• Photonics Research
  • Vol. 10, Issue 5, 1271 (2022)
Zhuohui Yang1, Zhengqing Ding1, Lin Liu1, Hancheng Zhong1, Sheng Cao1, Xinzhong Zhang1, Shizhe Lin1, Xiaoying Huang1, Huadi Deng1, Ying Yu1、*, and Siyuan Yu1、2
Author Affiliations
  • 1State Key Laboratory of Optoelectronic Materials and Technologies, School of Electronics and Information Technology, Sun Yat-sen University, Guangzhou 510275, China
  • 2e-mail: yusy@mail.sysu.edu.cn
  • show less
    DOI: 10.1364/PRJ.454200 Cite this Article Set citation alerts
    Zhuohui Yang, Zhengqing Ding, Lin Liu, Hancheng Zhong, Sheng Cao, Xinzhong Zhang, Shizhe Lin, Xiaoying Huang, Huadi Deng, Ying Yu, Siyuan Yu. High-performance distributed feedback quantum dot lasers with laterally coupled dielectric gratings[J]. Photonics Research, 2022, 10(5): 1271 Copy Citation Text show less

    Abstract

    The combination of grating-based frequency-selective optical feedback mechanisms, such as distributed feedback (DFB) or distributed Bragg reflector (DBR) structures, with quantum dot (QD) gain materials is a main approach towards ultrahigh-performance semiconductor lasers for many key novel applications, as either stand-alone sources or on-chip sources in photonic integrated circuits. However, the fabrication of conventional buried Bragg grating structures on GaAs, GaAs/Si, GaSb, and other material platforms has been met with major material regrowth difficulties. We report a novel and universal approach of introducing laterally coupled dielectric Bragg gratings to semiconductor lasers that allows highly controllable, reliable, and strong coupling between the grating and the optical mode. We implement such a grating structure in a low-loss amorphous silicon material alongside GaAs lasers with InAs/GaAs QD gain layers. The resulting DFB laser arrays emit at pre-designed 0.8 THz local area network wavelength division multiplexing frequency intervals in the 1300 nm band with record performance parameters, including sidemode suppression ratios as high as 52.7 dB, continuous-wave output power of 26.6 mW (room temperature) and 6 mW (at 55°C), and ultralow relative intensity noise (RIN) of <-165 dB/Hz (2.5–20 GHz). The devices are also capable of isolator-free operating under very high external reflection levels of up to -12.3 dB while maintaining high spectral purity and ultralow RIN qualities. These results validate the novel laterally coupled dielectric grating as a technologically superior and potentially cost-effective approach for fabricating DFB and DBR lasers free of their semiconductor material constraints, which are thus universally applicable across different material platforms and wavelength bands.

    1. INTRODUCTION

    Embedding semiconductor lasers with Bragg gratings as wavelength-selective feedback mechanisms is a well-established approach to achieving high-quality single-frequency lasing. In conjunction with the distinctive properties of various compound semiconductor gain materials, distributed feedback (DFB) and distributed Bragg reflector (DBR) lasers are finding a wide range of applications in both classical and quantum domains, such as with InGaAs emitting in the near infrared for optical communication [1], with GaAs or GaAsP in the red spectral range [2] for atomic clocks [3], atom interferometry [4], and efficient optical pumping [5], with GaSb or InAs/AlSb in the middle to far infrared [6] for trace-gas sensing [7], and with nitride semiconductors in the green to ultraviolet for absorption spectroscopy [8] and high-density data storage [9].

    For optical data interconnect applications, the 1310 nm band is of particular interest for low-cost wavelength division multiplexing (WDM) systems [10], 5G and 6G optical networks [11], as well as for LiDAR [12,13] and sensing [14,15]. Compared with conventional InGaAs/InGaAlAs quantum well (QW) materials emitting in the same wavelength range, self-assembled InAs/GaAs quantum dots (QDs) have achieved superior performance, including lower threshold currents [16] and higher-temperature stability [17] due to the zero-dimensional (0D) carrier confinement in QDs. A highly desirable feature of QD lasers is their ultralow relative intensity noise (RIN) [1820] originating from their very low linewidth enhancement factor α, which affords great advantages in analog transmission [such as radio over fiber (RoF)] and sensing [21]. The low α also underpins their tolerance to high levels of external optical feedback [2224], making QD lasers very promising light sources for reliable, scalable, and isolator-free photonic integrated circuits (PICs) [25]. Furthermore, InAs/GaAs QD materials have also been proved to be highly tolerant to epitaxial defects [2628], yielding high-performance Fabry–Perot (FP)-type laser diodes epi-grown on silicon substrate [2830]. Similar advantages have also been observed in other QD laser systems such as InAs/InP [31,32].

    For DFB lasers, the buried Bragg gratings, as first realized in InP-based 1550 nm lasers, are conventionally placed on top of the active layer and fabricated through a regrowth process after grating definition and etching. The proximity to the waveguide allows the grating to intercept the optical field with a high coupling coefficient κ. This conventional approach has also been attempted in other materials and wavelengths such as GaAs-based QD lasers. In 2011, Tanaka et al. demonstrated a 1293 nm InAs/GaAs QD DFB laser with a GaAs grating buried in a metal organic vapor phase epitaxy (MOVPE) regrown InGaP cladding, achieving a κ of 40  cm1 and sidemode suppression ratio (SMSR) of 45 dB [33]. In 2020, Wan et al. demonstrated a 1310 nm QD DFB laser epitaxially on silicon by molecular beam epitaxy (MBE) regrowth. Using a GaAs grating buried in an Al0.4Ga0.6As upper cladding, a κ of 45  cm1 and an SMSR of >50  dB were achieved [34].

    However, the regrowth process represents poor productivity for many material systems. For GaAs-based laser devices, an InGaP upper cladding layer requires wafer transfer from MBE to MOCVD systems, while an AlGaAs upper cladding layer requires rigorous pretreatment before regrowth via an ultrahigh-vacuum MBE chamber. Harder still, the regrowth process for GaN(Sb)-based lasers suffers from lack of a contamination-free AlGaN(Sb) regrowth process or other latticed-matched cladding materials with sufficient bandgap and refractive index contrasts.

    An alternative, regrowth-free approach uses Bragg gratings etched alongside a ridge waveguide to form a laterally coupled DFB (LC-DFB) laser. For InP-based laser structures, the gratings can be fabricated simultaneously during the waveguide etching process. Using an aluminum-containing stop-etch layer and a chemically selective recipe [35,36], the grating penetration depth can be precisely delimited to just above the active layer to achieve a precise κ value. However, translating this approach to GaAs- or GaSb-based DFB laser structures has proven to be challenging due to the lack of a suitable selective etch-stop layer. Previously demonstrated GaAs [37], GaSb [38], or GaN [39] ridge waveguides with LC gratings fabricated using such a one-step reactive ion etching (RIE) approach suffer from the fact that, due to local chemical transportation and reaction rate variations caused by the etched ridge and the very narrow grating gaps, it is very difficult to control the etch depth at the foot of the ridge waveguide where the grating intercepts the optical mode. An undesirable feature known as “footing,” which is a gradual increase in etch depth away from the foot of the etched waveguide, and another feature known as “RIE-lag,” which is a decreased etch depth in narrow gaps, result in significant uncertainties in the grating κ value.

    To circumvent this problem, in a previous work, the authors chose to etch the lateral grating deeply through the active region so that the grating etch depth no longer affects κ [40]. However, a deep-etched active waveguide suffers from increased surface recombination and optical scattering loss, and a waveguide with practical widths can support more than one transverse mode, with the unwanted high-order modes as favored lasing modes due to their higher κ values [40]. For GaSb lasers, metal gratings deposited after waveguide etching [41] were also used. While providing strong optical coupling, metal gratings can introduce significant additional absorption loss in the laser cavity.

    In this paper, we demonstrate a novel dielectric grating structure placed alongside single transverse mode ridge waveguides that have a precisely controlled trapezoid cross-sectional profile etched to a depth just above the active layer. Fabricated in an amorphous silicon (α-Si) layer deposited after the formation of the ridge waveguide, the grating corrugations, plasma-etched into the α-Si, are precisely stopped at an underlying etch stop layer of Al2O3 deposited after the waveguide etching and before the α-Si layer. A high-contrast grating (with a refractive index difference of Δn2) is formed between the α-Si corrugations and a subsequently deposited silicon dioxide (SiO2) cladding layer, producing an LC grating with significantly enhanced and precisely controllable coupling coefficient κ.

    We implemented the novel structure on an InAs/GaAs QD gain material, producing LC-DFB laser arrays emitting across the 1300 nm band on a 0.8 THz local area network wavelength division multiplexing (LWDM) grid. The devices emit more than 26.6 mW of single-mode output power at room temperature and a typical SMSR greater than 52.7 dB. They also demonstrate ultralow RIN of <165  dB/Hz in the range of 2.5–20 GHz and isolator-free operation under external feedback levels of up to 12.3  dB (5.9%). The output power, SMSR, and RIN values are the best of the reported values of InAs/GaAs QD LC-DFB lasers, as far as we are aware of. These superior performances of the devices validate the novel LC grating as an effective regrowth-free approach to grating-based laser fabrication. In addition to the elimination of regrowth, the deployment of the novel LC grating structure decouples its fabrication from specific laser materials, and therefore the scheme could serve as a universal alternative approach for high-performance semiconductor laser devices employing grating structures as optical feedback mechanisms.

    2. DESIGN AND FABRICATION OF THE LC-DFB QD LASER

    The InAs/GaAs QD laser structure, in a typical p-i-n configuration [Fig. 1(a)], was grown on 3-inch (1 inch = 2.54 cm) semi-insulating GaAs (001) substrates in a solid-source MBE chamber. First, a 500 nm Si-doped GaAs buffer layer and a 1.8 μm Al0.4Ga0.6As cladding layer were grown, followed by an active region that contains a five-layer InAs/GaAs QD separated by 35 nm GaAs barriers. Each QD layer comprises 2.4 ML InAs covered with a 3.5 nm In0.15Ga0.85As strain-reducing layer. Modulation p-doping with Be was implemented in a 6 nm GaAs layer located 10 nm beneath each QD layer to obtain a concentration of 20 acceptors per dot. Afterwards, 1.8 μm Be-doped Al0.4Ga0.6As and 100 nm GaAs layers were grown as p-cladding and p-contact layers, respectively. The InAs/GaAs QD gain materials with a density of 5.5×1010  cm2 were achieved, as indicated by the 1  μm×1  μm atomic force microscope (AFM) inset of Fig. 1(b) from an uncapped QD sample grown on a GaAs substrate under the same QD growth conditions. Room-temperature photoluminescence (PL) emission peaking at 1308 nm was observed with a narrow full-width at half-maximum (FWHM) of 30.9 meV. Additionally, large quantized-energy separation of 80 meV between the ground-state and the first exited state effectively suppressed carrier overflow at high temperatures. Traditional multi-mode ridge waveguide FP lasers (20 μm ridge width) with cleaved facets were fabricated to characterize the properties of the QD materials. Light–current–voltage (LIV) characteristics of the fabricated laser with a length of 2000 μm and its temperature dependence under a continuous-wave (CW) condition are shown in Fig. 1(c). The threshold currents are as low as 80 mA (200  A/cm2) at 25°C. Reduced temperature sensitivity is achieved with the characteristic temperature T0 as high as 98 K in the range of 25°C–55°C and 53 K in the range of 65°C–115°C. Since the measurements were carried out under CW mode, the extracted T0 should be an underestimate of the true value due to junction heating.

    Material properties of InAs/GaAs QD lasers. (a) Cross-sectional scanning electron microscope (SEM) image of layer stack of the epi-wafer. The inset is the transmission electron microscope (TEM) image of the five QD layers. (b) Photoluminescence spectrum of the QD active layers on GaAs. The inset shows the atomic force microscope (AFM) image of an uncapped QD layer. (c) Light–current–voltage (L–I–V) characteristics of the fabricated laser with a length of 2000 μm and its temperature dependence under continuous-wave (CW) condition ranging from 25°C to 115°C. The inset shows the natural logarithm of threshold current versus stage temperature. The dashed line represents linear fitting to the experimental data.

    Figure 1.Material properties of InAs/GaAs QD lasers. (a) Cross-sectional scanning electron microscope (SEM) image of layer stack of the epi-wafer. The inset is the transmission electron microscope (TEM) image of the five QD layers. (b) Photoluminescence spectrum of the QD active layers on GaAs. The inset shows the atomic force microscope (AFM) image of an uncapped QD layer. (c) Light–current–voltage (LIV) characteristics of the fabricated laser with a length of 2000 μm and its temperature dependence under continuous-wave (CW) condition ranging from 25°C to 115°C. The inset shows the natural logarithm of threshold current versus stage temperature. The dashed line represents linear fitting to the experimental data.

    The as-grown wafers were then processed into LC-DFB laser arrays. As shown in Fig. 2(a), a waveguide width of 2.1 μm and a depth of 1.7 μm are used to support only the lowest-order transverse mode (TE00). The ridge waveguides were patterned using electron beam lithography (EBL) and etched using an optimized chlorine-based GaAs inductively coupled plasma reactive-ion etching (ICP-RIE) process, with a trapezoid cross section, sidewall slope of θ=76°, and near-zero footing [Fig. 2(b)]. The near-ideal trapezoid waveguide profile is key to a deterministic grating coupling coefficient κ and low scattering loss, both very important for improved DFB laser performance. In addition, a small footing gives rise to a larger κ, which results from the increased evanescent field into the grating region [42].

    (a) Schematic of the DFB laser structure, including the near-zero “footing” trapezoid waveguide and the α-Si gratings (not to scale); (b) cross-sectional SEM image of the trapezoid waveguide with θ=76°, with the α-Si and the ARP6200 photoresist layers also present; (c) SEM images of the etched α-Si gratings with a λ/4 phase shift in the middle; (d) microscope image of laser array.

    Figure 2.(a) Schematic of the DFB laser structure, including the near-zero “footing” trapezoid waveguide and the α-Si gratings (not to scale); (b) cross-sectional SEM image of the trapezoid waveguide with θ=76°, with the α-Si and the ARP6200 photoresist layers also present; (c) SEM images of the etched α-Si gratings with a λ/4 phase shift in the middle; (d) microscope image of laser array.

    A 10 nm Al2O3 passivation layer and a 150 nm thick α-Si (n3.495) layer were deposited by atomic layer deposition (ALD) and ICP-CVD, respectively, with both layers covering the entire sample surface terrain, including the sloped sidewalls. First-order gratings with a period Λ in a range of 194.5–199.7 nm, grating duty cycle of 1:1, and extrusion of 8 μm from the waveguide foot were designed and patterned by an EBL resist (ARP6200) alongside the ridge. The grating duty cycle is defined as the ratio of the grating width to the groove width. A λ/4 phase shift was placed in the middle of the gratings to force lasing in the defect mode. LWDM-compatible laser arrays were achieved by adjusting the grating period Λ of adjacent lasers on the same bar, with a change of ΔΛ=0.727  nm resulting in a wavelength increment of 0.8 THz. The grating corrugations were etched through the α-Si using a fluoride-based RIE process. The SEM image of Fig. 2(c) reveals a high-quality LC grating with very sharp and smooth sidewalls. With near-zero footing and a naturally chemically selective etch recipe, this dielectric grating structure is less sensitive to processing variations and, thus, more manufacturable than a buried heterostructure DFB structure. Grating coupling coefficient κ is calculated from the lateral electric field distribution and effective index of the fundamental transverse electric mode via coupled-mode theory [4244], detailed in Appendix A. For the designed devices, a first-order grating with a duty cycle of 1:1 produces a calculated κ of 3.24  mm1.

    The corrugations were subsequently covered with a 200 nm layer of SiO2 (n1.46) to form a high-contrast grating prior to contact window opening and Ti/Pt/Au p-type contact deposition [Fig. 2(d)]. An AuGe/Ni/Au n-type contact was deposited in the back after thinning the GaAs substrate down to 200 μm. After being cleaved into bars, the two facets were covered with a six-layer SiO2/TiO2 high-reflection (HR, 96.8%) coating and a one-pair SiO2/TiO2 anti-reflection (AR, 1.7%) coating to suppress the facet backreflections and increase output power. Finally, the laser arrays were mounted epitaxy-side-up on gold-coated copper heat sinks.

    3. LASING CHARACTERISTICS OF THE FABRICATED LC-DFB QD LASER

    At room temperature (25°C), a typical 2.1  μm×1.5  mm DFB laser device [Fig. 3(a)] has a turn-on voltage of 0.93 V and a differential series resistance of 10  Ω. The measured CW threshold current of 18 mA corresponds to a current density of 571  Acm2. Above the threshold, the output power follows a near-linear curve with a slope efficiency of 0.2  WA1. The AR facet output power of 26.6 mW was obtained at an injection current of 150 mA or 8.3×Ith. Figure 3(b) shows CW lasing up to 55°C with output power of >6  mW, and it is believed that the operation temperature can be further increased by more effective junction heat dissipation, by either further thinning the substrate or flip-chip bonding.

    (a) Typical L–I–V characteristics of a DFB laser with a 2.1 μm×1500 μm cavity at room temperature; (b) temperature-dependent L–I curves from the DFB laser, showing lasing up to 55°C under CW operation; (c), (d) optical spectra of the single DFB laser operating just below threshold (c) and at a drive current of 80 mA (d).

    Figure 3.(a) Typical LIV characteristics of a DFB laser with a 2.1  μm×1500  μm cavity at room temperature; (b) temperature-dependent LI curves from the DFB laser, showing lasing up to 55°C under CW operation; (c), (d) optical spectra of the single DFB laser operating just below threshold (c) and at a drive current of 80 mA (d).

    The experimental value of κ is estimated from the photonic bandgap width λs=0.436  nm observable from the below-threshold amplified spontaneous emission (ASE) spectrum of Fig. 3(c). The total coupling strength κL for this 1.5 mm long device was estimated to be 3.0 (κ=2.0), which is consistent with the theoretical value when taking into consideration the deviations of rating duty cycle (1:1.9 as shown in the inset of Fig. 2(c)]. When increasing the current to 80 mA (4.4×Ith), the central defect mode seen in the ASE spectrum rises to be the dominant longitudinal mode with an SMSR of 52.7 dB [Fig. 3(d)].

    Stable single-mode operation was observed at CW currents up to 150 mA in the temperature range of 20°C–50°C, with a linear wavelength–temperature tuning rate of 0.12 nm/°C and a quadratic wavelength–current tuning curve [Figs. 4(a) and 4(b)]. It is noteworthy that the laser maintains single-mode operation across the entire current range. This high single-mode quality is credited to the novel α-Si grating along the trapezoid waveguide, which affords reliable deterministic optical coupling.

    (a) Wavelength shift with injection currents; (b) wavelength shift with heat-sink temperature; (c), (d) optical spectra and lasing frequencies of an LWDM DFB laser array measured at 100 mA.

    Figure 4.(a) Wavelength shift with injection currents; (b) wavelength shift with heat-sink temperature; (c), (d) optical spectra and lasing frequencies of an LWDM DFB laser array measured at 100 mA.

    Across each eight-device bar, channel spacing of 0.80±0.10  THz was measured, producing wavelengths ranging from 1300.05 to 1332.41 nm [Figs. 4(c) and 4(d)], matching well with the standard LWDM grid. This wavelength range was limited by the range of grating periods we fabricated. The high-precision EBL process together with the broad gain bandwidths of QD materials affords very promising applications for both LWDM and coarse wavelength division multiplexing within the O-band.

    4. RELATIVE INTENSITY NOISE AND EXTERNAL FEEDBACK SENSITIVITY

    The RIN across the frequency range of 2.5–20 GHz is assessed using a commercial system (SYCATUS A0010A), measuring <155  dB/Hz at 4×Ith and reducing to a saturated minimum level of 165  dB/Hz at 9×Ith, as shown in Fig. 5. The relaxation oscillation (RO) of this device is suppressed due to the large damping factor of the QD as reported [45]. To the best of our knowledge, this ultralow RIN is the best-reported result among QD DFB lasers and is in good agreement with reported values in both GaAs [20] and silicon [19] based QD-FP lasers. We consider that the high output power decreases the proportion of spontaneous emission, while the suitable grating κ value ensures single-mode operation without mode hopping or spatial hole burning at high injection currents. Both can effectively suppress the disturbance of photons and carriers and thus yield an ultralow RIN.

    Measured RIN spectra at several bias currents at 25°C.

    Figure 5.Measured RIN spectra at several bias currents at 25°C.

    Finally, the performance of the QD LC-DFB lasers under coherent optical feedback was assessed using the optical measurement setups shown schematically in Fig. 6(a). The emission from the QD laser AR facet is coupled into the feedback test system by a lensed fiber with coupling efficiency of 20%–30%, and divided into feedback and detection paths by a 90/10 fiber coupler. On the feedback path (10% of the coupled power), a fibered optical circulator is used to feed the light back to the laser cavity. Since the external cavity resonance frequency (17.25 MHz in this stage) is much less than the laser RO frequency, the impact of the feedback phase is negligible. The feedback strength rext is defined as the ratio of the returning power to the laser free-space output power, and precise returning power can be obtained by calculating the product of the power (detected by a power meter) and the lensed fiber coupling efficiency. The optical feedback intensity is controlled by changing the operating current of a polarization-maintaining boost optical amplifier (BOA, Thorlabs S9FC1132P). A filter with 0.8 nm bandwidth is employed to spectrally suppress the ASE from the BOA. A polarization controller is inserted in the external cavity to compensate for the polarization rotation in the fiber. The insertion losses produced in the lens fiber, BOA, beam splitter, and each connector are carefully calibrated. The remaining 90% of the coupled power is sent to a high-resolution (0.03 nm) optical spectrum analyzer (OSA, Anritsu MS9740A) or RIN measurement system (SYCATUS A0010A) to monitor the evolution of spectra and RIN as the feedback strength rext varies. For the whole measurement, the DFB laser is mounted on a thermo-electric cooler (TEC) operated at 25°C.

    (a) Experimental setup used for the long-delay feedback measurements. LF, lens fiber; PM, power meter; BOA, boost optical amplifier; OSA, optical spectrum analyzer; RIN, relative intensity noise; PC, polarization controller; ISO, optical isolator; BPF, bandpass filter. (b) Evolution of the SMSR with increasing feedback strength; the inset is the optical spectrum of the DFB laser as the feedback strength increases. (c) Change of RIN in the same DFB laser under 2.5×Ith, 3×Ith, and 4×Ith current injections. The inset is the frequency domain plot of RIN as a function of increasing feedback strength.

    Figure 6.(a) Experimental setup used for the long-delay feedback measurements. LF, lens fiber; PM, power meter; BOA, boost optical amplifier; OSA, optical spectrum analyzer; RIN, relative intensity noise; PC, polarization controller; ISO, optical isolator; BPF, bandpass filter. (b) Evolution of the SMSR with increasing feedback strength; the inset is the optical spectrum of the DFB laser as the feedback strength increases. (c) Change of RIN in the same DFB laser under 2.5×Ith, 3×Ith, and 4×Ith current injections. The inset is the frequency domain plot of RIN as a function of increasing feedback strength.

    Figures 6(b) and 6(c) present the evolution of SMSR and RIN with the laser operating at 4×Ith. This device shows RO peaks at a low injection current, which move from 2 GHz to about 5.5 GHz with the increasing bias current [inset of Fig. 6(c)]. Little sign of deterioration is observed until the optical feedback levels reach rext=5.9% (12.3  dB). In particular, the SMSR of the laser is found to be still above 50 dB under this feedback strength. RIN levels (at the frequency of 5 GHz) rise only slowly until rext=5.6% (12.5  dB), without any visible periodic or chaotic oscillations in the RIN spectra. A sharp increase in RIN indicates a transition to the coherence collapse regime beyond this critical level of optical feedback fext,c. For DFB lasers, fext,c is strongly associated with the normalized coupling coefficient κL (L is the length of the laser cavity), where an increased κL will lead to an increased fext,c [46,47].

    5. DISCUSSION

    Given the rapid development of QD-DFB lasers at 1310 nm on both GaAs and silicon substrates, it is useful to compare the performance of our device with those reported in the literature. Table 1 lists the threshold current (current density), power, highest work temperature, SMSR, RIN, and optical feedback tolerance of reported DFB lasers together with our device. In general, compared with commercial QW lasers, QD lasers exhibit excellent performance in terms of high operating temperature, low RIN, and high optical feedback tolerance owing to their stronger carrier confinement, larger damping factor, and smaller α factor. QD lasers with buried [24,52] gratings show good feedback tolerance, but the RIN needs to be further improved. Our device simultaneously achieves high output power (26.6 mW), ultralow RIN (165  dB/Hz), and high tolerance to optical feedback (12.3  dB).

    Comparison of the Performance of Our Device with Reference QD DFB Laser at 1310 nm

    YearSubstrateGratingκ  (mm1)Threshold Current (mA)Threshold Current Density (A/cm2)Power (mW)SMSR (dB)T(°C)RIN (dB/Hz)Anti-feedback (dB)Ref.
    2003GaAsGaAs sidewall39.350–14[48]
    2005GaAsMetal sidawall512>50[49]
    2011GaAsInGa/GaAs buried46.8104580[33]
    2011GaAsCr sidewall181500>105385–12[50]
    2014GaAsInGa/GaAs buried2.543.81830345860[51]
    2018GaAsGaAs sidewall3017102351[37]
    2018GaAsInGa/GaAs buried46.2110720>4070–150–8[52]
    2018Heterogenous integrated GaAs/SiSi7.79.52052.547100[53]
    2018Monolithic integrated GaAS/SiGaAs sidewall4.2125501.550[40]
    2020Monolithic integrated GaAs/SiAIGaAs/GaAs buried4.5204404.45070[34]
    2021Monolithic integrated GaAs/SiSi542502.86075[54]
    2021Heterogeneous GaAs/oxide/SiSi156.713476170–125[55]
    2021GaAsInGaP/GaAs buried1.69.3155055–150–6[24]
    2021GaAsAmorphous Si sidewall2.01857126.652.755–165–12.3This work

    To conclude, by implementing a novel first-order α-Si Bragg grating LC to a near-ideal trapezoid GaAs ridge waveguide, high-performance 1300 nm InAs/GaAs QD DFB laser arrays have been realized with high power, ultralow RIN, high robustness against optical feedback, and accurate LWDM grid. These excellent features make the device a very attractive candidate for high-performance digital and analog WDM optical transmission systems as well as on-chip sources for PICs where optical isolators are not readily available. In the near future, decreasing uniformities in the size or energy level of QDs while keeping the QD density [17], is expected to further improve the material modal gain and therefore thermal stability.

    The novel grating structure affords accurate and deterministic grating coupling coefficient κ, which can be engineered independent of the laser material epitaxy process. The scheme can therefore be readily implemented on other material systems such as InP-, GaAs/Si- and GaSb-based compound semiconductor lasers, and in other wavelength windows by scaling the size of the grating and using low absorption dielectric materials for those wavelengths. The simplicity and versatility of the regrowth-free scheme make it possible to establish a new paradigm of semiconductor laser manufacturing. In this paradigm, multiple types of grating-based semiconductor lasers (including DFB, DBR, and tunable lasers) can be jointly manufactured on the same fabrication platform, with different active materials and operating wavelengths, which would potentially reduce their manufacturing cost very significantly.

    APPENDIX A: CALCULATED AND EXPERIMENTAL VALUE OF κ

    According to coupled mode theory, coupling coefficient κ can be calculated by [42] κ=n22n12λ0neff·sin(πmΛ)m·Γ,where n1 and n2 are the refractive index of SiO2 and α-Si grating, respectively. λ0 is the Bragg wavelength, m=1 is the grating diffraction order, Λ is the duty cycle, and Γ is the electric field overlap factor in the grating region. neff is the effective refractive index of the mode traveling in the ridge waveguide. The average refractive index of the αSi/SiO2 grating can be obtained by navg=Λn22+(1Λ)n12.

    The neff and Γ of TE00 mode can be obtained by solving the transverse light field distribution after replacing the grating index by navg. Table 2 illustrates the refractive index used in the theoretical calculation for the α-Si first-order LC grating. Figure 7(a) illustrates the dependence of coupling coefficient κ on the grating duty cycle at different ridge waveguide etch depths. It can be seen that κ is asymmetric with respect to the duty cycle and has a peak value at the etch depth of 1.7 μm. Figures 7(b) and 7(c) show coupling coefficient κ as a function of grating thickness (d) and grating extrusion length (l) from the waveguide foot. It shows that the value of κ saturates as d or l increases, which is expected since the optical field in the grating decays rapidly away from the foot of the ridge. For our fabricated devices with d=150  nm and l=8  μm, κ is not sensitive to the variation in d and l, which should be preferable from the perspective of κ stability.

    Material Refractive Index Used in the Simulation at Wavelength of 1310 nm

    MaterialRefractive Index
    GaAs3.41
    Al0.4GaAs3.25
    Al2O31.75
    Si3.49
    SiO21.45
    BCBa1.56

    The material B-staged bisbenzocyclobutene (BCB, type: CYCLOTENE 4022-35) is used to planarize the laser ridge waveguide. Therefore, in our simulation, the background refractive index is set as 1.56.

    Coupling coefficient κ as a function of (a) grating duty cycle, (b) grating thickness, and (c) grating length.

    Figure 7.Coupling coefficient κ as a function of (a) grating duty cycle, (b) grating thickness, and (c) grating length.

    Experimental grating coupling coefficient κ is estimated from the bandgap width λs=0.436  nm [as shown in Fig. 3(c)] by using the relation [56]κ=(πngλsλB2)2(πLg)2,where κ is the grating coupling coefficient, ng is the group index, λB is the lasing mode, and Lg is the grating length. The total coupling strength κL for the 1.5 mm long device was estimated to be 3.0 (κ=2.0), which is very consistent with our theoretical design if taking into consideration the deviations of grating duty cycle [1:1.9 as shown in the inset of Fig. 2(c)]. Further increase in κ by a factor of >2 is therefore possible by increasing the duty cycle or grating thickness.

    References

    [1] D. Botez, G. J. Herskowitz. Components for optical communications systems: a review. Proc. IEEE, 68, 689-731(1980).

    [2] O. Brox, F. Bugge, A. Mogilatenko, E. Luvsandamdin, A. Wicht, H. Wenzel, G. Erbert. Distributed feedback lasers in the 760 to 810  nm range and epitaxial grating design. Semicond. Sci. Technol., 29, 095018(2014).

    [3] E. Di Gaetano, S. Watson, E. McBrearty, M. Sorel, D. J. Paul. Sub-megahertz linewidth 780.24  nm distributed feedback laser for 87Rb applications. Opt. Lett., 45, 3529-3532(2020).

    [4] V. Schkolnik, O. Hellmig, A. Wenzlawski, J. Grosse, A. Kohfeldt, K. Döringshoff, A. Wicht, P. Windpassinger, K. Sengstock, C. Braxmaier, M. Krutzik, A. Peters. A compact and robust diode laser system for atom interferometry on a sounding rocket. Appl. Phys. B, 122, 217(2016).

    [5] Y. He, H. An, J. Cai, C. Galstad, S. Macomber, M. Kanskar. 808 nm broad area DFB laser for solid-state laser pumping application. Electron. Lett., 45, 163-164(2009).

    [6] S. Stephan, D. Frederic, A. Markus-Christian. Novel InP- and GaSb-based light sources for the near to far infrared. Semicond. Sci. Technol., 31, 113005(2016).

    [7] M. Hoppe, C. Aßmann, S. Schmidtmann, T. Milde, M. Honsberg, T. Schanze, J. Sacher. GaSb-based digital distributed feedback filter laser diodes for gas sensing applications in the mid-infrared region. J. Opt. Soc. Am. B, 38, B1-B8(2021).

    [8] S. Najda, P. Perlin, M. Leszczyński, T. Slight, W. Meredith, M. Schemmann, H. Moseley, J. Woods, R. Valentine, S. Kalra, P. Mossey, E. Theaker, M. Macluskey, G. Mimnagh, W. Mimnagh. A multi-wavelength (u.v. to visible) laser system for early detection of oral cancer. Proc. SPIE, 9328, 932809(2015).

    [9] T. Miyajima, T. Tojyo, T. Asano, K. Yanashima, S. Kijima, T. Hino, M. Takeya, S. Uchida, S. Tomiya, K. Funato, T. Asatsuma, T. Kobayashi, M. Ikeda. GaN-based blue laser diodes. J. Phys.: Condens. Matter, 13, 7099(2001).

    [10] J. C. Palais. Fiber Optic Communications(1988).

    [11] T. Sudo, Y. Matsui, G. Carey, A. Verma, D. Wang, V. Lowalekar, M. Kwakernaak, F. Khan, N. Dalida, R. Patel, A. Nickel, B. Young, J. Zeng, Y. L. Ha, C. Roxlo. Challenges and opportunities of directly modulated lasers in future data center and 5G networks. Optical Fiber Communications Conference and Exhibition (OFC), 1-3(2021).

    [12] C. P. Hsu, B. Li, B. Solano-Rivas, A. R. Gohil, P. H. Chan, A. D. Moore, V. Donzella. A review and perspective on optical phased array for automotive LiDAR. IEEE J. Sel. Top. Quantum Electron., 27, 8300416(2021).

    [13] D. N. Hutchison, J. Sun, J. K. Doylend, R. Kumar, J. Heck, W. Kim, C. T. Phare, A. Feshali, H. Rong. High-resolution aliasing-free optical beam steering. Optica, 3, 887-890(2016).

    [14] M.-C. Amann, M. Ortsiefer. Long-wavelength (λ1.3μm) InGaAlAs–InP vertical-cavity surface-emitting lasers for applications in optical communication and sensing. Phys. Status Solidi A, 203, 3538-3544(2006).

    [15] A. Liu, P. Wolf, J. A. Lott, D. Bimberg. Vertical-cavity surface-emitting lasers for data communication and sensing. Photon. Res., 7, 121-136(2019).

    [16] H. Y. Liu, S. L. Liew, T. Badcock, D. J. Mowbray, M. S. Skolnick, S. K. Ray, T. L. Choi, K. M. Groom, B. Stevens, F. Hasbullah, C. Y. Jin, M. Hopkinson, R. A. Hogg. p-doped 1.3 μm InAs/GaAs quantum-dot laser with a low threshold current density and high differential efficiency. Appl. Phys. Lett., 89, 073113(2006).

    [17] T. Kageyama, K. Nishi, M. Yamaguchi, R. Mochida, Y. Maeda, K. Takemasa, Y. Tanaka, T. Yamamoto, M. Sugawara, Y. Arakawa. Extremely high temperature (220°C) continuous-wave operation of 1300-nm-range quantum-dot lasers. The European Conference on Lasers and Electro-Optics, PDA_1(2011).

    [18] Y.-G. Zhou, C. Zhou, C.-F. Cao, J.-B. Du, Q. Gong, C. Wang. Relative intensity noise of InAs quantum dot lasers epitaxially grown on Ge. Opt. Express, 25, 28817-28824(2017).

    [19] M. Liao, S. Chen, Z. Liu, Y. Wang, L. Ponnampalam, Z. Zhou, J. Wu, M. Tang, S. Shutts, Z. Liu, P. M. Smowton, S. Yu, A. Seeds, H. Liu. Low-noise 1.3 μm InAs/GaAs quantum dot laser monolithically grown on silicon. Photon. Res., 6, 1062-1066(2018).

    [20] A. Capua, L. Rozenfeld, V. Mikhelashvili, G. Eisenstein, M. Kuntz, M. Laemmlin, D. Bimberg. Direct correlation between a highly damped modulation response and ultra low relative intensity noise in an InAs/GaAs quantum dot laser. Opt. Express, 15, 5388-5393(2007).

    [21] D. A. I. Marpaung. High dynamic range analog photonic links(2009).

    [22] B. Dong, J.-D. Chen, F.-Y. Lin, J. C. Norman, J. E. Bowers, F. Grillot. Dynamic and nonlinear properties of epitaxial quantum-dot lasers on silicon operating under long- and short-cavity feedback conditions for photonic integrated circuits. Phys. Rev. A, 103, 033509(2021).

    [23] H. Huang, J. Duan, B. Dong, J. Norman, D. Jung, J. E. Bowers, F. Grillot. Epitaxial quantum dot lasers on silicon with high thermal stability and strong resistance to optical feedback. APL Photon., 5, 016103(2020).

    [24] B. Dong, J. Duan, H. Huang, J. C. Norman, K. Nishi, K. Takemasa, M. Sugawara, J. E. Bowers, F. Grillot. Dynamic performance and reflection sensitivity of quantum dot distributed feedback lasers with large optical mismatch. Photon. Res., 9, 1550-1558(2021).

    [25] J. C. Norman, D. Jung, Y. Wan, J. E. Bowers. Perspective: the future of quantum dot photonic integrated circuits. APL Photon., 3, 030901(2018).

    [26] C. Hantschmann, Z. Liu, M. Tang, S. Chen, A. J. Seeds, H. Liu, I. H. White, R. V. Penty. Theoretical study on the effects of dislocations in monolithic III-V lasers on silicon. J. Lightwave Technol., 38, 4801-4807(2020).

    [27] J. C. Norman, D. Jung, Z. Zhang, Y. Wan, S. Liu, C. Shang, R. W. Herrick, W. W. Chow, A. C. Gossard, J. E. Bowers. A review of high-performance quantum dot lasers on silicon. IEEE J. Quantum Electron., 55, 2000511(2019).

    [28] S. Chen, W. Li, J. Wu, Q. Jiang, M. Tang, S. Shutts, S. N. Elliott, A. Sobiesierski, A. J. Seeds, I. Ross, P. M. Smowton, H. Liu. Electrically pumped continuous-wave III–V quantum dot lasers on silicon. Nat. Photonics, 10, 307-311(2016).

    [29] J. C. Norman, R. P. Mirin, J. E. Bowers. Quantum dot lasers—history and future prospects. J. Vac. Sci. Technol. A, 39, 020802(2021).

    [30] D. Jung, R. Herrick, J. Norman, K. Turnlund, C. Jan, K. Feng, A. C. Gossard, J. E. Bowers. Impact of threading dislocation density on the lifetime of InAs quantum dot lasers on Si. Appl. Phys. Lett., 112, 153507(2018).

    [31] T. Septon, A. Becker, S. Gosh, G. Shtendel, V. Sichkovskyi, F. Schnabel, A. Sengül, M. Bjelica, B. Witzigmann, J. P. Reithmaier, G. Eisenstein. Large linewidth reduction in semiconductor lasers based on atom-like gain material. Optica, 6, 1071-1077(2019).

    [32] Z. Lu, K. Zeb, J. Liu, E. Liu, L. Mao, P. Poole, M. Rahim, G. Pakulski, P. Barrios, W. Jiang, D. Poitras. Quantum dot semiconductor lasers for 5G and beyond wireless networks. Proc. SPIE, 11690, 116900N(2021).

    [33] K. Takada, Y. Tanaka, T. Matsumoto, M. Ekawa, H. Z. Song, Y. Nakata, M. Yamaguchi, K. Nishi, T. Yamamoto, M. Sugawara, Y. Arakawa. Wide-temperature-range 10.3  Gbit/s operations of 1.3  μm high-density quantum-dot DFB lasers. Electron. Lett., 47, 206-208(2011).

    [34] Y. Wan, J. C. Norman, Y. Tong, M. J. Kennedy, W. He, J. Selvidge, C. Shang, M. Dumont, A. Malik, H. K. Tsang, A. C. Gossard, J. E. Bowers. 1.3  μm quantum dot-distributed feedback lasers directly grown on (001) Si. Laser Photon. Rev., 14, 2000037(2020).

    [35] C. B. Cooper, S. Salimian, H. F. Macmillan. Reactive ion etch characteristics of thin InGaAs and AlGaAs stop-etch layers. J. Electron. Mater., 18, 619-622(1989).

    [36] G. C. Desalvo, W. F. Tseng, J. Comas. ChemInform abstract: etch rates and selectivities of citric acid/hydrogen peroxide on GaAs, Al0.3Ga0.7As, In0.2Ga0.8As, In0.53Ga0.47As, In0.52Al0.48As, and InP. ChemInform, 23, 309(1992).

    [37] Q. Li, X. Wang, Z. Zhang, H. Chen, Y. Huang, C. Hou, J. Wang, R. Zhang, J. Ning, J. Min, C. Zheng. Development of modulation p-doped 1310  nm InAs/GaAs quantum dot laser materials and ultrashort cavity Fabry–Perot and distributed-feedback laser diodes. ACS Photon., 5, 1084-1093(2018).

    [38] S. Forouhar, R. M. Briggs, C. Frez, K. J. Franz, A. Ksendzov. High-power laterally coupled distributed-feedback GaSb-based diode lasers at 2  μm wavelength. Appl. Phys. Lett., 100, 031107(2012).

    [39] S. Masui, K. Tsukayama, T. Yanamoto, T. Kozaki, S.-I. Nagahama, T. Mukai. CW operation of the first-order AlInGaN 405  nm distributed feedback laser diodes. Jpn. J. Appl. Phys., 45, L1223-L1225(2006).

    [40] Y. Wang, S. Chen, Y. Yu, L. Zhou, L. Liu, C. Yang, M. Liao, M. Tang, Z. Liu, J. Wu, W. Li, I. Ross, A. J. Seeds, H. Liu, S. Yu. Monolithic quantum-dot distributed feedback laser array on silicon. Optica, 5, 528-533(2018).

    [41] C. A. Yang, S. W. Xie, Y. Zhang, J. M. Shang, S. S. Huang, Y. Yuan, F. H. Shao, Y. Zhang, Y. Q. Xu, Z. C. Niu. High-power, high-spectral-purity GaSb-based laterally coupled distributed feedback lasers with metal gratings emitting at 2  μm. Appl. Phys. Lett., 114, 021102(2019).

    [42] A. Laakso, J. Karinen, M. Dumitrescu. Modeling and design particularities for distributed feedback lasers with laterally-coupled ridge-waveguide surface gratings. Proc. SPIE, 7933, 79332K(2011).

    [43] W. Streifer, D. Scifres, R. Burnham. Coupling coefficients for distributed feedback single- and double-heterostructure diode lasers. IEEE J. Quantum Electron., 11, 867-873(1975).

    [44] W.-Y. Choi, J. C. Chen, C. G. Fonstad. Evaluation of coupling coefficients for laterally-coupled distributed feedback lasers. Jpn. J. Appl. Phys., 35, 4654-4659(1996).

    [45] J. Duan, H. Huang, B. Dong, J. C. Norman, Z. Zhang, J. E. Bowers, F. Grillot. Dynamic and nonlinear properties of epitaxial quantum dot lasers on silicon for isolator-free integration. Photon. Res., 7, 1222-1228(2019).

    [46] F. Grillot, B. Thedrez, D. Guang-Hua. Feedback sensitivity and coherence collapse threshold of semiconductor DFB lasers with complex structures. IEEE J. Quantum Electron., 40, 231-240(2004).

    [47] Q. Zou, K. Merghem, S. Azouigui, A. Martinez, A. Accard, N. Chimot, F. Lelarge, A. Ramdane. Feedback-resistant p-type doped InAs/InP quantum-dash distributed feedback lasers for isolator-free 10 Gb/s transmission at 1.55 μm. Appl. Phys. Lett., 97, 231115(2010).

    [48] H. Su, L. Zhang, A. L. Gray, R. Wang, T. C. Newell, K. J. Malloy, L. F. Lester. High external feedback resistance of laterally loss-coupled distributed feedback quantum dot semiconductor lasers. IEEE Photon. Technol. Lett., 15, 1504-1506(2003).

    [49] H. Su, L. F. Lester. Dynamic properties of quantum dot distributed feedback lasers: high speed, linewidth and chirp. J. Phys. D, 38, 2112-2118(2005).

    [50] S. Azouigui, D.-Y. Cong, A. Martinez, K. Merghem, Q. Zou, J.-G. Provost, B. Dagens, M. Fischer, F. Gerschütz, J. Koeth, I. Krestnikov, A. Kovsh, A. Ramdane. Temperature dependence of dynamic properties and tolerance to optical feedback of high-speed 1.3  μm DFB quantum-dot lasers. IEEE Photon. Technol. Lett., 23, 582-584(2011).

    [51] M. Stubenrauch, G. Stracke, D. Arsenijević, A. Strittmatter, D. Bimberg. 15  Gb/s index-coupled distributed-feedback lasers based on 1.3  μm InGaAs quantum dots. Appl. Phys. Lett., 105, 011103(2014).

    [52] M. Matsuda, N. Yasuoka, K. Nishi, K. Takemasa, T. Yamamoto, M. Sugawara, Y. Arakawa. Low-noise characteristics on 1.3-μm-wavelength quantum-dot DFB lasers under external optical feedback. IEEE International Semiconductor Laser Conference (ISLC), 1-2(2018).

    [53] S. Uvin, S. Kumari, A. De Groote, S. Verstuyft, G. Lepage, P. Verheyen, J. Van Campenhout, G. Morthier, D. Van Thourhout, G. Roelkens. 1.3  μm InAs/GaAs quantum dot DFB laser integrated on a Si waveguide circuit by means of adhesive die-to-wafer bonding. Opt. Express, 26, 18302-18309(2018).

    [54] Y. Wan, C. Xiang, J. Guo, R. Koscica, M. J. Kennedy, J. Selvidge, Z. Zhang, L. Chang, W. Xie, D. Huang, A. C. Gossard, J. E. Bowers. High speed evanescent quantum-dot lasers on Si. Laser Photon. Rev., 15, 210057(2021).

    [55] D. Liang, S. Srinivasan, A. Descos, C. Zhang, G. Kurczveil, Z. Huang, R. Beausoleil. High-performance quantum-dot distributed feedback laser on silicon for high-speed modulations. Optica, 8, 591-593(2021).

    [56] G. Liu, G. Zhao, J. Sun, D. Gao, Q. Lu, W. Guo. Experimental demonstration of DFB lasers with active distributed reflector. Opt. Express, 26, 29784-29795(2018).

    Zhuohui Yang, Zhengqing Ding, Lin Liu, Hancheng Zhong, Sheng Cao, Xinzhong Zhang, Shizhe Lin, Xiaoying Huang, Huadi Deng, Ying Yu, Siyuan Yu. High-performance distributed feedback quantum dot lasers with laterally coupled dielectric gratings[J]. Photonics Research, 2022, 10(5): 1271
    Download Citation