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

Raman silicon nanocavity laser with efficient light emission from the edge of an adjacent waveguide

Open Access Open Access

Abstract

A Raman nanocavity laser can emit light into free space and into a properly designed waveguide adjacent to the cavity by mode coupling. In common device designs, the emission from the edge of this waveguide is relatively weak. However, a Raman silicon nanocavity laser with strong emission from the waveguide edge would be advantageous for certain applications. Here we investigate the increase in the edge emission that can be achieved by adding photonic mirrors to the waveguides adjacent to the nanocavity. We experimentally compare devices with and without photonic mirrors: the edge emission for devices with mirrors is 4.3 times stronger on average. This increase is analyzed using coupled-mode theory. The results indicate that the control of the round-trip phase shift (between the nanocavity and the mirror) and an increase of the quality factors of the nanocavity are important for further enhancement.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Integrated optical circuits based on silicon that combine thin waveguides, small resonators, short optical modulators and photo-detectors are expected to be useful in information processing [1]. For example, optical transceivers for short-distance optical communication [2], all-optical routing for efficient high-speed networks [3], and quantum optical signal processing for large-scale quantum computing [4] have been intensively studied. Silicon-based laser sources using stimulated Raman scattering (SRS) and optical amplifiers are also attractive elements, because they can enhance the functionality of such optical circuits [513].

Raman silicon lasers that are based on high-quality factor (high-Q) photonic crystal (PC) nanocavities have resonator lengths on the order of several micrometers and lasing thresholds on the order of one microwatt [14], and it has been shown that continuous-wave laser operation with an excitation wavelength in the S-band, the O-band, or the T-band is also possible [15,16]. Moreover, the fabrication of such lasers by a CMOS-compatible approach has been demonstrated [17]. Such lasers can even be used to detect ionized air [18,19]. However, the device design used in these previous studies has a significant limitation: the power emitted from the edge of an adjacent waveguide (Pedge) is several times smaller than the surface emission power from the nanocavity (Psurface) [20]. On the other hand, for the applications described above, edge-emitting Raman silicon lasers can often be advantageous.

It has been shown that photonic mirrors that reflect photons by a mode-gap difference can be used to efficiently emit photons from the high-Q nanocavity to an adjacent waveguide [21]. Efficient edge emission has been demonstrated using devices that consist of a nanocavity, an input waveguide with a side mirror, and an output waveguide with another mirror [22,23]. Photonic mirrors have also been utilized in studies related to optical buffer memories for silicon optical circuits [2426]. In these previous studies, a single high-Q resonant mode in the nanocavity was used, and thus the input waveguide and the output waveguide had the same design. In contrast, a Raman laser uses two high-Q resonant modes (to confine the pump photons and the Stokes Raman-scattered photons). Since the wavelength of the Stokes photons is about 100 nm longer than that of the pump photons, the design of the output waveguide with the photonic mirror needs to be different from that of the input waveguide. Numerical simulations for a Raman laser with such waveguides have indicated a possible enhancement of Pedge [27].

Here, we experimentally and numerically investigate the increase in Pedge of a Raman silicon nanocavity laser that can be achieved by introducing heterointerface mirrors (HMs). We call this type of laser a HM Raman laser. In Section 2, the device configuration of our laser is explained. In Section 3, experimental results for HM and conventional Raman lasers are compared; the average Pedge of the HM Raman lasers is 4.3 times higher. Section 4 describes the used calculation framework, and the calculation results are described in Section 5. This numerical analysis can explain the experimentally observed increase in Pedge. In Section 6, we discuss how Pedge can be further increased.

2. Sample structure

Figure 1 clarifies the HM Raman laser device design. We consider the use of a silicon-on-insulator (SOI) substrate where the thickness of the top silicon layer is 220 nm and that of the buried oxide (BOX) layer is 3 µm. Figures 1(a) and (c) show the top view of the silicon layer, and the x-axis is along the [100] direction of the silicon layer. The basic PC consists of air holes arranged in a triangular lattice with a lattice constant (a) of 410 nm. The air hole radius (r) is about 127 nm. Such PC patterns can be formed by electron beam (EB) lithography and inductively coupled plasma (ICP) etching. To obtain a free-standing PC slab, the BOX layer below the PC pattern needs to be removed (for example, by using an 48% hydrofluoric acid solution). Regarding the samples used in this work, the waveguide edge was fabricated by manual cleaving.

 figure: Fig. 1.

Fig. 1. Device structure of the HM Raman laser. (a) Schematic of the heterostructure nanocavity and the two adjacent waveguides. (b) Energy diagram of the heterostructure nanocavity. (c) Entire HM-laser device layout. (d) and (e) clarify the structures of the photonic mirrors with a heterointerface for reflection in the cases of pump and Stokes waveguides, respectively.

Download Full Size | PDF

Figure 1(a) is used to explain the core-region design of the laser. The heterostructure nanocavity consists of 27 missing air holes in the PC pattern. To create the heterostructure, the lattice constant a is slightly adapted in the x-direction near the center of the cavity: a is increased from 410 nm to 420 nm in steps of 5 nm [28]. This slight increase in a shifts the propagation bands to lower frequencies, as shown in Fig. 1(b). The cavity width remains constant, $\sqrt 3 a$ (W1 = 710 nm). The resonant mode formed by the 2nd propagation band and the 2nd mode gap is the so-called pump mode (with resonance wavelength λp). The other cavity mode at lower energy is the Stokes mode (with resonance wavelength λS). The pump and Stokes modes are used to confine the pump light and the Stokes Raman-scattered light, respectively. The design Q values of the pump mode and the Stokes mode are Qp_design = 6.14 × 105 and QS_design = 5.18 × 106, respectively. It has been shown that high quality factors help to enhance both spontaneous and stimulated Raman scattering [11,12,29]. Note that the above design Q values were determined by three-dimensional (3D) finite-difference time-domain (FDTD) calculations without waveguides adjacent to the cavity.

To enhance SRS in this device, the frequency spacing between these two modes (Δf) should be close to the Raman shift of silicon (15.606 THz) [30]. In addition, according to previous results that clarify the relationship between the crystal direction and the effective modal volume for SRS, the cavity should extend along the [100] direction of silicon [14,31]. It is known that Δf almost linearly increases with r [32]. Therefore, we fabricated several samples with slightly different values of r and determined the sample that shows laser oscillation with the lowest threshold. Furthermore, we used an SOI wafer where the crystal direction of the top silicon layer is rotated by 45 degrees with respect to the supporting Si substrate. Therefore, the [100] direction of the top silicon layer and the [110] direction of the supporting substrate are parallel [20]. This allows us to fabricate the Raman laser fabricated along the [100] direction while we can cleave the sample vertically to the waveguide, which is important for optical circuits.

The line defect below the nanocavity in Fig. 1(a) is the pump waveguide for the excitation of the pump mode. The upper waveguide is the Stokes waveguide for the extraction of Stokes light. The widths of the two waveguides are 0.88 W1 and 1.10 W1. The pump mode of the cavity couples to the guided mode in the 1st propagation band of the pump waveguide, and the Stokes mode of the cavity couples to the guided mode in the 1st propagation band of the Stokes waveguide. This coupling leads to lower quality factors of the pump and Stokes modes. We use the term Qtotal to denote the calculated Q value including the coupling to the waveguides:

$$\frac{1}{Q_{i\_{\textrm{total}}}} = \frac{1}{Q_{i\_{\textrm{design}}}} + \frac{1}{Q_{i\_{\textrm{in}}}},$$
where we can replace the index i by either p or S to obtain the equation that describes the pump or Stokes mode, respectively. The physical meaning of a reciprocal Q value is an optical loss. Therefore, the 1/Qi_in in Eq. (1) corresponds to the coupling loss due to the waveguide. The distance between the cavity and the pump waveguide, dp, is 4 rows as shown in Fig. 1(a). The distance for the Stokes mode, dS, is 5 rows. From 3D FDTD calculations including the adjacent waveguides we obtain Qp_total = 2.85 × 105 and QS_total = 1.56 × 106. This QS_total is smaller than the corresponding values in previous studies due to the smaller dS (dS > 6 in previous studies). We chose a smaller dS to increase Pedge.

In an actual sample, the positions and radii of the air holes exhibit small random deviations (in the sub-nanometer range) with respect to the ideal values due to fluctuations during the air-hole fabrication process. These deviations can induce additional losses by optical scattering, leading to the factor 1/Qscat [33]. Furthermore, the absorption losses related to defects in the silicon nanocavity and surface water need to be considered (1/Qabs) [34,35]. Thus, the experimental Q value (Qexp) is smaller than Qtotal, which is expressed by the following relation:

$$\frac{1}{Q_{i\_{\textrm{exp}}}} = \frac{1}{Q_{i{\_\textrm{design}}}} + \frac{1}{Q_{i{\_{\textrm{in}}}}} + \frac{1}{Q_{i{\_{\textrm{scat}}}}} + \frac{1}{Q_{i{\_{\textrm{abs}}}}}.$$
The magnitude of Qi_scat depends on the actual pattern of air holes including the random displacements, and thus Qp_exp and QS_exp vary from sample to sample [32].

Figure 1(c) shows the entire HM-laser device layout. Between the two waveguides, ten nanocavities with the same r are placed with a spacing of about 20 µm. This sample design enables an efficient measurement of the individual laser samples [36]: Both the λp and λS values are expected to be slightly different for each of the ten cavities due to the above-mentioned air-hole fabrication fluctuations. The magnitudes of the variation in λp and λS are usually more than one order of magnitude larger than the full widths at half-maximum of the pump and Stokes modes [32]. Therefore, in a structure that contains only a few nanocavities, it is unlikely that even one of the resonance modes interferes with any of the other resonance modes [24]. This means that we are usually able to characterize the individual resonance peaks of the nanocavities even if we only use one set of waveguides. In our sample, the waveguide edge was formed by cleaving the sample. Note that we did not add any sophisticated structure that would allow us to control the radiation pattern of the edge emission, the reflection, and the degree of scattering.

The positions of the two HMs used to reflect the light propagating in the waveguides are indicated by the blue and red rectangles in Fig. 1(c) for the pump light and the Stokes light, respectively. The lattice constant of the holes in the colored areas in Figs. 1(d) and (e) is 380 nm in the x-direction (that in the y-direction is 710 nm to maintain structural uniformity). The propagation bands in the colored area shift to higher frequencies as shown in the diagrams at the bottom of Figs. 1(d) and (e). The mode-gap differences at the heterointerfaces result in the reflection of light at λp and λS.

For the evaluation of the performance increase that can be achieved by the HMs, ten conventional Raman laser cavities were also fabricated on the same chip. The design of the core region is the same as that shown in Fig. 1(a). The HM Raman laser array and conventional Raman laser array are displaced in the y-direction by 100 µm. Therefore, the magnitudes of Qscat and Qabs should be similar for all cavities.

3. Experimental results

We used a tunable laser to determine the following properties of each cavity: the resonance wavelengths λp and λS, the frequency spacing Δf, Qp_exp and QS_exp, the threshold for laser oscillation (Ith), and the maximum Pedge (Pedge_max). The details of the measurement procedure are described in Appendix A1. The measurement was performed at 24 °C in air. Table 1 summarizes the results obtained from the four HM Raman laser cavities in which laser oscillation was achieved. These four cavities are hereafter referred to as cavities #1–#4. Table 2 summarizes the results obtained from the five conventional Raman laser cavities in which laser oscillation occurred. These cavities are referred to as cavities #5–#9.

Tables Icon

Table 1. Measurement results for four HM Raman lasers (λp, λS: resonance wavelengths, Δf: frequency spacing, Qp_exp and QS_exp: actual Q values, Ith: threshold, Pedge_max: maximum Pedge)

Figure 2 shows more details of the experimental results of cavity #1, which provides the largest Pedge_max. In Figure 2(a), the resonance and transmission spectra of the pump mode are shown by the closed and open circles, respectively. To obtain these spectra, the output of the tunable laser was focused at the input port indicated in Fig. 1(c) to excite the pump mode. As illustrated in the inset of Fig. 2(a), the surface emission from the cavity in the vertical direction of the slab (+z direction) was measured, as well as the transmitted intensity at the waveguide edge opposite to the input port. Figure 2(b) shows the spectra for the Stokes mode. For this measurement, the tunable laser irradiated the output port indicated in Fig. 1(c).

 figure: Fig. 2.

Fig. 2. Details of the results of HM Raman laser cavity #1. (a) Resonance spectrum of the pump mode (closed circles) and the corresponding transmission spectrum (open circles). The solid curve is the fitting result. (b) The resonance and transmission spectra of the Stokes mode. (c) Pedge and Psurface as a function of I. (d)−(g) NIR camera images for I = 1.2 Ith obtained by using a long-pass filter to remove the pump light. The images show (d) the edge emission at the output port, (e) the scattered light at the HM of the Stokes waveguide, (f) the scattered light at the output port, and (f) the surface emission from the nanocavity.

Download Full Size | PDF

Due to the reflection at the heterointerfaces shown in Figs. 1(d) and 1(e), the transmitted intensities shown in Figs. 2(a) and 2(b) are very small. In each figure, the solid curve shows the fit of the resonance spectrum to a Lorentzian function, from which the resonance wavelength (λ) and the full width at half-maximum (Δλ) are evaluated. The experimental Q value is calculated using the relation Qexp = λλ, and the obtained Qp_exp and QS_exp values are 1.49 × 105 and 7.87 × 105, respectively. This QS_exp is smaller than the values reported in previous studies due to the smaller dS (see Appendix A2 regarding the relation between Qexp and Ith). The measured Δf is 15.602 THz, and thus the detuning from the Raman shift (|Δfdet.| = |Δf − 15.606 THz|) is only 0.004 THz [30,32]. These optical characteristics allow Raman laser oscillation.

Figure 2(c) shows the input–output characteristics of this HM Raman laser. The orange points represent Pedge (the Stokes light emitted via the output port), while the black points represent Psurface (the Stokes light emitted from the cavity into free space). The pump power I is the power coupled into the pump mode of the nanocavity (to clearly distinguish between the pump power and the output characteristics, the symbol I is used to denote the pump power while P is used for the output power). Pedge, Psurface, and I were estimated from the intensities detected by using photodiodes and taking into account the loss of the optical components (details are described in Appendix A1).

As shown in Fig. 2(c), Pedge and Psurface increase steeply when I starts to exceed the threshold Ith = 1200 nW, but the Raman laser output saturates as I increases further. This is due to free-carrier absorption (FCA) by the carriers generated by two photon absorption (TPA) [37]. FCA is responsible for an additional (nonlinear) loss term, 1/Qi_FCA, that needs to be added to Eq. (2) in the case of strong excitation. It has been shown that the Qp_exp and QS_exp values under lasing conditions are smaller [38]. The Qi_exp values presented in Table 1 and 2 were measured with low pump powers where TPA effects are negligible.

The maximum Pedge (8.8 nW) is obtained at I = 4580 nW and the experimentally obtained ratio of Pedge/Psurface is 0.83. Note that our experimental setup leads to an underestimation of this ratio, because the numerical aperture (NA) of the objective lens that collects the surface emission is 0.7 and this value is larger than that of the lens that collects the edge emission, NA = 0.4 (see Appendix A1). We calculated the angular distribution of the edge emission from a PC waveguide edge using 3D FDTD simulations. The intensity emitted at angles larger than 23.6 degrees (with respect to the waveguide axis), cannot be collected by a lens with NA = 0.4, but this intensity can be as much as the intensity emitted at angles smaller than 23.6 degrees. Since the radiation pattern changes depending on the shape of the air holes at the edge, the difference in the NA for Pedge and Psurface is not considered in the analysis.

Figures 2(d)−(g) present different near-infrared (NIR) camera images obtained during laser oscillation at I = 1.2 Ith. Figure 2(d) is an image of the edge emission at the output port (view direction: parallel to the x-axis shown in Fig. 1). A clear emission spot is obtained. Figure 2(e) is an NIR camera image at the HM of the Stokes waveguide (top view). The scattered light at the heterointerface is very small (we consider that this scattered light are Stokes photons with TM polarization [39]). This small intensity of the scattered light and the small intensity of the transmitted light in Fig. 2(b) indicates that the reflectivity of this HM is close to 100%.

Figure 2(f) is the top view of the silicon slab at the position of the output port. A spot with strong scattering to the z-direction is observed. The intensity of this emission spot is 0.80 × Psurface at I = 4580 nW (determined by a photodiode), and thus it is almost the same as Pedge_max. A similar amount of scattered light is probably also radiated into the direction of the silicon substrate (−z direction). If a sophisticated structure, for example a spot size converter (SSC), had been added to the output port to reduce the scattering and to control the radiation pattern, the Pedge_max of cavity #1 would have exceeded maybe even 30 nW (methods to increase Pedge are discussed in Section 6).

The above results are now compared with those of cavity #5, the cavity with the largest Pedge_max among the five conventional Raman lasers. Figures 3(a) and 3(b) indicate a Qp_exp and a QS_exp of 1.59 × 105 and 6.89 × 105, respectively. The Δf is 15.588 THz, and thus |Δfdet.| = 0.018 THz. These values are equivalent to those of cavity #1.

 figure: Fig. 3.

Fig. 3. Details of the results of conventional Raman laser cavity #5. (a) Resonance spectrum of the pump mode (closed circles) and the corresponding transmission spectrum (open circles). The solid curve is the fitting result. (b) The resonance and transmission spectra of the Stokes mode. (c) Input–output characteristics of the laser. (d)−(f) NIR camera images of (d) the edge emission at the output port, (e) the scattered light at output port, and (f) the surface emission from nanocavity obtained at I = 1.2 Ith.

Download Full Size | PDF

Figure 3(c) shows the input–output characteristics of this Raman laser and indicates an Ith of about 1600 nW. The saturation of the Stokes power due to TPA-induced FCA is also observed. We find that Pedge_max = 2.0 nW at I = 4520 nW and Pedge/Psurface = 0.11. The latter value is only 13% of the Pedge/Psurface of cavity #1. On the other hand, the sum Pedge + Psurface is almost the same (20 nW for cavity #5 and 19.5 nW for cavity #1).

Figures 3(d)−(f) present the NIR camera images obtained at I = 1.2 Ith. The intensities of the emission spots in Figs. 3(d) and 3(e) are much smaller than those in Fig. 3(f). The power of the light scattered at the output port into the z-direction is also small, 0.08 × Psurface at I = 4520 nW. This is 10 times smaller than that of cavity #1. These results strongly suggest that the conventional Raman laser emits most of the laser light from the cavity into free space.

Figure 4(a) shows the normalized input–output characteristics of the four HM Raman lasers presented in Table 1 (we normalized the pump power to the threshold power for each sample). The black and orange curves correspond to Psurface and Pedge, respectively. Figure 4(b) shows the normalized input–output characteristics of the five conventional Raman lasers shown in Table 2. These figures clearly show the tendency of a larger Pedge for Raman lasers with HMs. The average Pedge_max of the four HM Raman lasers is 4.3 times larger than that of the five conventional Raman lasers.

Tables Icon

Table 2. Measurement results for five conventional Raman lasers (λp, λS: resonance wavelengths, Δf: frequency spacing, Qp_exp and QS_exp: actual Q values, Ith: threshold, Pedge_max: maximum Pedge)

 figure: Fig. 4.

Fig. 4. (a) Normalized input–output characteristics of four HM Raman samples (cavities #1∼#4), (b) Normalized input–output characteristics of five conventional Raman lasers (cavities #5∼#9).

Download Full Size | PDF

On the other hand, the sum of Pedge and Psurface is similar for both types of lasers except for cavity #7. Furthermore, the average Qp and QS of the four HM Raman lasers are equivalent to those of the five conventional Raman lasers as shown in Tables 1 and 2. Therefore, it can be concluded that the HMs lead to an increase in the Pedge of a Raman laser.

It is noted that the Psurface_max of cavity #4 is much larger than its Pedge_max. This suggests that even in a Raman laser with HMs, the increase in Pedge is sometimes not as large as would be expected from simple considerations. The coefficient of variation of Pedge_max (the ratio of the standard deviation to the mean) for the HM Raman laser is 0.78, while that for the conventional Raman laser is 0.55. Although the number of measured samples is small, the Pedge of a HM Raman laser may be subject to larger changes. These characteristics are analyzed in Section 5 by numerical simulations.

4. Calculation method

This section presents the model based on coupled-mode theory used to describe the devices and to investigate the Pedge and Psurface of Raman silicon nanocavity lasers with HMs. Figure 5 clarifies the physical meaning of the parameters used in the calculation model. Port 1 and Port 4 correspond to the input and output ports in Fig. 1(c), respectively. The HMs are placed at Port 2 and at Port 3. A reflectivity of 100% is assumed. To simplify the calculations, the reflectivities at the input and output waveguide edges are assumed to be 0%. In the actually measured samples, each pump waveguide is used to excite ten nanocavities, but the calculation considers only one nanocavity.

 figure: Fig. 5.

Fig. 5. Calculation model of the Raman silicon nanocavity laser with two HMs.

Download Full Size | PDF

The nanocavity has a pump mode and a Stokes mode characterized by the angular frequency ωp and ωS, respectively. The photons from the excitation source (photons at ω = ωp in the case of resonant excitation) enter the system at Port 1, and the Stokes Raman scattered photons at ωS leave the system from Port 4. Furthermore, the nanocavity also emits photons at ωp and ωS in the direction vertical to the PC slab. The value of 1/Qp_in describes the strength of the coupling between the pump mode and the pump waveguide, and 1/QS_in is the strength of the coupling between the Stokes mode and the Stokes waveguide. The photons at ωp do not couple to the Stokes waveguide and similarly, the photons at ωS do not couple to the pump waveguide. The strength of the coupling between the nanocavity and free space is denoted as 1/Qi_v. In an ideal nanocavity, Qi_v is equal to Qi_design. On the other hand, actual nanocavities include scattering and absorption losses. Accordingly, Qi_v should be expressed in terms of Qi_design, Qi_scat, and Qi_abs as shown in the following relation:

$$\frac{1}{Q_{i{\_\textrm{v}}}} = \frac{1}{Q_{i{\_{\textrm{design}}}}} + \frac{1}{Q_{i{\_{\textrm{scat}}}}}\textrm{ + }\frac{1}{Q_{i{\_{\textrm{abs}}}}}.$$
On the other hand, to simplify the calculations, we assume that 1/Qi_abs is zero. Furthermore, we do not take into account TPA and related effects (FCA, the carrier plasma effect, the Kerr effect, and the thermo-optic effect), which lead to a reduction in Qi_v and a shift in ωi in the strong-excitation regime [38,40]. The differences between the experimental results and the theoretical results calculated using these assumptions are discussed in Section 6.

The amplitudes of light propagating from each port to the cavity are denoted as S + k (k = 1, 2, 3, 4) and the amplitudes of light propagating from the cavity to each port are denoted S-k. d1d4 are the distances between the cavity and Ports 1–4, respectively. In the case that S−1 and S + 2 interfere destructively, the number of photons at ωp that are trapped in the cavity (Np) increases [21,22], and if they interfere constructively, Np decreases. Furthermore, if S + 3 and S−4 interfere constructively, the magnitude of S−4 increases, and thus Pedge becomes larger. θin and θout denote the round-trip phase shifts of the light reflected at the HMs. The degrees of interference in the pump and Stokes waveguides depend on θin and the θout, respectively.

For the numerical simulations, we use Eqs. (4)−(10) provided below. The coupled-mode theory used to derive these equations is described in Appendix A3.

$$N_\textrm{p} = \frac{{\frac{{{\omega _\textrm{p}}}}{{Q_{{\rm p}\_{\rm in}}^\mathrm{{\prime}}}}}}{{{{({\omega - \omega_\textrm{p}^{{\prime}}} )}^2} + {{\left\{ {\frac{{{\omega_\textrm{p}}}}{{2{Q_{{\rm p}\_{\rm v}}}}} + \frac{{{\omega_\textrm{p}}}}{{2Q_{{\rm p}\_{\rm in}}^{{\prime}}}} + g_\textrm{R}^{\textrm{cav}}({N_\textrm{S}} + 1)} \right\}}^2}}}\frac{{|{S_{ + 1}}{|^2}}}{{\hbar {\omega _\textrm{p}}}}{ ,}$$
$${N_\textrm{S}} = \frac{{2g_\textrm{R}^{\textrm{cav}}{N_\textrm{P}}}}{{\frac{{{\omega _\textrm{S}}}}{Q_{{\rm S}\_{\rm v}}} + \frac{{{\omega _\textrm{S}}}}{{Q_{{{\rm S}\_{\rm in}}}^{{\prime}}}} - 2g_\textrm{R}^{\textrm{cav}}{N_\textrm{P}}}}.$$
Here, Np is the number of photons at ωp in the nanocavity, NS is the number of photons at ωS in the nanocavity, and $g_\textrm{R}^{\textrm{cav}}$ is the Raman gain coefficient of the nanocavity. ${Q^{\prime}_{{\rm p}\_{\rm in}}}$ and ${Q^{\prime}_{{{\rm S}\_{\rm in}}}}$ are effective Q values obtained by considering the effect of the HMs:
$${Q^{\prime}_{{\rm p}\_{\rm in}}} = \frac{{{Q_{{\rm p}\_{\rm in}}}}}{{1 + \cos {\theta _{\textrm{in}}}}},$$
$${Q^{\prime}_{{{\rm S}\_{\rm in}}}} = \frac{{{Q_{{{\rm S}\_{\rm in}}}}}}{{1 + \cos {\theta _{\textrm{out}}}}}.$$
${Q^{\prime}_{{\rm p}\_{\rm in}}}\,({Q^{\prime}_{{\rm S}\_{\rm in}}})$ reaches its minimum in the case of θin = 0 (θout = 0), and this means that the coupling between the pump waveguide (Stokes waveguide) and the pump mode (Stokes mode) is maximized. Accordingly, a large Pedge is expected. On the other hand, if θin = ±π (θout = ±π), there is no coupling between the pump waveguide (Stokes waveguide) and the pump mode (Stokes mode). The definition of ${\omega^{\prime}_{\text{p}}}$ in Eq. (4) is:
$${\omega ^{\prime}_\textrm{p}} = {\omega _\textrm{p}}\left( {1 + \frac{{\sin {\theta_{\textrm{in}}}}}{{2{Q_{{\rm p}\_{\rm in}}}}}} \right)\textrm{.}$$
In our calculation, we use the approximation ${\omega ^{\prime}_\textrm{p}} = {\omega _\textrm{p}}$ because of Qp_in >> 1. The Stokes output power at Port 4, Pedge, is given by the following equation:
$${P_{\textrm{edge}}} = {|{{S_{\textrm{ - 4}}}} |^2} = \frac{{{\omega _\textrm{S}}}}{{{{Q^{\prime}}_{{{\rm S}\_{\rm in}}}}}}{N_\textrm{S}}\hbar {\omega _\textrm{S}}.$$
Similarly, the power of the Stokes surface emission from the cavity (into the + z direction), Psurface, is given by
$${P_{\textrm{surface}}} = \frac{1}{2}\frac{{{\omega _\textrm{S}}}}{Q_{{\rm S}\_{\rm v}}}{N_\textrm{S}}\hbar {\omega _\textrm{S}}.$$
The factor 1/2 accounts for the emission towards the substrate. We wrote a program that calculates the Pedge and Psurface values for arbitrary values of S + 1 for ω = ωp using Eqs. (4)−(10). The parameters used in the calculations are shown in Table 4 in Appendix A3.

5. Calculation results

For the calculations in this section, we use Qp_v = 2.27 × 105, Qp_in = 5.32 × 105, QS_v = 1.02 × 106, and QS_in = 2.12 × 106. These values result in Qi_exp values equal to those of cavity #5 (the estimation method is described in Appendix A4). Figures 6(a) and (b) show the calculated input–output characteristics of the HM Raman laser and the corresponding conventional laser, respectively, for (θin, θout) = (0, 0). The x-axis represents the excitation power at Port 1, |S + 1|2. It should be noted that this definition is different from that for the x-axis in Figs. 2(c) or 3(c). In the case of the conventional Raman laser, we can roughly estimate that the power coupled into the pump mode of the nanocavity is about 0.4 × |S + 1|2. In Fig. 6, the number of pump photons in the cavity, Np, is shown by the dashed line. The orange curve is Pedge and the black curve is Psurface.

 figure: Fig. 6.

Fig. 6. Theoretical input–output characteristics for (θin, θout) = (0, 0). (a) Pedge, Psurface, and Np as a function of |S + 1|2 for a HM Raman laser and (b) the results for a conventional Raman laser.

Download Full Size | PDF

Similar to Figs. 2(c) or 3(c), Pedge and Psurface in Figs. 6(a) and (b) rapidly increase as |S + 1|2 becomes larger than the threshold (Ith_cal). Np increases linearly in the region |S + 1|2 < Ith_cal and becomes constant in the region |S + 1|2 > Ith_cal. The calculated values of Pedge and Psurface are much larger than the experimentally measured values in Fig. 4. This is probably due to the fact that the calculations ignore TPA-induced FCA and, in the experiments, it was only possible to measure a part of Pedge. The difference between the Pedge values in Figs. 6(a) and (b) is discussed after the explanation of the threshold behavior.

By comparing Figs. 6(a) and (b), we find that the Ith_cal of the HM Raman laser is about half of that of the conventional Raman laser. In a HM Raman laser with θin = 0, S + 2 and S−1 interfere destructively and the excitation efficiency of the pump mode is improved. Therefore, the rate of increase in Np in Fig. 6(a) is larger than that for Fig. 6(b) in the region |S + 1|2 < Ith_cal.

The number of pump photons in the cavity at Ith_cal, Np_th, in the case of the HM Raman laser is 1890, which is larger than that of the conventional laser. As explained below, this is due to the smaller QS_total of the HM Raman laser. Instead of Eq. (1), the Qi_total of the HM laser is determined by the following equation:

$$\frac{1}{{{Q_{i{\_{\rm{total}}}}}}} = \frac{1}{{{Q_{i{\_{\rm v}}}}}} + \frac{1}{{{{Q^{\prime}}_{i{\_{\rm in}}}}}}\textrm{ }\textrm{.}$$
From Eq. (11) we obtain QS_total = 5.20 × 105 and 6.89 × 105 for the HM and the conventional Raman lasers, respectively. This has an influence on the threshold because Raman laser oscillation becomes only possible if the Raman gain exceeds the total loss of the resonator for the Stokes mode (QS_total−1). Furthermore, the Raman gain is proportional to the Np. A laser with a smaller QS_total requires a larger Raman gain, and thus the Np_th for Fig. 6(a) is larger than that for Fig. 6(b).

Now we consider the influence of the HMs on Pedge. In a HM Raman laser with θout = 0, S + 3 and S−4 interfere constructively and thus the Pedge increases: the Pedge of the HM Raman laser in Fig. 6 is 938 nW at 3 Ith_cal while that of the conventional Raman laser is 197 nW. The ratio is 4.8, which is close to the ratio of the Pedge_max values of cavities #1 and #5 shown in Figs. 2(c) and 3(c). The calculation results in Fig. 6 are qualitatively consistent with the input–output characteristics of cavities #1 and #5.

To investigate the dependence of Pedge and Psurface on the round-trip phase shifts θin and θout, we calculated the input–output characteristics of the HM Raman laser for various sets of (θin, θout). Figure 7(a) corresponds to (θin, θout) = (0, 0) [the same graph as in Fig. 6(a)]. Figures 7(b)−(d) show the calculation results for (θin, θout) = (0, π), (π, 0), and (π, π), respectively.

 figure: Fig. 7.

Fig. 7. (a)−(d) Calculated input-output characteristics of the HM Raman laser for (θin, θout) = (0, 0), (0, π), (π, 0), and (π, π), respectively. (e)−(h) Dependence of Pedge on the (θin, θout) for four different values of |S + 1|2 (0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0). (i) (θin, θout) dependence of Psurface at 3.0 × Ith_cal_θ=0. (j) Total Stokes output at 3.0 × Ith_cal_θ=0.

Download Full Size | PDF

The HM Raman lasers with θout = π have a ${Q^{\prime}_{{{\rm S}\_{\rm in}}}} = \infty $, that is, no coupling between the Stokes mode and the Stokes waveguide (Pedge is zero). Therefore, the Psurface in Fig. 7(b) is larger than that in Fig. 7(a). Since Psurface is also emitted towards the substrate, the Psurface does not increase as rapidly as the Pedge in Fig. 7(a). The values of Ith_cal and Np_th in Fig. 7(b) are smaller than those in Fig. 7(a) because the QS_total for Fig. 7(b) is 1.02 × 106, which is larger than that for Fig. 7(a), 5.20 × 105. The larger QS_total leads to a decrease in Np_th, but the rate of increase in Np below Ith_cal is identical with that in Fig. 7(a) since the ${Q^{\prime}_{{\rm p}\_{\rm in}}}$ values are the same for both calculations. The calculation results shown in Fig. 7(b) with (θin, θout) = (0, π) exhibit the lowest Ith_cal and the largest Psurface.

In the case of θin = π, the excitation efficiency of the pump mode is zero because ${Q^{\prime}_{{\rm p}\_{\rm in}}} = \infty $. Therefore, no laser oscillation occurs in Figs. 7(c) and (d). The results of Figs. 7(a)−(d) indicate that the threshold and the Pedge of a HM Raman laser strongly depend on (θin, θout). Cavities #1−#4 in Fig. 4(a) probably exhibit a large variation in Pedge because the (θin, θout) values have a strong sample dependence.

The images in Figs. 7(e)−(h) show the (θin, θout) dependence of Pedge for different input powers. In each image, the red region shows the combinations of (θin, θout) that are possible to achieve a strong Pedge. The results of Pedge are a function of θin and θout with a period of 2π. Therefore, we hereafter discuss Pedge in the range −π < θin < π and −π < θout < π. We define Ith_cal_θ=0 as the threshold power for (θin, θout) = (0, 0) [the threshold in Fig. 7(a)]. Figures 7(e)−(h) are the results for |S + 1|2 = 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.

For |S + 1|2 = 0.1 × Ith_cal_θ=0, Pedge exhibits a maximum at (θin, θout) = (0, 0). The maximum value is only 16 pW because the input power is low. It is known that the excitation efficiency of a cavity with a HM is maximized under the condition ${Q_{{\rm p}\_{\rm v}}} = {Q^{\prime}_{{\rm p}\_{\rm in}}}$ if the excitation power is small enough to be able to ignore the nonlinear effects of SRS or TPA [21]. Here, this condition is not realized for any θin since Qp_in (5.32 × 105) is more than two times larger than Qp_v (2.27 × 105). The highest excitation efficiency is realized for θin = 0, because here the minimum ${Q^{\prime}_{{\rm p}\_{\rm in}}}$ (Qp_in/2) is obtained. The coupling between the Stokes mode and the Stokes waveguide is maximized at θout = 0, where ${Q^{\prime}_{{{\rm S}\_{\rm in}}}}$ reaches its minimum. As a result, Pedge reaches its maximum at (θin, θout) = (0, 0) in Fig. 7(e).

For |S + 1|2 = 1.0 × Ith_cal_θ=0, the maximum Pedge is obtained at (θin, θout) = (0, ±π/2). This is because the threshold for the HM Raman laser with (θin, θout) = (0, ±π/2) is smaller than Ith_cal_θ=0: as θout approaches π or −π from 0, the value of ${Q^{\prime}_{{{\rm S}\_{\rm in}}}}$ increases according to Eq. (7) and QS_total increases according to Eq. (11), and thus Np_th decreases. The lowest threshold is obtained at (θin, θout) = (0, π) and (0, −π), but Pedge becomes zero under such a condition as shown in Fig. 7(b). Therefore, in Fig. 7(f), the maximum Pedge is obtained at (θin, θout) = (0, ±π/2).

In Fig. 7(g), the value of (θin, θout) that is advantageous for Pedge shifts again to the region around (0, 0): as the excitation intensity become higher than 1.0 × Ith_cal_θ=0, strong coupling with the Stokes waveguide due to a smaller ${Q^{\prime}_{{{\rm S}\_{\rm in}}}}$ is more advantageous than a lower threshold. In Fig. 7(h), which is for 3.0 × Ith_cal_θ=0, a maximum Pedge of about 900 nW is obtained at (θin, θout) = (0, 0).

In the experimental results shown in Figs. 2(c) and 3(c), the maximum Pedge appears at excitation powers near 3.0 × Ith. Thus, the calculation condition of Fig. 7(h) should be considered in discussions on further increases in Pedge. Although (θin, θout) = (0, 0) does not minimize the threshold, these round-trip phase shifts are the most desirable values to increase the Pedge of HM Raman lasers. This feature is maintained even if the four Q values change, which is demonstrated in Section 6. Since cavity #1 has the highest Ith and the largest Pedge among the fabricated HM Raman lasers, the calculation result in Fig. 7(h) suggests that the (θin, θout) of cavity #1 is close to (0, 0).

Figure 7(i) shows the (θin, θout) dependence of Psurface at 3.0 × Ith_cal_θ=0. The maximum Psurface is obtained at (θin, θout) = (0, ±π). This clarifies that the HM laser with (θin, θout) = (0, ±π) realizes the lowest threshold and the largest Psurface. Cavity #4 exhibits the largest Psurface as shown in Table 1, while its Pedge is the lowest. Furthermore, this cavity has the highest QS_exp and the lowest Ith. Based on these features, we consider that cavity #4 has a (θin, θout) close to (0, ±π).

Figure 7(j) shows the (θin, θout) dependence of the total output (Pedge + 2 × Psurface) at 3.0 × Ith_cal_θ=0. In the case of θin = 0, the Stokes output is larger than 1900 nW, independent of θout. On the other hand, the intensities in the range 21π/25 ≤ θin ≤ π are less than 10 pW since laser oscillation does not occur due to the low excitation efficiency. This means that laser oscillation may not occur even if the Qexp values are sufficiently high and the Δf is in principle appropriate for laser oscillation.

To clarify the impact of a change in either θin or θout, we summarize the Ith_cal and the Np_th for (θin, θout) = (0, 0), (0, π/3), (0, 2π/3), (0, 5π/6), (π/3, 0), (2π/3, 0), and (5π/6, 0) in Table 3. The impact of θout on QS_total is described by Eqs. (7) and (11), and thus a variation of θout changes Np_th and Ith_cal. The impact of θin on the excitation efficiency is a result of a change in Qp_total according to Eqs. (6) and (11). As a result, a variation of θin changes only Ith_cal.

Tables Icon

Table 3. Ith_cal and Np_th for the seven input–output relations displayed in Fig. 8

Tables Icon

Table 4. Parameters used in the calculations

Figure 8(a) shows the normalized input–output characteristics for (θin, θout) = (0, 0), (0, π/3), (0, 2π/3), and (0, 5π/6) (we normalized the input power to the threshold for each sample). Pedge and Psurface are shown by the orange and black curves, respectively. Similarly, Fig. 8(b) shows the characteristics for (θin, θout) = (0, 0), (π/3, 0), (2π/3, 0), (5π/6, 0). Figure 8 shows that Pedge varies relatively strongly with (θin, θout). This is probably the reason why the variation in Pedge for the HM Raman lasers in Fig. 4 is 1.42 times larger than that for the conventional Raman lasers. Although the results in Fig. 4(a) are for cavities with different Q values and different Δf values, the degree of consistency with the results in Fig. 8 is good. The (θin, θout) values of the four HM Raman lasers shown in Table 1 are probably different from each other. The control of (θin, θout) is thus considered important to reduce the variation in Pedge.

 figure: Fig. 8.

Fig. 8. (a) Calculated normalized input–output characteristics of HM Raman lasers for θin = 0 and seven different values of θout. (b) The input–output characteristics for θout = 0 and seven different values of θin.

Download Full Size | PDF

The 2D plot in Fig. 9 shows the ratio of the Pedge of the HM Raman laser at 3.0 × Ith_cal_θ=0 [Fig. 7(h)] to the Pedge of the conventional Raman laser at 3.0 × Ith_cal [Fig. 6(b)]. The maximum value is 4.76, which is obtained at (θin, θout) = (0,0). The red region describes the (θin, θout) range where the Pedge of the HM Raman laser is higher than that of the conventional laser (Pedge_HM / Pedge_conv > 1.0). The area of this region is 56% of the total area. The area of the region where Pedge_HM / Pedge_conv > 3.0 is 22% of the total area. Cavities #1–#3 have a larger Pedge_max than the conventional laser cavities #5–#9, and thus their (θin, θout) pairs may be located in the red region. On the other hand, the (θin, θout) pair of cavity #4 may be located in the blue region. We conclude that it is important to choose a HM Raman laser design with a (θin, θout) close to (0, 0) and to fabricate the device accurately.

 figure: Fig. 9.

Fig. 9. (θin, θout) dependence of Pedge_HM / Pedge_conv. To prepare this plot, we used the Pedge_HM values for 3.0 × Ith_cal_θ=0 and the Pedge_conv values for 3.0 × Ith_cal. The red regions show conditions where Pedge_HM / Pedge_conv > 1.0.

Download Full Size | PDF

6. Discussion

The two types of Raman silicon nanocavity lasers investigated in Section 3 have the same nanocavity design and they were fabricated on the same chip with a separation of 100 µm. They have similar average Q values as shown in Tables 1 and 2, while the average Pedge_max of the HM Raman lasers is 4.3 times larger than that of the conventional Raman lasers. On the other hand, the average Psurface_max of the conventional Raman lasers is 1.84 times larger than that of the HM Raman lasers. The input–output characteristics presented in Figs. 6(a) and (b), which were calculated using Q values that are consistent with the experimental Q values of cavity #5, are in good agreement with the experimental results shown in Figs. 2(c) and 3(c). It is apparent that adding HMs to a Raman silicon nanocavity laser can lead to an increase in Pedge.

As shown in Fig. 4(a), the Pedge of the HM Raman laser varies significantly from sample to sample. This is most likely due to the difference in the round-trip phase shift (θin, θout), as explained in Figs. 7 − 9. Figure 9 clarified that the values of (θin, θout) should be close to (0,0) in order to obtain a high Pedge. To this end, numerical investigations of the relation between (θin, θout) and the distances between the cavity and the ports, dk, would be useful. Furthermore, the influence of the reflection at ports 1 and 4 should be studied. Such studies will also be useful for the design of a SSC for the output and input ports. In addition, we need to consider that the positions and radii of the air holes randomly deviate from the design values due to fabrication imperfections [32,33]. These variations can affect θin and θout. It will be important to perform device simulations by 3D FDTD calculations including random deviations of the air hole pattern.

The experimentally determined maximum value of Pedge shown in Fig. 2(c) is about two orders of magnitude smaller than the calculated one shown in Fig. 6(a). We can consider three possible causes for this result: Firstly, only a part of the total Pedge was measured. Secondly, the actual absorption losses in the fabricated sample (1/Qi_abs) are significantly large although they were assumed to be zero in the calculation. In the fabricated sample, the magnitude of QS_abs is expected to be larger than 5.0 × 106. On the other hand, the Qp_abs can be less than 5.0 × 105. This is comparable to Qp_v, and thus it is possible that about half of the pump photons were lost due to absorption. The third possible cause is that the QS_v in the experiment decreased by TPA-induced FCA. A QS_FCA of about 1 million has been previously reported in the high-excitation power regime [38], which is also comparable to QS_v. Therefore, it is possible that about half of the Stokes photons were lost due to FCA.

To increase the edge emission, not only the round-trip phase shifts θin and θout are important, but also Qp_v, Qp_in, QS_v , and QS_in. The values of Qp_v, Qp_in, QS_v, and QS_in used in Section 5 are 2.27 × 105, 5.32 × 105, 1.02 × 106, and 2.12 × 106, respectively. This set of Q values belongs to the regime where Qp_v < Qp_in and QS_v < QS_in. We define this as regime 1. Figures 10(a) and (b) are the results of calculations in which the first two values are exchanged (resulting in regime 2 with Qp_v < Qp_in and QS_v > QS_in), while Figs. 10(c) and (d) are for the calculations in which the last two values are exchanged (resulting in regime 3 with Qp_v > Qp_in and QS_v < QS_in). Figures 10(e) and (f) are the results of calculations in which both the first two and the last two values are exchanged, resulting in regime 4 with Qp_v > Qp_in and QS_v > QS_in.

 figure: Fig. 10.

Fig. 10. Calculation results for different Q-value sets. The left column shows the dependence of Pedge on (θin, θout) at 3.0 × Ith_cal_θ=0 for (a) regime 2, (c) regime 3, and (e) regime 4. The right column shows Pedge, Psurface, and Np as a function of |S + 1|2 for (b) regime 2, (d) regime 3, and (f) regime 4. Here, we used the condition (θin, θout) = (0, 0).

Download Full Size | PDF

Figures 10(a), (c), and (e) show the (θin, θout) dependence of Pedge at |S + 1|2 = 3.0 × Ith_cal_θ=0 for the sets of Q values in regimes 2 − 4, respectively (the dependence for other excitation intensities are shown in Figs. 12 − 14 in Appendix A5). These three figures and Fig. 7(h) clearly indicate that the condition (θin, θout) = (0, 0) maximizes Pedge in all four regimes.

Figures 10(b), (d), and (f) show the input–output characteristics for the regimes 2 − 4 under the condition (θin, θout) = (0, 0). In Fig. 10(b), the Pedge at 3.0 × Ith_cal is 1970 nW. This is 2.1 times larger than that in regime 1, because the condition QS_v > QS_in increases the output efficiency to the Stokes waveguide. In Fig. 10(d), the Pedge at 3.0 × Ith_cal is 1220 nW, which is 1.3 times larger than that in regime 1, because the excitation efficiency of the pump mode increases under the condition Qp_v > Qp_in. The enhancement of Pedge in regime 2 is larger than that in regime 3, indicating that the condition QS_v > QS_in is more important for Pedge. In Fig. 10(f), the Pedge at 3.0 × Ith_cal is 2540 nW, which is 2.7 times larger than that of the regime 1. The Pedge in regime 4 is the largest among the values in the four regimes.

The above shows that we should fabricate a HM Raman laser with (θin, θout) = (0, 0), Qp_v > Qp_in, and QS_v > QS_in in order to increase Pedge. Raman lasers with Qp_exp > 1.0 × 106 and QS_exp > 3.0 × 106 have already been reported [41]. The Qi_in can be controlled by the width of the excitation waveguide or the distance between the waveguide and the nanocavity, di. Therefore, it is possible to fabricate a HM Raman laser with Q values in regime 4.

Finally, we comment on the remaining challenges regarding the fabrication process. The Qp_v and QS_v, which are expressed in terms of Qdesign, Qscat, and Qabs in Eq. (3), should be as high as possible. The Qi_design can be further increased by using machine learning [42], and thus it is important to improve the fabrication process in such a way that 1/Qscat and 1/Qabs become smaller. This will be important not only for Raman lasers fabricated using electron-beam lithography, but also for samples fabricated by CMOS-compatible processes [43]. Since a reverse-biased diode structure adjacent to the nanocavity can shorten the lifetime of TPA carriers, the Pedge in the high-excitation regime can be improved by adding such a structure [10,11,44]. Recently, the combination of a p–i–n diode and a nanocavity, where a Q value larger than one million was maintained, has been reported [26]. Raman silicon nanocavity lasers with a reverse-biased diode can be developed. The addition of SSCs to the waveguide edges can further improve Pedge by several times [45].

7. Summary

We investigated the increase in the laser emission from the edge of the Stokes waveguide of a Raman silicon nanocavity laser that can be achieved by adding HMs. The average intensity of the edge emission of our Raman lasers with HMs is 4.3 times stronger than that of our conventional Raman lasers. Numerical calculations based on coupled-mode theory reproduce the experimentally observed increase in the edge emission and its variation well. The calculations reveal that the round-trip phase shifts (θin, θout) should be close to (0, 0) and the Qi_v values should be larger than the Qi_in values.

Appendix

A1. Experimental method

Figure 11 shows the measurement system used to obtain the results presented in Section 3. The figure illustrates the situation in which laser oscillation occurs. A continuous-wave tunable laser (Santec TSL-510) was used as the light source. The wavelength of the tunable laser was accurately measured by a wavelength meter (Agilent 86122A). The excitation light (Fig. 11; blue beam) was focused at the edge of the excitation waveguide (Fig. 1; input port) by an objective lens with a numerical aperture (NA) of 0.4. The surface emission was collected by an objective lens placed above the silicon slab (NA = 0.7) and detected by an InGaAs photodiode (Fig. 11; photodiode 1) with a lock-in amplifier. The edge emission was collected by the objective lens on the left-hand side of the sample (NA = 0.4) in Fig. 11 and detected by photodiode 2. Because of the limitations of the lens diameter and the working distance, the NA of the lens that collects the edge emission is smaller than that for the surface emission. As indicated in Fig. 11, the excitation light and the light emitted from the waveguide edge was passed through polarizers that transmit the TE component.

 figure: Fig. 11.

Fig. 11. Measurement system.

Download Full Size | PDF

The resonance spectra in Figs. 2 and 3 were obtained by measuring the wavelength of the tunable laser and the surface and edge emission while changing the wavelength gradually from short to long wavelengths. The pump power used in these measurements was small enough to be able to neglect the effect of TPA in the analysis. The long-pass filters (LPFs) were not used when we measured the resonance spectra.

TPA-induced carriers can lead to a shift of λp mainly through the thermo-optic effect and the carrier plasma effect [38,40]. When we measured the input–output characteristics of the Raman lasers [Figs. 2(c) and 3(c)], we used the excitation wavelength at which the laser intensity is maximized (in other words, we always adjusted the wavelength of the tunable laser to the shifted λp [40]). The LPFs with a cutoff wavelength of 1500 nm were inserted into the detection paths to remove the pump light. To obtain the NIR camera images, switching mirrors were inserted into the optical paths. The pump power was estimated from the surface emission at λp in the weak-excitation regime by assuming that the power coupled into the pump mode is linearly proportional to the power of the pump laser. This estimation method includes an uncertainty due to the variation in the collection efficiency of the emitted light (since the emission pattern is slightly different for each sample) [33]. Therefore, the estimated Ith should have a relative error equal to that of the collection efficiency.

A2. Relation between the threshold and the Q values

In the case that TPA-induced FCA loss can be ignored, the laser threshold Ith is inversely proportional to the product Qp_exp·QS_exp and the Raman gain coefficient of the nanocavity, $g_\textrm{R}^{\textrm{cav}}$ [12]:

$${I_{\textrm{th}}} \propto {({g_\textrm{R}^{\textrm{cav}}{Q_{{{\rm p}\_{\rm{exp}}}}}{Q_{\rm S}\_{\rm{exp}}}} )^{ - 1}}.$$
According to this relation, the Ith values presented in Tables 1 and 2 are larger than in the previous studies.

A3. Coupled-mode theory and calculation parameters

In Fig. 5, the time evolution of the amplitudes of the pump light confined in the nanocavity (ap) and the Stokes light (aS) can be calculated using the following rate equations [12,13]:

$$\begin{array}{l} \frac{{d{a_\textrm{p}}}}{{dt}} = (j{\omega _\textrm{p}} - \frac{{{\omega _\textrm{p}}}}{{2{Q_{{\rm v}\_{\rm p}}}}} - \frac{{{\omega _\textrm{p}}}}{{2{Q_{{\rm{in}}\_{\rm p}}}}}){a_\textrm{p}} - g_\textrm{R}^{\textrm{cav}}({N_\textrm{S}} + 1){a_\textrm{p}}\\ + \sqrt {\frac{{{\omega _\textrm{p}}}}{{2{Q_{{\rm{in}}\_{\rm p}}}}}} {e^{ - j\beta {d_1}}}{S_{ + 1}} + \sqrt {\frac{{{\omega _\textrm{p}}}}{{2{Q_{{\rm{in}}\_{\rm p}}}}}} {e^{ - j\beta {d_2}}}{S_{ + 2}}, \end{array}$$
$$\begin{array}{l} \frac{{d{a_\textrm{S}}}}{{dt}} = (j{\omega _\textrm{s}} - \frac{{{\omega _\textrm{s}}}}{{2{Q_{{\rm v}\_{\rm S}}}}} - \frac{{{\omega _\textrm{s}}}}{{2{Q_{{\rm{in}}\_{\rm S}}}}}){a_\textrm{s}} + g_\textrm{R}^{\textrm{cav}}{N_\textrm{P}}{a_\textrm{s}}\\ + \sqrt {\frac{{{\omega _\textrm{s}}}}{{2{Q_{{\rm{in}}\_{\rm S}}}}}} {e^{ - j\beta {d_3}}}{S_{ + 3}} + \sqrt {\frac{{{\omega _\textrm{s}}}}{{2{Q_{{\rm{in}}\_{\rm S}}}}}} {e^{ - j\beta {d_4}}}{S_{ + 4}}. \end{array}$$
Here, j is the imaginary unit and β is the propagation constant. Ni with i = p and S describes the number of pump photons and Stokes photons in the nanocavity, respectively. Ni has the following relationship with ai:
$${N_i} = \frac{{{{|{{a_i}} |}^2}}}{{\hbar {\omega _i}}}\textrm{(}i\textrm{ = p,S)}\textrm{.}$$
The Raman gain coefficient of the silicon nanocavity can be written as
$$g_\textrm{R}^{\textrm{cav}} = \frac{{\hbar {\omega _\textrm{p}}{c^2}{g_{{\rm R}\_{\rm{Si}}}}}}{{2{n_\textrm{p}}{n_\textrm{S}}{V_\textrm{R}}}}$$
Here, ni is the refractive index, ωp is the angular frequency of the pump light, c is the speed of light, and gR_Si is the Raman gain coefficient of bulk silicon. VR is the effective mode volume for Raman scattering. VR is determined by the electromagnetic field distributions of the two nanocavity modes and the Raman selection law of crystalline silicon. Our Raman silicon nanocavity laser extends along the [100] direction of silicon to reduce VR [31]. The amplitudes of the light propagating in the waveguides, S+k and Sk, are given by
$${S_{ - 1}} = {e^{ - j\beta ({d_1} + {d_2})}}{S_{ + 2}} - {e^{ - j\beta {d_1}}}\sqrt {\frac{{{\omega _\textrm{p}}}}{{2{Q_{{\rm{in}}\_{\rm p}}}}}} {a_\textrm{p}},$$
$${S_{ - 2}} = {e^{ - j\beta ({d_1} + {d_2})}}{S_{ + 1}} - {e^{ - j\beta {d_2}}}\sqrt {\frac{{{\omega _\textrm{p}}}}{{2{Q_{{\rm{in}}\_{\rm p}}}}}} {a_\textrm{p}},$$
$${S_{ - 3}} = {e^{ - j\beta ({d_3} + {d_4})}}{S_{ + 4}} - {e^{ - j\beta {d_3}}}\sqrt {\frac{{{\omega _\textrm{s}}}}{{2{Q_{{\rm{in}}\_{\rm S}}}}}} {a_\textrm{s}},$$
$${S_{ - 4}} = {e^{ - j\beta ({d_3} + {d_4})}}{S_{ + 3}} - {e^{ - j\beta {d_4}}}\sqrt {\frac{{{\omega _\textrm{s}}}}{{2{Q_{{\rm{in}}\_{\rm S}}}}}} {a_\textrm{s}}.$$
In our model, the amplitudes S + 2 and S + 3 are the result of HMs with 100% reflectivity:
$${S_{ + 2}} = {S_{ - 2}}{e^{ - j{\Delta _{\textrm{in}}}}},$$
$${S_{ + 3}} = {S_{ - 3}}{e^{ - j{\Delta _{\textrm{out}}}}}.$$
The parameters Δin and Δout are the phase shifts due to the reflection at the HMs. The reflection at Port 4 is zero and thus S + 4 = 0. The round-trip phase shifts θin and θout are defined by the following two equations:
$${\theta _{\textrm{in}}} = 2\beta {d_2} + {\Delta _{\textrm{in}}}.$$
$${\theta _{\textrm{out}}} = 2\beta {d_3} + {\Delta _{\textrm{out}}}.$$

A4. Estimation of the four Q values used in Section 5

In Section 5, we used Qp_v = 2.27 × 105, Qp_in = 5.32 × 105, QS_v = 1.02 × 106, and QS_in = 2.12 × 106. The experimental Qi_exp values have the following relationship with Qi_in and Qi_v (i = p, S):

$$\frac{1}{{{Q_{i{\_{\rm{exp}}}}}}} = \frac{1}{{{Q_{i{\_{\rm v}}}}}} + \frac{1}{{{Q_{i{\_{\rm{in}}}}}}}.$$
Furthermore, according to coupled-mode theory, Qi_v has the following relationship with Qi_exp [47]:
$$\frac{1}{{{Q_{i{\_{\rm v}}}}}} = \frac{{{Q_{i{\_{\rm{exp}}} }}}}{{\sqrt T }}$$
Here, T is the transmittance at either λp or λS of the transmission spectrum. From Fig. 3(b), we obtained QS_exp = 6.89 × 105 and T = 0.48. These values were used for the estimation of QS_v = 1.02 × 106 and QS_in = 2.12 × 106. On the other hand, the transmittance for the pump mode cannot be correctly estimated from Fig. 3(a). Therefore, we derived the Qp_in of 5.32 × 105 from Eq. (1) with Qp_design = 6.14 × 105 and Qp_total = 2.85 × 105 obtained by FDTD calculations. Then, the Qp_v of 2.27 × 105 was estimated from Eq. (25) with Qp_exp of 1.59 × 105 and the Qp_in estimated above.

A5. (θin, θout) dependence of Pedge in regimes 2–4 for various input powers

Figures 12(a)−(d) present the (θin, θout) dependence of Pedge in regime 2 for |S + 1|2 = 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively. The images in Figs. 13 and 14 correspond to the results in regimes 3 and 4, respectively. The calculation results for regime 1 are shown in Figs. 7(e)−(h).

 figure: Fig. 12.

Fig. 12. (a)−(d) Dependence of Pedge on (θin, θout) in regime 2 for 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.

Download Full Size | PDF

 figure: Fig. 13.

Fig. 13. (a)−(d) Dependence of Pedge on (θin, θout) in regime 3 for 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.

Download Full Size | PDF

 figure: Fig. 14.

Fig. 14. (a)−(d) Dependence of Pedge on (θin, θout) in regime 4 for 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.

Download Full Size | PDF

We can confirm that the images of regime 2 are similar to those of regime1. On the other hand, the images for 0.1 × Ith_cal_θ=0 to 1.5 × Ith_cal_θ=0 in regimes 3 and 4 differ significantly from those in regime1 [Figs. 7(e)−(g)]. This is because Qp_v is larger than Qp_in in regimes 3 and 4, whereas Qp_v is smaller than Qp_in in regime 1. As mentioned in the explanation of Fig. 7(e), in the case that Qp_v = ${Q^{\prime}_{{\rm p}\_{\rm in}}}$, the excitation efficiency of the pump mode is maximized [21]. In regimes 1 and 2, this matching condition cannot be satisfied with the given set of Q values. On the other hand, in regimes 3 and 4, Qp_v is equal to ${Q^{\prime}_{{\rm p}\_{\rm in}}}$ at about θin = 25π/36. Therefore, Pedge reaches the maximum near (θin, θout) = (25/36π, 0) at 0.1 ×, 1.0 ×, and 1.5 × Ith_cal_θ=0. However, at 3.0 × Ith_cal_θ=0, Pedge reaches its maximum under the condition (θin, θout) = (0, 0).

Funding

Program for Creating STart-ups from Advanced Research and Technology (JPMJST2032, JPMJST2111); Japan Society for the Promotion of Science (18H01479, 21H01373, 22H01988).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. D. A. B. Miller, “Device requirements for optical interconnects to silicon chips,” Proc. IEEE 97(7), 1166–1185 (2009). [CrossRef]  

2. C. Xiang, W. Jin, D. Huang, M. A. Tran, J. Guo, Y. Wan, W. Xie, G. Kurczveil, A. M. Netherton, D. Liang, H. Rong, and J. E. Bowers, “High-performance silicon photonics using heterogeneous integration,” IEEE J. Select. Topics Quantum Electron. 28(3), 1–15 (2022). [CrossRef]  

3. Y. Zhang, X. Xiao, K. Zhang, S. Li, A. Samanta, Y. Zhang, K. Shang, R. Proietti, K. Okamoto, and S. J. B. Yoo, “Foundry-enabled scalable all-to-all optical interconnects using silicon nitride arrayed waveguide router interposers and silicon photonic transceivers,” IEEE J. Select. Topics Quantum Electron. 25(5), 1–9 (2019). [CrossRef]  

4. N. C. Harris, D. Bunandar, M. Pant, G. R. Steinbrecher, J. Mower, M. Prabhu, T. Baehr-Jones, M. Hochberg, and D. Englund, “Large-scale quantum photonic circuits in silicon,” Nanophotonics 5(3), 456–468 (2016). [CrossRef]  

5. R. Claps, D. Dimitropoulos, V. Raghunathan, Y. Han, and B. Jalali, “Observation of stimulated Raman amplification in silicon waveguides,” Opt. Express 11(15), 1731–1739 (2003). [CrossRef]  

6. R. L. Espinola, J. I. Dadap, R. M. Osgood, S. J. McNab, and Y. A. Vlasov, “Raman amplification in ultrasmall silicon-on-insulator wire waveguides,” Opt. Express 12(16), 3713–3718 (2004). [CrossRef]  

7. A. Liu, H. Rong, M. Paniccia, O. Cohen, and D. Hak, “Net optical gain in a low loss silicon-on-insulator waveguide by stimulated Raman scattering,” Opt. Express 12(18), 4261–4268 (2004). [CrossRef]  

8. Q. Xu, V. R. Almeida, and M. Lipson, “Time-resolved study of Raman gain in highly confined silicon-on-insulator waveguides,” Opt. Express 12(19), 4437–4442 (2004). [CrossRef]  

9. M. Krause, H. Renner, and E. Brinkmeyer, “Analysis of Raman lasing characteristics in silicon-on-insulator waveguides,” Opt. Express 12(23), 5703–5710 (2004). [CrossRef]  

10. H. Rong, R. Jones, A. Liu, O. Cohen, D. Hak, A. Fang, and M. Paniccia, “A continuous-wave Raman silicon laser,” Nature 433(7027), 725–728 (2005). [CrossRef]  

11. H. Rong, S. Xu, Y. Kuo, V. Sih, O. Cohen, O. Raday, and M. Paniccia, “Low-threshold continuous-wave Raman silicon laser,” Nat. Photonics 1(4), 232–237 (2007). [CrossRef]  

12. X. Yang and C. W. Wong, “Coupled-mode theory for stimulated Raman scattering in high-Q/Vm silicon photonic band gap defect cavity lasers,” Opt. Express 15(8), 4763–4780 (2007). [CrossRef]  

13. X. Checoury, Z. Han, and P. Boucaud, “Stimulated Raman scattering in silicon photonic crystal waveguides under continuous excitation,” Phys. Rev. B 82(4), 041308 (2010). [CrossRef]  

14. Y. Takahashi, Y. Inui, M. Chihara, T. Asano, R. Terawaki, and S. Noda, “A micrometre-scale Raman silicon laser with a microwatt threshold,” Nature 498(7455), 470–474 (2013). [CrossRef]  

15. M. Kuwabara, S. Noda, and Y. Takahashi, “Ultrahigh-Q photonic nanocavity devices on a dual thickness SOI substrate operating at both 1.31- and 1.55-μm telecommunication wavelength bands,” Laser Photon. Rev. 13(2), 1800258 (2019). [CrossRef]  

16. H. Okada, M. Fujimoto, N. Tanaka, Y. Saito, T. Asano, S. Noda, and Y. Takahashi, “1.2-µm-band ultrahigh-Q photonic crystal nanocavities and their potential for Raman silicon lasers,” Opt. Express 29(15), 24396–24410 (2021). [CrossRef]  

17. T. Yasuda, M. Okano, M. Ohtsuka, M. Seki, N. Yokoyama, and Y. Takahashi, “Raman silicon laser based on a nanocavity fabricated by photolithography,” OSA Continuum 3(4), 814–823 (2020). [CrossRef]  

18. S. Yasuda, Y. Takahashi, T. Asano, Y. Saito, K. Kikunaga, D. Yamashita, S. Noda, and Y. Takahashi, “Detection of negatively ionized air by using a Raman silicon nanocavity laser,” Opt. Express 29(11), 16228–16240 (2021). [CrossRef]  

19. Y. Takahashi, S. Yasuda, M. Fujimoto, T. Asano, K. Kikunaga, S. Noda, and Y. Takahashi, “Oscillation interruption of a Raman silicon nanocavity laser induced by positively ionized-air irradiation,” in Proc. of Nonlinear Optics (2021), paper NF2B.1.

20. Y. Yamauchi, M. Okano, H. Shishido, S. Noda, and Y. Takahashi, “Implementing a Raman silicon nanocavity laser for integrated optical circuits by using a (100) SOI wafer with a 45-degree-rotated silicon top layer,” OSA Continuum 2(7), 2098–2112 (2019). [CrossRef]  

21. B. S. Song, T. Asano, Y. Akahane, and S. Noda, “Role of interfaces in heterophotonic crystals for manipulation of photons,” Phys. Rev. B 71(19), 195101 (2005). [CrossRef]  

22. H. Takano, B. S. Song, T. Asano, and S. Noda, “Highly efficient multi-channel drop filter in a two-dimensional hetero photonic crystal,” Opt. Express 14(8), 3491–3496 (2006). [CrossRef]  

23. A. Shinya, S. Mitsugi, E. Kuramochi, and M. Notomi, “Ultrasmall multi-port channel drop filter in two-dimensional photonic crystal on silicon-on-insulator substrate,” Opt. Express 14(25), 12394–12400 (2006). [CrossRef]  

24. Y. Sato, Y. Tanaka, J. Upham, Y. Takahashi, T. Asano, and S. Noda, “Strong coupling between distant photonic nanocavities and its dynamic control,” Nat. Photonics 6(1), 56–61 (2012). [CrossRef]  

25. R. Konoike, H. Nakagawa, M. Nakadai, T. Asano, Y. Tanaka, and S. Noda, “On-demand transfer of trapped photons on a chip,” Sci. Adv. 2(5), e1501690 (2016). [CrossRef]  

26. M. Nakadai, T. Asano, and S. Noda, “Electrically controlled on-demand photon transfer between high-Q photonic crystal nanocavities on a silicon chip,” Nat. Photonics 16(2), 113–118 (2022). [CrossRef]  

27. Y. Saito, T. Asano, S. Noda, and Y. Takahashi, “Design Characteristics of a Raman Silicon Nanocavity Laser for Efficient Emission of Light Into an Adjacent Waveguide,” in Proc. of Nonlinear Optics (2021), paper NM1A. 6.

28. B. S. Song, S. Noda, T. Asano, and Y. Akahane, “Ultra-high-Q photonic double-heterostructure nanocavity,” Nat. Mater. 4(3), 207–210 (2005). [CrossRef]  

29. Y. Takahashi, R. Terawaki, M. Chihara, T. Asano, and S. Noda, “First observation of Raman scattering emission from silicon high-Q photonic crystal nanocavities,” in Proc. of Conference on Lasers and Electro-Optics (2011), paper QWC3.

30. D. Yamashita, Y. Takahashi, T. Asano, and S. Noda, “Raman shift and strain effect in high-Q photonic crystal silicon nanocavity,” Opt. Express 23(4), 3951–3959 (2015). [CrossRef]  

31. Y. Takahashi, Y. Inui, M. Chihara, T. Asano, R. Terawaki, and S. Noda, “High-Q resonant modes in a photonic crystal heterostructure nanocavity and applicability to a Raman silicon laser,” Phys. Rev. B 88(23), 235313 (2013). [CrossRef]  

32. J. Kurihara, D. Yamashita, N. Tanaka, T. Asano, S. Noda, and Y. Takahashi, “Detrimental Fluctuation of Frequency Spacing Between the Two high-quality resonant modes in a Raman silicon nanocavity laser,” IEEE J. Select. Topics Quantum Electron. 26(2), 1–12 (2020). [CrossRef]  

33. H. Hagino, Y. Takahashi, Y. Tanaka, T. Asano, and S. Noda, “Effects of fluctuation in air hole radii and positions on optical characteristics in photonic crystal heterostructure nanocavities,” Phys. Rev. B 79(8), 085112 (2009). [CrossRef]  

34. H. Sekoguchi, Y. Takahashi, T. Asano, and S. Noda, “Photonic crystal nanocavity with a Q-factor of ∼9 million,” Opt. Express 22(1), 916–924 (2014). [CrossRef]  

35. T. Asano, Y. Ochi, Y. Takahashi, K. Kishimoto, and S. Noda, “Photonic crystal nanocavity with a Q factor exceeding eleven million,” Opt. Express 25(3), 1769–1777 (2017). [CrossRef]  

36. Y. Taguchi, Y. Takahashi, Y. Sato, T. Asano, and S. Noda, “Statistical studies of photonic heterostructure nanocavities with an average Q factor of three million,” Opt. Express 19(12), 11916–11921 (2011). [CrossRef]  

37. T. Liang and H. Tsang, “Role of free carriers from two-photon absorption in Raman amplification in silicon-on-insulator waveguides,” Appl. Phys. Lett. 84(15), 2745–2747 (2004). [CrossRef]  

38. D. Yamashita, Y. Takahashi, J. Kurihara, T. Asano, and S. Noda, “Lasing dynamics of optically-pumped ultralow-threshold Raman silicon nanocavity lasers,” Phys. Rev. Appl. 10(2), 024039 (2018). [CrossRef]  

39. Y. Tanaka, S. Takayama, T. Asano, Y. Sato, and S. Noda, “A polarization diversity two-dimensional photonic-crystal device,” IEEE J. Select. Topics Quantum Electron. 16(1), 70–76 (2010). [CrossRef]  

40. D. Yamashita, T. Asano, S. Noda, and Y. Takahashi, “Strongly asymmetric wavelength dependence of optical gain in nanocavity-based Raman silicon lasers,” Optica 5(10), 1256–1263 (2018). [CrossRef]  

41. T. Kawakatsu, T. Asano, S. Noda, and Y. Takahashi, “Sub-100-nW-threshold Raman silicon laser designed by a machine-learning method that optimizes the product of the cavity Q-factors,” Opt. Express 29(11), 17053–17068 (2021). [CrossRef]  

42. T. Asano and S. Noda, “Optimization of photonic crystal nanocavities based on deep learning,” Opt. Express 26(25), 32704–32716 (2018). [CrossRef]  

43. Y. Ota, M. Okano, and Y. Takahashi, “Increasing the Q-factor-product and efficiency of Raman silicon nanocavity lasers fabricated by photolithography,” in Proc. CLEO-PR (2022), paper CFA12F_02.

44. Y. Zhang, K. Zhong, and H. K. Tsang, “Raman Lasing in Multimode Silicon Racetrack Resonators,” Laser Photon. Rev. 15(2), 2000336 (2021). [CrossRef]  

45. M. Shinkawa, N. Ishikura, Y. Hama, K. Suzuki, and T. Baba, “Nonlinear enhancement in photonic crystal slow light waveguides fabricated using CMOS-compatible process,” Opt. Express 19(22), 22208–22218 (2011). [CrossRef]  

46. X. Checoury, Z. Han, M. El Kurdi, and P. Boucaud, “Deterministic measurement of the Purcell factor in microcavities through Raman emission,” Phys. Rev. A 81(3), 033832 (2010). [CrossRef]  

47. A. Chutinan, M. Mochizuki, M. Imada, and S. Noda, “Surface-emitting channel drop filters using single defects in two-dimensional photonic crystal slabs,” Appl. Phys. Lett. 79(17), 2690–2692 (2001). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (14)

Fig. 1.
Fig. 1. Device structure of the HM Raman laser. (a) Schematic of the heterostructure nanocavity and the two adjacent waveguides. (b) Energy diagram of the heterostructure nanocavity. (c) Entire HM-laser device layout. (d) and (e) clarify the structures of the photonic mirrors with a heterointerface for reflection in the cases of pump and Stokes waveguides, respectively.
Fig. 2.
Fig. 2. Details of the results of HM Raman laser cavity #1. (a) Resonance spectrum of the pump mode (closed circles) and the corresponding transmission spectrum (open circles). The solid curve is the fitting result. (b) The resonance and transmission spectra of the Stokes mode. (c) Pedge and Psurface as a function of I. (d)−(g) NIR camera images for I = 1.2 Ith obtained by using a long-pass filter to remove the pump light. The images show (d) the edge emission at the output port, (e) the scattered light at the HM of the Stokes waveguide, (f) the scattered light at the output port, and (f) the surface emission from the nanocavity.
Fig. 3.
Fig. 3. Details of the results of conventional Raman laser cavity #5. (a) Resonance spectrum of the pump mode (closed circles) and the corresponding transmission spectrum (open circles). The solid curve is the fitting result. (b) The resonance and transmission spectra of the Stokes mode. (c) Input–output characteristics of the laser. (d)−(f) NIR camera images of (d) the edge emission at the output port, (e) the scattered light at output port, and (f) the surface emission from nanocavity obtained at I = 1.2 Ith.
Fig. 4.
Fig. 4. (a) Normalized input–output characteristics of four HM Raman samples (cavities #1∼#4), (b) Normalized input–output characteristics of five conventional Raman lasers (cavities #5∼#9).
Fig. 5.
Fig. 5. Calculation model of the Raman silicon nanocavity laser with two HMs.
Fig. 6.
Fig. 6. Theoretical input–output characteristics for (θin, θout) = (0, 0). (a) Pedge, Psurface, and Np as a function of |S + 1|2 for a HM Raman laser and (b) the results for a conventional Raman laser.
Fig. 7.
Fig. 7. (a)−(d) Calculated input-output characteristics of the HM Raman laser for (θin, θout) = (0, 0), (0, π), (π, 0), and (π, π), respectively. (e)−(h) Dependence of Pedge on the (θin, θout) for four different values of |S + 1|2 (0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0). (i) (θin, θout) dependence of Psurface at 3.0 × Ith_cal_θ=0. (j) Total Stokes output at 3.0 × Ith_cal_θ=0.
Fig. 8.
Fig. 8. (a) Calculated normalized input–output characteristics of HM Raman lasers for θin = 0 and seven different values of θout. (b) The input–output characteristics for θout = 0 and seven different values of θin.
Fig. 9.
Fig. 9. (θin, θout) dependence of Pedge_HM / Pedge_conv. To prepare this plot, we used the Pedge_HM values for 3.0 × Ith_cal_θ=0 and the Pedge_conv values for 3.0 × Ith_cal. The red regions show conditions where Pedge_HM / Pedge_conv > 1.0.
Fig. 10.
Fig. 10. Calculation results for different Q-value sets. The left column shows the dependence of Pedge on (θin, θout) at 3.0 × Ith_cal_θ=0 for (a) regime 2, (c) regime 3, and (e) regime 4. The right column shows Pedge, Psurface, and Np as a function of |S + 1|2 for (b) regime 2, (d) regime 3, and (f) regime 4. Here, we used the condition (θin, θout) = (0, 0).
Fig. 11.
Fig. 11. Measurement system.
Fig. 12.
Fig. 12. (a)−(d) Dependence of Pedge on (θin, θout) in regime 2 for 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.
Fig. 13.
Fig. 13. (a)−(d) Dependence of Pedge on (θin, θout) in regime 3 for 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.
Fig. 14.
Fig. 14. (a)−(d) Dependence of Pedge on (θin, θout) in regime 4 for 0.1 × Ith_cal_θ=0, 1.0 × Ith_cal_θ=0, 1.5 × Ith_cal_θ=0, and 3.0 × Ith_cal_θ=0, respectively.

Tables (4)

Tables Icon

Table 1. Measurement results for four HM Raman lasers (λp, λS: resonance wavelengths, Δf: frequency spacing, Qp_exp and QS_exp: actual Q values, Ith: threshold, Pedge_max: maximum Pedge)

Tables Icon

Table 2. Measurement results for five conventional Raman lasers (λp, λS: resonance wavelengths, Δf: frequency spacing, Qp_exp and QS_exp: actual Q values, Ith: threshold, Pedge_max: maximum Pedge)

Tables Icon

Table 3. Ith_cal and Np_th for the seven input–output relations displayed in Fig. 8

Tables Icon

Table 4. Parameters used in the calculations

Equations (26)

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

1 Q i _ total = 1 Q i _ design + 1 Q i _ in ,
1 Q i _ exp = 1 Q i _ design + 1 Q i _ in + 1 Q i _ scat + 1 Q i _ abs .
1 Q i _ v = 1 Q i _ design + 1 Q i _ scat  +  1 Q i _ abs .
N p = ω p Q p _ i n ( ω ω p ) 2 + { ω p 2 Q p _ v + ω p 2 Q p _ i n + g R cav ( N S + 1 ) } 2 | S + 1 | 2 ω p ,
N S = 2 g R cav N P ω S Q S _ v + ω S Q S _ i n 2 g R cav N P .
Q p _ i n = Q p _ i n 1 + cos θ in ,
Q S _ i n = Q S _ i n 1 + cos θ out .
ω p = ω p ( 1 + sin θ in 2 Q p _ i n ) .
P edge = | S  - 4 | 2 = ω S Q S _ i n N S ω S .
P surface = 1 2 ω S Q S _ v N S ω S .
1 Q i _ t o t a l = 1 Q i _ v + 1 Q i _ i n   .
I th ( g R cav Q p _ e x p Q S _ e x p ) 1 .
d a p d t = ( j ω p ω p 2 Q v _ p ω p 2 Q i n _ p ) a p g R cav ( N S + 1 ) a p + ω p 2 Q i n _ p e j β d 1 S + 1 + ω p 2 Q i n _ p e j β d 2 S + 2 ,
d a S d t = ( j ω s ω s 2 Q v _ S ω s 2 Q i n _ S ) a s + g R cav N P a s + ω s 2 Q i n _ S e j β d 3 S + 3 + ω s 2 Q i n _ S e j β d 4 S + 4 .
N i = | a i | 2 ω i ( i  = p,S) .
g R cav = ω p c 2 g R _ S i 2 n p n S V R
S 1 = e j β ( d 1 + d 2 ) S + 2 e j β d 1 ω p 2 Q i n _ p a p ,
S 2 = e j β ( d 1 + d 2 ) S + 1 e j β d 2 ω p 2 Q i n _ p a p ,
S 3 = e j β ( d 3 + d 4 ) S + 4 e j β d 3 ω s 2 Q i n _ S a s ,
S 4 = e j β ( d 3 + d 4 ) S + 3 e j β d 4 ω s 2 Q i n _ S a s .
S + 2 = S 2 e j Δ in ,
S + 3 = S 3 e j Δ out .
θ in = 2 β d 2 + Δ in .
θ out = 2 β d 3 + Δ out .
1 Q i _ e x p = 1 Q i _ v + 1 Q i _ i n .
1 Q i _ v = Q i _ e x p T
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.