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

Anisotropic infrared plasmonic broadband absorber based on graphene-black phosphorus multilayers

Open Access Open Access

Abstract

Two-dimensional materials (2DMs) such as graphene and black phosphorus (BP) have aroused considerable attentions in the past few years. Engineering and enhancing their light-matter interaction is possible due to their support for localized surface plasmon resonances in the infrared regime. In this paper, we have proposed an infrared broadband absorber consisting of multilayer graphene-BP nanoparticles sandwiched between dielectric layers. Benefiting from the properties of graphene and BP, the absorber exhibits both perfect broadband responses and strong anisotropy beyond individual graphene and BP layers. The absorber is tunable with the variation of geometric parameters as well as the doping levels of graphene and BP. The physical insight is revealed by electric field distributions. Furthermore, the angular robustness for incident wave is investigated. The proposed anisotropic omnidirectional broadband absorber may have promising potential applications in various biosensing, communication and imaging systems.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Two-dimensional materials (2DMs) have attracted extensive attention as plasmonic materials due to their remarkable electrical and optical performance over the past decade [1]. Among these 2DMs, graphene is the most popular one since its isolation by the Manchester group in 2004 [2]. Graphene is a single-layered carbon atoms arranged in a honey comb lattice. In the terahertz and infrared regimes, graphene can support surface plasmons [3–7], which have been widely utilized to design novel optoelectronic devices, such as sensors [8], field effect transistors [9], polarizers [10], modulators [11], photodetectors [12], and absorbers [13].

Although graphene has the highest carrier mobility among 2DMs, its zero or near-zero band gap limits its applications that require high on-off ratio and strong light-matter interaction [14]. As an alternative 2DM, black phosphorus (BP) has also attracted enormous attentions because of its unique optoelectronic properties, such as thickness-dependent tunable band gap [15], high carrier mobility and carrier density [16]. Besides, the monolayer BP shows in-plane anisotropic properties due to its hexagonal lattice with a puckered structure formed by phosphorus atoms [17,18]. Recently, various structures and methods have been investigated to strengthen the light-BP interaction in the BP-based metamaterials in the terahertz and infrared range [19–29]. In [19], plasmonic excitations were realized in BP, and the plasmon frequency was related to the carrier concentration and effective couplings between conduction band and valence band. Liu et al. designed and theoretically demonstrated localized surface plasmon in monolayer BP using periodic nanoribbon and nanopatch [20]. Qing et al. proposed and investigated a metamaterial perfect absorber consisting of a BP monolayer, a photonic crystal, and a perfect electric conductor (PEC) to enhance light absorption at terahertz frequencies [21]. Han et al. exploited the behaviors of anisotropic plasmons inside BP in different systems including individual nanoribbon, vertically offset paired ones, and nanoribbon/sheet hybrid system [22]. Fang et al. numerically proposed a periodic bowtie structure based on BP, in which the localized surface plasmons can be excited in the BP nanoantennas at terahertz regime [23]. Wang et al. proposed an infrared absorber composing of multilayered BP sandwiched between dielectric layers, in which the incident infrared light can be efficiently dissipated by increasing the number of BP layers [24]. Afterwards, Wang et al. further designed a dual-band absorber consisting of orthogonally stacked BP nanoribbons, which absorbs incident infrared light at two frequency bands with polarization-independent property [25].

However, most of the above BP-based structures have the disadvantages of weak absorption rate with low doping concentration, or require relatively complicated fabrication technique. Moreover, the resonance band cannot be tuned effectively and flexibly. This is mainly because the plasmon resonances in individual doped and patterned BP are relatively weak, which limits its anisotropic potentials. Therefore, Nong et al. investigated the hybridization of surface plasmons between graphene and BP in the strong coupling regime [30]. Hong et al. combined graphene and BP plasmonic properties to realize polarization-dependent perfect absorption [31]. Both strong and anisotropic plasmonic absorption are numerically demonstrated in the graphene-BP hybridization structures. Nevertheless, absorption bandwidths of the mentioned structures are often narrow due to the single resonance generated from graphene and BP.

In this paper, we propose an anisotropic infrared plasmonic broadband absorber based on graphene-BP multilayers. Utilizing the advantages of graphene and BP plasmons, the absorber shows both strong and anisotropic plasmon resonance that are not available in either individual graphene or BP layer. By stacking multilayer graphene-BP pairs, the absorption band can be broadened with absorption rate larger than 90%. Furthermore, absorption spectrum can be tuned by geometric parameters and doping levels. The physical mechanism is also revealed by electric field distributions. To illustrate the significance and novelty of the proposed structure, main characteristics of state-of-the-art absorbers based on 2DMs are listed in Table 1 for comparisons.

Tables Icon

Table 1. Comparisons between plasmonic absorbers at infrared frequencies.

2. Models and method

Here, we propose and investigate a three-layer graphene-BP-sandwiched absorber (GBPSA) based on sub-wavelength plasmonic configuration as shown in Fig. 1. The proposed GBPSA consists of stacked graphene-BP bilayers supported by Al2O3 layers with a relative permittivity of ~3.2 [32]. The monolayer graphene and BP are separated by a hexagonal boron nitride (hBN) layer, which is used as an insulating spacer to maintain higher carrier mobility of graphene and BP and prevent carrier transport between graphene and BP. The optical parameter of lossless hBN in simulations is obtained from the literature [33]. The thickness t2 is much smaller than the wavelength in the entire range of interest, thus the top dielectric layer has a negligible influence on the absorption spectrum. However, it prevents the top graphene-BP bilayer from harmness induced by environment, which enable GBPSA to work in different circumstances with high stability for a long period. The finite element method (FEM)-based frequency domain solver of the commercial software COMSOL is used to investigate the broadband absorption properties of GBPSA. Floquet periodic boundary conditions are applied in the x- and y-axis directions. The incident infrared light is imposed from the top surface of structure. PEC boundary condition is applied in the bottom surface. Triangular meshes of user-defined size are used for graphene and BP layers due to their localized enhanced electromagnetic fields. Tetrahedral meshes are applied for the remaining domains of the structure. The absorption rate of the absorbers can be expressed as A = 1−TR, where transmission T = |S21|2, reflectance R = |S11|2, and S represents the scattering parameter, respectively. Owing to the PEC used in the substrate, the transmission T is equal to zero in the infrared frequency range. As a result, absorption rate expression can be simplified as A = 1−R. In this work, we assume graphene and BP as 2D conductive surfaces with zero thickness, which requires much less discrete meshes surrounding graphene and BP [34].

 figure: Fig. 1

Fig. 1 Schematic of the proposed three-layer graphene-BP-sandwiched absorber (GBPSA) in (a) perspective view and (b) cross-section view. t1 is the thickness of hBN layer between graphene and BP. t2 and d are the thickness of the upper three Al2O3 layers and bottom Al2O3 layer, respectively. w1, w2 and w3 are the widths of graphene-BP nanoparticles in different layers (Three layers can be ranked in arbitrary order). p is the periodicity of the periodic GBPSA structure.

Download Full Size | PDF

Kubo formulas are commonly used to describe the surface conductivity of graphene σ(ω) as below [35]:

σ(ω,μc,Г,T)=σintra+σinter
σintra=je2π2(ωj2Г)0ξ(fd(ξ,μc,T)ξfd(ξ,μc,T)ξ)dξ
σinter=je2(ωj2Г)π20fd(ξ,μc,T)fd(ξ,μc,T)(ωj2Г)24(ξ/)2dξ
fd(ξ,μc,T)=(e(ξμc)/kBT+1)1
where σintra and σinter represent the intraband and interband component of surface conductivity originated from the corresponding carrier transition, respectively. e denotes the electron charge, ω is the angular frequency, T is the Kelvin temperature, kB is the Boltzmann constant, ħ is the reduced Planck constant, Г = 1/(2τ) is the scattering rate where τ is the relaxation time of electron-phonon, ξ is the electrons energy, μc is the chemical potential and fd (ξ, μc, T) is the Fermi-Dirac distribution. Based on the experimental work from [36], τ is taken as 0.3 ps and μc is taken as 0.5 eV in the simulations below. Room temperature T = 300 K is assumed for all simulations. Particularly, in the mid-infrared and far-infrared range, the energy of the incident photon is much smaller than the Fermi level of doping graphene. Thus, intraband transition is dominated in the light-graphene interaction, while the interband transition can be ignored.

For BP model, its surface conductivity can be calculated by employing a simple semi-classical Drude model as [19]

σjj=iDjπ(ω+iη)
Dj=πe2nsmj
where j denotes the concerned direction and Dj is the Drude weight. The electron mass along the x-direction and y-direction can be evaluated by

mcx=22γ2Δ+ηc
mcy=22νc

For single layer BP, we have γ=4aπeVm, Δ = 2 eV, ηc=20.4m0,νc=21.4m0. The electron doping is chosen as ns = 2×1013 cm−2, while the relaxation rate is chosen as η = 10 meV. a denotes the scale length of the BP and πa denotes the width of the Brillouin Zone.

3. Results and discussions

In order to elucidate the anisotropic absorption properties of the proposed GBPSA, light absorption rate of single-layer GBPSA with different nanoparticle widths is plotted in Fig. 2(a). Both transverse electric (TE) and transverse magnetic (TM) polarized infrared incidence could excite the localized surface plasmons in the finite length of graphene-BP nanoparticles [20]. However, due to different effective mass along different crystal directions in BP, the optical loss under TM and TE incidence shows obvious differences. When the incident electric field is along the armchair direction of BP crystal (TE), the absorption peaks are red-shifted from 16.21 μm to 19.91 μm as w increases from 105 nm to 145 nm, with near-unity absorption. On the other hand, the resonant wavelength is red-shifted from 14.01 μm to 17.31μm under TM incidence, whose electric filed is along the zigzag direction of BP. The redshift is mainly attributed to the increase of the effective length for carrier vibration inside graphene-BP nanoparticles. The perfect absorption occurs when the impedance matching condition between the GBPSA and free space is satisfied. The width-tunable anisotropic absorption capacity of GBPSA could be utilized to design reflective polarizers or polarized light filters working in different frequency regions.

 figure: Fig. 2

Fig. 2 Absorption rate of (a) single-layer and (b) three-layer GBPSA. (c) Optical losses inside different materials of GBPSA under TE incidence. The parameters are w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm, under normal incidence. The width of single layer graphene-BP is 115 nm.

Download Full Size | PDF

By stacking graphene-BP nanoparticles of different widths layer by layer, the bandwidth can be significantly improved as shown in Fig. 2(b). For TE polarization, three-layer GBPSA has broadband absorption with a 90% absorption bandwidth of 2.01 μm, from 17.11 μm to 19.12 μm. The center wavelength λc can be obtained by λc = (λ - + λ +)/2 = 18.12 μm, where λ and λ + represent the low and upper wavelength edges of 90% absorption, respectively. The fractional bandwidth, the ratio of the absolute bandwidth to the center wavelength, is about 11.1%. For TM polarization, the corresponding bandwidth is 2.91 μm, from 13.71 μm to 16.62 μm. The center wavelength λc is 15.17 μm and the fractional bandwidth is 19.2%.

For comparison, Fig. 2(b) also plots the absorption rate in a three-layer BP-sandwiched structure without graphene. The plasmonic resonance of BP with low doping is relatively weak for both TE (2.3%) and TM (41.3%) polarization. The polarization-dependent capability is attributed to the anisotropic plasmonic dispersion of BP. Meanwhile, the absorption rate of three-layer graphene-sandwiched structure without BP is also shown in Fig. 2(b). Three strong resonances in graphene are observed with absorption rates of 84.2%, 92.1% and 86.3%, respectively. The optical response is insensitive to the incident polarization due to the in-plane isotropic surface conductivity of graphene. Accordingly, the advantages of graphene and BP are combined in the GBPSA configuration, which reveals both strong and anisotropic plasmon resonances.

As can be observed in Fig. 2(c), both the losses in BP and graphene contribute to the total absorption of GBPSA. Due to the strong plasmon resonance in graphene, the incident infrared light could be highly concentrated around graphene layers. BP layers are placed in the vicinity of graphene, so the confined and enhanced electromagnetic fields penetrate BP and dissipate in the lossy material. Therefore, the enhanced absorption inside BP is significant considering the weak plasmon resonance in individual BP as shown in Fig. 2(b). However, the loss in graphene is a little stronger than that in BP, which can be mainly attributed to the relatively larger surface conductivity of graphene than BP in the frequency range.

In order to reveal the physical mechanism of anisotropic broadband absorption in GBPSA, the electric field distributions of three-layer GBPSA are plotted in Fig. 3. For TE polarization, at the resonance wavelength of 18.8 μm as shown in Figs. 3(a) and 3(d), the incident wave could excite carriers to oscillate along the finite particle side (y-axis) and induce an independent electric dipole response both in graphene and BP layers. The dipole electric responses inside graphene layer and BP layer have an opposite sign, which will generate antiparallel surface current on the graphene and BP plane. Therefore, the incident electromagnetic fields are focused around the edges of the 1st graphene-BP layer and induce the effects of near field enhancement and energy concentration. The energy assumption inside lossy material such as graphene and BP can be calculated by the following formula:

A(λ)=2πcλε''V|El|2dV
where El is the electric field inside materials, V is the volume of lossy material, ε is the imaginary part of material permittivity and c is the speed of light in vacuum. Thus, the enhanced fields penetrating graphene-BP nanoparticle dissipate in the lossy material and contribute to the enhanced absorption inside graphene and BP. At λ = 18.1 μm and λ = 17.4 μm, the resonance wavelengths of the 3rd and 2nd layer nanoparticles respectively, the incident electric fields are dramatically enhanced and concentrated on the rims of 3rd and 2nd layer of graphene-BP nanoparticle as shown in Figs. 3(b) and 3(c), respectively. This is mainly because the resonance wavelength is related to the width of corresponding graphene-BP particle.

 figure: Fig. 3

Fig. 3 Electric field distributions of three-layer GBPSA, (a)-(c) are the cross-section views and (d)-(f) are the top views, where w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm.

Download Full Size | PDF

On the contrary, for λ = 15.8 μm as shown in Fig. 3(e), few enhanced near field appears in graphene-BP nanoparticles since this wavelength is far away from all the plasmonic resonance wavelengths of these three graphene-BP particles in the TE absorption spectrum. However, for TM polarization, λ = 15.8 μm is exactly right the resonance wavelength for the 1st graphene-BP particle. Hence, the incident TM wave could induce the carriers in graphene and BP to vibrate in the x-axis direction as shown in Fig. 3(f). The graphene and BP layers are positioned in the vicinity of each other, the evanescent near field scattered by graphene and BP nanoparticle is considerably intense compared to the exciting field, which leads to the strong coupling of scattered field from each graphene-BP nanoparticle. Consequently, the superposition of the inverse electromagnetic fields induced by electric dipole that is excited by the incident light makes contributions to the suppressed reflectance. Thus, perfect absorption occurs at the resonance wavelength of 15.8 μm for TM polarization as demonstrated in Fig. 2(b).

According to Figs. 2 and 3, therefore, it is predicted that an anisotropic ultra-broadband absorption spectrum can be further realized by increasing the number of stacked layers in the proposed GBPSA and changing the dimensions of each graphene-BP particle.

By altering the geometric dimensions of GBPSA configurations, the anisotropic broadband absorption spectra can be tuned as demonstrated in Fig. 4. The resonance properties of Fabry-Perot resonator formed by the graphene-BP layers and PEC substrate are highly sensitive to the thickness variation of the dielectric spacer layers, which facilitate the tuning of absorption band by slightly changing the spacer thickness, as shown in Figs. 4(a) and 4(b). When the thickness d<1.6 μm for TE polarization in Fig. 4(a), the middle absorption peak is relatively weak with an absorption rate below 90%. As d increases from 1.6 μm to 2.2 μm, the middle absorption peak is strong enough and the broad bandwidth decreases because the left and right peaks become closer to the middle peak. When d>2.2 μm, the broadband is divided into two separated bands. For TM polarization in Fig. 4(b), as d increases from 1 μm to 2.2 μm, the bandwidth increases at first and then decreases. There is an optimal thickness of the dielectric layer, at which the absorption bandwidth reaches the maximum. When d>2.2 μm, the bandwidth is divided into two narrow bands as well as the TE polarization. This is attributed to the increase of the effective thickness of these multiple Fabry-Perot resonators, which are strongly coupled between each graphene-BP layer and the gold substrate. The phenomenon can be utilized for designing both dual-band and broadband anisotropic absorbers for novel optoelectronic devices.

 figure: Fig. 4

Fig. 4 Absorption spectra as a function of geometric parameters: (a) and (b) for dielectric thickness d, (c) and (d) for insulator thickness t1, (e) and (f) for periodic spacing p, under TE and TM incidences, respectively. Other parameters are as in Fig. 2(b).

Download Full Size | PDF

The coupling strength between graphene and BP layers depends on the thickness of the insulator layer t1. Therefore, Figs. 4(c) and 4(d) plot the absorption spectra versus t1. In Fig. 4(c), the absorption band tends to be slightly red-shifted when t1 increases from 2 nm to 13 nm, and then keeps almost unchanged when t1 still rises up to 40 nm. The bandwidth remains almost constant with the variation of t1. In Fig. 4(d), the absorption band is obviously red-shifted as t1 increases from 2 nm to 42 nm, while the bandwidth gradually increases at first until it reaches the maximum and then decreases afterwards. This originates from the fact that for TM polarization, the plasmonic resonance in BP layer is stronger than TE polarization as shown in Fig. 2(b). As a consequence, the coupling between graphene and BP layer for TM polarization is more sensitive to the thickness of the insulator than TE polarization.

As a critical parameter of GBPSA configuration, periodic spacing p affects the anisotropic broadband performance, thus the absorption spectra under different values of p are illustrated in Figs. 4(e) and 4(f). When the periodic spacing is less than 280 nm, the coupling between adjacent graphene-BP nanoparticles is relatively strong, so the bandwidth above 90% absorption rate is broad for both of TE and TM polarizations. As p increases from 280 nm to 1000 nm, the absorption efficiency of GBPSA has a dramatic drop, while the broadband is divided into two narrow and weak absorption bands. The absorption efficiency is related to the filling factor of graphene-BP nanoparticles, which can be defined as F = w/p. F becomes smaller as p increases, then the intensity of the electric field concentration is significantly weaken inside graphene-BP nanoparticles, so the absorption efficiency remarkably drops.

It is an effective way to tune the absorption spectrum via w, d, t, p and other dimensions in the design of a GBPSA with specific requirements. Besides, according to graphene and BP model formulas in Eqs. (1)-(4) and Eqs. (5)-(8), the surface conductivities of graphene and BP are determined by μc and ns, respectively, which represent the doping level of corresponding material. Therefore, the performance of GBPSA can be manipulated by changing μc and ns as shown in Fig. 5. From a practical point of view, we choose μc between 0.4 eV and 0.8 eV which are verified by experiments in previous works [36]. A maximum theoretical carrier density was demonstrated to be ns = 2.6 × 1014 cm−2 [16,37], so we use a moderate value between 1013 cm−2 and 1014 cm−2. In Figs. 5(a) and 5(b), as μc increases from 0.4 eV to 0.8 eV, the absorption spectra for both TE and TM polarizations are blue-shifted obviously, with center wavelength changing from 20.1 μm to 14.8 μm for TE polarization, and from 16.6 μm to 12.7 μm for TM polarization. The bandwidth above 90% absorption gradually decreases from 2.8 μm to 0.8 μm for TE polarization, and from 3.9 μm to 2.3 μm for TM polarization. This is in accordance with the previous work [38], which can be explained by the higher conductivity of graphene with increasing μc,, resulting in electromagnetic energy loss and absorption. On the other hand, for monolayer BP, the plasmonic resonance wavelength λp can be calculated as λpw/ns, where w is the length of nanoparticles [19]. Therefore, while w is fixed, as ns increases from 1013 cm−2 to 1014 cm−2, the center wavelength is blue-shifted from 18.7 μm to 15.2 μm for TE incidence as shown in Fig. 5(c). For TM incidence, the bandwidth tends to be narrower and narrower as ns increases as demonstrated in Fig. 5(d). The explanation is similar with that of graphene above. Generally, both real and imaginary parts of BP surface conductivity become higher with the increase of ns [24,25], which means more optical loss and absorption in BP.

 figure: Fig. 5

Fig. 5 Absorption spectra as a function of doping level: (a) and (b) for different μc of graphene, (c) and (d) for different ns of BP, under TE and TM incidences, respectively, where w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm.

Download Full Size | PDF

In the discussions above, anisotropic broadband absorption properties of GBPSA were investigated under normal incidence. However, the tolerance of wide incident angles is necessary for broadband infrared absorbers in the application of novel photonic devices. Hence, we further investigate the optical response of GBPSA under oblique infrared incidences. In Fig. 6(a), the dependence of absorption spectra on the incident angles θ for TE polarization is demonstrated. It is observed that GBPSA has stable absorption bandwidth with θ up to 64°, while the middle absorption peak decreases gradually. When θ>70°, the broadband absorption is divided into two absorption peaks because the middle absorption peak drops sharply. For TM polarization shown in Fig. 6(b), as θ increases from 0° to 62°, the absorption bandwidth will be wider slightly while the broadband absorption rate remains larger than 80%. The peak absorption rate maintains larger than 70% with a sufficient bandwidth of 4 μm while the incident angle increases up to 74°. Besides, for both TE and TM incident infrared lights, the center wavelengths of the absorption band keep almost unchanged with the increase of θ. The excellent angular stability is attributed to the tightly confined surface plasmon resonances in graphene-BP nanoparticles, which is not sensitive to the incident angle.

 figure: Fig. 6

Fig. 6 (a) and (b) are absorption spectra as a function of incident angles θ for TE polarization and TM polarization respectively. (c) is absorption spectra for various polarization angles β under normal incidence. The parameters are w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm.

Download Full Size | PDF

To further investigate the polarization dependence of GBPSA, we plot the absorption spectra under incidence of different polarization angles β. The polarization angle of TE incidence is assumed to be 0°. As can be seen from Fig. 6(c), when β = 0°, the spectrum is the same as the TE polarization in Fig. 2(b). As β increases to 90°, the incident light turns out to be a TM incidence. Thus, the absorption spectrum is the same as the TM polarization in Fig. 2(b). When the polarization angle is between 0° and 90°, the incident electric field has both x and y- components. Therefore, the hybridization mode will induce the electrons in BP to vibrate in both armchair and zigzag directions. As a consequence, localized plasmon resonances in armchair and zigzag directions can be excited simultaneously with reduced light intensity at corresponding x and y-directions. Interestingly, when the polarization angle is between 30° to 60°, three absorption peaks can be obviously observed in the absorption spectra. This is mainly because the resonance modes in three graphene-BP particles are not closely positioned together.

4. Conclusions

In conclusion, we have theoretically proposed an anisotropic broadband infrared absorber based on multilayer graphene-BP nanoparticles in a sandwiched metamaterial configuration. Combining the advantages of graphene and BP localized surface plasmons, the GBPSA shows highly anisotropic plasmonic responses that are available in neither individual graphene nor BP sheets. By stacking slightly different widths of graphene-BP nanoparticles in a sandwiched configuration, the absorption bandwidth can be increased due to the different resonant modes closely positioned together. The absorption spectra can be tuned either by the geometric parameters of GBPSA or by the doping levels of graphene and BP. Moreover, GBPSA can tolerate a wide range of incident angles for both TE and TM polarizations. Our approach can be used in the design of novel infrared sensors and tunable reflective polarizer in symmetrical structure.

Funding

National Natural Science Foundation of China (61601390, 51702271), the Young and Middle-aged Teachers Education and Scientific Research Foundation of Fujian Province (JAT160324, JAT170405, JAT170407), the High Level Talent Project of Xiamen University of Technology (YKJ16016R, YKJ16011R), the Shenzhen Science and Technology Innovation Commission (JCYJ20170306141249935).

Acknowledgments

The authors thank Dr. Y. Zhou from Xi'an University of Science and Technology for language check.

References

1. F. Xia, H. Wang, D. Xiao, M. Dubey, and A. Ramasubramaniam, “Two-dimensional material nanophotonics,” Nat. Photonics 8(12), 899–907 (2014). [CrossRef]  

2. K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov, and A. K. Geim, “Two-dimensional atomic crystals,” Proc. Natl. Acad. Sci. U.S.A. 102(30), 10451–10453 (2005). [CrossRef]   [PubMed]  

3. F. J. G. D. Abajo, “Graphene plasmonics: challenges and opportunities,” ACS Photonics 1(3), 135–152 (2014). [CrossRef]  

4. A. Politano and G. Chiarello, “Plasmon modes in graphene: status and prospect,” Nanoscale 6(19), 10927–10940 (2014). [CrossRef]   [PubMed]  

5. T. Low and P. Avouris, “Graphene plasmonics for terahertz to mid-infrared applications,” ACS Nano 8(2), 1086–1101 (2014). [CrossRef]   [PubMed]  

6. A. N. Grigorenko, M. Polini, and K. S. Novoselov, “Graphene plasmonics,” Nat. Photonics 6(11), 749–758 (2012). [CrossRef]  

7. 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(8), 3370–3377 (2011). [CrossRef]   [PubMed]  

8. W. Tang, L. Wang, X. Chen, C. Liu, A. Yu, and W. Lu, “Dynamic metamaterial based on the graphene split ring high-Q Fano-resonnator for sensing applications,” Nanoscale 8(33), 15196–15204 (2016). [CrossRef]   [PubMed]  

9. L. Wang, X. Chen, A. Yu, Y. Zhang, J. Ding, and W. Lu, “Highly sensitive and wide-band tunable terahertz response of plasma waves based on graphene field effect transistors,” Sci. Rep. 4(1), 5470 (2014). [CrossRef]   [PubMed]  

10. Z. H. Zhu, C. C. Guo, K. Liu, J. F. Zhang, W. M. Ye, X. D. Yuan, and S. Q. Qin, “Electrically controlling the polarizing direction of a graphene polarizer,” J. Appl. Phys. 116(10), 104304 (2014). [CrossRef]  

11. Y. Cai, J. Zhu, Q. H. Liu, T. Lin, J. Zhou, L. Ye, and Z. Cai, “Enhanced spatial near-infrared modulation of graphene-loaded perfect absorbers using plasmonic nanoslits,” Opt. Express 23(25), 32318–32328 (2015). [CrossRef]   [PubMed]  

12. J. Zhang, Z. Zhu, W. Liu, X. Yuan, and S. Qin, “Towards photodetection with high efficiency and tunable spectral selectivity: graphene plasmonics for light trapping and absorption engineering,” Nanoscale 7(32), 13530–13536 (2015). [CrossRef]   [PubMed]  

13. Y. Cai, J. Zhu, and Q. H. Liu, “Tunable enhanced optical absorption of graphene using plasmonic perfect absorbers,” Appl. Phys. Lett. 106(4), 043105 (2015). [CrossRef]  

14. Y. Chen, G. Jiang, S. Chen, Z. Guo, X. Yu, C. Zhao, H. Zhang, Q. Bao, S. Wen, D. Tang, and D. Fan, “Mechanically exfoliated black phosphorus as a new saturable absorber for both Q-switching and Mode-locking laser operation,” Opt. Express 23(10), 12823–12833 (2015). [CrossRef]   [PubMed]  

15. L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen, and Y. Zhang, “Black phosphorus field-effect transistors,” Nat. Nanotechnol. 9(5), 372–377 (2014). [CrossRef]   [PubMed]  

16. Y. Saito and Y. Iwasa, “Ambipolar insulator-to-metal transition in black phosphorus by ionic-liquid gating,” ACS Nano 9(3), 3192–3198 (2015). [CrossRef]   [PubMed]  

17. T. Low, A. S. Rodin, A. Carvalho, Y. Jiang, H. Wang, F. Xia, and A. H. C. Neto, “Tunable optical properties of multilayers black phosphorus thin films,” Phys. Rev. B Condens. Matter Mater. Phys. 90(7), 075434 (2014). [CrossRef]  

18. H. Liu, A. T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tománek, and P. D. Ye, “Phosphorene: an unexplored 2D semiconductor with a high hole mobility,” ACS Nano 8(4), 4033–4041 (2014). [CrossRef]   [PubMed]  

19. T. Low, R. Roldán, H. Wang, F. Xia, P. Avouris, L. M. Moreno, and F. Guinea, “Plasmons and screening in monolayer and multilayer black phosphorus,” Phys. Rev. Lett. 113(10), 106802 (2014). [CrossRef]   [PubMed]  

20. Z. Liu and K. Aydin, “Localized surface plasmons in nanostructured monolayer black phosphorus,” Nano Lett. 16(6), 3457–3462 (2016). [CrossRef]   [PubMed]  

21. Y. M. Qing, H. F. Ma, and T. J. Cui, “Tailoring anisotropic perfect absorption in monolayer black phosphorus by critical coupling at terahertz frequencies,” Opt. Express 26(25), 32442–32450 (2018). [CrossRef]   [PubMed]  

22. L. Han, L. Wang, H. Xing, and X. Chen, “Active tuning of midinfrared surface plasmon resonance and its hybridization in black phosphorus sheet array,” ACS Photonics 5(9), 3828–3837 (2018). [CrossRef]  

23. C. Fang, Y. Liu, G. Han, Y. Shao, J. Zhang, and Y. Hao, “Localized plasmon resonances for black phosphorus bowtie nanoantennas at terahertz frequencies,” Opt. Express 26(21), 27683–27693 (2018). [CrossRef]   [PubMed]  

24. J. Wang and Y. Jiang, “Infrared absorber based on sandwiched two-dimensional black phosphorus metamaterials,” Opt. Express 25(5), 5206–5216 (2017). [CrossRef]   [PubMed]  

25. J. Wang, Y. Jiang, and Z. Hu, “Dual-band and polarization-independent infrared absorber based on two-dimensional black phosphorus metamaterials,” Opt. Express 25(18), 22149–22157 (2017). [CrossRef]   [PubMed]  

26. X. Ni, L. Wang, J. Zhu, X. Chen, and W. Lu, “Surface plasmons in a nanostructured black phosphorus flake,” Opt. Lett. 42(13), 2659–2662 (2017). [CrossRef]   [PubMed]  

27. X. Wang, Q. Ma, L. Wu, J. Guo, S. Lu, X. Dai, and Y. Xiang, “Tunable terahertz/infrared coherent perfect absorption in a monolayer black phosphorus,” Opt. Express 26(5), 5488–5496 (2018). [CrossRef]   [PubMed]  

28. Y. M. Qing, H. F. Ma, and T. J. Cui, “Strong coupling between magnetic plasmons and surface plasmons in a black phosphorus-spacer-metallic grating hybrid system,” Opt. Lett. 43(20), 4985–4988 (2018). [CrossRef]   [PubMed]  

29. D. T. Debu, S. J. Bauman, D. French, H. O. H. Churchill, and J. B. Herzog, “Tuning infrared plasmon resonance of black phosphorene nanoribbon with a dielectric interface,” Sci. Rep. 8(1), 3224 (2018). [CrossRef]   [PubMed]  

30. J. Nong, W. Wei, W. Wang, G. Lan, Z. Shang, J. Yi, and L. Tang, “Strong coherent coupling between graphene surface plasmons and anisotropic black phosphorus localized surface plasmons,” Opt. Express 26(2), 1633–1644 (2018). [CrossRef]   [PubMed]  

31. Q. Hong, F. Xiong, W. Xu, Z. Zhu, K. Liu, X. Yuan, J. Zhang, and S. Qin, “Towards high performance hybrid two-dimensional material plasmonic devices: strong and highly anisotropic plasmonic resonances in nanostructured graphene-black phosphorus bilayer,” Opt. Express 26(17), 22528–22535 (2018). [CrossRef]   [PubMed]  

32. E. D. Palik, Handbook of Optical Constants of Solids 3 (Academic, 1998).

33. T. Ishii and T. Sato, “Growth of single crystals of hexagonal boron nitride,” J. Cryst. Growth 61(3), 689–690 (1983). [CrossRef]  

34. J. Zhu, J. Cheng, L. Zhang, and Q. H. Liu, “Modeling of 2D graphene material for plasmonic hybrid waveguide with enhanced near-infrared modulation,” Mater. Lett. 186, 53–56 (2017). [CrossRef]  

35. G. W. Hanson, “Dyadic Green’s functions and guided surface waves for a surface conductivity model of graphene,” J. Appl. Phys. 103(6), 064302 (2008). [CrossRef]  

36. Y. Jiang, H. Zhang, J. Wang, C. N. Gao, J. Wang, and W. P. Cao, “Design and performance of a terahertz absorber based on patterned graphene,” Opt. Lett. 43(17), 4296–4299 (2018). [CrossRef]   [PubMed]  

37. Z. C. Luo, M. Liu, Z. N. Guo, X. F. Jiang, A. P. Luo, C. J. Zhao, X. F. Yu, W. C. Xu, and H. Zhang, “Microfiber-based few-layer black phosphorus saturable absorber for ultra-fast fiber laser,” Opt. Express 23(15), 20030–20039 (2015). [CrossRef]   [PubMed]  

38. Y. Cai, K. D. Xu, R. Guo, J. Zhu, and Q. H. Liu, “Graphene-based plasmonic tunable dual-band bandstop filter in the far-infrared region,” IEEE Photonics J. 10(6), 1 (2018). [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 Schematic of the proposed three-layer graphene-BP-sandwiched absorber (GBPSA) in (a) perspective view and (b) cross-section view. t1 is the thickness of hBN layer between graphene and BP. t2 and d are the thickness of the upper three Al2O3 layers and bottom Al2O3 layer, respectively. w1, w2 and w3 are the widths of graphene-BP nanoparticles in different layers (Three layers can be ranked in arbitrary order). p is the periodicity of the periodic GBPSA structure.
Fig. 2
Fig. 2 Absorption rate of (a) single-layer and (b) three-layer GBPSA. (c) Optical losses inside different materials of GBPSA under TE incidence. The parameters are w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm, under normal incidence. The width of single layer graphene-BP is 115 nm.
Fig. 3
Fig. 3 Electric field distributions of three-layer GBPSA, (a)-(c) are the cross-section views and (d)-(f) are the top views, where w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm.
Fig. 4
Fig. 4 Absorption spectra as a function of geometric parameters: (a) and (b) for dielectric thickness d, (c) and (d) for insulator thickness t1, (e) and (f) for periodic spacing p, under TE and TM incidences, respectively. Other parameters are as in Fig. 2(b).
Fig. 5
Fig. 5 Absorption spectra as a function of doping level: (a) and (b) for different μc of graphene, (c) and (d) for different ns of BP, under TE and TM incidences, respectively, where w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm.
Fig. 6
Fig. 6 (a) and (b) are absorption spectra as a function of incident angles θ for TE polarization and TM polarization respectively. (c) is absorption spectra for various polarization angles β under normal incidence. The parameters are w1 = 135 nm, w2 = 115 nm, w3 = 125 nm, p = 250 nm, t1 = 5 nm, t2 = 100 nm and d = 1.65 μm.

Tables (1)

Tables Icon

Table 1 Comparisons between plasmonic absorbers at infrared frequencies.

Equations (9)

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

σ(ω, μ c ,Г,T)= σ intra + σ inter
σ intra = j e 2 π 2 (ωj2Г) 0 ξ( f d (ξ, μ c ,T) ξ f d (ξ, μ c ,T) ξ ) dξ
σ inter = j e 2 (ωj2Г) π 2 0 f d (ξ, μ c ,T) f d (ξ, μ c ,T) (ωj2Г) 2 4 (ξ/) 2 dξ
f d ( ξ, μ c ,T )= ( e (ξ μ c )/ k B T +1) 1
σ jj = i D j π(ω+ iη )
D j = π e 2 n s m j
m cx = 2 2 γ 2 Δ + η c
m cy = 2 2 ν c
A(λ)=2π c λ ε '' V | E l | 2 dV
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.