Next Article in Journal
Model of the Effects of Femtosecond Laser Pulse Energy on the Effective Z-Position of the Resulting Cut after Laser-Induced Optical Breakdown
Previous Article in Journal
Birefringence and Anisotropy of the Losses Due to Two-Photon Absorption of Femtosecond Pulses in Crystals
Previous Article in Special Issue
Optical Rogue Waves in Fiber Lasers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Nonlinear Dynamics of Silicon-Based Epitaxial Quantum Dot Lasers under Optical Injection

1
School of Physical Science and Technology, Southwest University, Chongqing 400715, China
2
Chongqing Key Laboratory of Micro & Nano Structure Optoelectronics, Southwest University, Chongqing 400715, China
*
Author to whom correspondence should be addressed.
Photonics 2024, 11(8), 684; https://doi.org/10.3390/photonics11080684 (registering DOI)
Submission received: 26 June 2024 / Revised: 20 July 2024 / Accepted: 22 July 2024 / Published: 23 July 2024
(This article belongs to the Special Issue Advanced Lasers and Their Applications II)

Abstract

:
For silicon-based epitaxial quantum dot lasers (QDLs), the mismatches of the lattice constants and the thermal expansion coefficients lead to the generation of threaded dislocations (TDs), which act as the non-radiative recombination centers through the Shockley–Read–Hall (SRH) recombination. Based on a three-level model including the SRH recombination, the nonlinear properties of the silicon-based epitaxial QDLs under optical injection have been investigated theoretically. The simulated results show that, through adjusting the injection parameters including injection strength and frequency detuning, the silicon-based epitaxial QDLs can display rich nonlinear dynamical states such as period one (P1), period two (P2), multi-period (MP), chaos (C), and injection locking (IL). Relatively speaking, for a negative frequency detuning, the evolution of the dynamical state with the injection strength is more abundant, and an evolution path P1-P2-MP-C-MP-IL has been observed. Via mapping the dynamical state in the parameter space of injection strength and frequency detuning under different SRH recombination lifetime, the effects of SRH recombination lifetime on the nonlinear dynamical state of silicon-based epitaxial QDLs have been analyzed.

1. Introduction

Semiconductor lasers exhibit rich dynamical behaviors after being subjected to external perturbations such as optical injection [1], optical feedback [2], and optoelectronic feedback [3], which can be applied in various scenarios, including chaotic secure communications [4], photonic microwave amplifiers [5], radio-over-fiber uplink transmission [6], reservoir computing [7], and random bit generation [8,9].
In recent years, with the development of fabrication technology, quantum dot lasers (QDLs) have shown many performance advantages over traditional semiconductor lasers, such as a lower threshold current density [10], improved stability against optical feedback [11,12], minimal impact from the environmental temperature [13], lower relative intensity noise (RIN) [14], and a small linewidth enhancement factor [15,16]. These above advantages arise from the unique self-assembly growth mode of quantum dots (QDs) structures. Three-dimensional quantum confinement of QDs leads to a complete discretization of the energy levels, with the atom-like degeneracy into the form of a delta-function-like state, which gives quantum dot lasers (QDLs) a different form of emission from other semiconductor lasers [17]. Similar to traditional semiconductor lasers, QDLs can also emit from the ground state (GS). Furthermore, it is possible to observe QDLs emitting from both the GS and the excited state (ES). In QDLs, the GS is excited when a slight bias current is applied, followed by the ES once the current surpasses the second threshold. As the current rises, the ES lasing becomes dominant while the GS lasing weakens. It has also been shown that the lasing state of a QDL can be directly switched from ES to GS under certain conditions when there is external optical feedback as an external perturbation [18].
The rapid increase in internet traffic has made the development of low-cost, high-speed, and low-power optical communication technology in data centers become a hot topic. As a result, scientists have turned their focus towards innovating the next generation of information and communication technology by looking into silicon-based photonic integrated circuits (PICs) [19]. For silicon-based PICs, both optically active and passive components are monolithically integrated on a single chip, offering a high bandwidth density, high energy efficiency, and low latency. Unfortunately, silicon, which is an indirect band gap material, is not suitable for the generation of lasers. In contrast, direct bandgap III-V compound semiconductors have robust optical properties that can be tailored for lasers operating at various wavelengths with high efficiency, a large modulation bandwidth, and sufficient optical power output [20]. Therefore, epitaxial growth of III-V QDLs directly on silicon substrates is considered as one of the most promising methods for achieving on-chip light sources in silicon-based PICs [21,22,23,24,25]. For QDLs directly grown on silicon, the lattice mismatch between silicon and QDs semiconductor materials and thermal expansion coefficients lead to the generation of threaded dislocations (TDs), which act as non-radiative recombination centers through the Shockley–Read–Hall (SRH) recombination. In silicon-based epitaxial QDLs, TDs densities are typically 106~108 cm−2. Despite significant efforts to improve the quality of silicon-based epitaxial QDLs, the achievable TDs densities are still around 105~106 cm−2 [26,27]. Previous investigations demonstrated that, for silicon-based epitaxial QDLs under optical feedback, the static and dynamic characteristics are influenced by the SRH recombination. For a shorter SRH recombination lifetime, a smaller linewidth enhancement factor but a larger damping factor are obtained. Furthermore, any reduction in the SRH recombination lifetime will shrink the chaotic region and shift the first Hopf bifurcation point to stronger feedback [28]. As a result, for inspecting the dynamic characteristics of silicon-based epitaxial QDLs under optical feedback, the SRH recombination should be taken into account. Since optical feedback can be regarded as a special light injection from a certain perspective, it can be predicted that, for analyzing the dynamic characteristics of silicon-based epitaxial QDLs under optical injection, the SRH recombination should also be considered. However, we have noticed that relevant reports on the nonlinear dynamical characteristics of silicon-based epitaxial QDLs under optical injection are almost based on the model excluding the SRH recombination.
In this paper, based on a three-level model including SRH recombination lifetime, we investigate the dynamical behaviors of silicon-based epitaxial QDLs under optical injection. By changing the injection strength and frequency detuning, different nonlinear dynamical states are observed, and the route for silicon-based epitaxial QDLs entering into chaos is revealed. In addition, the impact of the SRH recombination lifetime on the nonlinear dynamical state of silicon-based epitaxial QDLs under optical injection has also been analyzed.

2. Theoretical Model

Figure 1 shows a schematic diagram of the carrier dynamics of silicon-based epitaxial QDLs, considering charged electrons and holes as neutral excitons. The emission time of the spontaneous emission of carriers in the ground state (GS), excited state (ES) and reservoir state (RS) as shown in Figure 1 is τ G S s p o n , τ E S s p o n , τ R S s p o n . It should be emphasized that the rate of SRH recombination is characterized by the SRH recombination lifetime τSRH. Carriers are pumped directly into the RS through the Auger recombination process; some carriers are trapped from RS to the ES with a captured time τ E S R S , and some carriers relaxed from ES to GS with a relaxation time τ G S E S . In addition, through thermal excitations, some carriers in GS escape into ES with an escape time τ E S G S , and from ES escape into RS with an escape time τ R S E S ; stimulated radiation occurs only in the GS. After introducing optical injection, the rate equation for silicon-based epitaxial QDLs including the SRH recombination lifetime can be described as follows [29]:
d N R S d t = I q + N E S τ R S E S N R S τ E S R S ( 1 ρ E S ) N R S τ R S s p o n N R S τ S R H
d N E S d t = ( N R S τ E S R S + N G S τ E S G S ) ( 1 ρ E S ) N E S τ G S E S ( 1 ρ G S ) N E S τ R S E S N E S τ E S s p o n N E S τ S R H
d N G S d t = N E S τ G S E S ( 1 ρ G S ) N G S τ E S G S ( 1 ρ E S ) Γ p v g g G S S G S N G S τ G S s p o n N G S τ S R H
d ϕ d t = 1 2 Γ p v g ( g G S α H G S + g E S α H E S + g R S α H R S ) Δ ω i n j k c k i n j S i n j / S G S sin ϕ
d S G S d t = ( Γ p v g g G S 1 τ p ) S G S + β s p N G S τ G S s p o n + 2 k c k i n j S i n j S G S cos ϕ
where N, S, ϕ are the carrier number, photon number, and phase, respectively. I is the bias current; q is the electron charge; Γp is the optical confinement factor; vg is the group velocity, defined as vg = c/nr, where c is the speed of light and nr is the refractive index. α H G S , E S , R S is the GS, ES, and RS contribution to the αH factor, respectively. βsp is the spontaneous emission factor; Δωinj is the frequency detuning; kc is the coupling coefficient of master and slave lasers, defined as kc = vg(1 − R)/(2L(R)1/2) where L is the laser cavity length; R is the facet reflectivity; and τp is the photon lifetime. Sinj is the photon number of injection light, and kinj is the injection strength. The material gain coefficient gGS,ES,RS can be described by the following:
g G S = a G S 1 + ε S G S N B V B ( 2 ρ G S 1 )
g E S = a E S N B V B ( 2 ρ E S 1 )
g R S = a R S D R S V R S ( 2 ρ R S 1 )
where aGS, ES, RS are the differential gain of different states, ε is the gain compression factor, NB is the total dot number of QDs, VB is the volume of the active region, DRS is the total number of states in the RS, VRS is the volume of the RS, ρGS = NGS/(2NB), ρES = NES/(4NB) and ρRS = NRS/DRS represent the occupancy probabilities of carriers in GS, ES, and RS, respectively. For simplicity, the noises have been neglected. These rate equations can be numerically solved using the fourth-order Runge–Kutta method through MATLAB R2022a software.
The relationship between the SRH recombination lifetime and the threading dislocation (TD) density in silicon-based epitaxial QDLs can be written as follows [30]:
1 τ S R H = 1 τ S R H 0 + D π 3 σ 4
where τ S R H 0 = 1877 ns is the lifetime of dislocation-free GaAs-based QDLs, and D = 221 cm2 s−1 is the average diffusion coefficient. σ is the TD density value, and its value is within the range of (105~108) cm−2 [31], which is at least two orders of magnitude higher than in GaAs-based QDLs [26]. Therefore, the SRH recombination in silicon-based epitaxial QDLs is stronger than that in GaAs-based QDLs. The relationship between TD density and the SRH recombination lifetime is shown in Figure 2. As shown in this diagram, for the current attainable TD density of approximately (105~106) cm−2, the corresponding SRH recombination lifetime is about 5~0.5 ns. In this case, the SRH recombination lifetime is equivalent to the spontaneous emission lifetime, and, therefore, the SRH recombination cannot be ignored. The other parameters utilized during the simulation are as follows: τ E S R S = 6.3 ps, τ G S E S = 2.9 ps, τ R S E S = 2.7 ns, τ E S G S = 10.4 ps, τ G S s p o n = 1.2 ns, τ E S s p o n = 0.5 ns, τ R S s p o n = 0.5 ns, Sinj = 3 × 105, Γp = 0.06, nr = 3.5, α H G S = 0.5, α H E S = 0.122, α H R S = 0.037, βsp = 1.0 × 10−4, kc = 10 × 1010 s−1, τp = 4.1 ps, τSRH = 0.5~5 ns, aGS = 5.0 × 10−15 cm2, aES = 10.0 × 10−15 cm2, aRS = 2.5 × 10−15 cm2, ε = 2.0 × 10−16 cm3, NB = 1.0 × 107, VB = 5.0 × 10−11 cm3, DRS = 4.8 × 106, VRS = 1.0 × 10−11 cm3 [32].

3. Results and Discussion

In this work, we consider the case that the SRH recombination lifetime (τSRH) of a silicon-based epitaxial quantum dot laser (QDL) is set at 0.5 ns, 1 ns, 5 ns, respectively. For the convenience of comparison, the case of ignoring SRH recombination is also considered. It should be pointed out that, when ignoring the SRH recombination, the last term in Equations (1)–(3) is absence, and the rate equations are transferred into those reported in Ref. [33]. Figure 3 illustrates the function of the GS photon number associated with the bias current under different values of τSRH for a silicon-based epitaxial QDL at free-running. From this diagram, it can be seen that, with the increase in τSRH from 0.5 ns to 5 ns, the threshold current decreases from 50 mA to 20 mA. When τSRH takes 5 ns, the variation trend of the output power is very close to that obtained when ignoring the SRH recombination. In other words, we can ignore the effect of the SRH recombination when τSRH is greater than the spontaneous emission lifetime. In the following, the bias current is set at 150 mA.
The above results show that the SRH recombination lifetime affects the output photon number under a given current. Next, we will analyze the influence of the SRH recombination lifetime on the dynamical state of silicon-based epitaxial QDL under optical injection. Figure 4 presents the time series, power spectra, and phase portraits of the silicon-based epitaxial QDL under optical injection with Δωinj = −6.3 GHz for different injection strengths, where red and blue correspond to ignoring SRH recombination and τSRH = 0.5 ns, respectively. We firstly analyze the case that the effect of SRH recombination is ignored. For kinj = 0.1 (Figure 4(a1–a3)), the time series of the silicon-based epitaxial QDL exhibits periodic oscillations, and the fundamental frequency can be obtained in the power spectrum, which is about 14.5 GHz. At this time, the trajectory of the phase diagram is characterized by a clear limit ring. Therefore, it can be determined that the silicon-based epitaxial QDL operates at a period-one (P1) state. For kinj = 0.27 (Figure 4(b1–b3)), a periodic waveform with two peak intensities can be observed in the time series, and a subharmonic frequency of about 7.6 GHz appears in the power spectrum, and the corresponding phase diagram possesses two rings. As a result, the silicon-based epitaxial QDL exhibits a period-two (P2) state. For kinj = 0.5 (Figure 4(c1–c3)), similar to the analysis method of Figure 4(b1–b3), the laser also operates at a P2 state. For kinj = 1.1 (Figure 4(d1–d3)), multiple peaks with varying intensities emerge in the time series, new frequency components manifest in the power spectrum, and the phase diagram exhibits overlapping alternation of multiple loops. Hence, the dynamics of silicon-based epitaxial QDL can be characterized as a multi-period (MP) state. When kinj = 1.3 (Figure 4(e1–e3)), the peak intensity of the time series shows irregular fluctuation; at this point, the power spectrum broadens, and the phase diagrams display strange attractors. With these characteristics, we can determine that the dynamical state of silicon-based epitaxial QDL can be classified as the chaos (C) state. For kinj = 1.4 (Figure 4(f1–f3)), the dynamical state can be determined as MP, similar to Figure 4(d1–d3). For kinj = 2 (Figure 4(g1–g3)), the time series output is stable and there are no obvious peaks in the power spectrum, and the corresponding phase portrait is a stable point. Under this case, the laser frequency of the silicon-based epitaxial QDL is exactly equal to the frequency of the injected light. As a result, the silicon-based epitaxial QDL operates in the injection locking (IL) state. In the same way, as the above analyses, the dynamical states can be distinguished under different kinj when τSRH is equal to 0.5 ns. By comparing the dynamical states obtained under τSRH = 0.5 ns with those obtained under ignoring SRH recombination, it can be seen that for kinj taken as 0.5, 1, 1.3, 1.4, the obtained dynamical states are different. Therefore, we could conclude that different τSRH may cause different dynamical states in the silicon-based epitaxial QDLs under optical injection even if other parameters are the same. It should be pointed out that the determination of the dynamical state is referred to in Ref. [34].
Figure 5 gives the bifurcation diagram of the power extreme peak series of the silicon-based epitaxial QDL under optical injection with Δωinj = −6.3 GHz for τSRH taken as different values. For ignoring SRH recombination (as shown in Figure 5a), there exist two branches when kinj is varied within the range of (0.02~0.25), and, therefore, it can be judged that the silicon-based epitaxial QDL is working at P1 oscillation. For kinj located in the region of (0.25~1), there exist four branches, and it can be determined that the laser operates at P2 oscillation. For kinj located in the range of (1~1.1), the silicon-based epitaxial QDL exhibits an MP state. For kinj varied within (1.1~1.4), the silicon-based epitaxial QDL enters a C state due to coherent collapse. When kinj is within (1.4~1.75), the silicon-based epitaxial QDL enters the MP state again. Under a stronger injection strength, the laser will be locked. The dynamical evolution route of P1-P2-MP-C-MP-IL is observed. Similarly, the dynamical evolution paths for τSRH taken with other values can be obtained. As shown in Figure 5b–d, the dynamical evolution paths are P1-P2-MP-C-MP-IL, P1-P2-C-P2-MP-C-MP-IL, and P1-P2-C-P2-MP-C-MP-C-MP-IL, respectively.
Figure 6 displays the bifurcation diagram of the power extreme of peak series for the silicon-based epitaxial QDL under optical injection with Δωinj = −8 GHz for different τSRH. Via the same analysis as those in Figure 5, the dynamical evolution paths of Figure 6a–d are P1-P2-MP-C-MP-C-MP-IL, P1-P2-MP-C-MP-C-MP-IL, P1-P2-MP-P1-MP-IL, and P1-P2-MP-P2-MP-P2-MP-P1-MP-IL, respectively. Combining Figure 5 with Figure 6, it can be concluded that the laser has different evolution paths for different frequency detuning. However, some common characteristics can be observed. When τSRH is much larger than the spontaneous emission lifetime, the bifurcation diagram is similar to that obtained when ignoring the SRH recombination. When τSRH is less than or comparable to the spontaneous emission lifetime, the bifurcation diagram is different from that obtained when ignoring SRH recombination. Under this case, the SRH recombination cannot be not ignored.
The above result is based on the conditions of Δωinj = −6.3 GHz and Δωinj = −8 GHz. In order to systematically analyze the characteristics of silicon-based epitaxial QDL under optical injection, the distribution of the dynamical states of the laser in the parameter space of frequency detuning and injection strength are given in Figure 7, where the red dashed line indicates a frequency detuning of −6.3 GHz and the white dashed line indicates −8 GHz. As shown in this diagram, in the positive frequency detuning region, the C state is almost invisible, and the laser is directly transited from the P1 state to the IL state with the increase in kinj, and meanwhile, the dynamical states are much richer in the negative frequency detuning region. For τSRH = 5 ns (Figure 7b), the distribution of the dynamical state is almost the same as that obtained when ignoring the SRH recombination (Figure 7a). Therefore, the SRH recombination can be neglected when τSRH is much larger than the spontaneous emission lifetime. When τSRH and the spontaneous emission lifetime are comparable (Figure 7c,d), with the decrease in τSRH, the region for IL state shrinks, and meanwhile, the regions for C state and P2 state expand.

4. Conclusions

In summary, we have theoretically studied the nonlinear dynamical characteristics of a silicon-based epitaxial quantum dot laser (QDL) under optical injection after taking the influence of Shockley–Read–Hall (SRH) recombination into account. The results indicate that, for the silicon-based epitaxial QDL operating at free-running, its threshold increases with the decrease in the SRH recombination lifetime (τSRH). After introducing optical injection with different parameters, the silicon-based epitaxial QDL can exhibit a series of nonlinear dynamical states including period-one (P1), period-two (P2), multi-period (MP), chaos (C), and injection locking (IL). Through comparing the nonlinear dynamical state distributions of a silicon-based epitaxial QDL in the parameter space of injection strength and frequency detuning under different τSRH values, the influence of SRH recombination on the state of the laser is revealed. The results demonstrate that ignoring the SRH recombination may lead to an incorrect estimation of the laser state when τSRH is comparable to the spontaneous emission lifetime.

Author Contributions

Conceptualization, R.F. and Z.-M.W.; methodology, R.F.; validation, R.F., Y.-F.Z., G.-Q.X. and Z.-M.W.; investigation, R.F. and Q.-Q.W.; resources, G.-Q.X. and Z.-M.W.; data curation, R.F.; writing—original draft preparation, R.F.; writing—review and editing, G.-Q.X. and Z.-M.W.; visualization, R.F.; supervision, Z.-M.W.; project administration, G.-Q.X.; funding acquisition, G.-Q.X., Z.-M.W. and Y.-F.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (62335015, 61875167), the Chongqing Natural Science Foundation (CSTB2022NSCQ-MSX0313), and the Postgraduates’ Research and Innovation Project of Chongqing (CYB22111).

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lin, F.Y.; Tu, S.Y.; Huang, C.C.; Chang, S.M. Nonlinear dynamics of semiconductor lasers under repetitive optical pulse injection. IEEE J. Sel. Top. Quantum Electron. 2009, 15, 604–611. [Google Scholar] [CrossRef]
  2. Ohtsubo, J. Feedback induced instability and chaos in semiconductor lasers and their applications. Opt. Rev. 1999, 6, 1–15. [Google Scholar] [CrossRef]
  3. Abarbanel, H.D.I.; Kennel, M.B.; Illing, L.; Tang, S.; Chen, H.F.; Liu, J.M. Synchronization and communication using semiconductor lasers with optoelectronic feedback. IEEE J. Quantum Electron. 2001, 37, 1301–1311. [Google Scholar] [CrossRef]
  4. Jiang, N.; Xue, C.P.; Lv, Y.X.; Qiu, K. Physically enhanced secure wavelength division multiplexing chaos communication using multimode semiconductor lasers. Nonlinear Dyn. 2016, 86, 1937–1949. [Google Scholar] [CrossRef]
  5. Chang, D.; Zhong, Z.Q.; Valle, A.; Jin, W.; Jiang, S.; Tang, J.M.; Hong, Y.H. Microwave photonic signal generation in an optically injected discrete mode semiconductor laser. Photonics 2022, 9, 171. [Google Scholar] [CrossRef]
  6. Cui, C.; Fu, X.; Chan, S.C. Double-locked semiconductor laser for radio-over-fiber uplink transmission. Opt. Lett. 2009, 34, 3821–3823. [Google Scholar] [CrossRef]
  7. Romain, M.N.; Guy, V.; Jan, D.; Guy, V.S. Reducing the phase sensitivity of laser-based optical reservoir computing systems. Opt. Express 2016, 24, 1238–1252. [Google Scholar]
  8. Kawaguchi, Y.; Okuma, T.; Kanno, K.; Uchida, A. Entropy rate of chaos in an optically injected semiconductor laser for physical random number generation. Opt. Express 2021, 29, 2442–2457. [Google Scholar] [CrossRef]
  9. Zhang, L.; Pan, B.; Chen, G.; Guo, L.; Lu, D.; Zhao, L.; Wang, W. 640-Gbit/s fast physical random number generation using a broadband chaotic semiconductor laser. Sci. Rep. 2017, 7, 45900. [Google Scholar] [CrossRef] [PubMed]
  10. Bimberg, D.; Pohl, U.W. Quantum dots: Promises and accomplishments. Mater. Today 2011, 14, 388–397. [Google Scholar] [CrossRef]
  11. Liu, A.Y.; Komljenovic, T.; Davenport, M.L.; Gossard, A.C.; Bowers, J.E. Reflection sensitivity of 1.3 μm quantum dot lasers epitaxially grown on silicon. Opt. Express 2017, 25, 9535–9543. [Google Scholar] [CrossRef] [PubMed]
  12. Jin, Z.Y.; Huang, H.M.; Zhou, Y.G.; Zhao, S.Y.; Ding, S.H.; Wang, C.; Yao, Y.; Xu, X.C.; Grillot, F.; Duan, J.N. Reflection sensitivity of dual-state quantum dot lasers. Photonics Res. 2023, 11, 1713–1722. [Google Scholar] [CrossRef]
  13. Nishi, K.; Takemasa, K.; Sugawara, M.; Arakawa, Y. Development of quantum dot lasers for data-com and silicon photonics applications. IEEE J. Sel. Top. Quantum Electron. 2017, 23, 1–7. [Google Scholar] [CrossRef]
  14. Duan, J.N.; Zhou, Y.; Dong, B.Z.; Huang, H.M.; Norman, J.C.; Jung, D.; Zhang, Z.; Wang, C.; Bowers, J.E.; Grillot, F. Effect of p-doping on the intensity noise of epitaxial quantum dot lasers on silicon. Opt. Lett. 2020, 45, 4887–4890. [Google Scholar] [CrossRef] [PubMed]
  15. Ukhanov, A.A.; Stintz, A.; Eliseev, P.G.; Malloy, K.J. Comparison of the carrier induced refractive index, gain, and linewidth enhancement factor in quantum dot and quantum well lasers. Appl. Phys. Lett. 2004, 84, 1058–1060. [Google Scholar] [CrossRef]
  16. Ukhanov, A.A.; Wang, R.H.; Rotter, T.J.; Stintz, A.; Lester, L.F.; Eliseev, P.G.; Malloy, K.J. Orientation dependence of the optical properties in InAs quantum-dash lasers on InP. Appl. Phys. Lett. 2002, 81, 981–983. [Google Scholar] [CrossRef]
  17. Norman, J.C.; Jung, D.; Zhang, Z.; Wan, Y.; Liu, S.; Shang, C.; Herrick, R.W.; Chow, W.W.; Gossard, A.C.; Bowers, J.E. A review of high-performance quantum dot lasers on silicon. IEEE J. Quantum Electron. 2019, 55, 1–11. [Google Scholar] [CrossRef]
  18. Zheng, Y.F.; Xia, G.Q.; Lin, X.D.; Wang, Q.Q.; Wang, H.; Jiang, C.; Chen, H.; Wu, Z.M. Experimental investigation on the mode characteristics of an excited-state quantum dot laser under concave mirror optical feedback. Photonics 2023, 10, 166. [Google Scholar] [CrossRef]
  19. Yang, J.; Tang, M.; Chen, S.; Liu, H. From past to future: On-chip laser sources for photonic integrated circuits. Light Sci. Appl. 2023, 12, 16. [Google Scholar] [CrossRef]
  20. Tang, M.C.; Park, J.; Wang, Z.C.; Chen, S.; Jurczak, P.; Seeds, A.; Liu, H.Y. Integration of III-V lasers on Si for Si photonics. Prog. Quantum Electron. 2019, 66, 1–18. [Google Scholar] [CrossRef]
  21. Wan, Y.T.; Xiang, C.; Guo, J.; Koscica, R.; Kennedy, M.J.; Selvidge, J.; Zhang, Z.; Chang, L.; Xie, W.Q.; Huang, D.; et al. High speed evanescent quantum-dot lasers on Si. Laser Photonics Rev. 2021, 15, 2100057. [Google Scholar] [CrossRef]
  22. Wei, W.Q.; Huang, J.Z.; Ji, Z.T.; Han, D.; Yang, B.; Chen, J.J.; Qin, J.L.; Cui, Y.O.; Wang, Z.H.; Wang, T.; et al. Reliable InAs quantum dot lasers grown on patterned Si (001) substrate with embedded hollow structures assisted thermal stress relaxation. J. Phys. D Appl. Phys. 2022, 55, 40. [Google Scholar] [CrossRef]
  23. Liu, A.Y.; Chong, Z.; Norman, J.; Snyder, A.; Lubyshev, D.; Fastenau, J.M.; Liu, A.W.K.; Gossard, A.C.; Bowers, J.E. High performance continuous wave 1.3 µm quantum dot lasers on silicon. Appl. Phys. Lett. 2014, 104, 041104. [Google Scholar] [CrossRef]
  24. Liu, A.Y.; Bowers, J. Photonic integration with epitaxial III–V on silicon. IEEE J. Sel. Top. Quantum Electron. 2018, 24, 1–12. [Google Scholar] [CrossRef]
  25. Wei, W.Q.; Feng, Q.; Guo, J.J.; Guo, M.C.; Wang, J.H.; Wang, Z.H.; Wang, T.; Zhang, J.J. InAs/GaAs quantum dot narrow ridge lasers epitaxially grown on SOI substrates for silicon photonic integration. Opt. Express 2022, 28, 26555–26563. [Google Scholar] [CrossRef] [PubMed]
  26. Shang, C.; Wan, Y.T.; Norman, J.; Collins, N.; MacFarlane, I.; Dumont, M.; Liu, S.T.; Li, Q. Low-Threshold epitaxially grown 1.3-μm InAs quantum dot lasers on patterned (001) Si. IEEE J. Sel. Top. Quantum Electron. 2019, 25, 1–7. [Google Scholar] [CrossRef]
  27. Chen, S.; Li, W.; Wu, J.; Jiang, Q.; Tang, M.C.; Shutts, S.; Elliott, S.N.; Sobiesierski, A.; Seeds, A.J.; Ross, I.; et al. Electrically pumped continuous-wave III–V quantum dot lasers on silicon. Nat. Photonics 2016, 10, 307–311. [Google Scholar] [CrossRef]
  28. Zhao, S.Y.; Grillot, F. Effect of Shockley-Read-Hall recombination on the static and dynamical characteristics of epitaxial quantum-dot lasers on silicon. Phys. Rev. A 2021, 103, 9. [Google Scholar] [CrossRef]
  29. Chu, Q.; Zhao, S.Y.; Wang, J.W.; Sun, Y.X.; Yao, Y.; Xu, X.C.; Grillot, F.; Duan, J.N. Optical noise characteristics of injection-locked epitaxial quantum dot lasers on silicon. Opt. Express 2023, 31, 25177–25190. [Google Scholar] [CrossRef]
  30. Andre, C.; Boeckl, J.; Wilt, D.; Pitera, A.; Lee, M.L.; Fitzgerald, E.; Keyes, B.; Ringel, S. Impact of dislocations on minority carrier electron and hole lifetimes in GaAs grown on metamorphic SiGe substrates. Appl. Phys. Lett. 2004, 84, 3447–3449. [Google Scholar] [CrossRef]
  31. Saldutti, M.; Tibaldi, A.; Cappelluti, F.; Gioannini, M. Impact of carrier transport on the performance of QD lasers on silicon: A drift-diffusion approach. Photonics Res. 2020, 8, 1388–1397. [Google Scholar] [CrossRef]
  32. Duan, J.; Wang, X.G.; Zhou, Y.G.; Wang, C.; Grillot, F. Carrier-noise-enhanced relative intensity noise of quantum dot lasers. IEEE J. Quantum Electron. 2018, 54, 1–7. [Google Scholar] [CrossRef]
  33. Wang, C.; Zhuang, J.P.; Grillot, F.; Chan, S.C. Contribution of off-resonant states to the phase noise of quantum dot lasers. Opt. Express 2016, 24, 29872–29881. [Google Scholar] [CrossRef] [PubMed]
  34. Jiang, Z.F.; Wu, Z.M.; Jayaprasath, E.; Yang, W.Y.; Hu, C.X.; Xia, G.Q. Nonlinear dynamics of exclusive excited-state emission quantum dot lasers under optical injection. Photonics 2019, 6, 58. [Google Scholar] [CrossRef]
Figure 1. QD level structure and carrier dynamics of the three-level model.
Figure 1. QD level structure and carrier dynamics of the three-level model.
Photonics 11 00684 g001
Figure 2. Calculated SRH recombination lifetime as a function of TD density.
Figure 2. Calculated SRH recombination lifetime as a function of TD density.
Photonics 11 00684 g002
Figure 3. Dependence of output photon number on bias current under different values of τSRH.
Figure 3. Dependence of output photon number on bias current under different values of τSRH.
Photonics 11 00684 g003
Figure 4. Time series, power spectra, and phase portraits of the silicon-based epitaxial QDL under optical injection with Δωinj = −6.3 GHz and kinj = 0.1 (a1a6), 0.27 (b1b6), 0.5 (c1c6), 1 (d1d6), 1.3 (e1e6), 1.4 (f1f6), and 2 (g1g6), respectively. Red curves are for ignoring SRH recombination, and blue curves are for τSRH = 0.5 ns.
Figure 4. Time series, power spectra, and phase portraits of the silicon-based epitaxial QDL under optical injection with Δωinj = −6.3 GHz and kinj = 0.1 (a1a6), 0.27 (b1b6), 0.5 (c1c6), 1 (d1d6), 1.3 (e1e6), 1.4 (f1f6), and 2 (g1g6), respectively. Red curves are for ignoring SRH recombination, and blue curves are for τSRH = 0.5 ns.
Photonics 11 00684 g004
Figure 5. Bifurcation diagram of power extreme of peak series as a function of kinj for the silicon-based epitaxial QDL under optical injection with Δωinj = −6.3 GHz and ignoring SRH recombination (a), τSRH = 5 ns (b), τSRH = 1 ns (c), and τSRH = 0.5 ns (d).
Figure 5. Bifurcation diagram of power extreme of peak series as a function of kinj for the silicon-based epitaxial QDL under optical injection with Δωinj = −6.3 GHz and ignoring SRH recombination (a), τSRH = 5 ns (b), τSRH = 1 ns (c), and τSRH = 0.5 ns (d).
Photonics 11 00684 g005
Figure 6. Bifurcation diagram of the power extreme of peak series as a function of kinj for the silicon-based epitaxial QDL under optical injection with Δωinj = −8 GHz and ignoring SRH recombination (a), τSRH = 5 ns (b), τSRH = 1 ns (c), and τSRH = 0.5 ns (d).
Figure 6. Bifurcation diagram of the power extreme of peak series as a function of kinj for the silicon-based epitaxial QDL under optical injection with Δωinj = −8 GHz and ignoring SRH recombination (a), τSRH = 5 ns (b), τSRH = 1 ns (c), and τSRH = 0.5 ns (d).
Photonics 11 00684 g006
Figure 7. Mappings of the nonlinear dynamical states distribution of the silicon-based epitaxial QDL in the parameter space of injection strength and frequency detuning under ignoring SRH recombination (a), τSRH = 5 ns (b), τSRH = 1 ns (c), τSRH = 0.5 ns (d).
Figure 7. Mappings of the nonlinear dynamical states distribution of the silicon-based epitaxial QDL in the parameter space of injection strength and frequency detuning under ignoring SRH recombination (a), τSRH = 5 ns (b), τSRH = 1 ns (c), τSRH = 0.5 ns (d).
Photonics 11 00684 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fang, R.; Xia, G.-Q.; Zheng, Y.-F.; Wang, Q.-Q.; Wu, Z.-M. Nonlinear Dynamics of Silicon-Based Epitaxial Quantum Dot Lasers under Optical Injection. Photonics 2024, 11, 684. https://doi.org/10.3390/photonics11080684

AMA Style

Fang R, Xia G-Q, Zheng Y-F, Wang Q-Q, Wu Z-M. Nonlinear Dynamics of Silicon-Based Epitaxial Quantum Dot Lasers under Optical Injection. Photonics. 2024; 11(8):684. https://doi.org/10.3390/photonics11080684

Chicago/Turabian Style

Fang, Ruilin, Guang-Qiong Xia, Yan-Fei Zheng, Qing-Qing Wang, and Zheng-Mao Wu. 2024. "Nonlinear Dynamics of Silicon-Based Epitaxial Quantum Dot Lasers under Optical Injection" Photonics 11, no. 8: 684. https://doi.org/10.3390/photonics11080684

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop