Open Access
19 August 2013 Exact analytical solution for fields in gradient index metamaterials with different loss factors in negative and positive refractive index segments
Mariana Dalarsson, Martin Norgren, Tatjana Asenov, Nebojsa Doncov, Zoran Jaksic
Author Affiliations +
Abstract
Gradient refractive index metamaterials are of interest for various applications of transformation optics. Wave propagation through gradient index metamaterials using an exact analytical approach is investigated. Composite materials containing constituents with negative real and positive real indexes of refraction are considered. An exact analytical solution for the field distribution is obtained for the sinusoidal spatial variation of complex effective permittivity and permeability along a fixed direction, under the assumption that the wave impedance remains spatially uniform across the structure. Loss factors in the constituent materials can be different from each other corresponding to the realistic situations. Temporal dispersion can be arbitrary subject to the physical limitations imposed by the Kramers-Kronig relations. A numerical model based on the Z -transform is developed to verify the analytical results. The approach can be applied to arbitrary periodic refractive index profiles using the Fourier series method.

1.

Introduction

Electromagnetic metamaterials (MM) are artificial composites with electromagnetic properties not readily found in nature. A special class of MMs is the negative refractive index metamaterials (NRM), artificial structures with negative phase velocity.1 A number of practical implementations of optical MM have been reported.2,3

NRM are typically produced using arrays of subwavelength “particles” with negative effective relative permittivity and permeability as their structural units. The first proposed NRM particles were split-ring resonators and nanowires, furnishing negative permeability and permittivity of their composites.3 They are well understood and extensively used in the microwave domain. However, other particles such as complementary split-ring resonators,4 cut-wire pairs/plate pairs,5 and double fishnets68 are also investigated. The first NRM were experimentally confirmed in 2001,9 and recently the experimental fishnet-type NRM in the visible range of frequencies have been fabricated and investigated.10

The properties of NRM, such as the negative index of refraction (and negative phase velocity), inverse Doppler effect, radiation tension instead of pressure, etc.11,12 resulted in a number of proposed applications. Among those we mention superlenses and hyperlenses that enable imaging far below the diffraction limit,13,14 resonant cavities, and waveguides with geometrical dimension orders of magnitude smaller than the operating wavelength15 as well as invisibility cloaks and generally transformation optics.16

Most studies consider structures with constant effective permittivity and permeability within the NRM part and abrupt interfaces with the surrounding regular materials [positive refractive index media (PRM)]. However, there is both theoretical and practical interest in NRM with spatially varying effective permittivity and permeability within the NRM structure and with gradual transition from the PRM to NRM and vice versa. Graded refractive index is interesting for transformation optics including hyperlenses17 and invisibility cloaks.18 Various other proposed applications of gradient refractive index (GRIN) metamaterials include beam shaping and directing, enhancement of nonlinear effects,19 superlenses,20 etc.

The first paper dedicated to GRIN NRM was published in 2005.21 Analytical approaches to graded index metamaterial structures are of special interest, since they ensure fast, simple, and direct route to the determination of the field distribution and the calculation of the scattering parameters within such materials.2228

The present study is a generalization of our previous work27,28 in a sense that it allows different loss factors in PRM and NRM segments. An assumption of uniform loss factors throughout the structure has been made in Refs. 2728.29. The possibility of choosing arbitrary loss factors in PRM and NRM, independent of each other, is essential for a realistic description of composites involving NRM as building blocks, since it is empirically well known that the losses in the NRM part are significantly higher than those in the PRM. Thus, in this paper, we present an exact analytical solution of Helmholtz equations for the propagation of electromagnetic waves through a periodic gradient-index PRM–NRM composite with most general loss factors in the two materials and with sinusoidal periodicity for the case of constant impedance throughout the structure. A comparison of the obtained analytical solution to the results of numerical simulation using a Z-transform based model is given.

2.

Problem Formulation

The geometry of the present problem is shown in Fig. 1. The electric field points to the y-direction and has the form E(r)=E(x)ey, while the magnetic field points to the z-direction and has the form H(r)=H(x)ez. The wave propagates along the x-axis. The spatial variation of the refractive index along the x-axis is described by a cosine function. An implementation of a GRIN metamaterial obtained following an approach similar to that found in Ref. 29 is shown in Fig. 2. For this purpose, a two-dimensional array of single split rings is deposited onto a dielectric substrate. The example is given for illustration purposes only, as the structures with varying dimensions may be any of metamaterial unit structures or “particles” (“atoms”). Also, periodicity does not have to be sinusoidal (and actually may be represented by any graded structure, as long as a Fourier series representation is valid) and the gradient itself may be along one, two, or all three axes.

Fig. 1

Propagation of a wave through a graded index structure with a cosine profile: a zeroth-order component of a Fourier series expansion for an arbitrary spatial dispersion.

JNP_7_1_073086_f001.png

Fig. 2

An example of practical implementation of a metamaterial with sinusoidal profile along the x-axis: a two-dimensional array of circular single split ring structures forming a graded metamaterial media. The substrate is dielectric and the spatial dispersion is obtained by varying the split ring dimensions and spacing.

JNP_7_1_073086_f002.png

The NRM structures are of importance for transformation optics,30 the most well-known example being the invisibility cloaks. For instance, optical carpet cloaks31 were reported with effective index gradient obtained by drilling hole arrays with varying geometry.32 Other important applications include optical and generally electromagnetic concentrators based on metamaterials, beam shapers, and beam steering devices, as well as different kinds of metamaterial lenses including the hyperlenses17 for the transformation of near field into the far field. Finally, an important application is gradient index circuitry utilizing metamaterial waveguides.33

We write the Helmholtz equations for E(x) and H(x)28

Eq. (1)

d2Edx21μdμdxdEdx+ω2μεE(x)=0,d2Hdx21εdεdxdHdx+ω2μεH(x)=0,
where ε=ε(ω,x) and μ=μ(ω,x) are the frequency-dependent dielectric permittivity and magnetic permeability, respectively. The spatial dependency of the functions ε(x) and μ(x) may be completely arbitrary. The only limit of applicability of the present approach, in terms of radiation wavelength versus the structural period of the metamaterial, is posed by the requirement that the effective medium approximation remains valid, i.e., that the structure periodicity is much smaller than the operating wavelength.

3.

Analytical Solutions of the Field Equations

Let us now consider an infinite and inhomogeneous periodic structure, where the real parts of the effective dielectric permittivity and magnetic permeability vary as a cosine function from positive values (right-handed material) to negative ones (left-handed material) and back again. The thickness a of the positive part is equal to that of the negative part. For the sake of simplicity, we assume an impedance-matched case where real parts of effective permittivities and permeabilities of the two materials at a given frequency have opposite signs and equal absolute values. Thus, to determine their values, we use the functions

Eq. (2)

μ(ω,x)=μ0μR(ω)cos(πxa)iμ0[μIR+μIL2+μIRμIL2cos(πxa)],

Eq. (3)

ε(ω,x)=ε0εR(ω)cos(πxa)iε0[εIR+εIL2+εIRεIL2cos(πxa)],
where ±εR(ω) and ±μR(ω) are the frequency-dependent real parts of the effective permittivities and permeabilities in the two materials, respectively, where the plus sign applies to PRM and minus sign to NRM. Furthermore, εIR(ω) and εIL(ω) are the imaginary parts of the effective permittivities of PRM and NRM, respectively, whereas μIR(ω) and μIL(ω) are the imaginary parts of the effective permeabilities of PRM and NRM, respectively. For a material to be passive, i.e., without gain, and to satisfy causality, the conditions must be satisfied that the imaginary parts of both the permittivity and permeability are positive. It should be noted here that, so far, the imaginary parts of the permittivity [εIR(ω) and εIL(ω)] and permeability [μIR(ω) and μIL(ω)] in PRM and NRM media, respectively, are not related to each other in any way and can be chosen arbitrarily except for the causality requirement.

According to the Kramers-Kronig (KK) relations that are fully valid for metamaterials34,35 (i.e., due to causality), the real and the imaginary parts of such complex media are related, but it is an integral interdependence which by no means stipulates that they have to follow the same trends in the relatively narrow range of frequencies typically observed in metamaterials. On the contrary, in a more general case, these parts of the dependencies can assume very different forms, up to the point of assuring a possibility to compensate unavoidable losses in the negative refractive index part through the introduction of active media,36 where KK relations still remain valid.

The apparent interdependence between the parameters of the two media is due to the mathematical properties of the transition function model, sinusoidal in this case. However, further away from the transition region, the imaginary parts of the permittivity [εIR(ω) and εIL(ω)] and permeability [μIR(ω) and μIL(ω)] in PRM and NRM media, respectively, can be chosen independently from each other. The reason for giving Eqs. (2) and (3) is just a need to formulate a simple mathematical model of gradual transition between NRM and PRM. In a more elaborate model, which will be the subject of our future work, this distinction between the two materials will become more manifest.

Unlike some other functional dependences of PRM–NRM transitions studied so far, for instance tanh(x) model,24 the sinusoidal model provides only for relatively slow transitions and it is not equally obvious where we are “far away from the transition region.” In continuous models, the freedom of choice of parameters e.g., the permittivity εIR(ω) and εIL(ω) as well as permeability μIR(ω) and μIL(ω) in PRM and NRM media, respectively, is an asymptotic statement. For example, tanh()1 is only an asymptotic constant. The fact is that in a continuous model there is everywhere a spatial dependency and interdependency of material parameters. But at some asymptotic points, we calibrate the spatially constant and frequency-dependent material parameters to correspond to the actual PRM and NRM media far away from the transition region which we desire that they have. For instance, Eqs. (2) and (3) at the maxima of the cosine function (in the middle of PRM), where πx/a=2nπ or x=2na (n=0,1,2,3,), give

μ(ω)=μ0μR(ω)iμ0[μIR+μIL2+μIRμIL2]=μ0[μR(ω)iμIR(ω)],ε(ω)=ε0εR(ω)iε0[εIR+εIL2+εIRεIL2]=ε0[εR(ω)iεIR(ω)],
while at the minima of the cosine function (in the middle of NRM), where πx/a=(2n+1)π or x=(2n+1)a (n=0,1,2,3,), give
μ(ω)=μ0μR(ω)iμ0[μIR+μIL2+μIRμIL2]=μ0[μR(ω)iμIL(ω)],ε(ω)=ε0εR(ω)iε0[εIR+εIL2+εIRεIL2]=ε0[εR(ω)iεIL(ω)].

Thus, we see that in the middle of the PRM, we have the imaginary parts of permittivity and permeability εIR(ω) and μIR(ω), respectively, while in the middle of the NRM, we have the imaginary parts of permittivity and permeability εIL(ω) and μIL(ω), respectively, and in Eqs. (2) and (3) these parameters can be chosen arbitrarily.

In order to obtain a constant wave impedance throughout the structure, we now introduce a further requirement that the real and imaginary parts of the effective permittivity and permeability satisfy the condition

Eq. (4)

μIR(ω)+μIL(ω)2μR(ω)i[μIR(ω)μIL(ω)]=εIR(ω)+εIL(ω)2εR(ω)i[εIR(ω)εIL(ω)]=β(ω),
where we again stipulate that the Kramers-Kronig relations remain valid, and we note that β(ω) is a complex number, as can easily be seen from Eq. (4).

The condition (4) is a restrictive mathematical requirement on the complex permittivity and permeability that reduces our analysis to a special case. A justification for the requirement (4) is based on the fact that both permittivity and permeability of many of the NRM structures reported until now can be described by Drude or Lorentz models, i.e., that ε(ω) and μ(ω) are strongly resonant and thus quite narrow. In order to obtain the widest possible frequency range of negative refractive index, it is then useful to have the best possible overlap between the ranges of negative values of the real parts of ε(ω) and μ(ω). In an ideal situation their dispersions in the resonant range would be therefore identical. On the other hand, in order to preserve causality, the imaginary parts of both ε(ω) and μ(ω) must be positive and their dispersions are determined by the real parts—thus the imaginary parts should also overlap. Actually, Eq. (4) can be seen as a condition, although too stringent, for the maximum bandwidth of Drude- or Lorentz-type resonant NRM structures.

Regarding the fabrication of NRM structures with largely overlapping complex ε(ω) and μ(ω), it was our reasoning that since NRM are artificial media they allow us, in principle, to design ε(ω) and μ(ω) separately, as is the case with prototypical NRM, the split ring resonators combined with the wire media.3,9 Using Eq. (4), we readily obtain

Eq. (5)

μ=μ0μIR+μIL2β[cos(πxa)iβ],ε=ε0εIR+εIL2β[cos(πxa)iβ].

Except for condition Eq. (4), our method allows for arbitrary temporal dispersion. Upon condition Eq. (4), the wave impedance Z=Z0Z(ω)=μ(ω,x)/ε(ω,x) becomes constant throughout the entire structure; as a result, there is no reflection on the graded interfaces between the two materials. Equation (1) with μ=μ(ω,x) and ε=ε(ω,x) given by Eqs. (2) and (3), respectively, is easily transformed into hypergeometric equations. If such an approach is used, the derivation is rather simple and does not require the use of any specialized software. The solutions to these equations involving the appropriate hypergeometric functions, for the given set of parameters, are reduced to relatively simple elementary functions. For this particular graded index structure, the exact solutions to the two differential equations (2) and (3) reduce to a remarkably simple form

Eq. (6)

E(x)=E0eκβxexp(iκaπsinπxa),H(x)=H0eκβxexp(iκaπsinπxa),
where E0 and H0 are the amplitudes of the electric and magnetic fields at the point x=0, respectively, and

Eq. (7)

κ=k+iα=ωcμRεR+iω2cεRμR(μILμIR).

From Eq. (7), one may draw a superficial conclusion that the loss of the entire structure stems from the imaginary part of the permeability of both PRM and NRM only. However, this is not the case, and Eq. (7) is just one way of writing the loss parameters of the structure. Using Eq. (4), it is possible to formulate the equivalent result to Eq. (7) where only the imaginary part of the permittivity of both PRM and NRM appears. Thus, Eq. (7) by no means implies that the contribution from the imaginary part of permittivity is marginal and can be ignored. It is just a matter of an arbitrary choice of presenting the quantities interrelated by Eq. (4).

Although both κ and β are complex numbers, it should be noted that the product κ×β is a real number. From Maxwell’s equation (1), the field amplitudes are related by E0=Z0Z(ω)H0. The exact solutions (6) are valid for arbitrary loss factors in NRM and PRM. In the PRM slab around the origin, i.e., in the limit x0 we readily obtain the time-domain fields of the form (since E(x,t)=Re[E(x)eiωt], etc.)

Eq. (8)

E(x,t)E0eγ1xcos(ωtkx)H(x,t)H0eγ1xcos(ωtkx)
as expected for a regular plane wave in a right-handed material slab. In order to investigate the behavior of the wave in the left-handed material slab, let us translate the solutions (6) by a distance a, i.e., let us assume xx+a. Using sin(α±π)=sinα in the limit x0, the time-domain fields acquire the form

Eq. (9)

E(x,t)E0eγ2(xa)cos[ωt(k)x],H(x,t)H0eγ2(xa)cos[ωt(k)x].

In the results shown in Eqs. (8) and (9) for the fields, the loss factors γ1 and γ2 are given by

Eq. (10)

γ1=ωcεRμRμIR,γ2=ωcεRμRμIL,
and they can be set freely by a suitable modeling of μIR(ω) and μIL(ω).

From the results in Eq. (8), it follows that the asymptotic wavevector in the right-handed material is kRHM=+kex, i.e., the wave propagates in the +x-direction. On the other hand, from the results (9), it follows that the asymptotic wavevector in the left-handed material is kLHM=kex, i.e., the wave propagates in the x-direction. However, the energy flux (the Poynting vector) is in the +x-direction in both media. Thus, in the limit of small x, we have the correct wave behavior in both the PRM and the NRM slabs. The wave changes the direction periodically along the periodic structure, which will be obvious from the figures in Sec. 5.

4.

Numerical Model of Metamaterials

A dispersive transmission line matrix (TLM) Z-transform model of the lossy MM-composite, described in Ref. 37, is used here to verify the analytical solution for gradient index metamaterials with arbitrary loss factor in PRM and NRM presented in Secs. 2 and 3. This model follows the notation used in Refs. 38 and 39 to describe various types of conventional linear time-dependent materials with the purpose that it can be easily incorporated into the algorithm of the so-called Z-transform-based TLM method, given in Ref. 38.

The TLM Z-transform model of the lossy MM-composite is based on the Drude dispersive model as it allows to characterize MM-composite response in a much wider frequency range than, e.g., the Lorentz dispersion model. However, it could be easily adapted to describe any higher-order material responses. In this paper, the Drude model describing the frequency dependence of electric and magnetic conductivities is used

Eq. (11)

σe(ω)=σe01+jωτe=ωpe2τeε01+jωτe,σm(ω)=σm01+jωτm=ωpm2τmμ01+jωτm,
where ωpe,m, τe,m, and σe,m0 are the electric and magnetic plasma frequencies, and the corresponding collision times and static conductivities, respectively. For NRM matched to free space, the static electric and magnetic conductivities are related to σm0=η02σe0, where η0 is the wave impedance of free space.

As an alternative, the Drude model describing the frequency dependence of permittivity and permeability (i.e., electric and magnetic susceptibilities) of MM composites can be used, but both models give identical results as shown in Ref. 37. In addition, using the relations

Eq. (12)

ε(ω)=ε0[1jσe(ω)ωε0],μ(ω)=μ0[1jσm(ω)ωμ0]
it is possible to determine permittivity and permeability from known electric and magnetic conductivities and vice versa.

In Ref. 38, the Drude model was also used but only to describe electric conductivity of a nonmagnetized plasma with collisions. Also, instead of the exponential Z-transform employed in Ref. 38 to transfer frequency dependence of the considered material properties to the Z-domain, the bilinear Z-transform is adopted in Ref. 37 to develop TLM Z-transform model of lossy MM since the bilinear discretization provides a much more accurate scheme.

In this section, the dispersive TLM Z-transform model of the lossy MM composite, based on the Drude model for electric and magnetic conductivities of MM composite, will be briefly described and also presented in a more general form through signal flow diagrams. For the sake of brevity and clarity of the formulation to follow, only diagrams corresponding to the y-component of the electric field and the z-component of the magnetic field are shown in this paper. The calculation for other field components within the proposed numerical model can also be illustrated using similar diagrams.

In general, the TLM Z-transform method algorithm can be described by the following three equations:38

Eq. (13)

F̲r=R̲̲1TV̲i0.5V̲f,F̲=t̲̲(z)F̲r,V̲r=R̲̲F̲P̲̲V̲i,
where F̲r is the vector of reflected fields, R̲̲1T is the TLM process input matrix, V̲i is the vector of incident voltages, V̲f is the vector of free-source voltages, F̲ is the vector of total normalized fields, t̲̲(z) is the matrix of transmission coefficients, V̲r is the vector of reflected voltages, R̲̲ is the TLM process reflection matrix, and P̲̲ is the TLM process input/output matrix.

For the modeling of general linear isotropic frequency-dependent materials, only the transmission block t̲̲(z) needs to be calculated. The second term of Eq. (13) can be then reduced, for the considered field components, to

Eq. (14)

4Vyr=4Vy+ge(z)Vy+4(1z11+z1)χe(z)Vy,

Eq. (15)

4izr=4iz+rm(z)iz+4(1z11+z1)χm(z)iz.

Expressing electric and magnetic conductivities, given by the Drude model in Eqs. (11) and (12), by using the normalized conductivity ge(ω)=σe(ω)η0Δl and normalized resistivity rm(ω)=σm(ω)Δl/η0, respectively, and transforming them to the Z-domain using the bilinear transformation jω2(1z1)/[Δt(1+z1)] (Δl and Δt are the space and time discretization steps in TLM, respectively), the following representations in the Z-domain are obtained

Eq. (16)

ge(z)=(1+z1)gecBce(1z1Ace/Bce),rm(z)=(1+z1)rmcBcm(1z1Acm/Bcm),
where gec=σe0η0Δl, rmc=σm0Δl/η0, Ace=2τe/Δt1, Acm=2τm/Δt1, Bce=2τe/Δt+1, and Bcm=2τm/Δt+1. Assuming frequency-independent susceptibility terms in Eqs. (14) and (15), χe(z)=χe and χm(z)=χm, TLM scattering process incorporating the dispersive model of the lossy MM composite can be represented for considered field components with the flow graph shown in Fig. 3. In addition, the calculation of the accumulators Sce and Scm by using the blocks ge¯(z) and gm¯(z) in Fig. 3, respectively, is shown in Fig. 4.

Fig. 3

Dispersive transmission line matrix (TLM) Z-transform model for lossy electromagnetic metamaterials (MM): (a) calculation of Ey in the time-domain and (b) calculation of Hz in the time domain.

JNP_7_1_073086_f003.png

Fig. 4

Calculation of (a) accumulator Sce and (b) accumulator Scm.

JNP_7_1_073086_f004.png

The coefficients shown in Figs. 3 and 4 can be found after arranging that the frequency dependence of electric and magnetic conductivity is represented as a function of the field value at the previous time step by taking partial fraction expansions forms shown below

Eq. (17)

(1+z1)ge(z)=ge0+z1ge1+ge(z)¯,(1+z1)rm(z)=rm0+z1rm1+rm(z)¯.

Inserting Eq. (16) into Eq. (17) gives

Eq. (18)

ge0=gecBce,ge1=gecBce(2+Ace/Bce)=ge0(2+ace),

Eq. (19)

ge(z)¯=z1gec(1+2Ace/Bce+Ace2/Bce2)Bce(1z1Ace/Bce)=z1bce1z1ace,

Eq. (20)

rm0=rmcBmc,rm1=rmcBcm(2+Acm/Bcm)=rm0(2+acm),

Eq. (21)

rm(z)¯=z1rmc(1+2Acm/Bcm+Acm2/Bcm2)Bcm(1z1Acm/Bcm)=z1bcm1z1acm,
where we introduce the parameters ace=Ace/Bce, acm=Acm/Bcm, bce=ge0(1+2ace+ace2), and bcm=rm0(1+2acm+acm2). The coefficients Tce, Tcm, kce, and kcm can be easily found after multiplying Eqs. (14) and (15) with (1+z1) and replacing the products (1+z1)ge(z) and (1+z1)rm(z) with the right-hand side of Eqs. (17) and (18), respectively

Eq. (22)

Tce=(4+ge0+4χe)1,Tcm=(4+rm0+4χm)1,

Eq. (23)

kce=(4+ge14χe),kcm=(4+rm14χm).

The results of the numerical calculations using TLM Z-transform for the present problem are given and compared with the exact analytical results in the next section.

5.

Results and Discussion

The exact analytical solutions for the real part of the electric field E(x), given by Eq. (6), for two different values of the numerical parameters (blue lines) are presented and compared to the corresponding numerical solutions (brown lines) in Figs. 5 and 6.

Fig. 5

Analytical results (solid blue lines) versus numerical (dashed brown) for the real part of electric field E(x) as a function of x, with E0=1, a=105m, k=10,000cm1, γ1=0.25×102k, and γ2=7 γ1=1.75×102k.

JNP_7_1_073086_f005.png

Fig. 6

Analytical results (solid blue lines) versus numerical (dashed brown) for the real part of electric field E(x) as a function of x, with E0=1, a=105m, k=10,000cm1, γ1=0.80×102k and γ2=1.5 γ1=1.20×102k.

JNP_7_1_073086_f006.png

Both figures show that there is no reflection at the interfaces between the right-handed and the left-handed material slabs in this particular case. This is expected, since in our case, the impedance is constant throughout the entire space. From the presented curves, we see the obvious change of the direction of the wave at the boundaries between the slabs (e.g., at x=a/2 and x=a/2). Furthermore, we see in Fig. 5, for γ2=7γ1 there is considerably stronger attenuation of the signal in NRM compared to that in PRM. On the other hand, in Fig. 6, for γ2=1.5γ1 the attenuation of the signal in NRM is only slightly stronger compared to that in PRM. Furthermore, both Figs. 5 and 6 show an excellent agreement between the exact analytic results and the corresponding numerical results obtained using TLM Z-transform. The obtained numerical and analytical curves are practically indistinguishable.

From the results (6), and utilizing the definition of the Poynting vector,40 we readily see that the time average of the Poynting vector through the structure is given by

Eq. (24)

Π(x)=12Re[E(x)×H*(x)]=12E0H0e2κβxex.

This is an exponentially decaying, strictly positive function, showing that the energy flows in the positive x-direction throughout the entire periodical PRM-NRM structure, as expected. In particular, for the case without losses the time average of the Poynting vector is constant throughout the structure and equal to

Eq. (25)

Π(x)=12E0H0ex.

6.

Conclusion

We have presented a simple exact analytical solution to Helmholtz equation for periodic structures with graded permittivity and permeability profile changing according to a cosine function along the direction of propagation, with arbitrary loss factors in PRM and NRM. We analyzed a special case of matched impedance throughout the structure where the real parts of the effective permittivities and permeabilities have opposite signs and equal absolute values, while the imaginary parts are fully arbitrary. We compared the exact analytical results with the corresponding numerical results obtained using TLM Z-transform and obtained an excellent agreement between the analytical and numerical results.

The model is valid for arbitrary temporal dispersion and arbitrary losses as long as the general mathematical and physical constraints are satisfied (e.g., Kramers-Kronig relations).

It should be noted that, throughout the present paper, when we state that the loss factors can be chosen arbitrarily, we tacitly assume that they can be chosen arbitrarily as long as the general mathematical and physical constraints, e.g., Kramers-Kronig relations, are satisfied. Such constraints do not, in general, impose serious restrictions to (theoretically) match the impedance between the two media. The challenge is rather the practical realization of media which satisfy the conditions posed in Eq. (4). The model allows a straightforward generalization to any periodic refractive index profile using suitable Fourier series, including abrupt transitions between two materials, which will be the subject of coming studies. Other future challenges include the study of the extension of the present model to various two-dimensional and three-dimensional structures and the case of arbitrary index profiles.

Acknowledgments

The work of T.A. and N.D. was funded by the Serbian Ministry of Education, Science and Technological Development within the project TR-32024. The work of Z.J. was funded by the Serbian Ministry of Education, Science and Technological Development within the project TR-32008. M.D. developed the analytical solution and wrote a part of the manuscript. M.N. took part in developing the analytical model. T.A. developed numerical model and performed numerical verification of the results. N.D. developed numerical model and wrote a part of the manuscript. Z.J. took part in developing the concept and wrote a part of the manuscript.

References

1. 

R. A. DepineA. Lakhtakia, “A new condition to identify isotropic dielectric-magnetic materials displaying negative phase velocity,” Microw. Opt. Technol. Lett., 41 (4), 315 –316 (2004). http://dx.doi.org/10.1002/(ISSN)1098-2760 MOTLEO 0895-2477 Google Scholar

2. 

V. M. Shalaevet al., “Negative index of refraction in optical metamaterials,” Opt. Lett., 30 (24), 3356 –3358 (2005). http://dx.doi.org/10.1364/OL.30.003356 OPLEDP 0146-9592 Google Scholar

3. 

J. Valentineet al., “Three-dimensional optical metamaterial with a negative refractive index,” Nature, 455 (7211), 376 –379 (2008). http://dx.doi.org/10.1038/nature07247 NATUAS 0028-0836 Google Scholar

4. 

F. Falconeet al., “Babinet principle applied to the design of metasurfaces and metamaterials,” Phys. Rev. Lett., 93 (19), 197401 (2004). http://dx.doi.org/10.1103/PhysRevLett.93.197401 PRLTAO 0031-9007 Google Scholar

5. 

G. Dollinget al., “Cut-wire pairs and plate pairs as magnetic atoms for optical metamaterials,” Opt. Lett., 30 (23), 3198 –3200 (2005). http://dx.doi.org/10.1364/OL.30.003198 OPLEDP 0146-9592 Google Scholar

6. 

S. Zhanget al., “Experimental demonstration of near-infrared negative-index metamaterials,” Phys. Rev. Lett., 95 (13), 137404 (2005). https://doi.org/10.1103/PhysRevLett.95.137404 PRLTAO 0031-9007 Google Scholar

7. 

M. Kafesakiet al., “Left-handed metamaterials: the fishnet structure and its variations,” Phys. Rev. B, 75 (23), 235114 (2007). http://dx.doi.org/10.1103/PhysRevB.75.235114 PLRBAQ 0556-2805 Google Scholar

8. 

J. Valentineet al., “Three-dimensional optical metamaterial with a negative refractive index,” Nature, 455 (7211), 376 –379 (2008). http://dx.doi.org/10.1038/nature07247 NATUAS 0028-0836 Google Scholar

9. 

R. A. ShelbyD. R. SmithS. Schultz, “Experimental verification of a negative index of refraction,” Science, 292 (5514), 77 –79 (2001). http://dx.doi.org/10.1126/science.1058847 SCIEAS 0036-8075 Google Scholar

10. 

S. Xiaoet al., “Yellow-light negative-index metamaterials,” Opt. Lett., 34 (22), 3478 –3480 (2009). http://dx.doi.org/10.1364/OL.34.003478 OPLEDP 0146-9592 Google Scholar

11. 

W. CaiV. Shalaev, Optical Metamaterials: Fundamentals and Applications, Springer, Dordrecht, Germany (2009). Google Scholar

12. 

S. A. RamakrishnaT. M. Grzegorczyk, Physics and Applications of Negative Refractive Index Materials, SPIE Press Bellingham, WA & CRC Press, Taylor & Francis Group, Boca Raton, FL (2009). Google Scholar

13. 

J. B. Pendry, “Negative refraction makes a perfect lens,” Phys. Rev. Lett., 85 (18), 3966 –3969 (2000). http://dx.doi.org/10.1103/PhysRevLett.85.3966 PRLTAO 0031-9007 Google Scholar

14. 

N. Fanget al., “Sub-diffraction-limited optical imaging with a silver superlens,” Science, 308 (5721), 534 –537 (2005). http://dx.doi.org/10.1126/science.1108759 SCIEAS 0036-8075 Google Scholar

15. 

N. Engheta, “An idea for thin subwavelength cavity resonators using metamaterials with negative permittivity and permeability,” IEEE Antennas Wirel. Propag. Lett., 1 (1), 10 –13 (2002). http://dx.doi.org/10.1109/LAWP.2002.802576 IAWPA7 1536-1225 Google Scholar

16. 

J. B. PendryD. SchurigD. R. Smith, “Controlling electromagnetic fields,” Science, 312 (5781), 1780 –1782 (2006). http://dx.doi.org/10.1126/science.1125907 SCIEAS 0036-8075 Google Scholar

17. 

D. LuZ. Liu, “Hyperlenses and metalenses for far-field super-resolution imaging,” Nat. Commun., 3 1205 (2012). https://doi.org/10.1038/ncomms2176 NCAOBW 2041-1723 Google Scholar

18. 

W. Caiet al., “Optical cloaking with metamaterials,” Nat. Photonics, 1 (4), 224 –227 (2007). http://dx.doi.org/10.1038/nphoton.2007.28 1749-4885 Google Scholar

19. 

T. H. Funget al., “Controlling electric fields spatially by graded metamaterials: implication on enhanced nonlinear optical responses,” Opt. Commun., 282 (5), 1028 –1031 (2009). http://dx.doi.org/10.1016/j.optcom.2008.11.028 OPCOB8 0030-4018 Google Scholar

20. 

S. A. RamakrishnaJ. B. Pendry, “Spherical perfect lens: solutions of Maxwell’s equations for spherical geometry,” Phys. Rev. B, 69 (11), 115115 (2004). http://dx.doi.org/10.1103/PhysRevB.69.115115 PLRBAQ 0556-2805 Google Scholar

21. 

D. R. Smithet al., “Gradient index metamaterials,” Phys. Rev. E, 71 (3), 036609 (2005). http://dx.doi.org/10.1103/PhysRevE.71.036609 PLEEE8 1063-651X Google Scholar

22. 

A. O. PinchukG. C. Schatz, “Metamaterials with gradient negative index of refraction,” J. Opt. Soc. Am. A, 24 (10), A39 –A44 (2007). http://dx.doi.org/10.1364/JOSAA.24.000A39 JOAOD6 0740-3232 Google Scholar

23. 

N. M. Litchinitseret al., “Metamaterials: electromagnetic enhancement at zero-index transition,” Opt. Lett., 33 (20), 2350 –2352 (2008). http://dx.doi.org/10.1364/OL.33.002350 OPLEDP 0146-9592 Google Scholar

24. 

M. DalarssonP. Tassin, “Analytical solution for wave propagation through a graded index interface between a right-handed and a left-handed material,” Opt. Express, 17 (8), 6747 –6752 (2009). http://dx.doi.org/10.1364/OE.17.006747 OPEXFF 1094-4087 Google Scholar

25. 

M. DalarssonZ. JakšićP. Tassin, “Exact analytical solution for oblique incidence on a graded index interface between a right-handed and a left-handed material,” J. Optoelectron. Biomed. Mat., 1 (4), 345 –352 (2009). Google Scholar

26. 

M. DalarssonZ. JakšićP. Tassin, “Structures containing left-handed metamaterials with refractive index gradient: exact analytical versus numerical treatment,” Microwave Rev., 15 (2), 2 –5 (2009). Google Scholar

27. 

M. DalarssonM. NorgrenZ. Jakšić, “Lossy wave propagation through a graded interface to a negative index material—case of constant impedance,” Microwave Rev., 17 (2), 2 –6 (2011). Google Scholar

28. 

M. DalarssonM. NorgrenZ. Jakšić, “Lossy gradient index metamaterial with sinusoidal periodicity of refractive index: case of constant impedance throughout the structure,” J. Nanophoton., 5 (1), 051804 (2011). http://dx.doi.org/10.1117/1.3590251 JNOACQ 1934-2608 Google Scholar

29. 

Z. L. MeiJ. BaiT. J. Cui, “Gradient index metamaterials realized by drilling hole arrays,” J. Phys. D, 43 (5), 055404 (2010). http://dx.doi.org/10.1088/0022-3727/43/5/055404 JPAPBE 0022-3727 Google Scholar

30. 

V. N. Smolyaninovaet al., “Experimental demonstration of birefrigent transformation optics devices,” Phys. Rev. B, 87 (7), 075406 (2013). http://dx.doi.org/10.1103/PhysRevB.87.075406 PLRBAQ 0556-2805 Google Scholar

31. 

T. Erginet al., “Three dimensional invisibility cloak at optical wavelengths,” Science, 328 (5976), 337 –339 (2010). http://dx.doi.org/10.1126/science.1186351 SCIEAS 0036-8075 Google Scholar

32. 

J. Valentineet al., “An optical cloak made of dielectrics,” Nat. Mater., 8 (7), 568 –571 (2009). http://dx.doi.org/10.1038/nmat2461 NMAACR 1476-1122 Google Scholar

33. 

R. Liuet al., “Gradient index circuit by waveguided metamaterials,” Appl. Phys. Lett., 94 (7), 073506 (2009). http://dx.doi.org/10.1063/1.3081399 APPLAB 0003-6951 Google Scholar

34. 

K.-E. Peiponenet al., “Kramers-Kronig relations and sum rules of negative refractive index media,” Eur. Phys. J. B, 41 (1), 61 –65 (2004). http://dx.doi.org/10.1140/epjb/e2004-00294-6 EPJBFY 1434-6028 Google Scholar

35. 

Z. Szaboet al., “A unique extraction of metamaterial parameters based on Kramers–Kronig relationship,” Microwave Theory Technol., 58 (10), 2646 –2653 (2010). https://doi.org/10.1109/TMTT.2010.2065310 Google Scholar

36. 

A. FangTh. KoschnyC. M. Soukoulis, “Self-consistent calculations of loss-compensated fishnet metamaterials,” Phys. Rev. B, 82 (12), 121102 (2010). https://doi.org/10.1103/PhysRevB.82.121102 PLRBAQ 0556-2805 Google Scholar

37. 

N. Dončovet al., “TLM modelling of left-handed metamaterials by using digital filtering techniques,” Microwave Rev., 16 (2), 2 –7 (2010). Google Scholar

38. 

J. PaulC. ChristopoulosD. W. P. Thomas, “Generalized material models in TLM—part I: materials with frequency-dependent properties,” IEEE Trans. Antennas Propag., 47 (10), 1528 –1534 (1999). http://dx.doi.org/10.1109/8.805895 IETPAK 0018-926X Google Scholar

39. 

J. PaulC. ChristopoulosD. W. P. Thomas, “Generalized material models in TLM—part II: materials with anisotropic properties,” IEEE Trans. Antennas Propag., 47 (10), 1535 –1542 (1999). http://dx.doi.org/10.1109/8.805896 IETPAK 0018-926X Google Scholar

40. 

L. SolymarE. Shamonina, Waves in Metamaterials, Oxford University Press, Oxford (2009). Google Scholar

Biography

JNP_7_1_073086_d001.png

Mariana Dalarsson is a teaching and research assistant at Royal Institute of Technology, School of Electromagnetic Engineering. She received her MSc in microelectronics and applied physics from the Royal Institute of Technology, Stockholm, Sweden. Her interests include inverse problems in electromagnetics, electromagnetics of stratified media, double-negative metamaterials, mathematical physics, etc. She is an author of 18 peer-reviewed publications, including 10 journal papers and a textbook in theoretical physics.

JNP_7_1_073086_d002.png

Martin Norgren received his PhD in electrical engineering-electromagnetic theory from the Royal Institute of Technology, Stockholm, Sweden. His interests include electromagnetic theory of stratified media, chiral materials, double-negative metamaterials, etc. He is an author of a number of peer-reviewed publications, including 35 journal papers. He is a professor at Royal Institute of Technology, School of Electromagnetic Engineering.

JNP_7_1_073086_d003.png

Tatjana Asenov received his Dipl.-Ing. degree from the Faculty of Electronic Engineering, University of Niš, Serbia, in 2008. Since 2009, she has been a research assistant and a PhD candidate at the Department of Telecommunications, Faculty of Electronic Engineering, Serbia. She conducts research, funded by the Serbian Ministry of Education and Science, on metamaterials and computational and applied electromagnetics with strong emphasis on microwave applications.

JNP_7_1_073086_d004.png

Nebojšа Dončov was born in Niš, Serbia, in 1970. He received his Dipl.-Ing., MSc, and PhD degrees from the Faculty of Electronic Engineering, University of Nis, Serbia, in 1995, 1999, and 2002, respectively. From 1995 to 2001, he was with the Department of Telecommunications, Faculty of Electronic Engineering, Serbia, as a research assistant. From 2001 to 2004, he was working with Flomerics Ltd., Electromagnetics Division, UK, as a research and development engineer. In 2004, he joined the Department of Telecommunications, Faculty of Electronic Engineering, Serbia, where he is now an associate professor. His current research interests are in computational and applied electromagnetics with a particular emphasis on TLM and network methods applications in microwaves and EMC. He was the recipient of the International Union of Radio Science (URSI) Young Scientist Award in 2002.

JNP_7_1_073086_d005.png

Zoran Jakšić received his Dipl.-Ing., Mag. Sci., and PhD degrees in electrical engineering-engineering physics from the School of Electrical Engineering, University of Belgrade. His interests are nanophotonics and nanoplasmonics, including optical metamaterials, NEMS and MEMS sensors and infrared detectors. He is an author of about 260 peer-reviewed publications including 60 journal papers and five book chapters. He is a full research professor and the science director with the Center of Microelectronic Technologies and Single Crystals, a department of the Institute of Chemistry, Technology and Metallurgy, University of Belgrade, Serbia.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Mariana Dalarsson, Martin Norgren, Tatjana Asenov, Nebojsa Doncov, and Zoran Jaksic "Exact analytical solution for fields in gradient index metamaterials with different loss factors in negative and positive refractive index segments," Journal of Nanophotonics 7(1), 073086 (19 August 2013). https://doi.org/10.1117/1.JNP.7.073086
Published: 19 August 2013
Lens.org Logo
CITATIONS
Cited by 10 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Metamaterials

Gradient-index optics

Refractive index

Magnetism

Composites

Wave propagation

Mathematical modeling

RELATED CONTENT


Back to Top