* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download Light-Matter Interaction: Conversion of Optical Energy and
Survey
Document related concepts
Transcript
Light-Matter Interaction: Conversion of Optical Energy and Momentum to Mechanical Vibrations and Phonons Masud Mansuripur, College of Optical Sciences, The University of Arizona, Tucson [Published in Quantum Sensing and Nano Electronics and Photonics XIII, edited by M. Razeghi, G. J. Brown, and J. S. Lewis, Proceedings of SPIE Vol. 9755, 975521~1-34 (2016).] Abstract. Reflection, refraction, and absorption of light by material media are, in general, accompanied by a transfer of optical energy and momentum to the media. Consequently, the eigen-modes of mechanical vibration (phonons) created in the process must distribute the acquired energy and momentum throughout the material medium. However, unlike photons, phonons do not carry momentum. What happens to the material medium in its interactions with light, therefore, requires careful consideration if the conservation laws are to be upheld. The present paper addresses some of the mechanisms by which the electromagnetic momentum of light is carried away by mechanical vibrations. 1. Introduction. When a light pulse arrives at the surface of a mirror, it bounces back and transfers twice its initial momentum to the mirror.1 The mechanical motion thus set off at the front facet of the mirror travels forward and gives rise to elastic vibrations through the entire thickness of the mirror and its substrate.2,3 Similarly, the passage of a light pulse through a transparent dielectric slab sets in motion elastic waves of mechanical vibration, which mediate the exchange of momentum between the optical wave and the material medium of the slab.4-11 The goal of the present paper is to analyze the aforementioned problems and to derive expressions for the elastic waves excited by the electromagnetic (EM) fields, emphasizing in particular the energy and momentum associated with these elastic vibrations. In preparation for the analysis, we begin in Sec. 2 with an elementary treatment of the continuum mechanics of solid objects, including the properties of elastic waves excited within these bodies. Several examples illustrate the one-dimensional motion of a solid object in the presence of external excitations, when the objectβs boundaries are fixed, and also when one or both of its boundaries are free. Then, in Sec. 3, the results of Sec. 2 pertaining to longitudinal elastic waves are extended to the case of transverse waves, where the conservation of angular momentum introduces additional complications, which require extra care and attention. Section 4 is devoted to an analysis of systems in a steady-state of motion, after the initial vibrations have died down, and the effects of external forces are reduced to static deformations of the solid object experiencing a constant and uniform acceleration. In Sec. 5 we derive the general equation of motion for a flexible one-dimensional rod that is free to move in three-dimensional space while undergoing arbitrary elastic deformations. The motion of the rod under these circumstances is quite complicated, and the corresponding equation is not amenable to analytic solutions for any specific examples. However, we will show the consistency of the equation of motion with the conservation laws of energy and linear as well as angular momentum. In Sec. 6 we return to the simpler problems associated with longitudinal vibrations of elastic rods in one dimension, and show the effects of a light pulse impinging on a high-reflectivity metallic mirror, as well as those of a light pulse entering a transparent rod from one end (with and without anti-reflection coating), then propagating along the length of the rod. Elastic vibrations of one-dimensional inhomogeneous media are taken up in Sec. 7. Here we derive general expressions for acoustic and optical phonons in periodic structures, and discuss the corresponding dispersion relations as well as issues related to energy and momentum conservation. Also discussed are the extension of these methods to more general situations using the theory of eigen-functions and eigen-values. Some concluding remarks appear in Sec. 8. The three appendices describe the Fourier transform theorems used throughout the paper, the optics of plane EM waves in simple optical systems, and technical details related to the topic of Sec. 7. 1 2. Continuum mechanics of acoustic vibrations and phonons. This section describes the basic equation of motion and its solution for longitudinal vibrations of a homogeneous, onedimensional medium modelled as a string of point-particles connected by short, identical springs. With reference to Fig.1, let π’π’(π₯π₯, π‘π‘) be the displacement from the equilibrium position at location π₯π₯ and time π‘π‘, ππ the mass-density (i.e., mass per unit-length along the π₯π₯-axis), and ππ the spring constant (or stiffness coefficient). Newtonβs second law of motion may be written as follows: ππ οΏ½ π’π’(π₯π₯+βπ₯π₯) β π’π’(π₯π₯) βπ₯π₯ β οΏ½ β ππ οΏ½ π’π’(π₯π₯) β π’π’(π₯π₯ββπ₯π₯) βπ₯π₯ οΏ½ = (ππβπ₯π₯)πππ‘π‘2 π’π’(π₯π₯, π‘π‘) (ππβππ )πππ₯π₯2 π’π’(π₯π₯, π‘π‘) = πππ‘π‘2 π’π’(π₯π₯, π‘π‘). (1) u(x,t) x xββx x x+βx Fig.1. Approximating a continuous, one-dimensional slab of homogeneous material with a linear chain of point-particles connected via short springs of length βπ₯π₯. The equilibrium position of the nth particle is π₯π₯ = ππβπ₯π₯, and the particleβs displacement from equilibrium at location π₯π₯ and time π‘π‘ is denoted by π’π’(π₯π₯, π‘π‘). The mass-density of the continuum being ππ, each point-particle has mass ππβπ₯π₯. Denoting the spring constant by ππ, the force exerted by each spring on its adjacent particles is given by ±πππππ₯π₯ π’π’(π₯π₯, π‘π‘). Defining the phase-velocity of wave propagation along the π₯π₯-axis by π£π£ππ = οΏ½ππβππ , we may solve Eq.(1) by the method of separation of variables, where π’π’(π₯π₯, π‘π‘) is written as ππ(π₯π₯)ππ(π‘π‘), and the separated equations become πππ‘π‘2 ππ(π‘π‘) = βππ2 ππ(π‘π‘) and πππ₯π₯2 ππ(π₯π₯) = β(ππβπ£π£ππ )2 ππ(π₯π₯). Here we have defined the real-valued constant ππ as the (arbitrary) frequency of vibrations in the time domain. Similarly, the real-valued constant ππ = ππβπ£π£ππ is the frequency of vibrations in the space domain. The general form of the separable solution of Eq.(1) is thus written π’π’± (π₯π₯, π‘π‘) = ππ± (ππ) exp[i(ππππ ± ππππ)]. (2) The initial state of the medium at π‘π‘ = 0, that is, its position π’π’(π₯π₯, π‘π‘ = 0) and its local velocity profile π£π£(π₯π₯, π‘π‘ = 0) = πππ‘π‘ π’π’(π₯π₯, π‘π‘ = 0) may thus be expressed as follows: 1 β π’π’(π₯π₯, π‘π‘ = 0) = 2ππ β«ββ[ππ+ (ππ) + ππβ (ππ)] exp(iππππ) ππππ, π£π£ππ β π£π£(π₯π₯, π‘π‘ = 0) = 2ππ β«ββ iππ[ππ+ (ππ) β ππβ (ππ)] exp(iππππ) ππππ. (3a) (3b) The functions ππ± (ππ) are readily derived from the Fourier-transformed initial conditions (ππ) ππ0 and ππ0 (ππ). In particular, if π’π’(π₯π₯, π‘π‘ = 0) = 0, we will have ππ+ (ππ) = βππβ (ππ) = ππ0 (ππ)β(2iπππ£π£ππ ). In the absence of dispersion, that is, when π£π£ππ is independent of the frequency ππ, the general solution of Eq.(1) will be π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + π’π’β (π₯π₯ β π£π£ππ π‘π‘). The velocity profile will then be π£π£(π₯π₯, π‘π‘) = π£π£ππ πππ₯π₯ [π’π’+ (π₯π₯ + π£π£ππ π‘π‘) β π’π’β (π₯π₯ β π£π£ππ π‘π‘)]. For a uniform, homogeneous, dispersionless medium that is also infinitely-long along the propagation direction π₯π₯, one can prove the conservation of energy (kinetic β°πΎπΎ plus potential β°ππ ) and of linear momentum πππ₯π₯ , as follows: 2 β β β° = β°πΎπΎ + β°ππ = ½ππ β«ββ π£π£ 2 (π₯π₯, π‘π‘)ππππ + ½ππ β«ββ[πππ₯π₯ π’π’(π₯π₯, π‘π‘)]2 ππππ β 2 = ½πππ£π£ππ2 β«ββοΏ½πππ₯π₯ π’π’+ (π₯π₯ + π£π£ππ π‘π‘) β πππ₯π₯ π’π’β (π₯π₯ β π£π£ππ π‘π‘)οΏ½ ππππ β 2 +½ππ β«ββοΏ½πππ₯π₯ π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + πππ₯π₯ π’π’β (π₯π₯ β π£π£ππ π‘π‘)οΏ½ ππππ β β = ππ β«ββ{[πππ₯π₯ π’π’+ (π₯π₯)]2 + [πππ₯π₯ π’π’β (π₯π₯)]2 }ππππ. (4) β πππ₯π₯ = β«ββ ππππ(π₯π₯, π‘π‘)ππππ = πππ£π£ππ β«ββ πππ₯π₯ [π’π’+ (π₯π₯ + π£π£ππ π‘π‘) β π’π’β (π₯π₯ β π£π£ππ π‘π‘)]ππππ = πππ£π£ππ [π’π’+ (β) β π’π’β (β) β π’π’+ (ββ) + π’π’β (ββ)]. (5) Both β° and πππ₯π₯ are thus seen to be time-independent, which is proof that the total energy and linear momentum of the system are conserved. Conservation of angular momentum is trivial to prove in the present case of longitudinal vibrations along π₯π₯; just pick an arbitrary reference point β οΏ½ × ππππ(π₯π₯, π‘π‘)ππ οΏ½ ππππ is zero at all times. π₯π₯0 on the π₯π₯-axis and observe that π³π³ = β«ββ(π₯π₯ β π₯π₯0 )ππ Example 1. Suppose that the initial position and velocity of our one-dimensional medium are specified as π’π’(π₯π₯, π‘π‘ = 0) = 0 and π£π£(π₯π₯, π‘π‘ = 0) = π£π£0 β[1 + (π₯π₯βπ€π€0 )2 ], respectively. As shown in Appendix A, the Fourier transform of the initial velocity will be ππ0 (ππ) = πππ£π£0 π€π€0 exp(βπ€π€0 |ππ|) and, consequently, ππ± (ππ) = ± πππ€π€0 π£π£0 exp(βπ€π€0 |ππ|)β(2iπππ£π£ππ ), which leads to π’π’± (π₯π₯, 0) = ±½(π€π€0 π£π£0 βπ£π£ππ ) arctan(π₯π₯βπ€π€0 ), as depicted in Fig.2. The temporal evolution of our system having the postulated initial conditions may thus be written as follows: π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯, π‘π‘) + π’π’β (π₯π₯, π‘π‘) = (a) π€π€0 π£π£0 2π£π£ππ π£π£ππ (b) β β«ββ ππ β1 exp(βπ€π€0 |ππ| + iππππ) sin(ππππ) ππππ. (6) π£π£± (π₯π₯) π£π£ππ π’π’± (π₯π₯) π₯π₯ π£π£ππ π£π£ππ π₯π₯ Fig.2. Plots of π£π£± (π₯π₯) and π’π’± (π₯π₯) at π‘π‘ = 0 inside an infinitely-wide slab. The plot of π’π’β (π₯π₯), shown with broken lines, is similar to that of π’π’+ (π₯π₯), except for being flipped around the horizontal axis. As time progresses, π’π’+ (π₯π₯) and π£π£+ (π₯π₯) move to the left, while π’π’β (π₯π₯) and π£π£β (π₯π₯) move to the right, all at the constant velocity π£π£ππ . The overall displacement and velocity are π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + π’π’β (π₯π₯ β π£π£ππ π‘π‘) and π£π£(π₯π₯, π‘π‘) = π£π£+ (π₯π₯ + π£π£ππ π‘π‘) + π£π£β (π₯π₯ β π£π£ππ π‘π‘), respectively. 3 As expected, we find π’π’(π₯π₯, π‘π‘ = 0) = 0, that is, the lattice has no dislocations at π‘π‘ = 0. However, recalling that ππ = π£π£ππ ππ, we will have β π£π£ πππ‘π‘ π’π’(π₯π₯, π‘π‘ = 0) = ½π€π€0 π£π£0 β«ββ exp(βπ€π€0 |ππ|) exp(iππππ) ππππ = 1+(π₯π₯β0π€π€ 2 0) . (7) In other words, the lattice has an initial velocity in a region of width ~πππ€π€0 centered at the origin of coordinates, as originally postulated. The linear momentum πππ₯π₯ of the lattice along the π₯π₯-axis is thus given by πππ₯π₯ (π‘π‘ = 0) = οΏ½ β ββ πππ£π£0 1+(π₯π₯ βπ€π€0 )2 ππππ = πππ€π€0 πππ£π£0 . (8) As time progresses, the initial momentum causes the following lattice deformation: β π’π’(π₯π₯, π‘π‘) = ½(π€π€0 π£π£0 βπ£π£ππ ) β«ββ ππ β1 exp(βπ€π€0 |ππ|) exp(iππππ) sin(πππ£π£ππ π‘π‘) ππππ = ½(π€π€0 π£π£0 βπ£π£ππ ) οΏ½ β exp(βπ€π€0 |ππ|) ββ 2iππ οΏ½exp[iππ(π₯π₯ + π£π£ππ π‘π‘)] β exp[iππ(π₯π₯ β π£π£ππ π‘π‘)]οΏ½ππππ = ½(π€π€0 π£π£0 βπ£π£ππ )οΏ½arctan[(π₯π₯ + π£π£ππ π‘π‘)βπ€π€0 ] β arctan[(π₯π₯ β π£π£ππ π‘π‘)βπ€π€0 ]οΏ½. (9) In other words, the deformation starts at π₯π₯ = 0, then spreads to the right and to the left with constant velocity π£π£ππ . The velocity profile of the lattice at time π‘π‘ is now given by β π£π£(π₯π₯, π‘π‘) = πππ‘π‘ π’π’(π₯π₯, π‘π‘) = ½π€π€0 π£π£0 β«ββ exp(βπ€π€0 |ππ| + iππππ) cos(πππ£π£ππ π‘π‘) ππππ β = ¼π€π€0 π£π£0 οΏ½ = ββ ½π£π£0 exp(βπ€π€0 |ππ|) οΏ½exp[iππ(π₯π₯ + π£π£ππ π‘π‘)] + exp[iππ(π₯π₯ β π£π£ππ π‘π‘)]οΏ½ππππ 1+οΏ½(π₯π₯+π£π£ππ π‘π‘)βπ€π€0 οΏ½ 2 + ½π£π£0 2 1+οΏ½(π₯π₯βπ£π£ππ π‘π‘)βπ€π€0 οΏ½ . (10) Clearly, the initial momentum of the lattice is preserved, although it is now divided between two regions at the leading and trailing edges of the expanding pulse. In between these two edges, the individual atoms/molecules of the lattice are displaced by ½πππ€π€0 π£π£0 βπ£π£ππ along the π₯π₯-axis. 2.1. Energy content of the elastic wave. In the absence of initial displacement at π‘π‘ = 0, the entire energy of the system is initially contained in the kinetic energy of atoms/molecules constituting the medium, that is, β π£π£ β°πΎπΎ (π‘π‘ = 0) = οΏ½ ½ππ οΏ½1+(π₯π₯β0π€π€ ββ ππβ2 0 2 οΏ½ ππππ = ½πππ£π£02 π€π€0 οΏ½ )2 β ββ = ½πππ£π£02 π€π€0 β«βππβ2 cos 2 ππ ππππ = ¼πππππ£π£02 π€π€0 . ππππ = ½πππ£π£02 π€π€0 οΏ½ (1 + π₯π₯ 2 )2 ππβ2 βππβ2 ππππ 1 + tan2 ππ (11) At later times, when π‘π‘ β« 0, the kinetic energy is contained in two well-separated pulses given by the two terms on the right-hand side of Eq.(10). This, however, accounts for only onehalf of the initial energy of the pulse; the remaining half is carried by the potential energy of elastic springs that connect adjacent atoms/molecules, as follows: 4 β°ππ (π‘π‘ β« 0) = β ½ππ β«ββ[πππ₯π₯ π’π’(π₯π₯, π‘π‘)]2 πππ₯π₯ β ¼ππ(π£π£02 οΏ½π£π£ππ2 )π€π€0 οΏ½ β ββ β = ½ππ οΏ½ οΏ½ ππππ (1 + π₯π₯ 2 )2 ββ ½(π£π£0 βπ£π£ππ ) 1+[(π₯π₯+π£π£ππ π‘π‘)βπ€π€0 = β πππππ£π£02 π€π€0 . β ]2 ½(π£π£0 βπ£π£ππ ) 1+[(π₯π₯βπ£π£ππ π‘π‘)βπ€π€0 2 οΏ½ ππππ ]2 (12) It must be obvious from the preceding analysis that, when the leading and trailing edges of the (expanding) pulse are less than fully separated, the kinetic energy would contain a crossterm, namely, the product of the two terms in Eq.(10), which cancels out the cross-term associated with the potential energy contained in the first line of Eq.(12). Example 2. Consider a slab of homogeneous material having width 2πΏπΏ along the π₯π₯-axis, and infinite dimensions along the π¦π¦ and π§π§ axes. The slab boundaries at π₯π₯ = ±πΏπΏ are fixed, so that π’π’(π₯π₯ = ±πΏπΏ, π‘π‘) = 0 at all times π‘π‘. Let π’π’(π₯π₯, π‘π‘ = 0) = 0 and π£π£(π₯π₯, π‘π‘ = 0) = π£π£0 β[1 + (π₯π₯βπ€π€0 )2 ] in the interval βπΏπΏ β€ π₯π₯ β€ πΏπΏ, the implicit assumption here being that π€π€0 βͺ πΏπΏ. In order to satisfy the boundary conditions at π₯π₯ = ±πΏπΏ, the initial distributions π’π’± (π₯π₯, π‘π‘ = 0) and π£π£± (π₯π₯, π‘π‘ = 0) must be expanded in a Fourier series (rather than a Fourier integral). In the present example, where the walls of the slab at π₯π₯ = ±πΏπΏ are rigidly affixed to external mounts, the initial velocity profiles π£π£± (π₯π₯, π‘π‘ = 0) must have the periodic structure depicted in Fig.3(a). As time progresses, π£π£+ (π₯π₯) moves to the left and π£π£β (π₯π₯) moves to the right, both at the constant velocity π£π£ππ . A quick glance at Fig.3(a) confirms that the walls at π₯π₯ = ±πΏπΏ remain motionless at all times. (a) π£π£± (π₯π₯) π₯π₯ (b) π’π’± (π₯π₯) π£π£ππ π£π£ππ 2πΏπΏ π₯π₯ Fig.3. Plots of π£π£± (π₯π₯) and π’π’± (π₯π₯) at π‘π‘ = 0, when the slabβs walls at π₯π₯ = ±πΏπΏ are rigidly affixed to external mounts. The plot of π’π’β (π₯π₯), shown with broken lines, is that of π’π’+ (π₯π₯) flipped around the horizontal axis. As time progresses, π’π’+ (π₯π₯) and π£π£+ (π₯π₯) travel to the left, while π’π’β (π₯π₯) and π£π£β (π₯π₯) move to the right, all at the constant velocity π£π£ππ . Observe that at the boundaries π₯π₯ = ±πΏπΏ, the overall displacement π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + π’π’β (π₯π₯ β π£π£ππ π‘π‘) and velocity π£π£(π₯π₯, π‘π‘) = π£π£+ (π₯π₯ + π£π£ππ π‘π‘) + π£π£β (π₯π₯ β π£π£ππ π‘π‘) remain zero at all times. In the absence of dispersion, one may obtain the displacement function π’π’+ (π₯π₯, π‘π‘ = 0) by integrating the velocity profile π£π£+ (π₯π₯, π‘π‘ = 0) over π₯π₯; the result is shown in Fig.3(b). Clearly, the derivative with respect to π₯π₯ of π’π’+ (π₯π₯, π‘π‘ = 0) is equal to π£π£+ (π₯π₯, π‘π‘ = 0). Considering that the initial displacement is zero, we will have π’π’β (π₯π₯, π‘π‘ = 0) = βπ’π’+ (π₯π₯, π‘π‘ = 0). Now, the periodic function π£π£(π₯π₯, π‘π‘ = 0), may be written 5 π£π£ π£π£(π₯π₯, π‘π‘ = 0) = οΏ½1+(π₯π₯β0π€π€ 0 )2 1 π₯π₯ ππππ οΏ½ β οΏ½2πΏπΏ comb οΏ½2πΏπΏοΏ½ cos οΏ½ 2πΏπΏ οΏ½οΏ½. (13) With the aid of the convolution and multiplication theorems of the Fourier transform theory (see Appendix A), the Fourier series representation of π£π£(π₯π₯, π‘π‘ = 0) is found to be πΏπΏπΏπΏ ππ ππ ππ0 (ππ) = πππ€π€0 π£π£0 exp(βπ€π€0 |ππ|) οΏ½comb οΏ½ ππ οΏ½ β οΏ½½πΏπΏ οΏ½ππ + 2πΏπΏοΏ½ + ½πΏπΏ οΏ½ππ β 2πΏπΏοΏ½οΏ½οΏ½ = (ππ 2 π€π€0 π£π£0 βπΏπΏ) exp(βπ€π€0 |ππ|) ββ ππ=ββ πΏπΏ[ππ β (ππ + ½)ππβπΏπΏ]. Consequently, the Fourier transform of π’π’± (π₯π₯, π‘π‘ = 0) will be given by ππ± (ππ) = ± ππ0 (ππ) 2iπππ£π£ππ =± ππ 2 π€π€0 π£π£0 exp(βπ€π€0 |ππ|) 2iπππππ£π£ππ ββ ππ=ββ πΏπΏ[ππ β (ππ + ½)ππβπΏπΏ]. (14) (15) The spatio-temporal evolution of π’π’(π₯π₯, π‘π‘) and π£π£(π₯π₯, π‘π‘) may now be determined as was done in the case of unbounded media. Of course, for a slab having a finite-width, the spatial and temporal frequencies ππππ = (ππ + ½)ππβπΏπΏ and ππππ are discrete, but they continue to satisfy ππππ = π£π£ππ ππππ , where π£π£ππ = οΏ½ππβππ is the phase velocity of propagation inside the material medium. π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯, π‘π‘) + π’π’β (π₯π₯, π‘π‘) 1 = 2ππ οΏ½ = β ββ πππ€π€0 π£π£0 πΏπΏπ£π£ππ ππ 2 π€π€0 π£π£0 exp(βπ€π€0 |ππ|) ββ ππ=0 πππππ£π£ππ sinοΏ½ππππ π£π£ππ π‘π‘οΏ½ ππππ {ββ ππ=ββ πΏπΏ[ππ β (ππ + ½)ππβπΏπΏ]} exp(iππππ) sin(ππππ) ππππ exp(βπ€π€0 ππππ ) cos(ππππ π₯π₯) ; [ππππ = (ππ + ½)ππβπΏπΏ]. (16) Note that the elastic wave bouncing back and forth within the (finite-width) slab maintains its energy, even though its momentum reverses direction each time the lattice exchanges momentum with the rigid support mounts located at π₯π₯ = ±πΏπΏ. Example 3. Here we examine the finite-width slab of the preceding example in the absence of the constraints on its sidewalls at π₯π₯ = ±πΏπΏ. In other words, the sidewalls are now completely free to move along the π₯π₯-axis. The corresponding plots of π£π£± (π₯π₯, π‘π‘ = 0) and π’π’± (π₯π₯, π‘π‘ = 0) are shown in Fig.4. The lattice retains its initial energy and momentum while moving forward along the π₯π₯axis, albeit in discrete steps. The initial velocity π£π£(π₯π₯, π‘π‘ = 0), a periodic function, is given by π£π£ π£π£(π₯π₯, π‘π‘ = 0) = οΏ½1+(π₯π₯β0π€π€ 2 0) 1 π₯π₯ οΏ½ β οΏ½2πΏπΏ comb οΏ½2πΏπΏοΏ½οΏ½. (17) The Fourier transform of the above function is obtained with the aid of the convolution theorem (see Appendix A), as follows: πΏπΏπΏπΏ ππ0 (ππ) = πππ€π€0 π£π£0 exp(βπ€π€0 |ππ|)comb οΏ½ ππ οΏ½ = (ππ 2 π€π€0 π£π£0 βπΏπΏ) exp(βπ€π€0 |ππ|) ββ ππ=ββ πΏπΏ(ππ β ππππ βπΏπΏ). Consequently, the Fourier transform of π’π’± (π₯π₯, π‘π‘ = 0) will be given by ππ± (ππ) = ± ππ0 (ππ) 2iπππ£π£ππ =± ππ 2 π€π€0 π£π£0 exp(βπ€π€0 |ππ|) 2iπππππ£π£ππ 6 ββ ππ=ββ πΏπΏ(ππ β ππππ βπΏπΏ). (18) (19) π£π£± (π₯π₯) (a) π₯π₯ (b) π’π’± (π₯π₯) π£π£ππ π£π£ππ 2πΏπΏ π₯π₯ Fig.4. Plots of π£π£± (π₯π₯) and π’π’± (π₯π₯) at π‘π‘ = 0, when the slabβs walls at π₯π₯ = ±πΏπΏ are unconstrained. The plot of π’π’β (π₯π₯), shown with broken lines, is similar to that of π’π’+ (π₯π₯), except for being flipped around the horizontal axis. As time progresses, π’π’+ (π₯π₯) and π£π£+ (π₯π₯) move to the left, while π’π’β (π₯π₯) and π£π£β (π₯π₯) move to the right, all at the constant velocity π£π£ππ . Note that at the boundaries π₯π₯ = ±πΏπΏ, the overall displacement π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + π’π’β (π₯π₯ β π£π£ππ π‘π‘) and velocity π£π£(π₯π₯, π‘π‘) = π£π£+ (π₯π₯ + π£π£ππ π‘π‘) + π£π£β (π₯π₯ β π£π£ππ π‘π‘) do not vanish. The spatio-temporal evolution of π’π’(π₯π₯, π‘π‘) and π£π£(π₯π₯, π‘π‘) may now be determined as was done in the case of unbounded media. Needless to say, the spatial and temporal frequencies ππππ = ππππβπΏπΏ and ππππ = π£π£ππ ππππ are discrete. π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯, π‘π‘) + π’π’β (π₯π₯, π‘π‘) 1 = 2ππ οΏ½ β ββ πππ€π€0 π£π£0 =οΏ½ 2πΏπΏ ππ 2 π€π€0 π£π£0 exp(βπ€π€0 |ππ|) οΏ½ π‘π‘ + πππππ£π£ππ πππ€π€0 π£π£0 πΏπΏπ£π£ππ ββ ππ=1 ββ ππ=ββ πΏπΏ(ππ β ππππ βπΏπΏ) exp(iππππ) sin(ππππ) ππππ sinοΏ½ππππ π£π£ππ π‘π‘οΏ½ ππππ exp(βπ€π€0 ππππ ) cos(ππππ π₯π₯) ; (ππππ = ππππβπΏπΏ). (20) The elastic wave bouncing back and forth inside the (finite-width) slab maintains not only its energy but also its linear momentum in the present example, as there are no rigid mounts at π₯π₯ = ±πΏπΏ with which the excited vibrational modes could exchange momentum. Example 4. Figure 5 shows the case of an asymmetrically excited finite-width slab, having one fixed and one free boundary, and an initial velocity profile π£π£(π₯π₯, π‘π‘ = 0). The periodic functions π£π£± (π₯π₯) and π’π’± (π₯π₯), constructed to satisfy the boundary conditions and to allow reflections at the boundaries, have a period of 8πΏπΏ. As expected, the material displacement at each point within the slab changes sign repeatedly and periodically, while the total energy of the system is conserved. Example 5. Figure 6 provides another example of an asymmetrically excited finite-width slab, having an initial velocity profile π£π£(π₯π₯, π‘π‘ = 0). Unlike the preceding example, the boundaries of the slab in the present case are free. The periodic (or quasi-periodic) functions π£π£± (π₯π₯) and π’π’± (π₯π₯), all having the same period, 4πΏπΏ, are now constructed to satisfy the boundary conditions by allowing reflections without a sign change at the floating boundaries. The material displacements 7 at each location within the slab are such that the slab in its entirety moves forward in a step-bystep fashion, while maintaining its total energy and linear momentum. π£π£± (π₯π₯) (a) β3πΏπΏ π£π£ππ βπΏπΏ (b) πΏπΏ 2πΏπΏ π£π£ππ π’π’± (π₯π₯) βπΏπΏ β3πΏπΏ π₯π₯ = 0 πΏπΏ π₯π₯ 3πΏπΏ 3πΏπΏ π£π£ππ π£π£ππ π₯π₯ Fig.5. Plots of π£π£± (π₯π₯) and π’π’± (π₯π₯) at π‘π‘ = 0, when the initial disturbance is asymmetric, the slabβs left wall at π₯π₯ = βπΏπΏ is unconstrained, and the slabβs right wall at π₯π₯ = πΏπΏ is rigidly affixed to an external mount. The plot of π’π’β (π₯π₯), shown with broken lines, is obtained by flipping π’π’+ (π₯π₯) around the π₯π₯-axis. As time progresses, π’π’+ (π₯π₯) and π£π£+ (π₯π₯) move to the left, while π’π’β (π₯π₯) and π£π£β (π₯π₯) move to the right, all at the constant velocity π£π£ππ . Note that the overall displacement π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + π’π’β (π₯π₯ β π£π£ππ π‘π‘) and velocity π£π£(π₯π₯, π‘π‘) = π£π£+ (π₯π₯ + π£π£ππ π‘π‘) + π£π£β (π₯π₯ β π£π£ππ π‘π‘) always vanish at the right-hand boundary. The free boundary on the left, however, oscillates back and forth at regular intervals. π£π£± (π₯π₯) (a) β3πΏπΏ π£π£ππ βπΏπΏ (b) πΏπΏ 2πΏπΏ π£π£ππ π’π’± (π₯π₯) β3πΏπΏ βπΏπΏ π₯π₯ = 0 πΏπΏ π₯π₯ 3πΏπΏ 3πΏπΏ π£π£ππ π£π£ππ π₯π₯ Fig.6. Plots of π£π£± (π₯π₯) and π’π’± (π₯π₯) at π‘π‘ = 0, when the initial disturbance is asymmetric and the slabβs walls at π₯π₯ = ±πΏπΏ are unconstrained. The plot of π’π’β (π₯π₯), shown with broken lines, is obtained by flipping π’π’+ (π₯π₯) around the π₯π₯-axis. As time progresses, π’π’+ (π₯π₯) and π£π£+ (π₯π₯) move to the left, while π’π’β (π₯π₯) and π£π£β (π₯π₯) move to the right, all at the constant velocity π£π£ππ . Note that at the boundaries π₯π₯ = ±πΏπΏ, the overall displacement π’π’(π₯π₯, π‘π‘) = π’π’+ (π₯π₯ + π£π£ππ π‘π‘) + π’π’β (π₯π₯ β π£π£ππ π‘π‘) and velocity π£π£(π₯π₯, π‘π‘) = π£π£+ (π₯π₯ + π£π£ππ π‘π‘) + π£π£β (π₯π₯ β π£π£ππ π‘π‘) do not vanish. 8 2.2. Dispersive media. As discussed in Sec.2.1, the general solution for elastic wave propagation in a uniform, homogeneous, infinite medium may be expressed as follows: 1 β π’π’(π₯π₯, π‘π‘) = 2ππ β«ββ{ππ+ (ππ) exp[i(ππππ + ππππ)] + ππβ (ππ) exp[i(ππππ β ππππ)]}ππππ, 1 β π£π£(π₯π₯, π‘π‘) = 2ππ β«ββ{iππππ+ (ππ) exp[i(ππππ + ππππ)] β iππππβ (ππ) exp[i(ππππ β ππππ)]}ππππ. (21a) (21b) Fourier transforming the initial conditions π’π’(π₯π₯, π‘π‘ = 0) and π£π£(π₯π₯, π‘π‘ = 0) into ππ0 (ππ) and ππ0 (ππ), the functions ππ+ (ππ) and ππβ (ππ) can be determined as follows: ππ+ (ππ) = ½[ππ0 (ππ) β iππ0 (ππ)/ππ], ππβ (ππ) = ½[ππ0 (ππ) + iππ0 (ππ)/ππ]. (22a) (22b) In order to introduce dispersion, one would like to relate ππ to ππ via some arbitrary function, say, ππ(ππ), then continue to use Eqs.(21) and (22) to determine the temporal evolution of the initial conditions. This particular generalization of the solution of the wave equation, which now includes dispersion, does not pose any problems as far as momentum conservation is concerned; β the reason being that the linear momentum of the medium is given by πππ₯π₯ (π‘π‘) = ππ β«ββ π£π£(π₯π₯, π‘π‘)ππππ , which, in accordance with the central value theorem (see Appendix A), is equal to ππ times the Fourier transform of π£π£(π₯π₯, π‘π‘) at ππ = 0. Invoking Eq.(21b), we write πππ₯π₯ (π‘π‘) = iππ(0)ππππ+ (0) exp[iππ(0)π‘π‘] β iππ(0)ππππβ (0) exp[βiππ(0)π‘π‘] = ππππ0 (0) cos[ππ(0)π‘π‘] β ππππ(0)ππ0 (0) sin[ππ(0)π‘π‘]. (23) The momentum of the system will, therefore, be time-independent if ππ(0) happens to be zero, which is generally the case. We conclude that dispersion does not violate the conservation β of momentum. Next, we examine the kinetic energy of the system β°πΎπΎ (π‘π‘) = ½ππ β«ββ π£π£ 2 (π₯π₯, π‘π‘)ππππ. Using Parsevalβs theorem (see Appendix A) in conjunction with Eq.(21b), we find ππ β β°πΎπΎ (π‘π‘) = 4ππ β«ββ|iππππ+ (ππ)exp(iππππ) β iππππβ (ππ) exp(βiππππ)|2 ππππ ππ β = 4ππ β«ββ|ππ0 (ππ) cos(ππππ) β ππ(ππ)ππ0 (ππ)sin(πππ‘π‘)|2 ππππ. (24) β As for the potential energy, we define β°ππ (π‘π‘) = ½ β«ββ[ππ(π₯π₯) β π’π’(π₯π₯, π‘π‘)]2 ππππ, where β signifies the convolution operation, and ππ(π₯π₯) is a real-valued weight function to be determined shortly. Invoking Parsevalβs theorem and denoting the Fourier transform of ππ(π₯π₯) by Ξ£(ππ), we write 1 β β°ππ (π‘π‘) = 4ππ β«ββ|Ξ£(ππ)ππ+ (ππ) exp(iππππ) + Ξ£(ππ)ππβ (ππ) exp(βiππππ)|2 1 β = 4ππ β«ββ|[Ξ£(ππ)βππ(ππ)]ππ0 (ππ) sin(ππππ) + Ξ£(ππ)ππ0 (ππ) cos(ππππ)|2 . (25) Comparison of Eqs.(24) and (25) reveals that the total energy β°πΎπΎ + β°ππ will be timeindependent provided that Ξ£(ππ) = iβππππ(ππ). 9 (26) Since the mass-density ππ and the oscillation frequency ππ(ππ) are real-valued, we conclude that ππ(π₯π₯) must be an odd function of π₯π₯, which in turn requires that ππ(ππ) be an odd function of ππ. In the special case of a dispersionless medium, ππ(ππ) = π£π£ππ ππ yields ππ(π₯π₯) = βπππ£π£ππ πΏπΏ β² (π₯π₯). We have β β β°ππ (π‘π‘) = ½ β«ββ[ππ(π₯π₯) β π’π’(π₯π₯, π‘π‘)]2 ππππ = ½πππ£π£ππ2 β«ββ[πππ₯π₯ π’π’(π₯π₯, π‘π‘)]2 ππππ . (27) Thus we have recovered the result for the simple case studied in Sec.1, where πππ£π£ππ2 is the spring constant ππ of the simple (i.e., dispersionless) material. 3. Transverse vibrations of a one-dimensional slab of homogeneous and uniform material. A one-dimensional string stretched along the π₯π₯-axis and vibrating along the transverse direction π¦π¦, provides a good model for the transverse vibrations of a homogeneous, uniform slab of material. The diagram in Fig.7 helps to explain that the equation of motion along the π¦π¦-axis is identical with that for longitudinal vibrations given by Eq.(1), provided that the tensile stress ππ within the string is substituted for the spring constant ππ in Eq.(1). This is because, assuming the vibration amplitude is not too great, the slope of the displacement curve πππ₯π₯ π’π’(π₯π₯, π‘π‘) = tan ππ may be approximated with sin ππ. The vertical component of the tensile force acting within the wire at location π₯π₯ and time π‘π‘ is then written as πππ¦π¦ (π₯π₯, π‘π‘) = ±ππ sin ππ β ±ππ πππ₯π₯ π’π’(π₯π₯, π‘π‘). The tensile stress ππ thus plays the same role in transverse vibrations of a string as the stiffness coefficient ππ does in longitudinal vibrations. The phase velocity of propagation along the string is then π£π£ππ = οΏ½ππβππ . y ΞΈ u(x,t) Ο Ο x xββx x x+βx Fig.7. Transverse vibrations of a one-dimensional wire of uniform mass-density ππ subject to a constant tension ππ. The amplitude of the vibrations is small enough to allow the approximation sin ππ β πππ₯π₯ π’π’(π₯π₯, π‘π‘). 3.1. Potential energy of the string. Suppose a short segment βπ₯π₯ of the string is stretched and expanded along the π¦π¦-axis, so that its length has become οΏ½(βπ₯π₯)2 + (βπ¦π¦)2 β βπ₯π₯ + ½ (βπ¦π¦)2ββπ₯π₯. Considering that the elongation by ββ β ½ (βπ¦π¦)2ββπ₯π₯ = ½[πππ₯π₯ π’π’(π₯π₯, π‘π‘)]2 βπ₯π₯ takes place under the constant tension ππ, the potential energy per unit-length stored in the string must be given by β°ππ (π₯π₯, π‘π‘) = ½ππ[πππ₯π₯ π’π’(π₯π₯, π‘π‘)]2 . Once again, it is seen that the tensile stress ππ plays the same role in transverse vibrations of the string as the spring constant ππ does in longitudinal vibrations. 3.2. Similarities and differences between longitudinal and transverse vibrations. For an infinitely long string aligned with the π₯π₯-axis and under a constant tensile stress ππ, transverse vibrations in the π¦π¦ direction behave similarly to longitudinal vibrations along π₯π₯. (A short pulse of light passing through can deposit its momentum and set off the vibrations.) All the ππ-vectors will be along the π₯π₯-axis, but of course the overall momentum of the string will be directed along π¦π¦. Also, for a finite-length string with fixed end-points, the transverse behavior is similar to the longitudinal behavior. The energy of the string remains conserved, but its momentum continually gets exchanged with the fixed support structures to which the end-points are attached. 10 In the case of a finite-length string with free boundaries, one must create a contraption, such as that shown in Fig.8, in order to maintain a tensile stress within the string as the string moves away from its initial location. Also, if the initial velocity profile happens to be asymmetric relative to the center of the string, conservation of angular momentum will add a spinning motion to an (untethered) string, which would complicate the analysis. The contraption of Fig.8 allows periodic exchanges of angular momentum between the frame and the string, in such a way as to maintain the total angular momentum of the system while preventing the spinning of the string. Fig.8. A string of length πΏπΏ and negligible diameter is kept under a constant tensile stress ππ by being mounted, tightly and horizontally, between the side-walls of a rectangular frame. The sliding mounts on both ends of the string are frictionless, leaving the string free to move up or down as the need arises. An initial upward kick, described by the initial conditions π’π’(π₯π₯, π‘π‘ = 0) and π£π£(π₯π₯, π‘π‘ = 0), is imparted to the string, starting it on an upward journey. The mounts, being essentially massless and frictionless, maintain the tensile stress ππ along the length of the string at all times, while allowing the end-points to behave as if they were free to move in the vertical direction. No torque acts on the frame whenever the end-points of the string happen to be at the same height. However, at those instants when one end of the string rises above the other, the tensile force within the string exerts a torque that tends to rotate the frame β albeit ever so slightly, because of the frameβs large moment of inertia. The string thus maintains its energy and linear momentum, while the system as a whole maintains its angular momentum by frequent exchanges between the string and the frame. L π’π’(π₯π₯, π‘π‘) x In accordance with Eq.(9), when a localized displacement of the string of Fig.8 reaches a free end-point and reverses direction, the end-point rises vertically by βπ¦π¦ = πππ€π€0 π£π£0 βπ£π£ππ . (A factor of 2 has been incorporated into this result because the height at the end-point doubles upon a reversal in the waveβs propagation direction.) If one end of the string rises before the other end, the torque acting on the frame will be ππβπ¦π¦ = πππ€π€0 πππ£π£0 π£π£ππ . This torque precisely accounts for the time rate-of-change of angular momentum of the vibrational motion of the string during the time when the leading and trailing edges of a deformation travel in the same direction β see Eq.(8), which gives the momentum of the entire spring along the π¦π¦-axis, namely, πππ¦π¦ = πππ€π€0 πππ£π£0 . When the leading and trailing edges travel in the same direction with velocity π£π£ππ , the time rate-ofchange of angular momentum, πππ¦π¦ π£π£ππ , is precisely cancelled out by the torque acting on the frame. 4. Solid object experiencing a constant external force. Returning to longitudinal motion and οΏ½ be applied on the one-dimensional excitation along the π₯π₯-axis, let a constant, uniform force πΉπΉ0 ππ left-hand side of a solid rod of length πΏπΏ, linear mass-density ππ, and stiffness coefficient ππ, as depicted in Fig.9. The situation is akin to a front-surface mirror subject to radiation pressure from a continuous wave (cw) light beam impinging from the left-hand-side. For a perfectly electrically conducting mirror, the light gets fully reflected at the front facet, which is the only place on which the external force acts. In the steady state, the rod will be slightly compressed β in order to produce the necessary internal forces along its length β and moves as a whole with οΏ½. The displacement function may thus be written as follows: constant acceleration ππ = (πΉπΉ0 /ππππ)ππ πΉπΉ π’π’(π₯π₯, π‘π‘) = 2πΏπΏ0 οΏ½ (π₯π₯ β πΏπΏ)2 ππ π‘π‘ 2 + ππ οΏ½ ; 11 0 β€ π₯π₯ β€ πΏπΏ, π‘π‘ β« 0. (28) Note that the above π’π’(π₯π₯, π‘π‘) satisfies the wave equation = πππππ₯π₯2 π’π’(π₯π₯, π‘π‘) throughout the rod at all times π‘π‘. Moreover, the rodβs internal stress πππππ₯π₯ π’π’(π₯π₯, π‘π‘) = (πΉπΉ0 βπΏπΏ)(π₯π₯ β πΏπΏ) vanishes at π₯π₯ = πΏπΏ, but equals βπΉπΉ0 at π₯π₯ = 0, as it must. πππππ‘π‘2 π’π’(π₯π₯, π‘π‘) Fig.9. Uniform rod of length πΏπΏ, mass-density ππ, and stiffness coefficient ππ οΏ½ on its left-hand facet. In the steady subject to a constant external force πΉπΉ0 ππ state, when all transient internal vibrations have settled down, the rod will be uniformly accelerating while slightly compressed, as dictated by the quadratic displacement function π’π’(π₯π₯, π‘π‘) depicted here. x π’π’(π₯π₯, π‘π‘0 ) πΉπΉ0 π‘π‘02 β(2πΏπΏπΏπΏ) 0 x πΏπΏ οΏ½ acting Next, let us examine the case of an external force-density ππ(π₯π₯) = ππ0 sin(2ππππ βπΏπΏ) ππ πΏπΏ within the body of a rod of length πΏπΏ. Since the total force ππ = β«0 ππ(π₯π₯)ππππ is zero in this case, the rod will not move as a whole, but it will develop internal stresses to cancel out the external force. This is a model for a transparent dielectric slab of refractive index ππ and thickness πΏπΏ, illuminated οΏ½ and at normal incidence by a linearly-polarized, monochromatic plane-wave of amplitude πΈπΈ0 ππ wavelength ππ0 ; see Appendix B for details. If the slab thickness happens to be πΏπΏ = ππ0 β2ππ, the EM force distribution within the slab will be ππ(π₯π₯) = β ππππ0 πΈπΈ02 (ππ2 β1)2 2ππππ0 4ππππππ sin οΏ½ ππ0 οΏ½. οΏ½ ππ (29) In the above equation, ππ0 is the permittivity of free space. The displacement function is now obtained as the second integral of the external force-density, namely, ππ πΏπΏ πΏπΏ 2ππππ 0 π’π’(π₯π₯, π‘π‘) = 2ππππ οΏ½2ππ sin οΏ½ πΏπΏ οΏ½ β π₯π₯οΏ½. (30) Note that the integration constants have been chosen to ensure that the internal stress πππππ₯π₯ π’π’(π₯π₯, π‘π‘) vanishes at both boundaries located at π₯π₯ = 0 and π₯π₯ = πΏπΏ. The plot in Fig.10 shows the displacement function of Eq.(30) for the case of a transparent dielectric slab under cw illumination, where ππ0 < 0; see Eq.(29). Fig.10. A dielectric slab of refractive index ππ and thickness πΏπΏ = ππ0 β2ππ is illuminated from the left-hand-side by a plane-wave of wavelength ππ0 . The radiation force acting on the dielectric medium is given by Eq.(29), and the resulting material displacement is given by Eq.(30). Since the integrated force of radiation on the slab is zero, the function π’π’(π₯π₯, π‘π‘) is time-independent. Note that the curvature of the left half of the displacement function is positive, while that of the right half is negative. The internal stresses of the slab thus precisely cancel out the external force exerted by the EM field. πππ₯π₯ (π₯π₯) ππ0 πΏπΏ2 β 2ππππ π’π’(π₯π₯, π‘π‘0 ) 0 πΏπΏ x Our next example pertains once again to a solid, uniform rod of length πΏπΏ, but now subject to πΏπΏ οΏ½. The total force ππ = β«0 ππ(π₯π₯)ππππ = (2ππ0 πΏπΏβππ)ππ οΏ½ the external force-density ππ(π₯π₯) = ππ0 sin(ππππ βπΏπΏ) ππ being nonzero in the present example, the rod as a whole will acquire an acceleration ππ = οΏ½β(ππππ), but it will also develop internal stresses to cancel out the residual external force. 2ππ0 ππ This is a model for a transparent dielectric slab of refractive index ππ and thickness πΏπΏ = ππ0 β4ππ, illuminated at normal incidence by a linearly polarized, monochromatic plane-wave of amplitude οΏ½ and wavelength ππ0 , for which the EM force-density, as shown in Appendix B, is given by πΈπΈ0 ππ 12 ππ(π₯π₯) = 2ππππππ0 πΈπΈ02 ππ2 β1 ππ0 2 4ππππππ οΏ½ππ2 +1οΏ½ sin οΏ½ ππ0 οΏ½. οΏ½ ππ (31) The displacement function is now obtained as the second integral of the residual external force-density, that is, π’π’(π₯π₯, π‘π‘) = ππ0 πΏπΏ πΏπΏ ππππ π₯π₯ ππ οΏ½ sin οΏ½ πΏπΏ οΏ½ + πΏπΏ (π₯π₯ β πΏπΏ)οΏ½ + οΏ½ππππ0 οΏ½ π‘π‘ 2 , ππππ ππ (32) where the integration constants have been chosen to ensure that the internal stress πππππ₯π₯ π’π’(π₯π₯, π‘π‘) vanishes at both ends of the rod located at π₯π₯ = 0 and π₯π₯ = πΏπΏ, while the average force-density πΏπΏ (1βπΏπΏ) β«0 πππππ₯π₯2 π’π’(π₯π₯, π‘π‘)ππππ survives. The plot of π’π’(π₯π₯, π‘π‘) of Eq.(32) shown in Fig.11 corresponds to a quarter-wave-thick dielectric slab under cw illumination, for which ππ0 > 0; see Eq.(31). Fig.11. Dielectric slab of refractive index ππ and thickness πΏπΏ = ππ0 β4ππ, illuminated from the left-hand-side by a plane-wave of wavelength ππ0 . The radiation force on the dielectric is given by Eq.(31), and the resulting material displacement is given by Eq.(32). The integrated force of radiation being positive, the slab moves forward with a net acceleration. Note that the displacement functionβs curvature is negative in the mid-section and positive near the entrance and exit facets of the slab. The internal stresses of the slab thus precisely cancel out the residual external force exerted by the EM field. πππ₯π₯ (π₯π₯) π’π’(π₯π₯, π‘π‘0 ) )2 (1 β ππβ4)(πΏπΏβππ ππ0 βππ 0 ππ0 π‘π‘02 βππππ πΏπΏ x Finally, we consider the case of an antireflection coating layer atop a semi-infinite substrate of refractive index ππ. The layer has index of refraction βππ and thickness πΏπΏ = ππ0 β(4βππ), where οΏ½. In accordance with ππ0 is the wavelength of the normally-incident plane-wave of amplitude πΈπΈ0 ππ the analysis in Appendix B, the EM force-density acting within the coating layer is ππ(π₯π₯) = β ππ(ππβ1)2 ππ0 πΈπΈ02 2βππππ0 4ππβπππ₯π₯ sin οΏ½ ππ0 οΏ½. οΏ½ ππ (33) οΏ½, acting inside a oneWe thus consider the external force-density ππ(π₯π₯) = ππ0 sin(ππππ βπΏπΏ) ππ dimensional rod of length πΏπΏ, as shown in Fig.12. The second integral of this force-density yields the displacement function of the rod, as follows: π’π’(π₯π₯, π‘π‘) = ππ0 πΏπΏ πΏπΏ οΏ½ sin(ππππ βπΏπΏ) + (πΏπΏ β π₯π₯)οΏ½. (34) ππππ ππ Fig.12. Antireflection coating layer of refractive index βππ and thickness πΏπΏ = ππ0 β(4βππ) ensures that an incident plane-wave of wavelength ππ0 enters the substrate with no reflection losses at the entrance facet. The EM force density inside the layer is given by Eq.(33), and the corresponding displacement function appears in Eq.(34). The internal force-density cancels out the EM force-density throughout the coating layer, while the slope of π’π’(π₯π₯, π‘π‘) at π₯π₯ = πΏπΏ οΏ½ pulls at guarantees that the total EM force ππ = (2ππ0 πΏπΏβππ)ππ the interface between the coating layer and the substrate. 13 Substrate (Index = n) πππ₯π₯ (π₯π₯) π’π’(π₯π₯, π‘π‘0 ) ππ0 πΏπΏ2 βππππ π₯π₯ = 0 πΏπΏ x Note that the integration constants have been chosen to ensure that the internal force completely cancels the external force, that the internal stress πππππ₯π₯ π’π’(π₯π₯, π‘π‘) at π₯π₯ = 0 vanishes, that the internal stress at π₯π₯ = πΏπΏ pulls at the substrate interface with a force of 2ππ0 πΏπΏβππ (this is a pull force when ππ0 < 0), and that the displacement at π₯π₯ = πΏπΏ is equal to zero, since the substrate has been taken to be immobile. 5. Equation of motion for one-dimensional rod under internal stress. So far, we have examined fairly simple situations involving either longitudinal or transverse motion in one dimension associated with thin solid rods. In the present section we analyze more complex mechanical vibrations and deformations of a one-dimensional medium that is free to move in three dimensional space. We will find that, although the resulting equation of motion does not admit simple analytical solutions, it is nonetheless possible to verify that its solutions comply with the conservation laws of energy and linear as well as angular momentum. Consider a thin, solid rod of length πΏπΏ, mass-density ππ(ππ), and stiffness coefficient ππ(ππ), as shown in Fig.13. Here ππ is the distance from the left-end of the rod when the rod is at rest. The rod is initially at rest on the π¦π¦-axis, occupying the interval [0, πΏπΏ]. Denoting the position of the point ππ of the rod at time π‘π‘ with the vector function ππ(ππ, π‘π‘), the equation of motion of the rod may be written as follows: ππππ ππ ππππ οΏ½ππ(ππ) οΏ½ππππ ππ β |ππ οΏ½οΏ½ = ππ(ππ)πππ‘π‘2 ππ(ππ, π‘π‘). (35) ππ ππ| The mass-density ππ(ππ) and the stiffness coefficient ππ(ππ) of the rod are positive everywhere. Note that ππππ ππ(ππ, π‘π‘) represents both the normalized length and orientation at time π‘π‘ of the infinitesimal length of the spring initially located at ππ. From this, the unit-vector ππππ ππβ|ππππ ππ| has been subtracted in Eq.(35) to yield the normalized elongation of the local spring at time π‘π‘. Multiplication by ππ(ππ) yields the tensile force within the spring, and the final differentiation with respect to ππ results in the net force (per unit length) acting on the local mass-density ππ(ππ). z ππ(ππ, π‘π‘) x ππ = 0 y ππ = πΏπΏ Fig.13. When π‘π‘ < 0, the one-dimensional rod of length πΏπΏ and negligible thickness rests on the π¦π¦-axis, overlaying the interval [0, πΏπΏ]. Each point on the rod is uniquely identified by its distance ππ from the leftend of the rod in its rest state. The rodβs mass-density and stiffness coefficient are denoted by ππ(ππ) and ππ(ππ), respectively. At times π‘π‘ β₯ 0, the location of the point ππ on the rod in three-dimensional Euclidean space is specified by the vector ππ(ππ, π‘π‘); the rodβs instantaneous velocity profile is thus ππ(ππ, π‘π‘) = πππ‘π‘ ππ(ππ, π‘π‘). The initial conditions are specified as ππ(ππ, π‘π‘ = 0) and ππ(ππ, π‘π‘ = 0). 14 The overall linear momentum of the rod at time π‘π‘ may now be evaluated as follows: π‘π‘ πΏπΏ ππ(π‘π‘) β ππ(0) = β«π‘π‘ β² =0 β«ππ=0 ππ(ππ)πππ‘π‘2 ππ(ππ, π‘π‘ β² )πππππππ‘π‘ β² π‘π‘ πΏπΏ ππππ ππ =οΏ½ οΏ½ =οΏ½ οΏ½ππ(πΏπΏ) οΏ½ππππ ππ β |ππ π‘π‘ β² =0 π‘π‘ π‘π‘ β² =0 ππ=0 ππππ οΏ½ππ(ππ) οΏ½ππππ ππ β |ππ ππππ ππ οΏ½οΏ½ ππ ππ| οΏ½οΏ½ πππππππ‘π‘ β² ππ ππ| ππ=πΏπΏ ππππ ππ β ππ(0) οΏ½ππππ ππ β |ππ οΏ½οΏ½ ππ ππ| ππ=0 οΏ½ πππ‘π‘ β² . (36) Conservation of linear momentum thus demands that |ππππ ππ(0, π‘π‘)| = |ππππ ππ(πΏπΏ, π‘π‘)| = 1.0 at all times. These are the boundary conditions for a rod whose end-points are free to move about. As for the conservation of angular momentum for an unconstrained rod, it suffices to show that the overall torque acting on the rod remains zero at all times. We will have πΏπΏ ππππ ππ π»π»(π‘π‘) = οΏ½ ππ(ππ, π‘π‘) × ππππ οΏ½ππ(ππ) οΏ½ππππ ππ β |ππ Integration by parts 0 ππππ ππ = ππ(πΏπΏ, π‘π‘) × ππ(πΏπΏ) οΏ½ππππ ππ β |ππ πΏπΏ οΏ½οΏ½ ππ ππ| οΏ½οΏ½ ππππ ππ ππ| ππ=πΏπΏ ππππ ππ β ππ(0, π‘π‘) × ππ(0) οΏ½ππππ ππ β |ππ β οΏ½ ππππ ππ(ππ, π‘π‘) × ππ(ππ)οΏ½1 β |ππππ ππ(ππ, π‘π‘)|β1 οΏ½ππππ ππ(ππ, π‘π‘)ππππ. 0 οΏ½οΏ½ ππ ππ| ππ=0 (37) When the end-points of the rod are free to move, we will have |ππππ ππ(0, π‘π‘)| = |ππππ ππ(πΏπΏ, π‘π‘)| = 1.0, which causes the first two terms on the right-hand side of Eq.(37) to vanish. The remaining term also vanishes because the cross-product of ππππ ππ(ππ, π‘π‘) into itself is equal to zero. The conservation of angular momentum for a free-floating rod is thus confirmed. Next, we consider the energy of the rod as a function of time, and demonstrate that the sum of its kinetic and potential energies remain constant regardless of whether the rod is free-floating, fixed at one end-point, or fixed at both end-points. The kinetic and potential energies of the rod at time π‘π‘ are written as follows: πΏπΏ πΏπΏ β°πΎπΎ (π‘π‘) = ½ β«0 ππ(ππ)π£π£ 2 (ππ, π‘π‘)ππππ = ½ β«0 ππ(ππ)πππ‘π‘ ππ(ππ, π‘π‘) β πππ‘π‘ ππ(ππ, π‘π‘)ππππ. πΏπΏ ππππ ππ β°ππ (π‘π‘) = ½ οΏ½ ππ(ππ) οΏ½ππππ ππ β |ππ 0 ππ ππππ ππ οΏ½ β οΏ½ππππ ππ β |ππ ππ| οΏ½ ππππ. ππ ππ| (38) (39) Taking the time-derivative of the kinetic energy β°πΎπΎ (π‘π‘) given by Eq.(38) and invoking the equation of motion, Eq.(35), we find ππ ππππ πΏπΏ β°πΎπΎ (π‘π‘) = β«0 ππ(ππ)πππ‘π‘2 ππ(ππ, π‘π‘) β πππ‘π‘ ππ(ππ, π‘π‘)ππππ πΏπΏ ππππ ππ = οΏ½ ππππ οΏ½ππ(ππ) οΏ½ππππ ππ β |ππ 0 οΏ½οΏ½ β πππ‘π‘ ππ(ππ, π‘π‘)ππππ ππ ππ| 15 ππππ ππ = ππ(πΏπΏ) οΏ½ππππ ππ β |ππ Integration by parts πΏπΏ οΏ½οΏ½ ππ ππ| ππ=πΏπΏ ππππ ππ β οΏ½ ππ(ππ) οΏ½ππππ ππ β |ππ 0 ππ ππππ ππ β πππ‘π‘ ππ(πΏπΏ, π‘π‘) β ππ(0) οΏ½ππππ ππ β |ππ οΏ½β ππ| ππ2 ππ(ππ,π‘π‘) ππππππππ ππππ. οΏ½οΏ½ ππ ππ| ππ=0 β πππ‘π‘ ππ(0, π‘π‘) (40) The first two terms on the right-hand-side of Eq.(40) vanish, whether the end-points of the rod are free or fully constrained (i.e., fixed). The remaining term is cancelled by the time-rate-ofchange of the potential energy β°ππ (π‘π‘), which is evaluated as follows: πΏπΏ ππ ππππ ππ β° (π‘π‘) = οΏ½ ππ(ππ) οΏ½ππππ ππ β |ππ ππππ ππ 0 πΏπΏ ππ ππππ ππ = οΏ½ ππ(ππ) οΏ½ππππ ππ β |ππ 0 ππππ ππ οΏ½ β πππ‘π‘ οΏ½ππππ ππ β |ππ ππ| ππ οΏ½β ππ| ππ2 ππ(ππ,π‘π‘) ππππππππ ππππ. οΏ½ ππππ ππ ππ| (41) The reason for removing the term πππ‘π‘ (ππππ ππβ|ππππ ππ|) from Eq.(41) is that the unit-vector ππππ ππβ|ππππ ππ| can only change in a direction perpendicular to itself, thus making πππ‘π‘ (ππππ ππβ|ππππ ππ|) orthogonal to (1 β |ππππ ππ|β1 )ππππ ππ(ππ, π‘π‘), causing the dot-product of the two vectors to vanish. We conclude that, irrespective of whether the end-points of the rod are free or fixed, its total energy is conserved. 6. Elastic wave propagation in the transient regime. Having observed the complications that could arise when elastic motion is unconstrained, we return to simpler problems involving longitudinal displacements and deformations in one dimension. This time, however, we change the focus of our attention to transient effects, which cannot be handled by simply introducing static deformations. The examples in the following subsections show how deformations initiate and then propagate through homogeneous semi-infinite rods in one-dimensional space. 6.1. Semi-infinite rod under constant external pressure. Shown in Fig.14 is a straight, uniform, homogeneous, semi-infinite rod aligned with the π₯π₯-axis. The rod has cross-sectional area π΄π΄, mass-density ππ, stiffness coefficient ππ, and phase velocity π£π£ππ = οΏ½ππβππ. Starting at π‘π‘ = 0, a uniform pressure ππ0 , acting on the left-hand facet of the rod, pushes the rod forward along the π₯π₯-axis. The displacement of the rod π’π’(π₯π₯, π‘π‘) is given by the following function, which satisfies the wave equation πππππ₯π₯2 π’π’(π₯π₯, π‘π‘) = πππππ‘π‘2 π’π’(π₯π₯, π‘π‘), as it has the standard form of π’π’(π₯π₯ β π£π£ππ π‘π‘). β(π£π£0 βπ£π£ππ )(π₯π₯ β π£π£ππ π‘π‘); π’π’(π₯π₯, π‘π‘) = οΏ½πΌπΌ sin2 [π½π½(π₯π₯ β π£π£ππ π‘π‘)] ; 0; 0 β€ π₯π₯ β€ π£π£ππ π‘π‘ β β, π£π£ππ π‘π‘ β β β€ π₯π₯ β€ π£π£ππ π‘π‘, π₯π₯ β₯ π£π£ππ π‘π‘. (42) In Eq.(42), β is the width of the leading edge of the wavefront, which, together with the shape of the leading edge, is determined by the way in which the external pressure rises from zero to ππ0 in the vicinity of π‘π‘ = 0. The functional form of the leading edge is chosen rather arbitrarily here, with the parameters πΌπΌ and π½π½ intended to be adjusted to ensure a smooth transition from the linear segment of π’π’(π₯π₯, π‘π‘) to the nonlinear segment representing the leading edge. The velocity π£π£(π₯π₯, π‘π‘) = πππ‘π‘ π’π’(π₯π₯, π‘π‘) is seen to be constant at π£π£ = π£π£0 in the linear segment between π₯π₯ = 0 and π₯π₯ = π£π£ππ π‘π‘ β β. The internal force acting on the left-hand facet of the rod, that 16 ππ0 π₯π₯ π’π’(π₯π₯, π‘π‘0 ) β π₯π₯ = 0 π₯π₯ = π£π£ππ π‘π‘0 π₯π₯ Fig.14. Starting at π‘π‘ = 0, a uniform, semi-infinite rod, having cross-sectional area π΄π΄, mass-sensity ππ, and stiffness coefficient ππ, is subjected to a constant pressure ππ0 on its left-hand facet. In the absence of dispersion, the pressure wave inside the rod propagates at the constant velocity π£π£ππ , producing the displacement π’π’(π₯π₯, π‘π‘) along the length of the rod at times π‘π‘ β₯ 0. The displacement is a linear function of π₯π₯, except in the short interval β immediately before the leading edge of the wavefront, where it has the nonlinear form given in Eq.(42). The nonlinearity is rooted in the fact that the pressure acting on the lefthand facet must rise from zero to ππ0 in a short time interval in the vicinity of π‘π‘ = 0. At any given time, say, π‘π‘ = π‘π‘0 , the displacement is largest at π₯π₯ = 0 and drops off linearly with increasing π₯π₯ up to the point π₯π₯ = π£π£ππ π‘π‘0 β β, where it merges with the nonlinear section representing the leading edge of the wavefront, and proceeds to vanish smoothly at π₯π₯ = π£π£ππ π‘π‘0 . is, πππππ₯π₯ π’π’(π₯π₯ = 0, π‘π‘) = βπππ£π£0 /π£π£ππ , must be cancelled out by the external push force πΉπΉπ₯π₯ = π΄π΄π΄π΄0 . Therefore, π£π£0 = (π΄π΄ππ0 βππ)π£π£ππ is the rodβs velocity in the linear region 0 β€ π₯π₯ β€ π£π£ππ π‘π‘ β β. At π₯π₯ = π£π£ππ π‘π‘ β β both π’π’(π₯π₯, π‘π‘) and πππ₯π₯ π’π’(π₯π₯, π‘π‘) must be continuous, that is, οΏ½ πΌπΌ sin2 (π½π½β) = βπ£π£0 βπ£π£ππ 2πΌπΌπΌπΌ sin(π½π½β) cos(π½π½β) = π£π£0 βπ£π£ππ β οΏ½ π½π½β β 1.1655 β οΏ½ (43) πΌπΌ β 1.184(π΄π΄ππ0 ββππ). πΌπΌπ½π½ sin(2π½π½β) = π΄π΄ππ0 βππ tan(π½π½β) = 2π½π½β Thus, given the width β of the leading edge, the parameters πΌπΌ and π½π½ are readily obtained from Eq.(43). The linear momentum of the rod at time π‘π‘ may now be calculated as follows: β π£π£ π‘π‘ πππ₯π₯ (π‘π‘) = β«0 ππππ(π₯π₯, π‘π‘)ππππ = β«0 ππ πππππ‘π‘ π’π’(π₯π₯, π‘π‘)ππππ β = πππ£π£0 (π£π£ππ π‘π‘ β β) + πππππππ£π£ππ β«0 sin(2π½π½π½π½) ππππ = πππ£π£0 (π£π£ππ π‘π‘ β β) + πππππ£π£ππ sin2 (π½π½β) = πππ£π£0 (π£π£ππ π‘π‘ β β) + πππ£π£0 β = πππ£π£0 π£π£ππ π‘π‘ = π΄π΄ππ0 π‘π‘. (44) The above result may appear surprising at first, considering that the initial pressure during the period 0 β€ π‘π‘ β€ ββπ£π£ππ has not been equal to ππ0 . Note, however, that the pressure history, which is inherent to the displacement profile of Eq.(42), is given by 0; ππ(π‘π‘) = β(ππβπ΄π΄)πππ₯π₯ π’π’(π₯π₯ = 0, π‘π‘) = οΏ½(πΌπΌπΌπΌπΌπΌβπ΄π΄) sin(2π½π½π£π£ππ π‘π‘) ; πππ£π£0 βπ΄π΄π£π£ππ ; π‘π‘ β€ 0, 0 β€ π‘π‘ β€ ββπ£π£ππ , π‘π‘ β₯ ββπ£π£ππ . (45) Considering that 2π½π½β β 2.331 > ½ππ, it is clear that ππ(π‘π‘) rises above ππ0 during the initial interval 0 β€ π‘π‘ β€ ββπ£π£ππ before settling down at ππ0 . In general, any pressure function ππ(π‘π‘) produces its own displacement function π’π’(π₯π₯ β π£π£ππ π‘π‘) via a simple integral, and the mechanical momentum of the rod can easily be shown to satisfy the following conservation law: 17 β π£π£ π‘π‘ π£π£ π‘π‘ πππ₯π₯ (π‘π‘) = β«0 ππππ(π₯π₯, π‘π‘)ππππ = β«0 ππ πππππ‘π‘ π’π’(π₯π₯ β π£π£ππ π‘π‘)ππππ = β β«0 ππ πππ£π£ππ πππ₯π₯ π’π’(π₯π₯ β π£π£ππ π‘π‘)ππππ π‘π‘ π‘π‘ = β β«0 πππ£π£ππ2 πππ₯π₯ π’π’(π£π£ππ π‘π‘ β³ β π£π£ππ π‘π‘)πππ‘π‘ β³ = β β«0 πππππ₯π₯ π’π’(βπ£π£ππ π‘π‘ β² )πππ‘π‘ β² π‘π‘ π‘π‘ = β β«0 πππππ₯π₯ π’π’(π₯π₯ = 0, π‘π‘ β² )πππ‘π‘ β² = π΄π΄ β«0 ππ(π‘π‘ β² )πππ‘π‘ β² . (46) To appreciate the generality of the above result, we give another example of the displacement function constructed from a simple ππ(π‘π‘), which rises linearly with time during the initial interval 0 β€ π‘π‘ β€ ββπ£π£ππ , before settling at the constant pressure ππ0 . We will have πΌπΌ β (π£π£0 βπ£π£ππ )(π₯π₯ β π£π£ππ π‘π‘); 0 β€ π₯π₯ β€ π£π£ππ π‘π‘ β β, π’π’(π₯π₯, π‘π‘) = οΏ½ π½π½(π₯π₯ β π£π£ππ π‘π‘)2 ; π£π£ππ π‘π‘ β β β€ π₯π₯ β€ π£π£ππ π‘π‘, 0; The continuity of π’π’(π₯π₯, π‘π‘) and πππ₯π₯ π’π’(π₯π₯, π‘π‘) at π₯π₯ = π£π£ππ π‘π‘ β β now yields πΌπΌ + (π£π£0 βπ£π£ππ )β = π½π½β2 οΏ½ π£π£0 βπ£π£ππ = 2π½π½β (47) π₯π₯ β₯ π£π£ππ π‘π‘. πΌπΌ = β½(π£π£0 βπ£π£ππ )β οΏ½ π½π½β = ½(π£π£0 βπ£π£ππ ). β (48) The linear momentum of the rod at time π‘π‘ is now easily calculated, as follows: β π£π£ π‘π‘ πππ₯π₯ (π‘π‘) = β«0 ππππ(π₯π₯, π‘π‘)πππ₯π₯ = β«0 ππ πππππ‘π‘ π’π’(π₯π₯, π‘π‘)ππππ β = πππ£π£0 (π£π£ππ π‘π‘ β β) + 2πππ£π£ππ π½π½ β«0 π₯π₯π₯π₯π₯π₯ = πππ£π£0 (π£π£ππ π‘π‘ β β) + πππ£π£ππ π½π½β2 = πππ£π£0 (π£π£ππ π‘π‘ β β) + ½πππ£π£0 β = π΄π΄ππ0 [π‘π‘ β ββ(2π£π£ππ )]. (49) Finally, we demonstrate that the energy of the rod is, in general, equally split between kinetic and potential energies. This is done by invoking the general formulas for the kinetic and potential energies, namely, β°πΎπΎ = οΏ½ π£π£ππ π‘π‘ 0 =οΏ½ π£π£ππ π‘π‘ 0 2 ½πποΏ½πππ‘π‘ π’π’(π₯π₯ β π£π£ππ π‘π‘)οΏ½ ππππ = οΏ½ π£π£ππ π‘π‘ 0 2 ½πποΏ½πππ₯π₯ π’π’(π₯π₯ β π£π£ππ π‘π‘)οΏ½ ππππ = β°ππ . 2 ½πππ£π£ππ2 οΏ½πππ₯π₯ π’π’(π₯π₯ β π£π£ππ π‘π‘)οΏ½ ππππ (50) 6.2. Transparent semi-infinite medium illuminated by a light pulse. Inside the transparent rod shown in Fig.15, the light beamβs leading edge propagates at the speed of light π£π£ππ . The pressure of the light on the medium is confined to the vicinity of the leading edge, where the displacement function π’π’(π₯π₯, π‘π‘) has a non-zero curvature. Since no radiation pressure is exerted at the entrance facet of the medium, the slope of π’π’(π₯π₯, π‘π‘) at π₯π₯ = 0 must vanish. The general form of the displacement function shown in Fig.15 has the following mathematical expression: 18 0; β§ ππ(π₯π₯ β π£π£ππ π‘π‘); βͺ βͺ π₯π₯ β₯ π£π£ππ π‘π‘, π’π’(π₯π₯, π‘π‘) = π’π’0 β (π£π£0 βπ£π£ππ )(π₯π₯ β π£π£ππ π‘π‘); β¨ π’π’0 + (1 β π£π£ππ βπ£π£ππ )π£π£0 π‘π‘ + ππ(π₯π₯ β π£π£ππ π‘π‘); βͺ βͺ β©π’π’1 + (1 β π£π£ππ βπ£π£ππ )π£π£0 π‘π‘; π£π£ππ π‘π‘ β βf β€ π₯π₯ β€ π£π£ππ π‘π‘, π£π£ππ π‘π‘ β€ π₯π₯ β€ π£π£ππ π‘π‘ β βf , (51) π£π£ππ π‘π‘ β βg β€ π₯π₯ β€ π£π£ππ π‘π‘, 0 β€ π₯π₯ β€ π£π£ππ π‘π‘ β βg . The continuity of π’π’(π₯π₯, π‘π‘) and its derivative πππ₯π₯ π’π’(π₯π₯, π‘π‘) requires that ππ(0) = ππ β² (0) = 0, ππ(ββf ) = π’π’0 + (π£π£0 βπ£π£ππ )βππ , ππ β² (ββf ) = β(π£π£0 βπ£π£ππ ), ππ(0) = 0, ππβ² (0) = β(π£π£0 βπ£π£ππ ), πποΏ½ββg οΏ½ = π’π’1 β π’π’0 , and ππβ² οΏ½ββg οΏ½ = 0. ππ π£π£ππ ππ π₯π₯ π’π’(π₯π₯, π‘π‘0 ) βg π₯π₯ = 0 βf π₯π₯ = π£π£ππ π‘π‘0 π₯π₯ = π£π£ππ π‘π‘0 π₯π₯ Fig.15. Light Pulse entering a semi-infinite, homogeneous, transparent, dispersionless rod from its left facet located at π₯π₯ = 0. Also shown is the fraction of the incident pulse reflected at the entrance facet. In the free space region outside the rod, the light propagates at velocity ππ, but inside the rod its velocity drops to π£π£ππ = ππ βππ, where ππ is the rodβs refractive index. The leading-edge of the pulse exerts a force on the material medium, causing the one-dimensional motion described by the displacement function π’π’(π₯π₯, π‘π‘). While the intrinsic disturbances of the elastic medium propagate at the speed of sound, π£π£ππ , the mechanical motion induced by the leading-edge of the light pulse propagates at the much larger speed of light, π£π£ππ . The function π’π’(π₯π₯, π‘π‘) of Eq.(51) satisfies the homogeneous wave equation everywhere except in the interval of length βf immediately before the leading edge of the light pulse, where the mass-density times the acceleration, πππππ‘π‘2 π’π’(π₯π₯, π‘π‘) = πππ£π£ππ2 ππ β³ (π₯π₯ β π£π£ππ π‘π‘), is significantly greater than the internal force-density πππππ₯π₯2 π’π’(π₯π₯, π‘π‘) = πππ£π£ππ2 ππ β³ (π₯π₯ β π£π£ππ π‘π‘). The difference between these two expressions, namely, ππ(π£π£ππ2 β π£π£ππ2 )ππ β³ (π₯π₯ β π£π£ππ π‘π‘), must of course be equal to the optical forcedensity exerted by the leading edge of the light beam. Integrating this optical force-density over the interval of length βf yields the total force exerted by the leading edge of the pulse as πΉπΉπ₯π₯ = ππ(π£π£ππ2 β π£π£ππ2 )[ππ β² (0) β ππ β² (ββf )] = πποΏ½1 β (π£π£ππ βπ£π£ππ )2 οΏ½π£π£0 π£π£ππ . (52) Next, we compute the total momentum of the slab at time π‘π‘, which is readily obtained from Eq.(51), as follows: β πππ₯π₯ (π‘π‘) = β«0 πππππ‘π‘ π’π’(π₯π₯, π‘π‘)ππππ 19 π£π£ππ π‘π‘ = ππ(1 β π£π£ππ βπ£π£ππ )π£π£0 (π£π£ππ π‘π‘ β βg ) + οΏ½ π£π£ππ π‘π‘ββg π£π£ π‘π‘ πποΏ½(1 β π£π£ππ βπ£π£ππ )π£π£0 β π£π£ππ ππβ² (π₯π₯ β π£π£ππ π‘π‘)οΏ½ππππ +πππ£π£0 οΏ½(π£π£ππ β π£π£ππ )π‘π‘ β βf οΏ½ β πππ£π£ππ β«π£π£ πππ‘π‘ββ ππ β² (π₯π₯ β π£π£ππ π‘π‘)ππππ ππ f = πποΏ½(π’π’1 β π’π’0 )π£π£ππ + π’π’0 π£π£ππ οΏ½ + πποΏ½1 β (π£π£ππ βπ£π£ππ )2 οΏ½π£π£0 π£π£ππ π‘π‘. (53) The rate of flow of mechanical momentum into the rod, namely, πππ‘π‘ πππ₯π₯ (π‘π‘), is thus seen to be precisely equal to the total optical force πΉπΉπ₯π₯ exerted by the leading edge of the light pulse; see Eq.(52). The small constant term on the right-hand side of Eq.(53) accounts for the momentum picked up by the rod during the early stages of entry, when the leading-edge of the light pulse arrives at the entrance facet. Note that the displacement π’π’(π₯π₯, π‘π‘) depicted in Fig.15 is the superposition of two functions, one that travels with the speed of light, π£π£ππ , and another that travels behind the first one at the speed of sound, π£π£ππ . Building on this idea, we may plot π’π’(π₯π₯, π‘π‘) at an instant of time when the pulse, whose duration is ππ, has fully entered the rod. In Fig.16, which corresponds to π‘π‘0 > ππ, the positive displacement represents that part of the function π’π’(π₯π₯, π‘π‘) which moves forward at the speed of light, π£π£ππ , whereas the negative displacement represents the part that travels along π₯π₯ at the speed of sound, π£π£ππ . 2π’π’0 + π£π£0 ππ 2[π’π’0 β π’π’1 + (π£π£0 βπ£π£ππ )β1 ] β(π£π£ππ βπ£π£ππ )π£π£0 ππ π’π’(π₯π₯, π‘π‘0 ) π₯π₯ = π£π£ππ (π‘π‘0 β ππ) βf π₯π₯ = π£π£ππ (π‘π‘0 β ππ) βg βg π₯π₯ = π£π£ππ π‘π‘0 βf π₯π₯ = π£π£ππ π‘π‘0 π₯π₯ Fig.16. Two contributions to the displacement π’π’(π₯π₯, π‘π‘) at an instant of time π‘π‘0 when the light pulse of duration ππ is fully inside the rod. The positive (upper) displacement travels along the π₯π₯-axis at the speed of light, π£π£ππ , with the negative (lower) displacement trailing behind at the speed of sound, π£π£ππ . Suppose now that the entrance facet of the rod in Fig.15 is antireflection coated, so that a net οΏ½ pulls on the front facet, while the leading edge of the pulse travels inside the rod force of βπΉπΉ0 ππ with velocity π£π£ππ , as before. Figure 17 shows that the general form of π’π’(π₯π₯, π‘π‘) will now differ from that given by Eq.(51) only when 0 β€ π₯π₯ β€ π£π£ππ π‘π‘. Specifically, the linear segment of π’π’(π₯π₯, π‘π‘) in the interval 0 β€ π₯π₯ β€ π£π£ππ π‘π‘ β βg is now given by π’π’1 + (1 β π£π£ππ βπ£π£ππ )π£π£0 π‘π‘ + (πΉπΉ0 βππ)(π₯π₯ β π£π£ππ π‘π‘), and the left-hand boundary conditions of ππ(π₯π₯) are ππ(ββg ) = π’π’1 β π’π’0 β (πΉπΉ0 βππ )βg and ππβ² (ββg ) = πΉπΉ0 βππ. Fig.17. When the entrance facet of the rod in Fig.15 is antireflection coated, a constant οΏ½ pulls on that facet, causing the force βπΉπΉ0 ππ displacement function π’π’(π₯π₯, π‘π‘) to acquire a positive slope, πΉπΉ0 βππ , in the interval between π₯π₯ = 0 and π₯π₯ = π£π£ππ π‘π‘ β βg . π’π’(π₯π₯, π‘π‘0 ) βg π₯π₯ = π£π£ππ π‘π‘0 π₯π₯ = 0 20 βf π₯π₯ = π£π£ππ π‘π‘0 π₯π₯ The mechanical momentum of the slab at time π‘π‘ is now given by β πππ₯π₯ (π‘π‘) = β«0 πππππ‘π‘ π’π’(π₯π₯, π‘π‘)ππππ = πποΏ½(1 β π£π£ππ βπ£π£ππ )π£π£0 β π£π£ππ (πΉπΉ0 βππ)οΏ½(π£π£ππ π‘π‘ β βg ) +οΏ½ π£π£ππ π‘π‘ π£π£ππ π‘π‘ββg πποΏ½(1 β π£π£ππ βπ£π£ππ )π£π£0 β π£π£ππ ππβ² (π₯π₯ β π£π£ππ π‘π‘)οΏ½ππππ + πππ£π£0 [(π£π£ππ β π£π£ππ )π‘π‘ β βf ] π£π£ π‘π‘ βπππ£π£ππ β«π£π£ πππ‘π‘ββ ππ β² (π₯π₯ β π£π£ππ π‘π‘)ππππ ππ f = πποΏ½(π’π’1 β π’π’0 )π£π£ππ + π’π’0 π£π£ππ οΏ½ + (πΉπΉπ₯π₯ β πΉπΉ0 )π‘π‘. (54) As expected, the influx of mechanical momentum into the rod is somewhat reduced by the pull of the light at the antireflection coating layer. Nevertheless, the overall strength of the EM field traveling inside the rod is now greater as a result of the presence of the antireflection coating at the front facet, causing the magnitude of the force πΉπΉπ₯π₯ at the leading edge of the pulse to be greater than before. 7. Elastic vibrations of a one-dimensional inhomogeneous slab. In this section we generalize the results of Sec. 2 to cover the case of inhomogeneous slabs in one dimension. Consider the one-dimensional periodic medium depicted in Fig.18, where the mass-density and stiffness coefficient in the interval 0 β€ π₯π₯ < ππ1 are (ππ1 , ππ1 ), while the corresponding parameters in the interval ππ1 β€ π₯π₯ < ππ1 + ππ2 are (ππ2 , ππ2 ). Defining ππ1 = οΏ½ππ1 βππ1 ππ and ππ2 = οΏ½ππ2 βππ2 ππ (both positive), a separable solution for the displacement π’π’(π₯π₯, π‘π‘) may be written as follows: οΏ½ π’π’1 (π₯π₯, π‘π‘) = [π΄π΄1 exp(iππ1 π₯π₯) + π΅π΅1 exp(βiππ1 π₯π₯)] exp(βiππππ) ; π’π’2 (π₯π₯, π‘π‘) = [π΄π΄2 exp(iππ2 π₯π₯) + π΅π΅2 exp(βiππ2 π₯π₯)] exp(βiππππ) ; (0 β€ π₯π₯ β€ ππ1 ), (ππ1 β€ π₯π₯ β€ ππ1 + ππ2 ). (55) Here ππ, an arbitrarily chosen, real-valued, positive constant represents the temporal vibration frequency of the phonon under consideration. At π₯π₯ = ππ1 , the boundary conditions are π’π’1 (ππ1 , π‘π‘) = π’π’2 (ππ1 , π‘π‘), (56a) ππ1 πππ₯π₯ π’π’1 (ππ1 , π‘π‘) = ππ2 πππ₯π₯ π’π’2 (ππ1 , π‘π‘). π΄π΄1 π΄π΄2 π₯π₯ π΅π΅1 π΅π΅2 (ππ2 , ππ2 ) (56b) (ππ1 , ππ1 ) ππ1 ππ2 Fig.18. An infinitely long, periodic slab of material consists of repeated segments having widths ππ1 and ππ2 . The mass-density and stiffness coefficient of these segments are (ππ1 , ππ1 ) and (ππ2 , ππ2 ), respectively. Invoking Eq.(55), the above boundary conditions may be written π΄π΄1 exp(iππ1 ππ1 ) + π΅π΅1 exp(βiππ1 ππ1 ) = π΄π΄2 exp(iππ2 ππ1 ) + π΅π΅2 exp(βiππ2 ππ1 ), ππ1 ππ1 [π΄π΄1 exp(iππ1 ππ1 ) β π΅π΅1 exp(βiππ1 ππ1 )] = ππ2 ππ2 [π΄π΄2 exp(iππ2 ππ1 ) β π΅π΅2 exp(βiππ2 ππ1 )]. 21 (57a) (57b) Solving Eqs.(57) for (π΄π΄2 , π΅π΅2 ) in terms of (π΄π΄1 , π΅π΅1 ), we find οΏ½ π΄π΄2 exp(iππ2 ππ1 ) ππ2 ππ2 exp(iππ2 ππ1 ) β οΏ½ οΏ½ = ½οΏ½ π΅π΅2 = ½οΏ½ exp(βiππ2 ππ1 ) β ππ2 ππ2 exp(βiππ2 ππ1 ) exp(βππππ2 ππ1 ) exp(iππ2 ππ1 ) οΏ½οΏ½ π΄π΄1 exp(iππ1 ππ1 ) + π΅π΅1 exp(βiππ1 ππ1 ) οΏ½=οΏ½ οΏ½ π΅π΅2 ππ1 ππ1 [π΄π΄1 exp(iππ1 ππ1 ) β π΅π΅1 exp(βiππ1 ππ1 )] π΄π΄2 (ππ2 ππ2 )β1 exp(βiππ2 ππ1 ) β(ππ2 ππ2 )β1 exp(iππ2 ππ1 ) οΏ½οΏ½ π΄π΄1 exp(iππ1 ππ1 ) + π΅π΅1 exp(βiππ1 ππ1 ) οΏ½ ππ1 ππ1 [π΄π΄1 exp(iππ1 ππ1 ) β π΅π΅1 exp(βiππ1 ππ1 )] π΄π΄1 [1 + (ππ1 ππ1 βππ2 ππ2 )] exp[i(ππ1 β ππ2 )ππ1 ] + π΅π΅1 [1 β (ππ1 ππ1 βππ2 ππ2 )] exp[βi(ππ1 + ππ2 )ππ1 ] οΏ½. π΄π΄1 [1 β (ππ1 ππ1 βππ2 ππ2 )] exp[i(ππ1 + ππ2 )ππ1 ] + π΅π΅1 [1 + (ππ1 ππ1 βππ2 ππ2 )] exp[βi(ππ1 β ππ2 )ππ1 ] At π₯π₯ = ππ1 + ππ2 , the boundary conditions are π’π’2 (ππ1 + ππ2 , π‘π‘) = πππ’π’1 (0, π‘π‘), ππ2 πππ₯π₯ π’π’2 (ππ1 + ππ2 , π‘π‘) = ππππ1 πππ₯π₯ π’π’1 (0, π‘π‘). (58) (59a) (59b) Here ππ is an arbitrary constant, which must be subsequently determined. Invoking Eq.(55), the boundary conditions in Eq.(59) are written π΄π΄2 exp[iππ2 (ππ1 + ππ2 )] + π΅π΅2 exp[βiππ2 (ππ1 + ππ2 )] = ππ(π΄π΄1 + π΅π΅1 ), ππ2 ππ2 {π΄π΄2 exp[iππ2 (ππ1 + ππ2 )] β π΅π΅2 exp[βiππ2 (ππ1 + ππ2 )]} = ππ1 ππ1 ππ(π΄π΄1 β π΅π΅1 ). (60a) (60b) Solving Eqs.(60) for (π΄π΄1 , π΅π΅1 ) in terms of (π΄π΄2 , π΅π΅2 ), we obtain ππ οΏ½ β ππ οΏ½ β ππ οΏ½ π΄π΄1 π΅π΅1 π΄π΄1 π΅π΅1 1 1 ππ1 ππ1 οΏ½ = ½οΏ½ οΏ½ = ½οΏ½ βππ1 ππ1 1 1 οΏ½οΏ½ π΄π΄1 π΅π΅1 οΏ½=οΏ½ exp[iππ2 (ππ1 + ππ2 )] exp[βiππ2 (ππ1 + ππ2 )] π΄π΄2 οΏ½οΏ½ οΏ½ ππ2 ππ2 exp[iππ2 (ππ1 + ππ2 )] βππ2 ππ2 exp[βiππ2 (ππ1 + ππ2 )] π΅π΅2 π΄π΄2 exp[βiππ2 (ππ1 + ππ2 )] exp[iππ2 (ππ1 + ππ2 )] 1β(ππ1 ππ1 ) οΏ½οΏ½ οΏ½οΏ½ οΏ½ β1β(ππ1 ππ1 ) ππ2 ππ2 exp[iππ2 (ππ1 + ππ2 )] βππ2 ππ2 exp[βiππ2 (ππ1 + ππ2 )] π΅π΅2 [1 + (ππ2 ππ2 βππ1 ππ1 )] exp[iππ2 (ππ1 + ππ2 )] [1 β (ππ2 ππ2 βππ1 ππ1 )] exp[βiππ2 (ππ1 + ππ2 )] π΄π΄2 οΏ½ οΏ½ οΏ½. [1 β (ππ2 ππ2 βππ1 ππ1 )] exp[iππ2 (ππ1 + ππ2 )] [1 + (ππ2 ππ2 βππ1 ππ1 )] exp[βiππ2 (ππ1 + ππ2 )] π΅π΅2 (61) Next, we define the parameter ππ = ππ1 ππ1 βππ2 ππ2 = οΏ½ππ1 ππ1 /ππ2 ππ2 , and substitute from Eq.(58) into Eq.(61) to arrive at ππ οΏ½ π΄π΄1 π΅π΅1 οΏ½ = ¼οΏ½ (1 + ππβ1 ) exp[iππ2 (ππ1 + ππ2 )] (1 β ππβ1 ) exp[βiππ2 (ππ1 + ππ2 )] οΏ½ (1 β ππβ1 ) exp[iππ2 (ππ1 + ππ2 )] (1 + ππβ1 ) exp[βiππ2 (ππ1 + ππ2 )] ×οΏ½ π΄π΄1 (1 + ππ) exp[i(ππ1 β ππ2 )ππ1 ] + π΅π΅1 (1 β ππ) exp[βi(ππ1 + ππ2 )ππ1 ] οΏ½ π΄π΄1 (1 β ππ) exp[i(ππ1 + ππ2 )ππ1 ] + π΅π΅1 (1 + ππ) exp[βi(ππ1 β ππ2 )ππ1 ] π΄π΄1 (2 + ππ + ππβ1 ) exp[i(ππ1 ππ1 + ππ2 ππ2 )] β π΅π΅1 (ππ β ππβ1 ) exp[βi(ππ1 ππ1 β ππ2 ππ2 )] β1 β1 β+π΄π΄1 (2 β ππ β ππ ) exp[i(ππ1 ππ1 β ππ2 ππ2 )] + π΅π΅1 (ππ β ππ ) exp[βi(ππ1 ππ1 + ππ2 ππ2 )]β = ¼β β. π΄π΄1 (ππ β ππβ1 ) exp[i(ππ1 ππ1 + ππ2 ππ2 )] + π΅π΅1 (2 β ππ β ππβ1 ) exp[βi(ππ1 ππ1 β ππ2 ππ2 )] ββπ΄π΄1 (ππ β ππβ1 ) exp[i(ππ1 ππ1 β ππ2 ππ2 )] + π΅π΅1 (2 + ππ + ππβ1 ) exp[βi(ππ1 ππ1 + ππ2 ππ2 )]β 22 (62) From Eq.(62), the ratio π΄π΄1 βπ΅π΅1 is seen to satisfy the following equation: π΄π΄1 π΅π΅1 (π΄π΄ βπ΅π΅ )οΏ½2 cos(ππ ππ )+i(ππ+ππβ1 ) sin(ππ2 ππ2 )οΏ½ exp(iππ1 ππ1 ) β i(ππβππβ1 ) sin(ππ2 ππ2 ) exp(βiππ1 ππ1 ) 2 = (π΄π΄ 1βπ΅π΅ 1)[i(ππβππβ12) sin(ππ 1 1 2 ππ2 )] exp(iππ1 ππ1 ) + [2 cos(ππ2 ππ2 )βi(ππ+ππ This yields a quadratic equation for π΄π΄1 βπ΅π΅1, namely, β1 ) sin(ππ . (63) 2 ππ2 )] exp(βiππ1 ππ1 ) (ππ β ππβ1 ) sin(ππ2 ππ2 ) exp(iππ1 ππ1 ) (π΄π΄1 βπ΅π΅1 )2 β 2[2 sin(ππ1 ππ1 ) cos(ππ2 ππ2 ) + (ππ + ππβ1 ) cos(ππ1 ππ1 ) sin(ππ2 ππ2 )](π΄π΄1 βπ΅π΅1 ) +(ππ β ππβ1 ) sin(ππ2 ππ2 ) exp(βiππ1 ππ1 ) = 0. Solving Eq.(64), we obtain two solutions for π΄π΄1 βπ΅π΅1, as follows: π΄π΄1 π΅π΅1 = 2 sin(ππ1 ππ1 ) cos(ππ2 ππ2 )+(ππ+ππβ1 ) cos(ππ1 ππ1 ) sin(ππ2 ππ2 ) ± οΏ½[2 sin(ππ1 ππ1 ) cos(ππ2 ππ2 )+(ππ+ππβ1 ) cos(ππ1 ππ1 ) sin(ππ2 ππ2 )]2 β [(ππβππβ1 ) sin(ππ2 ππ2 )]2 (ππβππβ1 ) sin(ππ2 ππ2 ) exp(iππ1 ππ1 ) (64) . (65) Aside from the phase-factor exp(iππ1 ππ1 ) in the denominator of Eq.(65), the two solutions thus obtained for π΄π΄1 βπ΅π΅1 are readily seen to be inverses of each other. If the expression under the radical happens to be positive, these two values for π΄π΄1 βπ΅π΅1 will result in a pair of conjugate solutions for π’π’(π₯π₯, π‘π‘) throughout the entire medium. Returning now to Eq.(62), we solve for ππ and, with the aid of Eq.(65), we find ππ = ½i(π΄π΄1 βπ΅π΅1 )(ππ β ππβ1 ) sin(ππ2 ππ2 ) exp(iππ1 ππ1 ) + cos(ππ2 ππ2 ) exp(βiππ1 ππ1 ) β ½i(ππ + ππβ1 ) sin(ππ2 ππ2 ) exp(βiππ1 ππ1 ) β ππ = cos(ππ1 ππ1 ) cos(ππ2 ππ2 ) β ½(ππ + ππβ1 ) sin(ππ1 ππ1 ) sin(ππ2 ππ2 ) ±½iοΏ½[2 sin(ππ1 ππ1 ) cos(ππ2 ππ2 ) + (ππ + ππβ1 ) cos(ππ1 ππ1 ) sin(ππ2 ππ2 )]2 β [(ππ β ππβ1 ) sin(ππ2 ππ2 )]2 . (66) When the expression under the radical is positive, the complex number ππ will have unit magnitude and may be written as ππ = exp(iππ); the phase ππ, of course, depends on the chosen temporal frequency ππ, as well as on the material parameters (ππ1 , ππ1 , ππ1 ) and (ππ2 , ππ2 , ππ2 ). The real part of ππ given by Eq.(66) will be equal to cos ππ, that is, cos ππ = (ππ+1)2 cos(ππ1 ππ1 +ππ2 ππ2 ) β (ππβ1)2 cos(ππ1 ππ1 βππ2 ππ2 ) 4ππ . (67) At ππ = 0, where ππ1 ππ1 ± ππ2 ππ2 = οΏ½οΏ½ππ1 βππ1 ππ1 ± οΏ½ππ2 βππ2 ππ2 οΏ½ππ = 0, we have cos ππ = 1. As ππ increases, cos ππ drops at first, but eventually it exits the interval [β1, +1], at which point the expression under the radical in Eq.(66) becomes negative. The range of frequencies ππ in which both values of ππ are real represents a bandgap, within which long-range phonons cannot exist. As ππ continues to climb, cos ππ in Eq.(67) returns to the interval [β1, +1], thus creating another frequency band for phonons. Each allowed band, of course, is followed by a band-gap and then by a new band of allowed phonon frequencies. Needless to say, the existence of the phase-angle ππ is what enables the definition of a Bloch wave-vector ππ for the system of Fig.18, where ππ = ππ(ππ1 + ππ2 ). In the special case when ππ = 1, that is, ππ1 ππ1 = ππ2 ππ2 , the only solution of Eq.(64) turns out to be π΄π΄1 βπ΅π΅1 = 0. Even though ππ1 could still differ from ππ2 , the overall solution of the problem is trivial, and, as Eq.(67) reveals, there will be no bandgaps. When the expression under the radical in Eqs.(65) and (66) is negative, π΄π΄1 βπ΅π΅1 becomes a complex number of unit magnitude, having a phase-angle that depends on ππ as well as on material parameters. At the same time, the two values of ππ become real, one greater and the other less than unity. These values of π΄π΄1 βπ΅π΅1 and ππ represent two distinct solutions for π’π’ππ (π₯π₯), one 23 growing and the other declining exponentially along the π₯π₯-axis. Each of these eigen-functions has its own complex conjugate, representing two additional solutions associated with negative ππ. Note, with reference to Eqs.(55), (58) and (65), that a simultaneous reversal of the signs of ππ, ππ1 β (π₯π₯) and ππ2 yields the complex conjugate solution π’π’ππ without changing the corresponding value of ππ given by Eq.(66). In practice, one ignores the values of ππ that lead to exponentially growing or decaying eigen-functions π’π’ππ (π₯π₯), relegating them to βband-gapsβ in the frequency domain. 7.1. Finite-width periodic slab. The Bloch solutions of the wave equation associated with the phase-factor ππ = exp(iππ) = exp[iππ(ππ1 + ππ2 )] may be written as π’π’ππ (π₯π₯, π‘π‘) = π’π’ππ (π₯π₯) exp[i(ππππ β ππππ)], where the (generally complex) function π’π’ππ (π₯π₯) is periodic, having a periodicity of ππ1 + ππ2 . These are the same solutions as in Eq.(55), except that the linear phase-factor exp(iππππ) has been factored out in order to emphasize the fact that π’π’ππ (π₯π₯) is periodic. Considering that the sign of ππ β (π₯π₯) is immaterial β in the sense that π’π’ππ (π₯π₯) exp(iππππ) and its conjugate π’π’ππ exp(βiππππ) are equally valid spatial profiles satisfying the wave equation irrespective of whether ππ is positive or negative β we conclude that separable solutions of the wave equation may be equivalently written in the following form: π’π’ππ (π₯π₯, π‘π‘) = |π’π’ππ (π₯π₯)|{π¬π¬1 sin[ππππ + ππππ (π₯π₯)] + π¬π¬2 cos[ππππ + ππππ (π₯π₯)]}{π¬π¬3 sin ππππ + π¬π¬4 cos ππππ} = π¬π¬|π’π’ππ (π₯π₯)| sin[ππππ + ππππ (π₯π₯) + ππ0 ] sin(ππππ + ππ1 ). (68) In the above equation, π’π’ππ (π₯π₯) = |π’π’ππ (π₯π₯)| exp[iππππ (π₯π₯)] is periodic, ππ is a function of ππ, and the real-valued coefficients π¬π¬1, π¬π¬2 , π¬π¬3 , π¬π¬4 , π¬π¬, ππ0 , ππ1 are arbitrary constants. If a boundary, say, that at π₯π₯ = βπΏπΏ, is immobile, then ππ0 must be chosen to ensure that π’π’ππ (βπΏπΏ, π‘π‘) = 0. If the other boundary at π₯π₯ = πΏπΏ is also immobile, then only those values of ππ will be acceptable that satisfy sin[ππππ + ππππ (πΏπΏ) + ππ0 ] = 0. Similarly, if the boundaries at π₯π₯ = ±πΏπΏ happen to be free, then ππ0 and ππ must be chosen to ensure that πππ₯π₯ π’π’ππ (±πΏπΏ, π‘π‘) = 0. In this way, a subset of values of ππ is selected for which π’π’ππ (π₯π₯, π‘π‘) of Eq.(68) is a Bloch wave that satisfies the wave equation as well as the specified boundary conditions. Needless to say, all such values of ππ are positive, since the negative values of ππ are automatically incorporated into the eigen-functions defined by Eq.(68). The initial conditions π’π’(π₯π₯, π‘π‘ = 0) and π£π£(π₯π₯, π‘π‘ = 0) may now be expanded in terms of the spatial parts of the eigen-functions given by Eq.(68) which also satisfy the boundary conditions, namely, π’π’(π₯π₯, π‘π‘ = 0) = βππ π¬π¬ππ |π’π’ππ (π₯π₯)| sin[ππππ + ππππ (π₯π₯) + ππ0ππ ], π£π£(π₯π₯, π‘π‘ = 0) = βππ π¬π¬Μππ |π’π’ππ (π₯π₯)| sin[ππππ + ππππ (π₯π₯) + ππ0ππ ]. (69a) (69b) Consequently, the general solution of the wave equation is given by π’π’(π₯π₯, π‘π‘) = βππ|π’π’ππ (π₯π₯)| sin[ππππ + ππππ (π₯π₯) + ππ0ππ ] οΏ½π¬π¬ππ cos(ππππ) + π¬π¬Μππ sin(ππππ)βππ οΏ½. (70) Note in the above equations that π’π’ππ (π₯π₯) is dimensionless, whereas π¬π¬ππ and π¬π¬Μππ have the units of length and velocity, respectively. The overall linear momentum ππ(π‘π‘) of a slab of width 2πΏπΏ is readily derived from its mass and velocity profiles, as follows: πΏπΏ ππ(π‘π‘) = β«βπΏπΏ ππ(π₯π₯)πππ‘π‘ π’π’(π₯π₯, π‘π‘)ππππ 24 πΏπΏ = βπποΏ½π¬π¬Μππ cos(ππππ) β π¬π¬ππ ππ sin(ππππ)οΏ½ β«βπΏπΏ ππ(π₯π₯)|π’π’ππ (π₯π₯)| sin[ππππ + ππππ (π₯π₯) + ππ0ππ ] ππππ. (71) If the slab happens to be free at both ends, the integral appearing in Eq.(71) will identically vanish for all nonzero frequencies ππ β this is readily proven by integrating the wave equation πππ₯π₯ [ππ(π₯π₯)πππ₯π₯ π’π’ππ (π₯π₯, π‘π‘)] = βππ2 ππ(π₯π₯)π’π’ππ (π₯π₯, π‘π‘), satisfied by all eigen-functions in Eq.(68), over the interval [βπΏπΏ, πΏπΏ]. The only surviving term in Eq.(71) corresponds to ππ = 0, whose eigen-function πΏπΏ is π’π’0 (π₯π₯, π‘π‘) = 1. Consequently, ππ(π‘π‘) = π¬π¬Μ0 β«βπΏπΏ ππ(π₯π₯)ππππ, where π¬π¬Μ0 is the coefficient of the zeroπΏπΏ frequency term in Eq.(69b), while β« ππ(π₯π₯)ππππ is the total mass of the slab. Note that π¬π¬Μ0 is not βπΏπΏ necessarily the average initial velocity of the slab, as the eigen-functions π’π’ππ (π₯π₯, π‘π‘) may not have a vanishing integral over the interval [βπΏπΏ, πΏπΏ] when ππ β 0. 7.2. Eigen-function orthogonality. In general, an inhomogeneous one-dimensional medium is specified by its mass-density ππ(π₯π₯) and stiffness coefficient ππ(π₯π₯), both of which vary with position along the π₯π₯-axis. In such a medium, the wave equation governing the displacement π’π’(π₯π₯, π‘π‘) is derived from Newtonβs second law of motion, as follows: πππ₯π₯ [ππ(π₯π₯)πππ₯π₯ π’π’(π₯π₯, π‘π‘)] = ππ(π₯π₯)πππ‘π‘2 π’π’(π₯π₯, π‘π‘). (72) ππ(π₯π₯)π’π’Μ (π₯π₯) + ππΜ (π₯π₯)π’π’Μ (π₯π₯) + ππ2 ππ(π₯π₯)π’π’(π₯π₯) = 0. (73) The eigen-modes of the above equation are obtained by separation of variables in the form of π’π’(π₯π₯, π‘π‘) = π’π’(π₯π₯) exp(±iππππ), which leads to Here differentiation with respect to π₯π₯ is indicated by an over-dot. Note that π’π’(π₯π₯), a complex function in general, does not depend on the choice of sign for ππ. However, since π’π’β (π₯π₯), the conjugate of π’π’(π₯π₯), is also a solution of Eq.(73) β because ππ(π₯π₯), ππ(π₯π₯), and ππ are real β we may arbitrarily assign π’π’(π₯π₯) to positive ππ, in which case π’π’β (π₯π₯) gets automatically assigned to βππ. Suppose now that π’π’ππ1 (π₯π₯) and π’π’ππ2 (π₯π₯) are two eigen-functions of Eq.(73), corresponding respectively to distinct temporal frequencies ππ1 and ππ2 . We will have ππ(π₯π₯)π’π’Μ ππ1 (π₯π₯) + ππΜ (π₯π₯)π’π’Μ ππ1 (π₯π₯) + ππ12 ππ(π₯π₯)π’π’ππ1 (π₯π₯) = 0, ππ(π₯π₯)π’π’Μ ππ2 (π₯π₯) + ππΜ (π₯π₯)π’π’Μ ππ2 (π₯π₯) + ππ22 ππ(π₯π₯)π’π’ππ2 (π₯π₯) = 0. (74a) (74b) Multiplying the first of the above equations by π’π’ππ2 (π₯π₯) and the second by π’π’ππ1 (π₯π₯), then subtracting one equation from the other, yields π’π’ππ2 (π₯π₯) ππ [ππ(π₯π₯)π’π’Μ ππ1 (π₯π₯)] β ππππ π’π’ππ1 (π₯π₯) ππ [ππ(π₯π₯)π’π’Μ ππ2 (π₯π₯)] + ππππ (ππ12 β ππ22 )ππ(π₯π₯)π’π’ππ1 (π₯π₯)π’π’ππ2 (π₯π₯) = 0. (75) Next, we integrate Eq.(75) over the interval [ππ, ππ] and apply the method of integration by parts to arrive at ππ [ππ(ππ)π’π’Μ ππ1 (ππ)π’π’ππ2 (ππ) β ππ(ππ)π’π’Μ ππ1 (ππ)π’π’ππ2 (ππ)] β β«ππ ππ(π₯π₯)π’π’Μ ππ1 (π₯π₯)π’π’Μ ππ2 (π₯π₯)ππππ ππ β[ππ(ππ)π’π’ππ1 (ππ)π’π’Μ ππ2 (ππ) β ππ(ππ)π’π’ππ1 (ππ)π’π’Μ ππ2 (ππ)] + β«ππ ππ(π₯π₯)π’π’Μ ππ1 (π₯π₯)π’π’Μ ππ2 (π₯π₯)ππππ ππ +(ππ12 β ππ22 ) β«ππ ππ(π₯π₯)π’π’ππ1 (π₯π₯)π’π’ππ2 (π₯π₯)ππππ = 0. 25 (76) If the boundary conditions are such that the bracketed terms in Eq.(76) vanish (or cancel out), and if |ππ1 | β |ππ2 |, we observe that π’π’ππ1 (π₯π₯) and π’π’ππ2 (π₯π₯), over the interval [ππ, ππ] and in the presence of the weight function ππ(π₯π₯), are orthogonal to each other. In other words, ππ β«ππ ππ(π₯π₯)π’π’ππ1 (π₯π₯)π’π’ππ2 (π₯π₯)ππππ = 0. (77) A few words must be said at this point about the zero-frequency eigen-function, which is π₯π₯ readily obtained from Eq.(72) as π’π’0 (π₯π₯, π‘π‘) = ππ1 + ππ2 β«ππ ππ β1 (π₯π₯ β² )πππ₯π₯ β² ; here ππ1 and ππ2 are arbitrary (real-valued) constants. Considering that ππ(π₯π₯) > 0 throughout the interval [ππ, ππ], the usual boundary conditions, π’π’0 (ππ, π‘π‘) = 0, π’π’0 (ππ, π‘π‘) = 0, πππ₯π₯ π’π’0 (ππ, π‘π‘) = 0, or πππ₯π₯ π’π’0 (ππ, π‘π‘) = 0, ensure that ππ2 = 0. Thus, when one or both boundaries are immobile, π’π’0 (π₯π₯, π‘π‘) = 0, whereas π’π’0 (π₯π₯, π‘π‘) = ππ1 when both boundaries are free. The initial position π’π’(π₯π₯, π‘π‘ = 0) and velocity π£π£(π₯π₯, π‘π‘ = 0) of the medium may now be expanded in terms of the above eigen-functions π’π’ππ (π₯π₯), as follows: β π’π’(π₯π₯, π‘π‘ = 0) = β«ββ ππ0 (ππ)π’π’ππ (π₯π₯)ππππ, β π£π£(π₯π₯, π‘π‘ = 0) = β«ββ ππ0 (ππ)π’π’ππ (π₯π₯)ππππ. (78a) (78b) Note that some values of ππ (e.g., those within a bandgap) may not possess acceptable eigenfunctions, in which case the domain of integration in Eq.(78) will be a subset of the real axis. Considering that the integrals in Eq.(78) range over both positive and negative values of ππ, and β (π₯π₯), that π’π’βππ (π₯π₯) = π’π’ππ the functions ππ0 (ππ) and ππ0 (ππ) must be Hermitian. Since π’π’ππ (π₯π₯) is not β (π₯π₯), necessarily orthogonal to π’π’ππ care must be taken when determining the functions ππ0 (ππ) and ππ0 (ππ); this problem is addressed in some detail in Appendix C. Given that each eigen-function π’π’ππ (π₯π₯) is associated with both temporal functions exp(iππππ) and exp(βiππππ), we must write β π’π’(π₯π₯, π‘π‘) = β«ββ[ππ+ (ππ)π’π’ππ (π₯π₯) exp(iππππ) + ππβ (ππ)π’π’ππ (π₯π₯) exp(βiππππ)]ππππ, β π£π£(π₯π₯, π‘π‘) = πππ‘π‘ π’π’(π₯π₯, π‘π‘) = β«ββ iππ[ππ+ (ππ)π’π’ππ (π₯π₯) exp(iππππ) β ππβ (ππ)π’π’ππ (π₯π₯) exp(βiππππ)]ππππ. A quick glance at Eqs.(78) and (79) now yields the functions ππ± (ππ), as follows: ππ± (ππ) = ½[ππ0 (ππ) β i ππ0 (ππ)βππ ]. (79a) (79b) (80) Finally, substitution from Eq.(80) into Eqs.(79) yields β π’π’(π₯π₯, π‘π‘) = β«ββ[ππ0 (ππ) cos(ππππ) + ππβ1 ππ0 (ππ) sin(ππππ)]π’π’ππ (π₯π₯)ππππ, β π£π£(π₯π₯, π‘π‘) = β«ββ[ππ0 (ππ) cos(ππππ) β ππππ0 (ππ) sin(ππππ)]π’π’ππ (π₯π₯)ππππ. (81a) (81b) These are general formulas for computing the position and velocity profiles of the medium as functions of time starting with the initial conditions π’π’(π₯π₯, π‘π‘ = 0) and π£π£(π₯π₯, π‘π‘ = 0). 7.3. Momentum conservation. Integrating the fundamental Eq.(73) over the [ππ, ππ] interval, we find ππ ππ2 β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)ππππ = ππ(ππ)π’π’Μ ππ (ππ) β ππ(ππ)π’π’Μ ππ (ππ). (82) 26 For the overall momentum of the system to be conserved, the boundaries at π₯π₯ = ππ and π₯π₯ = ππ must be free, that is, π’π’Μ ππ (ππ) = π’π’Μ ππ (ππ) = 0. Therefore, ππ β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)ππππ = 0, (ππ β 0). ππ (83) In the special case of ππ = 0, the integral β«ππ ππ(π₯π₯)π’π’0 (π₯π₯)ππππ can be nonzero. In fact, π’π’0 (π₯π₯) should be a real-valued constant, as this choice satisfies both Eq.(73) and the free boundary conditions. Next, we invoke Eq.(78b) to write ππ β«ππ ππ(π₯π₯)π£π£(π₯π₯, π‘π‘ = 0)π’π’0 (π₯π₯)ππππ β ππ ππ = β«ββ ππ0 (ππ) β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’0 (π₯π₯)ππππ ππππ = ππ0 (0) β«ππ ππ(π₯π₯)π’π’02 (π₯π₯)ππππ. (84) Finally, with the aid of Eqs.(81b), (83), and (84), the overall momentum of the system at time π‘π‘ may be evaluated as follows: ππ β ππ ππ(π‘π‘) = β«ππ ππ(π₯π₯)π£π£(π₯π₯, π‘π‘)ππππ = β«ββ[ππ0 (ππ) cos(ππππ) β ππππ0 (ππ) sin(ππππ)] β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)ππππ ππππ ππ ππ β« ππ(π₯π₯)π’π’0 (π₯π₯)ππππ = ππ0 (0) β«ππ ππ(π₯π₯)π’π’0 (π₯π₯)ππππ = οΏ½ ππππ β«ππ ππ(π₯π₯)π’π’02 (π₯π₯)ππππ ππ οΏ½ β«ππ ππ(π₯π₯)π£π£(π₯π₯, π‘π‘ = 0)π’π’0 (π₯π₯)ππππ. (85) Considering that the time π‘π‘ has disappeared from the above equation, and that π’π’0 (π₯π₯) is a constant over the [ππ, ππ] interval, Eq.(85) confirms that the overall momentum of the system is conserved. 7.4. Energy conservation. Invoking Eq.(81b) and the orthogonality condition, Eq.(77), the kinetic energy of the system at time π‘π‘ may be expressed as follows: ππ β°πΎπΎ (π‘π‘) = ½ β«ππ ππ(π₯π₯)π£π£ 2 (π₯π₯, π‘π‘)ππππ β ππ = ½ β¬ββ[ππ0 (ππ) cos(ππππ) β ππππ0 (ππ) sin(ππππ)][ππ0 (ππβ² ) cos(ππβ² π‘π‘) β ππβ² ππ0 (ππβ² ) sin(ππβ² π‘π‘)] β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππβ² (π₯π₯)ππππππππππππβ² β ππ 2 (π₯π₯)ππππππππ = ½ β«ββ[ππ0 (ππ) cos(ππππ) β ππππ0 (ππ) sin(ππππ)]2 β«ππ ππ(π₯π₯)π’π’ππ β ππ β (π₯π₯)ππππππππ +½ β«ββ[ππ0 (ππ) cos(ππππ) β ππππ0 (ππ) sin(ππππ)][ππ0β (ππ) cos(ππππ) β ππππ0β (ππ) sin(ππππ)] β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππ . (86) As for the potential energy, we first need to prove the following identity using the fundamental Eq.(73) and the method of integration by parts: 0 ππ ππ β«ππ ππ(π₯π₯)π’π’Μ ππ (π₯π₯)π’π’Μ ππβ² (π₯π₯)ππππ = [ππ(ππ)π’π’Μ ππ (ππ)π’π’ππβ² (ππ) β ππ(ππ)π’π’Μ ππ (ππ)π’π’ππβ² (ππ)] β οΏ½ ππ ππ ππππ ππ [ππ(π₯π₯)π’π’Μ ππ (π₯π₯)]π’π’ππβ² (π₯π₯)ππππ = ππ2 β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππβ² (π₯π₯)ππππ. (87) The bracketed term is the same one that appeared in Eq.(76), and has been set to zero because of the imposed boundary conditions. The potential energy of the system at time π‘π‘ may now be evaluated using Eq.(81a), Eq.(87), and the orthogonality condition, Eq.(77), as follows: ππ β°ππ (π‘π‘) = ½ β«ππ ππ(π₯π₯)π’π’Μ 2 (π₯π₯, π‘π‘)ππππ β ππ = ½ β¬ββ[ππ0 (ππ) cos(ππππ) + ππ β1 ππ0 (ππ) sin(πππ‘π‘)][ππ0 (ππβ² ) cos(ππβ² π‘π‘) + ππβ²β1 ππ0 (ππβ² ) sin(ππβ² π‘π‘)] β«ππ ππ(π₯π₯)π’π’Μ ππ (π₯π₯)π’π’Μ ππβ² (π₯π₯)ππππππππππππβ² 27 β ππ 2 (π₯π₯)ππππππππ = ½ β«ββ[ππ0 (ππ) cos(ππππ) + ππβ1 ππ0 (ππ) sin(ππππ)]2 ππ2 β«ππ ππ(π₯π₯)π’π’ππ β ππ β (π₯π₯)ππππππππ +½ β«ββ[ππ0 (ππ) cos(ππππ) + ππβ1 ππ0 (ππ) sin(ππππ)][ππ0β (ππ) cos(ππππ) + ππβ1 ππ0β (ππ) sin(ππππ)] ππ2 β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππ . (88) The total energy of the system is thus given by β ππ 2 (π₯π₯)ππππππππ β°πΎπΎ (π‘π‘) + β°ππ (π‘π‘) = ½ β«ββ[ππ02 (ππ)+ππ2 ππ02 (ππ)] β«ππ ππ(π₯π₯)π’π’ππ β ππ β (π₯π₯)ππππππππ +½ β«ββ[ππ0 (ππ)ππ0β (ππ) + ππ2 ππ0 (ππ)ππ0β (ππ)] β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππ . (89) The fact that the sum of kinetic and potential energies in Eq.(89) is time-independent is proof that the total energy of the system is conserved. Note that, unlike momentum conservation, the conservation of energy proven here holds for both free and immobile boundary conditions. 8. Concluding remarks. We have shown that any momentum and kinetic energy deposited (e.g., by a short pulse of light) into a solid elastic medium, will evolve with time in such a way as to preserve the initial momentum and energy, despite the fact that the vibrational modes of the medium with non-zero frequency (i.e., ππ β 0) do not carry any momentum of their own. In other words, the entire mechanical momentum acquired by the medium is carried by the ππ = 0 vibrational mode. Also, in some instances in practice, the optical energy absorbed by the material medium is converted to kinetic energy of motion and/or thermal energy of elastic vibrations. In other instances, the exchanged optical energy may arise from the Doppler-shift of the reflected and/or transmitted photons. We have ignored the effects of such Doppler shifts in our analysis β see, for instance, the examples of Sec.4, where the light beam is either reflected from a mirror or transmitted through a dielectric slab. So long as the optical energy exchanged with matter is relatively small, the neglect of the Doppler shift could be justified β but of course, in extreme circumstances, there may occur situations where such effects need to be taken into consideration. References 1. M. Mansuripur, βRadiation pressure and the linear momentum of the electromagnetic field,β Optics Express 12, 5375-5401 (2004). 2. T. PoΕΎar and J. MoΕΎina, βMeasurement of elastic waves induced by the reflection of light,β Phys. Rev. Lett. 111, 185501 (2013). 3. T. PoΕΎar, βOblique reflection of a laser pulse from a perfect elastic mirror,β Opt. Lett. 39, 48-51 (2014). 4. C. Baxter and R. Loudon, βRadiation pressure and the photon momentum in dielectrics,β J. Mod. Opt. 57, 830842 (2010). 5. B. A. Kemp, βMicroscopic Theory of Optical Momentum,β Progress in Optics 60, 437-488 (2015). 6. M. Bethune-Waddell and K. J. Chau, βSimulations of radiation pressure experiments narrow down the energy and momentum of light in matter,β Reports on Progress in Physics 78, 122401 (2015). 7. K. J. Chau and H. J. Lezec, βRevisiting the Balazs thought experiment in the presence of loss: electromagneticpulse induced displacement of a positive-index slab having arbitrary complex permittivity and permeability,β Applied Physics A 105, 267-281 (2011). 8. M. Mansuripur, A. R. Zakharian, and E. M. Wright, βElectromagnetic-force distribution inside matter,β Phys. Rev. A 88, 023826 (2013). 9. N. G. C. Astrath, G. V. B. Lukasievics, L. C. Malacarne, S. E. Bialkowski, βSurface deformation effects induced by radiation pressure and electrostriction forces in dielectric solids,β Appl. Phys. Lett. 102, 231903 (2013). 10. T. PoΕΎar, A. Babnik, and J. MoΕΎina, βFrom laser ultrasonics to optical manipulation,β Optics Express 23, 79787990 (2015). 11. T. PoΕΎar and J. MoΕΎina, β1D problems of radiation pressure on elastic solids,β Optical Trapping and Optical Micromanipulation XII, edited by K. Dholakia and G. C. Spalding, Proceedings of SPIE 9548, 95480N (2015). 28 Appendix A The one-dimensional Fourier transform operator β± acting on the function ππ(π₯π₯) is defined as follows: β β±{ππ(π₯π₯)} = β«ββ ππ(π₯π₯) exp(βiππππ) ππππ = πΉπΉ(ππ). (A1) The inverse Fourier transform operator β± β1 then yields 1 β β± β1 {πΉπΉ(ππ)} = 2ππ β«ββ πΉπΉ(ππ) exp(iππππ) ππππ = ππ(π₯π₯). (A2) Some important theorems are easily proven using the above definitions. The scaling theorem, for instance, asserts that 1 β±{ππ(ππππ)} = |ππ| πΉπΉ(ππβππ); (ππ = arbitrary real number). According to the differentiation theorem, β±{ππ β² (π₯π₯)} = iππππ(ππ). A(3) (A4) The central-value theorem and its inverse are written β β«ββ ππ(π₯π₯)ππππ = πΉπΉ(0). β (A5) β«ββ πΉπΉ(ππ)ππππ = 2ππππ(0). (A6) The convolution theorem and its inverse, the multiplication theorem, read as follows: β±{ππ(π₯π₯) β ππ(π₯π₯)} = πΉπΉ(ππ)πΊπΊ(ππ), (A7) 1 β±{ππ(π₯π₯)ππ(π₯π₯)} = 2ππ πΉπΉ(ππ) β πΊπΊ(ππ). (A8) Here the convolution operation is defined as β ππ(π₯π₯) β ππ(π₯π₯) = β«ββ ππ(π₯π₯ β² )ππ(π₯π₯ β π₯π₯ β² )πππ₯π₯ β² . (A9) Finally, Parsevalβs theorem reads as follows: β 1 β β«ββ ππ(π₯π₯)ππβ (π₯π₯)ππππ = 2ππ β«ββ πΉπΉ(ππ)πΊπΊ β (ππ)ππππ. (A10) The following Fourier transform pairs are readily evaluated using straightforward integration: ππ(π₯π₯) = exp(βππ|π₯π₯|) 1 ππ(π₯π₯) = 1 + π₯π₯ 2 ππ(π₯π₯) = cos(ππππ) ππ(π₯π₯) = arctan(π₯π₯βππ) β 2ππ πΉπΉ(ππ) = ππ2 + ππ 2 ; β πΉπΉ(ππ) = ππ exp(β|ππ|). β πΉπΉ(ππ) = β (ππ > 0). πΉπΉ(ππ) = ππ[πΏπΏ(ππ + ππ) + πΏπΏ(ππ β ππ)]. 29 ππ exp(βππ|ππ|) iππ ; (ππ > 0). (A11) (A12) (A13) (A14) To prove the last identity note that tan ππ(π₯π₯) = π₯π₯βππ β ππ β² (π₯π₯)[1 + tan2 ππ(π₯π₯)] = 1βππ β 1βππ ππ β² (π₯π₯) = 1+(π₯π₯βππ)2 . (A15) Now, with reference to Eq.(A12) and the scaling theorem, the Fourier transform of ππ β² (π₯π₯) is ππ exp(βππ|ππ|), which must be the same as iππππ(ππ) in accordance with the differentiation theorem. Consequently, ππ(π₯π₯) = arctan(π₯π₯βππ) Fourier transforms to πΉπΉ(ππ) = (ππβiππ) exp(βππ|ππ|). Finally, we prove that the Fourier transform of comb(π₯π₯) = ββ ππ=ββ πΏπΏ(π₯π₯ β ππ) is the function comb(ππβ2ππ). The following proof takes the scaled comb function ππ(π₯π₯) = (1βππ) comb(π₯π₯βππ) = ππ ββ ππ=ββ πΏπΏ(π₯π₯ β ππππ), and proceeds to show that πΉπΉ(ππ) = limππββ βππ=βππ exp(βiππππππ) is equal to comb(ππππβ2ππ). 2ππ βππ ππ=βππ exp(βiππππππ) = exp(iππππππ) βππ=0 exp(βiππππππ) = exp(iππππππ) οΏ½ = exp[i(ππ+½)ππππ] β exp[βi(ππ+½)ππππ] exp(½iππππ) β exp(β½iππππ) β ββ ππ=ββ = ββ ππ=ββ = ββ ππ=ββ sinc(β ) approaches Ξ΄ function when Nβ β. = sin[(ππ+½)ππππ] sin(½ππππ) (β1)ππ sin[(ππ+½)ππ(ππβ2ππππβππ)] sin(ππππ) + ½ππ cos(ππππ)(ππβ2ππππβππ) (β1)ππ sin[(ππ+½)ππ(ππ β 2ππππ βππ)] ½ππ(β1)ππ (ππ β 2ππππβππ) 2(ππ+½)(ππβππ ) sin[(ππ+½)ππ(ππ β 2ππππβππ)] (ππ+½)ππ(ππ β 2ππππ βππ) (ππβππ ) 1 β exp[βi(2ππ+1)ππππ] 1 β exp(βiππππ) οΏ½ When Nββ, everything is negligible except the vicinity of k = 2mΟ /a, where the denominator vanishes. sinc(k) = sin(Ο k)/(Ο k). = ββ ππ=ββ(2ππ βππ )(ππ + ½)(ππ βππ) sinc[(ππ + ½)(ππ βππ)(ππ β 2ππππ βππ )] β ββ ππ=ββ(2ππ βππ ) πΏπΏ(ππ β 2ππππ βππ ) ππ = ββ ππ=ββ πΏπΏ οΏ½2ππ οΏ½ππ β The proof is now complete. 2ππππ ππ ππππ ππππ οΏ½οΏ½ = ββ ππ=ββ πΏπΏ οΏ½2ππ β πποΏ½ = comb οΏ½2ππ οΏ½. 30 (A16) Appendix B Shown in Fig.B1 is a semi-infinite substrate of refractive index πππ π , coated with a dielectric layer of refractive index ππ and thickness ππ. The normally-incident plane-wave of wavelength ππ0 οΏ½ and is linearly-polarized along the π¦π¦-axis, having electric and magnetic field amplitudes πΈπΈ0 ππ π»π»0 πποΏ½, where π»π»0 = πΈπΈ0 /ππ0 . Here ππ0 = οΏ½ππ0 βππ0 is the impedance of free space. Inside the dielectric οΏ½, whose magnitude relative to that of the layer, the forward-propagating beam has amplitude πΈπΈ1 ππ incident beam is found to be πΈπΈ1 πΈπΈ0 = 2β(1+ππ) . 1β[(ππβπππ π )β(ππ+πππ π )][(ππβ1)β(ππ+1)] exp(i4ππππππ βππ0 ) (B1) y Fig.B1. Monochromatic, linearly-polarized planewave is normally incident on a dielectric layer of thickness ππ and refractive index ππ, sitting atop a transparent, semi-infinite substrate of refractive index πππ π . Also shown are the reflected wave, the counter-propagating waves within the dielectric layer, and the transmitted wave into the substrate. οΏ½, The incident beamβs πΈπΈ-field amplitude is πΈπΈ0 ππ while that of the plane-wave propagating to the οΏ½. right inside the dielectric layer is πΈπΈ1 ππ d Reflected Transmitted Substrate (Index = ns) Incident n 0 x d The total electric and magnetic fields inside the dielectric layer are readily found to be πΈπΈπ¦π¦ (π₯π₯, π‘π‘) = πΈπΈ1 οΏ½exp οΏ½ π»π»π§π§ (π₯π₯, π‘π‘) = πππΈπΈ1 ππ0 i2ππππππ οΏ½exp οΏ½ ππ0 i2ππππππ ππ0 ππ β ππ οΏ½ + οΏ½ππ + πππ π οΏ½ exp οΏ½ π π ππ β ππ i4ππππππ οΏ½ exp οΏ½β i4ππππππ οΏ½ β οΏ½ππ + πππ π οΏ½ exp οΏ½ π π ππ0 ππ0 i2ππππππ οΏ½ exp οΏ½β ππ0 οΏ½οΏ½ exp(βiππππ), i2ππππππ ππ0 οΏ½οΏ½ exp(βiππππ). (B2) (B3) Considering that the polarization of the dielectric material is π·π·(π₯π₯, π‘π‘) = ππ0 (ππ2 β 1)π¬π¬(π₯π₯, π‘π‘), one may compute the time-averaged EM force-density acting on the dielectric layer as follows: β©ππ(π₯π₯, π‘π‘)βͺ = ½Re[πππ‘π‘ π·π·(π₯π₯, π‘π‘) × ππ0 π―π―β (π₯π₯, π‘π‘)] = ½ImοΏ½ππππ0 ππ0 (ππ2 β 1)πΈπΈπ¦π¦ π»π»π§π§β οΏ½ππ οΏ½. (B4) In what follows, we shall treat several special cases of the above problem. As a first example, let us consider a free-standing layer of refractive index ππ and thickness ππ = ππ0 β2ππ in the absence of the substrate. We will have ππ+1 i2ππππππ πΈπΈπ¦π¦ (π₯π₯, π‘π‘) = πΈπΈ0 οΏ½ 2ππ οΏ½ οΏ½exp οΏ½ π»π»π§π§ (π₯π₯, π‘π‘) = πππΈπΈ0 ππ+1 ππ0 ππ0 i2ππππππ οΏ½ 2ππ οΏ½ οΏ½exp οΏ½ β©ππ(π₯π₯, π‘π‘)βͺ = β ππ0 ππβ1 οΏ½ + οΏ½ππ+1οΏ½ exp οΏ½β ππβ1 i2ππππππ οΏ½ β οΏ½ππ+1οΏ½ exp οΏ½β ππππ0 πΈπΈ02 (ππ2 β1)2 2ππππ0 ππ0 οΏ½οΏ½ exp(βiππππ), i2ππππππ 4ππππππ sin οΏ½ ππ0 ππ0 οΏ½. οΏ½ ππ οΏ½οΏ½ exp(βiππππ), (B5) (B6) (B7) For our second example, we consider another free-standing dielectric layer of refractive index ππ and thickness ππ = ππ0 β4ππ, again without a substrate. We find 31 ππ+1 πΈπΈπ¦π¦ (π₯π₯, π‘π‘) = πΈπΈ0 οΏ½ππ2 +1οΏ½ οΏ½exp οΏ½ π»π»π§π§ (π₯π₯, π‘π‘) = πππΈπΈ0 π§π§0 ππ+1 i2ππππππ ππ0 ππβ1 οΏ½ β οΏ½ππ+1οΏ½ exp οΏ½β i2ππππππ οΏ½ππ2 +1οΏ½ οΏ½exp οΏ½ ππβ1 ππ0 οΏ½ + οΏ½ππ+1οΏ½ exp οΏ½β ππ0 ππ0 2 2ππππππ0 πΈπΈ02 ππ2 β1 β©ππ(π₯π₯, π‘π‘)βͺ = i2ππππππ ππ0 i2ππππππ 4ππππππ οΏ½ππ2 +1οΏ½ sin οΏ½ οΏ½οΏ½ exp(βiππππ), ππ0 οΏ½. οΏ½ ππ οΏ½οΏ½ exp(βiππππ), (B8) (B9) (B10) Finally, in the case of a single-layer antireflection coating on a substrate of refractive index πππ π = ππ2 , the coating layer must have refractive index ππ and thickness ππ = ππ0 β4ππ. We find πΈπΈπ¦π¦ (π₯π₯, π‘π‘) = πΈπΈ0 π»π»π§π§ (π₯π₯, π‘π‘) = (ππ+1) 2ππ πππΈπΈ0 (ππ+1) π§π§0 2ππ i2ππππππ οΏ½exp οΏ½ ππ0 i2ππππππ οΏ½exp οΏ½ ππ0 β©ππ(π₯π₯, π‘π‘)βͺ = β ππ β ππ2 οΏ½ β οΏ½ππ + ππ2 οΏ½ exp οΏ½β ππ β ππ2 i2ππππππ οΏ½ + οΏ½ππ + ππ2 οΏ½ exp οΏ½β ππ(ππ2 β1)2 ππ0 πΈπΈ02 2ππππ0 4ππππππ sin οΏ½ ππ0 ππ0 οΏ½οΏ½ exp(βiππππ), i2ππππππ ππ0 οΏ½. οΏ½ ππ οΏ½οΏ½ exp(βiππππ), (B11) (B12) (B13) Alternatively, we can denote the substrateβs refractive index by πππ π = ππ, in which case the antireflection coating layer will have a refractive index βππ and thickness ππ = ππ0 β4βππ. The force-density of Eq.(B13) must then be written as follows: β©ππ(π₯π₯, π‘π‘)βͺ = β ππ(ππβ1)2 ππ0 πΈπΈ02 2βππππ0 4ππ βπππ₯π₯ sin οΏ½ ππ0 οΏ½. οΏ½ ππ (B14) Integrating the force-density of Eq.(B14) over the thickness ππ of the layer now yields the total EM force acting on the antireflection coating as β©ππ0 βͺ = β (ππβ1)2 ππ0 πΈπΈ02 4ππ οΏ½. ππ (B15) The above expressions of EM force-density acting on various dielectric layers have been used in Sec.4 to arrive at the displacement function π’π’(π₯π₯, π‘π‘) under plane-wave illumination. 32 Appendix C As pointed out in Sec.7.2, computation of ππ0 (ππ) and ππ0 (ππ) from initial conditions involves a subtlety. Consider, for instance, the following expansion of the initial condition π’π’(π₯π₯, π‘π‘ = 0) in β² (π₯π₯) β³ (π₯π₯) terms of the eigen-functions π’π’ππ (π₯π₯) = π’π’ππ + iπ’π’ππ of an inhomogeneous medium: β π’π’(π₯π₯, π‘π‘ = 0) = β«ββ ππ0 (ππβ² )π’π’ππβ² (π₯π₯)ππππβ² . (C1) The first step in determining ππ0 (ππ) involves evaluating the (weighted) overlap integral between π’π’(π₯π₯, π‘π‘ = 0) and π’π’ππ (π₯π₯) over the interval [ππ, ππ], namely, ππ ππ(ππ) = β«ππ ππ(π₯π₯)π’π’(π₯π₯, π‘π‘ = 0)π’π’ππ (π₯π₯)ππππ β ππ = β«ββ ππ0 (ππβ² ) β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππβ² (π₯π₯)ππππ ππππβ² ππ ππ 2 (π₯π₯)ππππ β (π₯π₯)ππππ + ππ0 (βππ) β«ππ ππ(π₯π₯)π’π’ππ (π₯π₯)π’π’ππ = ππ0 (ππ) β«ππ ππ(π₯π₯)π’π’ππ ππ β²2 (π₯π₯) β³2 (π₯π₯) β² (π₯π₯)π’π’ β³ (π₯π₯)]ππππ = ππ0 (ππ) β«ππ ππ(π₯π₯)[π’π’ππ β π’π’ππ + 2iπ’π’ππ ππ ππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ . +ππ0 (βππ) β«ππ ππ(π₯π₯)[π’π’ππ + π’π’ππ (C2) In the second step, we evaluate the (weighted) overlap integral between π’π’(π₯π₯, π‘π‘ = 0) and β (π₯π₯) π’π’βππ (π₯π₯) = π’π’ππ over the same interval [ππ, ππ]. Noting that ππ(π₯π₯) and π’π’(π₯π₯, π‘π‘ = 0) are real-valued, we will have ππ β (π₯π₯)ππππ ππ(βππ) = ππ β (ππ) = β«ππ ππ(π₯π₯)π’π’(π₯π₯, π‘π‘ = 0)π’π’ππ β ππ β (π₯π₯)π’π’ β² (π₯π₯)ππππ = β«ββ ππ0 (ππβ² ) β«ππ ππ(π₯π₯)π’π’ππ ππππβ² ππ ππ ππ β (π₯π₯)π’π’ (π₯π₯)ππππ β2 (π₯π₯)ππππ = ππ0 (ππ) β«ππ ππ(π₯π₯)π’π’ππ + ππ0 (βππ) β«ππ ππ(π₯π₯)π’π’ππ ππ ππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ = ππ0 (ππ) β«ππ ππ(π₯π₯)[π’π’ππ + π’π’ππ ππ β²2 (π₯π₯) β³2 (π₯π₯) β² (π₯π₯)π’π’ β³ (π₯π₯)]ππππ . +ππ0 (βππ) β«ππ ππ(π₯π₯)[π’π’ππ β π’π’ππ β 2iπ’π’ππ ππ (C3) Combining Eqs.(C2) and (C3), we arrive at οΏ½ ππ β²2 (π₯π₯) β π’π’ β³2 (π₯π₯) + 2iπ’π’ β² (π₯π₯)π’π’ β³ (π₯π₯)]ππππ β«ππ ππ(π₯π₯)[π’π’ππ ππ ππ ππ ππ β²2 (π₯π₯) + π’π’ β³2 (π₯π₯)]ππππ β«ππ ππ(π₯π₯)[π’π’ππ ππ ππ0 (ππ) ππ(ππ) οΏ½οΏ½ οΏ½=οΏ½ οΏ½. ππ β²2 (π₯π₯) β π’π’ β³2 (π₯π₯) β 2iπ’π’ β² (π₯π₯)π’π’ β³ (π₯π₯)]ππππ β β«ππ ππ(π₯π₯)[π’π’ππ (βππ) ππ ππ ππ (ππ) ππ0 ππ ππ β²2 (π₯π₯) + π’π’ β³2 (π₯π₯)]ππππ β«ππ ππ(π₯π₯)[π’π’ππ ππ The values of ππ0 (ππ) and ππ0 (βππ) may now be found by solving Eq.(C4), namely, ππ0 (ππ) 1 οΏ½= οΏ½ ππππππ ππ0 (βππ) ×οΏ½ (C4) ππ ππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ β² (π₯π₯)π’π’ β³ (π₯π₯)ππππ β 2i β«ππ ππ(π₯π₯)π’π’ππ β π’π’ππ β«ππ ππ(π₯π₯)[π’π’ππ ππ ππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ β β«ππ ππ(π₯π₯)[π’π’ππ + π’π’ππ ππ(ππ) οΏ½, οΏ½οΏ½ ππ ππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ β² (π₯π₯)π’π’ β³ (π₯π₯)ππππ + 2i β«ππ ππ(π₯π₯)π’π’ππ β π’π’ππ β«ππ ππ(π₯π₯)[π’π’ππ ππ ππ β (ππ) ππ β²2 (π₯π₯) β³2 (π₯π₯)]πππ₯π₯ β β«ππ ππ(π₯π₯)[π’π’ππ + π’π’ππ 33 (C5) where the determinant (det) of the 2 × 2 matrix on the left-hand-side of Eq.(C4) is given by 2 ππ 2 ππ ππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ β² (π₯π₯)π’π’ β³ (π₯π₯)ππππ β²2 (π₯π₯) β³2 (π₯π₯)]ππππ β π’π’ππ + π’π’ππ ππππππ = οΏ½β«ππ ππ(π₯π₯)[π’π’ππ οΏ½ + 4 οΏ½β«ππ ππ(π₯π₯)π’π’ππ οΏ½ β οΏ½β«ππ ππ(π₯π₯)[π’π’ππ οΏ½ ππ ππ 2 ππ ππ β² (π₯π₯)π’π’ β³ (π₯π₯)ππππ β²2 (π₯π₯)ππππ β³2 (π₯π₯)ππππ = 4 οΏ½β«ππ ππ(π₯π₯)π’π’ππ οΏ½ β 4 οΏ½β«ππ ππ(π₯π₯)π’π’ππ οΏ½ οΏ½β«ππ ππ(π₯π₯)π’π’ππ οΏ½. ππ 2 (C6) Clearly, ππ0 (ππ) and ππ0 (βππ), which turn out to be a complex-conjugate pair, are uniquely determined via Eq.(C5) in terms of the overlap integral ππ(ππ) defined in the first line of Eq.(C2) and the various integrals involving the real and imaginary parts of the eigen-function π’π’ππ (π₯π₯). 34