Next Article in Journal
Lipid-Based Nano-Sized Cargos as a Promising Strategy in Bone Complications: A Review
Previous Article in Journal
Synthesis and Characterization of Biopolyol-Based Waterborne Polyurethane Modified through Complexation with Chitosan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Excitation of Surface Plasmon Polariton Modes with Double-Layer Gratings of Graphene

1
College of Physics Science and Engineering Technology, Yichun University, Yichun 336000, China
2
College of Literature, Journalism and Communication, Yichun University, Yichun 336000, China
3
Laboratory for Micro-Nano Physics and Technology of Hunan Province, School of Physics and Electronics, Hunan University, Changsha 410082, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(7), 1144; https://doi.org/10.3390/nano12071144
Submission received: 11 February 2022 / Revised: 27 March 2022 / Accepted: 27 March 2022 / Published: 30 March 2022
(This article belongs to the Topic Application of Graphene-Based Materials)

Abstract

:
A long-range surface plasmon polariton (SPP) waveguide, composed of double-layer graphene, can be pivotal in transferring and handling mid-infrared electromagnetic waves. However, one of the key challenges for this type of waveguide is how to excite the SPP modes through an incident light beam. In this study, our proposed design of a novel grating, consisting of a graphene-based cylindrical long-range SPP waveguide array, successfully addresses this issue using finite-difference time-domain simulations. The results show that two types of symmetric coupling modes (SCMs) are excited through a normal incident light. The transmission characteristics of the two SCMs can be manipulated by changing the interaction of the double-layer gratings of graphene as well as by varying various parameters of the device. Similarly, four SCMs can be excited and controlled by an oblique incident light because this light source is equivalent to two orthogonal beams of light. Furthermore, this grating can be utilized in the fabrication of mid-infrared optical devices, such as filters and refractive index sensors. This grating, with double-layer graphene arrays, has the potential to excite and manipulate the mid-infrared electromagnetic waves in future photonic integrated circuits.

1. Introduction

Surface plasmon polaritons (SPPs) [1], which are the collective oscillations of electrons at the interface between a conductor and a dielectric, are promising candidates that can be used to overcome the diffraction limit of light owing to their strong sub-wavelength confinement of electromagnetic (EM) waves. Noble metals are considered the most promising SPP materials [2]. Therefore, various noble metal-based optical devices [3,4,5,6,7] have been studied in the visible range of light for future photonic integrated circuits. However, in the mid-infrared range, noble metal-based optical devices exhibit significantly weak EM field confinement and strong transmission losses, making them inefficient in this wavelength range. Therefore, a new SPP material is needed that can be used in mid-infrared range applications, such as spectroscopy, communication, sensing, and homeland security [8].
Graphene [9], a typical two-dimensional material formed by a single layer of carbon atoms packed into a honeycomb lattice, has several distinct advantages over the noble metals in the mid-infrared range [10], such as the deep sub-wavelength confinement of EM waves, lower transmission losses, and graphene-Fermi energy-dependent EM properties [11]. Because of these unique EM properties, various graphene-based optical devices [12,13,14,15] have been proposed in recent years. However, since the wave vector of a graphene SPP wave is considerably larger than that of an incident light beam, exciting an SPP wave in graphene by matching these two wave vectors is significantly challenging. Several methods, such as tailoring graphene into various geometric-patterned nanostructures [16,17,18], near-field probe coupling [19,20], prism coupling [21,22,23], dielectric grating coupling [24,25,26], metal grating coupling [27], and graphene grating coupling [24,28,29,30,31], have been proposed to overcome this mismatch to explore the advantages of graphene further. Among them, graphene grating, consisting of a graphene-based SPP waveguide array placed on a substrate, effectively compensates for this mismatch using the wave vector of the incident light by diffraction. Currently, most graphene gratings are composed of the most common type of SPP waveguides with a single layer of graphene. Obviously, for a newly proposed graphene SPP waveguide with potentially enhanced performance, a novel graphene grating has to be designed to excite the corresponding SPP mode.
Recently, we investigated an intriguing type of graphene-based long-range SPP (LRSPP) waveguide [32,33], that is composed of one or more dielectric layers placed between two layers of graphene. The symmetric coupling modes (SCMs) and the anti-symmetric coupling modes (ASCMs) are transferred through this waveguide when the two graphene layers are effectively coupled. The SCMs exhibit an ultra-strong EM field confinement and short propagation length, whereas the ASCMs exhibit completely opposite characteristics. Thus, the SCMs and the ASCMs can also be named as short-range SPP (SRSPP) and LRSPP modes, respectively. Because of these interesting EM features, this type of waveguide can have a wide range of applications in future photonic integrated circuits [13,15]. However, in the above-mentioned studies [32,33], methods of exciting the two types of modes have not been introduced. In the present study, by using the finite-difference time-domain (FDTD) method, we propose a grating composed of the graphene-based cylindrical LRSPP waveguide array. The results demonstrate that two and four types of SCMs are excited through a normal and an oblique incident light by this grating, respectively. In addition, the transmission characteristics of the SCMs are effectively controlled by changing the interaction of the double-layer grating of graphene as well as by varying the other parameters of the device. Furthermore, we can design some mid-infrared optical devices, such as the filters and the refractive index sensors, using the grating.

2. Materials and Methods

A schematic diagram of the proposed grating, which is comprised of a graphene LRSPP waveguide array placed on a substrate, is shown in Figure 1a. This waveguide is composed of a cylindrical silicon nanowire core surrounded by an inner graphene layer, a silica layer, and an outer graphene layer as shown in Figure 1b. The radius of the nanowire and the thickness of the silica layer are labeled as RSi and tSiO2, respectively. To fabricate such a grating, one of the key challenges is how to obtain a feasible manufacturing technique for the cylindrical LRSPP waveguide. Li et al. [34] have reported how to transfer a graphene flake onto a microfiber in detail. Several experiments [35,36] have shown that the dielectric nanowire can be tightly surrounded by the graphene layer because of the van der Waals force. Thus, this waveguide can be formed by rolling a graphene flake around the dielectric nanowire from the inner to the outer layer, step by step. The silica layer can be formed using plasma-enhanced chemical vapor deposition (PECVD) technology and the thickness of the silica layer can be controlled by tuning the deposition conditions [37]. For simplicity, we have not considered the dispersive characteristics of the medium in this paper. The refractive indexes of the silicon nanowire and the silica layer are set as n1 = nSi = 3.455 and n2 = nSiO2 = 1.445, respectively. We also assume that this device is embedded in a uniform silica environment with a refractive index of n3 = n4 = nSiO2 = 1.445. The period of the waveguide array is labeled as p. The permittivity of graphene is expressed as a complex permittivity with a surface-normal effective permittivity of 2.5 and an in-plane effective permittivity of 2.5 + g/(ε0ωd) [38,39], where σg is the conductivity of graphene, ε0 is the permittivity of free space, ω is the angular frequency of the incident light, and d is the thickness of a graphene layer. Using the Drude-like formula [38,40], σg is expressed as σg(ω) = ie2Ef/[πħ2(ω + iτ−1)], where e is the electron charge, Ef is the Fermi energy, and τ is the electron relaxation time. The latter is calculated from the relation τ = μEf/(evF2), where μ is the electron mobility, and vF = 10−6 ms−1 is the Fermi velocity. In this paper, unless stated otherwise, the following values of different parameters are used: μ = 1.0 m2/(Vs), Ef = 0.8 eV, T = 300 K, and d = 1 nm.
In this study, we employed the commercial software of the Lumerical FDTD Solutions to investigate the features of the proposed grating. The two-dimensional simulation was performed for a single unit of the grating in xy-plane. The simulation region was truncated by periodic boundaries in the x-direction and perfectly matched layer boundaries in the y-direction. The mesh sizes were set as 0.1 nm along the x, y axes inside the grating unit, and gradually increased outside the grating unit.

3. Transmission Characteristics of This Grating with a Perpendicular Incident Light

First, a normal incident plane wave (θ = 0°) was projected onto the graphene LRSPP waveguide array along the negative y-direction. θ represents the angle between the direction of β and the negative y-direction, in which β is the propagation constant of the incident light. We investigated the physical mechanism of exciting the SPP modes through the grating. It is widely known that SPP oscillations are excited in a grating consisting of a graphene waveguide array [24,28], when the electric field polarization is perpendicular to the grating. Thus, the SPP modes can also be excited by transverse electric (TE) and transverse magnetic (TM) waves. In this study, when the SPP oscillations are excited by a plane wave, a sharp notch corresponding to the resonant frequency is observed in the transmission spectrum. This resonant behavior has been demonstrated experimentally in a graphene ribbon grating [41]. A similar physical phenomenon is also observed in our proposed grating using numerical simulation. To excite such an SPP wave in graphene with a free-space optical wave, their large difference of wave-vector has to be overcome. Using a grating is an effective way to compensate for the wave-vector mismatch. The wave-vector matching equation of this grating can be expressed as
k 0 sin θ + m k g r a t i n g = k s p p
where k0 = 2π/λ0 is the k-vector of the incident light, kgrating is the compensatory k-vector by the grating, kspp is the k-vector of the excited SPP mode, and m is the diffraction order. Due to kspp = nrk0, the Equation (1) can be expressed approximatively as
2 π λ 0 sin θ + m 2 π p = n r 2 π λ 0 ,
where nr is the real part of the effective refractive index of the excited SPP mode. However, as the graphene LRSPP waveguide has two graphene layers, our proposed grating is equivalent to two optical gratings. The first grating consists of the inner layer graphene array and the silicon nanowire array, whereas the second grating consists of the outer layer graphene array, the silica layer array, and the silicon nanowire array. For simplicity, the first and the second gratings are referred to as the inner and the outer gratings, respectively. Either of them can individually excite SPP resonant modes. However, according to Equation (2), their corresponding resonant wavelength is different for the same incident light, because the SPP mode excited by two gratings has a different nr value.
Figure 2a shows that two main SPP resonant modes are excited by a normal incident plane EM wave with an electric field polarized along the x-direction. Both modes are 1-order SCMs because their Ex and Ex phase distributions are symmetric with respect to the y-axis (Figure 2d,e,i,j). The SCMs at the shorter and longer wavelengths are referred to as the Sx1 mode and Sx′1 modes, respectively. Both the Sx1 and Sx′1 modes are neither TE modes nor TM modes, according to the work described in [33]. In addition, Figure 2a also shows that the Sx′1 mode corresponds to stronger transmission and weaker reflection than the Sx1 mode. However, the two modes have nearly identical absorptions.
The EM field energy of the Sx′1 mode is effectively confined between the two graphene layers (Figure 2h,l), whereas that of the Sx1 mode is not strictly confined between the two graphene layers and spreads outside the waveguide, gradually decaying with increasing distance from the waveguide (Figure 2c,g). As evident from Figure 2c,g,h,l, the Sx1 and Sx′1 modes are excited by the outer and the inner gratings, respectively, which results in the contrasting EM field energies of the two modes. In addition, there is an interaction between the two gratings. For example, when the Sx′1 mode is excited, the EM oscillations of the quadrupoles occur in the inner grating, which further induces quadrupole oscillations with an opposite charge distribution in the outer grating (Figure 2k). This interaction causes a shift in the resonant wavelength of the Sx′1 mode. A similar EM oscillations of the quadrupoles occur in the outer grating (Figure 2f), and a similar shift in the resonant wavelength of the Sx1 mode is caused by the interaction between the two gratings. Therefore, the resonant wavelengths of the two modes are determined by the individual characteristics of the gratings as well as by the interaction between the two gratings.
Based on the above principle of double graphene gratings, we investigated the dependence of the resonant wavelength on various geometrical parameters of this waveguide. For convenience, we defined the equivalent radii of the inner and outer gratings, Ri, and Ro, as Ri = RSi + d and Ro = RSi + tSiO2 + 2d, respectively. Therefore, Ro = R, where R is the total radius of the graphene LRSPP waveguide. We further defined the occupation rate of the inner and outer gratings, ηi and ηo, as ηi = 2Ri/p and ηo = 2Ro/p, respectively. The parameters of this device can be effectively estimated by combining Equation (2) and the result described in [33]. Then, we assumed that RSi = 46 nm and tSiO2 = 12 nm, and increased p from 180 nm to 360 nm. In this case, Ri and Ro are constants, whereas ηi and ηo decrease. According to previous reports [24,28,30], a change in the occupation rate of a grating significantly changes the coupling strength between the components of the grating and subsequently causes a considerable shift in the resonant wavelength. Although establishing a mathematical relation between the resonant wavelength and the occupation rate is difficult, several studies [24,28,30] qualitatively demonstrated that the resonant wavelength decreases (blue shift) with decreasing occupation rate. Our study verifies this feature. One can find that the resonant wavelengths of the Sx1 and Sx′1 modes decrease with increasing p because ηi and ηo decrease in this case (Figure 2a).
Next, we fixed tSiO2 as 12 nm and increased the RSi and p values, while simultaneously ensuring that ηo was fixed at 0.5 (i.e., p = 4R). However, ηi increases with increasing p because ηi = 2Ri/p = 0.5 − 2(tSiO2 + d)/p. Thus, the resonant wavelength of the Sx′1 mode increases (red shift). Further, Ri and Ro increase with increasing RSi. Therefore, a longer wavelength is required to wrap the waveguide when same-order modes are excited. We have derived an equation to calculate the resonant frequency for the grating formed via cylindrical graphene-coated nanowire array (GCNA) [28]. According to the equation, the resonant frequency, ωp, is inversely proportional to the square root of the radius of the cylindrical graphene waveguide, r, (i.e.,   ω p 1 / r ). This subsequently implies that the resonant wavelength, λ, is directly proportional to the square root of r (i.e., λ r ). Our proposed double layers of gratings are similar to two GCNAs. Figure 2b shows that the resonant wavelengths of the Sx1 and Sx′1 modes increase with increasing R (red shift). Moreover, the red shift magnitudes of the Sx′1 mode are significantly larger than those of the S1 modes, because the red shift in the Sx′1 mode originates not only from increasing R but also from increasing ηi.
The interaction between the two gratings can be adjusted by varying the distance between them. We fixed RSi as 46 nm and increased the values of tSiO2 and p simultaneously to ensure that ηo was fixed at 0.5. For the Sx′1 mode, Ri = 47 nm is a constant, and ηi decreases with increasing p. The resonant wavelength of the Sx′1 mode decreases with decreasing ηi. However, for the Sx1 mode, since Ro increases with increasing tSiO2, the resonant wavelength increases. Thus, the resonant wavelengths of the two modes exhibit opposite trends with increasing tSiO2 (Figure 3a).
To better understand the dependence of the resonant wavelength on the interaction between the two gratings, we investigated two other types of graphene gratings without interaction for comparison. The first type of grating is formed by graphene hybrid waveguide [42] array (GHWA) (Figure 3c), and the second type of grating consists of the GCNA (Figure 3d). We refer to the two type of gratings as GHWA and GCNA gratings, which are similar to the outer and inner gratings of Figure 1, respectively. The Mx1 and Mx′1 modes are excited by the GHWA and GCNA gratings (Figure 3e,g), respectively. Both modes are 1-order SCMs, which originate from the quadrupole oscillations in graphene (Figure 3f,h). The occupation rates of the GHWA and GCNA gratings, η1 and η2, were set as η1 = ηo = 0.5 and η2 = ηi =0.5 − 2(tSiO2 + d)/p, respectively. Further, we investigated the impact of the interaction between the inner and outer grating on the resonant wavelengths of the Sx1 and Sx′1 modes. For relatively smaller values of tSiO2, there are strong interactions between the inner and outer gratings, which cause a more obvious shift of the resonant wavelengths of the Sx1 and Sx′1 modes from the Mx1 and Mx′1 modes, respectively. In addition, the resonant wavelength of the Sx′1 mode is significantly larger than that of the Mx′1 mode because the Sx′1 mode excited by the inner grating is strongly confined within the outer grating. On the contrary, the Sx1 mode excited by the outer grating is subjected to weaker confinement by the inner grating because such a mode can freely transmit outside the waveguide. This results in approximately similar resonant wavelengths of the Sx1 and the Mx1 modes. Therefore, we conclude that for small tSiO2, the interaction between the two gratings has a greater impact on the resonant wavelength of the Sx′1 mode than that of the Sx1 mode. With increasing tSiO2, the interaction of the two gratings decreases, the differences between the resonant wavelengths of the S1 and M1 modes and between that of the Sx′1 and Mx′1 modes also gradually decreases, and finally converges on a stable value (Figure 3b).
The most intriguing feature of the SPP wave excited by graphene is its tunability. The resonant behaviors of this grating can be adjusted by varying the graphene conductivity, which is primarily determined by the Fermi energy and electron mobility of graphene. The Fermi energy of graphene depends on the carrier concentration, which can be varied by controlling its gate voltage or chemical doping [10]. Efetov et al. [43] reported a significantly high carrier concentration of 4 × 1014 cm−2 achieved experimentally by using a field-effect transistor (FET) type structure. This carrier concentration value is equivalent to a Fermi energy (Ef) of 1.17 eV. Therefore, we shift Ef from 0.4 eV to 1.0 eV. Previous studies have shown that for a grating formed by the graphene nanoribbon array [24,30] or via GCNA [28], the resonant wavelength is related to Ef via the relation λ (Ef)−1/2. As our proposed double graphene gratings are similar to two GCNA gratings, the resonant wavelengths of the Sx1 and Sx′1 modes show a similar trend (blue shift) with increasing Ef (Figure 4a).
Next, we investigated the resonant feature by varying the electron mobility of graphene. An electron mobility of μ > 10 m2/(Vs) with a peak value (μmax) of 23 m2/(Vs) has been experimentally achieved in suspended exfoliated graphene [44]. Chen et al. reported an experimentally achieved electron mobility as high as 4 m2/(Vs) in SiO2-supported monolayer graphene [20]. However, since graphene with lower electron mobility is more practical, we selected μ values in the range from 0.08 m2/(Vs) to 1 m2/(Vs). Figure 4b shows that the resonant wavelengths of the two modes are approximately identical for different μ values because the resonant behavior is mainly characterized by the imaginary part of conductivity, which is mainly related to the Fermi energy. The optical loss originates from the thermal loss of the current transferring in graphene. Due to the conductivity of graphene σg = ie2Ef/[πħ2(ω + iτ−1)], σg can be expressed as σg = σgr + gi, where σgr and σgi are the real and imaginary parts of the graphene conductivity, respectively. According to Maxwell’s equations, we can obtain
× H = i ( ω ε σ g   i ) E + σ g   r E = J
where J is the current density of graphene. The power density of optical loss, PJ, can be derived as
P J = 1 2 Re ( J * · E ) = 1 2 Re [ i ( ω ε σ g i ) E * · E + σ g r E * · E ] = 1 2 σ g r E 0 2 ,
where E0 is the amplitude of the electric field. Therefore, the optical loss in graphene is determined by the real part of the graphene conductivity, which is related to both the Fermi energy and the electron mobility. On the other hand, we can obtain the optical loss according to the equation L[dB/μm] = 8.86 nik0 [45], where ni is the imaginary part of the mode effective index. The ni values can be obtained by the mode solver of the “mode source” from the software Lumerical FDTD Solutions. For example, we can obtain ni = 0.5863 for the Sx1 mode of the cylindrical LRSPP waveguide with μ = 1.0 m2/(Vs), Ef = 0.8 eV, tSiO2 = 12 nm, RSi = 46 nm, and λ0 = 10 μm. In this case, the optical loss of this mode is 3.2638 dB/μm. With decreasing Ef and μ values, the optical loss in graphene increases, and the notch depth of the transmission spectra decreases. To describe the depth of the notch, we define the extinction ratio (ER) of the transmission spectra as ER = −10 × log(T) (dB), where T is the transmission. As can be seen in Figure 4c,d, the ER values in both modes increase significantly with increasing Ef and μ values. On the other hand, the ER value in the Sx1 mode is considerably larger than that in the Sx′1 mode.
On the other hand, we can also adjust the interaction of the two gratings by varying the graphene characteristics. It can be observed that with decreasing Ef, the resonant wavelengths of the Sx1 and Sx′1 modes exhibit a more obvious shift relative to the Mx1 and Mx′1 modes (Figure 5a). This can be explained as follows. The effective refractive indexes of the Mx1 and Mx′1 modes increase with decreasing Ef (Figure 5a). This improves the confinement of the EM field, and enhances the EM field distributing on the surface of the graphene waveguide. Then, the interaction is strengthened between the two gratings. However, the effective refractive indexes of the Mx1 and Mx′1 modes are almost unchanged with the varying electric mobility of graphene (Figure 5b). This means that the interaction of two gratings will be not changed with the electric mobility. Furthermore, the resonant wavelengths of the Sx1 and Sx′1 modes maintain a stable shift relative to the Mx1 and Mx′1 modes (Figure 5b).
To further improve the ER values of the graphene LRSPP waveguide array, a more realistic method is decreasing the refractive index of the dielectric nanowire instead of increasing the Ef and μ values. Usually, there is an insignificant loss in dielectric nanowires. However, the EM field confinement is further weakened by decreasing the refractive index of the dielectric nanowire. Subsequently, the interaction between the graphene layer and the EM field is abated. Therefore, the waveguide losses decrease. Figure 6a shows that the ER values of the two modes increase approximately 1.7 times when the n1 values decrease from 3.455 to 1.445. In addition, we showed in our previous study [28] that the resonant wavelength of the GCNA grating is directly proportional to the square root εAVG [i.e., λ (εAVG)1/2], where εAVG = (εin + εout)/2 is the average dielectric constant of the graphene layers, and εin and εout are the dielectric constant values of the inner and outer layers of graphene, respectively. We observed a similar physical mechanism and phenomena in this study. The resonant wavelengths of the two modes exhibit a red shift with increasing n1 (Figure 6b). Based on the same principle, the red shift of the resonant wavelengths is also observed with increasing n3 (Figure 6c). Evidently, the resonant wavelengths of the two modes are in approximately linear relation with n3. By fitting the simulated data, the two-line equations can be expressed by the following relation for the Sx1 and Sx′1 modes:
λSx1 = 2407.9 × n3 + 8265 (nm),
λSx′1 = 837.1× n3 + 20,833.7 (nm)
We can design a sensor to measure the refractive index of the surroundings by using these equations. We usually evaluate the sensing performance of a sensor based on two assessment factors. One is the sensitivity of the wavelength (S), which is defined as S = Δλ/Δn, where Δλ/Δn is expressed in nanometers per refractive index unit (nm/RIU). The S values of the Sx1 and Sx′1 modes are 2407.9 nm/RIU and 837.1 nm/RIU at n1 = 3.445, respectively. The other assessment factor is the figure of merit (FOM) [46,47], which is defined as the ratio of S to the full width at half maximum (FWHM) of the transmission peak (i.e., FOM = S/FWHM). The damping of the quadrupole oscillations increases with increasing n3, which results in an increase in the corresponding FWHM. Therefore, the FOM values of the two modes decrease with increasing n3 (Figure 6d). Nevertheless, compared to other graphene-based refractive index sensors [48,49], this sensor exhibits an excellent sensing performance with high S and FOM values. Most notably, this sensor has an obvious advantage in measuring various gas surroundings with low refractive index values, because the Sx1 mode exhibits higher FOM values than the other sensor [50] in this case.
The substrate also has an impact on the characteristics of the grating. From Figure 7a, one can find that the resonant wavelengths of the excited modes increase with the refractive index of the substrate n2. This can be explained as follows. A strong EM field gathers in the interface between the waveguide and the substrate (Figure 7b). This results in the nr value of the excited mode increasing. According to Equation (2), the resonant wavelength λ0 can be expressed as λ0 = nrP/m for a normal incident light (θ = 0°). Thus, the values of nr and λ0 increase with n2. On the other hand, exciting the high order SPP mode becomes easier in the case of increasing n2. The Sx′2 mode is effectively excited as n2 = 3.455 (Figure 7a,c).
Finally, we investigated the excitation approach of the high-order modes. Since the wave vectors of the high-order modes are smaller than that of the low order modes, lower compensation by a grating is required to achieve wave vector matching for higher-order modes. Therefore, a grating with a lower grating constant easily excites higher-order modes. Considering that the ER value of the high-order mode is improved with increasing Ef and μ and decreasing n1, we set the parameters of the grating as R = 60 nm, p = 150 nm, Ef = 1.0 eV, μ = 1.0 m2/(Vs), and n1 = 1.72. Consequently, we observed a transmission spectrum with two low-order modes (Sx1 and Sx′1 modes) and five high-order modes (Sx2, Sx′3, Sx3, Sx′4, and Sx′5 modes) (Figure 8a). The |E| and |H| distributions of the five high-order modes are shown in Figure 8b,c,d,e,f and in Figure 8g,h,i,j,k, respectively. Some high-order modes are missing in this transmission spectra for two reasons. First, some high-order modes, such as the Sx4 and Sx5 modes, cannot be excited using this set of parameters of the grating. Second, although some high-order modes, such as the Sx′2 mode, are excited, their corresponding transmission peaks cannot be observed because their resonant wavelengths are overlapped by the Sx1 mode region.

4. Transmission Characteristics of This Grating with an Oblique Incident Light

Finally, we investigate the transmission characteristics of the grating through an oblique incident light of θ ≠ 0° (Figure 1b). Usually, two types of modes are used to describe the SPP resonant for an oblique incident light, including the spectral mode (at fixed angle) and the angular mode (at fixed wavelength) [50,51]. According to Equation (2), the resonant wavelength, λres, can be expressed as
λ   r e s = ( n r sin θ ) p m
Since the nr values are much larger than one for the graphene SPP modes, the resonant wavelengths are almost unrelated to θ. Figure 9a shows that four SPP modes are excited, and their λres values are almost unchanged with varying θ. However, the notch depth of the transmission spectra is sensitive to θ. This can be explained as follows. The electric field of an oblique incident light can be divided into the two polarization components of Ex and Ey. This results in two types of SPP oscillations along the x-axis and the y-axis, respectively. Therefore, four 1-order SCMs can be excited through two graphene gratings, as shown in Figure 9a. The Sx1, Sx′1, and Sy1, Sy′1 modes originate from the SPP oscillations along the x-axis and the y-axis, respectively. The |E| and |H| distributions of the Sy1, Sy′1 modes are shown in Figure 9c–f, respectively. Upon increasing the incident angular θ, the Ex component decreases, while the Ey component increases. The SPP oscillations will be weakened and enhanced along the x-axis and the y-axis, respectively. Thus, the notch depths of the transmission spectra of the Sx1 and Sx′1 modes decrease, while those of the Sy1 and Sy′1 modes increase, as shown in Figure 9a. Furthermore, the transmission values of the angular mode are almost unchanged with θ, when λ0 is greatly different from λres, such as λ0 = 9.0 μm. However, the transmission values of the angular mode exhibit a significant change with θ when λ0 is close to λres, such as λ0 = 8.5 μm and λ0 = 9.5 μm (Figure 9b). On the other hand, the resonance angle [50,51] can’t be observed in Figure 9b; this can also be attributed to the fact that the resonant wavelengths are almost unrelated to θ.

5. Conclusions

In summary, we have proposed and investigated a novel design for a grating, which is composed of a graphene-based cylindrical LRSPP waveguide array. The numerical simulation results demonstrate that two types of SCMs can be excited by a normal incident light passing through this grating, because such a device with double-layer graphene arrays can be equated to two interacting graphene gratings. Therefore, the excited SCMs can be effectively controlled not only by varying the geometrical and physical parameters of the device, but also by changing the interaction between the two graphene gratings. Furthermore, four types of SCMs can be excited by an oblique incident light, which can be divided into two orthogonal beams of light. Based on the transmission features of the SCMs through this grating, several active optical devices, such as filters and refractive index sensors, can be designed. This grating device with double-layer graphene has the potential to excite and manipulate the mid-infrared waves in future photonic integrated circuits.

Author Contributions

Conceptualization, J.L.; methodology, J.L. and W.W.; validation, X.Z. (Xia Zhou), and Y.Y.; formal analysis, F.X. and X.Z. (Xiaoming Zhang); software, L.W.; data curation, F.X. and X.Z. (Xia Zhou); writing—original draft preparation, J.L.; writing—review and editing, J.L. and X.Z. (Xiaoming Zhang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 62065018, 61741515, and 62165016.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barnes, W.L.; Dereux, A.; Ebbesen, T.W. Surface plasmon subwavelength optics. Nature 2003, 424, 824–830. [Google Scholar] [CrossRef] [PubMed]
  2. West, P.R.; Ishii, S.; Naik, G.V.; Emani, N.K.; Shalaev, V.M.; Boltasseva, A. Searching for better plasmonic materials. Laser Photonics Rev. 2010, 4, 795–808. [Google Scholar] [CrossRef] [Green Version]
  3. Li, R.; Kong, X.K.; Liu, S.B.; Liu, Z.M.; Li, Y.M. Planar metamaterial analogue of electromagnetically induced transparency for a miniature refractive index sensor. Phys. Lett. A 2019, 383, 125947. [Google Scholar] [CrossRef]
  4. Burokur, S.N.; Lupu, A.; Lustrac, A. Direct dark mode excitation by symmetry matching of a single-particle-based metasurface. Phys. Rev. B 2015, 91, 035104. [Google Scholar] [CrossRef] [Green Version]
  5. Kazanskiy, N.L.; Khonina, S.N.; Butt, M.A.; Ka′zmierczak, A.; Piramidowicz, R. A numerical investigation of a plasmonic sensor based on a metal-insulator-metal waveguide for simultaneous detection of biological analytes and ambient temperature. Nanomaterials 2021, 11, 2551. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, J.P.; Wang, L.L.; Sun, B.; Li, H.J.; Zhai, X. Enhanced optical transmission through anano-slit based on a dipole source and an annular nano-cavity. Opt. Laser Technol. 2015, 69, 71–76. [Google Scholar] [CrossRef]
  7. Abdulhalim, I. Coupling configurations between extended surface electromagnetic waves and localized surface plasmons for ultrahigh field enhancement. Nanophotonics 2018, 7, 1891–1916. [Google Scholar] [CrossRef]
  8. Soref, R. Mid-infrared photonics in silicon and germanium. Nat. Photonics 2010, 4, 495–497. [Google Scholar] [CrossRef]
  9. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.V.; Dubonos, S.V.; Firsov, A.A. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438, 197–200. [Google Scholar] [CrossRef]
  10. Jablan, M.; Buljan, H.; Soljacic, M. Plasmonics in graphene at infrared frequencies. Phys. Rev. B 2009, 80, 245435. [Google Scholar] [CrossRef] [Green Version]
  11. Vakil, A.; Engheta, N. Transformation optics using graphene. Science 2011, 332, 1291–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Biswas, S.R.; Gutiérrez, C.E.; Nemilentsau, A.; Lee, I.H.; Oh, S.H.; Avouris, P.; Low, T. Tunable graphene metasurface reflect array for cloaking, illusion, and focusing. Phys. Rev. Appl. 2018, 9, 034021. [Google Scholar] [CrossRef] [Green Version]
  13. Liu, J.P.; Wang, W.L.; Xie, F.; Luo, X.; Zhou, X.; Lei, M.; Yuan, Y.J.; Long, M.Q.; Wang, L.L. Efficient directional coupling from multilayer-graphene-based long-range SPP waveguide to metal-based hybrid SPP waveguide in mid-infrared range. Opt. Express 2018, 26, 29509–29521. [Google Scholar] [CrossRef]
  14. Jonas, A.R.S.; Lima, R.F. Perfect valley filter controlled by Fermi velocity modulation in graphene. Carbon 2020, 160, 353–360. [Google Scholar]
  15. Huang, C.C.; Chang, R.J.; Cheng, C.W. Ultra-low-loss mid-infrared plasmonic waveguides based on multilayer graphene. Nanomaterials 2021, 11, 2981. [Google Scholar] [CrossRef] [PubMed]
  16. Fang, Z.; Thongrattanasiri, S.; Schlather, A.; Liu, Z.; Ma, L.; Wang, Y.; Ajayan, P.M.; Nordlander, P.; Halas, N.J.; Garcia de Abajo, F.J. Gated tunability and hybridization of localized plasmons in nanostructured graphene. ACS Nano 2013, 7, 2388–2395. [Google Scholar] [CrossRef]
  17. Fan, Y.; Shen, N.H.; Koschny, T.; Soukoulis, C.M. Tunable terahertz meta-surface with graphene cut wires. ACS Photonics 2015, 2, 151–156. [Google Scholar] [CrossRef]
  18. Hossain, M.S.; Huynh, D.H.; Jiang, L.; Rahman, S.; Nguyen, P.D.; Al-Dirini, F.M.A.; Hossain, F.M.; Bahk, J.; Skafidas, S. Investigating enhanced thermoelectric performance of graphene-based nano-structures. Nanoscale 2018, 10, 4786–4792. [Google Scholar] [CrossRef]
  19. Chen, J.; Badioli, M.; Alonso-Gonzalez, P.; Thongrattanasiri, S.; Huth, F.; Osmond, J.; Spasenovic, M.; Centeno, A.; Pesquera, A.; Godignon, P.; et al. Optical nano-imaging of gate-tunable graphene plasmons. Nature 2012, 487, 77–81. [Google Scholar] [CrossRef] [Green Version]
  20. Fei, Z.; Rodin, A.S.; Andreev, G.O.; Bao, W.; McLeod, A.S.; Wagner, M.; Zhang, L.M.; Zhao, Z.; Thiemens, M.; Dominguez, G.; et al. Gate-tuning of graphene plasmons revealed by infrared nano-imaging. Nature 2012, 487, 82–85. [Google Scholar] [CrossRef]
  21. Nikitin, A.Y.; Alonso-González, P.; Hillenbrand, R. Efficient coupling of light to graphene plasmons by compressing surface polaritons with tapered bulk materials. Nano Lett. 2014, 14, 2896–2901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Lee, S.; Kim, S. Practical Perfect Absorption in Monolayer Graphene by Prism Coupling. IEEE Photonics J. 2017, 9, 2700810. [Google Scholar] [CrossRef]
  23. Heo, H.; Lee, S.; Kim, S. Broadband absorption enhancement of monolayer graphene by prism coupling in the visible range. Carbon 2019, 154, 42–47. [Google Scholar] [CrossRef]
  24. Gao, W.L.; Shu, J.; Qiu, C.Y.; Xu, Q.F. Excitation of plasmonics waves in graphene by guided-mode resonances. ACS Nano 2012, 6, 7806–7813. [Google Scholar] [CrossRef]
  25. Lu, H.; Zeng, C.; Zhang, Q.M.; Liu, X.M.; Hossain, M.M.; Reineck, P.; Gu, M. Graphene-based active slow surface plasmon polaritons. Sci. Rep. 2015, 5, 8443. [Google Scholar] [CrossRef] [Green Version]
  26. Janfazaa, M.; Mansouri-Birjandia, M.A.; Tavousi, A. Proposal for a graphene nanoribbon assisted mid-infrared band-stop/ band-pass filter based on bragg gratings. Opt. Commun. 2019, 440, 75–82. [Google Scholar] [CrossRef]
  27. Ren, Y.X.; Guo, X.G.; Zhang, G.X.; Balakin, A.V.; Shkurinov, A.P.; Yu, A.Q.; Zhu, Y.M. Excitation of graphene surface plasmons polaritons by guided-mode resonances with high efficiency. Opt. Express 2020, 28, 13224–13233. [Google Scholar] [CrossRef]
  28. Xia, X.S.; Zhai, X.; Wang, L.L.; Liu, J.P.; Li, H.J.; Liu, J.Q.; Pan, A.L.; Wen, S.C. Excitation of surface plasmons in graphene-coated nanowire arrays. J. Appl. Phys. 2016, 120, 103204. [Google Scholar] [CrossRef]
  29. Ding, P.; Li, Y.; Li, M.Y.; Shao, L.; Wang, J.Q.; Fan, C.Z.; Zeng, F.G. Tunable dual band light trapping and localization in coupled graphene grating-sheet system at mid-infrared wavelengths. Opt. Commun. 2018, 407, 410–416. [Google Scholar] [CrossRef]
  30. Xia, X.S.; Zhai, X.; Wang, L.L.; Li, H.J.; Huang, Z.R.; Lin, Q. Dynamically tuning the optical coupling of surface plasmons in coplanar graphene nanoribbons. Opt. Commun. 2015, 352, 110–115. [Google Scholar] [CrossRef]
  31. Ukhtary, M.S.; Saito, R. Surface plasmon in graphene and carbon nanotubes. Carbon 2020, 167, 455–474. [Google Scholar] [CrossRef]
  32. Liu, J.P.; Zhai, X.; Wang, L.L.; Li, H.J.; Xie, F.; Xia, S.X.; Shang, X.J.; Luo, X. Graphene-based long-range SPP hybrid waveguide with ultra-long propagation length in mid-infrared range. Opt. Express 2016, 24, 5376–5386. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, J.P.; Zhai, X.; Xie, F.; Wang, L.L.; Xia, S.X.; Li, H.J.; Shang, X.J. Analytical model of mid-infrared surface plasmon modes in a cylindrical long-range waveguide with double-layer graphene. J. Lightwave Technol. 2017, 35, 1971–1979. [Google Scholar] [CrossRef]
  34. Li, W.; Chen, B.; Meng, C.; Fang, W.; Xiao, Y.; Li, X.; Hu, Z.; Xu, Y.; Tong, L.; Wang, H.; et al. Ultrafast all-optical graphene modulator. Nano Lett. 2014, 14, 955–959. [Google Scholar] [CrossRef] [PubMed]
  35. Wu, Y.; Yao, B.; Zhang, A.; Rao, Y.; Wang, Z.; Cheng, Y.; Gong, Y.; Zhang, W.; Chen, Y.; Chiang, K.S. Graphene-coated microfiber Bragg grating for high-sensitivity gas sensing. Opt. Lett. 2014, 39, 1235–1237. [Google Scholar] [CrossRef]
  36. He, X.; Liu, Z.; Wang, D.N.; Yang, M.; Hu, T.Y.; Tian, J.G. Saturable absorber based on graphene-covered microfiber. IEEE Photon. Technol. Lett. 2013, 25, 1392–1394. [Google Scholar] [CrossRef]
  37. Dai, D.X.; He, S.L. A silicon-based hybrid plasmonic waveguidewith a metal cap for a nano-scale light confinement. Opt. Express 2009, 17, 16646–16653. [Google Scholar] [CrossRef]
  38. Farhat, M.; Guenneau, S.; Bağcı, H. Exciting graphene surface plasmon polaritons through light and sound interplay. Phys. Rev. Lett. 2013, 111, 237404. [Google Scholar] [CrossRef] [Green Version]
  39. Gao, Y.; Ren, G.; Zhu, B.; Wang, J.; Jian, S. Single-mode graphene-coated nanowire plasmonic waveguide. Opt. Lett. 2014, 39, 5909–5912. [Google Scholar] [CrossRef]
  40. Hanson, G.W. Quasi-transverse electromagnetic modes supported by a graphene parallelplate waveguide. J. Appl. Phys. 2008, 104, 084314. [Google Scholar] [CrossRef]
  41. Ju, L.; Geng, B.; Horng, J.; Girit1, C.; Martin, M.; Hao, Z.; Bechtel, H.A.; Liang, X.; Zettl, A.; Shen, Y.R.; et al. Graphene Plasmonics for Tunable Terahertz Metamaterials. Nat. Nanotechnol. 2011, 6, 630–634. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, J.P.; Zhai, X.; Wang, L.L.; Li, H.J.; Xie, F.; Lin, Q.; Xia, S.X. Analysis of mid-infrared surface plasmon modes in a graphene-based cylindrical hybrid waveguide. Plasmonics 2016, 11, 703–711. [Google Scholar] [CrossRef]
  43. Efetov, D.K.; Kim, P. Controlling Electron-Phonon Interactions in Graphene at Ultrahigh Carrier Densities. Phys. Rev. Lett. 2010, 105, 256805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Bolotin, K.I.; Sikes, K.J.; Jiang, Z.; Klima, M.; Fudenberg, G.; Hone, J.; Kim, P.; Stormer, H.L. Ultrahigh electron mobility in suspended graphene. Solid State Commun. 2008, 146, 351–355. [Google Scholar] [CrossRef] [Green Version]
  45. Dai, D.X.; Shi, Y.C.; He, S.L.; Wosinski, L.; Thylen, L. Gain enhancement in a hybrid plasmonic nano-waveguide with a low-index or high-index gain medium. Opt. Express 2011, 19, 12925–12936. [Google Scholar] [CrossRef]
  46. Zafar, R.; Salim, M. Enhanced figure of merit in fano resonance-based plasmonic refractive index sensor. IEEE Sens. J. 2015, 15, 6313–6317. [Google Scholar] [CrossRef]
  47. Wang, B.Q.; Yu, P.; Wang, W.H.; Zhang, X.T.; Kuo, H.C.; Wang, Z.M. High-Q plasmonic resonances: Fundamentals and applications. Adv. Opt. Mate. 2021, 9, 2001520. [Google Scholar] [CrossRef]
  48. Dolatabady, A.; Asgari, S.; Granpayeh, N. Tunable mid-infrared nanoscale graphene-based refractive index sensor. IEEE Sens. J. 2017, 18, 569–574. [Google Scholar] [CrossRef]
  49. Cen, C.; Lin, H.; Huang, J.; Liang, C.; Chen, X.; Tang, Y.; Yi, Z.; Ye, X.; Liu, J.; Yi, Y.; et al. A tunable plasmonic refractive index sensor with nonoring-strip graphene arrays. Sensors 2018, 18, 4489. [Google Scholar] [CrossRef] [Green Version]
  50. Abutoama, M.; Abdulhalim, I. Angular and Intensity modes self-referenced refractive index sensor based on thin dielectric grating combined with thin metal film. IEEE J. Sel. Top. Quantum Electron. 2017, 23, 4600309. [Google Scholar] [CrossRef]
  51. Abutoama, M.; Abdulhalim, I. Self-referenced Biosensor based on thin dielectric grating combined with thin metal film. Opt. Express 2015, 23, 28667–28682. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Schematic of the proposed grating consisting of the graphene-based cylindrical LRSPP waveguide array. (b) Cross-section of the proposed grating.
Figure 1. (a) Schematic of the proposed grating consisting of the graphene-based cylindrical LRSPP waveguide array. (b) Cross-section of the proposed grating.
Nanomaterials 12 01144 g001
Figure 2. (a) Simulated transmission spectra with different periods of waveguide array p for a fixed waveguide radius R = 60 nm. The blue solid line and the red solid line represent the reflection spectra and absorption spectra as p = 180 nm, respectively. (b) Simulated transmission spectra with different waveguide radius R and period of waveguide array p following the relation p = 4R. (cg) show the |E|, Ex, Ex phase, Ey, and |H| distribution of the Sx1 mode on the cross-section of the waveguide, respectively. (hl) show the |E|, Ex, Ex phase, Ey, and |H| distribution of the Sx′1 mode on the cross-section of the waveguide, respectively. The white lines in (c,h) describe the |E| distribution of the Sx1 and Sx′1 modes along the x-axis of the waveguide, respectively. The “+” and “−” symbols in (f,k) represent the positive and negative charges, respectively. This charge distribution is obtained from the EM field boundary conditions of n · ( D 2 D 1 ) = σ .
Figure 2. (a) Simulated transmission spectra with different periods of waveguide array p for a fixed waveguide radius R = 60 nm. The blue solid line and the red solid line represent the reflection spectra and absorption spectra as p = 180 nm, respectively. (b) Simulated transmission spectra with different waveguide radius R and period of waveguide array p following the relation p = 4R. (cg) show the |E|, Ex, Ex phase, Ey, and |H| distribution of the Sx1 mode on the cross-section of the waveguide, respectively. (hl) show the |E|, Ex, Ex phase, Ey, and |H| distribution of the Sx′1 mode on the cross-section of the waveguide, respectively. The white lines in (c,h) describe the |E| distribution of the Sx1 and Sx′1 modes along the x-axis of the waveguide, respectively. The “+” and “−” symbols in (f,k) represent the positive and negative charges, respectively. This charge distribution is obtained from the EM field boundary conditions of n · ( D 2 D 1 ) = σ .
Nanomaterials 12 01144 g002
Figure 3. (a) Simulated transmission spectra with different tSiO2. (b) Resonant wavelength as a function of tSiO2 for different modes. (c,d) show the schematics of the components of the graphene hybrid waveguide array (GHWA) and the graphene-coated nanowire array (GCNA) gratings, respectively. (e,f) show the |E| and Ey distribution of the Mx1 mode on the cross-section of the graphene hybrid waveguide, respectively. (g,h) show the |E| and Ey distribution of the Mx′1 mode on the cross-section of the graphene-coated nanowire waveguide, respectively. The white lines in (e,g) describe the |E| distribution of the Mx1 and Mx′1 modes along the x-axis direction of the corresponding waveguides, respectively. The “+” and “−” symbols in (f,h) represent the positive and negative charges, respectively.
Figure 3. (a) Simulated transmission spectra with different tSiO2. (b) Resonant wavelength as a function of tSiO2 for different modes. (c,d) show the schematics of the components of the graphene hybrid waveguide array (GHWA) and the graphene-coated nanowire array (GCNA) gratings, respectively. (e,f) show the |E| and Ey distribution of the Mx1 mode on the cross-section of the graphene hybrid waveguide, respectively. (g,h) show the |E| and Ey distribution of the Mx′1 mode on the cross-section of the graphene-coated nanowire waveguide, respectively. The white lines in (e,g) describe the |E| distribution of the Mx1 and Mx′1 modes along the x-axis direction of the corresponding waveguides, respectively. The “+” and “−” symbols in (f,h) represent the positive and negative charges, respectively.
Nanomaterials 12 01144 g003
Figure 4. (a,b) show the simulated transmission spectra at different Ef and μ values, respectively. (c,d) show the extinction ratio (ER) of the Sx1 and Sx′1 modes with varying Ef and μ values, respectively.
Figure 4. (a,b) show the simulated transmission spectra at different Ef and μ values, respectively. (c,d) show the extinction ratio (ER) of the Sx1 and Sx′1 modes with varying Ef and μ values, respectively.
Nanomaterials 12 01144 g004
Figure 5. (a) Resonant wavelength and effective index as a function of Ef for different modes. (b) Resonant wavelength and effective index as a function of μ for different modes.
Figure 5. (a) Resonant wavelength and effective index as a function of Ef for different modes. (b) Resonant wavelength and effective index as a function of μ for different modes.
Nanomaterials 12 01144 g005
Figure 6. (a,b) show the ER values and the resonant wavelength as a function of n1 for the two modes, respectively. (c,d) show the resonant wavelength and the figure of merit (FOM) as a function of n3 for the two modes.
Figure 6. (a,b) show the ER values and the resonant wavelength as a function of n1 for the two modes, respectively. (c,d) show the resonant wavelength and the figure of merit (FOM) as a function of n3 for the two modes.
Nanomaterials 12 01144 g006
Figure 7. (a) Simulated transmission spectra with different n2. (b,c) show the |E| distribution of the Sx1 and Sx′2 modes, respectively.
Figure 7. (a) Simulated transmission spectra with different n2. (b,c) show the |E| distribution of the Sx1 and Sx′2 modes, respectively.
Nanomaterials 12 01144 g007
Figure 8. (a) Simulated transmission spectra for low and high-order modes. The insert shows the transmission spectra of the region surrounded by the pink short dash line. (bf) show the |E| distribution of the Sx2, Sx′3, Sx3, Sx′4, and Sx′5 modes, respectively. (gk) show the |H| distribution of the Sx2, Sx′3, Sx3, Sx′4, and Sx′5 modes, respectively.
Figure 8. (a) Simulated transmission spectra for low and high-order modes. The insert shows the transmission spectra of the region surrounded by the pink short dash line. (bf) show the |E| distribution of the Sx2, Sx′3, Sx3, Sx′4, and Sx′5 modes, respectively. (gk) show the |H| distribution of the Sx2, Sx′3, Sx3, Sx′4, and Sx′5 modes, respectively.
Nanomaterials 12 01144 g008
Figure 9. (a) Simulated transmission spectra with different incident angular θ for an oblique incident light. (b) Transmission values as a function of θ for the different wavelengths of the incident light. (ce) show the |E| distribution of the Sy1 and Sy′1 modes, respectively. (df) show the |H| distribution of the Sy1 and Sy′1 modes, respectively.
Figure 9. (a) Simulated transmission spectra with different incident angular θ for an oblique incident light. (b) Transmission values as a function of θ for the different wavelengths of the incident light. (ce) show the |E| distribution of the Sy1 and Sy′1 modes, respectively. (df) show the |H| distribution of the Sy1 and Sy′1 modes, respectively.
Nanomaterials 12 01144 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; Wang, W.; Xie, F.; Zhang, X.; Zhou, X.; Yuan, Y.; Wang, L. Excitation of Surface Plasmon Polariton Modes with Double-Layer Gratings of Graphene. Nanomaterials 2022, 12, 1144. https://doi.org/10.3390/nano12071144

AMA Style

Liu J, Wang W, Xie F, Zhang X, Zhou X, Yuan Y, Wang L. Excitation of Surface Plasmon Polariton Modes with Double-Layer Gratings of Graphene. Nanomaterials. 2022; 12(7):1144. https://doi.org/10.3390/nano12071144

Chicago/Turabian Style

Liu, Jianping, Weilin Wang, Fang Xie, Xiaoming Zhang, Xia Zhou, Yijun Yuan, and Lingling Wang. 2022. "Excitation of Surface Plasmon Polariton Modes with Double-Layer Gratings of Graphene" Nanomaterials 12, no. 7: 1144. https://doi.org/10.3390/nano12071144

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop