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

Tunable Goos-Hänchen shift from graphene ribbon array

Open Access Open Access

Abstract

The Goos-Hänchen (GH) shift of light beam incident on graphene ribbon array is investigated by Green’s function method. Due to the resonance effects of leaky surface plasmons on ribbons, the zeroth-order reflection field shows both giant positive and negative GH shifts. By tuning the graphene Fermi level, we can control the shift conveniently. This effect is important to graphene-based metasurface and electro-optical devices.

© 2017 Optical Society of America

1. Introduction

Goos-Hänchen (GH) shift refers to a lateral shift between the reflected and the incident beams, which is named after Goos and Hänchen who observed this effect firstly by using multiple reflections in a glass slab [1–15]. Afterwards, Artmann and other authors theoretically interpreted this effect [2]. In their theories, the stationary light beam is expanded as a power series of plan-wave components. After the reflection, different components undergo different phase and amplitude changes, which results in a total spatial lateral shift. Lots of structures for generating large GH shifts have been investigated [4–12], such as waveguide, photonic crystal, and left-handed metamaterials, including both positive and negative shifts. Additionally, obvious GH shift also occurs in a surface grating [7–12].

Graphene, a single layer of carbon atoms arranged in a honeycomb lattice, has attracted tremendous interest [16–36]. The linear dispersion relation near the Dirac point of the energy band induces a special optical response to the light. Especially the graphene plasmons (GPs) have emerged as a hot topic in recent years due to their new frequency region, tunability, long-lived and extreme light confinement. Additionally, due to Pauli-blocking effect, doped graphene has a low absorption in the terahertz to mid-infrared.

To investigate the reflectivity of light beam by graphene, we need to enhance the coupling between the beam and the graphene due to little light can be reflected by monolayer graphene at small incident angle. One practical way is to excite the GPs [21]. However, as we know, GPs have large wave numbers [22] and are hard to be excited by a beam incident on the graphene directly. Periodic array of graphene ribbons, which can be considered to be a grating [37], has attracted a lot of attention because the GPs can be easily excited in the system [30, 34]. Light beam incident on the array can be converted to the leaky resonant graphene plasmons (RGPs) efficiently if the ribbon width satisfies the resonance conditions, which will induce large reflection. Additionally, since light waves with different incident angles have different coupling strengths as well as phase changes after the reflection, we can expect large GH shifts of the reflection fields.

In this article, formulas to calculate the reflections as well as GH shifts of light beams by Green’s tensor method are derived. The numerical results indicates that light beams incident on graphene ribbon array have either large positive or negative GH shifts at certain incident angles, which is qute different with monolayer graphene [38]. And, more remarkable, due to the tunability of graphene Fermi level, the plasmonic resonance conditions can be manipulated and consequently the GH shift can be controlled conveniently. These effects are helpful to the investigation of graphene-based metasurface [30, 39].

The paper is organized as follows. In Sec. 2, we introduce our model and mathematic theory. In Sec. 3, we present some numerical results and simulations. In Sec. 4, we present the concluding remarks.

2. Green’s tensor method to solve the scattering

Our model is schematically shown in Fig. 1. The graphene ribbon array is located at z = 0 along y direction and between two dielectric half spaces with dielectric constants ε1 and ε2, respectively. The ribbon has a width b and period a. A Gaussian beam is incident on the array with pz plane as the incident plane. The beam direction has a polar angle θ (with respect to the z axis) and an azimuthal angle φ (with respect to the x axis).

 figure: Fig. 1

Fig. 1 A Gaussian beam is incident on the periodic graphene ribbon array. The Fermi level of the ribbons can be tuned by the gate voltage.

Download Full Size | PDF

In the long wavelength and high doping limit, i.e., ħω ⩾ 2EF, the in-plane conductivity of graphene can be described as [25, 28]

σ(ω)=ie2EFπ2(ω+iτ1)
under the random phase approximation (RPA). Here, EF is the Fermi level, e is the electron charge, and τ describes the momentum relaxation time. Under certain conditions, τ is identified with DC relaxation time and can be expressed as τ=μEF/evF2 [33], where μ is the DC mobility and can reach 104−6cm2/Vs [29]. The tunability of the graphene originates from the controllability of the Fermi level. For example as in Fig. 1, the Fermi level can be manipulated by controlling the gate voltage.

Here, we use Green’s tensor method to deal with the scattering of the light. A plane wave with an electric field component parallel to the xy plane is incident on the ribbon array can induce surface electric current η(x, y) in the graphene layer. The field produced by each surface element dxdy′ is E(x, y) = G(xx′, yy′; ω)·((x′, y′))dxdy′, where G is the Green’s tensor [25, 40, 41]. Summing over all of these surface current contributions, and including the effect of a substrate through its Fresnel coefficients, we obtain the self-consistent relations (see the Appendix)

ηx(x,y)σ(x,y)=Ex0(x,y,0)+12(2π)3iωdkxdkydxdy2πiβk02ε1{[ky2ρ2(1+rs)+1k02kx2β2ρ2(1rp)]ηx(x,y)+[kxkyρ2(1+rs)+1k02kxkyβ2ρ2(1rp)]ηy(x,y)}ei[kx(xx)+ky(yy)];
and
ηy(x,y)σ(x,y)=Ey0(x,y,0)+12(2π)3iωdkxdkydxdy2πiβk02ε1{[kxkyρ2(1+rs)+1k02kxkyβ2ρ2(1rp)]ηx(x,y)+[kx2ρ2(1+rs)+1k02ky2β2ρ2(1rp)]ηy(x,y)}ei[kx(xx)+ky(yy)],
where ηx and ηy are the induced surface electric currents along x and y directions, ρ=(kx2+ky2)1/2 and β=(ε1k02ρ2)1/2 is the wave vector perpendicular to the ribbon array. rs and rp are the Fresnel reflection coefficients of the substrate for TE and TM polarized waves. Ex0 and Ey0 are the external electric fields components which already include the reflection by the homogeneous dielectric substrate.

Since the conductivity is periodic along the x direction and homogenous along the y direction, we can expand the conductivity along the x direction as Fourier series given by

σ(x,y)=nσneignx,
with gn = nΛ, where Λ = 2π/a is the reciprocal wave vector of the ribbon array. The surface current can be expanded as
ηx,y(x,y)=nηx,ynei[(kx0+gn)x+ky0y],
where kx 0=ε1k0cosθ and ky 0=ε1k0sinθ sin θ are the wave number components of the incident field along x and y directions, respectively. After some straightforward algebra, we project Eqs. (2) and (3) into
ηxn=1a0bdxEx0(x,y,0)σei(kxx+kyy)12ωnk02ε1β{[ky2ρ2(1+rs)+1k02kx2β2ρ2(1rp)]ηxnσnn+[kxkyρ2(1+rs)+1k02kxkyβ2ρ2(1rp)]ηynσnn};
and
ηyn=1a0bdxEy0(x,y,0)σei(kxx+kyy)12ωnk02ε1β{[kxkyρ2(1+rs)+1k02kxkyβ2ρ2(1rp)]ηxnσnn+[kx2ρ2(1+rs)+1k02ky2β2ρ2(1rp)]ηynσnn}.
Here, β and ρ are the same as those in Eqs. (2) and (3) but for kx = kx0 + gn and ky = ky0. Finally, we solve Eqs. (4) and (5) by using standard linear algebra with a finite number of waves M. The zeroth-order reflection field is given in terms of the η0 coefficients as
Ex(x,y,z)=Exr(x,y,z)i2ωk02ε1βei[kx0x+ky0y+βz]{[ky02ρ2(1+rs)+1k02kx02β2ρ2(1rp)]ηx0+[kx0ky0ρ2(1+rs)+1k02kx0ky0β2ρ2(1rp)]ηy0}=|rx|eiϕEx0(x,y,z)e2iβz;
and
Ey(x,y,z)=Eyr(x,y,z)i2ωk02ε1βei[kx0x+ky0y+βz]{[kx0ky0ρ2(1+rs)+1k02kx0ky0β2ρ2(1rp)]ηx0+[kx02ρ2(1+rs)+1k02ky02β2ρ2(1rp)]ηy0}=|ry|eiϕEy0(x,y,z)e2iβz.
Here, Exr(x,y,z) and Eyr(x,y,z) are the direct reflection of the incident field by the dielectric substrate, and ϕ is the phase shift.

For the incident beam with a sufficiently large beam waist (i.e., the beam with a narrow angular spectrum, Δkk0, where Δk = 1/w0 and w0 is the half width of the beam at waist), the GH shift of the zeroth-order reflection beam can be expressed as [4]

S=λ02πdϕdθ,
where λ0 is the vacuum wavelength.

3. Numerical calculations of Goos-Hänchen shift

In this section, we numerically calculate the reflectivities, phase changes, and GH shifts of the reflection fields. In the calculations, we set M to be 700, which can give us precious numerical results. We focus on two cases that TM modes with azimuthal angle ϕ to be 0 and TE mode with ϕ to be π/2.

In Fig. 2, we show the zeroth-order reflection spectrums for light waves incident on the array perpendicularly. Here, b = 4.8μm, a = 49.8μm, and τ = 1ps. For the convenience of calculations, we set ε1 = ε2 = 1. Since ab, we can neglect the interference between the ribbons. Remarkably, the ribbon is wide enough that we don’t need to consider the edge effect of the ribbon and adopt the conductivity shown in Eq. (1) [36]. The peak frequencies satisfy the plasmonic resonance conditions

nπb(1+ε2)ω24α0cωF.
Here ωF = EF, α0 is the fine-structure constant, and integers n = 1, 2, 3… measure the number of half wavelengths that fit within the ribbon width. The right three high peak frequencies correspond to n = 1 (electric dipole resonance), denoted by ωEF1, and the three left low peak frequencies correspond to n = 2 (electric quadrupole resonance), denoted by ωEF2 [34]. The high reflectivity originates from the strong coupling strength between the incident field and the RGPs. When EF = 0.6eV, the vacuum wave number of the peak frequency ω0.61 is larger than Λ, thus the first-order diffraction field is still in the far field region and has weak coupling with the RGPs. The reflection mainly comes from the coupling between the higher-order diffraction modes and the RGPs. However, when EF = 0.4, 0.5eV, the first-order diffraction fields with frequencies ω0.4,0.51 are evanescent waves and concentrated in the proximity of the ribbon array. This leads to stronger coupling with the RGPs and consequently higher reflectivities.

 figure: Fig. 2

Fig. 2 The zeroth-order reflection spectrum for the ribbon array with different Fermi levels.

Download Full Size | PDF

In Fig. 3(a), we show the reflection of TM waves incident on the array with an incident angle θ. Two peaks appear in the spectrum on the Rayleigh anomalies [42]. The left (right) peak position corresponds to θ1 = arcsin[(k0 − Λ)/k0] (θ−2 = arcsin[(2Λ − k0)/k0]), where the plus first (minus second)-order diffraction component of the incident wave becomes evanescent wave and has large coupling strength with the RGPs. As θ increases, the x-direction electric field of the incident beam decreases, which leads to smaller surface currents [25] and consequently decreasing reflectivity as shown in Fig. 3(a).

 figure: Fig. 3

Fig. 3 (a) The zeroth-order reflectivities of TM waves with φ = 0 at different Fermi levels. The labels (1) and (2) represent the plus first and minus second-order diffraction positions. The wavelength of the incident beam is 45.2μm and the other parameters are the same as Fig. 2. (b) The arguments ϕ of the zeroth-order reflection field. (c) and (d) The corresponding GH shifts.

Download Full Size | PDF

The GH shift can be calculated by using Eqs. (810). If ε2 = 1 and we just consider the zeroth-order reflection, ηx0 dominates the reflection field that we are interested in, thus we can obtain the phase shift

ϕ=Arg(ηx0).

In Fig. 3(b), we plot ϕ. When EF = 0.6eV, intense variations and even sudden jumps happen, especially near the Rayleigh anomalies. The corresponding GH shifts are shown in Figs. 3(c) and (d). We can see giant GH shifts when EF = 0.6eV. These results can be understood by Artmann’s theory. As the incident angular approaches and exceeds θ1, the z direction wave vector of the first-order diffraction field varies from real to zero and then to pure imaginary. This means that the coupling strength between the diffraction field and the RGPs increases intensely and the phase ϕ varies sharply. As a consequence, large GH shift appears. However, when the incident angular is much larger than θ1, the coupling strength and phase ϕ vary relatively slowly, therefore the total GH shift approaches zero [4]. It can be interpreted similarly for the case that the incident angle approaches θ−2. Additionally, as we know, the derivation of Eq. (10) in the article is based on the condition that the phase gradient is almost a constant in the angular spectrum. This means that at the points θ = θ1 and θ−2, Eq. (10) is not valid anymore. This is why there are poles in Figs. 3(c) and (d). However, a real optical beam can be expended as a series of plane waves around these angles and the total GH shift is the coherent summation of the contributions from all the plane waves. The waves with the incident angles θ1 and θ−2 will have little influence on the total GH shifts and the GH shifts will always be finite. In the following, we will give some simulations for real optical beams.

When EF = 0.4 or 0.5eV, the incident fields are not resonant with the ribbon array and there are no RGPs. The coupling strengths between the incident beams and array are weak and the reflectivities are small as shown in Fig. 3(a). Meanwhile, since all the components of the incident beams have small coupling strengths with the ribbons, which are similar to the cases that the beams are reflected by low reflection dielectric interfaces, the beams have small GH shifts.

Additionally, the Fermi level of the ribbon can be tuned easily, such as by static electric gate shown in Fig. 1. This gives us a way to manipulate the GH shift. In Fig. 4, we plot the GH shifts as a function of the Fermi level. The curves shows that the GH shifts are highly dependent on the Fermi level. Compared to a small ε2, the resonance condition of Eq. (11) requires large EF for a large ε2. We also plot the GH shifts for the ribbons with different relaxation time. For the electric dipole resonance, the polarizability is inverse proportion to the dissipation [34]. As a consequence, large dissipation weakens the coupling between the incident field and the RGPs and slows the φ variation. The GH shift decreases with the dissipation as shown in Fig. 4.

 figure: Fig. 4

Fig. 4 (The GH shifts versus the Fermi level under different ε2 and τ. The parameters are the same as the Fig. 3(a) and θ = 0.0769.

Download Full Size | PDF

The reflection spectrums and GH shifts for the TE waves with φ = π/2, i.e., the waves are incident in the yz plane, are shown in Fig. 5. In contrast to Fig. 3(a) where there are two peaks, there is only one valley in the spectrum at position θ=arcsin1Λ2/k02. The reflectivities approach zero near this point. Meanwhile, both positive and negative GH shifts occur when EF = 0.6eV and only positive shifts exist when EF = 0.4 and 0.5eV. Similar to the case that we discussed in Fig. 4, we can also manipulate the reflectivity and GH shift by controlling the Fermi level of the graphene ribbons.

 figure: Fig. 5

Fig. 5 (a) The zeroth-order reflectivity of TE wave along y direction asa function the incidence angle at different Fermi levels. (b) The corresponding GH shifts.

Download Full Size | PDF

In order to verify our previous theoretical predictions, we simulate the reflection fields of TM polarized Gaussian beams in Fig. 6. The vector electric field of the beam is expressed in terms of its angular spectrum as follows [3, 43]:

E(r)=A(kx0,ky0)eikrdkx0dk0,
where time dependence exp(−iωt) is assumed and suppressed. This beam has a principal axial direction kG = (k0 sin θ0, 0, k0 cos θ0). k = (kx0, ky0, β0) is the wave vector satisfying kx02+ky02+β02=k02. The element A(kx0, ky0) with direction vector (cos θ cos φ, − cos θ sin φ, sin θ) has the following amplitude [3, 43]
A(kx0,ky0)=(wxwyπ)1/2exp[wx22(kx0x0G)2wy22ky02],
where wx = w0/cos θ0 and wy = w0. In our simulation, w0=2502λ0, θ0 = 0.0771 and the other parameters are the same as those in Fig. 3. We utilize the method in the previous section to calculate the reflection fields for all the plane waves k and then summarize these fields associated with the element amplitudes A(kx0, ky0) to obtain the beam intensity distribution.

 figure: Fig. 6

Fig. 6 (a) The normalized field intensity distribution of the incident TM-polarized Gaussian beam on the z = 0 plane. (b, c) The normalized field distributions of the zeroth-order reflection fields for the cases that EF = 0.6eV and 0.5eV.

Download Full Size | PDF

In Fig. 6(b), we show that when EF = 0.6eV, the zeroth-order reflection beam has a lateral shift 3.14λ0, which is approximately equal to the previous theoretical result, i.e., 3.02λ0/cos θ0 = 3.27λ0. While for EF = 0.5eV, the reflection beam nearly has the same position as the incident beam, which also agrees with the previous prediction. These results evidently prove that the GH shift from graphene ribbon array is highly dependent on the Fermi level.

4. Conclusion

In conclusion, we calculate the reflections and GH shifts from graphene ribbon array by Green’s tensor method. Both positive and negative GH shifts can be found. Compared to the monolayer graphene where prism is required to excite the GPs and consequently GH shift [44–46], graphene ribbon array is a simpler and more practical choice to realize giant GH shifts. Additionally, it is easy to control the graphene Fermi level [30], meanwhile, the GH shift is strongly dependent on the Fermi level. Therefore, we can control the GH shifts conveniently. Our investigation is important to the tunable metasurfaces and electro-optical devices based on graphene. Compared with the monolayer graphene used in Refs. 45 and 46, realizing the periodical ribbon array is technologically harder. However, periodical ribbon array systems are widely investigated in other topics and the technology is not a big issue now, such as in Ref. 30 in the revised manuscript. We hope and also believe the experiment can be realized in the near future

Appendix

In this Appendix, we present the derivations of equations (2) and (3). According to the Maxwell’s equation, the electric field E(r, t) = E(r, ω)exp(−iωt) of the oscillatory electric current η(r, t) = η(r, ω)exp(−iωt) satisfies the equation [41]

(××ε(r,ω)ω2c2)E(r,ω)=iωη(r,ω).
In the absence of an external field, E(r, ω) is formally given by
E(r,ω)=iωd3rG(r,r;ω)η(r,ω).
In our system, the Green’s tensor can be described as the summation of plane waves
Gxx(r,r;ω)=12(2π)3dkxdky2πiε1β[ky2ρ2(1+rs)+kx2β2k02ρ2(1rp)]ei[kx(xx)+ky(yy)];
Gxy(r,r;ω)=12(2π)3dkxdky2πiε1β[kxkyρ2(1+rs)+kxkyβ2k02ρ2(1rp)]ei[kx(xx)+ky(yy)];
Gyx(r,r;ω)=Gxy(r,r;ω);
Gyy(r,r;ω)=12(2π)3dkxdky2πiε1β[kx2ρ2(1+rs)+ky2β2k02ρ2(1rp)]ei[kx(xx)+ky(yy)].
Here r|| = (x, y, 0) denotes the spot on the z = 0 plane. The reflection coefficients
rs=(βε2k02ρ2)/(β+ε2k02ρ2);
and
rp=(ε2βε1ε2k02ρ2)/(ε2β+ε1ε2k02ρ2).

Eqs. (1619) can be understood as the x(y) direction field at position r induced by a unit x(y) direction current at position r. By using relation η = σE, we can obtain equations (2) and (3).

Funding

Qatar National Research Fund (QNRF) NPRP Grant No. 8-352-1-074; King Abdulaziz City for Science and Technology (KACST).

References and links

1. F. Goos and H. Hänchen, “Ein neuer und fundamentaler Versuch zur Totalreflexion,” Ann. Phys. (Leipzig) 1, 333 (1947). [CrossRef]  

2. K. Artmann, “Berechnung der Seitenversetzung des totalreflektierten Strahles,” Ann. Phys. (Leipzig) 2, 87 (1948). [CrossRef]  

3. C. F. Li, “Unified theory for Goos-Hänchen and Imbert-Fedorov effects,” Phys. Rev. A 76, 013811 (2007). [CrossRef]  

4. L. G. Wang, H. Chen, and S. Y. Zhu, “Large negative Goos-Hänchen shift from a weakly absorbing dielectric slab, – Opt. Lett. 30(21), 2936–2938 (2005). [CrossRef]   [PubMed]  

5. X. Yin, L. Hesselink, Z. Liu, N. Fang, and X. Zhang, “Large positive and negative lateral optical beam displacements due to surface plasmon resonance,” Appl. Phys. Lett. 85, 372 (2004). [CrossRef]  

6. S. Asiri, J. Xu, M. Al-Amri, and M. S. Zubairy, “Controlling the Goos-Hänchen and Imbert-Fedorov shifts via pump and driving fields,” Phys. Rev. A 93, 013821 (2016). [CrossRef]  

7. R. Yang, W. Zhu, and J. Li, “Giant positive and negative Goos-Hänchen shift on dielectric gratings caused by guided mode resonance,” Opt. Express 22(2), 2043–2050 (2014). [CrossRef]   [PubMed]  

8. M. Merano, A. Aiello, G. W. ’t Hooft, M. P. van Exter, E. R. Eliel, and J. P. Woerdman, “Observation of Goos-Hänchen shifts in metallic reflection,” Opt. Express 15(24), 15928–15934 (2007). [CrossRef]   [PubMed]  

9. G.-Y. Oh, D. G. Kim, and Y. -W. Choi, “The characterization of GH shifts of surface plasmon resonance in a waveguide using the FDTD method,” Opt. Express 17(23), 20714–20720 (2009). [CrossRef]   [PubMed]  

10. C. Bonnet, D. Chauvat, O. Emile, F. Bretenaker, A. L. Floch, and L. Dutriaux, “Measurement of positive and negative Goos-Hänchen effects for metallic gratings near Wood anomalies,” Opt. Lett. 26(10), 666–668 (2001). [CrossRef]  

11. T. Tamir and H. L. Bertoni, “"Lateral Displacement of Optical Beams at Multilayered and Periodic Structures,” J. Opt. Soc. Am. 61(10), 1397–1413 (1971). [CrossRef]  

12. S. Zhang and T. Tamir, “Spatial modifications of Gaussian beams diffracted by reflection gratings,” J. Opt. Soc. Am. 6(9), 1368–1381 (1989). [CrossRef]  

13. L. G. Wang, S. Y. Zhu, and M. S. Zubairy, “Goos-Hänchen Shifts of Partially Coherent Light Fields,” Phys. Rev. Lett. 111, 223901 (2013). [CrossRef]  

14. R. Macedo, R. L. Stamps, and T. Dumelow, “Spin canting induced nonreciprocal Goos-Hänchen shifts,” Opt. Express 22(23), 28467–28478 (2014). [CrossRef]   [PubMed]  

15. L. G. Wang, M. Ikram, and M. S. Zubairy, “Control of the Goos-Hänchen shift of a light beam via a coherent driving field,” Phys. Rev. A 77, 023811 (2008). [CrossRef]  

16. A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nat. Mater. 6, 183–191 (2007). [CrossRef]   [PubMed]  

17. A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, “The electronic properties of graphene,” Rev, Mod. Phys. 81, 109 (2009). [CrossRef]  

18. 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]  

19. J. Chen, M. Badioli, P. Alonso-González, S. Thongrattanasiri, F. Huth, J. Osmond, M. Spasenović, A. Centeno, A. Pesquera, P. Godignon, A. Z. Elorza, N. Camara, F. J. Garcia de Abajo, R. Hillenbrand, and F. H. L. Koppens, “Optical nano-imaging of gate-tunable graphene plasmons,” Nature (London) 487, 77–81 (2012).

20. Z. Fei, A. S. Rodin, G. O. Andreev, W. Bao, A. S. McLeod, M. Wagner, L. M. Zhang, Z. Zhao, M. Thiemens, G. Dominguez, M. M. Fogler, A. H. Castro Neto, C. N. Lau, F. Keilmann, and D. N. Basov, “Gate-tuning of graphene plasmons revealed by infrared nano-imaging,” Nature (London) 487, 82–85 (2012).

21. A. Khavasi, “Fast convergent Fourier modal method for the analysis of periodic arrays of graphene ribbons,” Opt. Lett. 38(16), 3009–3012 (2013). [CrossRef]   [PubMed]  

22. X. D. Zeng, M. Al-Amri, and M. S. Zubairy, “Nanometer-scale microscopy via graphene plasmons,” Phys. Rev. B 90, 235418 (2014). [CrossRef]  

23. X. D. Zeng, L. F. Fan, and M. Suhail Zubairy, “Deep-subwavelength lithography via graphene plasmons,” Phys. Rev. A 95, 053850 (2017). [CrossRef]  

24. Z. Fang, S. Thongrattanasiri, A. Schlather, Z. Liu, L. Ma, Y. Wang, P. M. Ajayan, P. Nordlander, N. J. Halas, and F. J. Garcia de Abajo, “Gated Tunability and Hybridization of Localized Plasmons in Nanostructured Graphene,” ACS. Nano 7(3), 2388–2395 (2013). [CrossRef]   [PubMed]  

25. F. H. L. Koppens, D. E. Chang, and F. J. Garcia de Abajo, “Graphene Plasmonics: A Platform for Strong LightâĂŞMatter Interactions,” Nano Lett. 11(8), 3370–3377 (2011). [CrossRef]   [PubMed]  

26. X. Zhou, X. Ling, H. Luo, and S. Wen, “Identifying graphene layers via spin Hall effect of light,” Appl. Phys. Lett. 101, 251602 (2012). [CrossRef]  

27. L. Cai, M. Liu, S. Chen, Y. Liu, W. Shu, H. Luo, and S. Wen, “Quantized photonic spin Hall effect in graphene,” Phys. Rev. A 95013809 (2017). [CrossRef]  

28. X. D. Zeng, Z. Y. Liao, M. Al-Amri, and M. S. Zubairy, “Controllable waveguide via dielectric cylinder covered with graphene: Tunable entanglement,” Europhys. Lett. 115, 14002 (2016). [CrossRef]  

29. T. Low and P. Avouris, “Graphene Plasmonics for Terahertz to Mid-Infrared Applications,” ACS. Nano 8(2), 1086–1101 (2014). [CrossRef]   [PubMed]  

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

31. A. N. Grigorenko, M. Polini, and K. S. Novoselov, “Graphene plasmonics,” Nat. Photon. 6, 749–758 (2012). [CrossRef]  

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

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

34. A. Y. Nikitin, F. Guinea, F. J. Garcia-Vidal, and L. Martin-Moreno, “Surface plasmon enhanced absorption and suppressed transmission in periodic arrays of graphene ribbons,” Phys. Rev. B 85081405 (2012). [CrossRef]  

35. W. Wang and J. M. Kinaret, “Plasmons in graphene nanoribbons: Interband transitions and nonlocal effects,” Phys. Rev. B 87, 195424 (2013). [CrossRef]  

36. S. Thongrattanasiri, A. Manjavacas, and F. J. Garcia de Abajo, “Quantum Finite-Size Effects in Graphene Plasmons,” ACS Nano 6(2), 1766–1775 (2012). [CrossRef]   [PubMed]  

37. M. Sarrazin, J. P. Vigneron, and J. M. Vigoureux, “Role of Wood anomalies in optical properties of thin metallic films with a bidimensional array of subwavelength holes,” Phys. Rev. B 67, 085415 (2003). [CrossRef]  

38. M. Merano, “Optical beam shifts in graphene and single-layer boron-nitride,” Opt. Lett. 41, 5780–5783 (2016). [CrossRef]   [PubMed]  

39. Z. B. Li, K. Yao, F. N. Xia, S. Shen, J. G. Tian, and Y. M. Liu, “Graphene Plasmonic Metasurfaces to Steer Infrared Light,” Sci. Rep. 5, 12423 (2015). [CrossRef]   [PubMed]  

40. G. W. Hanson, “Dyadic Green’s functions and guided surface waves for a surface conductivity model of graphene,” J. Appl. Phys. 103, 064302 (2008) [CrossRef]  

41. M. S. Tomaš, “Green function for multilayers: Light scattering in planar cavities,” Phys. Rev. A 51, 2545 (1995). [CrossRef]  

42. A. A. Maradudin, I. Simonsen, J. Polanco, and R. M. Fitzgerald, “Rayleigh and Wood anomalies in the diffraction of light from a perfectly conducting reflection grating,” J. Opt. 18024004 (2016). [CrossRef]  

43. A. K. Ghatak and K. Thyagarajan, Contemporary Optics (Plenum, New York, 1978). [CrossRef]  

44. M. Cheng, P. Fu, M. H. Weng, X Y. Chen, X. H. Zeng, S. Y. Feng, and R. Chen, “Spatial and angular shifts of terahertz wave for the graphene metamaterial structure,” J. Phys. D 48, 285105 (2014). [CrossRef]  

45. X. Li, P. Wang, F. Xing, X. D. Chen, Z. B. Liu, and J. G. Tian, “Experimental observation of a giant Goos-Hänchen shift in graphene using a beam splitter scanning method,” Opt. Lett. 39(19), 5574–5577 (2014). [CrossRef]   [PubMed]  

46. S. Chen, C. Mi, L. Cai, M. Liu, H. Luo, and S. Wen, “Observation of the Goos-Hänchen shift in graphene via weak measurements,” Appl. Phys. Lett. 110, 031105 (2017). [CrossRef]  

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 (6)

Fig. 1
Fig. 1 A Gaussian beam is incident on the periodic graphene ribbon array. The Fermi level of the ribbons can be tuned by the gate voltage.
Fig. 2
Fig. 2 The zeroth-order reflection spectrum for the ribbon array with different Fermi levels.
Fig. 3
Fig. 3 (a) The zeroth-order reflectivities of TM waves with φ = 0 at different Fermi levels. The labels (1) and (2) represent the plus first and minus second-order diffraction positions. The wavelength of the incident beam is 45.2μm and the other parameters are the same as Fig. 2. (b) The arguments ϕ of the zeroth-order reflection field. (c) and (d) The corresponding GH shifts.
Fig. 4
Fig. 4 (The GH shifts versus the Fermi level under different ε2 and τ. The parameters are the same as the Fig. 3(a) and θ = 0.0769.
Fig. 5
Fig. 5 (a) The zeroth-order reflectivity of TE wave along y direction asa function the incidence angle at different Fermi levels. (b) The corresponding GH shifts.
Fig. 6
Fig. 6 (a) The normalized field intensity distribution of the incident TM-polarized Gaussian beam on the z = 0 plane. (b, c) The normalized field distributions of the zeroth-order reflection fields for the cases that EF = 0.6eV and 0.5eV.

Equations (22)

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

σ ( ω ) = i e 2 E F π 2 ( ω + i τ 1 )
η x ( x , y ) σ ( x , y ) = E x 0 ( x , y , 0 ) + 1 2 ( 2 π ) 3 i ω d k x d k y d x d y 2 π i β k 0 2 ε 1 { [ k y 2 ρ 2 ( 1 + r s ) + 1 k 0 2 k x 2 β 2 ρ 2 ( 1 r p ) ] η x ( x , y ) + [ k x k y ρ 2 ( 1 + r s ) + 1 k 0 2 k x k y β 2 ρ 2 ( 1 r p ) ] η y ( x , y ) } e i [ k x ( x x ) + k y ( y y ) ] ;
η y ( x , y ) σ ( x , y ) = E y 0 ( x , y , 0 ) + 1 2 ( 2 π ) 3 i ω d k x d k y d x d y 2 π i β k 0 2 ε 1 { [ k x k y ρ 2 ( 1 + r s ) + 1 k 0 2 k x k y β 2 ρ 2 ( 1 r p ) ] η x ( x , y ) + [ k x 2 ρ 2 ( 1 + r s ) + 1 k 0 2 k y 2 β 2 ρ 2 ( 1 r p ) ] η y ( x , y ) } e i [ k x ( x x ) + k y ( y y ) ] ,
σ ( x , y ) = n σ n e i g n x ,
η x , y ( x , y ) = n η x , y n e i [ ( k x 0 + g n ) x + k y 0 y ] ,
η x n = 1 a 0 b d x E x 0 ( x , y , 0 ) σ e i ( k x x + k y y ) 1 2 ω n k 0 2 ε 1 β { [ k y 2 ρ 2 ( 1 + r s ) + 1 k 0 2 k x 2 β 2 ρ 2 ( 1 r p ) ] η x n σ n n + [ k x k y ρ 2 ( 1 + r s ) + 1 k 0 2 k x k y β 2 ρ 2 ( 1 r p ) ] η y n σ n n } ;
η y n = 1 a 0 b d x E y 0 ( x , y , 0 ) σ e i ( k x x + k y y ) 1 2 ω n k 0 2 ε 1 β { [ k x k y ρ 2 ( 1 + r s ) + 1 k 0 2 k x k y β 2 ρ 2 ( 1 r p ) ] η x n σ n n + [ k x 2 ρ 2 ( 1 + r s ) + 1 k 0 2 k y 2 β 2 ρ 2 ( 1 r p ) ] η y n σ n n } .
E x ( x , y , z ) = E x r ( x , y , z ) i 2 ω k 0 2 ε 1 β e i [ k x 0 x + k y 0 y + β z ] { [ k y 0 2 ρ 2 ( 1 + r s ) + 1 k 0 2 k x 0 2 β 2 ρ 2 ( 1 r p ) ] η x 0 + [ k x 0 k y 0 ρ 2 ( 1 + r s ) + 1 k 0 2 k x 0 k y 0 β 2 ρ 2 ( 1 r p ) ] η y 0 } = | r x | e i ϕ E x 0 ( x , y , z ) e 2 i β z ;
E y ( x , y , z ) = E y r ( x , y , z ) i 2 ω k 0 2 ε 1 β e i [ k x 0 x + k y 0 y + β z ] { [ k x 0 k y 0 ρ 2 ( 1 + r s ) + 1 k 0 2 k x 0 k y 0 β 2 ρ 2 ( 1 r p ) ] η x 0 + [ k x 0 2 ρ 2 ( 1 + r s ) + 1 k 0 2 k y 0 2 β 2 ρ 2 ( 1 r p ) ] η y 0 } = | r y | e i ϕ E y 0 ( x , y , z ) e 2 i β z .
S = λ 0 2 π d ϕ d θ ,
n π b ( 1 + ε 2 ) ω 2 4 α 0 c ω F .
ϕ = A r g ( η x 0 ) .
E ( r ) = A ( k x 0 , k y 0 ) e i k r d k x 0 d k 0 ,
A ( k x 0 , k y 0 ) = ( w x w y π ) 1 / 2 exp [ w x 2 2 ( k x 0 x 0 G ) 2 w y 2 2 k y 0 2 ] ,
( × × ε ( r , ω ) ω 2 c 2 ) E ( r , ω ) = i ω η ( r , ω ) .
E ( r , ω ) = i ω d 3 r G ( r , r ; ω ) η ( r , ω ) .
G x x ( r , r ; ω ) = 1 2 ( 2 π ) 3 d k x d k y 2 π i ε 1 β [ k y 2 ρ 2 ( 1 + r s ) + k x 2 β 2 k 0 2 ρ 2 ( 1 r p ) ] e i [ k x ( x x ) + k y ( y y ) ] ;
G x y ( r , r ; ω ) = 1 2 ( 2 π ) 3 d k x d k y 2 π i ε 1 β [ k x k y ρ 2 ( 1 + r s ) + k x k y β 2 k 0 2 ρ 2 ( 1 r p ) ] e i [ k x ( x x ) + k y ( y y ) ] ;
G y x ( r , r ; ω ) = G x y ( r , r ; ω ) ;
G y y ( r , r ; ω ) = 1 2 ( 2 π ) 3 d k x d k y 2 π i ε 1 β [ k x 2 ρ 2 ( 1 + r s ) + k y 2 β 2 k 0 2 ρ 2 ( 1 r p ) ] e i [ k x ( x x ) + k y ( y y ) ] .
r s = ( β ε 2 k 0 2 ρ 2 ) / ( β + ε 2 k 0 2 ρ 2 ) ;
r p = ( ε 2 β ε 1 ε 2 k 0 2 ρ 2 ) / ( ε 2 β + ε 1 ε 2 k 0 2 ρ 2 ) .
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.