Understanding the dispersive properties of photonic crystals is a fundamental and well-studied problem. However, the introduction of singular permittivities and damping complicates the otherwise straightforward theory. In this paper, we study photonic crystals with a Drude–Lorentz model for the permittivity, motivated by halide perovskites. We demonstrate how the introduction of singularities and damping affects the spectral band structure and show how to interpret the notion of a “band gap” in this setting. We study a one-dimensional model for which we present explicit solutions.

Photonic crystals present a variety of interesting and useful wave transmission properties and have become ubiquitous in wave physics. Even very simple photonic crystals, such as those composed of periodically alternating layers of non-dispersive materials, can display exotic dispersive properties. As a result, they are able to support band gaps: ranges of frequencies that are unable to propagate through the material.25 These band gaps are the fundamental building blocks of the many different wave guides and wave control devices that have been conceived. Notable examples include flat lenses,20 invisibility cloaks,17 rainbow trapping filters27 and topological waveguides.11 

When working at certain electromagnetic frequencies (which often includes the visible spectrum), it is important to take into account the oscillatory behaviour of the free electrons in a metal. This behaviour leads to resonances at characteristic frequencies and gives metals a highly dispersive character (even before the introduction of macroscopic structure, as in a photonic crystal). Several different models exist to describe this behaviour. Most models are variants of the Lorentz oscillator model, whereby electrons are modelled as damped harmonic oscillators due to electrostatic attractions with nuclei.15 A popular special case of this is the Drude model, in which case the restoring force is neglected (to reflect the fact that most electrons in metals are not bound to any specific nucleus, so lack a natural frequency of oscillation). Many other variants of these models exist, for instance by adding or removing damping from the various models, cf. Ref. 21 or Ref. 14, and by taking linear combinations of the different models, as in Ref. 22.

A key feature that unites dispersive permittivity models is the existence of singularities in the permittivity. The position of these poles in the complex plane, which correspond to resonances, are one of the crucial properties that determines how a metal interacts with an electromagnetic wave. In conventional Lorentz models the poles appear in the lower complex plane.15 The imaginary part of the singular frequency is determined by the magnitude of the damping, and the singularities accordingly fall on the real line if the damping is set to zero. In the Drude model, the removal of the restorative force causes the singularities to fall at the origin and on the negative imaginary axis.

A particularly important example of dispersive materials, that are central to the motivation for this study, are halide perovskites. They have excellent optical and electronic properties and are cheap and easy to manufacture at scale.16 As a result, they are being used in many applications, including optical sensors,10 solar cells24 and light-emitting diodes.28 The dielectric permittivity of halide perovskites has been shown to depend heavily on excitonic transitions, leading to a permittivity that has symmetric poles in the lower complex plane.16 

There is a range of methods that can be used to capture the spectra of photonic crystals. For one-dimensional systems, explicit solutions typically exist and transfer matrices are particularly convenient. These were used for Drude materials in Ref. 23 and for undamped Lorentz materials in Ref. 14, for example. In multiple dimensions, studies often resort to numerical simulation (for instance with finite elements). A valuable approximation strategy is a multi-scale asymptotic method known as high-frequency homogenisation,9 which can be extended to approximate the dispersion curves in dispersive media.26 

In this work, we will study photonic crystals composed of metals with permittivity inspired by that of halide perovskites, in the sense that it has symmetric poles in the lower complex plane. After setting out the Floquet–Bloch formulation of the periodic problem in Sec. II, we will study the one-dimensional periodic Helmholtz problem in Subsection II D. In Sec. III, we retrieve the dispersion relation which characterizes the halide perovskite system and in Sec. IV, we show how its properties depend on the characteristics of the dispersive permittivity (namely, being real or being complex and having poles either on or below the real axis).

Let us consider NN particles D1, D2, …, DN which together occupy a bounded domain ΩRd, for d ∈ {1, 2, 3}. The collection of particles Ω will be the repeating unit of the periodic photonic crystal. We suppose that permittivity of the particles is given by a Drude–Lorentz-type model, given by
(2.1)
where ɛ0 denotes the background dielectric constant and α, β, γ are positive constants. α describes the strength of the interactions, β determines the natural resonant frequency and γ is the damping factor. This is motivated by the measured permittivity of halide perovskites, as reported in Ref. 16. We choose to use this expression as a canonical model for dispersive materials whose permittivities have singularities in the complex frequency space. Notice that (2.1) is singular at two complex values of ω. These are given by
(2.2)
By varying the parameters α, β and γ we can force these singularities to lie in the lower half of the complex plane (γ > 0), on the real line (γ = 0) or to vanish completely (β = γ = 0). We will make use of this property when trying to interpret the dispersion diagrams we obtain in the following analysis. We suppose that the particles are surrounded by a non-dispersive medium with permittivity ɛ0. We assume that the particles are non-magnetic, meaning the magnetic permeability μ0 is constant on all of Rd.
We consider the Helmholtz equation as a model for the propagation of time-harmonic waves with frequency ω. This is a reasonable model for the scattering of transverse magnetic polarised light (see e.g., Ref. 18, Remark 2.1 for a discussion). The wavenumber in the background Rd\Ω̄ is given by k0ωɛ0μ0 and we will use k to denote the wavenumber within Ω. Let us note here that, from now on, we will suppress the dependence of k0 and k on ω for brevity. We, then, consider the system of equations
(2.3)
where uin is the incident wave, assumed to satisfy (Δ+k02)uin=0, and the appropriate outgoing radiation condition depends on the dimension of the problem and of the periodic lattice.
We will assume that the collection of N particles is repeated in a periodic lattice Λ. We suppose that the lattice has dimension dl, in the sense that there are lattice vectors l1,,ldlRd which generate Λ according to
(2.4)
The fundamental domain of the lattice Λ is the set YRd given by
(2.5)
An example of a lattice of N=6 particles with lattice dimension dl=1 is shown in Fig. 1. The dual lattice of Λ, denoted by Λ*, is generated by the vectors α1,,αdl satisfying αi · lj = 2πδij for i, j = 1, …, dl. Finally, the Brillouin zone Y* is defined by
(2.6)
where 0 is the zero vector in Rddl. The Brillouin zone Y* is the space that the reduced unit cell of reciprocal space.
FIG. 1.

A periodic array of halide perovskite particles. Here, we have six particles D1, …, D6 repeated periodically in one dimension. Each of them has the halide perovskite permittivity ɛ(ω) defined by (2.1).

FIG. 1.

A periodic array of halide perovskite particles. Here, we have six particles D1, …, D6 repeated periodically in one dimension. Each of them has the halide perovskite permittivity ɛ(ω) defined by (2.1).

Close modal
The periodic structure, denoted by D, is given by
Hence, the problem we wish to study is the following:
(2.7)

In order to study the problem (2.7), we will make use of Floquet–Bloch theory.13 Let us first give certain definitions which will help with the analysis of the problem.

Definition 2.1.

A function f(x)L2(Rd) is said to be κ-quasiperiodic, with quasiperiodicity κY*, if e·xf(x) is Λ-periodic.

Definition 2.2
(Floquet transform). Let fL2(Rd). The Floquet transform of f is defined as

We have that F[f] is κ-quasiperiodic in x and periodic in κ. The Floquet transform is an invertible map F:L2(Rd)L2(Y×Y*), with inverse given by
where g(x, κ) is extended quasiperiodically for x outside of the unit cell Y.
Let us define uκ(x)F[u](x,κ). Then, applying the Floquet transform to (2.7), we obtain the following system:
(2.8)
The appropriate κ-quasiperiodic radiation condition is subtle to define in general, particularly since it depends on the dimensions of both the lattice and the physical space. The condition is required so that the solution (which, of course, is κ-quasiperiodic in the dl directions in which the lattice is periodic) contains only outgoing radiation in the ddl directions perpendicular to the axes of periodicity. See, for example Refs. 4–7, for more details. Clearly, as we will see in Sec. II D, this condition is not needed for the one-dimensional periodic transmission problems that will be the centrepiece of this work.

The solutions to (2.8) typically take the form of a countable collection of spectral bands, each of which depends continuously on the Bloch parameter κ. The goal of our analysis is identifying and explaining the gaps between the spectral band. At frequencies within these band gaps, waves do not propagate in the material and their amplitude decays exponentially. As a result, they are the starting point for building waveguides and other wave control devices.

For real-valued permittivities, it is straightforward to define band gaps as the intervals between the real-valued bands.

Definition 2.3

(Band gap for real permittivities). A frequency ωR is said to be in a band gap of the periodic structure D with real permittivities if it is such that (2.8) admits non-trivial solutions only for purely imaginary Bloch parameter κ.

We are interested in materials for which the permittivity takes complex values, corresponding to the introduction of damping to the model. We elect to keep the frequency ωR as a real number but allow the Bloch parameter κ to take complex values. In which case, the imaginary part of κ describes the rate at which the waves amplitude decays as it propagates in space. This choice is made for mathematical convenience (as will become clear below) and can be understood as encoding the attention of a wave propagating through the system as a function of space rather than time. It should be noted that it is also quite common to do the opposite and force κ to be real while allowing ω to be complex valued. This was done, for example in Refs. 3 and 26, and corresponds to viewing the attenuation as a function of time rather than space. Different conventions are typically adopted by different communities. For the one-dimensional problems considered in Sec. II D, it is clear that both approaches are mathematically equivalent.

In the real-valued case, it is clear that κ belongs to the Brillouin zone Y* (which has the topology of a torus, due to the periodicity in κ). When κ is complex valued, its real part still lives in Y* but its imaginary part can take arbitrary values. Thus, κ lives in a subset of the complex plane that is isomorphic to Y*×R. This can be thought of as a “generalised” Brillouin zone; this idea has been used to describe the spectral convergence of non-Hermitian systems in Ref. 2.

When there is damping in a system (characterized by a complex permittivity) and the frequency ω is chosen to be real, the Bloch parameter κ will typically have a non-zero imaginary part. As a result, it is less clear how to distinguish between spectral bands and band gaps. Nevertheless, we can understand a band gap as a region of frequency space in which the wave experiences attenuation beyond the material damping alone, due to the geometric structure. In which case, the imaginary part of κ will experience a local peak. With this in mind, we make the following definition of a band gap for a system with complex material parameters.

Definition 2.4

(Band gap for complex permittivities). We define a band gap for complex permittivities to be frequencies ωR for which (2.8) admits a non-trivial solution with quasiperiodicity κC that is such that |I(κ)| is in a neighbourhood of a local maximum.

We will study the problem in the one-dimensional setting. In one dimension, the problem is easier to manipulate and we are able to retrieve explicit expressions. Hence, we can get a variety of results concerning the characteristics of the quasiperiodic system. In particular, our main goal is to obtain the dispersion relation, an expression which relates the quasiperiodicities κC with the frequencies ωR, and study its properties.

Let us treat the Helmholtz problem (2.8) in the one-dimensional case. We will work on the interval [−1, 1], with [−1, 0) denoting the background and [0, 1) the particle. A schematic depiction of this is given in Fig. 2. Hence, the problem reads as follows:
(2.9)
on the domain [−1, 1], where
(2.10)
with the boundary conditions
(2.11)
FIG. 2.

The one-dimensional setting. The periodically repeated cell is of length 2. Here the interval [−1, 0) is the background and the interval [0, 1) is the particle.

FIG. 2.

The one-dimensional setting. The periodically repeated cell is of length 2. Here the interval [−1, 0) is the background and the interval [0, 1) is the particle.

Close modal
We will now retrieve an expression for the solution to (2.9). Let us define the quantities
(3.1)
For many of the results that follow, the crucial quantity will be the contrast between the material inside the particles and the background medium. With this in mind, we introduce the frequency-dependent contrast ρ as
(3.2)
Then, the following expression holds for the solution to (2.9).

Lemma 3.1.
Let u denote a solution to (2.9). Then, u is given by
(3.3)
where A,BC are two constants.

Proof.
We know that a solution to (2.9) must be given by
(3.4)
where A1,A2,B1,B2C are constants to be defined. This, also, gives
Now, from the boundary transmission conditions in (2.7), we require
These conditions mean we must have that B1=B2 and A1=ε(ω)/ε0A2=ρ(ω)A, which gives the desired result.□

Using the boundary conditions (2.11), we can obtain the dispersion relation for the one-dimensional problem. This is a well-known result, that first appeared in a quantum-mechanical setting12 and has since been shown to describe a range of periodic classical wave systems also Refs. 1 and 19. We include a brief proof, for completeness.

Theorem 3.2
(Dispersion relation). Let u denote the solution to (2.9) along with the boundary conditions (2.11). Then, for u to be non-trivial, the quasiperiodicities κC satisfies the dispersion relation
(3.5)

Proof.
From Lemma 3.1, we have that u is given by (3.3). Then, using (2.11), we have
(3.6)
We observe that for (2.9) to have a non-zero solution, it should hold
(3.7)
(3.8)
which gives
(3.9)
Making some algebraic rearrangements, we observe that
(3.10)
Finally, making the substitutions σc = ρσ0 and ε(ω)=ρε0, we obtain the desired result.□

The dispersion relation (3.5) can be used to plot the dispersion curves. For a given frequency ω, ρ(ω) can be calculated to yield the right hand side of (3.5), which can subsequently be solved to find κ. This is shown in Fig. 3. Since ɛ(ω) is complex valued, κ will generally take complex values. We plot only the absolute values of both the real and imaginary parts; as we will see below, this is sufficient to characterise the full dispersion relation. Notice also that R(κ)Y*=[π/2,π/2).

FIG. 3.

The dispersion relation of the halide perovskite photonic crystal. We model a material with permittivity given by (2.1) with α = 1, β = 1 and γ = 0.5. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is singular at two points, which are in the lower complex plane and are symmetric about the imaginary axis, as indicated in the sketch on the right and (the real parts) by the crosses on the frequency axes of the plots.

FIG. 3.

The dispersion relation of the halide perovskite photonic crystal. We model a material with permittivity given by (2.1) with α = 1, β = 1 and γ = 0.5. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is singular at two points, which are in the lower complex plane and are symmetric about the imaginary axis, as indicated in the sketch on the right and (the real parts) by the crosses on the frequency axes of the plots.

Close modal

The dispersion relation (3.5) describes the behaviour of the periodic system and reveals the relationship between the quasiperiodicities κC, the frequencies ωR and the permittivity ɛ(ω) of the material. We can use it to derive some simple results about the dispersion curves. The first thing to understand is the symmetries of the dispersion curves.

Lemma 3.3

(Opposite quasiperiodicities). Let κC be a complex quasiperiodicity satisfying the dispersion relation (3.5) for a given frequency ωR. Then, the opposite quasiperiodicity, i.e.,κ, satisfies the same dispersion relation.

Proof.

We just have to use that cos(·) is an even function. Then, if κC is such that (3.5) holds, from the fact that cos(−2κ) = cos(2κ), we get that κC also satisfies (3.5). This concludes the proof.□

It is with Lemma 3.3 in mind that we are able to plot only the absolute values of the imaginary parts in Fig. 3 and the subsequent figures.

1. Real and imaginary parts

In the analysis that will follow, it will be useful to be able to describe the behaviour of the real and imaginary part of the quasiaperiodicity with respect to the permittivity. In particular, we will decompose both the quasiperiodicity κ and ρ into real and imaginary parts, and we will derive this dependence from the dispersion relation. Since κC and ρC, let us define:
(3.11)
with κ1,κ2,ρ1,ρ2R. We will also define L1 and L2, which depend on ωR, as follows:
(3.12)
and
(3.13)
where we note that ρ1, ρ2 and σ0 all depend on the frequency ω, as specified in (3.1) and (3.2). Then, we have the following result.

Proposition 3.4.
Let κC, given by (3.11), satisfying the dispersion relation (3.5) for a given frequency ωR. Then, its real and imaginary parts are given by
(3.14)
and
(3.15)
where L1 and L2 are given by (3.12) and (3.13), respectively. We, also, note that the choice of + orshould be the same in (3.14) and (3.15).

Proof.
From (3.11), the dispersion relation (3.5) becomes
which is,
Taking real and imaginary parts, we obtain, for the real part,
(3.16)
and, for the imaginary part,
(3.17)
So, from (3.12) and (3.13), we obtain the system
(3.18)
From the first equation, we immediately see that
then, substituting into the second equation gives
We know that for x ∈ [−1, 1], we have the identity sin[arccos(x)]=1x2. Hence, from the above, we get
(3.19)
Similarly, using the fact that cosh2(x) − sinh2(x) = 1 for xR, we find that
(3.20)
Hence, we have
Using the quadratic formula, this gives
and so, we get
This gives the desired result.□

Remark.

Another way of viewing that the choice of + orin (3.14) is the same as the one in (3.15) is from the fact that we have shown that if κC satisfies (3.5), thenκ does as well, but κ̄ does not.

2. Imaginary part decay

From (3.15), we obtain a result on the decay of the imaginary part of the quasiperiodicity κ as ω. We will first state some preliminary results, before proving the main theorem.

Lemma 3.5.
Let the frequency-dependent contrast ρC be given by (3.2). Then, it holds
(3.21)
and
(3.22)

Proof.
From (3.1), we have
which gives directly limω|ρ|=1. This can be rewritten as
To ease the notation, let us write
(3.23)
Then, we have
(3.24)
We observe, from (3.23), that, as ω,
(3.25)
which gives
Also, since α, γ, ɛ0 > 0, it holds that
Hence, combining these results, we get
and
This concludes the proof.□

Lemma 3.6.
As ω, we have that
(3.26)
where L1=L1(ω) and L2=L2(ω) were defined in (3.12) and (3.13).

Proof.
From Lemma 3.5, we have that, as ω,
and from (3.1), we have that
So, it is essential to understand the behaviour of σ0ρ2 as ω. Using the same notations as in the Proof of Lemma 3.5, we have, without loss of generality on the ± of (3.24),
and hence, from (3.1), we have
From (3.25), we see that, as ω,
Hence, as ω,
which gives,
and so
(3.27)
Thus, (3.12) gives
which is the desired bound for L1. Similarly, from the triangle inequality applied on (3.13), we have
(3.28)
Using Lemma 3.5 and (3.27), we obtain
This concludes the proof.□

Using these results, we will describe the behaviour of the imaginary part κ2 of the quasiperiodicity κC as the frequency tends to infinity, i.e., ω.

Proposition 3.7.
Let us consider a complex quasiperiodicity κC satisfying the dispersion relation (3.5) with α,β,γR>0. Then, it holds that
(3.29)

Proof.
Indeed, since κC, let us define κ1R(κ) and κ2I(κ). Then, from (3.15), we have that κ2 is given by
From Lemma 3.6, we have that, as ω, L1 remains bounded, whereas L20. Thus, the following holds
since we have that |L1|1 also from Lemma 3.6. Then, from the continuity of the arcsinh(·) function, the desired result follows.□

The decay predicted by Proposition 3.7 is shown in Fig. 3. Due to the damping in the model, the imaginary part has discernible peaks at the first few gaps, but then decays steadily to zero at higher frequencies.

As mentioned before, the dispersion relation of the halide perovskite particles leads to dispersion curves which are not trivial to understand in terms of the traditional viewpoint of band gaps. In order to understand the behaviour, we will examine each distint feature of the halide perovskite permittivity, to understand the effect it has on the spectrum of the photonic crystal.

In particular, we will begin with the simplest case when the permittivity is real and constant with respect to the frequency ω. Then, we will introduce a dispersive behaviour to the permittivity by adding singularities at non-zero frequencies. We will initially suppose that these poles lie on the real axis and will study the behaviour close to these regions. Finally, we will study the effect of introducing a complex permittivity, corresponding to damping. Taken together, these results will allow us to explain the spectra the halide perovskite photonic crystal.

The first case we will consider is the one of a real-valued, non-dispersive permittivity, constant with respect to the frequency ω. In our setting this translates into having (2.1) with β = γ = 0 and α > 0, i.e.,
(4.1)
This setting has been studied quite extensively. We refer to Ref. 8, as a classical reference for studying the dispersive nature of waves in periodic systems. In Fig. 4, we provide an example for the dispersion curves when the permittivity is constant and real. It is worth noting the following result.
FIG. 4.

The dispersion relation of a photonic crystal with frequency-independent material parameters. We model a material with permittivity given by (2.1) with α = 1, β = 0 and γ = 0. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is never singular in this case.

FIG. 4.

The dispersion relation of a photonic crystal with frequency-independent material parameters. We model a material with permittivity given by (2.1) with α = 1, β = 0 and γ = 0. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is never singular in this case.

Close modal

Lemma 4.1.

Let ɛ(ω) be the real-valued, non-dispersive permittivity given by (4.1). Then, if κC is a quasiperiodicity satisfying (3.5) for a given frequency ωR, then so does κ̄C.

Proof.
Indeed, since ε(ω)R>0, then ρR>0. Let us take κC satisfying (3.5). Then, we can write κ = κ1 + 2, with κ1,κ2R. Now, since ρ > 0, we can write
Thus, (3.5) gives us
which becomes the following system
This implies that
If κ2 = 0, then κ=κ1R. Thus, κ=κ̄, which gives the desired result. If κ1=m2π, for mZ, then κ satisfies (3.5) if and only if
(4.2)
Since cosh(·) is an even function, we have that −κ2 satisfies (4.2) for the same frequency ω. Hence, κ̄=κ1iκ2 satisfies (3.5) and this concludes the proof.□

Crucially, the dispersion curves shown in Fig. 4 consist of a countable sequence of disjoint bands in which κ is real valued. Between each band there is a band gap, defined in the sense of Definition 2.3, in which κ is purely imaginary, corresponding to the decay of the wave. The occurence of κ being either purely real or purely imaginary is the mechanism behind Lemma 4.1. As we will see below, when we add singularities or damping to the model, the band gap structure is less straightforward to interpret.

Let us now study the case where the permittivity has a dispersive (and singular) character with respect to the frequency ω, but there is no damping, i.e., we consider α, β > 0 and γ = 0. This implies that
(4.3)
The interesting aspect in this setting is the existence of real poles for the permittivity. They are given by
In Fig. 5, we observe that near the pole of the permittivity there are infinitely many band-gaps. This was similarly observed recently by Ref. 26. Noting that a band gap occurs when the magnitude of the right-hand side of (3.5) is greater than one. We define the function
(4.4)
which is the right-hand side of (3.5). We will prove that this takes values greater than one on a countably infinite number of disjoint intervals within any neighbourhood of the singularity. To do so, we will introduce the following notation, which will be used in our analysis.
FIG. 5.

The dispersion relation of a photonic crystal with frequency-dependent material parameters that have a singularity. We model a material with permittivity given by (2.1) with α = 1, β = 1 and γ = 0. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is singular when ω = 1. The lower two plots are display the same dispersion curves, zoomed into the region around the singularity.

FIG. 5.

The dispersion relation of a photonic crystal with frequency-dependent material parameters that have a singularity. We model a material with permittivity given by (2.1) with α = 1, β = 1 and γ = 0. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is singular when ω = 1. The lower two plots are display the same dispersion curves, zoomed into the region around the singularity.

Close modal

Notation.

Let x,yR. Then, we use xy when xy and x > y. Similarly, we use xy when xy and x < y.

Then, close to a permittivity pole, the following holds.

Theorem 4.2.

Let ω* denote a pole of the permittivity ɛ(ω) given by (4.3), i.e., ω*±1β. Then, for δ > 0, the intervals [ω* − δ, ω*) and (ω*, ω* + δ] contain infinitely many disjoint sub-intervals, denoted by Ii and Ji, i = 1, 2, …, respectively, that are band gaps.

Proof.
We will first prove this result for the interval [ω* − δ, ω), for δ > 0. It suffices to show that there are infinitely many points ωi[ω*δ,ω*), i = 1, 2, …, for δ > 0 for which f(ωi)>1 or f(ωi)<1. Then, the continuity of f around these points gives us the existence of intervals of the form Ii[ωis,ωi+s], for i = 1, 2, …, for small s > 0, such that,
From (3.5) and (4.4), this gives
This is equivalent to the Ii’s, i = 1, 2, … being band gaps, since κ becomes complex in these intervals, i.e., |I(κ)|0. Hence, since |I(κ)| is continuous with respect to ω, we get that it has a local maximum in each of the Ii’s, for i = 1, 2, ….
We observe that limωω*ε(ω)=+. Then, this implies that limωω*ρ(ω)=+, and so, we get
Also, as ωω*, we have that σ0 is constant and so, without loss of generality, we can assume that sin(σ0) > 0 [the same argument holds for taking sin(σ0) < 0]. Hence, there exists δ1 > 0 such that for all ω ∈ [ω* − δ1, ω*), we have that
(4.5)
Now, since limωω*ρ(ω)=+, we have that
(4.6)
This implies that, for all K > 0, there exists δ2 > 0 such that for all ω ∈ [ω* − δ2, ω*) it holds that |ρ(ω)σ0(ω)| > K.
Now, let δ ≔ max{δ1, δ2} and let Iδ()[ω*δ,ω*). Then, (4.6) implies that the exist two families of infinitely many points in Iδ(), denoted by {ωi(+)}i=1,, and {ωi()}i=1,,, such that, for all i = 1, …, , we have
(4.7)
We also note that this implies, for all i = 1, …, , that
(4.8)
Thus, for all i = 1, …, , we have
(4.9)
and
(4.10)
In particular, without loss of generality, let us assume that ω0(+) is the smallest of the elements in both families {ωi(+)}i=1,, and {ωi()}i=1,,. Then, the periodicity of sin(·) shows that the elements of these families respect the following ordering:
(4.11)
Now, the continuity of f around these points allows us to take s > 0 such that
and
Finally, the infinity of elements in the families {ωi(+)}i=1,, and {ωi()}i=1,, gives us the desired result.

We note that for the neighborhood of the form ω ∈ (ω*, ω* + δ] for δ > 0, the proof remains the same with the slight change of taking the limits as ωω*.□

In addition to the occurrence of a countable number of band gaps close to the pole, in Fig. 5, we observe that there is an interesting behaviour of the imaginary part I(κ) of the quasiperiodicity κ as the frequency ωR approaches a permittivity pole. In fact, we see that close to a pole, |I(κ)| becomes arbitrarily big. This due to the resonance occurring here and is strongly related to the existence of infinitely many band gaps close to the pole. Actually, it is a corollary of Theorem 4.2.

Corollary 4.2.1.

Let ωR and κC be the associated quasiperiodicity satisfying the dispersion relation (3.5) and let f(ω) be the function defined in (4.4). Let ω*R denote a pole of the undamped permittivity ɛ(ω), given by (4.3). Then, for all K > 0, there exists δ > 0, such that for all p ∈ [−K, K], there exists ωp ∈ [ω* − δ, ω*) such that I(κ(ωp))=p. That is, |I(κ(ω))| takes arbitrarily large values as ωω*. The same result holds as ωω*.

Proof.
From Theorem 4.2, we have that for all K > 0, there exists δ > 0, such that, for all ω ∈ [ω* − δ, ω),
We have that cos(σ0(ω))cos(ρ(ω)σ0(ω)) remains bounded close to ω* and because of the continuity of sin(·), we can take sin(σ0(ω)) > 0 in [ω* − δ, ω). Then, it follows that, for all K > 0, in [ω* − δ, ω),
But, from Theorem (4.2), we have the existence of infinitely many points ω0(+)<ω0()<ω1(+)<ω1()< in [ω* − δ, ω), for which, sinρ(ω)σ0(ω) oscillates between 1 and −1 in each of the intervals of the form [ω0(+),ω0()], [ω0(),ω1(+)], …, denoted by Ii, i = 1, 2, …. This implies that, for all K > 0, for all p ∈ [−K, K], there exists ωpIi, for i = 1, 2, …, such that f(ωp) = p. Since this holds for all K > 0, it translates to f oscillating and taking all values between + and − in [ω* − δ, ω*) as we get closer to ω*.
Now, from (3.5) and (4.4), we get that
(4.12)
where ln(·) denotes the complex logarithm. We see that
Although, since we are using the complex logarithm, we have that
Hence, since we have shown that in [ω* − δ, ω*), the function f(ω) oscillates between + and −, we have that f(ω)±f(ω)21 has the same behaviour in [ω* − δ, ω*), but the oscillation takes place between 0 and +. Finally, since ln(·) is an increasing function, we obtain the desired result. Let us note that the proof is the same when we consider ωω*, with the slight change that we consider neighborhoods of the form (ω*, ω* + δ].□

We will now study the effect that introducing damping though allowing the permitivitty to be complex has on our one-dimensional system. Starting from the straightforward real-valued, non-dispersive model considered in Sec. IV A, we subsequently add damping. For this, we take αC and β = γ = 0, i.e.,
(4.13)
The dispersion curves for this setting are shown in Fig. 6. They are plotted for α with successively larger imaginary parts, to show the effect of gradually increasing the damping. We see that the clear structure of successive bands and gaps is gradually blurred out, eventually to the point that the spectrum bears no clear relation to the original undamped spectrum.
FIG. 6.

The dispersion relation of a photonic crystal with frequency-independent complex-valued material parameters. We model a material with permittivity given by (2.1) with αC, β = 0 and γ = 0. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is never singular in this case.

FIG. 6.

The dispersion relation of a photonic crystal with frequency-independent complex-valued material parameters. We model a material with permittivity given by (2.1) with αC, β = 0 and γ = 0. The frequency ω is chosen to be real and the Bloch parameter κ allowed to take complex values. The permittivity is never singular in this case.

Close modal

It is noticeable in Fig. 6 that, although the introduction of damping has distorted the curves (compared to those for the undamped system in Fig. 4), there are still intervals of frequency in which behaviour reminiscient of a band gap will be observed. That is, in Figs. 6(a)6(c) there are regions where the imaginary part of the Bloch parameter has clear peaks. As a result, a wave propagating through the system will experience stronger attenuation at these frequencies. This is due to effects from the periodic structure of the material, on top of the damping in the materials parameters, so it makes sense to refer to these regions as band gaps, as per the Definition 2.4.

In Fig. 6(d), the material damping is sufficiently large that its effect dwarfs any attenuation due to band gap effects. As a result, there are only minimal fluctuations of I(κ) in the corresponding regions and no local maxima. Hence, in the sense of Definition 2.4, there are no band gaps when the damping is too large. This aligns with what would be experienced in practice, in settings where the material damping is so large that any damping due to band-gap effects cannot be observed.

In many ways, the spectrum we obtain in this setting appears to be similar to the actual halide perovskite particles, as plotted in Fig. 3. Indeed, all of the results proved in Subsection III B hold, with the exception of the imaginary part decay. In fact, the converse is true, as made precise by the following result.

Lemma 4.3.
Let κC and ωR be a quasiperiodicity and a frequency, respectively, satisfying the dispersion relation (3.5) with complex-valued, non-dispersive permittivity given by (4.13). Then,
(4.14)

Proof.
Let us recall that κ2, denoting the imaginary part of κC, is given by (3.15), where L1 and L2 are given by (3.12) and (3.13), respectively. We have that
(4.15)
since σ0 is linear with respect to ω and ρ does not depend on ω in this setting. This implies
We note that it is enough to show that
since applying this on (3.15) gives that |κ2| → + as ω → +. Indeed, (3.12) gives that
where
and
Using the exponential formulation of the hyperbolic trigonometric functions, we get
Now, we observe that, as ω → +, C1 and C2 are both bounded and C1C2 ≠ 0. Then, from (4.15), we get that
Similarly, (3.13) gives that
where
and
Then, we can write
As before, we observe that, as ω → +, C̃1 and C̃2 are both bounded and C̃1C̃20. Then, from (4.15), we get that
This concludes the proof.□

The analysis in this section can be used to understand the dispersion diagram for the halide perovskite photonic crystal that was presented in Fig. 3. There are two crucial observations. First, we saw in Sec. IV B that the introduction of singularities in the permittivity led to the creation of countably infinitely many band gaps in a neigbourhood of the pole, when the pole falls on the real axis. However, this exotic behaviour is not seen in Fig. 3, due in part to the introduction of damping causing the poles to fall below the real axis. This effect can also be exaplined in terms of the results in Sec. IV C, where we saw that the introduction of damping to a simple non-dispersive model smoothed out the band gaps. The behaviour shown in Fig. 3 is a combination of these phenomena.

We have used analytic methods to understand the dispersive nature of photonic crystals fabricated from metals with singular permittivities. In particular, we considered a Drude–Lorentz model inspired by halide perovskites that has poles in the lower complex plane. For a one-dimensional system, we characterised the effect that each feature of this model has on the dispersion relation. In particular, we showed that the introduction of singularities leads to the creation of countably many band gaps near the poles, whereas the introduction of damping smooths out the band gap structure.

The work of K.A. was supported by ETH Zürich under the Project No. ETH-34 20-2. The work of B.D. was funded by the Engineering and Physical Sciences Research Council through a fellowship with Grant No. EP/X027422/1.

The authors have no conflicts to disclose.

Konstantinos Alexopoulos: Formal analysis (equal); Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal). Bryn Davies: Conceptualization (equal); Funding acquisition (equal); Methodology (equal); Resources (equal); Software (equal); Supervision (equal); Visualization (equal); Writing – review & editing (equal).

The data that support the findings of this study are openly available in Zenodo at http://doi.org/10.5281/zenodo.8055547. No other data were generated in this project.

1.
Adams
,
S. D. M.
,
Craster
,
R. V.
, and
Guenneau
,
S.
, “
Bloch waves in periodic multi-layered acoustic waveguides
,”
Proc. R. Soc. A
464
(
2098
),
2669
2692
(
2008
).
2.
Ammari
,
H.
,
Barandun
,
S.
,
Cao
,
J.
,
Davies
,
B.
, and
Hiltunen
,
E. O.
, “
Mathematical foundations of the non-Hermitian skin effect
,”
Arch. Ration. Mech. Anal.
248
(
3
),
33
(
2024
).
3.
Ammari
,
H.
,
Davies
,
B.
,
Hiltunen
,
E. O.
,
Lee
,
H.
, and
Yu
,
S.
, “
Exceptional points in parity–time-symmetric subwavelength metamaterials
,”
SIAM J. Math. Anal.
54
(
6
),
6223
6253
(
2022
).
4.
Ammari
,
H.
,
Fitzpatrick
,
B.
,
Kang
,
H.
,
Ruiz
,
M.
,
Yu
,
S.
, and
Zhang
,
H.
,
Mathematical and Computational Methods in Photonics and Phononics
,
Mathematical Surveys and Monographs Vol. 235
(
American Mathematical Society
,
Providence
,
2018
).
5.
Bao
,
G.
,
Dobson
,
D. C.
, and
Cox
,
J. A.
, “
Mathematical studies in rigorous grating theory
,”
J. Opt. Soc. Am. A
12
(
5
),
1029
1042
(
1995
).
6.
Bonnet-Bendhia
,
A.-S.
and
Starling
,
F.
, “
Guided waves by electromagnetic gratings and non-uniqueness examples for the diffraction problem
,”
Math. Methods Appl. Sci.
17
(
5
),
305
338
(
1994
).
7.
Botten
,
L. C.
,
Cadilhac
,
M.
,
Derrick
,
G.
,
Maystre
,
D.
,
McPhedran
,
R.
,
Nevière
,
M.
, and
Vincent
,
P.
,
Electromagnetic Theory of Gratings
(
Springer Science & Business Media
,
2013
), Vol.
22
.
8.
Brillouin
,
L.
,
Wave Propagation in Periodic Structures: Electric Filters and Crystal Lattices
(
Dover Publications
,
1953
), Vol.
2
.
9.
Craster
,
R. V.
,
Kaplunov
,
J.
, and
Pichugin
,
A. V.
, “
High-frequency homogenization for periodic media
,”
Proc. R. Soc. A
466
(
2120
),
2341
2362
(
2010
).
10.
Gu
,
L.
,
Poddar
,
S.
,
Lin
,
Y.
,
Long
,
Z.
,
Zhang
,
D.
,
Zhang
,
Q.
,
Shu
,
L.
,
Qiu
,
X.
,
Kam
,
M.
,
Javey
,
A.
, and
Fan
,
Z.
, “
A biomimetic eye with a hemispherical perovskite nanowire array retina
,”
Nature
581
(
7808
),
278
282
(
2020
).
11.
Khanikaev
,
A. B.
,
Hossein Mousavi
,
S.
,
Tse
,
W.-K.
,
Kargarian
,
M.
,
MacDonald
,
A. H.
, and
Shvets
,
G.
, “
Photonic topological insulators
,”
Nat. Mater.
12
(
3
),
233
239
(
2013
).
12.
Kronig
,
R. d. L.
and
Penney
,
W. G.
, “
Quantum mechanics of electrons in crystal lattices
,”
Proc. R. Soc. A
130
(
814
),
499
513
(
1931
).
13.
Kuchment
,
P.
, “
An overview of periodic elliptic operators
,”
Bull. Am. Math. Soc.
53
(
3
),
343
414
(
2016
).
14.
Li
,
J.
,
Zhou
,
L.
,
Chan
,
C. T.
, and
Sheng
,
P.
, “
Photonic band gap from a stack of positive and negative index materials
,”
Phys. Rev. Lett.
90
(
8
),
083901
(
2003
).
15.
Maier
,
S. A.
,
Plasmonics: Fundamentals and Applications
(
Springer
,
2007
), Vol.
1
.
16.
Makarov
,
S.
,
Furasova
,
A.
,
Tiguntseva
,
E.
,
Hemmetter
,
A.
,
Berestennikov
,
A.
,
Pushkarev
,
A.
,
Zakhidov
,
A.
, and
Kivshar
,
Y.
, “
Halide-perovskite resonant nanophotonics
,”
Adv. Opt. Mater.
7
(
1
),
1800784
(
2019
).
17.
Milton
,
G. W.
and
Nicorovici
,
N. A.
, “
On the cloaking effects associated with anomalous localized resonance
,”
Proc. R. Soc. A
462
(
2074
),
3027
3059
(
2006
).
18.
Moiola
,
A.
and
Spence
,
E. A.
, “
Acoustic transmission problems: Wavenumber-explicit bounds and resonance-free regions
,”
Math. Models Methods Appl. Sci.
29
(
02
),
317
354
(
2019
).
19.
Movchan
,
A. B.
,
Movchan
,
N. V.
, and
Poulton
,
C. G.
,
Asymptotic Models of Fields in Dilute and Densely Packed Composites
(
World Scientific
,
2002
).
20.
Pendry
,
J. B.
, “
Negative refraction makes a perfect lens
,”
Phys. Rev. Lett.
85
(
18
),
3966
(
2000
).
21.
Schulkin
,
B.
,
Sztancsik
,
L.
, and
Federici
,
J. F.
, “
Analytical solution for photonic band-gap crystals using Drude conductivity
,”
Am. J. Phys.
72
(
8
),
1051
1054
(
2004
).
22.
Sehmi
,
H. S.
,
Langbein
,
W.
, and
Muljarov
,
E. A.
, “
Applying the resonant-state expansion to realistic materials with frequency dispersion
,”
Phys. Rev. B
101
(
4
),
045304
(
2020
).
23.
Sigalas
,
M. M.
,
Chan
,
C. T.
,
Ho
,
K.
, and
Soukoulis
,
C. M.
, “
Metallic photonic band-gap materials
,”
Phys. Rev. B
52
(
16
),
11744
(
1995
).
24.
Snaith
,
H. J.
, “
Present status and future prospects of perovskite photovoltaics
,”
Nat. Mater.
17
(
5
),
372
376
(
2018
).
25.
Soukoulis
,
C. M.
,
Photonic Band Gap Materials
(
Springer Science & Business Media
,
2012
), Vol.
315
.
26.
Touboul
,
M.
,
Vial
,
B.
,
Assier
,
R.
,
Guenneau
,
S.
, and
Craster
,
R.
, “
High-frequency homogenization for periodic dispersive media
,”
SIAM J. Multisc. Model. Simul.
(to be published).
27.
Tsakmakidis
,
K. L.
,
Boardman
,
A. D.
, and
Hess
,
O.
, “‘
Trapped rainbow’ storage of light in metamaterials
,”
Nature
450
(
7168
),
397
401
(
2007
).
28.
Wang
,
H.
et al, “
Perovskite-molecule composite thin films for efficient and stable light-emitting diodes
,”
Nat. Commun.
11
(
1
),
891
(
2020
).
Published open access through an agreement with JISC Collections