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

Continuous-wave terahertz quantum cascade microlaser arrays operating on various bound states in the continuum

Open Access Open Access

Abstract

We report on the continuous-wave (CW) operation of 1D terahertz quantum cascade (THz QC) microlaser arrays working on various bound states in the continuum (BICs). We first created a quasi-BIC state by breaking the inversion symmetry of the microlaser array, which enables flexible control of the radiation efficiency. The optimized multi-periods array exhibits single-mode emission with the maximum output power of 21 mW (at 30 K), and the maximum operation temperature (Tcw) of 45 K. To further increase Tcw, we created a hybrid-BIC state by hybridizing a quasi-BIC generated in a few-periods array and a high-Q surface plasmon polariton mode formed in an unbiased array. The hybridization minimizes the pumping area with no obvious degradation of the threshold current density. The reduced pumping area, together with the discrete distribution of microlasers, significantly decreases the device thermal resistance. Such scheme improves the Tcw up to 79 K with a low beam divergence of 17°×17°, and the output power remains 3.4 mW at 20 K. Our work provides a novel solution to control the output power, the operating temperature, and the beam quality of THz QC lasers in CW mode.

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

1. Introduction

Electrically pumped continuous-wave (EPCW) operation is one of the most demanded characteristics for semiconductor lasers. Electric pumping is the key for miniaturization, integration, and practicability, and continuous-wave (CW) operation is indispensable for ultranarrow linewidth and ultra-stable output power, which are the core of applications including optical communication, sensing, spectral analysis, metrology [1], etc. EPCW semiconductor lasers – including multi-quantum well lasers [2], quantum cascade lasers [3], and interband cascade lasers [4] – usually operate with high input power density, and the nanometer-thick multilayers of the active region imply low vertical thermal conductivity [5,6]. Therefore, the power consumption and thermal conduction are the fundamental issues that limit the laser performances such as operating temperature and output power. The situation becomes more challenging at long operation wavelength because the device size and thus the heat accumulation increases dramatically with the wavelength [79].

Up to now, THz QC lasers generate coherent radiation with the longest available wavelengths in the context of semiconductor lasers [1013]. Innovations of the active region and the laser resonator have witnessed substantial progresses of THz QCLs, including near room temperature operation in pulsed mode [1416], high output power and high beam quality [1721]. Nevertheless, the maximum CW operating temperature (Tcw) has been stuck at cryogenic temperature for many years [12,22,23], mainly because of the high input power density and the poor thermal conductivity of the active region. Recent advances to improve Tcw focus on the wire lasers, where the ridge width is much less than the wavelength [2328]. The wire lasers enable small pumping area but relatively large specific surface area, which reduces the power consumption and facilitates the lateral heat dissipation. Novel resonators based on the wire laser have been developed to realize high CW output power and tight beam pattern. Successful examples include third-order distributed feedback (DFB) THz QC lasers [23,24], phased array of antenna coupled third-order DFB lasers [25], unidirectional DFB lasers [26], wire lasers hybridizing a DFB grating with an emission grating [27], metasurface-based vertical-external-cavity surface-emitting lasers [28], etc. Nevertheless, there is still huge room for the development in terms of the Tcw, the output power, and the beam quality.

Recently, optical bound states in the continuum (BICs) or quasi-BICs have attracted substantial attention as versatile tools to manipulate the photon behaviors [2932], and have greatly boosted the research of ultra-low threshold lasers [33], sensing [34], nonlinear optics [35], topological photonics [36], etc. Optical BICs refer to the special electromagnetic modes that locate inside the continuum and coexist with extended waves, but they remain perfectly confined without any radiation. BICs are thus spatially bound and spectrally discrete with an infinite quality factor (Q-factor). True BICs exist only in an infinite periodic structure due to either symmetric incompatibility (symmetry-protected BICs) or destructive interference (non-symmetry-protected BICs). In a finite system and by symmetry breaking, BICs can be converted into quasi-BICs with controllable radiation quality factor (Qrad) [2933].

Symmetry-protected BICs were also regarded as the non-radiative band-edge modes and used in surface-emitting semiconductor lasers, including DFB or photonic crystal (PhC) lasers [3739]. Recent efforts intentionally break the inversion symmetry to excite the quasi-BICs with controllable Qrad, which leads to high output power, high beam quality, and still relatively low threshold [4042]. However, all these efforts were conducted in multi-periods lasers where the resonator (grating or PhC) contains many periods. By minimizing the laser resonator – which is important for EPCW operation – the quasi-BICs inevitably suffer from severe in-plane leakage [43]. Although super-BICs or accident-BICs have recently been proposed that enable high Qrad in few-periods or subwavelength resonators [31,44], they were only demonstrated in passive devices or optically pumped lasers [45].

In this work, we demonstrate the CW operation of 1D THz QC microlaser arrays that work on various BIC states. First, in a multi-periods microlaser array, we prove that symmetry breaking can activate a quasi-BIC as the laser mode, leading to high output power with a tight beam under the CW condition. To further increase Tcw, we hybridize the quasi-BIC generated in a biased few-periods microlaser array with a high Q-factor spoof surface plasmon polariton (SPP) mode formed in another unbiased array. The so-called hybrid-BIC greatly reduces the pumping area while keeping the in-plane leakage substantially low. Moreover, the discretely distributed microlasers considerably reduce the device thermal resistance. The resultant microlaser array exhibits single-mode emission with significantly improved Tcw, high beam quality, and moderate output power under the CW condition.

2. Device concept, design, and results

We first investigate the quasi-BIC in the multi-periods THz QC microlaser arrays, and then the hybrid-BIC in the few-periods (biased) microlaser arrays. For simplicity, we name them as the quasi-BIC and hybrid-BIC microlaser arrays, respectively.

2.1 Quasi-BIC microlaser arrays

Figures 1(a) and (b) respectively show the schematic of a quasi-BIC microlaser array, and the 2D cross-section of one period. The device is based on a metal-metal waveguide, where the active region is sandwiched between the bottom and top metallic layers. The device consists of N periods with a periodicity of Λ. A single unit contains two microlasers, each of which is deep-subwavelength in thickness (tAR, z-direction) and in width (x-direction) but is elongated in length (Ly, y-direction). The widths of the two microlasers in each unit are respectively w + δw and w-δw, where δw<<w, hence breaking the in-plane inversion symmetry. For current injection, each microlaser is linked to the electric pads via two tapered waveguides along the y-direction. In the region of the electric pads and the tapered waveguides, there is a SiO2 isolating layer underneath the top metallization, and therefore the current injection occurs only in the microlasers.

 figure: Fig. 1.

Fig. 1. Operation principle of a quasi-BIC microlaser array. (a) 3D schematic of the quasi-BIC array. (b) 2D cross-section of one period. The array contains N periods with the periodicity of Λ. Each period contains two microlasers with slightly different widths (w + δw and w-δw). Each microlaser is tAR in thickness, Ly in length, and Λ/2 (center-to-center) from the neighbor. The red and blue regions refer to the biased and unbiased active regions, respectively. (c) Calculated photonic band structure of the array (Λ=60.0µm, w = 12.1µm, δw = 0.3µm), where the quasi-BIC is marked as the red point. (d) Electric fields (Ez and Ex) of the quasi-BIC in one periodic unit. (e) Dependence of Qcav on the asymmetric factor (δw/w) for the quasi-BIC state in an infinite-periods array, a multi-periods (N = 22) array, and a few-periods (N = 3) array. Note that logarithmic axes are used for both horizontal and vertical axes. (f) The calculated αrad, gth, and αradtot as a function of δw/w for a multi-periods quasi-BIC array (N = 22, Λ=60.0µm, w = 12.1µm).

Download Full Size | PDF

In the quasi-BIC array, w is chosen so that the microlaser supports a dipole mode oscillating in x-direction. The unbiased tapered waveguides act as absorbing boundaries and eliminate high-order modes in y-direction. Here, we use tapered waveguide since – compared to the straight waveguide – it slightly improves the mode confinement along the y-direction, mainly due to the impedance mismatch between the microlaser and the tapered waveguide. As such, each microlaser acts as a dipole antenna and delivers significant radiation in the vertical direction, and because of the fringing effect, the electric field extends considerably out of the microlaser and contributes to the coupling between the adjacent microlasers. On the other hand, Λ is chosen to make the device operate on the Γ-point of the band structure, as shown in Fig. 1(c). Figure 1(d) presents the electric field of the quasi-BIC [marked by the red point in Fig. 1(c)] in one unit. The numerical calculations were conducted by means of full-wave finite element method (see the Method section). Evidently, the broken symmetry makes the electric field (Ex, the dominant radiation field) of the quasi-BIC not perfectly odd symmetric and thus couple to the free space in a controllable manner. Figure 1(e) plots the relationship between the cavity quality factor (Qcav) and the asymmetric factor (δw/w) for an infinite-periods array, a multi-periods (N = 22) array and a few-periods (N = 3) array, respectively. Here, ${({{Q_{cav}}} )^{ - 1}} = {({{Q_\parallel }} )^{ - 1}} + {({{Q_{rad}}} )^{ - 1}}$, where Q// is the in-plane Q-factor. For the infinite-periods array (${Q_{cav}} = {Q_{rad}}$), the relationship between Qrad and δw/w perfectly consists with the perturbation theory [32]:

$${Q_{rad}} = {Q_0}{({\delta w/w} )^{ - 2}}$$
where Q0 is a constant. For the multi-periods array (N = 22), Qcav is dominated by Q// for small δw/w, and is dominated by Qrad for large δw/w. However, for the few-periods array (N = 3), Qcav is predominantly determined by Q// and is too low to lase. It is interesting to recognize the similarities and differences between the quasi-BIC array and the second-order grating. Both structures operate on the Γ-point of the photonic band structure, and thus deliver surface emission. However, the inversion symmetry is broken in the quasi-BIC array but is maintained in the second-order grating. Moreover, the width of the microlaser is determined to support a dipole mode, whereas in the usual second-order grating there is no rigorous requirement on the width of the high- or low-index region.

In a practical THz QC microlaser array, the waveguide loss (αw), the in-plane loss (α//), and the radiation loss (αrad) must be judiciously designed to balance the threshold and the power efficiency. Because a significant part of the electromagnetic (EM) field extends out of the microlasers, an accurate estimation of the effective index of the microlaser array is difficult. For this reason, a simple relationship between the loss components and the quality factors is not available. In the Supplement 1, we detail the method to numerically calculate the relevant quantities of the quasi-BIC array. For a quasi-BIC array (N = 22) with designed operation frequency of 3.2 THz, Fig. 1(f) illustrates the influence of δw/w on the αrad, the threshold gain (gth), and the ratio αradtot (αtot is the total loss, ${\alpha _{tot}} = {\alpha _w} + {\alpha _\parallel } + {\alpha _{rad}}$, and αradtot is proportional to the slope efficiency of the output power). When δw increases from 0 to 0.4 µm, αrad increases significantly from 0.05 cm−1 to 5.24 cm−1, confirming that the symmetry breaking is a powerful tool to tailor the power extraction. However, due to the relatively low optical confinement factor Γ (∼21%), gth also increases from 20.1 cm−1 to 44.5 cm−1. Figure 1(f) also indicates that the increase of αradtot is not as significant as that of αrad, because of the small and constant αw. Table S1 in the Supplement 1 lists the structure parameters and the calculated physical quantities of an optimized quasi-BIC array, aiming at the high CW output power.

To force the device operate on the fundamental quasi-BIC mode, the number of microlasers (N), the length of the microlasers (Ly) and the length of the unbiased tapered waveguide need to be carefully optimized to make its threshold gain sufficiently lower than the high-order modes. Fig. S4 in the Supplement 1 shows the dependence of operation frequency of the fundamental quasi-BIC mode on the width (w) or the periodicity (Λ) of the array, indicating that lithographically tunable single-mode emission is expectable in the quasi-BIC devices.

In experiments, the gain of the active region covers the frequency range from 2.8 to 3.5 THz and peaks near 3.15 THz. Quasi-BIC devices with different w, δw and Λ were systematically fabricated and measured. For all the devices, the values of N and Ly are 22 and 400 µm, respectively. Details of the device fabrication and measurements are described in the Method section.

To study the mechanism of the quasi-BIC, we first measured the quasi-BIC microlaser arrays in pulsed mode, and the results are presented in Fig. 2. Figure 2(a) presents the SEM picture of a quasi-BIC microlaser array. Figures 2(b) and 2(c) show the lasing spectra and light-current-voltage (L-I-V) curves of 4 devices with different w but the same δw (0.1 µm) and Λ (60.0 µm). Single mode emission is observed in all devices in the whole laser dynamic range, and the lasing frequency varies from 3.19 THz to 3.31 THz when w decreases from 12.5 µm to 11.9 µm. The L-I-V curves of these 4 devices illustrate very close peak power around 35 mW. We also investigated the spectra and L-I-V curves of another set of quasi-BIC devices, where w and δw are respectively fixed as 12.1 µm and 0.1 µm, but Λ varies from 56.0 µm to 64.0 µm, and the results are shown in Fig. S5 of Supplement 1. Again, stable single-mode emissions and similar output powers were observed. Fig. S4 in Supplement 1 compares the calculated and measured relationship between the laser frequency and the parameters of w and Λ, confirming good agreement.

 figure: Fig. 2.

Fig. 2. Quasi-BIC microlaser arrays operating in pulsed mode (the pulse width is 1µs, the repeat frequency is 10 kHz) at 20 K. (a) SEM picture of a microlaser array. (b) and (c) show the measured emission spectra and L-I-V curves of 4 quasi-BIC devices with different w marked in the figure (δw = 0.1 µm, Λ=60.0 µm). Emission spectra at different injection current are given. (d) and (e) plot the measured far-field beam pattern and polarization of a quasi-BIC device (w = 12.1 µm, δw = 0.3 µm, Λ=60.0 µm). (f) Measured L-I-V curves of 5 quasi-BIC devices, where δw varies from 0 to 0.4µm with an interval of 0.1µm (w = 12.1 µm, Λ=60.0 µm). The threshold current densities of these 5 devices are 375, 395, 408, 422, 435 A cm−2, respectively. The corresponding slope efficiencies are 27, 102, 286, 545, 511 mW A−1, respectively.

Download Full Size | PDF

Figures 2(d) and 2(e) show the measured 2D far field beam and the polarization of a representative device (w = 12.1 µm, δw = 0.3 µm, Λ = 60.0 µm). Single-lobed vertical emission with a low divergence of 5°× 15° was observed, which is linearly polarized along the array direction. Fig. S1 in Supplement 1 presents the calculated beam pattern, as well as the definition of the scanning angles during the measurement of beam pattern. Perfect agreement between the measurement and calculation was observed. The results confirm that the devices operate on the fundamental quasi-BIC mode.

Figure 2(f) shows the pulsed L-I-V curves of 5 devices, where w is fixed as 12.1 µm and δw increases from 0 to 0.4 µm, measured at the heat-sink temperature of 20 K. With the increase of δw, the slope efficiency (dP/dI) increases from 27 mW A−1 to 511 mW A−1, validating the influence of symmetry breaking on the radiation efficiency. Because of the increased αrad, the pulsed threshold current density (Jth,p) increases from 375 A cm−2 to 435 A cm−2, and the maximum pulsed operating temperature (Tp) decreases from 130 K to 90 K. Figure 2(f) also points out that, after the threshold, the differential resistance (dV/dI) increases with δw which shortens the laser dynamic range. The reason is that the increased αrad decreases the photonic density in the resonator and suppresses the photon assisted electrical transition. Combining the slope efficiency and the laser dynamic range, the pulsed output power and the wall plug efficiency (WPE) reach to the maximum (120 mW and 1%) when δw is 0.3 µm, as shown in Fig. 2(f).

We systematically investigated the spectral and L-I-V characteristics of an optimized quasi-BIC array (w = 12.1 µm, δw = 0.2 µm, Λ = 60.0 µm) in pulsed and CW mode, and the results are presented in Fig. 3. Figures 3(a) and 3(b) plots the pulsed and CW emission spectrum, respectively, confirming single-mode emission with the SMSR above 22 dB. Figures 3(c) and 3(d) show the pulsed and CW L-I-V curves at different operation temperatures. In pulsed mode and at 20 K, the threshold current and threshold current density (Ith,p and Jth,p) are respectively 0.86 A and 407 A cm−2, the maximum output power (Pout,p) is 109 mW, the power consumption (Pin,p) is 13.10 W. The maximum pulsed operation temperature (Tp) is 120 K. Whereas, in CW mode and at 30 K, the values of the threshold current (Ith,cw), threshold current density (Jth,cw), the maximum output power (Pout,cw), the power consumption (Pin,cw) are 0.94 A, 452 A cm−2, 21 mW, and 11.10 W, respectively. The maximum CW operation temperature (Tcw) is 45 K. The device performances degrade considerably in CW, implying the severe influence of heat accumulation. Considering the intrinsic low thermal conductivity of the active region, a practical approach to improve Tcw is to reduce the power consumption. However, simply decreasing the number of microlasers in the array will cause severe in-plane leakage, as already implied in Fig. 1(e).

 figure: Fig. 3.

Fig. 3. Comparison of the pulsed and CW operations of a quasi-BIC microlaser array, where w = 12.1 µm, δw = 0.2 µm, Λ=60.0 µm. (a) and (b) are the pulsed and CW emission spectra, respectively. (c) and (d) show respectively the pulsed and CW L-I-V curves measured at different temperatures.

Download Full Size | PDF

2.2 Hybrid quasi-BIC microlaser arrays

To improve Tcw, it is necessary to minimize the pumping area and keep low threshold current density. To this aim, we propose a hybrid-BIC microlaser array. As schematically shown in Figs. 4(a) and 4(b), the hybrid-BIC device contains a biased few-periods quasi-BIC array in the center, and an unbiased array on each side. In the unbiased array, there is a SiO2 isolating layer underneath the top metallic layer of the ridge. The unbiased array consists of Np periods of ridge with the periodicity Λp and the ridge width wp, and supports a high-Q band-edge spoof SPP mode (see the Supplement 1 for more details). Since the SPP mode is far below the light cone, the Ez field is tightly bound on the surface of the array, with only a weak part penetrating into the unbiased active region. Because of the nearly vanishing group velocity, the band-edge SPP mode is perfectly confined with an extremely high Qcav and a high Qtot (see the Supplement 1). Moreover, the frequency of the SPP mode can be precisely tuned by changing the periodicity Λp.

 figure: Fig. 4.

Fig. 4. Concept of the hybrid-BIC microlaser array. (a) and (b) schematics of the device and its 2D cross-section. A biased few-periods quasi-BIC array is surrounded by an unbiased array (periodicity Λp, ridge width wp, number of period Np) on each side. (c) and (d) calculated electric field (Ez and Ex) of the hybrid-BIC. The parameters of the central array are N = 3, Λ=60.0 µm, w = 12.1 µm, δw = 0.1 µm. The parameters of the unbiased array are Λp = 34.3 µm, wp =17.6 µm, and Np =14. The eigenfrequency and the Qcav of the hybrid-BIC are 3.16 THz and 2400, respectively.

Download Full Size | PDF

The device is designed such that the quasi-BIC is in resonance with the high-Q SPP mode, i.e., the two modes have nearly the same frequency and their field profiles match partially with each other. As a result, the field leaking out of the central quasi-BIC array (in the array direction) will be transformed efficiently into the SPP mode in the unbiased array. Thanks to the high Qcav of the latter, the field oscillates in the unbiased array and is reinjected into the central quasi-BIC array. We name this supermode as the hybrid-BIC.

Figures 4(c) and 4(d) show the field distribution of a 3-periods hybrid-BIC (the central quasi-BIC array contains only N = 3 periods), with the device parameters given in the caption of Fig. 4. Figure 4(c) shows that, the Ez field is mainly confined in the near field of the whole structure, which results in a low calculated αw of 1.28 cm−1, a low α// of 0.02 cm−1, but also a low Γ of 5.2%. Because of the low Γ, δw is selected as 0.1 µm to guarantee relatively low αrad (0.1 cm−1) and thus the low threshold. Consequently, the threshold gain of the hybrid-BIC is calculated to be 27 cm−1, which is only slightly higher than that of the optimized 22-periods quasi-BIC array (δw = 0.2 µm). Strikingly, the pumping area of the 3-periods hybrid-BIC device is only about one seventh of the 22-periods quasi-BIC array.

In experiment, we fabricated and measured a series of hybrid-BIC devices with different N (N= 2, 3 or 4), while all the other parameters are fixed and given in the caption of Fig. 5. Figure 5 shows the SEM picture of a 3-periods hybrid-BIC device and its performances measured in pulsed and CW modes, respectively. Figure 5(b) shows the pulsed L-I-V curves of the device. At 20 K, the values of Ith.p, Jth.p, Pin,p and Pout,p are 0.14 A, 413 A cm−2, 2.28 W and 3.64 mW, respectively. Tp reaches to 109 K, close to that of the optimized 22-periods quasi-BIC device [see Fig. 3(c)]. Figure 5(c) shows the pulsed emission spectrum, confirming that the device operates in single mode with side mode suppression ratio (SMSR) about 24 dB.

 figure: Fig. 5.

Fig. 5. Laser performance of a 3-periods hybrid-BIC microlaser array, where Λ=60.0 µm, w = 12.1 µm, δw = 0.1 µm, Λp = 34.3 µm, wp =17.6 µm, and Np =14. (a) SEM pictures of the device. (b) L-I-V curves measured at different temperatures in pulsed mode. (c) shows the pulsed emission spectra. (d) and (e) are L-I-V curves and emission spectra measured in CW mode, respectively. (f) Far-field beam pattern measured at 20 K.

Download Full Size | PDF

Figure 5(d) shows the L-I-V curves of the same device measured in CW mode. At 20 K, the values of Ith,cw, Jth,cw, Pin,cw and Pout,cw are 0.14 A, 422 A cm−2, 2.31 W, and 3.4 mW, respectively. More importantly, Tcw reaches to 79 K. Compared with the optimized 22-periods quasi-BIC device [see Fig. 3(d)], the 3-periods hybrid-BIC device does not degrade the threshold current density, but significantly reduces the threshold current and power consumption on the expense of output power. This explains the substantially increased Tcw for the 3-periods hybrid-BIC device. Figure 5(e) presents the CW emission spectrum, confirming single-mode emission with the SMSR about 29 dB. Figure 5(f) shows the measured beam pattern of the device, which features a main lobe with a low divergence of 17° × 17° and very weak side emission along the array direction. The weak side emission roots from the EM field distributed in the passive array.

2.3 Heat dissipation in the microlaser array

We quantitatively compared the heat dissipation of the quasi-BIC device (N = 22), the hybrid-BIC devices (N = 2, 3 or 4), and a reference FP-MM laser (a Fabry-Perot THz QC laser based on the metal-metal waveguide), all fabricated with the same material. The laser performances of the FP-MM laser and the hybrid-BIC devices (N = 2 or 4) are given in Supplement 1.

For the 3-periods hybrid-BIC device, Fig. 6(a) presents the variation of threshold current density as a function of heat-sink temperature in pulsed and CW modes, respectively. In pulsed mode (the pulse width is 1 µs, the repeat frequency is 10 kHz), the self-heating of the device is negligible and the temperature of the active region equals approximately to the heat-sink temperature. With the same threshold current density, the heat-sink temperatures under the CW condition (Ths,cw) is less than that under the pulsed condition (Ths,p), and the difference ($\Delta T = {T_{hs,p}} - {T_{hs,cw}}$) reflects the temperature increase in the active region due to the self-heating and can be described as [46]

$$\Delta T = {P_{th}} \times {R_{therm}}$$

Here, Pth is the threshold power density, and Rtherm is area-normalized thermal resistance reflecting the temperature increase per unit power density loaded on the device. The triangles in Fig. 6(a) illustrate a transformation from the CW to the pulsed characteristics using Rtherm as the fitting parameter. Figures 6(b) and 6(c) show the situations for the 22-periods quasi-BIC device and the FP-MM laser, respectively. Table 1 lists the deduced Rtherm and other measured parameters of the related devices.

 figure: Fig. 6.

Fig. 6. Dependence of the threshold current density on the operating temperature for the 3-periods hybrid-BIC device (a), the 22-periods quasi-BIC device (b), and the FP-MM laser (c). The black squares and the red circles correspond to the measured Jth as a function of temperature under the pulsed and CW conditions, respectively. The triangles represent the transformation from the CW to the pulsed characteristics with the fitting value of Rtherm.

Download Full Size | PDF

Tables Icon

Table 1. The measured parameters and the deduced thermal resistance of the FP-MM, the quasi-BIC, and the hybrid-BIC devices. The values of Ith,cw, Jth,cw, Pin,cw, and Pout,cw were measured at 30 K for the quasi-BIC device, and measured at 20 K for the others.

The variation tendency of Rtherm reveals the distinctions of heat dissipation among the three types of devices. First, for all devices, the lower Rtherm corresponds to the lower (Tp -Tcw). Therefore, an important direction to increase Tcw is to reduce the area-normalized thermal resistance Rtherm. Second, the values of Rtherm of the quasi-BIC and hybrid-BIC devices are substantially lower than that of the FP-MM device, meaning that the discrete distribution of the heat sources (i.e., microlasers) facilitates the heat dissipation. Third, the less pumping area, the lower Rtherm. Consequently, the few-periods hybrid-BIC microlaser array strongly suggests the laser configuration for high temperature CW operation. Balancing Rtherm and the threshold current density, it turns out that the 3-periods hybrid-BIC device exhibits the highest Tcw of 79 K and its output power remains 3.4 mW in CW mode at 20 K. The comprehensive performances of the optimized hybrid-BIC device are comparable to the state-of-the-art of THz QCLs [2328], in terms of the CW operating temperature, the output power, and the beam quality. It is worth noting that the two parameters – the maximum output power at low temperature (Pout,cw) and the maximum operating temperature (Tcw) – restrict each other. We can improve Pout,cw by increasing the asymmetric factor δw/w, but it will in turn increase the radiation loss and thus the threshold gain, resulting in the decrease of Tcw.

3. Conclusion

In conclusion, we realized continuous wave operation of 1D THz QC microlaser arrays that operate on various BIC states. Our work demonstrates the capability to comprehensively tailor the BIC states, including the radiation and in-plane Q-factors, the mode volume, and the radiation pattern. We also proved that the discrete distribution of the microlasers and the small pumping area facilitates the heat dissipation of the device. The results suggest a novel solution to improve the CW performance of THz QC lasers in terms of the operating temperature, the output power, and the beam quality.

4. Methods

4.1 Simulations

The simulations were implemented by full-wave finite element method (FEM) with a commercial solver of COMSOL Multiphysics. The complex refraction index of the active region and the metallic layers were computed by the Drude model. Frequency dependence of the material index was considered. During the calculation of Qcav, the material losses were neglected, and only the field leakage was considered. During the calculation of Qtot, the material loss, the radiation loss and the in-plane wave leakage are all considered.

4.2 Fabrication

The QCL active region is close to that reported work [47], and contains 180 stages with a total thickness of 11.8 µm. The fabrication was similar to that described in Ref. [46]. The only difference is that, in the unbiased region, an isolating SiO2 layer was deposited on the top of the active region by means of PECVD and dry etching.

4.3 Measurement details

The spectral characteristics of the devices were performed using a Fourier transform infrared spectrometer (Bruker 80 V). The pulsed output power was measured by a Golay cell, and the CW output power was measured by Ophir 3A-P-THz detector. The detectors were calibrated by a Thomas Keating absolute THz power meter. The output power reported in this work is that collected by the detectors, without considering the collection efficiency of the measurement setup which is about 50% [48]. The far-field beam patterns were measured with the Golay cell detector, which was scanned on a 15-cm-radius sphere centered on the device surface.

Funding

National Natural Science Foundation of China (61974151, 12074249, 62235010, 12274285); Natural Science Foundation of Shanghai (20ZR1466200, 21ZR1474000).

Acknowledgments

We acknowledge Dr. Chun Lin for the technical supports on the device fabrications.

Disclosures

The authors declare no conflicts of interest related to this article.

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.

Supplemental document

See Supplement 1 for supporting content.

References

1. G. P. Agrawal and N. K. Dutta, Semiconductor Lasers (Springer, 2013).

2. P. Liao and P. Kelley, Quantum well lasers (Elsevier, 2012).

3. J. Faist, Quantum cascade laser (Oxford University, 2013).

4. R. Q. Yang, “Infrared-laser based on intersub-band transitions in quantum-wells,” Superlattices Microstruct. 17(1), 77–83 (1995). [CrossRef]  

5. G. Chen and M. Neagu, “Thermal conductivity and heat transfer in superlattices,” Appl. Phys. Lett. 71(19), 2761–2763 (1997). [CrossRef]  

6. W. S. Capinski, H. J. Maris, T. Ruf, M. Cardona, K. Ploog, and D. S. Katzer, “Thermal-conductivity measurements of GaAs/AlAs superlattices using a picosecond optical pump-and-probe technique,” Phys. Rev. B 59(12), 8105–8113 (1999). [CrossRef]  

7. A. N. Baranov, M. Bahriz, and R. Teissier, “Room temperature continuous wave operation of InAs-based quantum cascade lasers at 15 mu m,” Opt. Express 24(16), 18799–18806 (2016). [CrossRef]  

8. C. A. Evans, D. Indjin, Z. Ikonic, P. Harrison, M. S. Vitiello, V. Spagnolo, and G. Scamarcio, “Thermal modeling of terahertz quantum-cascade lasers: Comparison of optical waveguides,” IEEE J. Quantum Electron. 44(7), 680–685 (2008). [CrossRef]  

9. S. Fathololoumi, D. Y. Ban, H. Luo, E. Dupont, S. R. Laframboise, A. Boucherif, and H. C. Liu, “Thermal behavior investigation of terahertz quantum-cascade lasers,” IEEE J. Quantum Electron. 44(12), 1139–1144 (2008). [CrossRef]  

10. R. Kohler, A. Tredicucci, F. Beltram, H. E. Beere, E. H. Linfield, A. G. Davies, D. A. Ritchie, R. C. Iotti, and F. Rossi, “Terahertz semiconductor-heterostructure laser,” Nature 417(6885), 156–159 (2002). [CrossRef]  

11. B. S. Williams, “Terahertz quantum-cascade lasers,” Nat. Photonics 1(9), 517–525 (2007). [CrossRef]  

12. G. Z. Liang, T. Liu, and Q. J. Wang, “Recent developments of terahertz quantum cascade lasers,” IEEE J. Sel. Top. Quantum Electron. 23(4), 1 (2017). [CrossRef]  

13. G. Scalari, C. Walther, M. Fischer, R. Terazzi, H. Beere, D. Ritchie, and J. Faist, “THz and sub-THz quantum cascade lasers,” Laser & Photon. Rev. 3(1-2), 45–66 (2009). [CrossRef]  

14. A. Khalatpour, A. K. Paulsen, C. Deimert, Z. R. Wasilewski, and Q. Hu, “High-power portable terahertz laser systems,” Nat. Photonics 15(1), 16–20 (2021). [CrossRef]  

15. L. Bosco, M. Franckie, G. Scalari, M. Beck, A. Wacker, and J. Faist, “Thermoelectrically cooled THz quantum cascade laser operating up to 210 K,” Appl. Phys. Lett. 115(1), 5 (2019). [CrossRef]  

16. A. Khalatpour, M. C. Tam, S. J. Addamane, J. Reno, Z. Wasilewski, and Q. Hu, “Terahertz semiconductor laser source at-12 C,” arXiv, arXiv:2211.08125 (2022). [CrossRef]  

17. Y. Jin, J. L. Reno, and S. Kumar, “Phase-locked terahertz plasmonic laser array with 2 W output power in a single spectral mode,” Optica 7(6), 708–715 (2020). [CrossRef]  

18. H. Zhu, F. F. Wang, Q. Yan, C. R. Yu, J. X. Chen, G. Y. Xu, L. He, L. H. Li, L. Chen, A. G. Davies, E. H. Linfield, J. M. Hao, P. B. Vigneron, and R. Colombelli, “Terahertz master-oscillator power-amplifier quantum cascade lasers,” Appl. Phys. Lett. 109(23), 5 (2016). [CrossRef]  

19. L. H. Li, L. Chen, J. R. Freeman, M. Salih, P. Dean, A. G. Davies, and E. H. Linfield, “Multi-Watt high-power THz frequency quantum cascade lasers,” Electron. Lett. 53(12), 799–800 (2017). [CrossRef]  

20. J. Perez-Urquizo, Y. Todorov, L. H. Li, A. G. Davies, E. H. Linfield, C. Sirtori, J. Madeo, and K. M. Dani, “Monolithic patch-antenna THz lasers with extremely low beam divergence and polarization control,” ACS Photonics 8(2), 412–417 (2021). [CrossRef]  

21. W. J. Wan, H. Li, and J. C. Cao, “Homogeneous spectral broadening of pulsed terahertz quantum cascade lasers by radio frequency modulation,” Opt. Express 26(2), 980–989 (2018). [CrossRef]  

22. B. S. Williams, S. Kumar, Q. Hu, and J. L. Reno, “Operation of terahertz quantum-cascade lasers at 164 K in pulsed mode and at 117 K in continuous-wave mode,” Opt. Express 13(9), 3331–3339 (2005). [CrossRef]  

23. M. Wienold, B. Roben, L. Schrottke, R. Sharma, A. Tahraoui, K. Biermann, and H. T. Grahn, “High-temperature, continuous-wave operation of terahertz quantum-cascade lasers with metal-metal waveguides and third-order distributed feedback,” Opt. Express 22(3), 3334–3348 (2014). [CrossRef]  

24. M. I. Amanti, M. Fischer, G. Scalari, M. Beck, and J. Faist, “Low-divergence single-mode terahertz quantum cascade laser,” Nat. Photonics 3(10), 586–590 (2009). [CrossRef]  

25. A. Khalatpour, J. L. Reno, and Q. Hu, “Phase-locked photonic wire lasers by pi coupling,” Nat. Photonics 13(1), 47–53 (2019). [CrossRef]  

26. A. Khalatpour, J. L. Reno, N. P. Kherani, and Q. Hu, “Unidirectional photonic wire laser,” Nat. Photonics 11(9), 555–559 (2017). [CrossRef]  

27. S. Biasco, K. Garrasi, F. Castellano, L. H. Li, H. E. Beere, D. A. Ritchie, E. H. Linfield, A. G. Davies, and M. S. Vitiello, “Continuous-wave highly-efficient low-divergence terahertz wire lasers,” Nat. Commun. 9(1), 1122 (2018). [CrossRef]  

28. C. A. Curwen, J. L. Reno, and B. S. Williams, “Broadband continuous single-mode tuning of a short-cavity quantum-cascade VECSEL,” Nat. Photonics 13(12), 855–859 (2019). [CrossRef]  

29. S. I. Azzam and A. V. Kildishev, “Photonic bound states in the continuum: from basics to applications,” Adv. Opt. Mater. 9, 2001469 (2021). [CrossRef]  

30. C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljacic, “Bound states in the continuum,” Nat. Rev. Mater. 1(9), 16048 (2016). [CrossRef]  

31. A. Kodigala, T. Lepetit, Q. Gu, B. Bahari, Y. Fainman, and B. Kante, “Lasing action from photonic bound states in continuum,” Nature 541(7636), 196–199 (2017). [CrossRef]  

32. K. Koshelev, S. Lepeshov, M. K. Liu, A. Bogdanov, and Y. Kivshar, “Asymmetric metasurfaces with high-q resonances governed by bound states in the continuum,” Phys. Rev. Lett. 121(19), 193903 (2018). [CrossRef]  

33. M. S. Hwang, H. C. Lee, K. H. Kim, K. Y. Jeong, S. H. Kwon, K. Koshelev, Y. Kivshar, and H. G. Park, “Ultralow-threshold laser using super-bound states in the continuum,” Nat. Commun. 12(1), 4135 (2021). [CrossRef]  

34. A. Leitis, A. Tittl, M. K. Liu, B. H. Lee, M. B. Gu, Y. S. Kivshar, and H. Altug, “Angle-multiplexed all-dielectric metasurfaces for broadband molecular fingerprint retrieval,” Sci. Adv. 5(5), 8 (2019). [CrossRef]  

35. K. Koshelev, S. Kruk, E. Melik-Gaykazyan, J. H. Choi, A. Bogdanov, H. G. Park, and Y. Kivshar, “Subwavelength dielectric resonators for nonlinear nanophotonics,” Science 367(6475), 288–292 (2020). [CrossRef]  

36. Y. Chen, H. Deng, X. Sha, W. Chen, R. Wang, Y.-H. Chen, D. Wu, J. Chu, Y. S. Kivshar, S. Xiao, and C.-W. Qiu, “Observation of intrinsic chiral bound states in the continuum,” Nature 613(7944), 474–478 (2023). [CrossRef]  

37. C. H. Henry, R. F. Kazarinov, R. A. Logan, and R. Yen, “Observation of destructive interference in the radiation loss of 2nd-order distributed feedback lasers,” IEEE J. Quantum Electron. 21(2), 151–154 (1985). [CrossRef]  

38. M. Meier, A. Mekis, A. Dodabalapur, A. Timko, R. E. Slusher, J. D. Joannopoulos, and O. Nalamasu, “Laser action from two-dimensional distributed feedback in photonic crystals,” Appl. Phys. Lett. 74(1), 7–9 (1999). [CrossRef]  

39. M. Imada, A. Chutinan, S. Noda, and M. Mochizuki, “Multidirectionally distributed feedback photonic crystal lasers,” Phys. Rev. B 65(19), 195306 (2002). [CrossRef]  

40. M. Yoshida, M. De Zoysa, K. Ishizaki, Y. Tanaka, M. Kawasaki, R. Hatsuda, B. Song, J. Gelleta, and S. Noda, “Double-lattice photonic-crystal resonators enabling high-brightness semiconductor lasers with symmetric narrow-divergence beams,” Nat. Mater. 18(2), 121–128 (2019). [CrossRef]  

41. R. Morita, T. Inoue, M. De Zoysa, K. Ishizaki, and S. Noda, “Photonic-crystal lasers with two-dimensionally arranged gain and loss sections for high-peak-power short-pulse operation,” Nat. Photonics 15(4), 311–318 (2021). [CrossRef]  

42. K. Hirose, Y. Liang, Y. Kurosaka, A. Watanabe, T. Sugiyama, and S. Noda, “Watt-class high-power, high-beam-quality photonic-crystal lasers,” Nat. Photonics 8(5), 406–411 (2014). [CrossRef]  

43. A. Taghizadeh and I. S. Chung, “Quasi bound states in the continuum with few unit cells of photonic crystal slab,” Appl. Phys. Lett. 111(3), 5 (2017). [CrossRef]  

44. M. V. Rybin, K. L. Koshelev, Z. F. Sadrieva, K. B. Samusev, A. A. Bogdanov, M. F. Limonov, and Y. S. Kivshar, “High-Q supercavity modes in subwavelength dielectric resonators,” Phys. Rev. Lett. 119(24), 243901 (2017). [CrossRef]  

45. K. Koshelev and Y. Kivshar, “Dielectric resonant metaphotonics,” ACS Photonics 8(1), 102–112 (2021). [CrossRef]  

46. J. Faist, F. Capasso, C. Sirtori, D. L. Sivco, A. L. Hutchinson, and A. Y. Cho, “Continuous wave operation of a vertical transition quantum cascade laser above T=80 K,” Appl. Phys. Lett. 68(14), 2024 (1996). [CrossRef]  

47. M. I. Amanti, G. Scalari, R. Terazzi, M. Fischer, M. Beck, J. Faist, A. Rudra, P. Gallo, and E. Kapon, “Bound-to-continuum terahertz quantum cascade laser with a single-quantum-well phonon extraction/injection stage,” New J. Phys. 11(12), 125022 (2009). [CrossRef]  

48. K. Wang, H. Z. Bai, C. R. Yu, H. Q. Zhu, P. P. Chen, W. Lu, L. H. Li, G. Davies, E. H. Linfield, H. Li, J. C. Cao, C. Chen, H. E. Beere, D. A. Ritchie, L. He, and G. Y. Xu, “Independent control of mode selection and power extraction in terahertz semiconductor lasers,” ACS Photonics 9(6), 1973–1983 (2022). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Measurements and simulation details

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 (6)

Fig. 1.
Fig. 1. Operation principle of a quasi-BIC microlaser array. (a) 3D schematic of the quasi-BIC array. (b) 2D cross-section of one period. The array contains N periods with the periodicity of Λ. Each period contains two microlasers with slightly different widths (w + δw and w-δw). Each microlaser is tAR in thickness, Ly in length, and Λ/2 (center-to-center) from the neighbor. The red and blue regions refer to the biased and unbiased active regions, respectively. (c) Calculated photonic band structure of the array (Λ=60.0µm, w = 12.1µm, δw = 0.3µm), where the quasi-BIC is marked as the red point. (d) Electric fields (Ez and Ex) of the quasi-BIC in one periodic unit. (e) Dependence of Qcav on the asymmetric factor (δw/w) for the quasi-BIC state in an infinite-periods array, a multi-periods (N = 22) array, and a few-periods (N = 3) array. Note that logarithmic axes are used for both horizontal and vertical axes. (f) The calculated αrad, gth, and αradtot as a function of δw/w for a multi-periods quasi-BIC array (N = 22, Λ=60.0µm, w = 12.1µm).
Fig. 2.
Fig. 2. Quasi-BIC microlaser arrays operating in pulsed mode (the pulse width is 1µs, the repeat frequency is 10 kHz) at 20 K. (a) SEM picture of a microlaser array. (b) and (c) show the measured emission spectra and L-I-V curves of 4 quasi-BIC devices with different w marked in the figure (δw = 0.1 µm, Λ=60.0 µm). Emission spectra at different injection current are given. (d) and (e) plot the measured far-field beam pattern and polarization of a quasi-BIC device (w = 12.1 µm, δw = 0.3 µm, Λ=60.0 µm). (f) Measured L-I-V curves of 5 quasi-BIC devices, where δw varies from 0 to 0.4µm with an interval of 0.1µm (w = 12.1 µm, Λ=60.0 µm). The threshold current densities of these 5 devices are 375, 395, 408, 422, 435 A cm−2, respectively. The corresponding slope efficiencies are 27, 102, 286, 545, 511 mW A−1, respectively.
Fig. 3.
Fig. 3. Comparison of the pulsed and CW operations of a quasi-BIC microlaser array, where w = 12.1 µm, δw = 0.2 µm, Λ=60.0 µm. (a) and (b) are the pulsed and CW emission spectra, respectively. (c) and (d) show respectively the pulsed and CW L-I-V curves measured at different temperatures.
Fig. 4.
Fig. 4. Concept of the hybrid-BIC microlaser array. (a) and (b) schematics of the device and its 2D cross-section. A biased few-periods quasi-BIC array is surrounded by an unbiased array (periodicity Λp, ridge width wp, number of period Np) on each side. (c) and (d) calculated electric field (Ez and Ex) of the hybrid-BIC. The parameters of the central array are N = 3, Λ=60.0 µm, w = 12.1 µm, δw = 0.1 µm. The parameters of the unbiased array are Λp = 34.3 µm, wp =17.6 µm, and Np =14. The eigenfrequency and the Qcav of the hybrid-BIC are 3.16 THz and 2400, respectively.
Fig. 5.
Fig. 5. Laser performance of a 3-periods hybrid-BIC microlaser array, where Λ=60.0 µm, w = 12.1 µm, δw = 0.1 µm, Λp = 34.3 µm, wp =17.6 µm, and Np =14. (a) SEM pictures of the device. (b) L-I-V curves measured at different temperatures in pulsed mode. (c) shows the pulsed emission spectra. (d) and (e) are L-I-V curves and emission spectra measured in CW mode, respectively. (f) Far-field beam pattern measured at 20 K.
Fig. 6.
Fig. 6. Dependence of the threshold current density on the operating temperature for the 3-periods hybrid-BIC device (a), the 22-periods quasi-BIC device (b), and the FP-MM laser (c). The black squares and the red circles correspond to the measured Jth as a function of temperature under the pulsed and CW conditions, respectively. The triangles represent the transformation from the CW to the pulsed characteristics with the fitting value of Rtherm.

Tables (1)

Tables Icon

Table 1. The measured parameters and the deduced thermal resistance of the FP-MM, the quasi-BIC, and the hybrid-BIC devices. The values of Ith,cw, Jth,cw, Pin,cw, and Pout,cw were measured at 30 K for the quasi-BIC device, and measured at 20 K for the others.

Equations (2)

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

Q r a d = Q 0 ( δ w / w ) 2
Δ T = P t h × R t h e r m
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.