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

2.1 µm multi-quantum well laser epitaxially grown on on-axis (001) InP/SiO2/Si substrate fabricated by ion-slicing

Open Access Open Access

Abstract

A cost-effective method to achieve a 2-3 µm wavelength light source on silicon represents a major challenge. In this study, we have developed a novel approach that combines an epitaxial growth and the ion-slicing technique. A 2.1 µm wavelength laser on a wafer-scale heterogeneous integrated InP/SiO2/Si (InPOI) substrate fabricated by ion-slicing technique was achieved by epitaxial growth. The performance of the lasers on the InPOI are comparable with the InP, where the threshold current density (Jth) was 1.3 kA/cm2 at 283 K when operated under continuous wave (CW) mode. The high thermal conductivity of Si resulted in improved high-temperature laser performance on the InPOI. The proposed method offers a novel means of integrating an on-chip light source.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Given the increasing communication demands of next-generation optical networks in terms of supercomputing and cloud computing, an innovative platform that integrates electronics and photonics has been identified as a key enabling technology. The use of photonic integrated circuits can facilitate compact, functional and high-performance optoelectronics, drawing on existing advanced processes, low material costs, and the inherent compatibility of Si platform. However, the generation of an effective light source on silicon represents a major challenge [1]. Significant advances have been made with all-Si laser systems, but a highly integrated all-Si laser is still limited by the excitation mode (optical pump) and low output power [2]. However, III-V compound semiconductors can serve as effective materials in laser fabrication due to the associated direct bandgap, broad wavelength tuning range, high electron mobility and photoelectric conversion efficiency. Consequently, integrating an III-V laser on Si represents a feasible approach to realizing a Si-based light source with respect to the all-Si laser scheme [311].

The 2-3 µm wavelength spectral range is critical in military, spectroscopic sensing, medical and industrial applications as this wavelength band encompasses transparent atmospheric windows and spans the absorption bands of numerous chemical molecules [12,13]. Many works have been devoted to achieve the 2-3 µm lasers on the Si [1416]. Moreover, the InP-based lasers emitting near 2.1 grown on InP have been reported [17,18]. However, most integration schemes rely on the use of wafer bonding to transfer lasers onto a Si platform. Wafer bonding requires substrate removal after transfer, resulting in substrate wastage and low integration density. The monolithic integration of a laser on a Si-based InP film represents a very promising means of achieving large-scale and effective photonic integration.

The fabrication of a high-quality Si-based InP film has been the subject of significant research activity. Heterogeneous epitaxial growth is an effective means of integrating InP film on the Si substrate [19,20]. In 2010, G. Wang et al. demonstrated the selective area growth of InP film in submicron trenches on (001) oriented Si substrates using a thin Ge buffer layer [21] . In 2020, Dmitrii V. Viazmitinov et al. reported the monolithic integration of InP on Si by molten alloy driven selective area epitaxial growth [12]. However, due to the physical mismatch between InP and Si, such as large lattice mismatch (8% mismatch), different polarity and different thermal expansion coefficients (73.1% mismatch), a high density of defects are introduced in the epitaxial InP layer, resulting in poor performance and low reliability of devices compared with those on commercial InP substrates. Alternatively, ion-slicing technique as a novel integration route can achieve a high-quality integration of InP film on the Si substrate regardless of the physical mismatch as our earlier reports [2224]. Moreover, the residual InP substrate can be recycled for other transferring after being polished by chemical mechanical polishing (CMP), which provides a cost-effective monolithic integration route.

In this work, we report the monolithic integration of a 2.1 µm InAs/In0.53Ga0.47As multi-quantum well (MQW) laser on a wafer-scale heterogeneous integrated InP/SiO2/Si (InPOI) substrate fabricated by ion-slicing technique. Under pulsed mode, the laser on the InPOI substrate can sustain lasing up to 293 K with a threshold current density (Jth) of 1.06 kA/cm2. Under CW mode, laser operation was up to 283 K with an associated Jth of 1.3 kA/cm2. The enhanced high-temperature performance of the laser on the InPOI can be attributed to the high thermal conductivity of Si. The results confirm that a combination of epitaxial growth and ion-slicing technology offers a feasible route to achieve the integration of mid-infrared light source on a Si substrate.

2. Device structure and fabrication

The InPOI substrate was fabricated by ion-slicing as outlined in an earlier report [17]. A sequential co-implantation of He and H ions was initially carried out on a 2-inch bulk InP wafer at room temperature, and the ion energy/fluence of the He and H ion implantation were 115 keV/2 × 1016 cm−2 and 75 keV/5 × 1016 cm−2, respectively. The implanted InP wafer was bonded to the Si/SiO2 substrate at room temperature through hydrophilic wafer bonding. A thin monocrystalline InP layer was transferred to the Si/SiO2 substrate after annealing at 400°C for 0.5 hours in N2 atmosphere. Chemical mechanical polishing (CMP) was finally employed to remove the damaged layer introduced by ion implantation and to smooth the surface for epitaxial growth. However, there are many particles appearing on the surface due to the poor post-wash. These particles will cause the failure of the lasers so that the yield of the laser is low. If the surface treating can be improved to remove the residual particles after CMP, the yield will be greatly increased. The prepared 2-inch InPOI substrate is shown in Fig. 1(a), where the thickness of the InP layer is about 473 nm.

 figure: Fig. 1.

Fig. 1. (a) Image of the as-prepared InPOI substrate; (b) schematic representation of the lasers on InP and InPOI; (c) 3D diagram of lasers on InP and InPOI; optical microscopy images of as-fabricated laser bars on (d) InP and (e) InPOI; (f) cross-section SEM image of an as-cleaved laser on InPOI.

Download Full Size | PDF

An InAs/In0.53Ga0.47As MQW laser structure was then grown on the as-prepared InPOI substrate using the VG Semicon V80 H gas source molecular beam epitaxy (GSMBE) system. As a counterpart, the same structure was grown on a commercial bulk InP wafer using the same growth conditions. It should be noted that the epitaxial growth process standardized for the bulk InP substrate was not optimized for InPOI, when considering the differences in thermal conductivity and thermal expansivity between Si and InP. The detailed structure on both substrates is illustrated in Fig. 1(b). The thickness of the InP buffer layer as well as the cladding layer on InP and InPOI substrates is 1 µm and 2 µm, respectively. The thicker N+ InP buffer layer was introduced on InPOI to reduce the series resistance caused by the heterogeneous substrate. The active region consists of three periods of 8 nm thick InAs/In0.53Ga0.47As triangular quantum wells separated by 20 nm thick n-In0.53Ga0.47As barrier layers. In each triangular QW region, the indium composition was increased from 0.53 to 1 and then decreased to 0.53 by using InAs/In0.53Ga0.47As digital alloy growth technology to form the triangular QW shape. The thickness d of a short period including one InAs layer and one In0.53Ga0.47As layer is designed to be 1 nm and 8 periods are included in each well layer. It is sandwiched between two 100 nm thick n-In0.53Ga0.47As barrier layers, embedded in two 120 nm thick n-InGaAsP waveguide layers. The top cladding and top contact layers are 1.7 µm thick InP and 300 nm thick P+ In0.53Ga0.47As, respectively.

Both epitaxial structures were processed into ridge lasers using standard photolithography and wet etching. The ridge widths are 12 µm, 14 µm, 16 µm and 18 µm, respectively. Applying the same process, the p-contacts were taken on the top ridge and the n-contact were taken on the top of InP buffer layers, i.e. the “top-top” contact scheme. The top ridges were formed by selective wet etching and the etch stops at the interface between the upper 1.7 µm thick InP cladding layer and 120 nm thick InGaAsP waveguide layer to avoid damage to the MQWs. Exposure of the n-contact layer required etching through the entire structure as far as the InP buffer layer. Sidewall passivation with a 300 nm thick Si3N4 layer was deposited using plasma enhanced chemical vapor deposition (PECVD). Reactive ion etching (RIE) was used to expose both contact windows. The Ti/Pt/Au metal layers were first evaporated on the top p-contact in a lift-off process, followed by Ge/Au/Ni/Au metal layers deposition on the n-contact layer. Following rapid thermal annealing and substrates thinning, the laser bars were cleaved into lasers with a cavity length of 1 mm. The 3D device structure is represented schematically in Fig. 1(c). The surface morphology of the devices fabricated on the InP and InPOI substrates prior to cleaving are illustrated in Figs. 1(d) and 1(e), respectively. No obvious surface defects were observed. The cross-section scanning electron microscopy (SEM) images of an as-prepared FP laser on InPOI is presented in Fig. 1(f). The cavity surface of the laser on the InPOI is sufficiently smooth to facilitate the lasing of light.

3. Results and discussions

3.1 Material characterization

The crystallinity and surface roughness of the epitaxial layers have a major impact on the performance of Si-based lasers. The surface morphologies (5 µm × 5 µm area) of the two laser structures were evaluated by atomic force microscope (AFM), as shown in Figs. 2(a) and 2(b). The root mean square (RMS) roughness value for the laser structure on the InPOI substrate (2.7 nm) is larger than that for the InP substrate (0.138 nm). The highest point on the epitaxial surface of the InPOI substrate reaches 15 nm due to the residual particles following CMP. It is expected that epitaxial layers with lower RMS roughness values can be achieved under well optimized CMP conditions for the InPOI heterogeneous substrate.

 figure: Fig. 2.

Fig. 2. AFM images of the laser structures on (a) InP and (b) InPOI for a 5 × 5 µm2 scan area; (c) HRXRD (004) ω/2θ profiles for the laser structures on InP and InPOI; (d) RT-PL spectra of the MQWs on InP and InPOI.

Download Full Size | PDF

The crystallinity of the epitaxial layers on both InP and InPOI substrates were characterized by high-resolution X-ray diffraction (HRXRD). Figure 2(c) shows the (004) ω/2θ profiles. Both profiles show obvious satellite peaks where this feature is less pronounced in the case of InPOI, indicating a slightly inferior quality of the associated epitaxial layers. The full width at half maximum (FWHM) of the InP layer on the InPOI substrate is 104 arcsec, which is 4.5 times greater than that of InP substrate (23 arcsec), and can be attributed to residual defect propagation and evolution during epitaxial growth.

In order to further evaluate the optical quality of the quantum well structure, photoluminescence (PL) measurements were conducted at room temperature. Before the PL measurement, the 300 nm thick In0.53Ga0.47As contact layer and 1.7 µm thick InP cladding layer were removed by selective wet etching. As shown in Fig. 2(d), the MQW structure on InP is characterized by two peaks in the PL spectrum. The peak around 1.7 µm represents an In0.53Ga0.47A emission peak, and the peak around 2.1 µm is the MQW emission peak. In the case of the PL spectrum associated with MQW on InPOI, four emission peaks (2.14 µm, 1.94 µm, 1.75 µm, 1.61 µm) are evident, where the highest intensity is four times greater than that of InP substrate. The enhanced PL intensity is mainly due to the strong reflections from the Si and oxide layer coupled with the reflection from the top III/V surface to form a resonance cavity that increases PL pump efficiency in the InPOI sample [25]. The additional peaks (for the emission wavelength of 1.94 µm, 1.75 µm) as compared to the reference structure on InP and the MQW emission peak (for the emission wavelength of 2.14 µm) show the same wavelength spacing (∼0.2 µm). Hence, it is reasonable to think that the additional peaks also caused by the resonance effect in the formed resonance cavity. The principal peak for the epitaxial layer on InPOI is located at 2140 nm compared with 2099nm for the InP substrate. The residual thermal strain in the MQWs associated with InPOI is the likely cause of the difference in PL wavelength. In addition, a slight difference in the growth temperature on the top surface of InPOI relative to InP due to the different thermal conductivities may cause a composition change, resulting in the PL shifts.

The microstructure of the epitaxial layers grown on InP and InPOI were characterized using transmission electron microscopy (TEM) and scanning TEM (STEM). Cross-sectional TEM images of the overall epitaxial structures on InP and InPOI substrate are shown in Figs. 3(a) and (b), respectively. In both cases, the multilayer interfaces can be clearly distinguished without obvious threading dislocations in the epitaxial layers. The active regions, as the most crucial components, were examined by STEM, as shown in Figs. 3(c)-3(f). The interfaces between the InAs/In0.53Ga0.47As QW layer and In0.53Ga0.47As barrier layers are well distinguished by the different contrast of atoms, and the atoms of the active regions are arranged in a regular lattice structure without any visible dislocation on both the InP and InPOI substrates. The STEM image of the interface between InPOI substrate and epitaxial layer, shown in Fig. 3(g), reveals obvious void defects in the transferred InP film, which were introduced during ion implantation. However, the void defects did not propagate or evolve significantly during epitaxial growth.

 figure: Fig. 3.

Fig. 3. Cross-sectional TEM images of laser structures on (a) InP and (b) InPOI, where the dashed lines represent the interfaces; STEM images of the entire MQW structures on (c) InP and (d) InPOI, and single QW structures on (e) InP and (f) InPOI; (g) STEM image of the interface between the InPOI substrate and epitaxial layers.

Download Full Size | PDF

3.2 Laser performance

The lasing properties of lasers with different ridge widths on InP and InPOI substrates under pulsed mode (duty cycle of 5%, 2 kHz) at 293 K are presented in Figs. 4(a) and 4(b), respectively. The variation of threshold current (Ith) and threshold current density (Jth) as a function of ridge width is shown in Fig. 4(c). As the ridge width was increased, the associated Ith exhibited a near-liner increase due to the increased current injection area, whereas Jth remains essentially constant. In the case of 12 µm ×1 mm lasers on InP and InPOI, Jth is about 0.79 kA/cm2 and 1.06 kA/cm2, respectively. The lasing spectra of 14 µm × 1 mm lasers on both substrates under pulsed mode at 293 K are presented in Fig. 4(d). The peak wavelengths of lasers on InP substrate and InPOI substrate are located at 2.125 µm and 2.1 µm, respectively. The difference in lasing wavelengths may be attributed to the residual thermal strain in the MQWs of InPOI [25] and the composition changes caused by the variation in growth temperature as a result of the different thermal conductivity of Si and InP.

 figure: Fig. 4.

Fig. 4. Lasing properties of lasers with different ridge widths on (a) InP and (b) InPOI operated under pulsed mode at 293 K; (c) linear relationship of threshold current and threshold current density as a function of ridge width; (d) lasing spectra for 14 µm×1 mm lasers on both substrates.

Download Full Size | PDF

In order to evaluate the performance of the lasers operated under CW mode, the temperature-dependent light output versus current (L-I) characteristics were measured for the lasers on InP and InPOI substrates with an identical ridge width and cavity length (14 µm and 1000 µm). A thermoelectric cooler (TEC) system was used to control the stage temperature. The lasers on both substrates under CW mode exhibited a comparable temperature performance, as shown in Figs. 5(a) and 5(b). The lasers on InP and InPOI can operate up to 313 K and 283 K under CW mode, respectively. At 283 K, the Jth associated with the laser on InPOI substrate is 1.3 kA/cm2, which is 1.7 times larger than that of the laser on InP (0.76 kA/cm2). The single-facet output power for the lasers on InP and InPOI are 11.5 mW and 3.1 mW, respectively. Pulsed operation was performed to avoid a self-heating effect. Under pulsed mode, the operating temperatures of both lasers are increased relative to CW by about 20 K, as shown in Figs. 5(c) and 5(d). At 283 K, the Jth of the lasers on InP and InPOI are 0.65 kA/cm2 and 1.03 kA/cm2, respectively. When compared with CW mode, pulsed operation resulted in a significant reduction in threshold current density on both substrates, indicating that device self-heating dominates thermal stability.

 figure: Fig. 5.

Fig. 5. I-P characteristics of 14 µm ×1 mm LDs on (a) InP and (b) InPOI under CW mode at different temperatures; I-P characteristics of 14 µm ×1 mm LDs on (c) InP and (d) InPOI under pulsed mode at different temperatures; temperature dependence of threshold current density for LD on (e) InP and (f) InPOI.

Download Full Size | PDF

The average characteristic temperature T0, which is an important measure of the temperature sensitivity of semiconductor laser, can be calculated using the following equation:

$$\frac{{{J_{th}}({{T_1}} )}}{{{J_{th}}({{T_2}} )}} = \textrm{exp}\left( {\frac{{{T_1} - {T_2}}}{{{T_0}}}} \right)$$
where Jth(T1) and Jth(T2) represent the threshold current density at operating temperature T1 and T2, respectively. The threshold current densities at various temperature are shown in Figs. 5(e) and 5(f). The characteristic temperature T0 was obtained in different temperature intervals. For the laser on InP under CW mode, the calculated T0 was 98.5 K between 150 K and 225 K, 59.2 K between 225 K and 293 K, and 30.7 K between 293 K and 313 K. Under pulsed mode, the T0 value was lower (96.8 K for 150-225 K and 50 K for 225-293 K) than CW mode in the temperature range 150 K to 293 K. However, the reverse applies between 293 K and 313 K with a higher T0 in pulsed mode. In the case of lasers on InPOI, the T0 was lower under pulsed mode relative to CW mode in the 150-250 K (65.8 K/52.9 K) and 250-203 K (40.3 K/39.6 K) intervals, and lower than the corresponding T0 for the InP substrate. As the temperature was increased, the laser T0 value for both substrates gradually converged as a result of the higher thermal conductivity of Si relative to InP that facilitates heat dissipation at higher temperature. The incorporation of Si serves to enhance the temperature characteristics of the lasers on the InPOI substrate.

The current-voltage (I-V) characteristics at various temperatures for the 14 µm×1 mm ridge lasers fabricated on InP and InPOI are shown in Fig. 6. The lasers on both InP and InPOI substrates have nearly equal turn voltage of about 0.7 V, while the extracted series resistance for InPOI (9.2 Ω) at 293 K is slightly higher than that of bulk InP (7.2 Ω). When compared with the thick InP layer (greater than 100 µm) remaining after lapping in the case of lasers on InP, the n-InP contact layer for lasers on the InPOI substrate is significantly thinner (1 µm), which contributes to extra lateral resistance in the total series resistance.

 figure: Fig. 6.

Fig. 6. The current-voltage (I-V) characteristics at various temperatures for the 14 µm×1 mm ridge lasers fabricated on InP and InPOI.

Download Full Size | PDF

The lasing spectra for lasers on InP and InPOI operated at 1.2 times the threshold current at various temperatures under CW mode are shown in Figs. 7(a) and 7(b). A clear red-shift to longer emission wavelengths is evident for the lasers on both substrates. As the operating temperature was increased, the band gap of the active area narrowed. Besides, the self-heating effect caused by the increasing injection currents can account for the wavelength red-shift. The average red-shift rates (Δλ/ΔT) of the lasers on both substrates under CW mode are shown in Fig. 7(c). The laser on InPOI exhibited a lower red-shift rate (0.89 nm/K) than the laser on the InP (0.91 nm/K). During operation, the high thermal conductivity of Si reduces the influence of thermal effects in the active area on the red-shift caused by self-heating. In order to further assess the impact of self-heating, the lasing spectra at different temperatures for both substrates at 1.2 times the threshold current under pulsed mode were measured, as shown Figs. 7(d) and 7(e). The increase in temperature was accompanied by the same wavelength red-shift on both substrates as the laser under the CW mode. As shown in Fig. 7(f), a lower average red-shift rate (0.87 nm/K) was recorded under pulsed mode for lasers on InP and InPOI. The equivalent red-shift rates under pulsed mode suggests that the self-heating effect can be ignored and the quality of the transferred InP layer is comparable to the InP substrate. In general, the laser on InPOI exhibited better thermostability.

 figure: Fig. 7.

Fig. 7. Lasing spectra for 14 µm×1 mm lasers on (a) InP and (b) InPOI operated under CW mode at different temperatures; (c) temperature dependence of the lasing wavelength for lasers on InP and InPOI under CW mode; lasing spectra for 14 µm×1 mm lasers on (d) InP and (e) InPOI under pulsed mode at different temperatures; (f) temperature dependence of the lasing wavelength for lasers on InP and InPOI under pulsed mode.

Download Full Size | PDF

4. Conclusions

Applying epitaxial growth, we have achieved CW and pulsed lasing of electrically-driven InAs/In0.53Ga0.47As MQW lasers emitting at 2.1 µm on the InPOI substrate fabricated by ion-slicing. Although some void defects are present in the transferred InP layer, the epitaxial layers on InPOI exhibit high crystallinity, suggesting that the defects do not propagate and evolve during epitaxial growth. The lasers on InPOI deliver a comparable performance to the lasers on InP. The operating temperature of the laser on InPOI extended to 283 K under CW mode with an associated Jth of 1.3 kA/cm2. Moreover, the lasers on InPOI exhibited better high-temperature performance with a higher average characteristic temperature T0 under both CW and pulsed modes. As the temperature was increased, the lower red-shift rate for the laser on InPOI under CW mode can be attributed to the higher thermal conductivity of Si than InP. The combination of ion-slicing and epitaxial growth offers a new route to cost-effective mid-infrared photonic integration on a Si substrate.

Funding

National Natural Science Foundation of China (12205119, 62174167, 62293521); Shanghai Rising-Star Program (22QA1410700); China Postdoctoral Science Foundation (2022M723282); Natural Science Foundation of Zhejiang Province (LQ23F040002); Jiashan County Scientific and Technological projects (2022A03); Jiaxing Municipal Public Welfare Research Project (2022AY10027).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (62293521, 62174167, 12205119), Shanghai Rising-Star Program (22QA1410700), China Postdoctoral Science Foundation (2022M723282), Zhejiang Provincial Natural Science Foundation of China (LQ23F040002), Jiaxing Municipal Public Welfare Research Project (2022AY10027), Jiashan County Scientific and Technological projects (2022A03).

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. S. Pan, V. Cao, M. Liao, et al., “Recent progress in epitaxial growth of III-V quantum-dot lasers on silicon substrate,” J. Semicond. 40(10), 101302 (2019). [CrossRef]  

2. D. C. Wang, C. Zhang, P. Zeng, et al., “An all-silicon laser based on silicon nanocrystals with high optical gains,” Sci. Bull. 63(2), 75–77 (2018). [CrossRef]  

3. S. Chen, M. Liao, M. Tang, et al., “Electrically pumped continuous-wave 13 µm InAs/GaAs quantum dot lasers monolithically grown on on-axis Si (001) substrates,” Opt. Express 25(5), 4632 (2017). [CrossRef]  

4. P. Mrowiński, P. Holewa, A. Sakanas, et al., “Optimization of heterogeneously integrated InP-Si on-chip photonic components,” Opt. Express 31(2), 1541 (2023). [CrossRef]  

5. J. Kwoen, B. Jang, J. Lee, et al., “All MBE grown InAs/GaAs quantum dot lasers on on-axis Si (001),” Opt. Express 26(9), 11568 (2018). [CrossRef]  

6. J. Kwoen, B. Jang, K. Watanabe, et al., “High-temperature continuous-wave operation of directly grown InAs/GaAs quantum dot lasers on on-axis Si (001),” Opt. Express 27(3), 2681 (2019). [CrossRef]  

7. W. Wei, Q. Feng, J. Guo, et al., “InAs/GaAs quantum dot narrow ridge lasers epitaxially grown on SOI substrates for silicon photonic integration,” Opt. Express 28(18), 26555 (2020). [CrossRef]  

8. Z. Lv, S. Wang, S. Wang, et al., “Ultra-high thermal stability InAs/GaAs quantum dot lasers grown on on-axis Si (001) with a record-high continuous-wave operating temperature of 150 °C,” Opt. Express 31(15), 24173 (2023). [CrossRef]  

9. H. Liang, T. Jin, C. Chi, et al., “InAs/GaAs quantum dot laser epitaxially grown on on-axis (001) GaAsOI substrate,” Opt. Express 29(23), 38465 (2021). [CrossRef]  

10. T. Aihara, T. Hiraki, K. Takeda, et al., “Membrane buried-heterostructure DFB laser with an optically coupled III-V/Si waveguide,” Opt. Express 27(25), 36438 (2019). [CrossRef]  

11. C. Besancon, D. Néel, D. Make, et al., “Algainas multi-quantum well lasers on silicon-on-insulator photonic integrated circuits based on inp-seed-bonding and epitaxial regrowth,” Appl. Sci. 12(1), 1 (2022). [CrossRef]  

12. D. V. Viazmitinov, Y. Berdnikov, S. Kadkhodazadeh, et al., “Monolithic integration of InP on Si by molten alloy driven selective area epitaxial growth,” Nanoscale 12(46), 23780–23788 (2020). [CrossRef]  

13. A. Z. Subramanian, E. Ryckeboer, A. Dhakal, et al., “Silicon and silicon nitride photonic circuits for spectroscopic sensing on-a-chip [Invited],” Photonics Res. 3(5), B47 (2015). [CrossRef]  

14. N. Hattasan, A. Gassenq, L. Cerutti, et al., “Integrated thin-film GaSb-based Fabry-Perot lasers: towards a fully integrated spectrometer on a SOI waveguide circuit,” Quantum Sens. Nanophotonic Devices X 8631, 863114 (2013). [CrossRef]  

15. M. Rio Calvo, L. Monge Bartolomé, M. Bahriz, et al., “Mid-infrared laser diodes epitaxially grown on on-axis (001) silicon,” Optica 7(4), 263 (2020). [CrossRef]  

16. A. Spott, M. Davenport, J. Peters, et al., “Heterogeneously integrated 20 µm CW hybrid silicon lasers at room temperature,” Opt. Lett. 40(7), 1480 (2015). [CrossRef]  

17. M. Mitsuhara, M. Ogasawara, M. Oishi, et al., “2.05-µm wavelength InGaAs-InGaAs distributed-feedback multiquantum-well lasers with 10-mW output power,” IEEE Photonics Technol. Lett. 11(1), 33–35 (1999). [CrossRef]  

18. Y. Ueda, T. Shindo, T. Kanai, et al., “2-µm band active distributed Bragg reflector laser for CO2 gas sensing,” Appl. Phys. Express 12(9), 092011 (2019). [CrossRef]  

19. B. Shi and J. Klamkin, “Defect engineering for high quality InP epitaxially grown on on-axis (001) Si,” J. Appl. Phys. 127(3), 1 (2020). [CrossRef]  

20. J. Li, Y. Xue, L. Lin, et al., “Telecom InGaAs/InP Quantum Well Lasers Laterally Grown on Silicon-on-Insulator,” J. Lighteave Technol. 40(16), 5631–5635 (2022). [CrossRef]  

21. G. Wang, M. R. Leys, R. Loo, et al., “Selective area growth of high quality InP on Si (001) substrates,” Appl. Phys. Lett. 97(12), 19–22 (2010). [CrossRef]  

22. J. Lin, T. You, M. Wang, et al., “Efficient ion-slicing of InP thin film for Si-based hetero-integration,” Nanotechnology 29(50), 504002 (2018). [CrossRef]  

23. J. Lin, T. You, T. Jin, et al., “Wafer-scale heterogeneous integration InP on trenched Si with a bubble-free interface,” APL Mater. 8(5), 051110 (2020). [CrossRef]  

24. T. Jin, J. Lin, T. You, et al., “Efficient heterogeneous integration of InP/Si and GaSb/Si templates with ultra-smooth surfaces,” Sci. China Inf. Sci. 65(8), 1–9 (2022). [CrossRef]  

25. Y. Hu, D. Liang, K. Mukherjee, et al., “III/V-on-Si MQW lasers by using a novel photonic integration method of regrowth on a bonding template,” Light: Sci. Appl. 8(1), 93 (2019). [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 (7)

Fig. 1.
Fig. 1. (a) Image of the as-prepared InPOI substrate; (b) schematic representation of the lasers on InP and InPOI; (c) 3D diagram of lasers on InP and InPOI; optical microscopy images of as-fabricated laser bars on (d) InP and (e) InPOI; (f) cross-section SEM image of an as-cleaved laser on InPOI.
Fig. 2.
Fig. 2. AFM images of the laser structures on (a) InP and (b) InPOI for a 5 × 5 µm2 scan area; (c) HRXRD (004) ω/2θ profiles for the laser structures on InP and InPOI; (d) RT-PL spectra of the MQWs on InP and InPOI.
Fig. 3.
Fig. 3. Cross-sectional TEM images of laser structures on (a) InP and (b) InPOI, where the dashed lines represent the interfaces; STEM images of the entire MQW structures on (c) InP and (d) InPOI, and single QW structures on (e) InP and (f) InPOI; (g) STEM image of the interface between the InPOI substrate and epitaxial layers.
Fig. 4.
Fig. 4. Lasing properties of lasers with different ridge widths on (a) InP and (b) InPOI operated under pulsed mode at 293 K; (c) linear relationship of threshold current and threshold current density as a function of ridge width; (d) lasing spectra for 14 µm×1 mm lasers on both substrates.
Fig. 5.
Fig. 5. I-P characteristics of 14 µm ×1 mm LDs on (a) InP and (b) InPOI under CW mode at different temperatures; I-P characteristics of 14 µm ×1 mm LDs on (c) InP and (d) InPOI under pulsed mode at different temperatures; temperature dependence of threshold current density for LD on (e) InP and (f) InPOI.
Fig. 6.
Fig. 6. The current-voltage (I-V) characteristics at various temperatures for the 14 µm×1 mm ridge lasers fabricated on InP and InPOI.
Fig. 7.
Fig. 7. Lasing spectra for 14 µm×1 mm lasers on (a) InP and (b) InPOI operated under CW mode at different temperatures; (c) temperature dependence of the lasing wavelength for lasers on InP and InPOI under CW mode; lasing spectra for 14 µm×1 mm lasers on (d) InP and (e) InPOI under pulsed mode at different temperatures; (f) temperature dependence of the lasing wavelength for lasers on InP and InPOI under pulsed mode.

Equations (1)

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

J t h ( T 1 ) J t h ( T 2 ) = exp ( T 1 T 2 T 0 )
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.