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

Optical bistability and nonlinearity of coherently coupled exciton-plasmon systems

Open Access Open Access

Abstract

We theoretically investigated optical third-order nonlinearity of a coherently coupled exciton-plasmon hybrid system under a strong control field with a weak probe field. The analytic formulas of exciton population and effective third-order optical susceptibility of the hybrid of a metal nanoparticle (MNP) and a semiconductor quantum dot (SQD) were deduced. The bistable exciton population and the induced bistable nonlinear absorption and refraction response were revealed. The bistability region can be tuned by adjusting the size of metal nanoparticle, interparticle distance and intensity of control field. Our results have perspective applications in optical information processing based on resonant coupling of exciton-plasmon.

© Optical Society of America

1. Introduction

Interactions between exciton systems and plasmonic structures have attracted considerable attention [120]. Because of these interactions, one can enhance emission and fluorescence [3,4], control energy transfer [5], generate single plasmons [6], induce exciton-plasmon-photon conversion [7], modify the spontaneous emission in semiconductor quantum dots (SQDs) [8], etc. Moreover, the exciton coherent dynamics building on the exciton-plasmon interaction have been extensively theoretically studied, including nonlinear Fano resonances [9,10], formation and control of Rabi oscillation [11,12], enhancement of Rabi flopping [11], mid-infrared generation [14] and SQD-induced transparency [15] and so on. Many photonic devices based on exciton-plasmon hybrid structures have already been predicted in theory and demonstrated experimentally, including optical switches and transistors [16,17], nanoscale lasers [1820], photodetectors [21] and optical modulators [22].

Many investigations of the third-order optical nonlinearity and bistability that arise in the coherently coupled exciton-plasmon system have also been reported. Recently, Wang et al. experimentally demonstrated that the effective nonlinear absorption and refraction of CdS-Ag core-shell QDs comparing with CdS QDs were greatly enhanced [23]. Furthermore, Ji’s group observed the intensity-dependent enhancement of saturable absorption in a PbS-Au nanohybrid composite [24]. In theoretical research, Xiong et al. reported that third-order nonlinear optical susceptibility of CdTe QDs in the hybrid SQD-MNP system can be effectively modified by local field enhancement and dipole interaction [25]. Enhancement of Kerr nonlinearity in the exciton-plasmon hybrid system has also been theoretically demonstrated [26]. Optical bistability has been exhibited in some strong coupled exciton-plasmon hybrid systems, which could not been manifested in strongly coupled two-level molecules [27]. Artuso et al. found a regime of bistability in the hybrid system when SQD strongly coupled to MNP by increasing the sizes of both the SQD and the MNP. Then, they probed the transition to bistability that existed in such system, and saw a discontinuous jump in the response of the system [28,29]. Most recently, optical bistability and hysteresis in a hybrid SQD-MNP nanodimer could be realized by adjusting the incident field intensity and was manifested by the internal degrees of freedom of such system, there properties could be revealed by measuring the intensity of the Rayleigh scattering [30]. However, the nonlinear absorption and refraction response related to the optical bistability arose from the exciton-plasmon interaction have not been reported.

In this paper, we theoretically deduced the analytic formulas for exciton population and effective third-order optical susceptibility of the SQD-MNP hybrid system under a strong control field with a weak probe field. We studied the third-order optical nonlinearity of the system and found the bistable exciton population and the induced bistable nonlinear absorption and refraction response.

2. Model and formalism

We consider a system comprising a spherical SQD with radius r and a spherical MNP with radius R, separated by a surface-to-surface distance d (Fig. 1(a) ). The SQD and MNP are described by a density matrix formalism and classical electrodynamics, respectively. The hybrid is subject to a strong control field and a weak signal field (Fig. 1(b)).

 figure: Fig. 1

Fig. 1 (a) Schematic diagram of the hybrid system driven by a strong control field with amplitude Ec and frequency ωc and probed by a weak signal field with amplitude Es and frequency ωs (Ec >>Es). r and R are the radii of SQD and MNP, respectively. d is the surface-to-surface distance between SQD and MNP. ε0, εs and εm are the dielectric constants of the background medium, SQD and MNP, respectively. (b) Energy level diagram of the system. ω10 and ωsp are the frequencies of exciton and surface plasmon, respectively.

Download Full Size | PDF

The Hamiltonian of the system in the rotating-wave approximation can be written as

H^=Δσ^zμ(E˜SQDσ^01+E˜SQDσ^10),
where Δ = ω10 - ωc is the difference between the exciton frequency ω10 and the control field frequency ωc.σ^ij(i, j = 0,1) is the dipole transition operator between |i〉 and |j〉 of SQD, σ^z=(σ^11σ^00)/2. µ is the interband dipole moment element. SQD and MNP are the total electric field felt by the SQD and MNP, respectively. Here, SQD = Ec + Ese-iδt + SαPMNP /[εeff (d + r + R)3], MNP = Ec + Ese-iδt + SαPSQD /[εeff (d + r + R)3] [9,31]. εeff = (2ε0 + εs)/3, where ε0 and εs are the dielectric constants of the background medium and SQD, respectively. δ = ωsωc is the frequency difference of the signal field and the control field. Geometric factor Sα is equal to 2 (−1) when E is parallel to the z (x, y) axis. The z direction corresponds to the axis of the hybrid system. The dipole of the MNP PMNP = αMNPMNP derives from the charge induced on the surface of the MNP, where αMNP = ε0R3(εm-ε0)/(2ε0 + εm) [12], and εm is the dielectric constant of the Au nanoparticle [32]. The dipole moment of the SQD is PSQD = µσ10 [33]. Based on the Heisenberg equation of motion, if we set p = µσ10 and w = 2σz, and ignore thequantum properties of σ^10 and σ^z [34], then the equations give
p˙=(1/T2+iΔ)piμ2E˜SQDw/w˙=(w+1)/T1+4Im(pE˜SQD)/,
where T1 and T2 are the exciton lifetime and the exciton dephasing time, respectively. SQD = A(Ec + Ese-iδt) + Bp [31], where A = 1 + SααMNP /[εeff (d + r + R)3] and B = Sα2αMNP /[εeff2(d + r + R)6]. To obtain a steady-state solution for Eq. (2), we make the ansatz [33,35] p(t) = p0 + p1e-iδt + p-1eiδt, w(t) = w0 + w1e-iδt + w-1eiδt, where p0, p1 and p-1 are the dipole moments of the SQD oscillating at frequencies ωc, ωs and 2ωc - ωs, respectively. Similarly, w0, w1 and w-1 are the population inversions oscillating at frequencies ωc, ωs and 2ωc - ωs, respectively. Here, p-1 represents a wave-mixing response, because it gives rise to generation of an optical wave with frequency 2ωc - ωs. We assume that the other terms are small in the sense that |p0|>>|p1|, |p-1| and |w0|>>|w1|, |w-1|. Upon working to the lowest order in Es but to all orders in Ec, we can get p-1, which are related to the effective nonlinear optical susceptibility as follows:
χeff(3)=Np13ε0Ec2Es=2NA3μ4T23w03ε03(iΔc)(2iδc)[i+(Δc+Bcw0)][i(Δc+Bcw0)]D(δc),
where

D(δc)=(δciT2/T1)[1+(Δc+Bcw0)2][(δci)2(Δc+Bcw0)2]+4A2Ωc2[(iδc)(1+Δc2)Bcw0(iBcw0+Δcδc)].

Here, N is the number density of the hybrid system, Ωc2 = µ2|Ec|2T22 /ћ2 is the generalized intensity of the control field, δc = δT2, Δc = ΔT2 and Bc = µ2BT2 /ћ2. Imχeff(3) and Reχeff(3) represent the nonlinear absorption and refraction, respectively. The population inversion of the exciton w0 is determined by the cubic equation

w0=4A2Ωc2(T1/T2)w0/[1+(Δc+Bcw0)2]1.

The Eq. (5) is of the third order in w0 and therefore may have one or three real solutions, strongly depending on the radius of MNP, surface-to-surface distance d and the intensity of the control field Ωc2. The latter case corresponds to the optical bistability arising from the strong coupling between SQD and MNP.

3. Results and discussion

To obtain an insight from these complicated expressions, we graphically display them for a wide range of some important parameters. We take the typical values ε0 = 2.25, εs = 6, T1 = 0.8 ns, T2 = 0.3 ns and µ = 10−28 C∙m [9], and N = 1014 m−3. The bare exciton frequency is chosen to be 2.5 eV, which is close to the broad plasmon frequency of gold (peak near 2.4 eV with a width of approximately 0.25 eV). As for the Au nanoparticle size region we consider, the plasmon resonance frequency changes little with the size.

Figures 2(a) and 2(b) show the dynamic evolution of nonlinear absorption and refraction spectra with four different detunings (ω10 - ωc)T2. In Fig. 2(a), the absorption spectra which are located in regions I, II, III are originated from the three-photon resonance, stimulated Rayleigh resonance and ac-Stark resonance, respectively [35]. In region I, the absorption spectrum attributed to the three-photon effect evolves from a Fano-like lineshape into a negative peak and has a peak redshift as the detuning (ω10 - ωc)T2 increases. In region II, the absorption spectrum assigned to the Rayleigh resonance evolves from a negative peak into a Fano-like lineshape and has an increased magnitude. In region III, the absorption spectrum induced by the ac-Stark effect evolves from the Fano-like lineshape into a positive peak, and the peak first redshifts and then blueshifts. The corresponding refraction spectra are shown in Fig. 2(b). As required by the nonlinear Kramers-Kronig relation, the complementary behavior between the absorptive and refractive features is evident. From resonant excitation to near-resonant excitation, the dynamic evolution of the absorption and refraction spectra implies that the dominant role of three resonance mechanisms changes continuously. In the following discussion, we focus on the nonlinear absorption and refraction response of the hybrid under near-resonant excitation when ωs = ωc corresponding to the stimulated Rayleigh resonance.

 figure: Fig. 2

Fig. 2 Nonlinear absorption Imχeff(3) (a) and refraction Reχeff(3) (b) as a function of the detuning (ωs - ωc)T2 with d = 15 nm and (ω10 - ωc)T2 = 0, 1, 2, 5.

Download Full Size | PDF

Figure 3(a) shows the bistability of exciton population on the interparticle distance between SQD and MNP d for the radius of MNP R = 5, 7.5, 10 nm. There are three real roots for w0 in Eq. (5) for a given d, which implies that a regime of bistability emerges. A region of optical bistability is formed by adjusting the value of d. In the bistability region of db0ddb1 (db0 and db1 are the critical values that the optical bistable response exists), population inversion w0 versus d manifests a standard S-shaped curve, while the induced nonlinear absorption and refraction spectra are exotic coiled curves. When d < db0 or d > db1, the bistability effect disappears. The nonlinear refraction of SQD-MNP system compared with SQD system is obviously enhanced. Speaking clearly, the maximum value of |Reχeff(3)| of the hybrid system is 3.1, 3.6, 4.3 times that of SQD system, corresponding to R = 5, 7.5, 10 nm, respectively. But the enhancement of nonlinear refraction is negligible. Take R = 7.5 nm for example, Figs. 3(c) and 3(e) show the nonlinear absorption and refraction spectra in the bistability region (12.52 nm ≤ d ≤ 13.09 nm), respectively. The bistable nonlinear absorption response curve traces out a crossed path ①→②→③→④→⑤→⑥→⑦, and the bistable refraction response curve displays a complementary behavior. When the interaction between SQD and MNP is strong enough by adjusting the interparticle distance d, the interaction can result in bistable optical nonlinear absorption and refraction response.

 figure: Fig. 3

Fig. 3 (a) Population inversion w0 as a function of interparticle distance between SQD and MNP d for R = 5, 7.5, 10 nm with ωc = ωs and (ω10 - ωc)T2 = 1. (b) Nonlinear absorption Imχeff(3) versus d. (c) Imχeff(3) versus d for R = 7.5 nm in the bistability region of (b). (d) Nonlinear refraction Reχeff(3) versus d. (e) Reχeff(3) versus d for R = 7.5 nm in the bistability region of (d).

Download Full Size | PDF

Figure 4(a) shows the bistability of exciton population on intensity of control field Ωc2 for R = 7.5 nm and d = 12.8 nm. Population inversion w0 versus Ωc2 manifests a standard S-shaped curve by adjusting the control field intensity, implying a standard bistable behavior. When the control field intensity slowly increases, the system firstly follows the lower (stable) branch and then jumps to the upper (stable) branch at the critical intensity. With sweeping the intensity back, the system remains on the upper branch and then makes a transition to the lower branch when the intensity passes through the other critical value. A hysteresis loop has been completed. The intermediate branch is unstable. Similar optical hysteresis curve has also been observed by Malyshev et al. in the metal-semiconductor nanodimer when they measured the Rayleigh scattering intensity as a function of excitation intensity [30]. Both Imχeff(3) and Reχeff(3) behave in a different manner, and display bistable curves (see Figs. 4b4e). Obviously, the nonlinear absorption and refraction response (Imχeff(3) and Reχeff(3)) could be used to manifest optical bistability and optical hysteresis. Out of the bistability region, the maximum value of |Reχeff(3)| of the hybrid system is 2.8 times that of SQD system.

 figure: Fig. 4

Fig. 4 (a) Population inversion w0 as a function of control field intensity Ωc2 for R = 7.5 nm and d = 12.8 nm with ωc = ωs and (ω10 - ωc)T2 = 1. (b) Nonlinear absorption Imχeff(3) versus Ωc2. (c) Imχeff(3) versus Ωc2 in the bistability region of (b). (d) Nonlinear refraction Reχeff(3) versus Ωc2. (e) Reχeff(3) versus Ωc2 in the bistability region of (d).

Download Full Size | PDF

4. Conclusion

In conclusion, we theoretically investigated the third-order optical nonlinearity of a coherently coupled SQD-MNP system in the presence of a strong control field with a weak probe field. We deduced the analytic formulas of exciton population and effective third-order optical susceptibility of the hybrid, and found the bistable exciton population and the induced bistable nonlinear absorption and refraction response. We showed that the bistability region can be tuned by adjusting the size of metal nanoparticle, interparticle distance and intensity of control field. The optical bistability promised possible applications as optical memory cells and optical switches. We hoped that our predictions in this work can be experimentally demonstrated using the optical pump-probe method in the near future.

Acknowledgments

This work was supported by NSFC (10874134, 61008043 and 11004001), National Program on Key Science Research of China (2011CB922201), and Key Project of Ministry of Education of China (708063).

References and links

1. D. E. Chang, A. S. Sørensen, P. R. Hemmer, and M. D. Lukin, “Quantum optics with surface plasmons,” Phys. Rev. Lett. 97(5), 053002 (2006). [CrossRef]   [PubMed]  

2. N. T. Fofang, N. K. Grady, Z. Y. Fan, A. O. Govorov, and N. J. Halas, “Plexciton dynamics: exciton-plasmon coupling in a J-Aggregate-Au nanoshell complex provides a mechanism for nonlinearity,” Nano Lett. 11(4), 1556–1560 (2011). [CrossRef]   [PubMed]  

3. A. O. Govorov and I. Carmeli, “Hybrid structures composed of photosynthetic system and metal nanoparticles: plasmon enhancement effect,” Nano Lett. 7(3), 620–625 (2007). [CrossRef]   [PubMed]  

4. S. Kühn, U. Håkanson, L. Rogobete, and V. Sandoghdar, “Enhancement of single-molecule fluorescence using a gold nanoparticle as an optical nanoantenna,” Phys. Rev. Lett. 97(1), 017402 (2006). [CrossRef]   [PubMed]  

5. A. O. Govorov, G. W. Bryant, W. Zhang, T. Skeini, J. Lee, N. A. Kotov, J. M. Slocik, and R. R. Naik, “Exciton-plasmon interaction and hybrid excitons in semiconductor-metal nanoparticle assemblies,” Nano Lett. 6(5), 984–994 (2006). [CrossRef]  

6. A. V. Akimov, A. Mukherjee, C. L. Yu, D. E. Chang, A. S. Zibrov, P. R. Hemmer, H. Park, and M. D. Lukin, “Generation of single optical plasmons in metallic nanowires coupled to quantum dots,” Nature 450(7168), 402–406 (2007). [CrossRef]   [PubMed]  

7. Y. Fedutik, V. V. Temnov, O. Schöps, U. Woggon, and M. V. Artemyev, “Exciton-plasmon-photon conversion in plasmonic nanostructures,” Phys. Rev. Lett. 99(13), 136802 (2007). [CrossRef]   [PubMed]  

8. G. Y. Chen, Y. N. Chen, and D. S. Chuu, “Spontaneous emission of quantum dot excitons into surface plasmons in a nanowire,” Opt. Lett. 33(19), 2212–2214 (2008). [CrossRef]   [PubMed]  

9. W. Zhang, A. O. Govorov, and G. W. Bryant, “Semiconductor-metal nanoparticle molecules: hybrid excitons and the nonlinear fano effect,” Phys. Rev. Lett. 97(14), 146804 (2006). [CrossRef]   [PubMed]  

10. W. Zhang and A. O. Govorov, “Quantum theory of the nonlinear Fano effect in hybrid metal-semiconductor nanostructures: The case of strong nonlinearity,” Phys. Rev. B 84, 081405 (R)(2011).

11. S. M. Sadeghi, “Plasmonic metaresonances: molecular resonances in quantum dot-metallic nanoparticle conjugates,” Phys. Rev. B 79(23), 233309 (2009). [CrossRef]  

12. M. T. Cheng, S. D. Liu, H. J. Zhou, Z. H. Hao, and Q. Q. Wang, “Coherent exciton-plasmon interaction in the hybrid semiconductor quantum dot and metal nanoparticle complex,” Opt. Lett. 32(15), 2125–2127 (2007). [CrossRef]   [PubMed]  

13. S. M. Sadeghi, “The inhibition of optical excitations and enhancement of Rabi flopping in hybrid quantum dot-metallic nanoparticle systems,” Nanotechnology 20(22), 225401 (2009). [CrossRef]   [PubMed]  

14. J. Y. Yan, W. Zhang, S. Duan, and X. G. Zhao, “Plasmon-enhanced midinfrared generation from difference frequency in semiconductor quantum dots,” J. Appl. Phys. 103(10), 104314 (2008). [CrossRef]  

15. R. D. Artuso, G. W. Bryant, A. Garcia-Etxarri, and J. Aizpurua, “Using local fields to tailor hybrid quantum-dot/metal nanoparticle systems,” Phys. Rev. B 83(23), 235406 (2011). [CrossRef]  

16. D. E. Chang, A. S. Sørensen, E. A. Demler, and M. D. Lukin, “A single-photon transistor using nanoscale surface plasmons,” Nat. Phys. 3(11), 807–812 (2007). [CrossRef]  

17. S. M. Sadeghi, “Tunable nanoswitches based on nanoparticle meta-molecules,” Nanotechnology 21(35), 355501 (2010). [CrossRef]   [PubMed]  

18. D. J. Bergman and M. I. Stockman, “Surface plasmon amplification by stimulated emission of radiation: quantum generation of coherent surface plasmons in nanosystems,” Phys. Rev. Lett. 90(2), 027402 (2003). [CrossRef]   [PubMed]  

19. M. A. Noginov, G. Zhu, A. M. Belgrave, R. Bakker, V. M. Shalaev, E. E. Narimanov, S. Stout, E. Herz, T. Suteewong, and U. Wiesner, “Demonstration of a spaser-based nanolaser,” Nature 460(7259), 1110–1112 (2009). [CrossRef]   [PubMed]  

20. R. F. Oulton, V. J. Sorger, T. Zentgraf, R. M. Ma, C. Gladden, L. Dai, G. Bartal, and X. Zhang, “Plasmon lasers at deep subwavelength scale,” Nature 461(7264), 629–632 (2009). [CrossRef]   [PubMed]  

21. J. S. White, G. Veronis, Z. F. Yu, E. S. Barnard, A. Chandran, S. H. Fan, and M. L. Brongersma, “Extraordinary optical absorption through subwavelength slits,” Opt. Lett. 34(5), 686–688 (2009). [CrossRef]   [PubMed]  

22. D. Pacifici, H. J. Lezec, and H. A. Atwater, “All-optical modulation by plasmonic excitation of CdSe quantum dots,” Nat. Photonics 1(7), 402–406 (2007). [CrossRef]  

23. H. M. Gong, X. H. Wang, Y. M. Du, and Q. Q. Wang, “Optical nonlinear absorption and refraction of CdS and CdS-Ag core-shell quantum dots,” J. Chem. Phys. 125(2), 024707 (2006). [CrossRef]   [PubMed]  

24. H. I. Elim, W. Ji, J. Yang, and J. Y. Lee, “Intensity-dependent enhancement of saturable absorption in PbS–Au4 nanohybrid composites: evidence for resonant energy transfer by Auger recombination,” Appl. Phys. Lett. 92(25), 251106 (2008). [CrossRef]  

25. X. Zhang and G. G. Xiong, “Metal nanoparticle-induced variation of nonlinear optical susceptibility of a CdTe semiconductor quantum dot,” Physica E 41(7), 1258–1262 (2009). [CrossRef]  

26. Z. E. Lu and K. D. Zhu, “Enhancing Kerr nonlinearity of a strong coupled exciton-plasmon in hybrid nanocrystal molecules,” J. Phys. B 41(18), 185503 (2008). [CrossRef]  

27. V. A. Malyshev, H. Glaeske, and K.-H. Feller, “Absence of bistable behavior in the optical response of a dimer,” Phys. Rev. A 58(2), 1496–1500 (1998). [CrossRef]  

28. R. D. Artuso and G. W. Bryant, “Optical response of strongly coupled quantum dot-metal nanoparticle systems: Double peaked Fano structure and bistability,” Nano Lett. 8(7), 2106–2111 (2008). [CrossRef]   [PubMed]  

29. R. D. Artuso and G. W. Bryant, “Strongly coupled quantum dot-metal nanoparticle systems: Exciton-induced transparency, discontinuous response, and suppression as driven quantum oscillator effects,” Phys. Rev. B 82(19), 195419 (2010). [CrossRef]  

30. A. V. Malyshev and V. A. Malyshev, “Optical bistability and hysteresis of a hybrid metal-semiconductor nanodimer,” Phys. Rev. B 84(3), 035314 (2011). [CrossRef]  

31. H. Wang and K. D. Zhu, “Coherent optical spectroscopy of a hybrid nanocrystal complex embedded in a nanomechanical resonator,” Opt. Express 18(15), 16175–16182 (2010). [CrossRef]   [PubMed]  

32. E. D. Palik, Handbook of Optical Constants of Solids (Academic, Orlando, Fla., 1985).

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

34. B. I. Greene, J. F. Mueller, J. Orenstein, D. H. Rapkine, S. Schmitt-Rink, and M. Thakur, “Phonon-mediated optical nonlinearity in polydiacetylene,” Phys. Rev. Lett. 61(3), 325–328 (1988). [CrossRef]   [PubMed]  

35. 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]  

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 (4)

Fig. 1
Fig. 1 (a) Schematic diagram of the hybrid system driven by a strong control field with amplitude Ec and frequency ωc and probed by a weak signal field with amplitude Es and frequency ωs (Ec >>Es). r and R are the radii of SQD and MNP, respectively. d is the surface-to-surface distance between SQD and MNP. ε0, εs and εm are the dielectric constants of the background medium, SQD and MNP, respectively. (b) Energy level diagram of the system. ω10 and ωsp are the frequencies of exciton and surface plasmon, respectively.
Fig. 2
Fig. 2 Nonlinear absorption Imχeff(3) (a) and refraction Reχeff(3) (b) as a function of the detuning (ωs - ωc)T2 with d = 15 nm and (ω10 - ωc)T2 = 0, 1, 2, 5.
Fig. 3
Fig. 3 (a) Population inversion w0 as a function of interparticle distance between SQD and MNP d for R = 5, 7.5, 10 nm with ωc = ωs and (ω10 - ωc)T2 = 1. (b) Nonlinear absorption Imχeff(3) versus d. (c) Imχeff(3) versus d for R = 7.5 nm in the bistability region of (b). (d) Nonlinear refraction Reχeff(3) versus d. (e) Reχeff(3) versus d for R = 7.5 nm in the bistability region of (d).
Fig. 4
Fig. 4 (a) Population inversion w0 as a function of control field intensity Ωc2 for R = 7.5 nm and d = 12.8 nm with ωc = ωs and (ω10 - ωc)T2 = 1. (b) Nonlinear absorption Imχeff(3) versus Ωc2. (c) Imχeff(3) versus Ωc2 in the bistability region of (b). (d) Nonlinear refraction Reχeff(3) versus Ωc2. (e) Reχeff(3) versus Ωc2 in the bistability region of (d).

Equations (5)

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

H ^ =Δ σ ^ z μ( E ˜ SQD σ ^ 01 + E ˜ SQD σ ^ 10 ),
p ˙ =( 1/ T 2 +iΔ )pi μ 2 E ˜ SQD w/ w ˙ =( w+1 )/ T 1 +4Im( p E ˜ SQD )/,
χ eff ( 3 ) = N p 1 3 ε 0 E c 2 E s = 2N A 3 μ 4 T 2 3 w 0 3 ε 0 3 ( i Δ c )( 2i δ c )[ i+( Δ c + B c w 0 ) ] [ i( Δ c + B c w 0 ) ]D( δ c ) ,
D( δ c )=( δ c i T 2 / T 1 )[ 1+ ( Δ c + B c w 0 ) 2 ][ ( δ c i ) 2 ( Δ c + B c w 0 ) 2 ] +4 A 2 Ω c 2 [ ( i δ c )( 1+ Δ c 2 ) B c w 0 ( i B c w 0 + Δ c δ c ) ] .
w 0 = 4 A 2 Ω c 2 ( T 1 / T 2 ) w 0 / [ 1+ ( Δ c + B c w 0 ) 2 ] 1 .
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.