Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Nonlocal optical properties in periodic lattice of graphene layers

Open Access Open Access

Abstract

Based on the effective medium model, nonlocal optical properties in periodic lattice of graphene layers with the period much less than the wavelength are investigated. Strong nonlocal effects are found in a broad frequency range for TM polarization, where the effective permittivity tensor exhibits the Lorentzian resonance. The resonance frequency varies with the wave vector and coincides well with the polaritonic mode. Nonlocal features are manifest on the emergence of additional wave and the occurrence of negative refraction. By examining the characters of the eigenmode, the nonlocal optical properties are attributed to the excitation of plasmons on the graphene surfaces.

© 2014 Optical Society of America

1. Introduction

Graphene is a single layer of carbon atoms packed into a honeycomb lattice, which exhibits extraordinary electronic transport properties as strong ambipolar electric field effect [1], massless Dirac Fermions [2], and half-integer quantum Hall effect [3]. It is considered the thinnest material that combines aspects of semiconductors and metals with ultrahigh electron mobility [4] and superior thermal conductivity [5]. These properties are exploited in a variety of applications as transparent electrodes [6], ultrafast photodetectors and lasers [7, 8], broadband optical modulators [9], and even metamaterials [10, 11].

The optical properties of a single layer graphene are mainly determined by its sheet conductivity. A universal optical conductance is found for undoped graphene and, interestingly, is determined by the fine structure constant [1215]. By applying a gate voltage or doping in graphene, the optical properties can be modified due to a shift in the onset of interband transition [16, 17]. In particular, plasmons relations are identified in graphene layers and graphene plasmons provide a suitable alternative to metal plasmons [1822]. Compared to the latter, graphene plasmons exhibit much tighter field confinement and relatively longer propagation distances, with the advantage of being highly tunable through, for instance, the electrostatic gating [23].

If a number of decoupled graphene layers are arranged as a periodic lattice [2426], with the period much less than the wavelength, the collection of layers can be regarded as a medium with the optical properties characterized by its effective parameters. Considering the effective medium to be nonmagnetic (the effective permeability being unity) as a viewpoint, the effective permittivity is likely to be spatially dispersive or nonlocal, as in the crystals with excitations [27, 28]. This is true even when the wavelength is much greater than the characteristic length (e.g. lattice period) of the medium [29, 30], although the nonlocal effects are usually not important at long wavelengths. In a periodic structure, the nonlocal effects arise from the coupling of fields between neighboring unit cells and tend to be significant when the resonance (e.g. surface plasmon) occurs. These features are present in metal layers [31, 32] and expected to appear as well in graphene layers.

In the present study, we investigate the nonlocal optical properties in the periodic lattice of graphene layers, with emphasis on the concept of effective medium. Based on a nonlocal effective medium model, the effective permittivity tensor for the graphene layers is derived and expressed in approximate formulas that show explicit dependence on the wave vector. Strong nonlocal effects are indicated by the Lorentzian character of the effective permittivity and are manifest on the emergence of additional wave and the occurrence of negative refraction in the graphene layers. These nonlocal properties are attributed to the excitation of plasmon polaritons, in the form of nonlocal resonance, on the graphene surface that strongly modulate the dispersion of the lattice.

2. Effective medium model

The basic idea of the effective medium model in this study is to extract the effective parameters of a structure from its dispersion relations. This is achieved when analytical relations are available for the structure. A simple way to extract the effective parameters is by expanding the dispersion relations with respect to the wave number [32, 33]. Nonlocal effective parameters, in particular, can be approximately obtained when the expansion order is larger than two, the order for a local medium. For periodic structures, the effective medium model is valid provided that the period is much less than the wavelength.

2.1. Dispersion relations

Consider a periodic lattice of graphene layers with surface conductivity σ and period a embedded in a background with dielectric constant ε, as schematically shown in Fig. 1. Let the wave vector lie on the xz plane, that is, k = (kx, 0, kz), without loss of generality. With the time-harmonic dependence eiωt, the dispersion relations of the graphene layers are given by (see Appendix for details)

cos(kxa)=cos(qa)iσq2ωε0εsin(qa),
cos(kxa)=cos(qa)iσωμ02qsin(qa),
for TM and TE polarizations, respectively, where q=εk02kz2 with k0 = ω/c. Here, TM refers to transverse magnetic, where E = (Ex, 0, Ez) and H = (0, Hy, 0), and TE to transverse electric, where E = (0, Ey, 0) and H = (Hx, 0, Hz). The relations (1)(2) are equivalent to those obtained by the transfer matrix method [24, 26].

 figure: Fig. 1

Fig. 1 Schematic diagram of the periodic lattice of graphene layers with period a embedded in a background with dielectric constant ε. In this study, a = 0.1 h̄c/μ (≈ 98.9 nm) with μ = 0.2 eV and ε = 1.5 will be used as the parameters. A small area of the graphene feature is shown on the top layer for illustration.

Download Full Size | PDF

At sufficiently low temperature, kTμ, where k is the Boltzmann constant, T is the temperature, and μ is the chemical potential of graphene, the surface conductivity of graphene is given, within the random phase approximation, by the following relation [3438]:

σε0c=4αiΩ+πα[θ(Ω2)+iπln|Ω2Ω+2|],
where Ω = h̄ω/μ is the dimensionless frequency, is the reduced Planck constant, α ≈ 1/137 is the fine structure constant, and θ(x) is the Heaviside step function. The first and second terms on the right side of Eq. (3) stem from the intraband and interband contributions, respectively. The intraband conductivity is of the Drude-like form as in the metal. The imaginary part of σ, however, become negative for Ω > Ω* ≈ 1.667 due to the interband contribution [38]. For Ω > 2, σ is no longer purely imaginary, leading to the absorption of light in graphene.

2.2. Effective permittivity tensor

Assume that the lattice period a is much less than the wavelength λ, so that the graphene layers can be regarded as an effective medium, characterized by the dispersion relations of a homogeneous yet anisotropic medium as

kx2εzeff+kz2εxeff=k02,
kx2+kz2=εyeffk02
for TM and TE polarizations, respectively, where εxeff, εyeff, and εzeff are components of the effective permittivity tensor. By expanding the dispersion relations of the graphene layers [Eqs. (1) and (2)] with respect to k0, kx, and kz up to fourth order, and rearranging the expanding terms according to the forms in Eqs. (4) and (5), we arrive at the approximate formulas for the effective permittivity components as
εzeff=εz0γ12k02a21112kx2a2,εxeff=ε(1γ12εz0k02a2)1γ6εz0(k02a212εkz2a2),
εyeff=εy0(1+16kz2a2)δ6ε2k02a2+a212k02(kx4kz4),
where εz0=εy0=ε(1+δ), γ = ε2 (1 + 2δ), δ = iσ̃/(εk0a), and σ̃ = σ/(ε0c).

In this study, we choose a = 0.1 h̄c/μ (≈ 98.9 nm) with μ = 0.2 eV and ε = 1.5 to be the geometric and material parameters. The wavelength, λ = 2πh̄c/(μΩ), corresponds to the interband transition (Ω = 2) is around 3.1 μm. The working wavelength is therefore much greater than the lattice period up to the interband transition, which fulfills the necessary condition for the effective medium to be valid. Theoretically, the expansion is valid when the wave vector component is small. In practice, the effective permittivity tensor based on this expansion successfully recovers the dispersion relations even when the wave vector component is no longer small if the expansion order is suitably increased.

The in-plane effective permittivity component εzeff [cf. Eq. (6)], which is located in the plane where the wave vector lies, depends on the wave vector component kx, indicating the nonlocal nature of the graphene layers. The nonlocal effect, however, is weak, as |kx| is bound by π/a in a periodic lattice. It is shown in Fig. 2(a) that εzeff (blue solid line) differs not much from εz0 (blue dashed line). Here, εz0 (or εy0) is considered the quasistatic effective parameters along the parallel (to graphene surface) direction. This parameter exhibits the Drude-like behavior as in the metal, with an effective plasma frequency Ω0 determined by the zero of εz0, that is, εz0(Ω0)=0. Using Eq. (3) in εz0, we have

Ω02(1+εa˜α)1/2,
where ã = μa/(h̄c) is the dimensionless lattice period. Here, Ω0 is very close to the zero of εzeff and Eq. (8) also serves as a good approximation to εzeff(Ω0)=0, as shown in Fig. 2(b).

 figure: Fig. 2

Fig. 2 (a) In-plane effective permittivity components εzeff and εxeff as the functions of Ω for the same periodic lattice of graphene layers as in Fig. 1, where Ω = h̄ω/μ, Kx = kxa, Kz = kza, and ã = μa/(h̄c). (b) Effective plasma frequency Ω0 and its approximate formula [Eq. (8)] as the functions of ã.

Download Full Size | PDF

Another in-plane component εxeff (green solid line), on the other hand, distinctly deviates from its quasistatic parameter, εx0=ε (green dashed line). In particular, there exists a pole frequency Ωp, where εxeff±. Around Ωp, εxeff exhibits a Lorentzian resonance feature [cf. Fig. 2(a)]. Using Eq. (3) in εxeff, we have at small Kz

Ωp2[1+εa˜(12+Kz2)2α(6+Kz2)]1/2,
where Kz = kza is the dimensionless wave number. The pole frequency Ωp is very close to Ω0 at Kz = 0 and gradually moves toward higher frequencies as Kz increases (also shown in Fig. 6). As there is no constraint for Kz, the nonlocal effects due to εxeff are expected to be strong in the graphene layers, as in the lattice of metal layers [32].

The out-of-plane effective permittivity component εyeff, which is perpendicular to the plane where the wave vector lies, depends on both kx and kz, as shown in Fig. 3. Note that εyeff and εzeff are no longer equal as in the quasistatic case, although the layered structure makes no difference between the y and z directions in the geometry. The medium properties, therefore, depend on the polarization. Near Kx = Kz = 0 (where Kx = kxa), εyeff is characterized by the effective plasma frequency Ω0 and the critical frequency Ω* ≈ 1.667 [38]. Below Ω0, εyeff is negative and Ω0 also serves as the cutoff frequency of the graphene layers for TE polarization [cf. Eq. (5)]. Above Ω*, where the imaginary part of σ is negative [38], εyeff is larger than the background dielectric constant. At Kx = 0, εyeff basically decreases with Kz [Fig. 3(a)], while at Kz = 0, εyeff increases with Kx [Fig. 3(b)]. Compared to εxeff, the nonlocal effect due to εyeff is weaker.

 figure: Fig. 3

Fig. 3 Out-of-plane effective permittivity component εyeff for the same periodic lattice of graphene layers as in Fig. 1, where (a) Kx = 0 and (b) Kz = 0. Red and green lines correspond to εyeff=0 and εyeff=ε, respectively.

Download Full Size | PDF

Note also that there are no first-order kx and kz terms in all the effective permittivity components. This is a consequence of inversion symmetry of the underlying structure, an invariance property of a system when the coordinates are inverted [39, 40].

3. Nonlocal optical properties

The optical properties of the periodic lattice of graphene layers are closely related to its dispersion characteristics. In the present problem, the nonlocal effects, in particular, are manifest on the emergence of additional wave, the occurrence of negative refraction, and the excitation of graphene plasmons. These features are well characterized by the nonlocal effective parameters of the graphene layers derived in the preceding section.

3.1. Dispersion characteristics

Figure 4 shows the equifrequency surfaces of the dispersion relations for the same graphene layers as in Fig. 1. For TM polarization [Fig. 4(a)], the dispersion surface consists of two parts: the upper surface is the photonic mode, largely conformed to the light cone: Ωa˜=(Kz2+Kx2)/ε; the lower surface is the polaritonic mode, arising from the mixing of light wave with the excitations (i.e. plasmons) in the graphene layers. The two modes intersect at a single point on either side of the half space (Kz > 0 or Kz < 0). Along the plane of constant Kx ≠ 0, the polaritonic band forms an anticrossing (avoided crossing) scheme with the photonic band, as shown in Fig. 5(a), indicating the existence of couplings between the two modes.

 figure: Fig. 4

Fig. 4 Equifrequency surfaces of (a) TM and (b) TE dispersion relations for the same periodic lattice of graphene layers as in Fig. 1. Black circle in (b) is the section of light cone at Ω = 2.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 (a) TM frequency bands at Kx = 0, 0.03 and (b) TE frequency band at Kx = 0 for the same periodic lattice of graphene layers as in Fig. 1. Insets in (a) are typical magnetic field patterns of photonic mode (P mode) and polaritonic mode (PL mode). Red dot in (b) corresponds to the critical frequency Ω* ≈ 1.667 that separates P mode and N mode.

Download Full Size | PDF

The photonic and polaritonic bands, however, cross on the plane of Kx = 0 [dashed lines in Fig. 5(a)], which is considered a degeneracy due to symmetry. In this situation, the polaritonic mode possesses a completely different symmetry than the photonic mode. The former is a surface-like mode with odd symmetry (along the x direction) in the magnetic field, whereas the latter is a bulk mode with even symmetry [see the insets in Fig. 5(a)]. Here, the surface-like mode is not a surface mode in the usual sense. It is the surface part of the mode that exists in the bulk of the medium [41]. Notice that the crossing point for the photonic and polaritonic bands corresponds to the zero of εxeff, across which the Lorentzian resonance reverses the order of maximum and minimum. Therefore, the band crossing occurs when the numerator and denominator of εxeff become zero simultaneously. When Kx ≠ 0, the symmetries in the two modes are no longer completely different. A certain similar symmetry allows for the mode interaction and gives rise to the anticrossing scheme [42].

For TE polarization, the dispersion surface is approximately conformed to the light cone (similar to the photonic mode for TM polarization), but with a cutoff at the bottom, as in the waveguide or cavity mode. The cutoff frequency is very close to Ω0 [cf. Eq. (8)], which does not scale like 1/(εa) as in the case of periodic lattice of thin metal films [43], although the graphene conductivity exhibits the Drude-like character below the interband transition [cf. Eq. (3)]. A distinct feature is observed in the frequency range Ω* ≈ 1.667 < Ω < 2, where the dispersion surface is located outside the light cone [cf. Fig. 4(b)], indicating that light wave in the graphene layers becomes a slow wave. In this range, the imaginary part of graphene conductivity becomes negative [cf. Eq. (3)], which is considered a very different property than the metal with a positive imaginary part of the conductivity. This wave has been identified as a new electromagnetic mode [denoted by N mode in Fig. 5(b)] in the two-dimensional electron gas for TE wave [38] that does not appear in metal. At the critical frequency Ω*, the TE frequency band coincides with the light line [indicated by red dot in Fig. 5(b)].

For Ω > 2, the surface conductivity of graphene σ is no longer purely imaginary [cf. Eq. (3)], the dispersion relations of the graphene layers [Eqs. (1)(2)] do not allow for real frequency with real wave vectors. The corresponding eigenmode is thus a virtual mode with finite lifetime. In this situation, light absorption occurs in the graphene layers.

3.2. Additional wave

The nonlocal optical properties of the periodic lattice of graphene layers come largely from the polaritonic mode. On the one hand, the polaritonic band at Kx = 0 coincides well with the pole frequency Ωp of the effective permittivity component εxeff, as shown in Fig. 6. At larger values of Kz, however, a higher order of expansion on the dispersion relation [Eq. (1)] is necessary to deliver a more accurate formula for εxeff. The polaritonic band, or the permittivity pole, varies with the wave vector component Kz over a wide range of frequency from Ω0 to Ω*. This is considered an important feature that characterizes the nonlocal optical properties in graphene layers, as in the case of insulating crystals with excitons [44].

On the other hand, the nonlocal properties associated with the polaritonic band give rise to additional wave. For a wave of frequency Ω, there may exists two Kz components in the graphene layers for a given Kx: a smaller one with the photonic mode and a larger one with the polaritonic mode. This corresponds to the fact that there are two mechanisms of propagating energy in a nonlocal medium: one electromagnetic and one mechanical [44]. As a result, Maxwell’s boundary conditions are insufficient to solve the reflection and transmission coefficients for a nonlocal medium. The additional boundary condition is therefore needed to complete the problem [45].

 figure: Fig. 6

Fig. 6 Pole frequency Ωp for εxeff and its approximate formula [Eq. (9)] as the functions of Kz for the same periodic lattice of graphene layers as in Fig. 1. Black solid lines are photonic and polaritonic modes.

Download Full Size | PDF

3.3. Negative refraction

The nonlocal nature of the polaritonic mode and the associated additional wave results in strong modulation of fields, leading to negative refraction in the graphene layers. In a nonlocal medium with the permittivity tensor εij, the time-averaged Poynting vector is written as [46]:

S=12Re[E×H*]ω4εijkEiEj*.
The energy transport is modified by an extra term related to the change rate of the permittivity tensor with respect to the wave vector, which is also termed as the mechanical Poynting vector [47]. The energy flow may deviate much from the electromagnetic Poynting vector when the nonlocal effect is significant. Consider a wave incident from vacuum onto the graphene layers with the interface located on the xy plane. The dispersion curves on the wave vector domain, along with the Poynting and wave vectors, are plotted in Fig. 7. Above the frequency Ω0 ≈ 0.4306, the dispersion curves exhibit strong modulations and separate into two parts [Fig. 7(a)]. The inner part belongs to the photonic mode and becomes flattened in shape. The outer part belongs to the polaritonic mode and orients in a manner that for a wave incident from vacuum with a certain angle of incidence, the Poynting vector in the medium orients toward the different side of the interface normal (the z axis) than the wave vector, which indicates the occurrence of negative refraction.

 figure: Fig. 7

Fig. 7 Dispersion curves on the wave vector domain at (a) Ω = 0.435 and (b) Ω = 0.35 for the same periodic lattice of graphene layers as in Fig. 1. Black and gray contours are equifrequency curves for vacuum and graphene layers, respectively. Dashed lines indicate the continuity of Kx at the interface.

Download Full Size | PDF

If is worthy of noting that the negative refraction in the underlying structure differs from that in the left-handed metamaterial with a negative refractive index [48], which is accompanied by a backward wave (Kz < 0). Here, the graphene layers are highly anisotropic and the negative refraction occurs as a forward wave (Kz > 0). The negative refraction, however, is different from that in an ordinary (local) anisotropic medium with a hyperbolic dispersion relation, where the normal (to interface) permittivity component εzeff<0 and the tangential component εxeff>0 [49]. In Fig. 7(a), where Ω = 0.435 is slightly above Ω0, the negative refraction in graphene layers occurs when εzeff>0 and εxeff<0 (see the insets). In a local medium, the same condition would result in a backward wave with ordinary (positive) refraction.

The above feature can be explained by the parabolic-like dispersion associated with the polaritonic mode (or additional wave), which is characterized by a quartic curve (in the xy plane) [32]:

x2(1εx2)a2+y2b2=1,
where a and b are constants, and ε is a small positive number. The parabolic-like character of the quartic curve comes from the x4 term at larger x, corresponding to the wave vector dependence of the effective permittivity in the present problem. At smaller x, on the other hand, the dispersion is dominant by the elliptic-like character. In Fig. 7(b), where Ω = 0.35 is well below Ω0, the nonlocal effect is weak and the negative refraction is basically similar to that in a local medium (see the insets for εzeff and εxeff).

3.4. Graphene plasmons

The nonlocal optical properties discussed in the preceding subsections are pertaining to the resonance with varying frequency on the wave vector domain. The physical origin of the nonlocal resonance can be identified as the excitation of plasmons on the graphene surface in two aspects. First, the field patterns on the polaritonic band depicts a typical feature of surface wave. The contours of magnetic field Hy, overlaid with the electric field vectors (Ez, Ex), are plotted in Fig. 8(a) for Ω = 0.6. Both fields decay exponentially away from the graphene surface and become more evanescent as the frequency increases [Fig. 8(b)]. This wave, however, is not a surface mode in the usual sense. It is rather the surface part of a bulk mode that exists in the medium [41]. This feature manifests the couplings of fields between the graphene layers [50]. Across the graphene surface, Hy is discontinuous (and antisymmetric) by the amount of surface current Js = σE [Fig. 8(b)]. Note that this wave is similar the bonding mode in the periodic lattice of thin metal films [32], with a symmetric alignment of surface charges along the interface. The graphene indeed can be treated as a one-atom-layer thick metal as the chemical potential μ ≠ 0 (away from the Dirac point). The crucial difference is that the graphene layers can only support the bonding mode due to the effective zero thickness for the two-dimensional electronic system, while both the bonding and antibonding modes exist in the layers of metal films with finite thickness.

 figure: Fig. 8

Fig. 8 Magnetic field (Hy) (a) on the xz plane and (b) along the x axis for the surface mode at the polaritonic band for the same periodic lattice of graphene layers as in Fig. 1, where Kx = 0, Kz = 3.919, and Ω = 0.6. In (a), red and green colors correspond to positive and negative values of Hy, respectively, gray arrows are the electric field vectors (Ez, Ex), and black lines are the locations of graphene layers. In (b), Hy is normalized by its maximum value (at the graphene surface). The profile for Ω = 0.9 is also shown for comparison.

Download Full Size | PDF

Second, the surface-like wave on the polaritonic band has points of resemblance to the plasmons in a single graphene layer, characterized by the relation [18, 19, 34, 36]:

kz2εk02ε=2ik0σ˜.
It is shown in Fig. 9 that the polaritonic band of the graphene layers approaches toward the plasmon dispersion of a single graphene layer as kz increases, where kz = Kz/a. The two curves almost coincide at sufficiently large kz. There is, however, a discrepancy at small kz, where the plasmon dispersion of the single layer (red dashed line) scales as kz and is approximated, using Eq. (3), as
Ω=2αβkzε,
where β = h̄c/μ. Nevertheless, the polaritonic band of the graphene layers has a cutoff, owing to the interaction of fields between adjacent layers. At small kz, the fields spread between the graphene layers and tend to be tightly bound on the graphene surface as kz increases. At large kz, where the retardation effects are not important ( Ωβkz/ε), both dispersion curves approach asymptotically to the critical frequency Ω* ≈ 1.667, at which σ = 0 [cf. Eq. (3)]. Beyond Ω*, σ < 0 and the graphene plasmons corresponding to TM modes no longer exist [38].

 figure: Fig. 9

Fig. 9 Plasmon relation of a single graphene layer with the same chemical potential and dielectric background as in Fig. 1. Photonic and polaritonic modes for the periodic lattice of graphene layers are shown for comparison.

Download Full Size | PDF

4. Concluding remarks

In conclusion, we have investigated the nonlocal optical properties in the periodic lattice of graphene layers from the view point of effective medium. Approximate formulas of the effective permittivity tensor derived from the nonlocal effective medium model show explicit dependence on the wave vector, indicating the nonlocal nature of the graphene layers. Strong nonlocal effects are found in a wide frequency range (from the effective plasma frequency Ω0 to the critical frequency Ω*) and manifest on the emergence of additional wave and the occurrence of negative refraction. These nonlocal properties are attributed to the excitation of graphene plasmons and are well characterized by the nonlocal effective permittivity for the graphene layers.

Appendix

Let the graphene surface locate at x = 0. For TM polarization, the y component of magnetic field in a unit cell (ξa < x < ξ, ξ > 0) is given by

Hy(x)={Aeiqx+Beiqx,0<x<ξCeiqx+Deiqx,ξa<x<0
where q=εk02kz2 and A, B, C, and D are coefficients to be determined. The z component of the corresponding electric field is determined by Ampere-Maxwell’s law: ∇ × H = −iωε0εE as
Ez(x)={qωε0ε(AeiqxBeiqx),0<x<ξqωε0ε(CeiqxDeiqx),ξa<x<0
The boundary conditions at the graphene surface are given by
Hy(0+)σEz(0+)=Hy(0),
Ez(0+)=Ez(0),
where the second term on the left side of Eq. (16) accounts for the surface current. At the unit cell boundary, the fields satisfy
Hy(ξ)=eikxaHy(ξa),
Ez(ξ)=eikxaEz(ξa).
The nontrivial solutions for A, B, C, and D require that
|1+σqωε0ε1σqωε0ε111111eiqξeiqξeikxaeiq(ξa)eikxaeiq(ξa)eiqξeiqξeikxaeiq(ξa)eikxaeiq(ξa)|=0,
which is simplified to the equation:
cos(kxa)=cos(qa)iσq2ωε0εsin(qa).
This is the dispersion relation o the periodic lattice of graphene layers for TM polarization. Similar procedures can be performed for TE polarization and give rise to the corresponding dispersion relation:
cos(kxa)=cos(qa)iσωμ02qsin(qa).

Acknowledgments

The authors thank Prof. C. T. Chan and Prof. Z. Q. Zhang at Hong Kong University of Science and Technology for helpful discussion. This work was supported in part by National Science Council of the Republic of China under Contract No. NSC 102-2221-E-002-202-MY3 and by the National Natural Science Foundation of China under Grant No. 11304038.

References and links

1. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, “Electric Field Effect in Atomically Thin Carbon Films,” Science 306, 666–669 (2004). [CrossRef]   [PubMed]  

2. K. Novoselov, A. K. Geim, S. Morozov, D. Jiang, M. K. I. Grigorieva, S. Dubonos, and A. Firsov, “Two-dimensional gas of massless Dirac fermions in graphene,” Nature 438, 197–200 (2005). [CrossRef]   [PubMed]  

3. Y. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim, “Experimental observation of the quantum Hall effect and Berry’s phase in graphene,” Nature 438, 201–204 (2005). [CrossRef]   [PubMed]  

4. K. I. Bolotin, K. J. Sikes, Z. Jiang, M. Klima, G. Fudenberg, J. Hone, P. Kim, and H. L. Stormer, “Ultrahigh electron mobility in suspended graphene,” Solid State Commun. 146, 351–355 (2008). [CrossRef]  

5. A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao, and C. N. Lau, “Superior thermal conductivity of single-layer graphene,” Nano Lett. 8, 902–907 (2008). [CrossRef]   [PubMed]  

6. K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn, P. Kim, J.-Y. Choi, and B. H. Hong, “Large-scale pattern growth of graphene films for stretchable transparent electrodes,” Nature 457, 706–710 (2009). [CrossRef]   [PubMed]  

7. F. Xia, T. Mueller, Y.-m. Lin, A. Valdes-Garcia, and P. Avouris, “Ultrafast graphene photodetector,” Nat. Nanotechnol. 4, 839–843 (2009). [CrossRef]   [PubMed]  

8. Z. Sun, T. Hasan, F. Torrisi, D. Popa, G. Privitera, F. Wang, F. Bonaccorso, D. M. Basko, and A. C. Ferrari, “Graphene mode-locked ultrafast laser,” ACS Nano 4, 803–810 (2010). [CrossRef]   [PubMed]  

9. M. Liu, X. Yin, E. Ulin-Avila, B. Geng, T. Zentgraf, L. Ju, F. Wang, and X. Zhang, “A graphene-based broadband optical modulator,” Nature 474, 64–67 (2011). [CrossRef]   [PubMed]  

10. P.-Y. Chen and A. Alù, “Atomically thin surface cloak using graphene mnolayers,” ACS Nano 5, 5855–5863 (2011). [CrossRef]   [PubMed]  

11. A. Vakil and N. Engheta, “Transformation optics using graphene,” Science 332, 1291–1294 (2011). [CrossRef]   [PubMed]  

12. R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R. Peres, and A. K. Geim, “Fine structure constant defines visual transparency of graphene,” Science 320, 1308 (2008). [CrossRef]   [PubMed]  

13. K. F. Mak, M. Y. Sfeir, Y. Wu, C. H. Lui, J. A. Misewich, and T. F. Heinz, “Measurement of the optical conductivity of graphene,” Phys. Rev. Lett. 101, 196405 (2008). [CrossRef]   [PubMed]  

14. A. B. Kuzmenko, E. van Heumen, F. Carbone, and D. van der Marel, “Universal optical conductance of graphite,” Phys. Rev. Lett. 100, 117401 (2008). [CrossRef]   [PubMed]  

15. T. Stauber, N. M. R. Peres, and A. K. Geim, “Optical conductivity of graphene in the visible region of the spectrum,” Phys. Rev. B 78, 085432 (2008). [CrossRef]  

16. Z. Li, E. A. Henriksen, Z. Jiang, Z. Hao, M. C. Martin, P. Kim, H. Stormer, and D. N. Basov, “Dirac charge dynamics in graphene by infrared spectroscopy,” Nat. Phys. 4, 532–535 (2008). [CrossRef]  

17. F. Wang, Y. Zhang, C. Tian, C. Girit, A. Zettl, M. Crommie, and Y. R. Shen, “Gate-variable optical transitions in graphene,” Science 320, 206–209 (2008). [CrossRef]   [PubMed]  

18. Y. Liu, R. F. Willis, K. V. Emtsev, and T. Seyller, “Plasmon dispersion and damping in electrically isolated two-dimensional charge sheets,” Phys. Rev. B 78, 201403 (2008). [CrossRef]  

19. M. Jablan, H. Buljan, and M. Soljačić, “Plasmonics in graphene at infrared frequencies,” Phys. Rev. B 80, 245435 (2009). [CrossRef]  

20. L. Ju, B. Geng, J. Horng, C. Girit, M. Martin, Z. Hao, H. A. Bechtel, X. Liang, A. Zettl, and Y. R. Shen, “Graphene plasmonics for tunable terahertz metamaterials,” Nat. Nanotechnol. 6, 630–634 (2011). [CrossRef]   [PubMed]  

21. A. Y. Nikitin, F. Guinea, F. J. García-Vidal, and L. Martín-Moreno, “Edge and waveguide terahertz surface plasmon modes in graphene microribbons,” Phys. Rev. B 84, 161407 (2011). [CrossRef]  

22. T. Stauber and G. Gómez-Santos, “Plasmons and near-field amplification in double-layer graphene,” Phys. Rev. B 85, 075410 (2012). [CrossRef]  

23. F. H. L. Koppens, D. E. Chang, and F. J. García de Abajo, “Graphene plasmonics: a platform for strong light-matter interactions,” Nano Lett. 11, 3370–3377 (2011). [CrossRef]   [PubMed]  

24. L. A. Falkovsky and S. S. Pershoguba, “Optical far-infrared properties of a graphene monolayer and multilayer,” Phys. Rev. B 76, 153410 (2007). [CrossRef]  

25. R. E. V. Profumo, M. Polini, R. Asgari, R. Fazio, and A. H. MacDonald, “Electron-electron interactions in decoupled graphene layers,” Phys. Rev. B 82, 085443 (2010). [CrossRef]  

26. T. Zhan, X. Shi, Y. Dai, X. Liu, and J. Zi, “Transfer matrix method for optics in graphene layers,” J. Phys.: Condens. Matter 25, 215301 (2013).

27. V. L. Ginzburg, A. A. Rukhadze, and V. P. Silin, “The electrodynamics of crystals and the theory of excitons,” J. Phys. Chem. Solids 23, 85–97 (1962). [CrossRef]  

28. R. M. Hornreich and S. Shtrikman, “Theory of gyrotropic birefringence,” Phys. Rev. 171, 1065–1074 (1968). [CrossRef]  

29. P. A. Belov, R. Marques, S. I. Maslovski, I. S. Nefedov, M. Silveirinha, C. R. Simovski, and S. A. Tretyakov, “Strong spatial dispersion in wire media in the very large wavelength limit,” Phys. Rev. B 67, 113103 (2003). [CrossRef]  

30. C. R. Simovski and P. A. Belov, “Low-frequency spatial dispersion in wire media,” Phys. Rev. E 70, 046616 (2004). [CrossRef]  

31. A. V. Chebykin, A. A. Orlov, C. R. Simovski, Y. S. Kivshar, and P. A. Belov, “Nonlocal effective parameters of multilayered metal-dielectric metamaterials,” Phys. Rev. B 86, 115420 (2012). [CrossRef]  

32. R.-L. Chern, “Spatial dispersion and nonlocal effective permittivity for periodic layered metamaterials,” Opt. Express 21, 16514–16527 (2013). [CrossRef]   [PubMed]  

33. J. Elser, V. A. Podolskiy, I. Salakhutdinov, and I. Avrutsky, “Nonlocal effects in effective-medium response of nanolayered metamaterials,” Appl. Phys. Lett. 90, 191109 (2007). [CrossRef]  

34. B. Wunsch, T. Stauber, F. Sols, and F. Guinea, “Dynamical polarization of graphene at finite doping,” New J. Phys. 8, 318 (2006). [CrossRef]  

35. V. P. Gusynin, S. G. Sharapov, and J. P. Carbotte, “Unusual Microwave Response of Dirac Quasiparticles in Graphene,” Phys. Rev. Lett. 96, 256802 (2006). [CrossRef]   [PubMed]  

36. E. H. Hwang and S. Das Sarma, “Dielectric function, screening, and plasmons in two-dimensional graphene,” Phys. Rev. B 75, 205418 (2007). [CrossRef]  

37. L. A. Falkovsky and A. A. Varlamov, “Space-time dispersion of graphene conductivity,” Eur. Phys. J. B 56, 281–284 (2007). [CrossRef]  

38. S. A. Mikhailov and K. Ziegler, “New electromagnetic mode in graphene,” Phys. Rev. Lett. 99, 016803 (2007). [CrossRef]   [PubMed]  

39. L. D. Landau, E. M. Lifshitz, and L. P. Pitaevski, Electrodynamics of Continuous Media, 2 (Butterworth-Heinenan, 1984).

40. V. M. Agranovich and V. L. Ginzburg, Crystal Optics with Spatial Dispersion, and Excitons (Springer-Verlag, 1984). [CrossRef]  

41. G. D. Mahan and G. Obermair, “Polaritons at surfaces,” Phys. Rev. 183, 834–841 (1969). [CrossRef]  

42. L. D. Landau and E. M. Lifshitz, Quantum Mechanics: Non-relativistic Theory, 3 (Butterworth-Heinenan, 1977).

43. X. Xu, Y. Xi, D. Han, X. Liu, J. Zi, and Z. Zhu, “Effective plasma frequency in one-dimensional metallic-dielectric photonic crystals,” Appl. Phys.Lett. 86, 091112 (2005).

44. J. J. Hopfield and D. G. Thomas, “Theoretical and experimental effects of spatial dispersion on the optical properties of crystals,” Phys. Rev. 132, 563–572 (1963). [CrossRef]  

45. S. I. Pekar, “Theory of electromagnetic waves in a crystal with excitons,” J. Phys. Chem. Solids 5, 11–22 (1958). [CrossRef]  

46. V. M. Agranovich and Y. N. Gartstein, “Spatial dispersion and negative refraction of light,” Phys. Usp. 49, 1029 (2006). [CrossRef]  

47. M. F. Bishop and A. A. Maradudin, “Energy flow in a semi-infinite spatially dispersive absorbing dielectric,” Phys. Rev. B 14, 3384–3393 (1976). [CrossRef]  

48. S. Foteinopoulou, E. N. Economou, and C. M. Soukoulis, “Refraction in Media with a Negative Refractive Index,” Phys. Rev. Lett. 90, 107402 (2003). [CrossRef]   [PubMed]  

49. P. A. Belov, “Backward waves and negative refraction in uniaxial dielectrics with negative dielectric permittivity along the anisotropy axis,” Microw. Opt. Technol. Lett. 37, 259–263 (2003). [CrossRef]  

50. B. Wang, X. Zhang, F. J. García-Vidal, X. Yuan, and J. Teng, “Strong coupling of surface plasmon polaritons in monolayer graphene sheet arrays,” Phys. Rev. Lett. 109, 073901 (2012). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1
Fig. 1 Schematic diagram of the periodic lattice of graphene layers with period a embedded in a background with dielectric constant ε. In this study, a = 0.1 h̄c/μ (≈ 98.9 nm) with μ = 0.2 eV and ε = 1.5 will be used as the parameters. A small area of the graphene feature is shown on the top layer for illustration.
Fig. 2
Fig. 2 (a) In-plane effective permittivity components ε z eff and ε x eff as the functions of Ω for the same periodic lattice of graphene layers as in Fig. 1, where Ω = h̄ω/μ, Kx = kxa, Kz = kza, and ã = μa/(h̄c). (b) Effective plasma frequency Ω0 and its approximate formula [Eq. (8)] as the functions of ã.
Fig. 3
Fig. 3 Out-of-plane effective permittivity component ε y eff for the same periodic lattice of graphene layers as in Fig. 1, where (a) Kx = 0 and (b) Kz = 0. Red and green lines correspond to ε y eff = 0 and ε y eff = ε , respectively.
Fig. 4
Fig. 4 Equifrequency surfaces of (a) TM and (b) TE dispersion relations for the same periodic lattice of graphene layers as in Fig. 1. Black circle in (b) is the section of light cone at Ω = 2.
Fig. 5
Fig. 5 (a) TM frequency bands at Kx = 0, 0.03 and (b) TE frequency band at Kx = 0 for the same periodic lattice of graphene layers as in Fig. 1. Insets in (a) are typical magnetic field patterns of photonic mode (P mode) and polaritonic mode (PL mode). Red dot in (b) corresponds to the critical frequency Ω* ≈ 1.667 that separates P mode and N mode.
Fig. 6
Fig. 6 Pole frequency Ωp for ε x eff and its approximate formula [Eq. (9)] as the functions of Kz for the same periodic lattice of graphene layers as in Fig. 1. Black solid lines are photonic and polaritonic modes.
Fig. 7
Fig. 7 Dispersion curves on the wave vector domain at (a) Ω = 0.435 and (b) Ω = 0.35 for the same periodic lattice of graphene layers as in Fig. 1. Black and gray contours are equifrequency curves for vacuum and graphene layers, respectively. Dashed lines indicate the continuity of Kx at the interface.
Fig. 8
Fig. 8 Magnetic field (Hy) (a) on the xz plane and (b) along the x axis for the surface mode at the polaritonic band for the same periodic lattice of graphene layers as in Fig. 1, where Kx = 0, Kz = 3.919, and Ω = 0.6. In (a), red and green colors correspond to positive and negative values of Hy, respectively, gray arrows are the electric field vectors (Ez, Ex), and black lines are the locations of graphene layers. In (b), Hy is normalized by its maximum value (at the graphene surface). The profile for Ω = 0.9 is also shown for comparison.
Fig. 9
Fig. 9 Plasmon relation of a single graphene layer with the same chemical potential and dielectric background as in Fig. 1. Photonic and polaritonic modes for the periodic lattice of graphene layers are shown for comparison.

Equations (22)

Equations on this page are rendered with MathJax. Learn more.

cos ( k x a ) = cos ( q a ) i σ q 2 ω ε 0 ε sin ( q a ) ,
cos ( k x a ) = cos ( q a ) i σ ω μ 0 2 q sin ( q a ) ,
σ ε 0 c = 4 α i Ω + π α [ θ ( Ω 2 ) + i π ln | Ω 2 Ω + 2 | ] ,
k x 2 ε z eff + k z 2 ε x eff = k 0 2 ,
k x 2 + k z 2 = ε y eff k 0 2
ε z eff = ε z 0 γ 12 k 0 2 a 2 1 1 12 k x 2 a 2 , ε x eff = ε ( 1 γ 12 ε z 0 k 0 2 a 2 ) 1 γ 6 ε z 0 ( k 0 2 a 2 1 2 ε k z 2 a 2 ) ,
ε y eff = ε y 0 ( 1 + 1 6 k z 2 a 2 ) δ 6 ε 2 k 0 2 a 2 + a 2 12 k 0 2 ( k x 4 k z 4 ) ,
Ω 0 2 ( 1 + ε a ˜ α ) 1 / 2 ,
Ω p 2 [ 1 + ε a ˜ ( 12 + K z 2 ) 2 α ( 6 + K z 2 ) ] 1 / 2 ,
S = 1 2 Re [ E × H * ] ω 4 ε i j k E i E j * .
x 2 ( 1 ε x 2 ) a 2 + y 2 b 2 = 1 ,
k z 2 ε k 0 2 ε = 2 i k 0 σ ˜ .
Ω = 2 α β k z ε ,
H y ( x ) = { A e i q x + B e i q x , 0 < x < ξ C e i q x + D e i q x , ξ a < x < 0
E z ( x ) = { q ω ε 0 ε ( A e i q x B e i q x ) , 0 < x < ξ q ω ε 0 ε ( C e i q x D e i q x ) , ξ a < x < 0
H y ( 0 + ) σ E z ( 0 + ) = H y ( 0 ) ,
E z ( 0 + ) = E z ( 0 ) ,
H y ( ξ ) = e i k x a H y ( ξ a ) ,
E z ( ξ ) = e i k x a E z ( ξ a ) .
| 1 + σ q ω ε 0 ε 1 σ q ω ε 0 ε 1 1 1 1 1 1 e i q ξ e i q ξ e i k x a e i q ( ξ a ) e i k x a e i q ( ξ a ) e i q ξ e i q ξ e i k x a e i q ( ξ a ) e i k x a e i q ( ξ a ) | = 0 ,
cos ( k x a ) = cos ( q a ) i σ q 2 ω ε 0 ε sin ( q a ) .
cos ( k x a ) = cos ( q a ) i σ ω μ 0 2 q sin ( q a ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.