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

Spherical symmetry and nonmagnetic dielectrics in analogue models of gravity

Eduardo Bittencourt bittencourt@unifei.edu.br    Renato Klippert Retired Federal University of Itajubá, Itajubá, Minas Gerais 37500-903, Brazil    Diego Renan da Silva São Paulo State University, Guaratinguetá, São Paulo 12516-410, Brazil    Érico Goulart Federal University of São João d’El Rei, C.A.P., Ouro Branco, Minas Gerais 36420-000, Brazil
Abstract

Nonmagnetic and spherically symmetric dielectric material media are investigated in their generality as analogue models, in the domains of geometrical optics, for a few relevant static and spherically symmetric solutions of gravitation, including black holes and wormholes. Typic and desirable gravitational features are systematically linked to the nonlinear properties of the phenomenological dielectric permittivity of the medium. From this, the limitations of such analogies are explored in terms of their corresponding geometric flexibility, dielectric configurations and causal structure.

analogue models of gravity; nonlinear electromagnetism; geometric optics.

I Introduction

Analogue models of gravity are the easiest way for testing experimentally the kinematical aspects of any geometric model of gravity in laboratory. The first example of these models was developed by Gordon [1] in the optical context, with further developments by De Felice [2] and Plebanski [3]. However, the turning point came with Unruh’s model on black hole evaporation in the context of fluid mechanics [4], where the quantum analysis became possible. Since then, several new models have been explored in the realm of condensed matter systems [5, 6]. More recently, it has been emphasized that the optical properties of material media can be fairly easily designed nearly at will by the present day technology on materials [7, 8, 9]. This suggests that the optical analogue models could be favoured as natural candidates for mimicking the behavior of gravitational systems from the experimental of view.

The present work aims to draw the basic lines to guide material designers, from the perspective that each kind of matter can be used to produce the desired effective gravitational potential. In order to obtain definite results of broad significance, some simplification is needed; the discussion applies to nonmagnetic isotropic dielectric media, which are used to model some properties of static and spherically symmetric solutions of the theory of gravitation.

For completeness, Sec. II reviews the theory of optical analogue models through the generalized Fresnel eigenvalue problem up to obtaining the ordinary and extraordinary modes of light propagation and the two corresponding optical analogue metrics. In Sec. III the general analogue models are obtained and the corresponding electromagnetic quantities are provided and compared with two paradigmatic gravitational models: the black hole and the wormhole models. Section IV discusses the causality issues related to Cauchy developments. The effective light cones were constructed in order to provide one possible physical interpretation for the optical analogue metrics.

Notation is such that Latin indices i,j,l,i,j,l,\ldots run from 11 to 33 and repeated indices are summed upon. Partial derivatives are denoted as i=/xi\partial_{i}=\partial/\partial x^{i}, t=/t\partial_{t}=\partial/\partial t, and ηijl\eta_{ijl} is the completely skew-symmetric Levi-Civita symbol with η123=1\eta_{123}=1.

II Geometric optics inside a nonlinear dielectric

To begin with, recall that Maxwell equations inside a material medium can be written, in a Cartesian coordinate system, as

iDi=ρ,ηijljEl=tBi,iBi=0,ηijljHl=Ji+tDi,\begin{array}[]{l}\partial_{i}D_{i}=\rho,\qquad\eta_{ijl}\partial_{j}E_{l}=-\partial_{t}B_{i},\\[4.30554pt] \partial_{i}B_{i}=0,\qquad\eta_{ijl}\partial_{j}H_{l}=J_{i}+\partial_{t}D_{i},\end{array} (1)

where DiD_{i} and HiH_{i} describe the field excitations and EiE_{i} and BiB_{i} stand for the corresponding electromagnetic field strengths. The nonlinear behavior of the medium is taken into account by replacing the usual linear constitutive relations by the following nonlinear ones:

Di=ϵ0Ei+Pi,\displaystyle D_{i}=\epsilon_{0}E_{i}+P_{i}, \displaystyle\quad\longrightarrow\quad Di=ϵijEj,\displaystyle D_{i}=\epsilon_{ij}E_{j}, (2)
Hi=1μ0BiMi,\displaystyle H_{i}=\frac{1}{\mu_{0}}B_{i}-M_{i}, \displaystyle\quad\longrightarrow\quad Hi=μij1Bj.\displaystyle H_{i}=\mu^{{}_{-1}}_{ij}B_{j}. (3)

Here, PiP_{i} is the polarization, MiM_{i} is the magnetization, and ϵ0\epsilon_{0} and μ0\mu_{0} are the vacuum dielectric constants. In general cases, the coefficients of permittivity ϵij\epsilon_{ij} and the inverse permeability μij1\mu^{{}_{-1}}_{ij} both can depend on the fields EiE_{i} and BiB_{i}. Note that extra terms accounting for the electric/magnetic excitation due to magnetic/electric field strength can both be included, but we shall not deal with this here (see Ref. [10] for details).

In the regime of geometric optics, the field strengths, the charge ρ\rho and the current JiJ_{i} are continuous, but their derivatives may have a finite step through the wavefront Σt(xi)=const.\Sigma_{t}(x_{i})=\mbox{const.}, for a given instant of time tt. According to Hadamard’s theorem [11, 12], [Ei]Σ=0=[Bi]Σ\left[E_{i}\right]_{\Sigma}=0=\left[B_{i}\right]_{\Sigma} implies

[tEi]Σt=ωei,[jEi]Σt=eikj,[tBi]Σt=ωbi,[jBi]Σt=bikj,\begin{array}[]{l}\left[\partial_{t}E_{i}\right]_{\Sigma_{t}}=-\omega\,e_{i},\quad\left[\partial_{j}E_{i}\right]_{\Sigma_{t}}=e_{i}\,k_{j},\\[4.30554pt] \left[\partial_{t}B_{i}\right]_{\Sigma_{t}}=-\omega\,b_{i},\quad\left[\partial_{j}B_{i}\right]_{\Sigma_{t}}=b_{i}\,k_{j},\\[4.30554pt] \end{array} (4)

where the symbol [f]Σt(p)limδ0+[f(p+)f(p)]\left[f\right]_{\Sigma_{t}}(p)\equiv\lim_{\delta\to 0^{+}}[f(p_{+})-f(p_{-})] indicates how the step of the electromagnetic quantities are evaluated on Σt\Sigma_{t}, with pΣtp\in\Sigma_{t} and p±Σt±δp_{\pm}\in\Sigma_{t\pm\delta} such that limδ0+p±=p\lim_{\delta\to 0^{+}}p_{\pm}=p. We also denote the wave frequency by ω\omega and the wave vector by kik_{i}. The polarization modes of the electric and magnetic components of the light rays are represented by eie_{i} and bib_{i}, respectively.

Once applied to the Maxwell equations (1), the procedure described above yields a linearly polarized wave with the magnetic polarization being given by ωbi=ηijlkjel\omega b_{i}=\eta_{ijl}k_{j}e_{l} , while the electric polarization eie_{i} satisfies the eigenvalue problem [13]

Zijej=0,Z_{ij}e_{j}=0, (5)

where ZijZ_{ij} represent the components of the Fresnel matrix [14]. The existence of nontrivial solutions for this eigenvalue problem is equivalent to the requirement

det[Zij]=0.\det[Z_{ij}]=0. (6)

From this equation, the dispersion relation associated with the light rays can then be obtained and the nonlinear behavior of the system is encoded in the dielectric parameters.

It is worth to mention that the vectors EiE_{i} and BiB_{i} represent the total field, that is, the composition of an external field plus a wave field. As long the limit of geometric optics is valid, the fields associated with the propagating waves are much weaker than the controllable fields—those produced by a given source distribution in the medium or by imposed external fields. Thus, wave fields were consistently neglected in these sums.

The majority of the optical media respond linearly to external magnetic fields. This means that the permeability can be considered as μij1=(1/μ0)δij\mu^{{}_{-1}}_{ij}=(1/\mu_{0})\delta_{ij}, where μ0\mu_{0} is the vacuum permeability and δij\delta_{ij} is the Kronecker delta, which is either 11 for i=ji=j or 0 otherwise. Besides, we shall consider the simplest case of an isotropic nonlinear electric medium such that the permittivity matrix is the same for all spatial directions and depends only on the norm E=EiEiE=\sqrt{E_{i}E_{i}} of the electric field, that is, ϵij=ϵ(E)δij\epsilon_{ij}=\epsilon(E)\delta_{ij}. In this case, the Fresnel matrix is given by [15]

Zij=ϵδij+ϵEEiEj1μ0v2(δijk^ik^j),Z_{ij}=\epsilon\delta_{ij}+\frac{\epsilon^{\prime}}{E}E_{i}E_{j}-\frac{1}{\mu_{0}v^{2}}(\delta_{ij}-\hat{k}_{i}\hat{k}_{j}), (7)

where ϵ=dϵ/dE\epsilon^{\prime}=d\epsilon/dE, v=ω/kv=\omega/k is the phase velocity and k^i=ki/k\hat{k}_{i}=k_{i}/k, with k=kikik=\sqrt{k_{i}k_{i}}.

Straightforward calculations show that Eq. (6) can be written as a polynomial equation for the phase velocity and a simple manipulation of such equation casts it as a pair of equations of the form gμνkμkν=0g^{\mu\nu}k_{\mu}k_{\nu}=0, from which we can read out the effective optical metrics [16, 17]. With Greek indices running from 0 to 33 and the Minkowski background metric in Cartesian coordinates taking the form [ημν]=diag(1,+1,+1,+1)[\eta_{\mu\nu}]={\rm diag}(-1,+1,+1,+1), these effective metric matrices can be written as

gαβ(+)=γαβ+(11μ0ϵ)vαvβ,\displaystyle g^{(+)}_{\alpha\beta}=\gamma_{\alpha\beta}+\left(1-\frac{\textstyle{1}}{\textstyle{\mu_{0}\epsilon}}\right)v_{\alpha}v_{\beta}, (8)
gαβ()=γαβ+[11μ0(ϵ+ϵE)]vαvβϵEϵ+ϵEl^l^α,β\displaystyle g^{(-)}_{\alpha\beta}=\gamma_{\alpha\beta}+\left[1-\frac{\textstyle{1}}{\textstyle{\mu_{0}(\epsilon+\epsilon^{\prime}E)}}\right]v_{\alpha}v_{\beta}-\frac{\textstyle{\epsilon^{\prime}E}}{\textstyle{\epsilon+\epsilon^{\prime}E}}\hat{l}{}_{\alpha}\hat{l}{}_{\beta}, (9)

where γαβ\gamma_{\alpha\beta} denotes the Minkowski metric in an arbitrary coordinate system, vμ=δ0μv^{\mu}=\delta^{\mu}_{0} represents the class of observers comoving with the laboratory, that is, it measures the electric and magnetic components of the electromagnetic fields, and l^α\hat{l}{}^{\alpha} is a unit vector pointing along the direction of the electric field. The effective metric gαβ(+)g^{(+)}_{\alpha\beta} is the well-known Gordon metric [1, 16] and it easily is recognizable by its isotropicity. This metric is responsible for describing the propagation of the ordinary modes. On the other hand, the optical metric gαβ()g^{(-)}_{\alpha\beta} does depend on the nonlinear behavior of the dielectric and it shall be responsible for the propagation of the extra-ordinary modes. This is the metric we are interested in.

It is easy to see that, in the particular case of vacuum, where the dielectric coefficients are ϵ0\epsilon_{0} and μ0\mu_{0}, these two metrics degenerate to a unique optical metric gμνg_{\mu\nu} given by the Minkowski one. The relevance of these optical metrics lies on the fact that the integral curves of the wave vector kμk_{\mu} correspond to null geodesic curves in the artificial spacetime endowed with a geometry given by the effective optical metric [18, 19]. As far as solely kinematic aspects of the gravitation are considered, the effective geometry could be compared with the metric of the curved spacetime. Thus, the sort of phenomena predicted in gravitational systems could be scrutinized in the realm of optics inside material media. There are many applications of this analogy including tests of bending of light, cosmological models in terrestrial laboratories, or the probe of predictions in quantum gravity phenomenology [5].

III Static and spherically symmetric optical metrics

Let us consider a dielectric material with the features described above, subjected to an external radial electric field and without external magnetic fields. For convenience, we choose coordinates adapted to the observers such that the four-velocity is vα=δ0αv^{\alpha}=\delta^{\alpha}_{0}. Then, the static and spherically symmetric situation we are dealing with admits solely a non-vanishing charge density ρ\rho. Therefore, Maxwell equations (1) essentially reduce to

r(r2ϵE)r2=ρ.\frac{\partial_{r}(r^{2}\epsilon E)}{r^{2}}=\rho. (10)

Thus, the effective geometry Eq. (9) expressed in spherical-like coordinates (t,r,θ,ϕ)(t,r,\theta,\phi) reads

gαβ=diag(1μ0(ϵ+ϵE),ϵϵ+ϵE,r2,r2sin2θ).g_{\alpha\beta}={\rm diag}\left(-\frac{\textstyle{1}}{\textstyle{\mu_{0}(\epsilon+\epsilon^{\prime}E)}},\,\frac{\epsilon}{\epsilon+\epsilon^{\prime}E},\,r^{2},\,r^{2}\sin^{2}\theta\right). (11)

This expression allows one to seek for analogue static and spherically symmetric metrics of the following form

gαβ=diag(e2Φ(r),11b(r)r,r2,r2sin2θ).g_{\alpha\beta}={\rm diag}\left(-e^{2\Phi(r)},\,\frac{\textstyle{1}}{\textstyle{1-\frac{b(r)}{r}}},\,r^{2},\,r^{2}\sin^{2}\theta\right). (12)

Inspired by the Morris-Thorne Ansatz concerning wormholes [20], the redshift function Φ(r)\Phi(r) and the shape function b(r)b(r) are enough to describe any static and spherically symmetric spacetime in a unified form. In particular, the well-known solutions of Einstein equations such as Schwarzschild, Reissner-Nordström, de-Sitter as well as the class of Morris-Thorne wormholes. The matching of Eqs. (11) and (12) gives the two independent equations

μ0(ϵ+ϵE)=e2Φ,\displaystyle\mu_{0}(\epsilon+\epsilon^{\prime}E)=e^{-2\Phi}, (13)
ϵ+ϵEϵ=1br.\displaystyle\frac{\epsilon+\epsilon^{\prime}E}{\epsilon}=1-\frac{b}{r}. (14)

Once combined, these conditions provide the radial dependence of the dielectric permittivity, that is

ϵ(r)=ϵ0e2Φ(1br)1,\epsilon(r)=\epsilon_{0}\,e^{-2\Phi}\left(1-\frac{b}{r}\right)^{-1}, (15)

where we used the identification μ01=ϵ0\mu_{0}^{-1}=\epsilon_{0}, since the speed of light in vacuum is set to unity. Substitution of Eq. (15) back in Eq. (13) yields the electric field in terms of the radial coordinate, as follows

E(r)=E0|rb1|exp(2rbdΦdr𝑑r),E(r)=E_{0}\left|\frac{r}{b}-1\right|\exp\left(2\int\frac{r}{b}\,\frac{d\Phi}{dr}dr\right), (16)

where E0E_{0} is an integration constant. This equation is valid for any signs of rbr-b and bb. However, we are mainly interested in the case rb0r-b\geq 0, in order to preserve the Lorentzian signature of metric in Eq. (12).

Finally, the expressions of the electric displacement D=ϵED=\epsilon E and the charge density ρ\rho in terms of rr read as

D\displaystyle D =\displaystyle= ϵ0E0rbexp[2(rb1)dΦdr𝑑r],\displaystyle\epsilon_{0}E_{0}\,\frac{\textstyle{r}}{\textstyle{b}}\exp{\left[2\int\left(\frac{\textstyle{r}}{\textstyle{b}}-1\right)\,\frac{\textstyle{d\Phi}}{\textstyle{dr}}dr\right]}, (17)
ρ\displaystyle\rho =\displaystyle= ϵ0E0[3brb2dbdr+2dΦdr(rb1)]×\displaystyle\epsilon_{0}E_{0}\,\left[\frac{\textstyle{3}}{\textstyle{b}}-\frac{\textstyle{r}}{\textstyle{b^{2}}}\frac{\textstyle{db}}{\textstyle{dr}}+2\,\frac{\textstyle{d\Phi}}{\textstyle{dr}}\left(\frac{\textstyle{r}}{\textstyle{b}}-1\right)\right]\times (18)
×exp[2(rb1)dΦdr𝑑r].\displaystyle\times\exp{\left[2\int\left(\frac{\textstyle{r}}{\textstyle{b}}-1\right)\,\frac{\textstyle{d\Phi}}{\textstyle{dr}}dr\right]}.

Therefore, for any choice of the functions Φ(r)\Phi(r) and b(r)b(r) describing the gravitational metric to be mimicked, the corresponding analogue model can be constructed, at least theoretically, via equations (15)-(18). In what follows, two important cases of static and spherically symmetric geometries in the context of general relativity are considered. Some possible experimental limitations for reproducing such metrics in laboratory are also discussed latter, by assuming a nonmagnetic and nonlinear isotropic material.

III.1 Black hole analogue models

First, consider the case in which the functions Φ(r)\Phi(r) and b(r)b(r) correlate as

Φ(r)=12ln(1b(r)r).\Phi(r)=\frac{1}{2}\ln\left(1-\frac{b(r)}{r}\right). (19)

This choice leads to a gravitational metric of the form given in Eq. (12) where the condition g00=1/g11g_{00}=-1/g_{11} holds. This condition includes all black hole metrics like Schwarzschild, Kottler and Reissner-Nordström (De Sitter and anti-De Sitter spacetimes are also solutions of this, but without event horizons).

Accordingly, when Eqs. (15)-(18) are restricted to the case of Eq. (19), the corresponding analogue model presents the following electromagnetic configuration

ϵ(r)=ϵ0(1br)2,E(r)=E0(rb1)2,D(r)=ϵ0E0r2b2,ρ(r)=2ϵ0E0r2b2(2r1bdbdr).\begin{array}[]{lcl}\epsilon(r)&=&\epsilon_{0}\left(1-\frac{\textstyle{b}}{\textstyle{r}}\right)^{-2},\\[4.30554pt] E(r)&=&E_{0}\left(\frac{\textstyle{r}}{\textstyle{b}}-1\right)^{2},\\[4.30554pt] D(r)&=&\frac{\textstyle{\epsilon_{0}E_{0}r^{2}}}{\textstyle{b^{2}}},\\[4.30554pt] \rho(r)&=&\frac{\textstyle{2\epsilon_{0}E_{0}r^{2}}}{\textstyle{b^{2}}}\left(\frac{\textstyle{2}}{\textstyle{r}}-\frac{\textstyle{1}}{\textstyle{b}}\frac{\textstyle{db}}{\textstyle{dr}}\right).\end{array} (20)

In particular, for this case one has

ϵ(E)=ϵ0(E0E±1)2.\epsilon(E)=\epsilon_{0}\left(\sqrt{\frac{E_{0}}{E}}\pm 1\right)^{2}. (21)

Thus, the permittivity diverges as EE approaches zero. Therefore, we have a large permittivity for small values of EE, maintaining always the electric displacement finite (see Eqs. 20). This cannot be achieved by using materials one can find in nature, but possibly with short-band human made media like metamaterials. Besides, the physically reasonable expression for ϵ(E)\epsilon(E) is the one with the plus sign, avoiding the situation ϵ|E0=0\epsilon|_{E_{0}}=0.

III.2 Wormhole analogue models

The (intuitive) conditions for the existence of a wormhole without an event horizon are that Φ(r)\Phi(r) is finite everywhere and b(r0)=r0b(r_{0})=r_{0} for some r=r0>0r=r_{0}>0 [20]. This implies that the permittivity from Eq. (15) diverges at the throat, while the electric field (16) goes to zero, at least as far as the argument of the exponential remains finite. For the sake of illustration, take the special class of Morris-Thorne wormholes with zero redshift function. Thus,

ϵ(r)=ϵ0(1br)1,E(r)=E0(rb1),D(r)=ϵ0E0rb,ρ(r)=ϵ0E0(3brb2dbdr).\begin{array}[]{lcl}\epsilon(r)&=&\epsilon_{0}\left(1-\frac{\textstyle{b}}{\textstyle{r}}\right)^{-1},\\[4.30554pt] E(r)&=&E_{0}\left(\frac{\textstyle{r}}{\textstyle{b}}-1\right),\\[4.30554pt] D(r)&=&\frac{\textstyle{\epsilon_{0}E_{0}r}}{\textstyle{b}},\\[4.30554pt] \rho(r)&=&\epsilon_{0}E_{0}\,\left(\frac{\textstyle{3}}{\textstyle{b}}-\frac{\textstyle{r}}{\textstyle{b^{2}}}\frac{\textstyle{db}}{\textstyle{dr}}\right).\end{array} (22)

The elimination of the variable rr from the first two relations above yields

ϵ(E)=ϵ0(E0E+1).\epsilon(E)=\epsilon_{0}\left(\frac{E_{0}}{E}+1\right). (23)

Although the equations are slightly different from the previous case, the conclusion is quite similar: a divergent permittivity and a finite electric displacement as the electric field goes to zero, which happens at the throat at least. Again, metamaterials would be required in order to reproduce this sort of geometries.

III.3 Power-law media

The previous subsections have shown that any gravitational black-hole or wormhole (with zero redshift function) cannot be reproduced in laboratory by conventional materials - those media whose electric permittivity is proportional to a positive power of the magnitude of the electric field, i.e. ϵEα\epsilon\propto E^{\alpha} with α>0\alpha>0. Then, the natural question raises: what kind of metrics can be produced with conventional media? Thus, consider a power law relation between the electric permittivity and the electric field. With this assumption we cover the well-known Pockels and Kerr effects, first and second order corrections in EE, respectively, to the electric susceptibility of a given medium.

Thus, from the substitution of

ϵ(E)=ϵ0(EE0)α,\epsilon(E)=\epsilon_{0}\left(\frac{\textstyle{E}}{\textstyle{E_{0}}}\right)^{\alpha}, (24)

with α>0\alpha>0 in Eqs. (15) and (16), it follows that

Φ=12ln(1br),\Phi=-\frac{1}{2}\ln\left(1-\frac{b}{r}\right), (25)

for any α>0\alpha>0. Please, note a sign on comparing this with Eq. (19). Returning back to Eqs. (15) and (16), one finds out that ϵ\epsilon and EE are actually constant and that the effective metric reduces to the flat Minkowski spacetime. Once again, nontrivial effective metrics that are static and spherically symmetric demand a more ellaborated relationship between ϵ\epsilon and EE.

III.4 Neutral dielectric media

Finally, assume that the charge density vanishes identically, such that the profile of the electric displacement is directly obtained from Gauss law, that is

D=kr2,D=\frac{k}{r^{2}}, (26)

where kk is an integration constant. The comparison of this pattern with the product ϵE\epsilon E given by Eqs. (15) and (16) yields a first order differential equation involving Φ(r)\Phi(r) and b(r)b(r), which can be solved, leading to

Φ(r)=12ln(1br)+lnrr0dr(rb).\Phi(r)=-\frac{1}{2}\ln\left(1-\frac{b}{r}\right)+\ln\frac{r}{r_{0}}-\int\frac{dr}{(r-b)}. (27)

The last two terms on the right-hand side of this relation distort the solution in Eq. (25), yielding a nontrivial effective metric. The dielectric parameter and the electric field strength are, respectively, given by

ϵ(r)=ϵ0(r0r)2exp(2dr(rb)),E(r)=E0exp(2dr(rb)).\begin{array}[]{lcl}\epsilon(r)&=&\epsilon_{0}\left(\frac{\textstyle{r_{0}}}{\textstyle{r}}\right)^{2}\exp\left(2\int\frac{dr}{(r-b)}\right),\\[8.61108pt] E(r)&=&E_{0}\exp\left(-2\int\frac{dr}{(r-b)}\right).\end{array} (28)

From this, it is clear that k=ϵ0E0r02k=\epsilon_{0}E_{0}r_{0}^{2}. Moreover, the permittivity is ϵ=ϵ0E0r02r2E1\epsilon=\epsilon_{0}E_{0}r_{0}^{2}\,r^{-2}E^{-1}. The actual dependence of ϵ\epsilon on EE requires the explicit behavior of E(r)E(r) to be algebraically inverted, which in turns demands the specification of the shape function b(r)b(r).

IV Hyperbolicity and causal structure

In this Section, the interconnected issues of hyperbolicity and causal structure for the models described so far are briefly discussed. Hyperbolicity lies at the roots of physics, since it amounts to the predictable power of the theory by asserting that solutions exist, are unique and depend continuously on the initial data. For physical models described by quasi-linear partial differential equations with constraints, as is the case under investigation, additional care is needed since the principal part of the equations will depend explicitly on the electromagnetic fields. Roughly speaking, this means that the Cauchy problem may be well-posed for some field intensities while ill-posed for others. Needless to say, if hyperbolicity was to be violated, one could have dispersion relations endowed with pathologies, which would compromise the propagation of waves assumed at the very beginning of our analysis. These pathologies are often associated with non-real eigenvalues of the characteristic polynomial (evanescent modes) or cones of influence with non-convex topologies.

The hyperbolicity of Maxwell equations with local, but otherwise arbitrary, constitutive laws has been analyzed in [21, 10]. In particular, one inescapable condition for well-posedness to hold for Eqs. (1)-(3) is that, at each spacetime point, the homogeneous multivariate polynomial of fourth order

P(ξ)(gαβ(+)ξαξβ)(gαβ()ξαξβ)P(\xi)\equiv(g_{\alpha\beta}^{(+)}\xi^{\alpha}\xi^{\beta})(g_{\alpha\beta}^{(-)}\xi^{\alpha}\xi^{\beta}) (29)

is hyperbolic with respect to a direction in the tangent space [22, 23]. In other words, wμTpM\exists\ w^{\mu}\in T_{p}M such that P(w)>0P(w)>0 and the map λP(u+λw)\lambda\mapsto P(u+\lambda w), itself a univariate polynomial of fourth order, has only real roots λi\lambda_{i}, for all uμTpMu^{\mu}\in T_{p}M. Geometrically, hyperbolicity in the direction of wμw^{\mu} is the requirement that every line parallel to wμw^{\mu} in TpMT_{p}M intersects the null cones of gαβ(+)g_{\alpha\beta}^{(+)} and gαβ()g_{\alpha\beta}^{(-)} at exactly four points (counting multiplicities). Therefore, one obtains the conditions

  1. 1.

    the effective metrics gαβ(+)g_{\alpha\beta}^{(+)} and gαβ()g_{\alpha\beta}^{(-)} both must have a Lorentzian signature equal to (,+,+,+)(-,+,+,+);

  2. 2.

    the set of future-directed time-like vectors with respect to both metrics must form a unique connected, open, convex set.

Needless to say, if the above requirements are satisfied, the causal structure of the theory will be well-behaved: wave excitations or discontinuities over any background configuration will propagate in a predictable way with a finite velocity of propagation.

Now, a direct calculation using Eqs. (9) shows that the effective metrics will be Lorentzian, whenever the following inequalities are satisfied

g(+)=γμ0ϵ<0,g()=γμ0ϵ(1+ϵEϵ)2<0.g^{(+)}=\frac{\gamma}{\mu_{0}\epsilon}<0,\qquad g^{(-)}=\frac{\gamma}{\mu_{0}\epsilon(1+\frac{\epsilon^{\prime}E}{\epsilon})^{2}}<0. (30)

where g(±)=det(gαβ(±))g^{(\pm)}=\det\left(g_{\alpha\beta}^{(\pm)}\right) and γdet(γαβ)\gamma\equiv\mbox{det}(\gamma_{\alpha\beta}). The latter imply that, in order to preserve the correct signature, the following two conditions must be fulfilled:

ϵ>0,ϵ+ϵE>0.\epsilon>0,\quad\quad\quad\epsilon+\epsilon^{\prime}E>0. (31)

These two conditions are automatically satisfied by all spherically symmetric models described so far. That condition 2 above is also fulfilled is immediate. Since the class of observers comoving with the laboratory, vμv^{\mu}, is time-like with respect to both effective metrics then, by continuity, there will be an open set of time-like vectors containing vμv^{\mu} which is necessarily connected, open and convex, as required. Hence, these models are not plagued by mathematical inconsistencies.

Refer to caption
Figure 1: Null cones of gμν(+)g^{(+)}_{\mu\nu} are inside the Minkowski light cones, shown as a function of rr. The vertical plane on left represents the worldsheet of the horizon, i.e. r=rHr=r_{H}, for which the effective light cones degenerate. As the rr-coordinate grows, the vertices of the other light cones are placed at 2rH,3rH2r_{H},3r_{H} and 4rH4r_{H}, respectively. Also, μ01=ϵ0=1=E0\mu_{0}^{-1}=\epsilon_{0}=1=E_{0} were adopted for simplicity. Notice the isotropic pattern of the propagation.
Refer to caption
Figure 2: Null cones of the effective metric gμν()g^{(-)}_{\mu\nu} also are inside the Minkowski light cones, here shown as a function of rr. All numerical choices are the same as in Fig. 1. The anisotropic pattern is due to the term ϵE\epsilon^{\prime}E in the effective metric. Since this term vanishes for large values of rr, the effective cones gradually approach Minkowski light cones from inside.

Another interesting feature of the effective metrics, Eqs. (8)-(9), is that either their null cones coincide or they share two (and only two) common directions. Indeed, in Cartesian coordinates adapted to the laboratory frame i.e., vμ=δ0μv^{\mu}=\delta^{\mu}_{0} and γμν=ημν\gamma_{\mu\nu}=\eta_{\mu\nu}, we may write ξμ=(ξ0,ξ)\xi^{\mu}=(\xi^{0},\vec{\xi}) and the corresponding null cone conditions read as

(ξ0)2μ0ϵ(ξξ)=0,(ξ0)2μ0(ϵ+ϵEsin2ψ)(ξξ)=0(\xi^{0})^{2}-\mu_{0}\epsilon(\vec{\xi}\cdot\vec{\xi})=0,\qquad(\xi^{0})^{2}-\mu_{0}(\epsilon+\epsilon^{\prime}E\mbox{sin}^{2}\psi)(\vec{\xi}\cdot\vec{\xi})=0 (32)

with ψ\psi denoting the angle between ξ\vec{\xi} and the electric field E\vec{E}. Therefore, if ϵE=0\epsilon^{\prime}E=0, the cones are precisely the same, whereas if ϵE0\epsilon^{\prime}E\neq 0, these cones agree along the directions of ±E\pm\vec{E}. This means that, in a static and spherically symmetric situation, it would be impossible to distinguish between the two effective geometries, and consequently between the ordinary and the extra-ordinary light rays, by solely studying the behavior of radial light propagation.

For a concrete example, consider the causal structure in the simple case of a Schwarzschild geometry, for which b(r)=rHb(r)=r_{H}. A direct computation using Eqs. (20) gives the following electromagnetic quantities

ϵ(r)=ϵ0(1rH/r)2,E(r)=E0(r/rH1)2,D(r)=ϵ0E0(r/rH)2,ρ(r)=4ϵ0E0(r/rH2),\begin{array}[]{lcl}\epsilon(r)&=&\epsilon_{0}\left(1-r_{H}/r\right)^{-2},\\[4.30554pt] E(r)&=&E_{0}\left(r/r_{H}-1\right)^{2},\\[4.30554pt] D(r)&=&\epsilon_{0}E_{0}\left(r/r_{H}\right)^{2},\\[4.30554pt] \rho(r)&=&4\epsilon_{0}E_{0}(r/r^{2}_{H}),\end{array} (33)

which are finite in the regime rH<r<r_{H}<r<\infty. Thus, one concludes that the electric permittivity is a positive definite, monotonically decreasing function, diverging as rrHr\rightarrow r_{H} and approaching ϵ0\epsilon_{0} as rr\rightarrow\infty. Conversely, the electric field intensity vanishes as rrHr\rightarrow r_{H} and diverges as rr\rightarrow\infty. Also, by noticing that

ϵE=(dϵdr)(dEdr)1E=ϵ0rHr(1rHr)2\epsilon^{\prime}E=\left(\frac{d\epsilon}{dr}\right)\left(\frac{dE}{dr}\right)^{-1}E=\epsilon_{0}\frac{r_{H}}{r}\left(1-\frac{\textstyle{r_{H}}}{\textstyle{r}}\right)^{-2} (34)

one easily shows that the anisotropic term, given by ϵE\epsilon^{\prime}E, vanishes as rr\rightarrow\infty, as expected. Therefore, the effective null cones, Eqs. (32), coincide at spatial infinity and increasingly disagree as the radial coordinate is gradually decreased. Nevertheless, the two cones always intersect for rays propagating along the radial direction, as discussed before. Clearly, both cones coincide with the Minkowski light cone for rrHr\gg r_{H}, and they degenerate at r=rHr=r_{H}, which is avoided here by considering only the exterior region.

The qualitative behavior of the light cones for both effective metrics are depicted in Figs. (1)-(2) for different values of the radii, that is, r/rH=1,2,3,4r/r_{H}=1,2,3,4. Figure (1) displays the effectively causal cones defined by gμ(+)g_{\mu}^{(+)}. The effective light cone lies inside the Minkowski background causal cone, and approaches it as rr goes to infinity. The isotropic feature of the effective cones is evident. In the case of gμ()g_{\mu}^{(-)} in Fig. (2), the same qualitative behavior for the effective light cones of gμν()g_{\mu\nu}^{(-)} for the same values of the parameter, despite their anisotropy.

V Concluding Remarks

Most of the literature on analogue models present some interesting properties of a certain kind of material medium, without actually providing what is the specific feature of the material which allows the obtained behavior. The present work is a first step to construct a global picture, by discussing what it can be obtained, and what cannot, from these properties. In particular, it was shown that most of the non standard optical properties of non magnetic isotropic dielectrics demand a medium whose electric permittivity diverges in the limit of E0E\rightarrow 0. The discussion of the causal structure provides the basic requirements on the effective metrics in order to mimic some reliable spacetime geometry.

Acknowledgements.
DRS would like to thank Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) for the financial support.

References

  • Gordon [1923] W. Gordon, Zur lichtfortpflanzung nach der relativitätstheorie, Annalen der Physik 72, 421 (1923).
  • de Felice [1971] F. de Felice, On the gravitational field acting as an optical medium, Gen. Rel. and Grav. 2, 1396 (1971).
  • Plebanski [1960] J. Plebanski, Electromagnetic waves in gravitational fields, Phys. Rev. 118, 1396 (1960).
  • Unruh [1981] W. G. Unruh, Experimental black-hole evaporation?, Phys. Rev. Lett. 46, 1351 (1981).
  • Barceló et al. [2005] C. Barceló, S. Liberati, and M. Visser, Analogue gravity, Living Reviews in Relativity 8, 12 (2005).
  • Novello et al. [2002] M. Novello, M. Visser, and G. Volovik, Artificial Black Holes (World Scientific, 2002).
  • Pendry et al. [2006] J. B. Pendry, D. Schurig, and D. R. Smith, Controlling electromagnetic fields, Science 312, 1780 (2006).
  • Smith et al. [2000] D. R. Smith, W. J. Padilla, D. C. Vier, S. C. Nemat-Nasser, and S. Schultz, Composite medium with simultaneously negative permeability and permittivity, Phys. Rev. Lett. 84, 4184 (2000).
  • Fleischhauer et al. [2005] M. Fleischhauer, A. Imamoglu, and J. P. Marangos, Electromagnetically induced transparency: Optics in coherent media, Rev. Mod. Phys. 77, 633 (2005).
  • Perlick [2011] V. Perlick, On the hyperbolicity of maxwell’s equations with a local constitutive law, Journal of Mathematical Physics 52, 042903 (2011).
  • Hadamard [1903] J. Hadamard, Leçons sur la propagation des ondes et les équations de hydrodynamique (Hermann, Paris, France, 1903).
  • Zakharov [1973] V. D. Zakharov, Gravitational waves in Einstein’s theory (John Wiley & Sons, New York, USA, 1973).
  • De Lorenci and Klippert [2006] V. A. De Lorenci and R. Klippert, Electromagnetic light rays in local dielectrics, Physics Letters A 357, 61 (2006).
  • De Lorenci and Goulart [2008] V. A. De Lorenci and G. P. Goulart, Magnetoelectric birefringence revisited, Physics Review D 78, 045015 (2008).
  • Bittencourt et al. [2017] E. Bittencourt, G. H. S. Camargo, V. A. De Lorenci, and R. Klippert, Controlled opacity in a class of nonlinear dielectric media, Phys. Rev. A 95, 033826 (2017).
  • Novello and Bittencourt [2012] M. Novello and E. Bittencourt, Gordon metric revisited, Physical Review D 86, 124024 (2012).
  • Bittencourt et al. [2016] E. Bittencourt, J. P. Pereira, I. I. Smolyaninov, and V. N. Smolyaninova, The flexibility of optical metrics, Classical and Quantum Gravity 33, 165008 (2016).
  • De Lorenci and Klippert [2002] V. A. De Lorenci and R. Klippert, Analogue gravity from electrodynamics in nonlinear media, Physical Review D 65, 064027 (2002).
  • De Lorenci [2002] V. A. De Lorenci, Effective geometry for light traveling in material media, Physical Review E 65, 026612 (2002).
  • Morris and Thorne [1988] M. S. Morris and K. S. Thorne, Wormholes in spacetime and their use for interstellar travel: a tool for teaching general relativity, American Journal of Physics 56, 395 (1988).
  • Rätzel et al. [2011] D. Rätzel, S. Rivera, and F. P. Schuller, Geometry of physical dispersion relations, Phys. Rev. D 83, 044047 (2011).
  • GÅRDING [1959] L. GÅRDING, An inequality for hyperbolic polynomials, Journal of Mathematics and Mechanics 8, 957 (1959).
  • Beig [2006] R. Beig, Concepts of hyperbolicity and relativistic continuum mechanics, in Analytical and Numerical Approaches to Mathematical Relativity, edited by J. Frauendiener, D. J. Giulini, and V. Perlick (Springer Berlin Heidelberg, Berlin, Heidelberg, 2006) pp. 101–116.