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

Switching freely between superluminal and subluminal light propagation in a monolayer MoS2 nanoresonator

Open Access Open Access

Abstract

We theoretically propose a feasible scheme to advance or slow the propagation of light in a monolayer MoS2 nanoresonator (NR). The scheme allows one to easily turn on or off the fast (superluminal) and slow (subluminal) light effects and switch freely between fast and slow light propagation by only adjusting the frequency or intensity of the pump field. As the exciton interacts strongly with the phonons in MoS2, the slow light effect will appear along with a large dispersion with a very steep negative slope and a sharp absorption peak. Especially, the maximal group velocity index of the slow light in the monolayer MoS2 NR can reach two orders of magnitude larger than that in a carbon nanotube resonator. These results provide a new way to measure the exciton-phonon coupling strength and may prove useful in device applications such as optical switching and optical signal processing.

© 2017 Optical Society of America

Corrections

27 June 2017: A typographical correction was made to the author affiliations.

1. Introduction

In recent decades, a great deal of interest has been focused on how to advance or slow the propagation of light in the atomic vapors [1], solid-state materials [2], and nanomechanical oscillator systems [3] due to their potential applications in electronics and optoelectronics [4]. Fast and slow light effects can be observed by using coherent population oscillations [2,5], electromagnetically induced transparency [1,6], and stimulated Brillouin scattering [7]. Recently, Zhu et al. have proposed a tunable slow and fast light device based on a carbon nanotube resonator [8]. Javed Akram et al. have reported tunable fast and slow light effects of a transmitted probe field in a hybrid optomechanical system consisting of a high Q Fabry-Perot cavity, a mechanical resonator and a two-level atom, and proposed a tunable switch from superluminal to slow light is achievable in their model by simply adjusting the atomic detuning [9]. Then they have also studied Fano resonances and slow light in a nanocavity using Bose-Einstein condensate. The results showed that the slow light effect can be enhanced by the strength of the atom-atom interaction and its robustness against the condensate fluctuations using presently available technology [10]. Slow and fast light effects can be switched in the microwave regime using a coupled nanomechanical resonator (NR)-Cooper-pair box (CPB) system [11]. Very recently, Dumeige’s group has demonstrated that it is feasible to enhance the photon lifetime by several orders of magnitude by using slow-light effects in an active whispering-gallery-mode (WGM) microresonator [12]. Based on slow light in cavity optomechanics, an ultrasensitive mass sensing method has also been presented [13].

Compared with one-dimensional carbon nanotube with small bandgap, molybdenum disulfide (MoS2) is an attractive semiconductor whose electronic structure depends on its layer number [14–20], which further promotes and extends the applications of MoS2. As a new member of two-dimensional materials, layered MoS2 with a large bandgap and surface-to-volume ratio possesses many excellent properties such as ultralow weight [15], exceptional strain limit [16], high elastic modulus [17], etc. Currently, significant research interest is devoted towards practical applications of such materials for transistors [18], photodetectors [20], small-signal amplifier [19], etc. Especially, nonlinear optical behaviors in MoS2 nanosheets have been extensively studied [21–25]. Degenerate two-photon absorption [22], ultrafast saturable absorption [23], and harmonic effects [24] in MoS2 nanosheets were reported one after another. Recent investigation has also shown that one can observe the dynamic self-diffraction in MoS2 nanoflake solutions with laser [25]. Despite these significant impacts and immense applications, though, very few reports have been published on fast and slow light effects in MoS2-based materials. Our main objective in this paper is to pave a new way to advance and slow light propagation using a suspended monolayer MoS2 NR and open a new horizon for switching freely between fast (superluminal) and slow (subluminal) light propagation. The dependencies of the group velocity of the probe light on the excitation frequency, pumping intensity and exciton-phonon coupling strength will be studied and the absorption and dispersion spectra strongly correlated to fast and slow light effects will be discussed.

2. Model and formalism

We consider an optomechanical system based on monolayer MoS2 suspended on a Si/SiO2 substrate [26]. As schematically represented in Fig. 1(a), the system is driven by a strong pump field accompanying a weak probe field. Epu (Epr) is the slowly varying envelope of the pump (probe) field, and ωpu (ωpr) is the frequency of the pump (probe) field. In such structure, the lowest-energy resonance corresponds to the fundamental in-plane flexural mode with the frequency ωn. The MoS2 NR is assumed to be characterized by a sufficiently high quality factor Q, so the lifetime of the resonator is long enough. In this case, the new mechanical resonances caused by the edge effects and irregular shapes can be neglected as compared to the fundamental flexural mode [27]. Similar with a carbon nanotube, due to quantum confinement, the MoS2 can act like as a QD when electrons are confined to a small region within a MoS2 nanosheet [28]. The eigenmode of MoS2 can be regarded as a quantum harmonic oscillator with two bosonic operators b+ and b (corresponding to the phonon creation and annihilation operators, respectively). The vibration Hamiltonian of the MoS2 NR can be written as hωnb+b. A localized exciton formed in the MoS2 NR can be modeled as a two-level system consisting of a ground state |0> and single exciton state |1>. Such an exciton can be characterized by three pseudospin operators σ01, σ10 and σz. As the exciton interacts with two applied fields [29, 30], the physical situation is illustrated in Fig. 1(b).

 figure: Fig. 1

Fig. 1 (a) Schematic representation of monolayer MoS2 suspended on a Si/SiO2 substrate by an optical pump-probe technique [31]. The system is driven by a strong pump laser and detected by a weak probe laser. (b) The level scheme of a localized exciton interacting with the phonons in the MoS2 NR.

Download Full Size | PDF

In a rotating frame, the total Hamiltonian of the system in the simultaneous presence of two laser fields can be written as [27]

H=Δpuσz+ωnb+b+gωnσz(b++b)μEpu(σ10+σ01)μ(σ10Epreiδt+σ01Epr*eiδt),
where Δpu = ω10ωpu is the exciton-pump field detuning, and δ = ωprωpu is the probe-pump frequency difference. σij = |i><j| (i, j = 0, 1) is the transition operator between the states |i> and |j>. w = 2<σz> in which σz = σ11σ00 represents the population inversion of the exciton, g denotes the coupling strength between the exciton in the MoS2 nanosheet and the phonons in the NR, and μ refers to the electric dipole moment. For simplicity, we set Ξ = < b+ + b> and p = <σ01>.

By applying the Heisenberg equation of motion, we obtain the corresponding quantum Langevin equations as follows [32]:

p˙=(Γ2+iΔpu)pigωnpΞiΩwiμEprweiδt,
w˙=Γ1(w+1)+2iΩ(p*p)+2iμ(p*EpreiδtpEpr*eiδt),
Ξ¨+γnΞ˙+ωn2Ξ=ωn2gw,
where Γ12) denotes the exciton relaxation rate (dephasing rate), Ω = μEpu/h is the effective Rabi frequency of the pump field, and γn = ωn/Q refers to the decay rate of the NR.

In order to solve Eqs. (2)-(4), we make the assumptions as follows: p = p0 + p1eδt + p–1 eiδt, w = w0 + w1eiδt + w–1eiδt and Ξ = Ξ0 + Ξ1eiδt + Ξ–1eiδt [33], where p0, w0, Ξ0 is the solution of Eqs. (2)-(4) for the case in which only the pump field is present. |p0| >> |p1|,|p–1|; |w0| >> |w1|,|w–1|; |Ξ0| >> |Ξ1|,|Ξ–1|. Upon substituting the above ansatz into Eqs. (2)-(4), we can calculate the linear susceptibility defined as

χ(1)=Nμp1ε0Epr=Nμ2ε0Γ2χ0(1)=Π1χ0(1),
χ0(1)=(C2C1)C3Γ2(1+C1C4C2C4)w0+p0*Γ2(1+C1C4C2C4)Ω,
where Π1 = 2/(ε02). N is the number density of MoS2 nanosheets, and ε0 is the dielectric constant of the surrounding medium. p0*=iΩw0/[Γ2i(Δpug2ωnw0)], C1=(δ+iΓ1)/2Ω,C2=i(g2ωnζp0*+Ω)/[Γ2i(Δpu+δg2ωnw0)],C3=1/(g2ωnζp0+Ω), C4=[Γ2+i(Δpuδg2ωnw0)]/[i(g2ωnζp0+Ω)], ζ=ωn2/(δ2+iγnδωn2).

The exciton-population inversion w0 can be obtained from the following third-order equation

Γ1(w0+1)[Γ22+(Δpugg0w0)2]+4Ω2Γ2w0=0,
In our system, the group velocity of the probe light has the following form:
vg=cn+ωpr(dn/dωpr),
where n ≈1 + 2πχ(1)(ωpr) [34]. Having this, the Eq. (8) will become

cvg1=2πRe(χ(1))ωpr=ω10+2πωprRe(dχ(1)dωpr)ωpr=ω10,

From Eq. (9) we can find that the dispersion is very steeply positive or negative when Re(χ(1))ωpr=ω10=0. Considering this, the effective group velocity index ng can be written as

ng=cvg1=Π2Γ2Re(dχ0(1)dωpr)ωpr=ω10.
where Π2 = 2πωex2/(ε022).

3. Numerical results and discussion

In this scheme, we choose a realistic monolayer MoS2 NR to illustrate the numerical results. All the parameters used here are accessible in experiment. We calculate the absorption and dispersion spectra, and the group velocity for the parameters [27, 35]: ωn = 1GHz, Q = 2000, and Γ2 = 0.25 GHz. In our calculations, we assume Γ1 = 2Γ2.

To study fast and slow light effects in a monolayer MoS2 NR, in Fig. 2(a) we show the variation of the probe absorption spectrum Imχ(1) as the exciton-pump detuning Δpu is changed. For Δpu = ωn, a deep dip appears in the absorption spectrum and the value of Imχ(1) is approximately equal to 0, suggesting the system almost becomes fully transparent to the probe light. This is rather expected, the system will behave like a three-level system in electromagnetically induced transparency [36]. For Δpu = −ωn, a sharp absorption peak occurs at the top of the absorption spectrum. For Δpu = 0, however, the deep dip and sharp peak both vanish and only a weak absorption peak is left. As we all know, when the MoS2 NR is subjected to two applied fields, a beat wave with a variable beat frequency δ = ωprωpu will appear. When the beat frequency δ is close to the vibrational frequency ωn, the MoS2 NR will oscillate coherently, inducing the occurrence of the anti-Stokes (δ = –ωn) and Stokes (δ = ωn) scattering of light via the localized exciton. The corresponding dispersion spectra are shown in Fig. 2(b). As required by the linear Kramers-Kronig relation, the complementarity between the absorption and dispersion spectra is evident.

 figure: Fig. 2

Fig. 2 The probe absorption Imχ(1) (a) and dispersion Reχ(1) (b) spectra (in units of Π1) as a function of the probe-pump detuning δ for different exciton-pump detunings Δpu. (c) The energy levels of the exciton dressed with the phonon modes of the MoS2 NR. Each energy level is split into a doublet with a separation Ω'. TP refers to the three-photon resonance, RL represents the Rayleigh resonance, and AC denotes the ac-Stark resonance.

Download Full Size | PDF

To understand these new features shown in the absorption and dispersion spectra, we depict a dressed-state picture. As the MoS2 NR is excited by a strong pump field, the original energy levels of the exciton (|0>, |1>) will be dressed by the phonon modes in the MoS2 NR. These two energy levels are split into four dressed states |0, n>, |0, n + 1>, |1, n> and |1, n + 1> [27,33], as shown in Fig. 2(c). The sharp peak in Fig. 2(a) is attributed to the three-photon resonance (TP) corresponding to a transition from the state |0, n> to |1, n + 1> by absorbing simultaneously two pump photons and emitting a photon with frequency ωpu – Ω'. The middle peak is assigned to the Rayleigh resonance (RL) corresponding to a transition from |0, n> to |1, n>. The deep dip is induced by the ac-Stark effect corresponding to a usual absorption resonance. Additionally, we note that the dominant role of the resonance mechanisms is strongly corresponded to the excitation frequency. In the following sections, we mainly study the propagation properties of the probe light near the anti-Stokes and Stokes sidebands.

In Figs. 3 we study the way that the effective group velocity index ng changes with the exciton-pump detuning Δpu when Ipu = 3.32 mW/cm2 and 6.64 mW/cm2. In Fig. 3(a), for the case of Ipu = 3.32 mW/cm2, as Δpu increases to the anti-Stokes sideband (Δpu0 = –ωn), ng will reach a maximum value 15831.7, suggesting the occurrence of an obvious slow light effect. Such large ng implies that a monolayer MoS2 NR would be an excellent candidate for slow light devices. As Δpu further increases to the Stokes sideband (Δpu1 = ωn), the peak value of ng is only 240.8. In addition, ng almost keeps invariant in the region between two sidebands. For Ipu = 6.64 mW/cm, however, the scenario becomes completely different. As Δpu increases, ng rises monotonically, reaching a maximum value 549.5, then starts to decrease abruptly, reaching a minimum value –3133.2 with a further increase of Δpu. From Fig. 3(b) we find that it is easy to switch freely between fast and slow light propagation by only adjusting the frequency of the pump field. The results also demonstrate that the fast and slow light effects are highly sensitive to the excitation frequency.

 figure: Fig. 3

Fig. 3 The group velocity index ng (in units of Π2) as a function of the exciton-pump detuning Δpu for Ipu = 3.32 mW/cm2 (a) and Ipu = 6.64 mW/cm2 (b). The insets show the magnification of these remarkable regions around Δpu = δ = ± ωn.

Download Full Size | PDF

To better visualize the Ipu–dependence of fast and slow light effects, in Fig. 4 we plot ng as a function of Ipu with and without the exciton-phonon coupling. As the exciton-phonon coupling is absent (g = 0), the slow light effect never occurs. Here we mainly consider the case of g = 0.08 and Δpu = –ωn. As shown in Fig. 4(a), the switch is turned off when Ipu = 0. With the increase of Ipu, ng first falls monotonously, reaching a minimum –5726 at Ipu = 2.74 mW/cm2. This suggests that the fast light effect emerges. In other word, the fast light is switched on. By further increasing Ipu, ng starts to increase abruptly, reaching a maximum 31536.6 at Ipu = 3.65 mW/cm2. This suggests that the slow light effect appears and the system is switched from fast light to slow light. However, the switch is turned off again at Ipu = 4.48 mW/cm2. Finally, ng declines gradually to a stable value. These results show that the system can act as a two-channel continuous-tunable switch.

 figure: Fig. 4

Fig. 4 The group velocity index ng (in units of Π2) as a function of the pumping intensity Ipu for Δpu = –ωn (a), Δpu = 0 (b), and Δpu = ωn (c) when the coupling between the exciton and the phonons turns on or off.

Download Full Size | PDF

The curves in Fig. 4(b) show that a fast light effect appears. The adjustable range of the fast light in MoS2 NR is a litter larger than that in a carbon nanotube resonator [8]. Also, g modification does not change the nature of the curves, which implies that the exciton-phonon coupling has a weak effect on the fast light effect as the pump field is resonant with the exciton (Δpu = 0). Similar results in Fig. 4(c) have also been obtained in the carbon nanotube resonator, and the achieved slowdown of the group velocity in the MoS2 nanoresonator is a bit higher than in the carbon nanotube resonator [8]. The results in Fig. 4 demonstrate that the fast and slow light effects depend strongly on the exciton-phonon coupling strength g.

To clarify why the group velocity index changes from negative to positive values and put some insight into the nature of the fast and slow light effects, we further explore the impact of the pumping intensity and exciton-phonon coupling strength on the absorption and dispersion spectra. In Fig. 5 we show how the absorption and dispersion spectra change with the pumping intensity when Δpu = −ωn. Figure 5(a) shows the results of calculations of Imχ(1) as a function of the pumping intensity. The magnification of the region around the anti-Stokes sideband is shown in Fig. 5(b). Note that the system supports a critical value of Ic (Ic = 3.65mW/cm2) such that when Ipu < Ic, the magnitude of the absorption peak increases and the width of this peak decreases with an increase of Ipu. When Ipu = Ic, the magnitude reaches a maximum value and the width reaches a minimum value. After this value the absorption spectrum changes suddenly from a sharp peak to a dip. As required by the linear Kramers-Kronig relation, the dispersion spectrum Reχ(1) exhibits a perfect complementary behavior of the absorption spectrum Imχ(1). The corresponding dispersion spectra are shown in Fig. 5(c). Figure 5(d) shows the magnification of the most remarkable region in Fig. 5(c). Please note that the curves in Figs. 5(b) and 5(d) are in correlation. For example, for Ipu = 3.65mW/cm2 the magnitude of the absorption peak has a maximum value and the slope of the dispersion spectrum at Re(χ(1))ωpr=ω10=0 reaches a minimum value. That is, the dispersion spectrum displays a very steep negative slope. By combining Fig. 4(a) with Fig. 5, we can draw a conclusion that a prominent slow light effect arises accompanying with a very sharp absorption peak and a large dispersion with a very steep negative sloop.

 figure: Fig. 5

Fig. 5 The absorption Imχ(1) (a) and dispersion Reχ(1) (c) spectra as a function of the probe-pump detuning δ for a different pumping intensity Ipu. Figures 5(b) and 5(d) show the magnification of the absorption and dispersion spectra around the anti-Stokes sideband (δ = –ωn), respectively. The other parameters used are g = 0.08 and Δpu = –ωn.

Download Full Size | PDF

Based on the above results, it is significant to further explore the dependence of the probe absorption and dispersion spectra near the anti-Stokes sideband (δ = –ωn) on the exciton-phonon coupling strength so as to achieve a better understanding about the fast and slow light effects. Herein, the probe absorption Imχ(1) and dispersion Reχ(1), versus the probe-pump detuning δ is plotted in Fig. 6 for Δpu = –ωn, against various values of the exciton-phonon coupling strength g. Here we mainly discuss that the amplification of the most remarkable regions in Figs. 6(a) and 6(c). When the coupling between the exciton and the phonons is absent (g = 0), the curve of the absorption spectrum keeps smooth, while when the coupling turns on (g > 0), it becomes very significant. In this case, by increasing g, the width of the absorption peak decreases and the magnitude increases until g approaches a critical value gc (gc = 0.08). When g passes this value, the sharp peak disappears and an attractive dip appears. The features presented in the dispersion spectra are similar to that in Fig. 5(d). This provides a feasible way to measure the exciton-phonon coupling strength g. The physics behind the slow light effect can be understood as follows: as the phonons in the MoS2 NR interact with the exciton, quantum interference between two applied fields and MoS2 NR via the exciton will occur, inducing the emergence of the slow light.

 figure: Fig. 6

Fig. 6 The absorption Imχ(1) (a) and dispersion Reχ(1) (c) spectra as a function of the probe-pump detuning δ for a different coupling strength g. Figures 6(b) and 6(d) show the magnification of the absorption and dispersion spectra around the anti-Stokes sideband (δ = –ωn), respectively. The other parameters used are Δpu = –ωn and Ipu = 3.65 mW/cm2.

Download Full Size | PDF

In this paper, a monolayer MoS2 NR is proposed to advance or slow the propagation of the light, which may lead to some significant improvements in the field of fast and slow light devices. As we all know, carbon nanotube, as an attractive functional material with low power consumption and high hardness, can be used as a good candidate for fast and slow light devices [8]. Compared with carbon nanotube resonator, the proposed MoS2 NR with a wider bandgap and larger surface-to-volume ratio has a greater advantage than carbon nanotube resonator in term of adjustment of the speed of light. By numerical calculation, it was found that the maximal group velocity index of the slow light in the monolayer MoS2 NR is nearly 150 times larger than that in the carbon nanotube resonator under the condition with the same number density. Specifically, it is easy to switch freely between fast and slow light propagation by only adjusting the pumping intensity in the monolayer MoS2 NR. In other word, such a system can act as a two-channel continuous-tunable switch.

4. Summary

In summary, we studied how to switch freely between fast (superluminal) and slow (subluminal) light propagation in a monolayer MoS2 NR. Our results showed that when the exciton-pump detuning approaches the anti-Stokes sideband, the fast and slow light effects are highly sensitive to the excitation frequency. In addition, we found that these two effects can be switched freely by only adjusting the frequency or intensity of the pump field and strong exciton-phonon coupling has a great impact on the slow light effect. Especially, the slow light effect will appear along with a large dispersion with a very steep negative slope and a sharp absorption peak. Finally, we hope that our results can be demonstrated experimentally in the near future.

Funding

National Natural Science Foundation of China (NSFC) under Grants Nos. 11404410, 11504105 and 11504434; Hunan Provincial Natural Science Foundation of China under Grants Nos. 14JJ3116 and 2015JJ3174; Foundation of Talent Introduction of Central South University of Forestry and Technology under Grant No.104-0260.

References and links

1. L. V. Hau, S. E. Harris, Z. Dutton, and C. H. Behroozi, “Light speed reduction to 17 metres per second in an ultracold atomic gas,” Nature 397(6720), 594–598 (1999). [CrossRef]  

2. A. V. Turukhin, V. S. Sudarshanam, M. S. Shahriar, J. A. Musser, B. S. Ham, and P. R. Hemmer, “Observation of ultraslow and stored light pulses in a solid,” Phys. Rev. Lett. 88(2), 023602 (2002). [CrossRef]   [PubMed]  

3. X. Zhou, F. Hocke, A. Schliesser, A. Marx, H. Huebl, R. Gross, and T. J. Kippenberg, “Slowing, advancing and switching of microwave signals using circuit nanoelectromechanics,” Nat. Phys. 9(3), 179–184 (2013). [CrossRef]  

4. R. W. Boyd, D. J. Gauthier, and A. L. Gaeta, “Applications of slow light in telecommunications,” Opt. Photonics News 17(4), 18–23 (2006). [CrossRef]  

5. E. Cabrera-Granado, E. Díaz, and O. G. Calderón, “Slow light in molecular-aggregate nanofilms,” Phys. Rev. Lett. 107(1), 013901 (2011). [CrossRef]   [PubMed]  

6. A. H. Safavi-Naeini, T. P. Mayer Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature 472(7341), 69–73 (2011). [CrossRef]   [PubMed]  

7. Y. Okawachi, M. S. Bigelow, J. E. Sharping, Z. Zhu, A. Schweinsberg, D. J. Gauthier, R. W. Boyd, and A. L. Gaeta, “Tunable all-optical delays via Brillouin slow light in an optical fiber,” Phys. Rev. Lett. 94(15), 153902 (2005). [CrossRef]   [PubMed]  

8. J. J. Li and K. D. Zhu, “Tunable slow and fast light device based on a carbon nanotube resonator,” Opt. Express 20(6), 5840–5848 (2012). [CrossRef]   [PubMed]  

9. M. J. Akram, M. M. Khan, and F. Saif, “Tunable fast and slow light in a hybrid optomechanical system,” Phys. Rev. A 92(2), 023846 (2015). [CrossRef]  

10. M. J. Akram, F. Ghafoor, M. M. Khan, and F. Saif, “Control of Fano resonances and slow light using Bose-Einstein condensates in a nanocavity,” Phys. Rev. A 95(2), 023810 (2017). [CrossRef]  

11. P. C. Ma, Y. Xiao, Y. F. Yu, and Z. M. Zhang, “Microwave field controlled slow and fast light with a coupled system consisting of a nanomechanical resonator and a Cooper-pair box,” Opt. Express 22(3), 3621–3628 (2014). [CrossRef]   [PubMed]  

12. V. Huet, A. Rasoloniaina, P. Guillemé, P. Rochard, P. Féron, M. Mortier, A. Levenson, K. Bencheikh, A. Yacomotti, and Y. Dumeige, “Millisecond photon lifetime in a slow-light microcavity,” Phys. Rev. Lett. 116(13), 133902 (2016). [CrossRef]   [PubMed]  

13. Y. He and M. Jiang, “Ultrasensitive mass sensing method based on slow light in cavity optomechanics,” Appl. Phys. Express 9(5), 052205 (2016). [CrossRef]  

14. K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS₂: a new direct-gap semiconductor,” Phys. Rev. Lett. 105(13), 136805 (2010). [CrossRef]   [PubMed]  

15. S. Bertolazzi, J. Brivio, and A. Kis, “Stretching and breaking of ultrathin MoS2.,” ACS Nano 5(12), 9703–9709 (2011). [CrossRef]   [PubMed]  

16. K. He, C. Poole, K. F. Mak, and J. Shan, “Experimental demonstration of continuous electronic structure tuning via strain in atomically thin MoS2.,” Nano Lett. 13(6), 2931–2936 (2013). [CrossRef]   [PubMed]  

17. A. Castellanos-Gomez, M. Poot, G. A. Steele, H. S. J. van der Zant, N. Agraït, and G. Rubio-Bollinger, “Elastic properties of freely suspended MoS2 nanosheets,” Adv. Mater. 24(6), 772–775 (2012). [CrossRef]   [PubMed]  

18. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-layer MoS2 transistors,” Nat. Nanotechnol. 6(3), 147–150 (2011). [CrossRef]   [PubMed]  

19. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, “Ultrasensitive photodetectors based on monolayer MoS2.,” Nat. Nanotechnol. 8(7), 497–501 (2013). [CrossRef]   [PubMed]  

20. B. Radisavljevic, M. B. Whitwick, and A. Kis, “Small-signal amplifier based on single-layer MoS2,” Appl. Phys. Lett. 101(4), 043103 (2012). [CrossRef]  

21. S. Bikorimana, P. Lama, A. Walser, R. Dorsinville, S. Anghel, A. Mitioglu, A. Micu, and L. Kulyuk, “Nonlinear optical responses in two- dimensional transition metal dichalcogenide multilayer: WS2, WSe2, MoS2 and Mo0.5 W0.5S2,” Opt. Express 24(18), 20685–20695 (2016). [CrossRef]   [PubMed]  

22. S. Zhang, N. Dong, N. McEvoy, M. O’Brien, S. Winters, N. C. Berner, C. Yim, Y. Li, X. Zhang, Z. Chen, L. Zhang, G. S. Duesberg, and J. Wang, “Direct observation of degenerate two-photon absorption and its saturation in WS2 and MoS2 monolayer and few-layer films,” ACS Nano 9(7), 7142–7150 (2015). [CrossRef]   [PubMed]  

23. K. Wang, J. Wang, J. Fan, M. Lotya, A. O’Neill, D. Fox, Y. Feng, X. Zhang, B. Jiang, Q. Zhao, H. Zhang, J. N. Coleman, L. Zhang, and W. J. Blau, “Ultrafast saturable absorption of two-dimensional MoS2 nanosheets,” ACS Nano 7(10), 9260–9267 (2013). [CrossRef]   [PubMed]  

24. N. Kumar, S. Najmaei, Q. Cui, F. Ceballos, P. M. Ajayan, J. Lou, and H. Zhao, “Second harmonic microscopy of monolayer MoS2,” Phys. Rev. B 87(16), 161403 (2013). [CrossRef]  

25. S. Xiao, B. Lv, L. Wu, M. Zhu, J. He, and S. Tao, “Dynamic self-diffraction in MoS(2) nanoflake solutions,” Opt. Express 23(5), 5875–5887 (2015). [CrossRef]   [PubMed]  

26. H. Zeng, J. Dai, W. Yao, D. Xiao, and X. Cui, “Valley polarization in MoS2 monolayers by optical pumping,” Nat. Nanotechnol. 7(8), 490–493 (2012). [CrossRef]   [PubMed]  

27. H. J. Chen and K. D. Zhu, “Coherent optical responses and their application in biomolecule mass sensing based on a monolayer MoS2 nanoresonator,” J. Opt. Soc. Am. B 31(7), 1684–1690 (2014). [CrossRef]  

28. S. Yasukochi, T. Murai, S. Moritsubo, T. Shimada, S. Chiashi, S. Maruyama, and Y. K. Kato, “Gate-induced blueshift and quenching of photoluminescence in suspended single-walled carbon nanotubes,” Phys. Rev. B 84(12), 121409 (2011). [CrossRef]  

29. J. J. Li, W. He, and K. D. Zhu, “All-optical Kerr modulator based on a carbon nanotube resonator,” Phys. Rev. B 83(11), 115445 (2011). [CrossRef]  

30. Y. He, C. Jiang, B. Chen, J. J. Li, and K. D. Zhu, “Optical determination of vacuum Rabi splitting in a semiconductor quantum dot induced by a metal nanoparticle,” Opt. Lett. 37(14), 2943–2945 (2012). [CrossRef]   [PubMed]  

31. B. Gu, W. Ji, P. S. Patil, and S. M. Dharmaprakash, “Ultrafast optical nonlinearities and figures of merit in acceptor-substituted 3,4,5-trimethoxy chalcone derivatives: Structure-property relationships,” J. Appl. Phys. 103(10), 103511 (2008). [CrossRef]  

32. D. F. Walls and G. J. Milburn, Quantum Optics (Springer, 1994).

33. R. W. Boyd, Nonlinear Optics (Academic Press, 2008).

34. R. W. Boyd, M. G. Raymer, P. Narum, and D. J. Harter, “Four-wave parametric interactions in a strongly driven two-level system,” Phys. Rev. A 24(1), 411–423 (1981). [CrossRef]  

35. J. Strait, P. Nene, H. Wang, C. Zhang, and F. Rana, “Carrier relaxation dynamics in MoS2 measured by optical/THz pump–probe spectroscopy,” in Conference on Lasers and Electro-optics (CLEO 2013): Laser Science to Photonic Applications, OSA Technical Digest (online) (Optical Society of America, 2013), paper JTh2A.37. [CrossRef]  

36. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77(2), 633–673 (2005). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 (a) Schematic representation of monolayer MoS2 suspended on a Si/SiO2 substrate by an optical pump-probe technique [31]. The system is driven by a strong pump laser and detected by a weak probe laser. (b) The level scheme of a localized exciton interacting with the phonons in the MoS2 NR.
Fig. 2
Fig. 2 The probe absorption Imχ(1) (a) and dispersion Reχ(1) (b) spectra (in units of Π1) as a function of the probe-pump detuning δ for different exciton-pump detunings Δpu. (c) The energy levels of the exciton dressed with the phonon modes of the MoS2 NR. Each energy level is split into a doublet with a separation Ω'. TP refers to the three-photon resonance, RL represents the Rayleigh resonance, and AC denotes the ac-Stark resonance.
Fig. 3
Fig. 3 The group velocity index ng (in units of Π2) as a function of the exciton-pump detuning Δ pu for Ipu = 3.32 mW/cm2 (a) and Ipu = 6.64 mW/cm2 (b). The insets show the magnification of these remarkable regions around Δ pu = δ = ± ωn.
Fig. 4
Fig. 4 The group velocity index ng (in units of Π2) as a function of the pumping intensity Ipu for Δ pu = –ωn (a), Δ pu = 0 (b), and Δ pu = ωn (c) when the coupling between the exciton and the phonons turns on or off.
Fig. 5
Fig. 5 The absorption Imχ(1) (a) and dispersion Reχ(1) (c) spectra as a function of the probe-pump detuning δ for a different pumping intensity Ipu. Figures 5(b) and 5(d) show the magnification of the absorption and dispersion spectra around the anti-Stokes sideband (δ = –ωn), respectively. The other parameters used are g = 0.08 and Δ pu = –ωn.
Fig. 6
Fig. 6 The absorption Imχ(1) (a) and dispersion Reχ(1) (c) spectra as a function of the probe-pump detuning δ for a different coupling strength g. Figures 6(b) and 6(d) show the magnification of the absorption and dispersion spectra around the anti-Stokes sideband (δ = –ωn), respectively. The other parameters used are Δ pu = –ωn and Ipu = 3.65 mW/cm2.

Equations (10)

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

H = Δ p u σ z + ω n b + b + g ω n σ z ( b + + b ) μ E p u ( σ 10 + σ 01 ) μ ( σ 10 E p r e i δ t + σ 01 E p r * e i δ t ) ,
p ˙ = ( Γ 2 + i Δ p u ) p i g ω n p Ξ i Ω w i μ E p r w e i δ t ,
w ˙ = Γ 1 ( w + 1 ) + 2 i Ω ( p * p ) + 2 i μ ( p * E p r e i δ t p E p r * e i δ t ) ,
Ξ ¨ + γ n Ξ ˙ + ω n 2 Ξ = ω n 2 g w ,
χ ( 1 ) = N μ p 1 ε 0 E p r = N μ 2 ε 0 Γ 2 χ 0 ( 1 ) = Π 1 χ 0 ( 1 ) ,
χ 0 ( 1 ) = ( C 2 C 1 ) C 3 Γ 2 ( 1 + C 1 C 4 C 2 C 4 ) w 0 + p 0 * Γ 2 ( 1 + C 1 C 4 C 2 C 4 ) Ω ,
Γ 1 ( w 0 + 1 ) [ Γ 2 2 + ( Δ p u g g 0 w 0 ) 2 ] + 4 Ω 2 Γ 2 w 0 = 0 ,
v g = c n + ω p r ( d n / d ω p r ) ,
c v g 1 = 2 π Re ( χ ( 1 ) ) ω p r = ω 10 + 2 π ω p r Re ( d χ ( 1 ) d ω p r ) ω p r = ω 10 ,
n g = c v g 1 = Π 2 Γ 2 Re ( d χ 0 ( 1 ) d ω p r ) ω p r = ω 10 .
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.