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

Enhanced spontaneous emission of quantum emitter in monolayer and double layer black phosphorus

Open Access Open Access

Abstract

We theoretically demonstrate that the spontaneous emission rate (SER) of quantum emitter is greatly enhanced in the proximity of single or double layers of black phosphorus (BP). The maximum enhancement factor is on the order of 106, about 3 times larger than that of graphene, due to the large photonic density of states near BP. We also study how the transition frequency, the BP doping level, the distance between emitter and BP, the existence of substrate underneath BP, and the orientation of transition dipole moment influence the SER. With properly designed parameters, both high SER of emitter and low propagation losses of BP surface plasmon polaritons can be achieved in the quantum system, making it a promising candidate for mid-infrared single-photon generation.

© 2017 Optical Society of America

1. Introduction

Since graphene was discovered as the first two-dimensional (2D) material in 2004, the variety of 2D materials has been increasing rapidly in the past decade. Great research efforts have been devoted to few-layer graphene [1,2], transition metal dichalcogenides (TMDCs) such as MoS2 and WS2 [3,4], hexagonal boron nitride (hBN) [5,6], etc. However, the existing 2D materials are hindered from high-performance optoelectronic and nanophotonic applications due to this mobility/bandgap trade-off [7]. For example, graphene holds extremely high mobilities but has a vanishing bandgap. On the contrary, TMDCs have large direct bandgaps, although their mobilities are usually orders of magnitude lower than those of graphene. Recently, black phosphorus (BP) was reintroduced into the community from the perspective of 2D materials, showing a direct bandgap at the Г point [8], excellent carrier mobility [9], strong light-matter interactions in the mid-infrared (mid-IR) range [10], and unique anisotropic physical properties [11]. These features make it possible to bridge gaps in the 2D material family, complementing important applications in mid-IR optoelectronics and photonics [12,13].

One of the most exciting merits of 2D materials is that they support highly confined surface plasmon polaritons (SPPs), which enable extremely strong light-matter interactions with quantum emitters. For example, the Purcell factors of dipole emitters in the vicinity of graphene can reach 105-107 due to tightly confined, low-loss graphene SPPs [14,15]. Dramatic modification of relaxation rate is also observed in excited erbium ions placed in subwavelength proximity of graphene [16]. Nanostructured graphene can support SPPs with hyperbolic topologies at THz and far-IR frequencies, which gives rise to stronger light-matter interactions with the Purcell factors being up to 107-1010 [17,18]. Furthermore, a novel theory points out that 2D plasmonics, with its unprecedented level of confinement, presents a unique opportunity to access radiative transitions in atomic-scale emitters that were considered forbidden [19]. Multiplasmon processes, spin-flip radiation, and very high-order multipolar transitions are shown to be competitive with the fastest atomic dipole transitions. Inspired by the intriguing properties of graphene, BP plasmonics promises strong light-matter interactions with quantum emitters as well. A recent study reported that spontaneous emission rate (SER) of emitter is greatly enhanced in BP thin films, comparable to that of graphene, while providing even higher upper limits [10]. However, the high SER comes at the cost of large SPP propagation losses, i. e., most of the emitted photons would be dissipated before being collected or utilized. For example, despite the fact that SER is higher in the hyperbolic regime, the propagation lengths of BP SPPs are hundreds of times smaller than the wavelength in free space, making it inappropriate for applications such as mid-IR photon sources. Comparing to the anisotropic elliptic regime, the SER is higher in the hyperbolic regime only when the emitter is several nanometers away from the BP (see Fig. 8 of Ref [10].). It drops dramatically with farther distance such as 10 nm. Since the non-radiative relaxation, e. g., resonance energy transfer from emitter to BP, boosts at nanoscale distances and is detrimental to the high quantum efficiency of photon sources, we restrict ourselves to separations larger than 10 nm in this work. Obviously, the anisotropic elliptic regime is more suitable than the hyperbolic regime for highly efficient photon sources in terms of SER, SPP propagation losses, and quantum efficiency.

In this paper, we systematically investigate strongly enhanced light-matter interactions in the quantum emitter-monolayer BP coupling system working at 10-30 THz. Since the bandgap of monolayer BP is ~2 eV [20,21], interband transitions in BP which are the primary origin of SPP losses are totally forbidden for the photon frequencies of concern [22]. We also explore the relationships between the SER and the BP doping level, the emitter-BP distance, the permittivity of substrate, and the orientation of transition dipole matrix element. In addition, it is verified that the emitter embedded in double layer BP exhibits even better radiation field confinement and thus higher SER. Therefore, with a judicious design of the emitter-BP coupling system, both high SER and low SPP losses can be achieved at the same time. Different from previous works in the literature studying the basic physics of BP plasmonics, we focus on the specific light-matter interactions in BP and how they can help to build single-photon sources with both record-high Purcell factor and near-unity β factor. Our findings pave the way to developing BP plasmonic platforms for new highly efficient, ultra-compact mid-IR photon sources which could find applications in single-plasmon devices and quantum plasmonics.

2. Device design and analysis

The quantum systems of interest are shown in Fig. 1. A quantum emitter such as a quantum dot or a rare-earth ion is suspended in air above a freestanding monolayer BP or between two layers of freestanding monolayer BP. The three-dimensional schematic diagrams of the emitter coupled to monolayer and double layer BP are illustrated in Figs. 1(a) and 1(c), respectively, while Figs. 1(b) and 1(d) show their cross sections in XOZ plane. The emitter is placed at a distance of d from each monolayer BP. Owing to its deep-subwavelength size, it can be modeled as an electric dipole which corresponds to the transition dipole matrix element of quantum emitter. The monolayer BP can be isolated from bulk BP by means of mechanical exfoliation and plasma thinning [23]. It has an orthorhombic lattice with two clearly distinguishable axes, as depicted in the inset of Fig. 1(a). Here we assume that the armchair and zigzag crystallographic directions of monolayer BP are oriented along x and y axis, respectively. The thickness of monolayer BP is about half of the lattice constant in the out-of-plane direction, i. e., 5.35 Å [24].

 figure: Fig. 1

Fig. 1 Three-dimensional view of a quantum emitter coupled to (a) monolayer and (c) double layer BP. Cross sections in XOZ plane of the systems with (b) monolayer and (d) double layer BP. The distance between emitter and each monolayer BP is denoted by d. Lattice structure and orientations of two crystal axes of monolayer BP are illustrated in the inset.

Download Full Size | PDF

For the mid-IR frequencies considered here, the conductivity of monolayer BP is dominated by intraband transitions and can be approximated by the Drude model [21, 22],

σjj=ine2/mjω+iη/,
where j = {x, y}, n is the electron density in the uniformly n-doped BP, -e is the charge of an electron, mj is the in-plane electron effective mass close to the Г point, ω is radian frequency, η is the scattering rate of electron, andis the reduced Planck’s constant. Note that here we assume a local conductivity, i. e., the conductivity is independent of the in-plane wave vector q. The assumption is valid as long as|q|<c/vfk0, where vf is the Fermi velocity in BP [25, 26]. Hereinafter, we mainly focus on the anisotropic elliptic regime where only the relatively small wave vectors are involved and therefore the nonlocality has a negligible effect on the SER of quantum emitter located nearby. In another word, the numerical results derived using Eq. (1) are accurate enough for our investigations. Figures 2(a)-2(d) explicitly show the frequency (10-30 THz) and electron density (1013-1014 cm−2) dependence of BP conductivity. The carrier concentration can be controlled either electrically by a gate or chemically by introducing donor or acceptor impurity atoms during the synthesis [21]. The parameters used in the calculations are as follows [10, 21]: mx = 0.18m0, my = 1.23m0, η = 3 meV, where m0 is the electron rest mass. As a consequence of orthorhombic crystal structure, monolayer BP exhibits high anisotropy of in-plane electron effective mass and conductivity. Moreover, sinceIm(σxx),Im(σyy)>0, only anisotropic elliptic quasi-transverse magnetic (TM) SPPs can be supported by BP in this Drude regime [10]. Under the quasi-static approximation, i. e.,q>>k0, the dispersion of quasi-TM SPPs is given by [22]
σxxcos2θ+σyysin2θ=2iε0ωq,
where q = |q| is the magnitude of the in-plane wave vector of plasmon pointing at an angle θ with respect to x, and k0 = ω/c is the wave vector of light in vacuum. Figures 2(e)-2(h) present the SPP wave vectors along two crystallographic axes as a function of frequency and electron density. The degree of anisotropyRe(qy)/Re(qx)=Im(qy)/Im(qx)=my/mx=6.83demonstrates its origin from the anisotropic in-plane electron effective mass. The real parts of qx and qy increase monotonically with higher frequency and/or lower electron density, providing a variety of high-k photonic states. Meanwhile, the imaginary parts which represent the loss of BP SPPs are independent of frequency and inversely proportional to electron density. So in addition to forbidding interband transitions by using monolayer BP, we can further reduce the losses from intraband transitions by inducing moderate doping. It is also worth pointing out that comparing to the BP thin film studied in Ref [10]. the losses of the anisotropic elliptic surface plasmons that are attributed to the intraband transitions are smaller in monolayer BP, leading to about 10 times longer propagation length in the proposed structures.

 figure: Fig. 2

Fig. 2 Optical conductivity of monolayer BP and in-plane wave vector of quasi-TM SPP supported by monolayer BP as a function of frequency and electron density. Real part of conductivity (a) σxx and (b) σyy and imaginary part of conductivity (c) σxx and (d) σyy. Real part of wave vector (e) qx and (f) qy and imaginary part of wave vector (g) qx and (h) qy. The parameters used in the calculations are mx = 0.18m0, my = 1.23m0, η = 3 meV.

Download Full Size | PDF

The magnitude of light-matter interactions can be quantified by the SER of nearby emitter, which is calculated by [27]

γγ0=npIm[G(r0,r0,ω)]npnpIm[G0(r0,r0,ω)]np,
where γis the SER of emitter located near BP, γ0is the one in free space, np is the unit orientation vector of transition dipole moment p, G(r0,r0,ω) is the dyadic Green function of the system with BP, G0(r0,r0,ω) is the dyadic Green function in free space, and r0 and ω are the location and oscillation frequency of emitter. The expression is also equivalent toγ/γ0=P/P0, where P and P0 are the power radiated by emitter in the proximity of BP and in free space, respectively [27]. In this work, we use commercial software Lumerical FDTD Solutions which is based on the finite-difference time domain (FDTD) techniques to simulate the quantum system, calculate the enhanced SER, and observe the photon emission as BP SPPs. The monolayer BP is approximated by a very thin layer of finite thickness with an effective permittivity which is related to the optical conductivity byεjj=1+iσjj/ε0ωΔ, whereΔis the thickness of monolayer BP. The total power radiated by emitter is evaluated by integrating outgoing power flow over the boundary of a closed volume which contains the emitter [28]. In this way, we derive the SER of quantum emitter in the proposed systems and conduct a comprehensive study of how the photon frequency, the doping level of BP, the emitter-BP distance, the existence of different substrates under BP, and the orientation of transition dipole moment may affect it.

3. Results and discussions

First, we consider the scenario in Figs. 1(a) and 1(b), with a z-oriented transition dipole moment of quantum emitter, i. e., p = pez. Figures 3(a) and 3(b) depict the Ez component in the monolayer BP at 10 and 15 THz, respectively. Here we assume that the emitter is located at a distance d = 10 nm above the monolayer BP, and the doping level of BP is n = 5 × 1013 cm−2. As a result of anisotropic in-plane conductivity, SPP propagation shows a significant canalization along x axis. The plasmon at higher frequency has shorter guided wavelength and propagation length, in agreement with the frequency dependence of in-plane wave vector q shown in Figs. 2(e)-2(f). The magnitude of electric field in XOZ plane is plotted in Fig. 3(c). Obviously, instead of going freely into space, the power generated by emitter is coupled to the monolayer BP and propagates along it as SPPs. Following the methods described in Ref [29], we calculate the percentage of power entering free space, coupling to SPPs, and dissipated in BP, respectively. We find that 6.82% of the total power is radiated into free space while 17.61% is dissipated after propagating along the BP for 1 μm. Only 75.57% is finally emitted in the form of SPPs. That is the reason why controlling SPP losses is of great importance in realizing high-yield photon sources based on BP. The SERs of quantum emitter close to monolayer BP and graphene are compared in Fig. 3(d). To make a fair comparison, the density and scattering rate of electrons in graphene are also n = 5 × 1013 and η = 3 meV, respectively, which correspond to the chemical potentialμc=vfπn=0.825eV and the electron mobilityμ=evf2/ημc=2660cm2/(V·s), wherevf106m/s is the Fermi velocity in graphene [2]. The SER of emitter with monolayer BP is about 3 times larger than that with graphene, which results from the higher photonic density of high-k states near BP. According to Eq. (2), monolayer BP supports evanescent waves with q ranging from qx to qy, while graphene only supports those with fixedq=2iε0ω/σg, where σg is the conductivity of graphene. Therefore, the spontaneous emission enhancement is stronger with monolayer BP due to additional coupling of the emitted evanescent waves to a wide variety of high-k SPP states. The mechanism can also explain the frequency dependence of SER. The coupling efficiency is proportional to the spatial overlap of the mode functions of the evanescent waves emitted by dipole and the SPPs supported by BP, henceeQd, where Qq is the out-of-plane decay coefficient in the z direction [16]. As pointed out in Figs. 2(e) and 2(f), the real part of q increases with frequency, which leads to the decrease in coupling efficiency, and consequently the drop in SER. However, for longer distances and higher frequencies graphene outperforms monolayer BP in the SER enhancement because graphene support less confined evanescent waves [18]. In the inset of Fig. 3(d), we show the SERs of emitter in the vicinity of monolayer BP and graphene at 25-30 THz with d = 100 nm. The SER with graphene is higher in this region and decays slower with frequency than that with BP, which is the feature of less confined SPPs sustained by graphene.

 figure: Fig. 3

Fig. 3 Ez component in monolayer BP at (a) 10 and (b) 15 THz. (c) Magnitude of electric field in XOZ plane at 10 THz. (d) SERs of quantum emitter near monolayer BP (red solid line) and graphene (blue solid line) versus frequency. The emitter-BP distance is d = 10 nm (100 nm in the inset), the electron density is n = 5 × 1013 cm−2, and the transition dipole moment is oriented in the z direction. The graphene has the same electron density and scattering rate as BP.

Download Full Size | PDF

Doping provides a degree of freedom to tune the optical properties of monolayer BP, adjust the dispersion relationship of BP SPPs, and in turn modify the SER of nearby emitter. Figure 4(a) illustrates the dependence of SER and SPP propagation length on the electron density in BP, with distance d = 10 nm and frequency f = 10 THz. The propagation length is calculated byLj=1/2Im(qj) [30]. Obviously, the SER is larger at a low doping level, which is attributed to the richness of high-k states. The high-k states include BP SPPs with q ranging from qx to qy. This range of q widens when decreasing the electron density (cf. Figures 2(e) and 2(f)), hence giving rise to more decay channels and higher SER of emitter in the vicinity of BP. However, BP SPPs suffer significant losses at a low doping level (cf. Figures 2(g) and 2(h)), which means the emitted photons have a high probability to be lost in the near field rather than be extracted or manipulated in the far field. To achieve a balanced trade-off between SER and propagation length (see Fig. 4(a)), a moderate doping level of BP is desirable for high-yield photon sources. Hereinafter, we assume that the electron density in monolayer BP is maintained at n = 5 × 1013 via appropriate doping.

 figure: Fig. 4

Fig. 4 (a) Electron density dependence of the SER of nearby emitter (red solid line), the propagation length along x, Lx (green dashed line), and the propagation length along y, Ly (blue dashed line) at f = 10 THz. (b) SER of emitter versus its distance to the monolayer BP. The frequency is chosen to be f = 10 (red solid line) and 50 THz (blue solid line). (c) Influence of substrates with εs = 2.25 (red solid line), 4 (blue solid line), and 6.25 (green solid line) on the SER of nearby emitter at f = 10 THz. Other parameters are the same as in Fig. 3.

Download Full Size | PDF

Then we investigate the SER of a z-oriented dipole versus its distance to the monolayer BP at a frequency of f = 10 THz. The distance dependence is shown in Fig. 4(b) where the SER decreases monotonically with growing gap. This is a clear signature of the aforementioned mechanism that the photons emitted by dipole are evanescently coupled to the monolayer BP as the coupling efficiency drops exponentially with increasing emitter-BP distance. To compare with Fig. 8 of Ref [10], we also plot the distance dependence of SER at f = 50 THz in Fig. 4(b). The SERs at d = 10 and 20 nm are 5.3 and 4.6, respectively, in reasonable agreement with the previously reported values [10]. Although the SER is supposed to be larger at even shorter distances, non-radiative relaxation such as resonance energy transfer from emitter to BP may prevail over plasmon-emitter coupling, considering that it scales as d−4 [16]. Thus, we do not study the system with the distance d smaller than 10 nm in this work.

In practice, monolayer BP always sits on a substrate, so it is necessary to consider the effect of substrate on the SER. In Fig. 4(c), we show the frequency dependence of SER with the permittivity of substrate εs = 2.25, 4, and 6.25. The distance is assumed to be d = 10 nm and the frequency is fixed at f = 10 THz. The SER increases with εs because higher permittivity leads to more confined SPPs and therefore stronger light-matter interactions. We also observe that the growth of SER at higher frequencies is more obvious than at lower frequencies, which is attributed to the fact that the in-plane wave vectors of SPPs at higher frequencies are more sensitive to the permittivity change of surrounding environment.

We proceed to study how the SER changes when the transition dipole moment is oriented in the x or y direction, i. e., p = pex or p = pey. Assuming d = 10 nm and f = 10 THz, Figs. 5(a) and 5(b) present the Ez component in the monolayer BP with an x- and y-polarized dipole, respectively. Apparently, for x (y) polarization the emitter is inclined to decay into BP SPPs propagating along x (y) axis. The SERs for x and y polarizations versus frequency are plotted in Fig. 5(c). As expected, they are smaller than that of z-polarized dipole which can radiate BP SPPs in all directions and, therefore, has a higher partial local density of states [27]. We also find that in spite of the canalization observed in the x direction, it is actually the SPPs propagating along y that make greater contributions to the spontaneous emission enhancement, given that the SER of y-polarized dipole is larger than that of x-polarized dipole. As a result, the importance of suppressing propagation losses is more pronounced because BP SPPs propagating along y are more vulnerable to dissipation (cf. Figure 4(a)).

 figure: Fig. 5

Fig. 5 Ez component in monolayer BP with (a) x- and (b) y-polarized dipole at 10 THz. (c) SERs of x- (blue solid line) and y-polarized (red solid line) dipole near monolayer BP versus frequency. Other parameters are the same as in Fig. 3.

Download Full Size | PDF

Next, we consider the scenario in Figs. 1(c) and 1(d), i. e., the emitter is coupled to two layers of monolayer BP. Figures 6(a) and 6(b) show the Ez component in individual BP layers and the magnitude of electric field in XOZ plane at 10 THz, respectively. The distance to each layer is d = 10 nm, the electron density in BP is n = 5 × 1013 cm−2, and the transition dipole moment is oriented in the z direction. Similar to the phenomenon observed in monolayer BP, the SPPs of double layer BP exhibit canalization along x axis, arising from the anisotropy of optical conductivity inherited from monolayer BP. In comparison with Fig. 3(c), the radiation field is now localized between two BP layers and thus more confined, resulting in stronger light-matter interactions between emitter and BP SPPs. The amount of power that leaks into free space is 1.32% of the total power radiated by emitter. The power lost in BP after propagation for 1 μm is 13.62% while the remaining 85.06% exists as surface plasmons. As most of the modal energy is concentrated in the air between two BP layers, the SPP propagation losses in double layer configuration are smaller than in single layer, making double layer BP more suitable for applications such as photon sources. The inset of Fig. 6(c) illustrates the Ez field profile of the symmetric (Es) and anti-symmetric (Ea) SPP modes supported by the double layer structure [31]. Since the transition dipole moment is z-oriented, only the symmetric mode can be excited in the double layer configuration. It is demonstrated by the mode profile monitored between two BP layers, as shown in Fig. 6(c). So the energy of quantum emitter is estimated to be radiated as symmetric mode SPPs at a rate of about 6.75 × 106γ0. It is then proved in Fig. 6(d) where the SERs of emitter between double layer BP and graphene are plotted as a function of frequency. The parameters of graphene are the same as in Fig. 3(d). As a matter of fact, the SER of emitter between two BP layers is more than 4 times larger than that near single BP layer, whereas the ratio between two SERs derived from double layer BP and graphene is weakly dependent on frequency and hovers around 3. We also depict the SERs of quantum emitter between double layer BP and graphene at 25-30 THz with d = 175 nm in the inset of Fig. 6(d). Since the radiation field is more confined in the double layer structure, the SER with graphene exceeds that with BP at a longer distance and a higher frequency, comparing to the situation in the single layer configuration (see the inset of Fig. 3(d)). The dependence of SER on the BP doping level, the distance between emitter and BP, the permittivity of substrate, and the orientation of transition dipole moment of emitter is quite similar in the cases of monolayer and double layer BP, so we do not repeat those discussions in the context of two BP layers in this paper.

 figure: Fig. 6

Fig. 6 (a) Ez component in each BP layer and (b) magnitude of electric field in XOZ plane at 10 THz. (c) Ez field profile of the SPP mode excited in the double layer BP. Symmetric (Es) and anti-symmetric (Ea) SPP modes supported by double layers are illustrated in the inset. (d) SERs of quantum emitter between two layers of BP (red solid line) and graphene (blue solid line) versus frequency. The emitter-BP distance is d = 10 nm (175 nm in the inset). Other parameters are the same as in Fig. 3.

Download Full Size | PDF

4. Conclusion

In summary, we have studied the enhanced spontaneous emission of quantum emitter in monolayer and double layer BP. Numerical results prove that the SER is multiplied by a factor of about 106, even higher than that of graphene, which is credited to the large density of high-k states supported by BP. We have also unveiled the relationships between the SER and the photon frequency, the BP doping level, the distance between emitter and BP, the permittivity of substrate, the transition dipole orientation, etc. By judiciously designing the system parameters, we have achieved an extremely efficient SPP source with relatively low propagation losses in BP. The proposed device can directly serve as a single-plasmon source or be coupled to a dielectric waveguide via a grating [32], showing promising potential for novel single-plasmon devices and future quantum photonic applications.

Funding

National Natural Science Foundation of China (NSFC) (61671306); Fundamental Research Project of Shenzhen (JCYJ) (20160328145357990); Project funded by China Postdoctoral Science Foundation (2017M612734).

References and links

1. A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nat. Mater. 6(3), 183–191 (2007). [CrossRef]   [PubMed]  

2. 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(1), 109–162 (2009). [CrossRef]  

3. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7(11), 699–712 (2012). [CrossRef]   [PubMed]  

4. X. Liu, T. Galfsky, Z. Sun, F. Xia, E. C. Lin, Y. H. Lee, S. Kéna-Cohen, and V. M. Menon, “Strong light-matter coupling in two-dimensional atomic crystals,” Nat. Photonics 9(1), 30–34 (2014). [CrossRef]  

5. R. Mas-Ballesté, C. Gómez-Navarro, J. Gómez-Herrero, and F. Zamora, “2D materials: to graphene and beyond,” Nanoscale 3(1), 20–30 (2011). [CrossRef]   [PubMed]  

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

7. T. Georgiou, R. Jalil, B. D. Belle, L. Britnell, R. V. Gorbachev, S. V. Morozov, Y. J. Kim, A. Gholinia, S. J. Haigh, O. Makarovsky, L. Eaves, L. A. Ponomarenko, A. K. Geim, K. S. Novoselov, and A. Mishchenko, “Vertical field-effect transistor based on graphene-WS2 heterostructures for flexible and transparent electronics,” Nat. Nanotechnol. 8(2), 100–103 (2012). [CrossRef]   [PubMed]  

8. A. S. Rodin, A. Carvalho, and A. H. Castro Neto, “Strain-induced gap modification in black phosphorus,” Phys. Rev. Lett. 112(17), 176801 (2014). [CrossRef]   [PubMed]  

9. 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]  

10. D. Correas-Serrano, J. S. Gomez-Diaz, A. A. Melcon, and A. Alù, “Black phosphorus plasmonics: anisotropic elliptical propagation and nonlocality-induced canalization,” J. Opt. 18(10), 104006 (2016). [CrossRef]  

11. F. Xia, H. Wang, and Y. Jia, “Rediscovering black phosphorus as an anisotropic layered material for optoelectronics and electronics,” Nat. Commun. 5, 4458 (2014). [CrossRef]   [PubMed]  

12. X. Ling, H. Wang, S. Huang, F. Xia, and M. S. Dresselhaus, “The renaissance of black phosphorus,” Proc. Natl. Acad. Sci. U.S.A. 112(15), 4523–4530 (2015). [CrossRef]   [PubMed]  

13. A. Castellanos-Gomez, “Black phosphorus: narrow gap, wide applications,” J. Phys. Chem. Lett. 6(21), 4280–4291 (2015). [CrossRef]   [PubMed]  

14. 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]  

15. A. Yu. Nikitin, F. Guinea, F. J. García-Vidal, and L. Martín-Moreno, “Fields radiated by a nanoemitter in a graphene sheet,” Phys. Rev. B 84(19), 195446 (2011). [CrossRef]  

16. K. J. Tielrooij, L. Orona, A. Ferrier, M. Badioli, G. Navickaite, S. Coop, S. Nanot, B. Kalinic, T. Cesca, L. Gaudreau, Q. Ma, A. Centeno, A. Pesquera, A. Zurutuza, H. de Riedmatten, P. Goldner, F. J. García de Abajo, P. Jarillo-Herrero, and F. H. L. Koppens, “Electrical control of optical emitter relaxation pathways enabled by graphene,” Nat. Phys. 11(3), 281–287 (2015). [CrossRef]  

17. J. S. Gomez-Diaz, M. Tymchenko, and A. Alù, “Hyperbolic plasmons and topological transitions over uniaxial metasurfaces,” Phys. Rev. Lett. 114(23), 233901 (2015). [CrossRef]   [PubMed]  

18. J. S. Gomez-Diaz, M. Tymchenko, and A. Alù, “Hyperbolic metasurfaces: surface plasmons, light-matter interactions, and physical implementation using graphene strips,” Opt. Mater. Express 5(10), 2313–2329 (2015). [CrossRef]  

19. N. Rivera, I. Kaminer, B. Zhen, J. D. Joannopoulos, and M. Soljačić, “Shrinking light to allow forbidden transitions on the atomic scale,” Science 353(6296), 263–269 (2016). [CrossRef]   [PubMed]  

20. V. Tran, R. Soklaski, Y. Liang, and L. Yang, “Layer-controlled band gap and anisotropic excitons in few-layer black phosphorus,” Phys. Rev. B 89(23), 235319 (2014). [CrossRef]  

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

22. 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]  

23. J. Lu, J. Wu, A. Carvalho, A. Ziletti, H. Liu, J. Tan, Y. Chen, A. H. Castro Neto, B. Özyilmaz, and C. H. Sow, “Bandgap engineering of phosphorene by laser oxidation toward functional 2D materials,” ACS Nano 9(10), 10411–10421 (2015). [CrossRef]   [PubMed]  

24. X. Wang and S. Lan, “Optical properties of black phosphorus,” Adv. Opt. Photonics 8(4), 618–655 (2016). [CrossRef]  

25. W. Yan, M. Wubs, and N. A. Mortensen, “Hyperbolic metamaterials: nonlocal response regularizes broadband supersingularity,” Phys. Rev. B 86(20), 205429 (2012). [CrossRef]  

26. D. Correas-Serrano, J. S. Gomez-Diaz, M. Tymchenko, and A. Alù, “Nonlocal response of hyperbolic metasurfaces,” Opt. Express 23(23), 29434–29448 (2015). [CrossRef]   [PubMed]  

27. L. Novotny and B. Hecht, Principles of Nano-Optics, 2nd ed. (Cambridge University, 2012).

28. https://kb.lumerical.com/en/diffractive_optics_cavity_purcell_factor.html.

29. H. Lian, Y. Gu, J. Ren, F. Zhang, L. Wang, and Q. Gong, “Efficient single photon emission and collection based on excitation of gap surface plasmons,” Phys. Rev. Lett. 114(19), 193002 (2015). [CrossRef]   [PubMed]  

30. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).

31. B. Wang, X. Zhang, X. Yuan, and J. Teng, “Optical coupling of surface plasmons between graphene sheets,” Appl. Phys. Lett. 100(13), 131111 (2012). [CrossRef]  

32. M. Gullans, D. E. Chang, F. H. L. Koppens, F. J. García de Abajo, and M. D. Lukin, “Single-photon nonlinear optics with graphene plasmons,” Phys. Rev. Lett. 111(24), 247401 (2013). [CrossRef]   [PubMed]  

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Three-dimensional view of a quantum emitter coupled to (a) monolayer and (c) double layer BP. Cross sections in XOZ plane of the systems with (b) monolayer and (d) double layer BP. The distance between emitter and each monolayer BP is denoted by d. Lattice structure and orientations of two crystal axes of monolayer BP are illustrated in the inset.
Fig. 2
Fig. 2 Optical conductivity of monolayer BP and in-plane wave vector of quasi-TM SPP supported by monolayer BP as a function of frequency and electron density. Real part of conductivity (a) σxx and (b) σyy and imaginary part of conductivity (c) σxx and (d) σyy. Real part of wave vector (e) qx and (f) qy and imaginary part of wave vector (g) qx and (h) qy. The parameters used in the calculations are mx = 0.18m0, my = 1.23m0, η = 3 meV.
Fig. 3
Fig. 3 Ez component in monolayer BP at (a) 10 and (b) 15 THz. (c) Magnitude of electric field in XOZ plane at 10 THz. (d) SERs of quantum emitter near monolayer BP (red solid line) and graphene (blue solid line) versus frequency. The emitter-BP distance is d = 10 nm (100 nm in the inset), the electron density is n = 5 × 1013 cm−2, and the transition dipole moment is oriented in the z direction. The graphene has the same electron density and scattering rate as BP.
Fig. 4
Fig. 4 (a) Electron density dependence of the SER of nearby emitter (red solid line), the propagation length along x, Lx (green dashed line), and the propagation length along y, Ly (blue dashed line) at f = 10 THz. (b) SER of emitter versus its distance to the monolayer BP. The frequency is chosen to be f = 10 (red solid line) and 50 THz (blue solid line). (c) Influence of substrates with εs = 2.25 (red solid line), 4 (blue solid line), and 6.25 (green solid line) on the SER of nearby emitter at f = 10 THz. Other parameters are the same as in Fig. 3.
Fig. 5
Fig. 5 Ez component in monolayer BP with (a) x- and (b) y-polarized dipole at 10 THz. (c) SERs of x- (blue solid line) and y-polarized (red solid line) dipole near monolayer BP versus frequency. Other parameters are the same as in Fig. 3.
Fig. 6
Fig. 6 (a) Ez component in each BP layer and (b) magnitude of electric field in XOZ plane at 10 THz. (c) Ez field profile of the SPP mode excited in the double layer BP. Symmetric (Es) and anti-symmetric (Ea) SPP modes supported by double layers are illustrated in the inset. (d) SERs of quantum emitter between two layers of BP (red solid line) and graphene (blue solid line) versus frequency. The emitter-BP distance is d = 10 nm (175 nm in the inset). Other parameters are the same as in Fig. 3.

Equations (3)

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

σ jj = in e 2 / m j ω+iη/ ,
σ xx cos 2 θ+ σ yy sin 2 θ= 2i ε 0 ω q ,
γ γ 0 = n p Im[G( r 0 , r 0 ,ω)] n p n p Im[ G 0 ( r 0 , r 0 ,ω)] n p ,
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.