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

Cavity enhanced nonlinear optics for few photon optical bistability

Open Access Open Access

Abstract

Weak material nonlinearity at optical frequencies poses a serious hurdle to realizing optical bistability at low optical powers, which is a critical component for digital optical computing. In this paper, we explore the cavity enhancement of the second-order optical nonlinearity in order to determine the feasibility of few photon optical bistability. Starting from a quantum optical formalism of a doubly resonant cavity (required to meet the condition of phase matching), we derive a dynamic classical model of a cavity that is bistable at the fundamental mode. We analyze the optical energy and the switching speed as a function of the cavity quality factors and mode volumes and identify the regime where only ten’s of photons are required to perform the switching. An unusual trend in the switching speed is also observed, where the speed monotonically decreases as the cavity linewidth increases. This is ascribed to the increase in the switching gain with increasing cavity linewidth.

© 2015 Optical Society of America

1. Introduction

It has long been a goal of optical physicists to utilize light to control the transmission of another optical signal, and thus constructing all optical logic gates [1]. One way to achieve such control is via optical bistability, in which two distinct output optical powers can be achieved for the same input power [2]. Thus, one can control the output power by transitioning between two stable states with a minimal change in the input signal. With an increasing trend of building optical interconnects for data transfer over short distances, one can envision that optical systems could be a new means of computation, rather than simply a channel for signal transportation [3, 4]. In such optical computing networks, a bistable optical device is an essential component.

Optical bistability can be achieved in several ways, including via thermo-optic effects [5], carrier injection [6], a combination of both [7], or via optoelectronic feedback [8]. Unfortunately, all these methods are inherently slow, as they implicitly or explicitly rely on carrier generation. Another way to achieve optical bistability is via nonlinear optical effects [9]. However, nonlinear optical effects in bulk materials are often very weak, requiring large optical powers (~mW) to observe the necessary bistability [10]. This increases the power consumption of such devices and thus, limits the technological practicality. While we can select different materials with greater nonlinear susceptibilities, we can also reduce the operating optical power by increasing the strength of the nonlinear light-matter interaction itself, e.g. by using high quality (Q) factor cavities with small mode volumes. However, the need for strong nonlinearity and possibility of fabricating very high Q-cavities poses an interesting trade-off. Silicon compatible materials (i.e. silicon or silicon nitride) have long formed the foundation of electronic devices and have recently emerged as a scalable technology for photonics [11,12]. However, these materials lack a second-order nonlinear susceptibility, which, when present, is often much stronger and subsequently, more effective in reducing the required optical input power. Silicon materials and their existing fabrication infrastructure could still form the basis for optically bistable devices if one can successfully incorporate strongly nonlinear materials with silicon photonic devices. In fact, several recently discovered materials, like transition metal dichalcogenides [13] and topological insulators [14], exhibit strong second-order nonlinear susceptibilities and can be easily integrated with silicon compatible platforms. The possibility of creating such hybrid systems consisting of strongly nonlinear materials integrated with a silicon compatible optical cavity, presents an opportunity to revisit the problem of low power optical bistability, based on second-order nonlinearity. In this paper, we seek to theoretically explore the lower limits of operating optical powers at which one can achieve optical bistability. Specifically, we analyze how the required optical power scales with cavity parameters as well as examine switching speeds by modeling the dynamics of the bistable switch.

2. Theoretical model

As explained before, in this paper, we limit our focus to second-order nonlinear optical effects. A similar model for third-order nonlinearity, where four wave mixing can give rise to optical bistability, is outlined in the appendix. The second-order nonlinear process relates two photons at the fundamental frequency to a photon at the second harmonic frequency. Since this device mediates an interaction between two different frequencies, one needs to ensure phase matching between the modes at both frequencies [15]. In a cavity structure, this equates to a non-zero spatial overlap between the cavity modes at the fundamental and the second harmonic frequencies [16,17]. However, as we show below, in the bistable device, both input and output light are at the fundamental mode frequency, while the second harmonic mode just mediates the nonlinear interaction. We emphasize that, this is fundamentally different from a third-order nonlinear cavity, where no such phase matching is required to observe optical bistability.

The dynamics of a nonlinear cavity with multiple resonances (at the fundamental frequency ωa and second harmonic frequency ωb) can be described by the Hamiltonian [18]:

H^s=ωaaa+ωbbb+gnl[b(a)2+ba2].

The coupling constant, gnl, can be expressed in terms of the classical electric fields as [18]:

gnl=Dε0(ωa2ε0)ωb2ε0drχ(2)(r)[ε(r)]3/2αa2(r)αb(r).

Here, a and b are the annihilation operators for the fundamental and second harmonic modes, respectively; D is a degeneracy factor, describing the number of terms in the χ(2) tensor that contribute to the nonlinearity; αa and αb are the normalized field profiles of the cavity modes such that the field-squared and integrated over the whole volume is unity, i.e., |αa,b|2dr=1. The volume integration in the expression of gnl, however, should be performed only over the space where there is a nonlinear material present. Such distinction is particularly important in a hybrid system, where a transition metal dichalcogenide or topological insulator are placed on top of an otherwise linear silicon or silicon nitride cavity, e.g., see Fig. 1(d). Assuming ωb = 2ωa, D = 2 and a perfect overlap between the cavity modes (αa(r) = αb(r) = α(r)), we find that the nonlinear interaction term reduces to

gnl=ε0(ωaε0εr)3/2χ(2)Vm
where, 1/Vm=NLα3(r)dr. Note that, the phase matching condition is implicit in the assumption of perfect overlap between the cavity modes. The input light can be modeled as driving the fundamental mode, a, by adding the term 2κraE(eiωlta+eiωlta) to the Hamiltonian. E denotes the amplitude of the electric field of the driving laser, (ωl is the laser frequency, and κra is the incoupling rate of the cavity mode a associated with the reflectance from the cavity. By transforming to a rotating frame (as the optical frequencies are much larger that the cavity loss rate or the nonlinear interaction strength) [19], we can write
Hrot=Δaaa+Δbbb+gnl[b(a)2+ba2]+2κraE(a+a).

 figure: Fig. 1

Fig. 1 (a) Schematic of a cavity with two modes with nonlinear interaction gnl. The losses in the cavity primarily arise from the leak rates in the reflection (κra) and transmission (κta) ports. There can be additional absorptive loss, which is neglected in our analysis. (b) At steady state, the output optical power is a bistable function of the input optical power. In practice, however, one observes a hysteretic behavior, depending on whether the input power is increased or decreased. The parameters for the simulation are: gnl/2π = 20 GHz; κta/2π = κra/2π = 3 GHz and Δa/κa = 8. (c) and (d) show different geometry of the nonlinear cavity: the cavity can itself be made of nonlinear material (c), or the cavity can be linear unto second-order, i.e., no χ(2) nonlinearity, but with a nonlinear material (e.g. 2D material or topological insulator) transferred on top of the cavity (d).

Download Full Size | PDF

In this representation, Δa,b describes the detuning of each cavity mode from the laser frequency. With a Hamiltonian, however, we can study only the lossless evolution of the system, whereas, a realistic system is lossy. All the cavity losses (from mirror losses and material absorption) can be incorporated by including the Lindblad terms in the Master equation describing the evolution of the density matrix, ρ of the double cavity system, as [20]:

dρdt=i[Hrot,ρ]+i=a,bκi[2AiρAiAiAiρρAiAi].

Here Ai represents the annihilation operators for either mode, a or b. Note that, for each of these cavity modes, we have assumed there are three loss channels: reflection, transmission and absorption, with field decay rates denoted as κra,rb, κta,tb and κla,lb, respectively (Fig. 1(a)). The total photon-loss rate for each cavity mode is then given by κa,b = κta,tb + κra,rb + κla,lb. Using the relation dAidt=Tr[Aidρdt], we derive the mean-field equations for the cavity fields:

dadt=iΔaa(κra+κta+κla)a2ignlba+i2κraE,
dbdt=iΔbb(κrb+κtb+κlb)bignla2.

Where 〈Ai〉 is a complex number denoting the expectation value of the operator Ai. One can numerically solve these mean-field equations to analyze the behavior of the nonlinear cavity. We however want to first understand the condition for optical bistability in the nonlinear system at the steady state. If ωb = 2ωa, then the rotating frame implies that Δb = 2Δa [18]. Furthermore, for the sake of simplicity we also assume that the quality factors of our cavity modes are identical at frequencies ωa and ωb leading to κb = 2κa. Therefore, in the steady state, by eliminating the mode b (from Eqn. 6 we can derive:

η2Ptrans3+2η(κa2Δa2)Ptrans2+(Δa2+κa2)2Ptrans=4κtaκra(Δa2+κa2)Pin,
where η=gnl2/2κta, Ptrans = 2κtaaa〉 is the transmitted power from the fundamental cavity mode, Pin = E2 is the input optical power and the intra-cavity photon number is given by Nc = 〈aa〉. The assumption of Δbb = 2(Δaa) is however difficult to achieve in the current technology and the effect of mismatch is analyzed in the appendix. However, the overall behavior of the bistable system does not change significantly when this condition is not satisfied. Note that Ptrans, as mentioned so far, does not have units of power, but rather photons per second. To obtain the actual power, we need to multiply Ptrans and Pin by the corresponding photon energy ħωa. Eq. (8), which describes the steady state behavior of the fundamental cavity mode, is a cubic equation (when η ≠ 0) and exhibits bistable behavior as long as |Δa|(2+3)κa (see Appendix), irrespective of the magnitude of η determined by the nonlinear interaction strength. However, η determines the input power at which the bistability appears. Simple inspection of Eq. (8) shows that by increasing η (e.g. by increasing the nonlinear coupling strength or by decreasing the transmission of the fundamental mode), one can decrease the required optical power by several orders of magnitude. We emphasize that η can also be interpreted as the strength of the positive feedback to observe the bistability, as explained in Ref. [8].

In a realistic dielectric cavity, the cavity loss primarily arises from loss in the transmission and reflection ports and one can neglect the absorptive loss. Moreover, the transmission and reflection ports are generally equivalent leading to κta = κra = κa/2. Under these assumptions we can simplify Eq. (8) to

η2Ptrans3+2η(κa2Δa2)Ptrans2+(Δa2+κa2)2Ptrans=κ2a(Δa2+κa2)Pin.

This assumption is used for all the numerical simulations present in the paper.

As previously discussed, optically bistable devices allow for large swings in the output power with slight changes to the input power. The steady state behavior of the input and output power as predicted by Eq. (8) shows two different values of Ptrans for the same Pin over a range of inout powers. In practice, however, such steady-state behavior is never observed. The bistable behavior manifests itself by a sudden jump of Ptrans between the two steady-state values, with slight changes in Pin. Depending on whether Pin is increasing or decreasing, we observe different behavior in the output power, characteristic of hysteretic behavior and a bistable system (see Fig. 1(b)). The critical points of the bistable system, where we observe sudden changes in output power from slight changes in the input power can be found by solving dPin/dPtrans = 0. These critical points are important to qualitatively understand the scaling of the optical power for the bistable system. To observe optical bistability, one typically requires a large laser detuning, relative to the loss rate. Thus, when Δaκa, we yield critical points at Ptrans=Δa2/η and Δa2/3η corresponding to the input powers Pin = 0 and 4Δa4/27ηκa2. This indicates that the signal swing is ~κa2/Δa2. From the expression of the critical input power 4Δa4/27ηκa2, we can clearly see that with an increasing η, the required power to observe bistability decreases, as pointed out earlier in the paper. The critical input power also scales with the detuning and cavity loss as ~Δa4/κa2. From our previous analysis, we also found out that to observe the bistability, the condition |Δa|>(2+3)κa needs to be satisfied. Hence, the input optical power where one can observe optical bistability scales as ~κa3/gnl2~Vm/Q3. We emphasize that, such scaling is more favorable for lowering the optical power by using a high-Q cavity, compared to other cavity-based bistable devices employing photo-refractive or thermo-optic effects, where the input power typically scales as ~ Vm/Q [6].

3. Performance of the switch

Using this model, we now analyze the performance of an optical switch based on the bistable device. The input signal consists of a fixed bias power Pbias and a modulated optical signal. For the best operation, Pbias should be in the grey region in Fig. 2(a). The modulated input signal is changed sinusoidally with a frequency Ω and amplitude, Pamp. This will modulate the output power with the same frequency, but a different amplitude. We define the ratio of the output and input power amplitude as the gain, G, of the switch. As discussed earlier, the output signal swing is proportional to κa2/Δa2, and for all subsequent analysis, we will let Δa/κa = 8. We analyze G as a function of Pbias and Pamp for a modulation frequency of Ω/2π = 500 MHz and find that the gain increases with decreasing Pamp (Fig. 2(b)). A bistable system can only swing between its two steady states, and by increasing Pamp, the output swing cannot be increased indefinitely. Hence a reduced Pamp increases the gain. At a very low value of Pamp, however, the system cannot switch, and the output becomes distorted (not shown here). Although, a large gain is desired for a good optical switch, we also want to have a large range of bias-points, i.e. absence of critical biasing [3]. This means we should have a range of Pbias, where a high gain can be achieved. Figure 2(b) clearly shows that this range decreases with decreasing Pamp. Hence to avoid critical biasing, one needs a Pamp that is significantly larger than the absolute minimum Pamp needed to switch between the two states.

 figure: Fig. 2

Fig. 2 (a) The steady state bistability plot is used to identify the bias-points, around which one can modulate the input power to observe the change in the output power. (b) The ratio between the output power amplitude and input power amplitude as a function of Pbias and Pamp. (c) The frequency response for two different Pamp, showing the bandwidth changes depending on the amplitude. (d) Gain, defined as the switching ratio at a low frequency, (e) Bandwidth, defined at the 3 dB point, is plotted as a function of the Pamp. The parameters for the simulation are: gnl/2π = 20 GHz; κta/2π = κra/2π = 3 GHz and ∆a/κa = 8.

Download Full Size | PDF

By changing the modulation frequency, Ω, we find that the switch behaves as a low-pass filter (Fig. 2(c)). However, the frequency response depends on the input amplitude Pamp. To understand such dependency, we analyze the frequency response of the switch as a function of the input amplitude Pamp keeping all other system parameters fixed (Figs. 2(d), and 2(e)). We find that with increasing Pamp, the gain decreases while the bandwidth increases. The change in bandwidth can be explained qualitatively via the constancy of the gain-bandwidth product. Next, we analyze the dependence of the input optical power on the cavity loss rate κa and nonlinear interaction strength gnl. For different system parameters, we calculate the input optical power Pbias where we observe the optical bistability. Figures 3(a) and 3(b) show the average Pbias and intra-cavity photon number N = 〈aa〉, respectively, as a function of κa and gnl. As previously derived, the scaling of input optical power should follow Vm/Q3~κa3/gnl2. However, the intra-cavity photon number N scales κa2/gnl2, as the photon number depends on the cavity loss rate κa. From these figures, we identify the parameter range (the region under the white line in Fig. 3(b)) where the photon number, N, is less than 10. With a nonlinear interaction strength of gnl/ = 20 GHz, and loss rate κa/ = 10 GHz, one can reach the N ~ 10 photons regime. For a target wavelength of ~ 1.5μm, this entails a quality factor of ~ 20,000, which is readily achievable in state-of-the-art cavities. To achieve gnl/ = 20 GHz using a dielectric cavity with mode-volume Vm ~ (λ/n)3, n being the refractive index of the cavity material, at a typical wavelength of λ ~ 1.5μm, the nonlinear second-order coefficient χ(2) would need to be ~ 5 nm/V. This is more than one order of magnitude larger than available bulk materials. To reach the necessary nonlinearity one either needs to use a material with larger optical nonlinearity or use a cavity with small mode volume. Although metallic structures can provide a small mode volume, the excess loss due to the metal is detrimental for the ultimate performance. Hence, a better avenue will be to use new materials, like 2D materials [21], perovskites [22], or topological insulators [14] instead. Another approach could be to compensate for the reduced nonlinear interaction strength by using a high-quality factor cavity. For example, one can reach the ten-photon regime with gnl/ = 1 GHz and a cavity quality factor of ~ 200000. Such a high quality factor is difficult to achieve in III–V materials, but can be reached in silicon cavities. This emphasizes the necessity of building the second-order nonlinear hybrid platform, as previously discussed.

 figure: Fig. 3

Fig. 3 (a) The bias point Pbias as a function of the total linewidth 2κa and nonlinear interaction strength gnl. (b) log10(N), N being the intra-cavity photon number plotted as a function of κa and gnl. (c) Gain as a function of κa for different gnl. (d) Bandwidth as a function of κ for different gnl.

Download Full Size | PDF

Using these bias-points, we calculate the frequency response of the bistable system. Specifically, we analyze the gain at low modulation frequency as well as the bandwidth of the switch. Figure 3(c) shows the gain as a function of the linewidth 2κa for different gnl values. A clear increase in the gain is observed with κa as can be explained from the simple model of gain ~κa2. From the constancy of the gain-bandwidth product, we expect the bandwidth to decrease with increasing κa (Fig. 3(d)). Note that such behavior is unusual for cavity based devices, where the speed is largely limited by the cavity lifetime. In those devices the speed generally increases with κa. Such observation is critical, as high-Q cavities are generally considered bad for high speed operation even though they are beneficial from the stand-point of power. However, for optically bistable devices, we find that increased quality factors improve the power consumption while maintaining high speed operation. The use of high-Q cavities is, however, not without drawbacks as it still poses problems for thermal stability, as well as limiting gain in the switching operation.

4. Conclusion

We have analyzed the performance of an optically bistable system based on second-order nonlinearity in terms of the power and the speed of operation via a simple dynamic classical model. We found that with high quality-factor and low mode volume Vm, one can push the energy required to observe bistability to the few photon level. The energy scales as Vm/Q3. Temporal analysis of the bistable switch also reveals that the speed of operation decreases with an increase in cavity linewidth, in contrast to other cavity-based devices. Such low power nonlinear operation will be useful to implement optical computing systems.

Appendix

4.1. Detailed derivation of the bistable system with χ(2) nonlinearity

Starting from the dynamic equations in the main text

dadt=iΔaa(κra+κta+κla)a2ignlba+i2κraE,dbdt=iΔbb(κrb+κtb+κlb)bignla2.
we find that at the steady-state,
b=ignla2iΔb(κrb+κtb+κlb)

Using this we can write the equation for operator a (at steady state) as:

iΔaa(κra+κta+κla)a+2gnl2iΔb(κrb+κtb+κlb)aa2+i2κraE=0

This can be rewritten as

iΔaaκaa+2gnl2iΔbκbPtrans2κtaa+i2κraE=0

Assuming 2(iΔaκa) = iΔbκb, we can write

η2Ptrans3+2η(κa2Δa2)Ptrans2+(Δa2+κa2)Ptrans=4κtaκra(Δa2+κa2)Pin,

The critical points are given by

dPindPtrans=3η2Ptrans2+4η(κa2Δa2)Ptrans+(κa2+Δa2)2=0

The system is bistable as long as 16η2(κa2Δa2)212η2(κa2+Δa2)2=4η2(κa4+Δa414κa2Δa2)>0. This condition can be simplified to the criteria: |Δa|<(2+3)κa. The critical points are given by

Ptranscr=2(Δa2κa2)±κa4+Δa414κa2Δa23η

These critical points are involved and provide little intuition. To qualitatively understand the system behavior, we assume lossless cavity with κta = κra = κa/2and Δa >> κa and simplify the equation as:

η2Ptrans32ηΔa2Ptrans2+Δa4Ptrans=κa2Δa2Pin

Here the critical points are the roots of

3η2Ptrans24ηΔa2Ptrans+Δa4=0

From this we calculate the critical points as previously reported.

4.2. Optical bistability in a cavity with third order nonlinear material

We assume a third order nonlinear cavity driven by an external laser. The Hamiltonian of the driven system is given by

=Δaa+χaaa+2κrE(a+a)

From the master equation (with losses added via the Lindblad formalism), the mean-field equations for the cavity annihilation operator a as:

dadt=Tr[adρdt]=iΔa(κr+κt+κl)a+2iχaaa+i2κrE

The different output powers are given by Ptrans = 2κtaa; Ploss = 2κlaa and Prefl = 2κraa. At steady state we can write

iΔa(κr+κt+κl)a+2iχaaa+i2κrE=0

This can be simplified to

a=i2κrEi(Δ+χPtransκt)(κr+κt+κl)

The transmitted power at steady state becomes:

Ptrans=4κtκr|E|2(Δ+χPtransκt)2+(κr+κt+κl)2

Denoting η = χ/κt, the equation can be written as

η2Ptrans3+2ΔηPtrans2+Ptrans(Δ2+(κr+κt+κl)2)4κrκtPin=0

The critical points can be found out by solving

dPindPtrans=3η2Ptrans2+4ΔηPtrans+Δ2+(κr+κt+κl)2=0
and critical points are
Ptranscr=2Δ±Δ23κ23η
Ptranscr=2Δ(Δ2+9κ2)2(Δ23κ2)3108ηκtκr

4.3. Effect of different quality factors in χ(2) nonlinear cavities

In the main text of the paper we assumed that Δbb = 2(Δaa), which is based on the assumptions that the mode near second harmonic frequency is exactly twice the fundamental frequency and the Q-factor of both cavities are exactly same. However, such assumptions do not always hold true. Without such an assumption the governing equation looks like:

4η2Ptrans3+4η(κaκbΔaΔb)Ptrans2+(κa2+Δa2)(κb2+Δb2)Ptrans=4κtaκra(Δb2+κb2)Pin

Using similar procedures as explained before, the critical points are given by

Ptranscr=κaκb6η[(ΔaκaΔbκb1)±(1ΔaκaΔbκb)23(Δaκa+Δbκb)2]

We note that, to observe the bistability, one needs to satisfy the condition,

|ΔaκaΔbκb1|3|Δaκa+Δbκb|

Note that, even if a condition Δbb = 2(Δa − iκa) is difficult to satisfy one can easily achieve ΔaκaΔbκb. This means, the bistability condition becomes

ΔaκaΔbκb2+3

Hence, we find that even if the quality factor of the second harmonic mode can be worse than the fundamental mode, as long as similar ratio between the detuning and the linewidth for the two modes. Under the assumptions Δaκa, Δbκb and κra = κta = κa/2, the critical points are Ptrans = ΔaΔb/6η and ΔaΔb/2η corresponding to the input powers of Pin = 0 and 2Δa3Δb/27ηκa2. Hence the input power scales as Pin~Δa3Δb/ηκa2. This shows that, the input power increases with a lower quality factor of the second harmonic mode, as it depends on Δb.

Acknowledgments

This work is supported by the National Science Foundation under grant NSF-EFRI-1433496 and the Air Force Office of Scientific Research under grant FA9550-15-1-0150. CD is supported by the startup fund provided by University of Washington, Seattle.

References and links

1. A. Sawchuk and T. C. Strand, “Digital optical computing,” Proceedings of the IEEE 72, 758–779 (1984). [CrossRef]  

2. H. Gibbs, Optical Bistability: Controlling Light with Light (Academic Press, 1985).

3. D. A. B. Miller, “Are optical transistors the logical next step?” Nature Photonics 4, 3–5 (2010). [CrossRef]  

4. H. J. Caulfield and S. Dolev, “Why future supercomputing requires optics,” Nature Photonics 4, 261–263 (2014). [CrossRef]  

5. V. R. Almeida and M. Lipson, “Optical bistability on a silicon chip,” Opt. Lett. 29, 2387–2389 (2004). [CrossRef]   [PubMed]  

6. K. Nozaki, A. Shinya, S. Matsuo, Y. Suzaki, T. Segawa, T. Sato, Y. Kawaguchi, R. Takahashi, and M. Notomi, “Ultralow-power all-optical ram based on nanocavities,” Nature Photonics 6, 248252 (2012). [CrossRef]  

7. M. Sodagar, M. Miri, A. A. Eftekhar, and A. Adibi, “Optical bistability in a one-dimensional photonic crystal resonator using a reverse-biased pn-junction,” Opt. Express 23, 2676–2685 (2015). [CrossRef]   [PubMed]  

8. A. Majumdar and A. Rundquist, “Cavity-enabled self-electro-optic bistability in silicon photonics,” Opt. Lett. 39, 3864–3867 (2014). [CrossRef]   [PubMed]  

9. Y.-D. Kwon, M. A. Armen, and H. Mabuchi, “Femtojoule-scale all-optical latching and modulation via cavity nonlinear optics,” Phys. Rev. Lett. 111, 203002 (2013). [CrossRef]   [PubMed]  

10. M. Soljačić, E. Lidorikis, M. Ibanescu, S. G. Johnson, J. Joannopoulos, and Y. Fink, “Optical bistability and cutoff solitons in photonic bandgap fibers,” Opt. Express 12, 1518–1527 (2004). [CrossRef]  

11. M. Hochberg and T. Baehr-Jones, “Towards fabless silicon photonics,” Nature Photonics 4, 492–494 (2010). [CrossRef]  

12. D. J. Moss, R. Morandotti, A. L. Gaeta, and M. Lipson, “New cmos-compatible platforms based on silicon nitride and hydex for nonlinear optics,” Nature Photonics 7, 597607 (2013). [CrossRef]  

13. L. M. Malard, T. V. Alencar, A. P. M. Barboza, K. F. Mak, and A. M. de Paula, “Observation of intense second harmonic generation from mos2 atomic crystals,” Phys. Rev. B 87, 201401 (2013). [CrossRef]  

14. J. Wang, H. Mabuchi, and X.-L. Qi, “Calculation of divergent photon absorption in ultrathin films of a topological insulator,” Phys. Rev. B 88, 195127 (2013). [CrossRef]  

15. R. Boyd, Nonlinear optics (Academic Press, 2008).

16. K. Rivoire, Z. Lin, F. Hatami, W. T. Masselink, and J. Vučković, “Second harmonic generation in gallium phosphide photonic crystal nanocavities with ultralow continuous wave pump power,” Opt. Express 17, 22609–22615 (2009). [CrossRef]  

17. S. Buckley, M. Radulaski, J. L. Zhang, J. Petykiewicz, K. Biermann, and J. Vučković, “Multimode nanobeam cavities for nonlinear optics: high quality resonances separated by an octave,” Opt. Express 22, 26498–26509 (2014). [CrossRef]   [PubMed]  

18. A. Majumdar and D. Gerace, “Single-photon blockade in doubly resonant nanocavities with second-order non-linearity,” Physical Review B 87235319 (2013). [CrossRef]  

19. A. Majumdar, M. Bajcsy, D. Englund, and J. Vuckovic, “All optical switching with a single quantum dot strongly coupled to a photonic crystal cavity,” Selected Topics in Quantum Electronics, IEEE Journal of 18, 1812–1817 (2012). [CrossRef]  

20. C. W. Gardiner and P. Zoller, Quantum Noise (Springer-Verlag, 2005).

21. C. Janisch, Y. Wang, D. Ma, N. Mehta, A. L. Elias, N. Perea-Lopez, M. Terrones, V. Crespi, and Z. Liu, “Extraordinary second harmonic generation in tungsten disulfide monolayers,” Scientific Reports 45530 (2010).

22. X. Wan, J. Dong, M. Qian, and W. Zhang, “Nonlinear optical properties of perovskite ymno3 studied by real-space recursion method,” Phys. Rev. B 61, 10664–10669 (2000). [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 (3)

Fig. 1
Fig. 1 (a) Schematic of a cavity with two modes with nonlinear interaction gnl. The losses in the cavity primarily arise from the leak rates in the reflection (κra) and transmission (κta) ports. There can be additional absorptive loss, which is neglected in our analysis. (b) At steady state, the output optical power is a bistable function of the input optical power. In practice, however, one observes a hysteretic behavior, depending on whether the input power is increased or decreased. The parameters for the simulation are: gnl/2π = 20 GHz; κta/2π = κra/2π = 3 GHz and Δa/κa = 8. (c) and (d) show different geometry of the nonlinear cavity: the cavity can itself be made of nonlinear material (c), or the cavity can be linear unto second-order, i.e., no χ(2) nonlinearity, but with a nonlinear material (e.g. 2D material or topological insulator) transferred on top of the cavity (d).
Fig. 2
Fig. 2 (a) The steady state bistability plot is used to identify the bias-points, around which one can modulate the input power to observe the change in the output power. (b) The ratio between the output power amplitude and input power amplitude as a function of Pbias and Pamp. (c) The frequency response for two different Pamp, showing the bandwidth changes depending on the amplitude. (d) Gain, defined as the switching ratio at a low frequency, (e) Bandwidth, defined at the 3 dB point, is plotted as a function of the Pamp. The parameters for the simulation are: gnl/2π = 20 GHz; κta/2π = κra/2π = 3 GHz and ∆a/κa = 8.
Fig. 3
Fig. 3 (a) The bias point Pbias as a function of the total linewidth 2κa and nonlinear interaction strength gnl. (b) log10(N), N being the intra-cavity photon number plotted as a function of κa and gnl. (c) Gain as a function of κa for different gnl. (d) Bandwidth as a function of κ for different gnl.

Equations (31)

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

H ^ s = ω a a a + ω b b b + g n l [ b ( a ) 2 + b a 2 ] .
g n l = D ε 0 ( ω a 2 ε 0 ) ω b 2 ε 0 d r χ ( 2 ) ( r ) [ ε ( r ) ] 3 / 2 α a 2 ( r ) α b ( r ) .
g n l = ε 0 ( ω a ε 0 ε r ) 3 / 2 χ ( 2 ) V m
H r o t = Δ a a a + Δ b b b + g n l [ b ( a ) 2 + b a 2 ] + 2 κ r a E ( a + a ) .
d ρ d t = i [ H r o t , ρ ] + i = a , b κ i [ 2 A i ρ A i A i A i ρ ρ A i A i ] .
d a d t = i Δ a a ( κ r a + κ t a + κ l a ) a 2 i g n l b a + i 2 κ r a E ,
d b d t = i Δ b b ( κ r b + κ t b + κ l b ) b i g n l a 2 .
η 2 P t r a n s 3 + 2 η ( κ a 2 Δ a 2 ) P t r a n s 2 + ( Δ a 2 + κ a 2 ) 2 P t r a n s = 4 κ t a κ r a ( Δ a 2 + κ a 2 ) P i n ,
η 2 P t r a n s 3 + 2 η ( κ a 2 Δ a 2 ) P t r a n s 2 + ( Δ a 2 + κ a 2 ) 2 P t r a n s = κ 2 a ( Δ a 2 + κ a 2 ) P i n .
d a d t = i Δ a a ( κ r a + κ t a + κ l a ) a 2 i g n l b a + i 2 κ r a E , d b d t = i Δ b b ( κ r b + κ t b + κ l b ) b i g n l a 2 .
b = i g n l a 2 i Δ b ( κ r b + κ t b + κ l b )
i Δ a a ( κ r a + κ t a + κ l a ) a + 2 g n l 2 i Δ b ( κ r b + κ t b + κ l b ) a a 2 + i 2 κ r a E = 0
i Δ a a κ a a + 2 g n l 2 i Δ b κ b P t r a n s 2 κ t a a + i 2 κ r a E = 0
η 2 P t r a n s 3 + 2 η ( κ a 2 Δ a 2 ) P t r a n s 2 + ( Δ a 2 + κ a 2 ) P t r a n s = 4 κ t a κ r a ( Δ a 2 + κ a 2 ) P i n ,
d P i n d P t r a n s = 3 η 2 P t r a n s 2 + 4 η ( κ a 2 Δ a 2 ) P t r a n s + ( κ a 2 + Δ a 2 ) 2 = 0
P t r a n s c r = 2 ( Δ a 2 κ a 2 ) ± κ a 4 + Δ a 4 14 κ a 2 Δ a 2 3 η
η 2 P t r a n s 3 2 η Δ a 2 P t r a n s 2 + Δ a 4 P t r a n s = κ a 2 Δ a 2 P i n
3 η 2 P t r a n s 2 4 η Δ a 2 P t r a n s + Δ a 4 = 0
= Δ a a + χ a a a + 2 κ r E ( a + a )
d a d t = T r [ a d ρ d t ] = i Δ a ( κ r + κ t + κ l ) a + 2 i χ a a a + i 2 κ r E
i Δ a ( κ r + κ t + κ l ) a + 2 i χ a a a + i 2 κ r E = 0
a = i 2 κ r E i ( Δ + χ P t r a n s κ t ) ( κ r + κ t + κ l )
P t r a n s = 4 κ t κ r | E | 2 ( Δ + χ P t r a n s κ t ) 2 + ( κ r + κ t + κ l ) 2
η 2 P t r a n s 3 + 2 Δ η P t r a n s 2 + P t r a n s ( Δ 2 + ( κ r + κ t + κ l ) 2 ) 4 κ r κ t P i n = 0
d P i n d P t r a n s = 3 η 2 P t r a n s 2 + 4 Δ η P t r a n s + Δ 2 + ( κ r + κ t + κ l ) 2 = 0
P t r a n s c r = 2 Δ ± Δ 2 3 κ 2 3 η
P t r a n s c r = 2 Δ ( Δ 2 + 9 κ 2 ) 2 ( Δ 2 3 κ 2 ) 3 108 η κ t κ r
4 η 2 P t r a n s 3 + 4 η ( κ a κ b Δ a Δ b ) P t r a n s 2 + ( κ a 2 + Δ a 2 ) ( κ b 2 + Δ b 2 ) P t r a n s = 4 κ t a κ r a ( Δ b 2 + κ b 2 ) P i n
P t r a n s c r = κ a κ b 6 η [ ( Δ a κ a Δ b κ b 1 ) ± ( 1 Δ a κ a Δ b κ b ) 2 3 ( Δ a κ a + Δ b κ b ) 2 ]
| Δ a κ a Δ b κ b 1 | 3 | Δ a κ a + Δ b κ b |
Δ a κ a Δ b κ b 2 + 3
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.