This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Transfer Matrix Formulation of Scattering Theory in Two and Three Dimensions

Farhang Loran and Ali Mostafazadeh
Department of Physics, Isfahan University of Technology, Isfahan, 84156-83111, Iran
Departments of Mathematics and Physics, Koç University, 34450 Sarıyer, Istanbul, Turkey
Abstract

In one dimension one can dissect a scattering potential v(x)v(x) into pieces vi(x)v_{i}(x) and use the notion of the transfer matrix to determine the scattering content of v(x)v(x) from that of vi(x)v_{i}(x). This observation has numerous practical applications in different areas of physics. The problem of finding an analogous procedure in dimensions larger than one has been an important open problem for decades. We give a complete solution for this problem and discuss some of its applications. In particular we derive an exact expression for the scattering amplitude of the delta-function potential in two and three dimensions and a potential describing a slab laser with a surface line defect. We show that the presence of the defect makes the slab begin lasing for arbitrarily small gain coefficients.

Pacs numbers: 03.65.Nk, 42.25.Bs

I Introduction

Scattering of waves has been one of the most important subjects of research in physics for over a century. Most students of physics begin their study of this phenomenon in undergraduate courses of quantum mechanics where they learn time-independent scattering theory and its application in solving simple one-dimensional problems. In more advanced courses they are exposed to the 𝐒\mathbf{S}-matrix formulation of scattering theory which is of basic importance in quantum field theories and particle physics weinberg .

In one dimension, there is an alternative to the 𝐒\mathbf{S}-matrix, called the transfer matrix, that similarly to the 𝐒\mathbf{S}-matrix stores all the scattering information about the scattering potential. To give the definition of the transfer matrix, we recall that for a scattering potential vv, every solution of the time-independent Schrödinger equation,

ψ′′(x)+v(x)ψ(x)=k2ψ(x),-\psi^{\prime\prime}(x)+v(x)\psi(x)=k^{2}\psi(x), (1)

has the asymptotic form:

ψ(x)=A±eikx+B±eikxforx±.\psi(x)=A_{\pm}e^{ikx}+B_{\pm}e^{-ikx}~~{\rm for}~~x\to\pm\infty.

The transfer matrix of vv is the 2×22\times 2 matrix 𝐌\mathbf{M} satisfying

[A+B+]=𝐌[AB].\left[\begin{array}[]{c}A_{+}\\ B_{+}\end{array}\right]=\mathbf{M}\left[\begin{array}[]{c}A_{-}\\ B_{-}\end{array}\right]. (2)

It is easy to show that the reflection and transmission coefficients of vv are given by the entries of 𝐌\mathbf{M}, razavy ; jpa-2009a ; sanchez .

The main advantage of the transfer matrix over the 𝐒\mathbf{S}-matrix is that it satisfies an extremely useful composition rule which allows for the reduction of the scattering problem for vv to that of any set of constituent potentials v1,v2,,vnv_{1},v_{2},\cdots,v_{n} provided that v=jvjv=\sum_{j}v_{j} and the support IjI_{j} of vjv_{j} (i.e., the smallest interval outside which vjv_{j} vanishes) do not intersect except possibly at a single point. If 𝐌j\mathbf{M}_{j} is the transfer matrix of vjv_{j} and IjI_{j} lies to the left of Ij+1I_{j+1}, the transfer matrix of vv is given by

𝐌=𝐌n𝐌n1𝐌1.\mathbf{M}=\mathbf{M}_{n}\mathbf{M}_{n-1}\cdots\mathbf{M}_{1}. (3)

This relation makes the transfer matrix into an indispensable tool for dealing with a variety of scattering problems sanchez . Typical examples are the study of multilayer optical systems yeh-LPP-1 ; yeh-LPP-2 ; yeh-LPP-3 ; yeh-LPP-4 , acoustic waves accoustics-1 ; accoustics-2 ; accoustics-3 , transport phenomena in mesoscopic systems MK , photonic band structure calculations pendry-92 ; pendry-94 , optical potential engineering pra-2014bc-2015b-2 ; pra-2014bc-2015b-3 ; pra-2014bc-2015b-4 , etc. These applications are however limited to (effectively) one-dimensional physical systems. This is because the very definition of a multidimensional transfer matrix, that contains all the information about the scattering features of the potential and fulfils the composition property (3), has been lacking. The purpose of the present article is to introduce such a notion of transfer matrix and use it to develop an alternative to standard approach to scattering theory sakurai ; weinberg-qm that is capable of dealing with a variety of intricate scattering problems.

The basic conceptual framework for the present work is provided by the recent observation that the transfer matrix in one dimension can be expressed as

𝐌=𝐔(,),\mathbf{M}=\mathbf{U}(\infty,-\infty), (4)

where 𝐔(x,x0)\mathbf{U}(x,x_{0}) is the evolution operator for a fictitious non-unitary two-level quantum system with xx playing the role of the evolution parameter ap-2014 ; pra-2014a . This was originally motivated by the fact that the evolution operator of every quantum system has the same composition (group) property as the transfer matrix, i.e., (3). It turns that one can define a two-component state vector Ψ(x)\Psi(x) and an xx-dependent effective matrix Hamiltonian 𝐇(x){\mathbf{H}}(x) of the form

Ψ(x)\displaystyle\Psi(x) :=\displaystyle:= 12[eikx[ψ(x)ik1ψ(x)]eikx[ψ(x)+ik1ψ(x)]],\displaystyle\frac{1}{2}\left[\begin{array}[]{c}e^{-ikx}[\psi(x)-ik^{-1}\psi^{\prime}(x)]\\[3.0pt] e^{ikx}[\psi(x)+ik^{-1}\psi^{\prime}(x)]\end{array}\right], (7)
𝐇(τ)\displaystyle{\mathbf{H}}(\tau) :=\displaystyle:= v(x)2k[1e2ikxe2ikx1],\displaystyle\frac{v(x)}{2k}\left[\begin{array}[]{cc}1&e^{-2ikx}\\ -e^{2ikx}&-1\end{array}\right], (10)

such that the time-independent Schrödinger equation (1) becomes equivalent to the time-dependent Schrödinger equation

ixΨ(x)=𝐇(x)Ψ(x),i\partial_{x}\Psi(x)=\mathbf{H}(x)\Psi(x), (11)

with xx playing the role of “time,” ap-2014 .

Clearly, there is an infinity of choices for Ψ(x)\Psi(x) (and the corresponding 𝐇(x)\mathbf{H}(x)) that allow for expressing (1) in the form (11), FV-1958 ; jmp-1998 . Note however that among these, (7) is the only choice that fulfils the asymptotic boundary conditions:

Ψ(x)[A±B±]forx±.\Psi(x)\to\left[\begin{array}[]{c}A_{\pm}\\ B_{\pm}\end{array}\right]~~{\rm for}~~x\to\pm\infty. (12)

This relation is the key to the identification of the transfer matrix 𝐌\mathbf{M} with 𝐔(,)\mathbf{U}(-\infty,\infty), where 𝐔(x,x0)\mathbf{U}(x,x_{0}) is the evolution operator for the Hamiltonian (10), i.e., the operator satisfying

Ψ(x)=𝐔(x,x0)Ψ(x0).\Psi(x)=\mathbf{U}(x,x_{0})\Psi(x_{0}). (13)

Equation (4) is a direct consequence of (2), (12), and (13) for x0x_{0}\to-\infty and xx\to\infty.

Another important consequence of (7) is that the corresponding effective Hamiltonian, namely (10), vanishes for the values of xx where v(x)=0v(x)=0. We can use this observation to give a derivation of the composition rule for transfer matrices. Let us denote by 𝐇j(x)\mathbf{H}_{j}(x) and 𝐔j(x,x0)\mathbf{U}_{j}(x,x_{0}) the effective Hamiltonian and the evolution operator for the constituent potentials vjv_{j}, respectively. Suppose that the support of vjv_{j} is given by Ij=[aj,aj+1]I_{j}=[a_{j},a_{j+1}] for some real numbers aja_{j} and aj+1a_{j+1}. Then for all xIjx\notin I_{j}, 𝐇j(x)=0\mathbf{H}_{j}(x)=0. This in turn implies that whenever the interval (x0,x)(x_{0},x) does not intersect IjI_{j}, we have 𝐔j(x,x0)=𝐈\mathbf{U}_{j}(x,x_{0})=\mathbf{I}, where 𝐈\mathbf{I} is the 2×22\times 2 identity matrix. In particular,

Mj=𝐔j(,)=𝐔j(aj+1,aj)=𝐔(aj+1,aj),M_{j}=\mathbf{U}_{j}(\infty,-\infty)=\mathbf{U}_{j}(a_{j+1},a_{j})=\mathbf{U}(a_{j+1},a_{j}), (14)

where the last equality follows from the fact that vj(x)v_{j}(x) is the restriction of v(x)v(x) to IjI_{j}. Next, we recall that because IjI_{j} lies to the left of Ij+1I_{j+1}, a1a2a3an+1a_{1}\leq a_{2}\leq a_{3}\leq\cdots\leq a_{n+1}. This together with (4), (13), and (14) imply

𝐔(,)=𝐔(an+1,an)𝐔(an,an1)𝐔(a2,a1).\mathbf{U}(\infty,\infty)=\mathbf{U}(a_{n+1},a_{n})\mathbf{U}(a_{n},a_{n-1})\cdots\mathbf{U}(a_{2},a_{1}).

In light of (4) and (14), this equation is equivalent to (3).

The idea of expressing the transfer matrix of one-dimensional scattering theory in terms of the evolution operator for a two-level quantum system has revealed an intriguing manifestation of geometric phases in semiclassical scattering jpa-2014a . It has also been used to devise a method of computing higher order corrections to semiclassical scattering calculations jpa-2014b and a generalization of the composition rule (3) to situations where the constituting potentials vjv_{j} have overlapping support ap-2015 . In this article, we use this idea as a guiding principle for developing a transfer matrix formulation of the scattering theory in two and three dimensions. Since the basic difficulty that has hindered the progress towards solving this problem is already present in two dimensions, we give a more thorough treatment of scattering theory in two-dimensions. In order to focus on the conceptual developments and their concrete applications, we give the technical details of most of the calculations in the appendices.

II Transfer matrix in 2D

Consider the Schrödinger equation,

[x2y2+v(x,y)]ψ(x,y)=k2ψ(x,y),\left[-\partial_{x}^{2}-\partial_{y}^{2}+v(x,y)\right]\psi(x,y)=k^{2}\psi(x,y), (15)

where vv is a real- or complex-valued scattering potential and (x,y)(x,y) are cartesian coordinates in the plane 2\mathbb{R}^{2}. Let us identify the xx-axis with the scattering axis, use 𝐫\mathbf{r} to denote the position vector associated with (x,y)(x,y), and take the Fourier transform of (15) with respect to yy. This gives

[x2+v(x,ip)]ψ~(x,p)=ω(p)2ψ~(x,p),\left[-\partial_{x}^{2}+v(x,i\partial_{p})\right]\tilde{\psi}(x,p)=\omega(p)^{2}\tilde{\psi}(x,p), (16)

where ψ~(x,p):=𝑑yeipyψ(x,y)\tilde{\psi}(x,p):=\int_{-\infty}^{\infty}dye^{-ipy}\psi(x,y) is the Fourier transform of ψ(x,y)\psi(x,y) over yy and ω(p):=k2p2\omega(p):=\sqrt{k^{2}-p^{2}}. We identify v(x,ip)v(x,i\partial_{p}) with the integral operator defined by v(x,ip)ϕ(p):=12π𝑑qv~(x,pq)ϕ(q)v(x,i\partial_{p})\phi(p):=\frac{1}{2\pi}\int_{-\infty}^{\infty}dq\,\tilde{v}(x,p-q)\phi(q), where ϕ\phi is a test function.

For the scattering problem, we are interested in the solutions of (16) that tend to a plane wave for x±x\to\pm\infty. These exist provided that |p|k|p|\leq k and |v(x,y)||v(x,y)| has a sufficiently rapid asymptotic decay rate. We therefore demand that as r:=x2+y2r:=\sqrt{x^{2}+y^{2}}\to\infty, |v(x,y)||v(x,y)| tends to zero at such a rate that every solution of (15) has the following asymptotic form for x±x\to\pm\infty:

12πkk𝑑peipy[A±(p)eiω(p)x+B±(p)eiω(p)x].\frac{1}{2\pi}\int_{-k}^{k}dp\,e^{ipy}\left[A_{\pm}(p)e^{i\omega(p)x}+B_{\pm}(p)e^{-i\omega(p)x}\right].

Here A±A_{\pm} and B±B_{\pm} are coefficient functions vanishing outside the interval [k,k][-k,k].

We identify the transfer matrix of the potential vv with the 2×22\times 2 matrix operator 𝐌\mathbf{M} satisfying (2). As we show in the sequel, this notion of transfer matrix not only fulfills the composition rule (3), but also possesses complete information about the scattering properties of vv.

We begin our analysis of the properties of 𝐌\mathbf{M} by introducing the following two-dimensional generalization of the two-component state vector (7) and effective Hamiltonian operator (10).

Ψ(x,p)\displaystyle\Psi(x,p) :=12[eiω(p)x[ψ~(x,p)iω(p)1ψ~(x,p)]eiω(p)x[ψ~(x,p)+iω(p)1ψ~(x,p)]],\displaystyle:=\frac{1}{2}\left[\begin{array}[]{c}e^{-i\omega(p)x}[\tilde{\psi}(x,p)-i\omega(p)^{-1}\tilde{\psi}^{\prime}(x,p)]\\[3.0pt] e^{i\omega(p)x}[\tilde{\psi}(x,p)+i\omega(p)^{-1}\tilde{\psi}^{\prime}(x,p)]\end{array}\right], (19)
𝐇(x,p)\displaystyle\boldsymbol{\mathbf{H}}(x,p) :=12ω(p)eiω(p)x𝝈3v(x,ip)𝓚eiω(p)x𝝈3,\displaystyle:=\frac{1}{2\omega(p)}e^{-i\omega(p)x\boldsymbol{\sigma}_{3}}v(x,i\partial_{p})\,\boldsymbol{\mathcal{K}}\,e^{i\omega(p)x\boldsymbol{\sigma}_{3}}, (20)

where ψ~(x,p):=xψ~(x,p)\tilde{\psi}^{\prime}(x,p):=\partial_{x}\tilde{\psi}(x,p), 𝝈i\boldsymbol{\sigma}_{i} are the Pauli matrices, and 𝓚:=𝝈3+i𝝈2\boldsymbol{\mathcal{K}}:=\boldsymbol{\sigma}_{3}+i\boldsymbol{\sigma}_{2}. Ψ(x,p)\Psi(x,p) has the following two important properties:

  1. 1.

    It satisfies ixΨ(x,p)=𝐇(x,p)Ψ(x,p)i\partial_{x}\Psi(x,p)=\boldsymbol{\mathbf{H}}(x,p)\Psi(x,p), which is equivalent to (16). In other words, Ψ(x,p)=𝐔(x,x0)Ψ(x0,p)\Psi(x,p)=\boldsymbol{\mathbf{U}}(x,x_{0})\Psi(x_{0},p), where 𝐔(x,x0)\boldsymbol{\mathbf{U}}(x,x_{0}) is the evolution operator for 𝐇(x,p)\boldsymbol{\mathbf{H}}(x,p).

  2. 2.

    It tends to [A±(p)B±(p)]\left[\begin{array}[]{c}A_{\pm}(p)\\ B_{\pm}(p)\end{array}\right] for x±x\to\pm\infty.

By virtue of these properties and Eq. (2), we can express 𝐌\mathbf{M} in the form

𝐌\displaystyle\mathbf{M} =\displaystyle= 𝐔(,)=𝒯expi𝐇(x,p)dx\displaystyle\boldsymbol{\mathbf{U}}(\infty,-\infty)=\mathscr{T}\exp\int_{-\infty}^{\infty}\!\!-i\boldsymbol{\mathbf{H}}(x,p)dx
=\displaystyle= 𝐈i𝑑x1𝐇(x1,p)\displaystyle\mathbf{I}-i\int_{-\infty}^{\infty}\!\!\!dx_{1}\boldsymbol{\mathbf{H}}(x_{1},p)
+(i)2𝑑x2x2𝑑x1𝐇(x2,p)𝐇(x1,p)+,\displaystyle+(-i)^{2}\int_{-\infty}^{\infty}\!\!\!dx_{2}\int_{-\infty}^{x_{2}}\!\!\!dx_{1}\boldsymbol{\mathbf{H}}(x_{2},p)\boldsymbol{\mathbf{H}}(x_{1},p)+\cdots,

where 𝒯\mathscr{T} is the “time-ordering” operation weinberg-qm with xx playing the role of “time.”

According to (20), for each xx satisfying v(x,y)=0v(x,y)=0, we have 𝐇(x,p)=0\boldsymbol{\mathbf{H}}(x,p)=0. Similarly to the case of one dimension, we can employ this property of 𝐇(x,p)\boldsymbol{\mathbf{H}}(x,p) together with Eq. (II) to establish the composition rule (3). This rule holds true if IjI_{j} do not have overlapping projections onto the scattering (xx-) axis except possibly in a discrete set of points. For example, we can slice the support of vv along the xx-axis, label the slices by IjI_{j} (with IjI_{j} to the left of Ij+1I_{j+1} as depicted in Fig. 1), and take

vj(𝐫)={v(𝐫)for𝐫Ij,0for𝐫Ij,v_{j}(\mathbf{r})=\left\{\begin{array}[]{cc}v(\mathbf{r})&{\rm for}~\mathbf{r}\in I_{j},\\ 0&{\rm for}~\mathbf{r}\notin I_{j},\end{array}\right.

where 𝐫:=(x,y)\mathbf{r}:=(x,y).

Refer to caption
Figure 1: Support II of a finite-range potential (the enclosed region) sliced along the xx-axis.

In order to see how 𝐌\mathbf{M} relates to the scattering amplitude of the potential vv, we consider a left-incident wave and write the corresponding scattered wave in the form

ψscat(𝐫)=Θ(x)ψ(𝐫)+Θ(x)ψ+(𝐫)asr,\psi_{\rm scat}(\mathbf{r})=\Theta(-x)\psi_{-}(\mathbf{r})+\Theta(x)\psi_{+}(\mathbf{r})~~~~{\rm as}~r\to\infty, (22)

where Θ\Theta is the Heaviside step function, i.e.,

Θ(x):={1forx0,0forx<0,\Theta(x):=\left\{\begin{array}[]{ccc}1&{\rm for}&x\geq 0,\\ 0&{\rm for}&x<0,\end{array}\right.

and ψ(𝐫)\psi_{-}(\mathbf{r}) and ψ+(𝐫)\psi_{+}(\mathbf{r}) are respectively the reflected and transmitted waves given by

ψ±(𝐫):=12πkk𝑑pT±(p)eipye±iω(p)xforr,\displaystyle\psi_{\pm}(\mathbf{r}):=\frac{1}{2\pi}\int_{-k}^{k}dp\,T_{\pm}(p)\,e^{ipy}e^{\pm i\omega(p)x}~~{\rm for}~r\to\infty, (23)
T+(p):=A+(p)A(p),T(p):=B(p).\displaystyle T_{+}(p):=A_{+}(p)-A_{-}(p),~~~~~~~~~T_{-}(p):=B_{-}(p). (24)

The functions T±(p)T_{\pm}(p) yield the momentum representation of the reflected and transmitted waves ψ±(𝐫)\psi_{\pm}(\mathbf{r}).

For a left-incident plane wave with a definite momentum kk along the positive xx-axis, we have

A(p)=2πδ(p),\displaystyle A_{-}(p)=2\pi\delta(p), B+(p)=0,\displaystyle B_{+}(p)=0, (25)

where δ\delta stands for the Dirac delta-function. Substituting (25) in (2) and using (24), we can express T±(p)T_{\pm}(p) in terms of the entries MijM_{ij} of 𝐌\mathbf{M} according to

T+(p)=2π(M111M12M221M21)δ(p),\displaystyle T_{+}(p)=2\pi(M_{11}-1-M_{12}M_{22}^{-1}M_{21})\delta(p), (26)
T(p)=2πM221M21δ(p).\displaystyle T_{-}(p)=-2\pi M_{22}^{-1}M_{21}\delta(p).

Next, we write the scattered wave in its standard form,

ψscat(𝐫)=ikreikrf(θ)asr,\psi_{{\rm scat}}(\mathbf{r})=\sqrt{\frac{i}{kr}}\,e^{ikr}f(\theta)~~~~{\rm as}~r\to\infty, (27)

where (r,θ)(r,\theta) are the polar coordinates of 𝐫\mathbf{r}, and f(θ)f(\theta) is the scattering amplitude of the potential adhikari-86 . A careful analysis of the right-hand side of (23) gives a closed form asymptotic expression for ψ±(𝐫)\psi_{\pm}(\mathbf{r}) (See Appendix A.) Using this in (22) and comparing the result with (27), we obtain

f(θ)=iω(p)T±(p)2π=ik|cosθ|2πT±(ksinθ),f(\theta)=-\frac{i\omega(p)T_{\pm}(p)}{\sqrt{2\pi}}=-\frac{ik|\cos\theta|}{\sqrt{2\pi}}\,T_{\pm}(k\sin\theta), (28)

where ±:=sgn(cosθ)\pm:={\rm sgn}(\cos\theta).

Equations (26) and (28) show that 𝐌\mathbf{M} includes all the information about the scattering properties of the potential vv. In other words, the solution of the scattering problem can be achieved by evaluating 𝐌\mathbf{M} using (II) and computing T±(p)T_{\pm}(p) and f(θ)f(\theta) as given by (26) and (28).

Because the effective Hamiltonian 𝐇(x,p)\mathbf{H}(x,p) scales with the potential vv, Eq. (II) provides a convenient method of perturbative solution of the scattering problem (similarly to its one-dimensional analog pra-2014a ). We have checked that if we perform the first order perturbation theory using our approach, we obtain the known expression for the scattering amplitude in the first Born approximation Lapidus-1982 . Here we do not wish to discuss the utility of our method in perturbation theory. Instead, we demonstrate its application in providing an exact (non-perturbative) solution of the old and difficult problem of delta-function potential in two dimensions. The perturbation series solution of the Lippmann-Schwinger equation sakurai for the delta-function potential in two dimensions is plagued with the emergence of infinities. The same problem arises in the study of the corresponding spectral problem and has led to the development of intricate renormalization schemes that amount to the introduction of a length scale for the problem thorn ; exner ; henderson-1997 ; henderson-1998 ; cabo-1998 ; phillips ; rajeev-99 ; frederico ; Camblong-2001a ; camblong-2002 . The problem of divergences also arises in the standard treatment of the delta-function potential in 3D. As we show in the sequel, our approach avoids this problem in both two and three dimensions.

III Delta-function potential in 2D

Consider

v(x,y)=𝔷δ(x)δ(y),v(x,y)=\mathfrak{z}\,\delta(x)\delta(y), (29)

where 𝔷\mathfrak{z} is a real or complex coupling constant. Then 𝐇=𝔷ω(p)1δ(x)δ(ip)𝓚/2\mathbf{H}=\mathfrak{z}\,\omega(p)^{-1}\delta(x)\delta(i\partial_{p})\boldsymbol{\mathcal{K}}/2 and (II) gives

𝐌=𝐈i𝔷2ω(p)δ(ip)𝓚.\displaystyle\mathbf{M}=\mathbf{I}-\frac{i\mathfrak{z}}{2\omega(p)}\,\delta(i\partial_{p})\boldsymbol{\mathcal{K}}. (30)

As we show in Appendix B, the calculation of T±(p)T_{\pm}(p) is more conveniently carried out by substituting (30) in (2) and making use of (24) and (25). This gives T±(p)=2i𝔷/(4+i𝔷)ω(p)T_{\pm}(p)=-2i\mathfrak{z}/(4+i\mathfrak{z})\omega(p) which together with (28) yield the following exact expression for the scattering amplitude of (29).

f(θ)=2π𝔷4+i𝔷.\displaystyle f(\theta)=-\sqrt{\frac{2}{\pi}}\frac{\mathfrak{z}}{4+i\mathfrak{z}}. (31)

The first Born approximation corresponds to the linear approximation of this equation, i.e., f(θ)𝔷/22πf(\theta)\approx-\mathfrak{z}/2\sqrt{2\pi}, which agrees with the results of Ref. Lapidus-1982 ; Camblong-2001a .

According to (31), f(θ)f(\theta) diverges for 𝔷=4i\mathfrak{z}=4i. This marks a zero-width resonance that is associated with what mathematicians call a spectral singularity prl-2009 . In optical applications pra-2011a , vv is related to the complex relative permittivity ε\varepsilon of the medium according to v=k2(1ε)v=k^{2}(1-\varepsilon). In particular, (29) corresponds to a thin wire of nonmagnetic material that is aligned along the zz-axis and described by ε=1+iζδ(x)δ(y)\varepsilon=1+i\zeta\delta(x)\delta(y) for some real parameter ζ\zeta. Denoting the unit vector along the positive zz-axis by 𝐞z\mathbf{e}_{z}, we can easily check that

𝐄(x,y,z,t)=eicktψ(x,y)𝐞z\mathbf{E}(x,y,z,t)=e^{-ickt}\psi(x,y)\mathbf{e}_{z} (32)

solves Maxwell’s equations for this system provided that ψ\psi satisfies (15) for the delta-function potential (29) with 𝔷=iζk2\mathfrak{z}=-i\zeta k^{2}. In terms of ζ\zeta, the condition for the emergence of a spectral singularity takes the form k2=4/ζk^{2}=-4/\zeta. This shows that the wire begins lasing pra-2011a for k=2/ζk=2/\sqrt{-\zeta} whenever ζ<0\zeta<0 (i.e., it is made of gain material), and functions as a coherent perfect absorber CPA ; CPA-Science ; longhi-1 ; longhi-3 for k=2/ζk=2/\sqrt{\zeta} whenever ζ>0\zeta>0 (it is made of lossy material).

IV A slab laser with a surface line defect

As a simple application of the composition property of our transfer matrix, we address the scattering problem for electromagnetic waves interacting with an infinite planar slab of optically active material with a delta-function line defect on one of its faces as depicted in Fig. 2.

Refer to caption
Figure 2: (Color online) A slab of thickness LL with a surface line defect shown by the thick blue line.

Identifying the location of the slab and the line defect respectively with 0xL0\leq x\leq L and x=y=0x=y=0 (zz-axis) and considering the linearly polarized waves of the form (32), we can reduce the problem to the two-dimensional scattering problem for the potential vδ+v~v_{\delta}+\tilde{v}, where vδv_{\delta} is the delta-function potential (31) which represents the defect, and v~\tilde{v} is a barrier potential of the form

v~(x,y):={𝔷~for0xL,0otherwise,\tilde{v}(x,y):=\left\{\begin{array}[]{cc}\tilde{\mathfrak{z}}&{\rm for}~0\leq x\leq L,\\ 0&{\rm otherwise},\end{array}\right. (33)

which represents the slab pra-2011a . Here 𝔷~:=k2(1ε)\tilde{\mathfrak{z}}:=k^{2}(1-\varepsilon), ε\varepsilon is the complex relative permittivity of the slab, and LL is its thickness.

Because, along the xx-axis, the support of vδv_{\delta} is to the left of that of v~\tilde{v}, the transfer matrix 𝐌\mathbf{M} of vδ+v~v_{\delta}+\tilde{v} is given by 𝐌=𝐌~𝐌δ\mathbf{M}=\tilde{\mathbf{M}}\mathbf{M}_{\delta}, where 𝐌~\tilde{\mathbf{M}} and 𝐌δ\mathbf{M}_{\delta} are respectively the transfer matrices of v~\tilde{v} and vδv_{\delta}.

We have already computed 𝐌δ\mathbf{M}_{\delta}. It has the form (30). We give the details of the calculation of 𝐌~\tilde{\mathbf{M}} in Appendix C. Its entries have the following expression.

M~11(ω)=M~22(ω)=[cos(𝔫Lω)+i𝔫+sin(𝔫Lω)]eiωL,\displaystyle\tilde{M}_{11}(\omega)=\tilde{M}_{22}(-\omega)=[\cos({\mathfrak{n}}L\omega)+i{\mathfrak{n}}_{+}\sin({\mathfrak{n}}L\omega)]e^{-i\omega L}, (34)
M~12(ω)=M~21(ω)=i𝔫sin(𝔫Lω)eiωL,\displaystyle\tilde{M}_{12}(\omega)=\tilde{M}_{21}(-\omega)=i{\mathfrak{n}}_{-}\sin({\mathfrak{n}}L\omega)e^{-i\omega L},

where 𝔫:=1𝔷~/ω2=(ε1)(k/ω)2+1{\mathfrak{n}}:=\sqrt{1-\tilde{\mathfrak{z}}/\omega^{2}}=\sqrt{(\varepsilon-1)(k/\omega)^{2}+1} and 𝔫±:=(𝔫±𝔫1)/2{\mathfrak{n}}_{\pm}:=({\mathfrak{n}}\pm{\mathfrak{n}}^{-1})/2. For p=0p=0, ω=k\omega=k, and 𝔫{\mathfrak{n}} coincides with the complex refractive index of the slab, namely ε\sqrt{\varepsilon}. In this case, (34) gives the known result obtained in one dimension pra-2011a ; pra-2015c .

Having determined 𝐌δ\mathbf{M}_{\delta} and 𝐌~\tilde{\mathbf{M}}, we can use the expression 𝐌=𝐌~𝐌δ\mathbf{M}=\tilde{\mathbf{M}}\mathbf{M}_{\delta} together with (2), (24) and (25) to compute T±T_{\pm}. As we show in Appendix D, this gives

ωT\displaystyle\omega\,T_{-} =2πk[𝒳(k)1]δ(p)i𝔷𝒳(k)𝒳(ω)𝒴(k),\displaystyle=2\pi k[\mathcal{X}(k)-1]\delta(p)-\frac{i\mathfrak{z}\,\mathcal{X}(k)\mathcal{X}(\omega)}{\mathcal{Y}(k)}, (35)
ωT+\displaystyle\omega\,T_{+} =2πk[1M~22(k)1]δ(p)i𝔷𝒳(k)𝒴(k)M~22(ω),\displaystyle=2\pi k\left[\frac{1}{\tilde{M}_{22}(k)}-1\right]\delta(p)-\frac{i\mathfrak{z}\,\mathcal{X}(k)}{\mathcal{Y}(k)\tilde{M}_{22}(\omega)},

where

𝒳(ω)\displaystyle\mathcal{X}(\omega) :=1M~21(ω)M~22(ω)=2(e2i𝔫Lω+𝔫1𝔫+1)(𝔫+1)𝒵(ω),\displaystyle:=1-\frac{\tilde{M}_{21}(\omega)}{\tilde{M}_{22}(\omega)}=\frac{2\left(e^{-2i{\mathfrak{n}}L\omega}+\frac{{\mathfrak{n}}-1}{{\mathfrak{n}}+1}\right)}{({\mathfrak{n}}+1)\mathcal{Z}(\omega)},
𝒴(k)\displaystyle\mathcal{Y}(k) :=2+i𝔷π0k𝒳(ω)dωk2ω2,\displaystyle:=2+\frac{i\mathfrak{z}}{\pi}\int_{0}^{k}\frac{\mathcal{X}(\omega)d\omega}{\sqrt{k^{2}-\omega^{2}}},
𝒵(ω)\displaystyle\mathcal{Z}(\omega) :=e2i𝔫Lω(𝔫1𝔫+1)2.\displaystyle:=e^{-2i{\mathfrak{n}}L\omega}-\left(\frac{{\mathfrak{n}}-1}{{\mathfrak{n}}+1}\right)^{2}.

Substituting (35) in (28), we find the scattering amplitude for the system.

In the absence of the line defect, 𝔷=0\mathfrak{z}=0, and the condition for the emergence of a spectral singularity takes the form M~22(k)=0\tilde{M}_{22}(k)=0 which is equivalent to 𝒵(k)=0\mathcal{Z}(k)=0. This equation turns out to be equivalent to the laser threshold and phase conditions for the slab pra-2011a . It is easy to see that the inclusion of the line defect yields other means of achieving a spectral singularity. In this case a spectral singularity occurs also for M~22(ω)=0\tilde{M}_{22}(\omega)=0 which is equivalent to 𝒵(ω)=0\mathcal{Z}(\omega)=0.

Recalling that the gain coefficient gg for the slab satisfies g=2kκg=-2k\kappa, where κ\kappa is the imaginary part of the refractive index ε\sqrt{\varepsilon}, we can follow the approach of pra-2011a to determine the threshold gain from 𝒵(ω)=0\mathcal{Z}(\omega)=0. Using the fact that for generic active material, |κ|η1|\kappa|\ll\eta-1, where η\eta is the real part of ε\sqrt{\varepsilon}, we then find the following expression for the threshold gain coefficient

g=4η2sin2θηLln[η2sin2θ+|cosθ|η21].g=\frac{4\sqrt{\eta^{2}-\sin^{2}\theta}}{\eta L}\ln\left[\frac{\sqrt{\eta^{2}-\sin^{2}\theta}+|\cos\theta|}{\sqrt{\eta^{2}-1}}\right]. (36)

This is precisely the laser threshold condition for the oblique TE modes of the slab in the absence of the defect, if we identify θ\theta with the incidence angle pra-2015a . Notice however that here we have a normally incident TE wave and θ\theta corresponds to the direction of the scattered wave.

Fig. 3 shows a plot of the threshold gain coefficient (36) as a function of θ\theta for a generic value of η>1\eta>1.

Refer to caption
Figure 3: Graph of the threshold gain coefficient gg as a function of θ\theta for a generic slab laser with a surface line defect.

As seen from this graph, gg attains its maximum value at θ=0,180\theta=0,180^{\circ}. If θ\theta deviates from these values, gg decreases monotonically until it reaches its minimum value, which is zero, at θ=±90\theta=\pm 90^{\circ}. This shows that if the slab consists of a gain medium, irrespective of how small its gain coefficient is, the defect will make it radiate laser light with a wavevector in some direction. This is actually easy to justify. The presence of the defect makes the waves scatter in different directions. Those that are scattered through angles closer to θ=±90\theta=\pm 90^{\circ} follow a longer optical path within the slab and lase more easily.

V Transfer matrix in 3D

First we introduce some useful conventions and notions. We take the scattering axis of our system to be aligned along the zz-axis of a cartesian coordinate system, label the position vector 𝐫\mathbf{r} with its cartesian coordinates (x,y,z)(x,y,z), and set ρ:=(x,y)\vec{\rho}:=(x,y) and (ρ,z):=(x,y,z)(\vec{\rho},z):=(x,y,z). In general overarrows identify two-dimensional vectors lying in the xx-yy plane.

Now, consider a potential v:3v:\mathbb{R}^{3}\to\mathbb{C} such that for z±z\to\pm\infty every solution of the Schrödinger equation,

[x2y2z2+v(x,y,z)]ψ(x,y,z)=k2ψ(x,y,z),\left[-\partial_{x}^{2}-\partial_{y}^{2}-\partial_{z}^{2}+v(x,y,z)\right]\psi(x,y,z)=k^{2}\psi(x,y,z), (37)

has the asymptotic form

1(2π)2𝒟kd2peipρ[A±(p)eiω(p)z+B±(p)eiω(p)z],\displaystyle\frac{1}{(2\pi)^{2}}\int_{\mathscr{D}_{k}}d^{2}p\,e^{i\vec{p}\cdot\vec{\rho}}\left[A_{\pm}(\vec{p})e^{i\omega(\vec{p})z}+B_{\pm}(\vec{p})e^{-i\omega(\vec{p})z}\right],

where 𝒟k\mathscr{D}_{k} stands for the disc {p2||p|k}\big{\{}\vec{p}\in\mathbb{R}^{2}\big{|}|\vec{p}|\leq k\big{\}}, A±A_{\pm} and B±B_{\pm} are coefficient functions vanishing outside 𝒟k\mathscr{D}_{k}, and ω(p):=k2p 2\omega(\vec{p}):=\sqrt{k^{2}-\vec{p}^{\>2}}. We can use (2) to define a transfer matrix 𝐌\mathbf{M} for the potential vv, and check that it satisfies (II) if we replace (x,p)(x,p) by (z,p)(z,\vec{p}), and define 𝐇(z,p)\mathbf{H}(z,\vec{p}) according to

𝐇(z,p):=12ω(p)eiω(p)z𝝈3v(ip,z)𝓚eiω(p)z𝝈3.\boldsymbol{\mathbf{H}}(z,\vec{p}):=\frac{1}{2\omega(\vec{p})}e^{-i\omega(\vec{p})z\boldsymbol{\sigma}_{3}}v(i\vec{\partial}_{p},z)\,\boldsymbol{\mathcal{K}}\,e^{i\omega(\vec{p})z\boldsymbol{\sigma}_{3}}. (38)

Here p:=(px,py)\vec{\partial}_{p}:=(\partial_{p_{x}},\partial_{p_{y}}), (px,py)(p_{x},p_{y}) are cartesian coordinates of p\vec{p}, and v(ip,z):=v(ipx,ipy,z)v(i\vec{\partial}_{p},z):=v(i\partial_{p_{x}},i\partial_{p_{y}},z). For details of the derivation of (38), see Appendix E.

According to (38), 𝐇(z,p)\boldsymbol{\mathbf{H}}(z,\vec{p}) vanishes for the values of zz for which v(x,y,z)=0v(x,y,z)=0. This together with Eq. (II) ensure that 𝐌\mathbf{M} satisfies the composition property (3).

In order to relate 𝐌\mathbf{M} to the scattering amplitude for the potential vv, we introduce T±(p)T_{\pm}(\vec{p}) by changing pp to p\vec{p} in (24). These have the same properties as their two-dimensional analog. We can express them in terms of 𝐌\mathbf{M} by enforcing (2) and setting

A(p)=4π2δ(px)δ(py),\displaystyle A_{-}({\vec{p}})=4\pi^{2}\delta(p_{x})\delta(p_{y}), B+(p)=0.\displaystyle B_{+}({\vec{p}})=0.

In terms of T±(p)T_{\pm}(\vec{p}) the reflected (ψ\psi_{-}) and transmitted (ψ+\psi_{+}) waves take the form

ψ±(𝐫)=14π2𝒟kd2pT±(p)eipρe±iω(p)zasr.\psi_{\pm}(\mathbf{r})=\frac{1}{4\pi^{2}}\int_{\mathscr{D}_{k}}d^{2}p\,T_{\pm}(\vec{p})e^{i\vec{p}\cdot\vec{\rho}}e^{\pm i\omega(\vec{p})z}~~~~{\rm as}~r\to\infty. (39)

These in turn give the scattered wave ψscat\psi_{\rm scat} according to (22) provided that we replace xx by zz.

Next, we recall that ψscat(𝐫)=f(ϑ,φ)eikr/r\psi_{\rm scat}(\mathbf{r})=f(\vartheta,\varphi)\,e^{ikr}/r, where ff is the scattering amplitude and (r,ϑ,φ)(r,\vartheta,\varphi) are the spherical coordinates of 𝐫\mathbf{r} with ϑ\vartheta and φ\varphi respectively denoting the polar and azimuthal angles. Comparing this expression with the one given by (22) and (39) and using an argument given in Appendix F, we find

f(ϑ,φ)=ik|cosϑ|2πT±(ksinϑcosφ,ksinϑsinφ),f(\vartheta,\varphi)=-\frac{ik|\cos\vartheta|}{2\pi}\,T_{\pm}(k\sin\vartheta\cos\varphi,k\sin\vartheta\sin\varphi), (40)

where ±:=sgn(cosϑ)\pm:={\rm sgn}(\cos\vartheta).

VI Delta-function potential in 3D

For the delta-function potential

v(x,y,z)=𝔷δ(x)δ(y)δ(z),v(x,y,z)=\mathfrak{z}\,\delta(x)\delta(y)\delta(z), (41)

Eq. (38) gives 𝐇(z,p)=𝔷ω(p)1δ(z)δ(ipx)δ(ipy)𝓚/2\boldsymbol{\mathbf{H}}(z,{\vec{p}})=\mathfrak{z}\,\omega(\vec{p})^{-1}\delta(z)\delta(i\partial_{p_{x}})\delta(i\partial_{p_{y}})\boldsymbol{\mathcal{K}}/2. Substituting this relation in (II) and making use of (2), (25), and (40), we find

𝐌=𝐈i𝔷2ω(p)𝓚δ(ipx)δ(ipy),\displaystyle\mathbf{M}=\mathbf{I}-\frac{i\mathfrak{z}}{2\omega({\vec{p}})}\,\boldsymbol{\mathcal{K}}\delta(i\partial_{p_{x}})\delta(i\partial_{p_{y}}), (42)
f(ϑ,φ)=𝔷4π+ik𝔷.\displaystyle f(\vartheta,\varphi)=\frac{-\mathfrak{z}}{4\pi+ik\mathfrak{z}}. (43)

In view of the latter relation the scattering length, i.e., ξ:=limk0f\xi:=-\lim_{k\to 0}f, is given by ξ=𝔷/4π\xi=\mathfrak{z}/4\pi. Therefore, we can express (43) as f=(ξ1+ik)1f=-(\xi^{-1}+ik)^{-1}. This coincides with Eq. (9) of Ref. rajeev-99 , which is obtained for general unitary interactions under the assumption that the range of the interaction is much smaller than the wavelength of the incident wave.

The standard physical application of delta-function potential is in the study of point defects in condensed matter systems. For a charged particle moving in a conducting medium, the presence of a point defect yields a non-vanishing resistivity, which is related to the scattering cross section of the defect |f(ϑ,φ)|2|f(\vartheta,\varphi)|^{2}. Eq. (43) suggests that it is an isotropic function of space that is inversely proportional to k2+μ2k^{2}+\mu^{2} where μ:=4π/|𝔷|\mu:=4\pi/|\mathfrak{z}|.

VII Concluding Remarks

Scattering of waves in one dimension is most conveniently studied in the framework of the transfer matrix formulation of scattering theory. This is mainly because of the practical advantages stemming from the composition property of the transfer matrix. Despite the enormous interest in the scattering of waves in two- and three-dimensions, there has been no successful attempt at developing a general transfer matrix formulation of scattering theory in these dimensions. In this article, we achieve this goal by introducing a genuine multi-dimensional notion of transfer matrix which enjoys the niceties of its one-dimensional analog.

In order to elucidate the practical aspects of our approach, we have explored its application in addressing the scattering problem for delta-function potentials in two and three dimensions. It is known that a proper treatment of the spectral problem for delta-function potential in two dimensions requires a coupling constant renormalization which makes it run with kk and introduces a length scale for the problem rajeev-99 ; camblong-2002 . As far as its physical applications are concerned, the delta-function potential is an abstraction for a physical potential whose range is much smaller than the relevant wavelengths. This provides an explanation for the length scale arising from the renormalization procedure and applies for the delta-function potential in three dimensions. It is remarkable that the direct application of our approach to delta function potentials yield finite results that do not rely on any approximation or renormalization schemes.

The composition property of the transfer matrix allows for extending our results to potentials consisting of finitely many delta-functions. In two dimensions, these describe finite arrays of parallel thin wires whose scattering properties are of direct interest in the study of metamaterials meta1 ; meta2 and nanophotonic devices nano1 ; nano2 ; nano3 . In three dimensions, delta-function potentials model point-like defects. This marks the significance of our approach for dealing with scattering problems in condensed matter physics. It is also not difficult to envisage a generalization of this approach that is capable of treating scattering of waves inside a waveguide. This will be of importance in the study of transport phenomena in mesoscopic systems MK .

To demonstrate the effectiveness of our approach in dealing with a realist physical system, we have used it to describe the scattering of electromagnetic ways from an optically active slab that has a surface line defect. Our results allow for the determination of the laser threshold condition for such a slab and reveal the fact that for any positive value of the gain coefficient the presence of the defect makes the slab begin emitting laser light in some direction. It is remarkable that this phenomenon does not depend on the strength of the delta-function interaction modeling the defect.

The main impact of the multidimensional transfer matrix formulation of scattering theory that we outline in this article will be in its numerical implementations. This is not only because of the composition property of the transfer matrix, which allows for the reduction of the scattering problem for a given potential to those of its truncations, but also because any discretization of this approach would yield matrix representations for the entries of the transfer matrix operator that could be substituted in (26) to determine T±T_{\pm} and consequently the scattering amplitude.

Acknowledgments

This project was initiated during F. L.’s visit to Koç University Aug. 11-24, 2015. We are indebted to the Turkish Academy of Sciences (TÜBA) for providing the financial support which made this visit possible and to Teoman Turgut for helpful discussions. This work has been supported by the Scientific and Technological Research Council of Turkey (TÜBİTAK) in the framework of the project no: 112T951 and by TÜBA.

Appendices

In the following we give the details of various calculations that we refer to in the text of this article.

A: Expressing ψ±(𝐫)\psi_{\pm}(\mathbf{r}) and f(θ)f(\theta) in terms of T±T_{\pm} in 2D

We wish to determine the asymptotic behavior of the factor eipye±iω(p)xe^{ipy}e^{\pm i\omega(p)x} appearing in (23) for rr\to\infty. First, we use (x,y)=(rcosθ,rsinθ)(x,y)=(r\cos\theta,r\sin\theta) and the identity

cos(α±β)=cosαcosβsinαsinβ,\cos(\alpha\pm\beta)=\cos\alpha\cos\beta\mp\sin\alpha\sin\beta, (44)

to express this quantity in the form

eipye±iω(p)x=eikrcos(θpθ±),e^{ipy}e^{\pm i\omega(p)x}=e^{ikr\cos(\theta_{p}-\theta^{\pm})}, (45)

where θ+:=θ\theta^{+}:=\theta, θ:=πθ\theta^{-}:=\pi-\theta, and θp\theta_{p} is the real variable determined by the conditions

|θp|π2,\displaystyle|\theta_{p}|\leq\frac{\pi}{2}, p=ksinθp,\displaystyle p=k\sin\theta_{p}, ω(p)=kcosθp.\displaystyle\omega(p)=k\cos\theta_{p}. (46)

Next, we recall from Ref. Lapidus-1982 that, for rr\to\infty,

eikrcosϕ2πkrm=0ϵmimcos(krmπ2π4)cos(mϕ),e^{ikr\cos\phi}\to\sqrt{\frac{2}{\pi kr}}\sum_{m=0}^{\infty}\epsilon_{m}i^{m}\cos\left(kr-\frac{m\pi}{2}-\frac{\pi}{4}\right)\cos(m\phi), (47)

where ϵ0=1\epsilon_{0}=1 and ϵm=2\epsilon_{m}=2 for m1m\geq 1. If we use the identity (44) to expand the factor cos(krmπ2π4)\cos\left(kr-\frac{m\pi}{2}-\frac{\pi}{4}\right) in (47), split the series into the sum of the terms with odd and even values of mm, and notice that for every integer \ell, sin(π)=cos(π+π2)=0\sin(\ell\pi)=\cos(\ell\pi+\frac{\pi}{2})=0 and sin(π+π2)=cos(π)=(1)\sin(\ell\pi+\frac{\pi}{2})=\cos(\ell\pi)=(-1)^{\ell}, we have

m=0ϵmimcos(ξmπ2)cos(mϕ)\displaystyle\sum_{m=0}^{\infty}\epsilon_{m}i^{m}\cos\left(\xi-\frac{m\pi}{2}\right)\cos(m\phi)
=cosξ=0ϵ2cos(2ϕ)+isinξ=0ϵ2+1cos[(2+1)ϕ]\displaystyle=\cos\xi\sum_{\ell=0}^{\infty}\epsilon_{2\ell}\cos(2\ell\phi)+i\sin\xi\sum_{\ell=0}^{\infty}\epsilon_{2\ell+1}\cos[(2\ell+1)\phi]
=eiξ2m=0ϵmcos(mϕ)+eiξ2m=0(1)mϵmcos(mϕ)\displaystyle=\frac{e^{i\xi}}{2}\sum_{m=0}^{\infty}\epsilon_{m}\cos(m\phi)+\frac{e^{-i\xi}}{2}\sum_{m=0}^{\infty}(-1)^{m}\epsilon_{m}\cos(m\phi)
=π[eiξδ(ϕ)+eiξδ(ϕ+π)],\displaystyle=\pi\left[e^{i\xi}\delta(\phi)+e^{-i\xi}\delta(\phi+\pi)\right], (48)

where we have introduced ξ:=krπ4\xi:=kr-\frac{\pi}{4} and employed the identities:

m=0ϵmcos(mϕ)=n=einϕ=2πδ(ϕ),\displaystyle\sum_{m=0}^{\infty}\epsilon_{m}\cos(m\phi)=\sum_{n=-\infty}^{\infty}e^{in\phi}=2\pi\delta(\phi),
m=0(1)mϵmcos(mϕ)=2πδ(ϕ+π).\displaystyle\sum_{m=0}^{\infty}(-1)^{m}\epsilon_{m}\cos(m\phi)=2\pi\delta(\phi+\pi).

Next, we change the variable of integration in (23) to θp\theta_{p} and use (45) – (48) to evaluate this integral. Note that for both θ[π2,π2]\theta\in[-\frac{\pi}{2},\frac{\pi}{2}] and θ[π2,3π2]\theta\in[\frac{\pi}{2},\frac{3\pi}{2}], we have θ±[π2,π2]\theta^{\pm}\in[-\frac{\pi}{2},\frac{\pi}{2}]. Therefore in both cases it is the first delta-function in Eq.(48) that contributes to the integral. The result is

ψ±(𝐫)=kcosθ±eiπ/4eikr2πkrT±(ksinθ±).\psi_{\pm}(\mathbf{r})=\frac{k\cos\theta^{\pm}e^{-i\pi/4}e^{ikr}}{\sqrt{2\pi kr}}T_{\pm}(k\sin\theta^{\pm}). (49)

Substituting this relation in (22) and comparing the result with (27), we obtain

f(θ)=ikcos(θ±)2πT±(ksinθ±)for±cosθ>0.f(\theta)=\frac{-ik\cos(\theta^{\pm})}{\sqrt{2\pi}}T_{\pm}(k\sin\theta^{\pm})~~~{\rm for}~\pm\cos\theta>0.

It is not difficult to see that we can express this relation as Eq. (28).

B: Calculation of T±T_{\pm} for delta-function potential in 2D

For v(x,y)=𝔷δ(x)δ(y)v(x,y)=\mathfrak{z}\delta(x)\delta(y), the effective Hamiltonian takes the form 𝐇(x,p)=𝔷ω(p)1δ(x)δ(ip)𝓚/2\mathbf{H}(x,p)=\mathfrak{z}\,\omega(p)^{-1}\delta(x)\delta(i\partial_{p})\,\boldsymbol{\mathcal{K}}/2. Substituting this equation in (II) and noting that 𝓚2=𝟎\boldsymbol{\mathcal{K}}^{2}=\mathbf{0}, we obtain Eq. (30), which reads 𝐌=1i𝔷ω(p)1δ(ip)𝓚/2\mathbf{M}=1-i\mathfrak{z}\omega(p)^{-1}\delta(i\partial_{p})\boldsymbol{\mathcal{K}}/2.

Next, we recall that

𝓚=𝝈3+i𝝈2=[1111],\displaystyle\boldsymbol{\mathcal{K}}=\boldsymbol{\sigma}_{3}+i\boldsymbol{\sigma}_{2}=\left[\begin{array}[]{cc}1&1\\ -1&-1\end{array}\right], (52)
δ(ip)ϕ(p)=12πkkdpϕ(p)=:ϕ~,\displaystyle\delta(i\partial_{p})\phi(p)=\frac{1}{2\pi}\int_{-k}^{k}dp\,\phi(p)=:\tilde{\phi}, (53)

where ϕ:\phi:\mathbb{R}\to\mathbb{C} is a test function such that ϕ(p)=0\phi(p)=0 for |p|>k|p|>k. If we insert (30) and (25) in (2) and make use of (52) and (53), we find

[A+(p)0]=[A(p)B(p)]i𝔷2ω(p)(1+B~)[11]\left[\begin{array}[]{c}A_{+}(p)\\ 0\end{array}\right]=\left[\begin{array}[]{c}A_{-}(p)\\ B_{-}(p)\end{array}\right]-\frac{i\mathfrak{z}}{2\omega(p)}(1+\tilde{B}_{-})\left[\begin{array}[]{c}1\\ -1\end{array}\right]

In light of this equation and (24),

T±(p)=i𝔷(B~+1)2ω(p).T_{\pm}(p)=-\frac{i\mathfrak{z}(\tilde{B}_{-}+1)}{2\omega(p)}. (54)

Because T(p)=B(p)T_{-}(p)=B_{-}(p), this relation in particular implies B(p)=i𝔷(B~+1)/2ω(p)B_{-}(p)=-i\mathfrak{z}(\tilde{B}_{-}+1)/2\omega(p). Applying δ(ip)\delta(i\partial_{p}) to both sides of this equation and employing Eq. (53) and

δ(ip)ω1(p)=12πkkdpk2p2=12,\delta(i\partial_{p})\omega^{-1}(p)=\frac{1}{2\pi}\int_{-k}^{k}\frac{dp}{\sqrt{k^{2}-p^{2}}}=\frac{1}{2}, (55)

we obtain B~=i𝔷/(4+i𝔷)\tilde{B}_{-}=-i\mathfrak{z}/(4+i\mathfrak{z}). Inserting this in (54) gives (31).

C: Calculation of the transfer matrix for a slab without a defect

Consider a scattering potential vv in two-dimensions that does not depend on yy. Then Eq. (16) and the effective Hamiltonian take the form

[x2+v(x)]ψ~(x,p)=ω2(p)ψ~(x,p),\displaystyle\left[-\partial_{x}^{2}+v(x)\right]\tilde{\psi}(x,p)=\omega^{2}(p)\tilde{\psi}(x,p),
𝐇(x,p)=12ω(p)v(x)eiω(p)x𝝈3𝓚eiω(p)x𝝈3,\displaystyle\mathbf{H}(x,p)=\frac{1}{2\omega(p)}v(x)\,e^{-i\omega(p)x\boldsymbol{\sigma}_{3}}\boldsymbol{\mathcal{K}}\,e^{i\omega(p)x\boldsymbol{\sigma}_{3}},

respectively. For the potential given by (33), these reduce to their one-dimensional analogs for a rectangular barrier potential pra-2015c provided that we replace kk with ω(p)\omega(p) in the latter. In particular, making this change in Eq. (32) of Ref. pra-2015c , we obtain Eq. (34). Alternatively, one can perform the xx-dependent transformation of two-component state vectors: Ψ(x)Ψˇ(x):=eiω(p)x𝝈3Ψ(x)\Psi(x)\to\check{\Psi}(x):=e^{i\omega(p)x\boldsymbol{\sigma}_{3}}\Psi(x) that gives rise to an xx-independent effective Hamiltonian 𝐇ˇ(p)\check{\mathbf{H}}(p), obtain the corresponding evolution operator, and transform back to compute 𝐌\mathbf{M}. This also leads to Eq. (34).

D: Calculation of T±T_{\pm} for a slab with a line defect

As we noted in the text of the paper, by virtue of the composition property (2), the transfer matrix for the potential vδ+v~v_{\delta}+\tilde{v} has the form

𝐌=𝐌~𝐌δ,\mathbf{M}=\tilde{\mathbf{M}}\,\mathbf{M}_{\delta}, (56)

where 𝐌δ\mathbf{M}_{\delta} and 𝐌~\tilde{\mathbf{M}} are respectively given by (30) and (34). Substituting (30) in (56) and enforcing (2) and (25), we find

A+(p)\displaystyle A_{+}(p) =\displaystyle= M~11(ω)[2πδ(p)i𝔷2ω(B~+1)]+\displaystyle\tilde{M}_{11}(\omega)\left[2\pi\delta(p)-\frac{i\mathfrak{z}}{2\omega}(\tilde{B}_{-}+1)\right]+ (57)
M~12(ω)[B(p)+i𝔷2ω(B~+1)],\displaystyle\tilde{M}_{12}(\omega)\left[B_{-}(p)+\frac{i\mathfrak{z}}{2\omega}(\tilde{B}_{-}+1)\right],
B(p)\displaystyle B_{-}(p) =\displaystyle= M~221(k)M~21(k)2πδ(p)+\displaystyle-\tilde{M}_{22}^{-1}(k)\,\tilde{M}_{21}(k)2\pi\delta(p)+
i𝔷2ω(p)(B~+1)(M~221(ω)M~21(ω)1),\displaystyle\frac{i\mathfrak{z}}{2\omega(p)}(\tilde{B}_{-}+1)\left(\tilde{M}_{22}^{-1}(\omega)\,\tilde{M}_{21}(\omega)-1\right),

where we have employed (53). Next, we determine B~\tilde{B}_{-} by operating δ(ip)\delta(i\partial_{p}) on both sides of (D: Calculation of T±T_{\pm} for a slab with a line defect) and making use of (55) and (34). This gives B~+1=2𝒳(k)/𝒴(k)\tilde{B}_{-}+1=2\mathcal{X}(k)/\mathcal{Y}(k) where 𝒳(k)\mathcal{X}(k) and 𝒴(k)\mathcal{Y}(k) are defined in the text of the paper. Substituting this expression for B~+1\tilde{B}_{-}+1 in (57) and (D: Calculation of T±T_{\pm} for a slab with a line defect) and using (24) and (34) we obtain (35).

E: Expressing 𝐌\mathbf{M} in terms of an evolution operator in 3D

Following the approach we pursued for treating the two-dimensional case, we first take the Fourier transform of both sides of the Schrödinger equation (37) over xx and yy. This gives

z2ψ~(p,z)+v(ip,z)ψ~(p,z)=ω(p)2ψ~(p,z),-\partial_{z}^{2}\tilde{\psi}({\vec{p}},z)+v(i{\vec{\partial}}_{p},z)\tilde{\psi}({\vec{p}},z)=\omega({\vec{p}})^{2}\tilde{\psi}({\vec{p}},z), (59)

where

ψ~(p,z)\displaystyle\tilde{\psi}({\vec{p}},z) :=\displaystyle:= 𝑑x𝑑yei(xpx+ypy)ψ(x,y,z)\displaystyle\int_{-\infty}^{\infty}dx\int_{-\infty}^{\infty}dy\,e^{-i(xp_{x}+yp_{y})}\psi(x,y,z)

is the Fourier transform of ψ(x,y,z)\psi(x,y,z) over xx and yy. Next, we introduce the two-component state vector

Ψ(z,p)\displaystyle\Psi(z,\vec{p}) :=12[eiω(p)z(ψ~(p,z)izψ~(x,p)ω(p))eiω(p)z(ψ~(p,z)+izψ~(p,z)ω(p))],\displaystyle:=\frac{1}{2}\left[\begin{array}[]{c}e^{-i\omega(\vec{p})z}\Big{(}\tilde{\psi}(\vec{p},z)-\displaystyle\frac{i\partial_{z}\tilde{\psi}(x,\vec{p})}{\omega(\vec{p})}\Big{)}\\[6.0pt] e^{i\omega(\vec{p})z}\Big{(}\tilde{\psi}(\vec{p},z)+\displaystyle\frac{i\partial_{z}\tilde{\psi}(\vec{p},z)}{\omega(\vec{p})}\Big{)}\end{array}\right],

and check that it has the same properties as its analogue in two-dimensions. In particular, (59) is equivalent to izΨ(z,p)=𝐇(z,p)Ψ(z,p)i\partial_{z}\Psi(z,\vec{p})=\mathbf{H}(z,\vec{p})\Psi(z,\vec{p}) for the effective Hamiltonian 𝐇(z,p)\mathbf{H}(z,\vec{p}) given by (38), and Ψ(z,p)[A±(p)B±(p)]\Psi(z,\vec{p})\to\left[\begin{array}[]{c}A_{\pm}(\vec{p})\\ B_{\pm}(\vec{p})\end{array}\right] for z±z\to\pm\infty. These together with (2) imply 𝐌=𝐔(,)\mathbf{M}=\mathbf{U}(\infty,-\infty), where 𝐔(z,z0)\mathbf{U}(z,z_{0}) is the evolution operator for the effective Hamiltonian 𝐇(z,p)\mathbf{H}(z,\vec{p}), i.e., 𝐔(z,z0):=𝒯expz0zi𝐇(z,p)dz\mathbf{U}(z,z_{0}):=\mathscr{T}\exp\int_{z_{0}}^{z}-i\mathbf{H}(z^{\prime},\vec{p})dz^{\prime}.

F: Expressing ψ±(𝐫)\psi_{\pm}(\mathbf{r}) and f(θ)f(\theta) in terms of T±T_{\pm} 3D

We wish to determine the asymptotic form of the factor eipρe±iω(p)ze^{i\vec{p}\cdot\vec{\rho}}e^{\pm i\omega(\vec{p})z} that appears in Eq. (39). First, we introduce θ+:=θ\theta^{+}:=\theta, θ:=πθ\theta^{-}:=\pi-\theta, and 𝐫±:=r(cosθ±,sinθ±cosϕ,sinθ±sinϕ){\bf r}_{\pm}:=r(\cos\theta^{\pm},\sin\theta^{\pm}\cos\phi,\sin\theta^{\pm}\sin\phi), where (r,θ,ϕ)(r,\theta,\phi) are the standard spherical coordinates, i.e., z=rcosθz=r\cos\theta and ρ=rsinθ(cosϕ,sinϕ)\vec{\rho}=r\sin\theta(\cos\phi,\sin\phi). Then, in view of 𝐩:=(ω(p),p){\bf p}:=(\omega(p),\vec{p}), it is not difficult to see that

eipρe±iω(p)z=ei𝐩𝐫±.e^{i\vec{p}\cdot\vec{\rho}}e^{\pm i\omega(\vec{p})z}=e^{i{\bf p}\cdot{\bf r}_{\pm}}. (60)

We can use the spherical wave expansion of plane waves Jackson to express the right-hand side of this equation in the form

ei𝐩𝐫±=4π=0ij(kr)m=Ym(θ±,ϕ)Ym(θp,ϕp),e^{i{\bf p}\cdot{\bf r}_{\pm}}=4\pi\sum_{\ell=0}^{\infty}i^{\ell}j_{\ell}(kr)\sum_{m=-\ell}^{\ell}Y_{\ell m}^{*}(\theta^{\pm},\phi)Y_{\ell m}(\theta_{p},\phi_{p}), (61)

where jj_{\ell} and YmY_{\ell m} are respectively the spherical Bessel functions and spherical harmonics, and the parameters θp[0,π2]\theta_{p}\in\left[0,\frac{\pi}{2}\right] and ϕp[0,2π)\phi_{p}\in[0,2\pi) are determined by

ω(p)=kcosθp,\displaystyle\omega({\vec{p}})=k\cos\theta_{p}, p=ksinθp(cosϕp,sinϕp).\displaystyle{\vec{p}}=k\sin\theta_{p}\left(\cos\phi_{p},\sin\phi_{p}\right). (62)

Next, we recall the following asymptotic expression for the spherical Bessel functions and the completeness relation for the spherical harmonics Jackson .

j(x)sin(xπ2)x=(i)eixieix2ixforx,j_{\ell}(x)\to\frac{\sin\left(x-\frac{\ell\pi}{2}\right)}{x}=\frac{(-i)^{\ell}e^{ix}-i^{\ell}e^{-ix}}{2ix}~~{\rm for}~~x\gg\ell, (63)
=0m=Ym(θ,ϕ)Ym(θp,ϕp)=δ(ϕpϕ)δ(cosθpcosθ).\sum_{\ell=0}^{\infty}\sum_{m=-\ell}^{\ell}Y_{\ell m}^{*}(\theta,\phi)Y_{\ell m}(\theta_{p},\phi_{p})=\delta(\phi_{p}-\phi)\delta(\cos\theta_{p}-\cos\theta).

Because Ym(πθ,π+ϕ)=(1)Ym(θ,ϕ)Y_{\ell m}(\pi-\theta,\pi+\phi)=(-1)^{\ell}Y_{\ell m}(\theta,\phi), the latter relation implies

=0m=(1)Ym(θ,ϕ)Ym(θp,ϕp)\displaystyle\sum_{\ell=0}^{\infty}\sum_{m=-\ell}^{\ell}(-1)^{\ell}Y_{\ell m}^{*}(\theta,\phi)Y_{\ell m}(\theta_{p},\phi_{p})
=δ(π+ϕpϕ)δ(cosθp+cosθ).\displaystyle\hskip 56.9055pt=\delta(\pi+\phi_{p}-\phi)\delta(\cos\theta_{p}+\cos\theta).~~~~~~ (64)

In light of (60)–(64), we have the asymptotic formula

eipρe±iω(p)z2iπkr[eikrδ(ϕpϕ)δ(cosθpcosθ±)\displaystyle e^{i\vec{p}\cdot\vec{\rho}}e^{\pm i\omega(\vec{p})z}\to-\frac{2i\pi}{kr}\Big{[}e^{ikr}\delta(\phi_{p}-\phi)\delta(\cos\theta_{p}-\cos\theta^{\pm})
eikrδ(ϕpϕ+π)δ(cosθp+cosθ±)]forr,\displaystyle-e^{-ikr}\delta(\phi_{p}-\phi+\pi)\delta(\cos\theta_{p}+\cos\theta^{\pm})\Big{]}~~{\rm for}~r\to\infty, (65)

which together with d2p=k2sinθpcosθpdθpdϕpd^{2}p=k^{2}\sin\theta_{p}\cos\theta_{p}d\theta_{p}d\phi_{p} allow us to evaluate the integral in (39). Because for both θ[0,π2]\theta\in[0,\frac{\pi}{2}] and θ[π2,π]\theta\in[\frac{\pi}{2},\pi], we have θ±[0,π2]\theta^{\pm}\in[0,\frac{\pi}{2}], in both cases it is the first term on the right-hand side of (65) that contributes to this integral. The result is ψ±(𝐫)=f(θ,ϕ)eikr/r\psi_{\pm}(\mathbf{r})=f(\theta,\phi)e^{ikr}/r, where

f(θ,ϕ):=i2πω(k)T±(k),f(\theta,\phi):=-\frac{i}{2\pi}\omega(\vec{k})T_{\pm}(\vec{k}), (66)

±:=sgn(cosθ)\pm:={\rm sgn}(\cos\theta), ω(k)=k|cosθ|\omega(\vec{k})=k\left|\cos\theta\right|, and k=ksinθ(cosϕ,sinϕ)\vec{k}=k\sin\theta\left(\cos\phi,\sin\phi\right). In view of (22), (66) is equivalent to (40).

References

  • (1) S. Weinberg, The Quantum Theory of Fields, Vol. I (Cambridge University Press, Cambridge, 1995).
  • (2) M. Razavy, Quantum Theory of Tunneling (World Scientific, Singapore, 2003).
  • (3) A.  Mostafazadeh and H. Mehri-Dehnavi, Spectral singularities, biorthonormal systems, and a two-parameter family of complex point interactions. J. Phys. A 42, 125303 (2009).
  • (4) L. L. Sánchez-Soto, J. J. Monzóna, A. G. Barriuso, and J. F. Carin~\tilde{\rm n}ena, The transfer matrix: A geometrical perspective. Phys. Rep. 513, 191 (2012).
  • (5) P. Yeh, A. Yariv, and C.-S. Hong, Electromagnetic propagation in periodic stratified media. I. General theory. J. Opt. Soc. Am. 67, 423 (1977).
  • (6) P. Pereyra, Resonant tunneling and band mixing in multichannel superlattices. Phys. Rev. Lett. 80, 2677 (1998).
  • (7) D. J. Griffiths and C. A. Steinke, Waves in locally periodic media. Am. J. Phys. 69, 137 (2001).
  • (8) P. Yeh, Optical Waves in Layered Media (Wiley, Hoboken, New Jersey, 2005).
  • (9) W. T. Thompson, Transmission of elastic waves through a stratified solid medium. J. Appl. Phys. 21, 89 (1950).
  • (10) D. Lévesque and L. J. Piché, A robust transfer matrix formulation for the ultrasonic response of multilayered absorbing media, J. Accous. Soc. Am. 92, 452 (1992).
  • (11) B. Hosten and M. Castaings, Transfer matrix of multilayered absorbing and anisotropic media. Measurements and simulations of ultrasonic wave propagation through composite materials. J. Accous. Soc. Am. 94, 1488 (1993).
  • (12) P. A. Mello and N. Kumar, Quantum Transport in Mesoscopic Systems (Oxford University Press, Oxford, 2004).
  • (13) J. B. Pendry and A. MacKinnon, Calculation of photon dispersion relations. Phys. Rev. Lett. 69, 2772 (1992).
  • (14) Pendry J. B. Photonic band structures. J. Mod. Opt. 41, 209 (1994).
  • (15) A. Mostafazadeh, Unidirectionally invisible potentials as local building blocks of all scattering potentials. Phys. Rev. A 90, 023833 (2014).
  • (16) A. Mostafazadeh, A. Addendum to Unidirectionally invisible potentials as local building blocks of all scattering potentials. Phys. Rev. A 90, 055803 (2014).
  • (17) A. Mostafazadeh, Active invisibility cloaks in one dimension. Phys. Rev. A 91, 063812 (2015).
  • (18) J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, Reading, Massachusetts, 1994).
  • (19) S. Weinberg, Lectures on Quantum Mechanics (Cambridge University Press, Cambridge, 2013).
  • (20) A. Mostafazadeh, A Dynamical formulation of one-dimensional scattering theory and its applications in optics. Ann. Phys. (N.Y.) 341, 77 (2014).
  • (21) A. Mostafazadeh, Transfer matrices as non-unitary S-matrices, multimode unidirectional invisibility, and perturbative inverse Scattering. Phys. Rev. A 89, 012709 (2014).
  • (22) H. Feshback and F. Villars, Elementary relativistic wave mechanics of spin 0 and spin 1/2 particles, Rev. Mod. Phys. 30, 24 (1958).
  • (23) A. Mostafazadeh, Two-component formulation of the Wheeler-DeWitt equation, J. Math. Phys. 39, 4499 (1998).
  • (24) A. Mostafazadeh, Adiabatic approximation, semiclassical scattering, and unidirectional invisibility. J. Phys. A 47, 125301 (2014).
  • (25) A. Mostafazadeh, Adiabatic series expansion and higher-order semiclassical approximations in scattering theory,. J. Phys. A 47, 345302 (2014).
  • (26) F. Loran and A. Mostafazadeh, Composition of transfer matrices for potentials with overlapping support. Ann. Phys. (NY) 359, 230 (2015).
  • (27) S. K. Adhikari, Quantum scattering in two dimensions. Amer. J. Phys. 54, 362 (1986).
  • (28) I. R. Lapidus, Quantum-mechanical scattering in two dimensions. Amer. J. Phys. 50, 45 (1982). Note that there is a factor of 4π4\pi error in Eqs. (33) and (36) of this reference. This is because the author has taken Eq. (30) from his Ref. [6] without noting that the latter uses a different convention for the definition of Green’s function.
  • (29) C. Thorn, Quark confinement in the infinite-momentum frame. Phys. Rev. D 19, 639 (1979).
  • (30) P. Exner and P. Seba, Point interactions in two and three dimensions as models of small scatterers, Physics Letters A 222, 1 (1996).
  • (31) R. J. Henderson and S. G. Rajeev, Renormalized path integral in quantum mechanics, J. Math. Phys. 38, 2171 (1997).
  • (32) R. J. Henderson and S. G. Rajeev, Renormalized contact potential in two dimensions, J. Math. Phys. 39, 749 (1998).
  • (33) A. Cabo, J. L. Lucio, and H. Mercado, On scale invariance and anomalies in quantum mechanics, Amer. J. Phys. 66, 240 (1998).
  • (34) D. R. Phillips, S. R. Beane, and T. D. Cohen, Nonperturbative regularization and renormalization: Simple examples from Nonrelativistic Quantum Mechanics, Ann. Phys. (N.Y.) 263, 255 (1998).
  • (35) S. G. Rajeev, Bound states in models of asymptotic freedom. Preprint arXiv:hep-th/9902025.
  • (36) T. Frederico, A. Delfino, and Lauro Tomio, Renormalization group invariance of quantum mechanics, Phys. Lett. B 481, 143 (2000).
  • (37) H. E. Camblong, L. N. Epele, H. Fanchiotti, and C. A. García Canal, Dimensional transmutation and dimensional regularization in quantum mechanics I. General theory. Ann. Phys. (N.Y.) 287, 14 (2001).
  • (38) H. E. Camblong and C. R. Ordóñez, Renormalized path integral for the two-dimensional delta-function interaction. Phys. Rev. A 65, 052123 (2002).
  • (39) A. Mostafazadeh, Spectral singularities of complex scattering potentials and infinite reflection and transmission coefficients at real energies. Phys. Rev. Lett. 102, 220402 (2009).
  • (40) A. Mostafazadeh, Optical spectral singularities as threshold resonances. Phys. Rev. A 83, 045801 (2011).
  • (41) Y. D. Chong, L. Cao, H. Ge, and A. D. Stone, Coherent perfect absorbers: time-reversed lasers. Phys. Rev. Lett. 105, 053901 (2010).
  • (42) W. Wan, Y. Chong, L. Ge, H. Noh, A. D. Stone, and H. Cao, Time-reversed lasing and interferometric control of absorption. Science 331, 889 (2011).
  • (43) S. Longhi, Backward lasing yields a perfect absorber. Physics 3, 61 (2010).
  • (44) S. Longhi, Coherent perfect absorption in a homogeneously broadened two-level medium. Phys. Rev. A 83, 055804 (2011).
  • (45) A. Mostafazadeh, Perturbative unidirectional invisibility. Phys. Rev. A 92, 023831 (2015).
  • (46) A. Mostafazadeh and M. Sarısaman, Lasing-threshold condition for oblique TE and TM modes, spectral singularities, and coherent perfect absorption. Phys. Rev. A 91, 043804 (2015).
  • (47) D. R. Smith, J. B. Pendry, and M. C. K. Wiltshire, Metamaterials and negative refractive index, Science 305, 788 (2004).
  • (48) V. M. Shalaev, Optical negative-index metamaterials, Nature Photonics 1, 41 (2007).
  • (49) O. Benson, Assembly of hybrid photonic architectures from nanophotonic constituents. Nature 480, 193-199 (2011).
  • (50) X. Guo, Y. Ma, Y. Wang, and L. Tong, Nanowire plasmonic waveguides, circuits and devices, Laser Photonics Rev. 7, 855 881 (2013).
  • (51) T. P. Sidiropoulos, et al. Ultrafast plasmonic nanowire lasers near the surface plasmon frequency, Nature Phys. 10, 870 876 (2014).
  • (52) J. D. Jackson, Classical Electrodynamics (Wiley, New York, 1975).