in various courses, workshops and summer schools. The feedback received from the students and participants xii Preface of these courses has been of immense value and has helped us in putting the book in its present form. During the preparation of the book, we have had numerous discussions with our colleagues, in particular with Professor M.S. Sodha, Professor I.C. Goyal, Dr B.P. Pal, Dr Arun Kumar, Dr Anurag Sharma, Dr Enakshi Sharma, Dr G. Umesh and Dr MLR. Shenoy; to them we are greatly indebted for many invaluable suggestions. Portions of the book have also been used by Professor M.S. Sodha, Professor IC. Goyal, Dr B.P. Pal, Dr Arun Kumar and Dr G. Umesh at IIT Delhi and Dr Enakshi Sharma at the University of Delhi. We are very grateful to them for their constructive criticisms. One of us (AG) used a part of this book in presenting a course of lectures at University of Karlsruhe, West Germany. The many stimulating discussions with Professor G. Grau, Dr W. Freude and Dr E.G. Sauter are gratefully acknowledged. We would also like to thank Ms Swagatha Banerjee, Mr U.K. Das, Ms Supriya Diggavi, Ms Vrinda Kalia, Ms Jacintha Kompella, Mr Verghese Paulose, Mr Saeed Pilevar, Mr Vishnu Priye, Mr M.R. Ramadas, Mr R.K. Sinha and Dr R.K. Varshney for their help during the preparation of the manuscript. We are grateful to Dr R.W. Terhune, Dr H. Kogelnik, Dr R.A. Phillips, Dr W. Freude and Dr M. Papuchon for providing some of the photographs appearing in the book. We are also grateful to Professor N.M. Swani, Director, IIT Delhi and Professor M.S. Sodha, Head of our department for their encouragement and support of this work. New Delhi Ajoy Ghatak 14 September 1987 K. Thyagarajan 1 propagation of electromagnetic waves 1d Introduction In this chapter we will use Maxwell's equations to derive the wave equation and study its solutions in a homogeneous, isotropic and linear medium; the medium could be either absorbing or nonabsorbing. The results derived in this chapter will be used almost throughout the book — in particular, the solutions will be the starting point in the next chapter in which we will study the reflection and refraction of electromagnetic waves by a dielectric and a metal surface. 1.2 Maxwell’s equations All electromagnetic phenomena can be said to follow from Maxwell's equations. These equations are based on experimental laws and are given by VG =p (1.1) V-B=0 (1.2) Vxé=—CB/ot (13) Vx H=JI+0G/ot (1.4) where p represents the charge density and J the current density; 6,7, 2 and # represent the electric field, electric displacement, magnetic induction and magnetic field respectively. We will consistently be using the MKS system of units. We will discuss the solution of above equations in a linear, isotropic and homogeneous medium where the following constitutive relations are satisfied Gab (1.5) B= iH (1.6) 2 Maxwell’s equations and propagation of waves and J=06 (1.7) The parameters ¢, jk and o are known as the dielectric permittivity, magnetic permeability and conductivity of the medium and since the medium has been assumed to be linear and homogeneous, these parameters have a constant value. Using the constitutive relations and assuming the medium to be charge free (i.e., p = 0), Eqs. (1.1)-(1.4) become V-E=0 (1.8) V-# =0 (19) Vx 8 = —new/ot) (1.10) Vx H = 08 +€(06/ct) (1.11) ing the curl of Eq. (1.10) and using Eq. (1.11) we obtain Vx(Vxd)= — Haw x H) (1.12) But* Vv =ViV-E)—-Vx(Vxé) (1.13) * For an anisotropic medium, the parameters form a tensor so that, for example, % and & are not in the same direction; see Chapter 3 We should mention here that (contrary to what is written in many books) Eq. (1.13) is not a vector identity. Eq. (1.13) defines the operator ting on a vector, However, simple vector manipulations show that if we take a Cartesian component of Eq. (1.13), we would obtain v6, where the V? operator on the RHS is now divgrad. Thus (V6), =VAVE) On the other hand, if we take a non-Cartesian component, the above equation is no longer valid. For example working in the cylindrical coordinates, it can easily be shown that 1Pé, C78, + rer ce 206, = divgrad 6 -[ : o" Le ag which contains wo extra terms in addition to divgrad 4, 1,3 Plane waves in a dielectric 3 Since V-& = 0 (see Eq. (1.8)), Eq. (1.12) becomes ap, 06 8 V6 = po's + Hea (1.14) Similarly, taking the curl of Eq. (1.11) and using Eqs. (1.9) and (1.10) we get (1.15) VW = p0' a é 13 Plane waves in a dielectric We consider a perfect dielectric for which o = 0. Thus Eqs. (1.14) and (1.15) simplify to VE = pe(02E /c1?) (1.16) VM = pe(0? MH /00) (1.17) which are known as vector wave equations. If we consider a Cartesian component of either of the two equations, we would obtain the scalar wave equation VOW = e(C2 W/et?) (1.18) where Y may represent é6,. 6, or &.. or #,, #, or #.. The solution of the above equation represents waves (see Appendix A) and therefore, Maxwell's equations (which were used to derive the wave equation) predict the existence of electromagnetic waves. The speed of these waves is given by (see Appendix A) v= (en)? (1.19) For free space € = € 9 = 8.854 x 10°? C/N m? 2 [= flo = 4x x 10°7 N/A? (1.20) The speed of electromagnetic waves in free space is denoted by the symbol c and is given by = I€otto)! = 2.99794 x 108 m/s (1.21) It may be worthwhile mentioning that except for the term corresponding to the displacement current (= 6%/ér) in Eq. (1.4), all the experimental laws which are described in Eqs. (1.1)-(1.4) were known before Maxwell. By introducing the concept of displacement current, Maxwell (around 1860) could derive the wave equation (Eqs. (1.16) and (1.17)) and predict the 4 Maxwell’s equations and propagation of waves existence of electromagnetic waves.’ Further, using the value of €9 available to him, Maxwell found that the velocity of these electromagnetic waves should be about 3.1074 x 10®m/s. During the time of Maxwell the best known value of the speed of light was 3.14858 x 10° m/s (measured by Fizeau in 1849) and with ‘faith in rationality of nature’, Maxwell said that these two numbers cannot be accidentally equal and therefore light must be an electromagnetic wave. In the words of Maxwell himself, the speed of electromagnetic waves ---calculated from the electromagnetic measurements of Kohlrausch and Weber, agrees so exactly with the velocity of light calculated from the optical experiments of M. Fizeau, that we can scarcely avoid the inference that light consists in the transverse undulations of the same medium which is the cause of electr phenomena. and magnetic Now, in a dielectric, the velocity of propagation of the electromagnetic wave can be written in the form (see Eqs. (1.19) and (1.20)) vec/n (1.22) where n, known as the refractive index of the medium, is given by ep} _ 1.23 ae 02 For most dielectrics jt is very close to fig and we have n=K? (1.24) where K=€/ey (1.25) is known as the dielectric constant of the medium. In Appendix A we have shown that the solution of the wave equation (Eq. (1.18)) can be written in the form w= Actor bo (1.26) where A is a constant and k,,k, and k, (which represent the components of the vector k) and w can take arbitrary values subject to the condition that Paki +k} +k? =07/0? =o7 ep (1.27) * It was only in 1888 that Hertz carried out experiments which could produce and detect clectromagnetic waves of frequencies much smaller than that of light 1,3 Plane waves in a dielectric 5 or o=k(ew!. (1.28) Eq. (1.26) represents a plane wave propagating in the direction of k and the phase fronts are normal to k (see Appendix A). It should be mentioned that for a given frequency «, the value of k is fixed (see Eq. (1.28)); however, we can have waves propagating in different directions depending on the relative values of k,,k, and k,. For example, for a wave propagating along the x-direction, k=kk and the phase fronts are parallel to the y~z plane. On the other hand, for, we have a plane wave which is propagating in the x-y plane making 30°, 60° and 90° with the x, y and z-axes respectively (see Fig. 1.1). For all points on a plane perpendicular to k, the quantity k-r, and therefore the phase, is a constant. Returning to Eqs. (1.16) and (1.17) we see that since each Cartesian component of & and .# satisfies the scalar wave equation, the plane wave Fig. 11 A propagating plane wave with its propagation vector k making angles 30°, 60° and 90° with the x,y and z-axes respectively. For all points on a plane perpendicular to k. the quantity k-r(=|k|-OL), and hence the phase, is a constant. 6 Maxwell’s equations and propagation of waves solutions of Eqs. (1.16) and (1.17) can be written in the form 6 = Eel kn) (1.29) = Hel (1.30) where the vectors E and H are assumed to be independent of space and time. The various components of & and ¥ are related to each other through Maxwell’s equations. Let us consider the x-component of Eq. (1.10): a 66,/dy — 06, = — (0%, /ét) (1.31) If we now substitute the various components of Eqs. (1.29) and (1.30) in Eq. (1.31) we obtain —ik,é, +ik,6,= —ione, or (kx &), =0n(H)s and similar equations for the y and z-components. Thus H = (l/ok x & (1.32) Similarly, by substituting the various components of Eqs. (1.29) and (1.30) in the various components of Eq. (1.11) (with o = 0) we would obtain €=—-kx He (1.33) From Egg. (1.32) and (1.33) it is obvious that the vectors & and # and k form a rectangular triad of vectors. & and # are at right angles to each other and also to k; thus the fields associated with a plane electromagnetic wave are transverse to the direction of propagation: the transverse character of waves could have been directly obtained by substituting Eqs. (1.29) and (1.30) in Eqs. (1.8) and (1.9) respectively to obtain k-6=0 (1.34) and k-#=0 (1.35) However, the above two equations do not show that & and # are at right angles to each other. It may be noted that if we substitute for W from Eq. (1.32) in Eq. (1.33) we obtain Paoreu which is consistent with Eq. (1.28). Without any loss of generality we may assume the propagation to be 1.3 Plane waves in a dielectric 7 along the z-axis, ie. k=%k (1.36) As a special case we assume the electric vector to be along the x-axis; thus the actual electric field variation is assumed to be of the form & =XE,cos(wt — kz) (1.37) Using Eq. (1.32) we readily obtain the corresponding magnetic field variation H = $H cos (wt — kz) (1.38) with the amplitudes of the electric and magnetic fields (Ey and Ho) related through the following equation Ho = (k/opEg = (e/p)5 Eo (1.39) Notice that the & and # fields are in phase, ie., at a particular value of time and on a plane z = Zo(zo being arbitrary), if the & field has attained its maximum value then the field will also be at its maximum value etc. The wave described by Egs. (1.37) and (1.38) is said to be linearly polarized because the & (or #) fields are always along a particular direction. It is also referred to as a plane polarized wave because the electric vector always lies in the x-z plane (and the magnetic vector in yz plane ~ see Fig. 1.2). The most common terminology (to describe the state of polarization corresponding to Eqs. (1.37) and (1.38)) is to refer it as an x-polarized wave. We next consider a y-polarized wave (with an additional phase of 47) described by & = JE cos(it — kz + 4x) = — PEpsin(wt — kz) (1.40) and H = —%H,cos(wt — kz + 42) = KH sin(wt — kz) (1.41) Fig. 1.2 An x-polarized plane electromagnetic wave propagating along the z-direction. The arrows represent the direction and magnitude of the € and H vectors at a particular instant of time. The electric vector always lies in the x-z plane and the magnetic vector always lies in the yz plane. 8 Maxwell’s equations and propagation of waves with Eg and Hy related through Eq. (1.39). Since Maxwell’s equations are linear, a linear superposition of Eqs. (1.37) and (1.40) (and a corresponding superposition of Eqs. (1.38) and (1.41)) will also be a valid solution leading, in general, to an elliptically polarized wave. In particular, just a simple addition of the two solutions would lead to & = E,[Xcos (wt — kz) — § sin (wt — kz)] and (1.42) H = Ho[¥cos(wt — kz) + Xsin(wt — kz)) If we consider any plane perpendicular to the direction of propagation ~ in particular, if we consider the plane z = 0 then at all points on the plane, the time dependence of the fields would be given by 6,=E,coswt and 6,= Heer} 1.43) HA,=Hysinwt and #,= Hy cost 7 Thus 6 =62462=B3 and H? =H? 4 HP MB implying that the electric and magnetic vectors rotate on the circumference of a circle and if we look along the direction of propagation of the wave (ie., along the z-axis) then the electric vector will rotate in an anticlockwise direction (see Fig. 1.3(a)). Such a wave is known as a left circularly polarized wave. On the other hand, at a particular instant of time (say 1 = 0), we have 6, =Epcoskz, 6, =Eosinkz (1.44) and 82 = 62482 = 6B Thus at a particular instant of time the tip of the electric vector (and similarly the tip of the magnetic vector) traces a right handed helix along the z-direction (see Fig. 1.3(b)). In general. if we superpose two mutually orthogonal linearly polarized waves propagating along the same direction but with different amplitudes and phases, we would have (at a particular value of 2, say 2 = 0) 6, = E,cos(wt — 0,) é, = E, cos(wt — 43)5 (1.45) Thus é,/E, =coswtcos@, + sin ct sin 0, 1.3 Plane waves in a dielectric 9 = cos wt cos 0, + sin wt sin 0, Simple manipulations give us (6,/E,)sin 0, —(6,/E,)sin 0, = cost sin (0, —0,) Fig. 1.3 (a) For the electric (and magnetic) fields given by Eq. (1.44), the electric and the magnetic vectors rotate on the circumference of a circle in an anticlockwise direction. The propagation is into the plane of the page and the wave is said to be left circularly polarized. (b) At a particular instant of time, the tip of the electric vector traces a right handed helix along the z-direction (@ () 10 Maxwell's equations and propagation of waves and (6,/E,)cos 0, —(6,/Ez)cos 0, = —sin et sin(0, — 0,) If we square and add we would obtain (6,/E,)° + (6,/E2)? — 2(6,/E,)(6,/E>) cos 0 = sin? (1.46) where 0 = 0, —0, represents the phase difference between &, and &,. Eq. (1.46) represents an ellipse (see Problem 1.1 and Fig. 1.4) and the electric Fig. 1.4 The superposition of two mutually orthogonal linearly polarized waves leads, in general, to elliptically polarized waves. The figure below shows the states of polarization corresponding to different values of @ for E, = Ey g n, 4 B O=!n (b) O=n O= fn te) wo e ™ O=2k Hy wo 1.3 Plane waves in a dielectric 11 vector rotates on the circumference of the ellipse; such an electromagnetic wave is said to be elliptically polarized. For 0 = $2, 3x,...,the major and minor axes of the ellipse are along the x and y-axes and in addition if E, =E, the ellipse becomes a circle. In general, the axes of the ellipse are tilted with respect to the x and y-axes (see Fig. 1.4). For 0=0, 7, 27,... the ellipse degenerates into a straight line 8,/8.= $E,/E which represents a linearly polarized wave. Fig. 1.4 shows the states of polarization corresponding to different values of @ for E,=E,. The production and detection of different types of polarized light is discussed in Chapter 3. From the above analysis and Fig. 1.4, it is obvious that any general elliptical polarization can be written as a superposition of two mutually orthogonal linearly polarized waves. Problem 1.1: Show that the equation xy cos (= sin? (1.47) ab represents an ellipse. Determine the direction of the major and minor axes and their lengths. Solution: We consider a rotated coordinate system x = Ecos @ —sing y (1.48) Esing +1 00s @ We substitute in Eq. (1.47) and equate the coefficient of the €7 term to zero which gives us P cos 0 (1.49) cl tan2¢ = anne a2 — b?) In the rotated coordinate system, Eq. (1.47) becomes 1 (1.50) where 1 (cos? sin? cos. \ - 2 x ( a tg ab M8 J neg 1 in? 26 0 1 a sin? cos? , cos e ( a@ top tab vn26) 70 Eq. (1.50) represents an ellipse with semiaxes of lengths x and f and oriented along 12 Maxwell’s equations and propagation of waves the € and y-axes. For example for the ellipse corresponding to Fig. 1.4(b), since a=b,o=4n. 14 The Poynting vector In this section we will discuss the power flow associated with an electromagnetic wave. We start with Eqs. (1.3) and (1.4) Vx&=-CB/it (1.52) and Vx H=I+eG/et (1.53) We next use the vector identity’ VE x HM) = —E(V x H)+ HAV x6) (1.54) and substitute Eqs. (1.52) and (1.53) on the RHS to obtain VE x H)= —S-E -(6-6D/it + H-CB/ét) (1.55) If we integrate over an arbitrary volume and use Gauss’ theorem‘ we would obtain } S-fida= -f J-6dV— I (6-CL Cr+ H-CBlojav (1.56) A v v where S=ExH (1.57) is known as the Poynting vector and the surface integral is over the surface bounding the volume V. Eq. (1.56) is usually referred to as the energy law associated with the electromagnetic field. It may be pointed out that Eq. (1.56) is rigorously correct in the sense that we have used only Maxwell's equations and not the constitutive relations. If we do use Eqs. (1.5) and (1.6) we can write Go _ 06, 06? 16 6G 8S = Ne G5 (ED) (1.58) * The vector identity can easily be proved in Cartesian coordinates. } According to Gauss’ theorem [re v where 4 represents the surface bounding the volume V and fi represents the unit (outward) normal to the surface. Feada 14 The Poynting vector 13 Similarly (1.59) Thus (1.60) where w= be6?=16-F (1.61) and Wy = eH? =1A-B (1.62) represent the energy densities associated with the electric and magnetic fields respectively. Thus, if we assume the validity of the constitutive relations, Eqs. (1.55) and (1.56) may be written in the form VS + éw/ét= —J-é (1.63) and f S-ida= -| J-&dV—(d/dt) | wav (1.64) 4 v v The term { J-EdV ¥ on the RHS of Eq. (1.64) represents the total dissipated power within the volume V. For example, in a conductor J=o6& and o&? represents the ohmic power dissipated per unit volume. We should mention here that the current density J in Eq. (1.53) can be considered to be the sum of two parts J=JI.+d, (1.65) where J. represents the conduction current density and J, the convection current density due to moving charges. Thus, if a charge Q (moving with velocity v) is acted on by an electromagnetic field then the work done by the field in moving it through a distance ds would be F-ds where F=Q(6 + vx B) (1.66) represents the Lorentz force. Thus the work done per unit time would be F-ds/dt = Q(6 +¥ x B)-v =08-v 14 Maxwell’s equations and propagation of waves If there are N particles per unit volume, each carrying a charge Q then the work done per unit volume per unit time would be NOv-6 =I. (1.67) where J,=Nov (1.68) represents the convection current density due to moving charges. The energy given by Eq. (1.67) appears in the form of kinetic (or heat) energy of the charged particles. Returning to Eq. (1.64), if we consider electromagnetic fields which vanish at large distances from the origin and if the surface area A is very far away (so that the surface integral vanishes), Eq. (1.64) can be written in the form dw, a-—[ J-EdV (1.69) ¥ where we | wav | (wv, + w,) dV (1.70) v v Since the RHS of Eq. (1.69) has been shown to represent the total dissipated power within the volume V, we may interpret W to represent the total electromagnetic energy contained within the volume V; w, and w,, may therefore be interpreted to represent the energy per unit volume associated with the electric and magnetic fields respectively. Thus in Eq. (1.64) (considering the integrals over an arbitrary volume V) we have the following physical interpretation of the various terms. (a) The term [,J-& dV represents the total (instantaneous) dissipated power in the volume V. (b) The term {,, wdV represents the total electromagnetic energy in the volume V. (c) Because of the above, the term s-aae= [ (E x #)-ida A A represents the total power flowing out of the volume V. For a better understanding, we consider a few specific cases: Case 1: For plane waves in a dielectric we can write for the electric and magnetic fields (see Eqs. (1.37) and (1.38)): & = XE, cos (wt — kz) (1.71) 1.4 The Poynting vector 15 H = FH cos (wt — kz) (1.72) Thus SeEZ cos? (cot — kz) (1.73) For optical frequencies, the cos? term fluctuates with extreme rapidity so that we should take a time average of Eq. (1.73) to obtain = 4€E2 (cos?(wt — kz)> where <---> denotes the time average of the quantity inside the angular brackets; the time average of a time dependent function is defined by the following equation =U 1 [roa o for a periodic function, one may choose T as the time period (= 27/cv). Thus cos? (cot — kz)) = (w/2n) [- * cos? (wt — kz) dt = 4 (1.74) Thus ° (we) = 2€ES Similarly (Wm > = 4MHG From Eq. (1.39) we) = CW) = 2S and therefore Cw) = 2.) = FEES (1.75) The Poynting vector is given by S=6 x H = EgH, cos*(wt — kz)é Taking the time average and using Eq. (1.39) we get (S) = Mk/cop) Eg = €/W)* E52 (1.76) Since the velocity of the wave, v is equal to «w/k(=«(es)?), the above expression for the Poynting vector can be written as (S> = Cw) vb (1.77) Thus for a plane wave in a dielectric we may interpret S-fida as the rate of energy flow through the area da. 16 Maxwell’s equations and propagation of waves Case 2: The quantity S-fida does not always represent the rate of energy flow through the area da; for example, we may have static electric and magnetic fields where & x # is finite but we know that there is no energy flow. However, the integral [ S-fida A over a closed surface rigorously represents the net energy flowing out of the surface. As a specific example we consider a long cylindrical conductor carrying a steady current J. Assuming constant current density (along the z-direction) the electric field inside the conductor is given by & =2J,/0) = a{I/oA) (1.78) where A represents the area of cross section of the conductor. Using Ampére’s law, the magnetic field at the outer surface of the conductor is given by (see, e.g., Johnk (1975) p. 272 and p. 405) H = $(1/2xp) (1.79) where p, represents the radius of the cylindrical conductor. We consider the integral fe x H)tida over the surface of a cylinder of length | whose curved surface coincides with the surface of the conductor. Since Pr ExXH=— Bony jad (using Eqs. (1.78) and (1.79)), & x # is normal to the flat surface at the ends and therefore there is no contribution (to the surface integral) from the ends. From the curved surface we have a ft ; Pp *ag)( (ee [PI a =r seetese~ —asea( fo 8e)( [i fue x H)yida= -P(I/Ac)= —PR (1.80) Thus where we have used the fact that the quantity |/Ao is the resistance R of the length / of the conductor. The quantity /?R is the power dissipated in 1.4 The Poynting vector 17 the length / of the conductor. Further [ssara0[eav—cr 0° A*)(Al) = ?-R (1.81) consistent with Eq. (1.80) Case 3: We consider an infinitesimal oscillating dipole at the origin whose dipole moment is along the z-axis and is assumed to be given by p=Zpocoswt (1.82) The far field radiation pattern is given by (see, e.g., Panofsky and Phillips (1962) p. 258): = K'Po sin gOS" “Kg (1.83) TLE w= ok Po sin) cos (wt — kr) é 4n r (1.84) where we have used spherical polar coordinates (r, 0,2) and have assumed free space surrounding the dipole (see Fig. 1.5). Eqs. (1.83) and (1.84) represent (outgoing) spherical waves and are valid when Ar > 1. Notice that Fig. 1.5 An infinitesimal oscillating dipole at the origin whose dipole moment is along the z-axis. At large distances from the dipole, &, % and § are mutually orthogonal and are along 6,6 and # respectively. 18 Maxwell’s equations and propagation of waves the & and # fields are in phase and the magnitudes are related through Eq. (1.39). The Poynting vector is given by _ ok pin? 0, “Re ae where we have taken the time average. The energy flows radially and the intensity falls off as 1/r? which is the inverse square law. The quantity S-fida would represent the energy flow per unit time through the area da and if we integrate over a sphere of radius R we would obtain (8)-ada = 28 PO_ (* (* sin? OR? sin 0 d0de . Bane R? |, J, _ ok3p§ ~ 12n€ (1.86) which represents the total radiated power. 1.5 The complex notation In many problems it is very convenient to use the complex notation in which the electric and magnetic fields are written in the form 6, =E,e br) (1.87) and He, = Hyer bron (1.88) where E, and H, are assumed to be real vectors. The actual electric and magnetic fields will be the real parts of the RHSs, ie, E, cos(wt—Kk-r+0,) and Hy, cos(wt—k-r + ¢,) respectively. Now, if we have another electromagnetic wave characterized by the fields 6, and #,, then the resultant electric and magnetic fields will be and (1.89) Reé, + Re&, = Re(é, + €;) } ReH, +ReH,=Re(H, +H) Thus we may superpose the complex fields and then take the real part. This leads to considerable simplification in many problems. One must, however, be careful in calculating any quantity which is not linear in & and # such as the energy density w or the Poynting vector S. For example, 1.6 Wave propagation in an absorbing medium 19 for the fields given by Eqs. (1.71) and (1.72), the complex fields are given by 8 =KE ee) and H=FHe™) The energy density is given by (see Eq. (1.75) =2 x $e(Re&-Reé > (1.90) where Reé represents the real part of &. Thus Kw) = heC(E + E*)-(E + 6*)> where &* represents the complex conjugate of &. The time average of &-& and &*-&* are zero and therefore (in the complex notation) Kw) = 486% = 1eE2 (1.91) We next consider the time average of the Poynting vector = Since KE x H)=0=(6* x H*) we obtain <«S> =}Re<(é x #*) (1.92) In Sec. 1.6 we will show that in an absorbing medium the electric and magnetic fields are given by 6 =k Ege etior 29 (1.93) w=s(8) [1 +(Z) | ee te =a (1.94) Substituting in Eq. (1.92) we obtain t 2 s>~45(£) [+(&) ] Bie 2 cos (1.95) 1.6 | Wave propagation in an absorbing medium Let us first consider propagation of electromagnetic waves through a conducting medium. The corresponding wave equation is given by (see and 20. Maxwell’s equations and propagation of waves Eq. (1.14)) V8 = po(CE/et) + pe(O26/et) (1.96) If we assume a solution of the form S =RE,eitt (1.97) we would obtain —k? =i@po — pew? (1.98) The above equation implies that k must be complex. We write k=a-ip (1.99) substitute in Eq. (1.98), and equate real and imaginary parts to obtain a? — f? = precy? a8 = Sono (1.100) The above two equations can readily be solved to obtain a= ofp)? [5 + 41 + 07/w7e)3 J! (1.101) _opo _(p\iofl 1 o \iys p= -(') 5] +5(\4+ a (1.102) Since k is now complex, Eq. (1.97) becomes E =KEce Peltor) (1.103) which shows attenuation in the z-direction. The quantity 1 2/e\f1 1 2 \i7 b= 9-5 (5) [+a(+at) | (1.104) is referred to as the ‘skin depth’ and represents the distance in which the amplitude falls off by a factor of e. It may be noted that assuming harmonic solutions (ie., time dependence of the form e“) Eq. (1.11) takes the form Vx H=(0+ iweé = i@(€ — io/a)E eae! Thus all the analysis in the previous section remains valid except for the fact that € has to be replaced by (€ — ic/c). Thus instead of Eq. (1.27) we have I? = @*(e —io/o)n (1.106) 1.6 Wave propagation in an absorbing medium 21 which is identical to Eq. (1.98). Furthermore, using Eq. (1.97), Eq. (1.10) gives §(08,/62) = —iopxt Thus H =F§(k/opE ge" ™ «\! io \! -o(< IO 5 a pegitor a2) a() (: we) Ege “e' t 278 -3(£) [++(&) | Ege Fetes (1.107) where g =Htan”! (o/oe) (1.108) Thus the electric and magnetic fields are no longer in phase. For a good conductor (at low frequencies) we may assume ofoe > 1 (1.109) so that a= B= (Sopo)* (1.110) and xin (1.1) As an example, we consider copper for which at low frequencies ¢ is a constant and = 5.8 x 10’ mhos/m, further y ¥ fo so that so4a {Ser 10-3m at 100Hz x |21x10°*m at 100kHz (1.112) implying that the skin depth decreases with increase in frequency. We next calculate the Poynting vector and for the sake of simplicity we assume /we > 1 so that 2 B and ¢ = 4n. In this limit, the actual fields are given by & =SE,e~* cos(wt — 22) (1.113) and H =JH e * cos(wt — xz — p) (1.114) where Ho = (€/H)3[1 + (o/we)P Eo (1.115) If we calculate & x # and then take the time average we obtain (S) =(8 x H) =4EgH oe 2% cos b% (1.116) which again shows the exponential attenuation of the power flow. 22 Maxwell's equations and propagation of waves We should point out here that for a metal (such as sodium, copper, silver, etc.) ¢ may be assumed to be a constant only for frequencies < 10!3 Hz. At higher frequencies « becomes both frequency dependent and complex, as does €, and we may write aio" and Substituting in Eq. (1.98) we obtain =iwp(o' ~ io") — ple’ —ie")e0? = ioe’ + we") — We — "Jojo? (1.117) (1.118) (1.119) Thus the entire analysis given earlier in this section remains valid provided Fig. 1.6 The variation of and « with wavelength for copper (adapted from Lynch and Hunter, 1985) 10? oth a rc of 10% 10% 10" 10" 0 Wavelength (um) 1.6 Wave propagation in an absorbing medium 23 we make the following transformations +0 toe } (1.120) eve —o"/o Often it is more convenient to write (cf. Eq. (1.99)) k= (w/c)n = (w/c)(q — ik) (1.121) where 9 and x represent the real and imaginary parts of the (complex) refractive index n. The plane wave solution is therefore written in the form & =Epexp fi[ot —(w/e)(n —ik)z]} = Egexp[—(w/dxz] exp (iLor — (w/c)nz]} (1.122) The quantity x is usually referred to as the extinction coefficient. Substituting for k from Eq. (1.121) in Eq. (1.119) we readily get 9? —K? =c2p(e' — 0") (1.123) 2nk =(c?n/o)(o' + we") (1.124) Figs. 1.6 and 1.7 show the variation of 7 and x with wavelength for copper Fig. 1.7 The variation of and x with wavelength for gold (adapted from Lynch and Hunter, 1985), 10° 10* 10° 10? 10! 10” tot Wavelength (um) 24 Maxwell’s equations and propagation of waves Table 1.1. Optical constants of some metals A= 6888A Metal Kk n K Copper 0.993 245x 1073 0.213 4.05 Gold 0995 615x10% 0.160 3.80 Indium 0992 814x109 2644.81 Molybdenum 0.995 5.09105 3.81 3.58 Nickel 0993 -203x 10-3 2144.00 Platinum 0993 808x103 2.51 4.43 Silver 0.998 = 882x10% 0140 4.44 Tungsten 0993 650x 10° 3.82291 Adapted from Lynch and Hunter (1985) and gold. Notice that at low wavelenghts (i.e, high frequencies) the metal becomes almost transparent. Table 1.1 gives the values of optical constants for some metals at 4 = 31 A and 6888 A. Detailed tables for such variations can be found in Lynch and Hunter (1985). 1.6.1 Refractive index of a metal In a metal the conductivity is approximately given by Nq@? 9) = FG 4 iw) (1.125) where N represents the number of free electrons per unit volume, q the electronic charge, m the electronic mass and v, the collision frequency. Thus Nqy, = Ss 1.126 & = nv? +0) ame) and Nqo Mas 1.127 7 mb? +0) (1.127) Assuming € = €9 and ju ¥ {lo (which is true for most metals) we obtain by using Eq. (1.98) (0? /e?\n? = — iepigo’ — 10”) + gE? 2 2\-17 _o@ v2 iv, 1 3(1+%) (1+ =) (1.128) Op = (NG? /meo): (1.129) where Additional problems 25 Table 1.2. Values of the wavelength corresponding to the plasma frequency for different metals p N = 10°Nop/A (Adin = 2rclmeE/G?NY! (Apert Element (gem) A (m=) (um) (um) Ney Lithium 0.534 6.94 4.634 x 10° 1.552 2.050 0.573 Sodium 09712 22,99 2.544 x 10 2.095 2.100 0.995 Potassium 0.870 39.10 1.340 x 10°* 2.887 3.150 0.840 Rubidium | 85.48 1.079 x 107 3.217 3.600 0.799 The values of various parameters used are m=9.109 x 10° “kg, q= 1.602 x 101°C, € = 8.854 x 10° 1? C?/Nm?, No = 6.023 x 107° mole! is known as the plasma frequency. The above equation tells us that for high frequencies («> v,), the refractive index is essentially real with a frequency dependence of the form we OP (1.130) Thus for o>o,, the refractive index is real. Indeed, in 1933 Wood discovered that alkali metals are transparent to ultraviolet light. Assuming that the refractive index is primarily due to the free electrons and that there is one free electron per atom, the calculated values of = 2nc/w, = 2nclmey/Nq?)! are tabulated in Table 1.2 for lithium, sodium, potassium and rubidium. The experimental values are also given and as can be seen although the calculated values are of the same order as the experimental values, the agreement is not really perfect (except for the case of sodium where the theoretical and experimental values agree very well). One of the short- comings of the theory is the assumption that there is one free electron per atom; indeed, one may use the experimental values of 4, to calculate the effective number of free electrons per atom from the formula Additional problems Problem 1.2: Discuss the state of polarization of a plane electromagnetic wave when the x and y-components of the electric field are given by the following equations: (a) Ey = $Ep cos(t — kz) 3 y= SS Eosin (oor — kz) 26 Maxwell's equations and propagation of waves (b) Ey = Eo cost — kz) SEocos (wt — kz +2) E, cos (wt — kz) — Eg sin (wt — kz) (Answer: (a) Right elliptically polarized; (b) linearly polarized; (c) left circularly polarized.) Problem 1,3: Determine the direction of propagation of the following waves: k (a) E= E00s( on ; (b) E=E,cos (or + N N 3° N Problem 1.4: Consider a displacement of the form y= 0.2c0s( + = ms) \ v2 v2 where ¢ is measured in seconds and w, x and y are measured in centimetres. Determine the direction of propagation and calculate the wavelength and frequency. Problem 1.5: A circularly polarized electromagnetic wave is propagating in the z- direction in free space and is described by the following equation & = XS cos (it — kz) + 9S sin (wt — kz) V/m The wavelength is 6 x 10~’m. Find the corresponding magnetic field and show that the average of the Poynting vector is given by = 6.64 x 10°72 W/m? Problem 1.6: For a plane clectromagnetic wave in a dielectric medium the ratio of the magnitude of the electric and magnetic fields is (u/ey* which has the units of ohms and is referred to as the intrinsic impedance of the medium and is denoted by Z. Show that for free space Z=Z, ~376.7Q~1200Q 2 Reflection and refraction of electromagnetic waves 21 Introduction In this chapter we will study the reflection and refraction of electro- magnetic waves from an interface separating two media and from a stack of films. Such studies are very important in understanding many practical optical devices such as Fabry-Perot etalons, interference filters, special optical coatings etc. Furthermore, by studying the state of polarization of a light beam reflected from a medium, one can obtain its optical character- istics; this forms the basis of the field of ellipsometry. In deriving the reflection and transmission coefficients we will use the following continuity conditions’ at the interface: (a) continuity of the tangential components of the electric vector &; (b) continuity of the normal components of the displacement vector 7; (c) continuity of the tangential components of the magnetic field vector HA and (d) continuity of the normal components of the magnetic induction vector 2. We will find that the equations determining the reflection and transmis- sion coefficients fall into two groups: one of the groups contains only the components of & parallel to the plane of incidence (and W perpendicular to the plane) and the other group contains only the components of & perpendicular to the plane of incidence (and # parallel to the plane). Therefore the two cases (being independent of each other) will be considered separately and using them we can study the reflection (and refraction) of electromagnetic waves which have arbitrary states of polarization. We will, for example, show that a circularly polarized wave on reflection from a * These boundary conditions can be derived from Maxwell's equations and since the field distributions that we would be using would satisfy Maxwell's equations, it will not be necessary to use all the boundary conditions 28 Reflection and refraction dielectric surface can become elliptically polarized. Similarly, a linearly polarized wave reflected from a metal surface may be linearly polarized or circularly polarized or elliptically polarized depending on the angle of incidence and the direction of the electric vector associated with the incident wave. 2.2 Reflection and refraction at the interface of two homogeneous nonabsorbing dielectrics We first consider the reflection and refraction of a plane electro- magnetic wave incident at the interface of two dielectrics characterized by (€,, 41) and (€2, 413) (see Fig. 2.1). We assume the media to be nonabsorbing, isotropic and homogeneous. Let 6, =E,et 9 6, =E,e 8" (2.1) and | eles bye represent the electric fields associated with the incident, refracted and reflected waves respectively; the vectors E,,E, and E; are independent of space and time. Let @,,0 and 04 represent the (acute) angles that the vectors k,,k,, and k, make with the normal to the interface (see Fig. 2.1). Fig. 2.1 also shows the direction of the Cartesian axes; the plane 0 represents the interface of the two dielectrics, and the direction of the y-axis é3= Fig. 2.1 The reflection and refraction of a plane wave incident at the interface of two dielectrics; the electric vector is assumed to be in the plane of incidence (ic., p-polarized), 2.2 Reflection and refraction at dielectric interfaces 29 is such that ky =0 (2.2) ie., the propagation vector associated with the incident wave is parallel to the x~z plane. Since the fields satisfy the wave equation we must have ki = wey (2.3a) Kp =03€2 Ky (2.3b) k2 =e, py (2.3c) We consider two special cases: one in which the electric vector lies in the plane of incidence and the other in which the electric vector lies perpendi- cular to the plane of incidence. By using appropriate boundary conditions it can easily be shown that if the electric vector associated with the incident plane wave lies in the plane of incidence then the electric vectors associated with the reflected and refracted waves will also lie in the plane of incidence (see Problem 2.1) and the same is true for the perpendicular case. Case 1: & lying in the plane of incidence. We will first consider the case when the electric field associated with the incident wave lies in the plane of incidence (see Fig. 2.1). We resolve the electric vector & along the x and z-axis and since the z-component is tangential to the surface we must have ‘ontinuous across the interface; thus 6,,463.=683. (24) or [—E,el" * cos 0, — Ese" "cos 05], -0 =[- Ee" *" cos 02), -0 where all expressions are to be evaluated at the interface x = 0. Now ker=kx+ky+k yh + Kz (at the interface x = 0) Thus, we must have — Eyel fis 2")eqs 0, — Bete’ "2 c08 05 = = Eel hs" 08 0, ) The above equation has to be valid at all times and fox all values of y and 30 Reflection and refraction z (on the plane x = 0) and therefore we must have O=0)=04 (2.6) Kay = hay = kay (27) Kys=ha= hs, (2.8) Thus the frequencies associated with the reflected and refracted waves must be the same as that of the incident wave ~ which is also physically obvious. Thus ky = le, 6y)) = ky (2.9) and ky = wl€2ft3)8 (2.10) Now, since k,, = 0 (see Eq. (2.2)) we must have kay =kyy=0 (2.11) ie. k,,k, and ky will all be parallel to the x-z plane. Eq. (2.8) gives k, sin 0, =k, sin, =k, sin 0, (2.12) where we have used Eq. (2.9). Thus 1= Os (2.13) which says that the angle of incidence equals the angle of reflection. Further sin, ky _ (€2f2 \? _ m2 sin, ky (2 ny (2.14) where nae -(ae) (2.15) yy €ollo and 6 _ (eke 3 me (22) (2.16) represent the refractive indices of media 1 and 2 respectively and C=(€ollo) } represents the velocity of light in free space. Eq. (2.14) gives us n, sin 0, =n sin0, (2.17) which is Snell’s law of refraction. We should mention here that in the derivation of the laws of reflection and refraction (viz. Eqs. (2.6)-(2.17)) we have only used the fact that a

Anda mungkin juga menyukai