HTML
-
Parallel to commercializing the first generation of heterogeneous innovation, primarily the InP-on-Si platform for pluggable optical transceivers used in data centres and high-performance computing, R&D continues to thrive on expanding the spectral coverage and developing novel device structures for broader and emerging applications. One obvious motivation is to expand beyond the 1310 and 1550 nm telecom windows. This is naturally accomplished by physically mating and optically coupling appropriate materials with respect to the spectral window of interest. In other words, transparent passive waveguiding materials are selected for combination with the Si substrate and active thin film materials with the appropriate band gap/epitaxial structure on III-V to develop an efficient active–passive optical coupling.
Silicon nitride (SiNx) has gained significant interest in the past ten years to challenge the dominant role of Si in the light propagation medium on Si substrates, particularly in scenarios where data modulation and detection are not strictly required. SiNx waveguides are attractive owing to their extremely low propagation loss, for example, 0.045 ± 0.04 dB/m in the low-confinement case69, and 0.13 ± 0.05 dB/m in the high-confinement case70, extremely large transparency window extending from visible to mid-infrared (MIR), medium refractive index ~2 for a good balance of compactness and sensitivity to fabrication imperfection, and favourable nonlinear characteristics71. A SiNx/SiO2/Si substrate structure serves as an ideal platform to extend heterogeneous Si photonics beyond the 1.1 μm cutoff line in Si for new applications in displays, AR/VR, metrology, spectroscopy, biomedical, sensing, quantum processing, etc. The shortest wavelength laser that has been heterogeneously integrated on Si72, 73 thus far adopted a vertical optical coupling scheme74, 75. As shown in Fig. 3a, a ‘half’ InGaAs-based 855 nm vertical cavity surface emitting laser (VCSEL) epitaxial structure was transferred onto a special dielectric-on-Si substrate by polymer bonding. The dielectric-on-Si substrates are composed of a 300 nm thick SiNx layer sandwiched between SiO2 cladding above 20 pairs of Ta2O6/SiO2 dielectric, which serves as the bottom distributed Bragg reflector (DBR) mirror for vertical directional lasing. A weak intra-cavity diffraction grating was patterned prior to bonding and resided inside the VCSEL cavity to tap off light into the in-plane SiNx waveguide. The grating was designed to enable a TE-polarization favoured intra-cavity grating/dielectric DBR combination with higher reflection and lower coupling, that is, mirror loss, into the SiNx waveguide. The III-V top DBR was covered by 100 nm Au to further suppress vertical emission. A 1 mA threshold and fibre-coupled bi-directional emission up to ~75 μW each were obtained in Fig. 3b owing to the grating symmetry with respect to the vertical FP lasing cavity. Fundamental transverse mode lasing at 856.5 nm was observed. Suppression of approximately 30 dB to higher order transverse modes is much more effective than conventional VCSEL designs without an intra-cavity grating because higher-order modes contain spatial frequency components with larger off-normal angles than the fundamental mode73.
Fig. 3
a Cross-sectional schematic of VCSEL with in-plane out-coupling into a SiNx waveguide, and b device light-current-voltage (LIV) characteristic showing symmetric coupling to the waveguides on the right and left and cw lasing at 856.5 nm73.More recently, a λ = 900 nm in-plane heterogeneous laser and photodetector were also developed by using molecular bonding to enable the GaAs-based gain medium onto the SiNx routing waveguide76, 77. Fig. 4a shows scanning electron microscopy (SEM) images and a simulated fundamental TE mode image of the 350 μm-thick SiNx passive waveguide and ~2 μm-thick GaAs active rib waveguide. A non-disclosed proprietary coupling scheme was implemented to overcome a large refractive index difference of over 1.5 between the SiNx and GaAs materials and allowed a coupling coefficient of as much as 70% in the experiment and over 90% in the case achieved by simulation. Efficient coupling is enabled without the need to use thin layers in the high-index region or prohibitively narrow taper tips to match the modes and their effective indices. The novel design concept and efficient coupling could be applied to GaN/InGaN material system for realizing visible lasers on Si in the future. Fig. 4c shows a top-view SEM image showing InGaAs multiple QWs-based lasers coupled to a SiNx waveguide and routed through an S-bend to an on-chip waveguide-coupled GaAs monitor photodetector (PD). The temperature-dependent light-current (LI) characteristic in Fig 4b shows a robust operation between 20 and 100 °C, benefiting from the larger band gap in the laser-active region. An output power of 10 mW was obtained for a 200 mA injection current at 20 °C. The spectra in Fig. 4b inset show the typical Fabry-Perot (FP) lasing characteristic, indicating the potential to develop more advanced laser cavities on this platform76, 77.
Fig. 4
a SEM images of fabricated SiNx and GaAs waveguides and their respective fundamental TE mode profiles, b temperature-dependent LIVs and room-temperature spectra at different injection currents, and c top-view microscopic image showing the integrated laser, passive SiNx waveguide, and PD77.On the other side of the spectrum, heterogeneous lasers with the longest wavelength at 4.8 μm were built by combining a quantum cascade laser (QCL) gain medium with a special Si-on-nitride-on-insulator (SONOI) substrate78. Si is optically transparent to 8 μm, whereas significant absorption in the BOX layer starts at approximately 4 μm79. The solution here is to use a 1.5 μm-thick top Si device layer and insert a layer of 400 nm-thick SiNx above the 3 μm-thick BOX layer such that the optical mode has minimal overlap with lossy SiO2. This platform can be readily employed to support ultra-broadband photonic integration from 350 nm to 6.5 μm80. Such heterogeneous III-V-on-SONOI layer stacks were prepared by depositing SiNx on a SiO2/Si substrate, followed by transfer of the top Si device layer from another regular SOI wafer using molecular bonding. Then, the distributed feedback (DFB) grating with a 1/4λ shift in the centre and Si waveguide were defined prior to transferring the QCL epitaxial thin film. Fig. 5a shows a schematic of the finished device after post-III-V bonding fabrication, where more than 70% of the fundamental TM lasing mode overlapped with the active region of the 30-pair InGaAs/InAlAs superlattices81. As shown in Fig. 5b, output powers of more than 100 and nearly 50 mW were measured under pulsed current injection at stage temperatures of 10 and 100 °C, respectively82. The conservatively designed 3 μm-thick BOX layer and 400 nm-thick SiNx prevented efficient thermal dissipation and cw operation. The corresponding threshold current densities are 1 and 1.59 kA/cm2. The wavelength shift of 0.25 nm/°C in Fig. 5c corresponds to the expected DFB lasing modal change instead of the gain peak shift. A multi-wavelength heterogeneous QCL DFB laser array with an array waveguide grating (AWG) at similar wavelengths was also demonstrated83. Numerous demonstrations based on different III-V thin film selections have been reported to cover light generation, amplification, and detection in the range of –2 to +4.5 μm for a variety of MIR applications84-90.
-
Innovations in heterogeneous materials and device engineering provide critical active functionalities that are inherently lacking in Si and extend the useful spectral window, as discussed above. Recently, the lower loss of silicon waveguides combined with heterogeneous materials has pushed device performance well beyond the abilities of their monolithic counterparts. A narrow-linewidth laser on Si is such an example that is progressing rapidly. The Lorentzian linewidth, that is, the Schawlow-Towns linewidth91 or intrinsic linewidth, is typically referred to as the ultimate, quantum noise-determined laser linewidth limit92. Quantum noise in diode lasers is caused by random spontaneous emission contamination into stimulated emission via photon field intrusion (phase noise Δνφ) and photon-carrier balance disturbance (frequency noise ΔνN). A modification of the Schawlow-Townes-Henry equation for the Lorentzian linewidth Δν 92-94 is shown in Eq.1 below:
$$ \Delta v=\Delta v_{\varphi}+\Delta v_{N}=\frac{R_{s p}}{4 \pi n_{p}}+\frac{R_{s p}}{4 \pi n_{p}} \alpha_{H}^{2} $$ (1) where αH is the linewidth enhancement factor, an important material-dependent parameter that represents the relationship between the imaginary and real parts of the refractive index95. Rsp and np represent the spontaneous emission rate coupling to the lasing mode and the total number of photons stored in the cavity, respectively. A more intuitive expression from the physical laser cavity design perspective is obtained by rewriting Eq. 1 as Eq. 292:
$$ \Delta v=\frac{\pi h v^{3} n_{s p}}{P Q_{E} Q}\left(1+\alpha_{H}^{2}\right) $$ (2) where h is Planck’s constant, ν is the laser frequency, nsp is the population inversion factor, P is the total laser output power, and Q (QE) is the loaded (external) quality factor of the laser cold cavity. The two quality factors here are directly correlated to the internal cavity loss and mirror loss. Therefore, the reduction of absorption in the laser active region and doped regions, passive waveguide loss, and mirror loss, all of which are quantified in a distributed manner through the entire cavity, are effective ways to realise a narrow laser linewidth.
Vertical optical confinement factors for III-V active region and doped layers in heterogeneous platforms can be conveniently engineered by adjusting the Si waveguide dimension, III-V epitaxial structure, and proximity in between16. As shown schematically in Fig. 6c, a 150 nm-thick SiO2 spacer between Si and III-V resulted in >95% lasing modal confinement in low-loss Si, 1.5% in III-V, and only 0.2% in QW96. In addition, a 1D photonic band gap engineered high-Q (~106) laser cavity in Fig. 6a introduced another layer of loss reduction. A Si grating section (Ld) with a varying grating groove width in the transverse direction created parabolically modulated frequency band edges along the mode propagation. Thus, the lasing mode can be localised tightly in this potential well ‘V’ with minimal coupling loss to the radiation modes. A shallow Si rib waveguide was also purposely designed to minimise scattering from the etched sidewalls. A Lorentzian linewidth of approximately 1.1 kHz with amplifier noise included was measured in this solitary heterogeneous DFB laser96, representing a linewidth reduction of 600× relative to the best monolithic III-V counterpart.
Fig. 6
a Schematic photonic band gap structure and top-view of the high-Q grating-based laser cavity geometry in96 and c its heterogeneous cross-sectional schematic with lasing mode profile, b schematic of a MQW-based heterogeneous extended cavity laser structure92, and d SEM cross-sectional image of a similar QD-based extended cavity laser105.In contrast to the solitary design, where the distributed internal cavity loss remains independent of the total cavity length, extended/external cavity design can effectively reduce the longitudinal optical confinement factor to absorptive III-V thin film by extending a large portion of the laser cavity to lower-loss Si (e.g. 0.16 dB/cm for a quasi-single mode waveguide97) or a much lower-loss SiNx waveguide98-100. Therefore, both distributed internal and mirror losses are reduced as the laser cavity extension. Further leverage of the detuned loading or optical negative feedback effect101 in designs with narrow mirror reflectivity bandwidth can dilute the negative impact of the linewidth enhancement factor αH. Without providing the details, one can comprehend detuned loading as an elegant way to build negative/favourable or positive/undesired feedback by setting the lasing frequency at a negative or positive detuned side with respect to the frequency at the maximum mirror reflectivity. Lasing frequency perturbation triggers a chain of events that affects the reflectivity, photon density, carrier density, refractive index, which are fed back to the lasing frequency. A negative feedback loop has been proved to be of great use to stabilise the lasing frequency and suppress quantum noise from the QW gain medium102. Because the αH factor is typically 2-5 for the QW structure (>3 in most of our devices) 95, 103, 104, this effect becomes a significant advantage over solitary lasers in which the Lorentzian linewidth scales with (1 + αH2) according to Eq. 1. Choices such as single-frequency Bragg gratings, sampled gratings, high-Q ring resonators, or a combination of them serve well as this extended passive section. High-Q ring resonators are particularly attractive for three reasons. Firstly, a high-Q ring resonator with a long photon lifetime at resonance frequency is analogous to the optical delay lines but with a much smaller physical footprint. Secondly, their extremely narrow bandwidth offers excellent spectral purity over a broad range. Finally, single or multiple coupled rings enable wide spectral tunability with low tuning power consumption. Fig. 6b illustrates such an extended heterogeneous laser design, where a coupled-quad ring structure was designed as the back mirror102. More than two coupled rings in this design help alleviate the trade-off between insertion loss and spectral bandwidth, both governed by the ring coupling coefficient. Therefore, the overall Q of the cavity can be enhanced without overly sacrificing the coupling coefficient102. At an output power of approximately 2 mW, optimal control of the gain section and rings with respect to the detuning loading effect led to the narrowest Lorentzian linewidth of 140 Hz at an emission wavelength of 1565 nm that was known at the time. The same device also exhibited impressive 120 nm-wide wavelength tuning from 1484 to 1604 nm 102.
If large spectral tunability is not required, a high-Q single-frequency Bragg mirror based on a low-loss waveguide is a much simpler choice. Xiang et al. recently fabricated a heterogeneous laser in the C-band using a 20 mm-long on-chip SiNx spiral-shaped DBR as the back mirror100. Except for the SiNx/Si taper region, the SiNx strip waveguide DBR was completely surrounded by SiO2 without Si and the III-V thin film above. This resulted in a highly thermally stable wavelength of 10.46 pm/°C, more than 7× smaller than that of a typical one with Si-based mirrors. A Lorentzian linewidth of 4 kHz was measured despite design and fabrication imperfection in the first run100.
We note that the detuned loading effect is a ‘double-edged sword’ capable of demolishing or amplifying the impact of the αH factor to cause subsequent linewidth reduction or broadening, depending on whether detuning of the lasing frequency with respect to the central frequency of the maximum mirror reflectivity is negative or positive. Thus, accurate and careful lasing frequency control is critical and challenging in practical applications with variable environmental factors. A trade-off between laser efficiency and lasing frequency detuning is also inevitable92. Another straightforward solution, particularly in heterogeneous integration, is to use a quantum dot (QD) gain medium instead. The αH factor of QD materials was experimentally and theoretically proved to be as low as zero or even negative106, 107. The αH factor in QD materials, which ranges from very small to zero, originates from their symmetrical gain spectrum108 and is strongly affected by dot uniformity and doping density107. A αH factor of approximately 1 was measured using an InAs/GaAs QD-on-Si heterogeneous laser109, 110, where the InAs/GaAs QD active region was not optimised towards a zero αH factor. A similar heterogeneous extended cavity laser based on a dual-coupled-ring reflector design and the same QD gain medium highlighted in the cross-sectional SEM image in Fig. 6d was fabricated recently for the first time105, 111. A minimal Lorentzian linewidth of 5.3 kHz was extracted from the measured white noise upper limit of 1.68 kHz2/Hz at approximately 1310 nm, indicating a nearly 10× improvement over previous O-band laser record, also from a heterogeneously integrated QW-on-Si laser112.
Fig. 7a shows the historical progress of Lorentzian linewidth reduction in diode lasers with different integration formats92. It is clear that progress in conventional III-V QW lasers started levelling off 10 years ago because of the fundamental properties of materials and design limits of external and solitary laser cavities on the III-V monolithic platform. However, in the past decade, a linewidth reduction of four orders of magnitude enabled by hybrid and heterogeneous platforms once again resulted in tremendous advancement because parameters such as Q, QE, and αH could be improved instantly and independently. As projected in Fig. 7b, the transition from the QW to the QD material system benefits from small or even zero αH dot gain medium with the largest and most straightforward impact without the complexity of accessory device design. Cavity loss reduction is of secondary significance, but is also effective, as the resulting linewidth reduction almost occurs at the same pace as loss reduction. A sub-100 Hz linewidth is expected to be within reach when low-αH QD materials are coupled with the design of distributed cavity loss as low as approximately 0.1 dB/cm. The effect of dilution on the αH factor by detuned loading is more potent for materials with large αH, that is, the QW gain medium. Diode lasers with such a narrow linewidth would be anticipated to have unprecedented impact in coherent communications and sensing, metrology113, microwave photonics114, etc.
-
In addition to frequency noise suppression or elimination, a low or zero αH factor in QDs largely raises the critical external feedback level for coherence collapse in diode lasers107, 115-117. Error-free laser operation under 100% external feedback has been shown in QD lasers118, 119, indicating the feasibility of removing bulky and expensive optical isolators in a laser or transmitter package120. Furthermore, 3D band gap structure in QDs provides a complete confinement for injected carriers as well as good crosstalk isolation among dots. This results in low transparency current density121, excellent optical gain thermal stability122, 123, low relative intensity noise (RIN)124, and large tolerance to material and fabrication defects125, 126. Owing to the inhomogeneous dot size distribution commonly seen in widely used Stranski-Krastanov growth method, wide spectral gain bandwidth127 is ideal for broad spectral lasing applications, for example, comb lasers or widely tunable lasers. All these favourable material properties encourage the development of heterogeneous QD comb lasers109 for DWDM communication128, 129, as the current CWDM4 standard with 20 nm channel spacing is becoming a bandwidth scalability bottleneck for post-400 Gb/s direct detection applications in data communications. Fig. 8a shows an FP laser built on this QD-on-Si platform and exhibits robust operation and impressively low cw threshold current density 670 A/cm2 at 100 °C (280 A/cm2, 7% wall-plug efficiency at 20 °C) 130. A minimal threshold current of 165 A/cm2 was recently measured. When adding a saturable absorber (SA) section to the laser cavity with appropriate bias for it and the gain section, four-wave mixing promotes stable mode-locked lasing. A clear multi-wavelength operation with extremely large 3, 6, and 10 dB comb widths of 12, 18, and 25 nm, respectively, as shown in Fig. 8b, is likely the result of the unique combination of spatial hole burning, group velocity dispersion, a linewidth enhancement factor, and four-wave mixing131. The enlarged spectrum in Fig. 8c reveals a channel spacing of 15.5 GHz, that is, the laser free spectral range (FSR), which is inversely proportional to the 2.6 mm-long laser cavity defined by photolithography. This channel spacing varies negligibly with changes in the device temperature. Within this cavity, a 1.4 mm long active section with an SA at the centre employed the same InAs/GaAs QD-on-Si configuration as that of the QD narrow-linewidth laser in Fig. 6d. Over 135 constantly spaced wavelengths within 3 dB power variation were realised from this single comb laser, which is only controlled by three terminals. A large channel spacing (e.g. 100 GHz) can be realised by a multiple-saturable absorber (SA) design or Vernier effect in an extended cavity132 rather than reducing the laser cavity excessively. Compared with the conventional approach to form a multi-wavelength source with a bank of single-wavelength lasers, the comb laser has a clear edge in terms of its footprint, control simplicity, and energy efficiency133. Inherently low RIN in the QD gain medium is critical to obtain high signal integrity in a full link, particularly for PAM4 and the more advanced modulation format. A 10-channel transceiver link based on a heterogeneous comb laser and SiGe avalanche photodetector was recently demonstrated to deliver a conservative aggregated 160 Gb/s error-free NRZ data transmission with an energy bill of 3 pJ/bit134. Current comb laser research focuses on further wall-plug efficiency enhancement and QD mode-locked laser model development. A similar InAs/GaAs QD gain medium has also been used to develop heterogeneous compact low-threshold microring lasers135, 136 and robust DFB lasers137. It is noted that despite numerous desirable material properties, the finite intra-band relaxation time and the gain saturation effect in QDs significantly limit the QD laser direct modulation bandwidth, resulting a bandwidth of only 5-10 GHz for 1310 nm InAs/GaAs QD lasers in general136. Optical injection locking was recently used to extend the bandwidth to more than 20 GHz to support 25 Gb/s NRZ modulation138, 139. The same heterogeneous platform can be expected to allow convenient master and slave co-integration, for example, using a single comb master laser to injection lock a series of microring slave lasers138.
Fig. 8
a Temperature-dependent LIV characteristic of a heterogeneous QD FP laser, b Optical spectrum of the comb laser. b optical spectrum of the comb laser over the full wavelength range, c enlargement of the range indicated in b to show individual lasing modes. The dashed line is drawn at −6 dB from the peak comb line131.Another interesting use case of such heterogeneous QD-on-Si integration is the discovery of extremely low dark current and avalanche gain when reversely biasing the same diode140, 141. A record low 1 × 10−6 A/cm2 dark current density at −1 V was measured in such a QD photodetector built on the same chip with comb lasers using an identical III-V epitaxial stack and fabrication process. Low crystal dislocation density among electrically isolated dots and well-passivated devices are considered to be the root causes. Because of the strong carrier confinement and small absorption volume, the responsivity is relatively low, for example, 0.1 A/W at −1 V bias for a 90 μm-long device with 1310 nm light as the input. However, a maximum avalanche gain of up to 350 for TM mode and 150 for TE was recently observed at a breakdown voltage of approximately −18 V. The corresponding gain-bandwidth-product is 300, also a record for QD avalanche photodetectors (APD) and comparable to state-of-the-art SiGe and Ge APDs142. Error-free operation at 20 Gb/s in NRZ mode was achieved on such a heterogeneous QD APD, indicating the potential to play the role of a high-speed receiver as well as a monitor PD.
-
Superior active functionalities in III-V materials used complementary with exceptional low-loss passive components in Si, SiNx, or SiO2 have proved to be a success formula in all the cases that have been discussed thus far, but are not the only success. Researchers are actively exploring the nonlinear properties of III-V thin films to enable high-quality heterogeneous passive III-V components on Si to be developed at the present time, and active-passive integration in the near future. The nonlinear coefficients in most III-V materials are usually several orders of magnitude larger than those of popular dielectrics such as SiNx and SiO2143. For example, the nonlinear Kerr index n2 of SiO2, SiNx, GaP and AlGaAs is 3 × 10−20, 2.5 × 10−19, 1.1 × 10−17 and 2.6 × 10−17 m2/W, respectively. The mode volume in an HIC III-V-on-insulator waveguide structure is also much smaller than that of its dielectric counterparts. Both factors contribute to a significantly lower parametric oscillation threshold power (Pth) to trigger nonlinear effects, as governed by Eq. 3144, 145:
$$ P_{t h} \approx 1.54 \frac{\pi n w A}{2 \eta n_{2} D_{1} Q_{T}^{2}} $$ (3) where n is the modal index, w is the angular frequency, A is the modal area, η is the coupling factor equivalent to the proportion of (external) coupling loss of the total (internal + external) resonator cavity loss, and QT is the total quality factor of the resonator. D1 equals c/Dng, a function of the resonator diameter D, the speed of light in vacuum c, and the group index ng. A smaller A, much larger n2 and smaller D all favour a III-V resonator despite the higher QT in much larger SiO2 and SiNx resonators70, 146. Furthermore, the third-order nonlinear effect on top of the second-order one in III-V materials enhances its attractiveness for frequency comb generation147. Therefore, AlGaAs/GaAs and GaP emerge as promising III-V thin-film candidates148, 149.
As shown in the inset of Fig. 9a, a 700 mm × 400 nm Al0.2Ga0.8As strip waveguide was fabricated upon transferring a layer of undoped Al0.2Ga0.8As on a SiO2/Si substrate via molecular bonding150, 151. The small mode volume (~0.28 μm2) is 4 × smaller than that of the SiNx counterpart, indicating a lower photon energy and more compact footprint for comb generation. Special care was taken to smooth out line-edge roughness in the photoresist pattern, optimise the III-V dry etch process, and passivate defect states on the III-V surface with 5 nm-thick Al2O3 to achieve an intrinsic quality factor of 1.53 × 106 in a microring resonator with a diameter of 12 μm (FSR=1 THz) (Fig. 9b inset) at approximately 1519 nm25, 150. This corresponds to an extremely low propagation loss of 0.4 dB/cm, 10 × smaller than that of typical III-V waveguides152, 153 and at least 3 × smaller than many reported loss values for Si waveguides of similar dimensions. Lattice-matched AlxGa1-xAs on the GaAs substrate provides the freedom to adjust the concentration of Al ‘x’, i.e. the band gap, for 1) a minimal two photon absorption loss and 2) anomalous group velocity dispersion (GVD)154 at the wavelength of interest, i.e. the C-band in this case. The onset of the frequency comb generation can be observed in the parametric oscillation spectrum of such a 1 THz resonator in Fig. 9a under a low pump power of 36 μW at the C-band. To the best of our knowledge, this represents record-low threshold pump power for comb generation among numerous nonlinear material platforms, for example, a reduction of approximately 100 × relative to previously reported AlGaAs-on-insulator resonators with similar FSR148, and an improvement of 10× compared with their integrated dielectric counterparts150, 154, 155. The higher nonlinear efficiency owing to the high n2 in this case is reflected by the >250 nm-wide spectral range under a pump power of 300 μW, as shown in Fig. 9b, which is also lower than the previous record held by a SiNx micro-comb70. A soliton-step transition has also been observed as a result of the small thermal impact of efficient pumping150. Outstanding performance in frequency conversion was also achieved in a low-loss GaAs-on-insulator waveguide for the same reasons discussed above151. Exceptional normalized second-harmonic efficiency of 13,000% W−1cm−2 at a fundamental wavelength of 2 μm was achieved by harnessing the advantages of a materials platform and an improved fabrication process25, 150.
Fig. 9
Measured spectra of 1 THz resonator under pump power of a 36 μW and b 300 μW. Inset. a cross-sectional and b side view SEM images of an AlGaAs resonator150.On-going efforts continue to focus on driving the waveguide loss down to further enhance the quality factor of III-V nano-waveguides. This will be a natural by-product once this platform is scaled up to 200 and 300 mm and processed in a commercial foundry with higher-resolution photolithography and tighter process control. Compared with the SiNx and SiO2 high-Q resonators, we note that another huge potential advantage inherent to high-Q III-V-on-insulator nonlinear components is the compatibility of their materials and process simplicity for many of the heterogeneous active devices discussed before. For example, an InAs/GaAs QD pump laser could be integrated with a high-Q AlGaAs-on-insulator resonator for potentially even lower pump thresholds, and this integration becomes an alternative multiple wavelength source apart from comb lasers. Alternatively, a more sophisticated integration could include tunable narrow-linewidth QD lasers, another high-efficiency QD pump laser, and APD that could be readily integrated with a high-Q AlGaAs-on-insulator resonator to form a fully integrated optical-frequency synthesiser based on the proof-of-concept demonstration using discrete parts113 but on a millimetre-scale chip. Briefly, the superior device performance and large potential for improvement discussed in this section are turning nonlinear optics into another irresistible force to drive heterogeneous Si photonics beyond the traditional optical commutations, and into applications of high-speed optical signal processing, metrology, and quantum communication and computation.
-
The last representative heterogeneous component, but definitely not the least, is a novel III-V/dielectric/Si MOS capacitor for ultra-efficient phase tuning and high-speed modulation156-159. As shown in the TEM cross-sectional image in the inset of Fig. 10a, this heterogeneous capacitor can be conveniently formed by sandwiching a layer of high-quality thin dielectric material, that is, gate oxides such as SiO2 or Al2O3, or even polymer, at the III-V/Si bonding interface during molecular or polymer bonding. When this capacitor is charged or discharged by applying an electrical field between the III-V material and Si, the carrier concentration around the gate oxide can change by several orders of magnitude to exceed 1020/cm244. By designing a ‘staircase’-shaped waveguide of which the optical mode overlaps with this capacitor (Fig. 10b inset), it would be possible to introduce the plasma dispersion effect to manipulate the real and imaginary parts of the modal index based on the Drude model6.
Fig. 10
a Measured spectra of a heterogeneous MOS microring resonator under different MOS bias. Insets: schematic device structure and a cross-sectional TEM image showing MOS capacitor region162. b Schematic device structure and cross-section of a heterogeneous MZI modulator. Insets: simulated mode profile at different waveguide stages and NRZ modulation eye diagram157. c Simplified circuit model and cross-sectional SEM image of a heterogeneous MOS capacitor laser, and d its spectra under different MOS voltages156.Prior to application to heterogeneous MOS devices, the same concept was employed on a pure Si platform to realise the first Gb/s Si modulators160. Because this capacitor (dis)charge process occurs under a fast electric field effect, rapid engineering progress soon led to the development of a decent Si MOS modulator capable of performing 40 Gb/s NRZ modulation161. However, compared with the pure Si design where deposited poly-Si is typically used as the top layer above the gate oxide, a wafer-bonded III-V layer offers three advantages with respect to its material characteristics. Firstly, crystalline III-V material has lower optical loss than poly-Si, even though it is highly doped. Secondly, the much smaller conductivity effective mass of electrons and larger electron mobility in a III-V material results in a larger (smaller) real (imaginary) part change in the modal index, which translates into larger phase change efficiency and lower free carrier absorption, respectively. Finally, the band-filling effect in a III-V material contributes to the phase change in addition to the plasma dispersion effect158. Furthermore, wafer bonding greatly simplifies fabrication by obviating the need for chemical mechanical polishing (CMP), which is typically required before poly-Si deposition in pure Si MOS waveguide formation.
Fig. 10a shows a heterogeneous InP/oxide/Si MOS microring resonator 40 μm in diameter162. The resonator consists of a 12 nm-thick gate oxide layer containing a combination of HfO2, Al2O3, and SiO2 sandwiched between 250 nm-thick p-Si and 150 nm-thick InP layers. Here, high-k HfO2 was selected to enhance the capacitance for a large plasma dispersion effect, and Al2O3 and SiO2 were used to form good interfaces with InP and Si, respectively. A resonance wavelength blue shift in the C-band in excess of 1 nm was obtained by applying bias of 4 V across this capacitor, translating to a tuning pace of 0.26 nm/V and a phase-shift/modulation efficiency VπL of 0.12 V-cm. Because of the extremely low DC leakage current, typically in the fA range44, a record tuning energy efficiency of 5.3 nm/pW was calculated. This is an improvement of nine orders of magnitude over conventional thermal and carrier injection tuning with a few or sub-nm/mW efficiency. The decreasing extinction ratio in the spectra when increasing bias is applied to the MOS capacitor is due to simultaneously rising FCA, which forces the resonator to drift away from the critical coupling condition at a bias of approximately 0 V. Another reported heterogeneous MOS MZI structure exhibited a VπL of 0.047 V cm and 0.23 dB attenuation across the 500 μm-long MOS phase shifter that are 5 and 10× lower than those of a Si MOS modulator, respectively158. This is attributable to the use of thin Al2O3 gate oxide with 5 nm equivalent oxide thicknesses and n-InGaAsP rather than n-InP for a larger (smaller) electron-induced refractive index (FCA) 158. Furthermore, MOS capacitor-enabled phase tuning is also an athermal and much faster process than the two popular methods, making it a more effective phase-tuning mechanism for negligible power consumption, zero thermal crosstalk, higher phase shift efficiency, lower insertion loss, and faster response.
Fig. 10b shows a schematic of one arm of a MOS MZI modulator157. A SiO2 gate oxide layer with a thickness of 10 nm was used to ensure a good trade-off between a small VπL (0.09 V-cm) and large RC time limit for high-speed modulation. Appropriately designed taper structures ensured low-loss transitions and minimal reflection at different stages of light propagation, i.e. an input/output Si strip waveguide, InGaAsP taper, InGaAsP/Si taper and InGaAsP/Si MOS capacitor, and an MOS capacitor phase shifter. Their corresponding simulated fundamental TE mode profiles are displayed in the insets in Fig. 10b. Although the RC-limited 3 dB bandwidth was below 1.5 GHz because of imperfections in the design and fabrication, NRZ signal modulation as high as 32 Gb/s and an extinction ratio over 3 dB in the C-band were achieved with the help of the pre-emphasis drive signal and maximum Vpp of 3.5 V. A much larger bandwidth (as large as 30 GHz in the O-band) was demonstrated in another MZI design by reducing the RC constant with lower series resistance and doubling the thickness of the SiO2 gate oxide layer (20 nm) at the cost of reduced modulation efficiency (VπL = 1.3 V-cm)159. NRZ signal modulation as high as 25 Gb/s was measured by applying a Vpp of 4 V without using a pre-emphasis driving signal. Improved design recently paved the way to relieve the trade-off between modulation efficiency and the RC limit, the primary speed limit for MOS-type modulators163. Excellent temperature insensitivity was also confirmed in these MOS capacitive modulators164. By the time of manuscript preparation, we also developed the first heterogeneous GaAs-on-Si MOS microring modulator with promising NRZ modulation in excess of 25 Gb/s without pre-emphasis in limited preliminary testing165.
All desirable MOS capacitor-enabled functionalities can be easily integrated in heterogeneous III-V-on-Si lasers, which then become a three-terminal configuration as shown in Fig. 10c, which shows a simplified circuit model and a cross-sectional SEM image of the device156. The terminals P1 on p-InGaAs and N on n-InP are the normal diode laser anode and cathode, respectively, to inject carriers into the active region of the MQW. The new terminal P2 on p-Si and the terminal N allow a voltage bias to be applied to the n-InP/Al2O3 (15 nm)/p-Si capacitor. A similar blue-shift in the wavelength as a function of the MOS voltage across the capacitor proves that it is a result of the plasma dispersion effect when the current injected into the MQW is held constant, as shown in Fig. 10d. On the other hand, increasing the FCA loss with higher MOS voltage leads to higher internal cavity loss and, subsequently, a higher laser threshold and lower efficiency. The laser output power was observed to decrease by more than 10 dB when the MOS bias is varied from 0 to 5 V, enabling high-speed frequency modulation and amplitude modulation. A 3 dB MOS modulation bandwidth of 14.5 GHz was measured. This is more than 2× larger than the conventionally preferred direct current modulation bandwidth (6.6 GHz) to the same device with the same DC injection current166. This is because MOS modulation perturbs the internal cavity loss, that is, the photon lifetime, which is a much faster process than injection carrier perturbation167.
This MOS capacitor structure is a direct and unique product of wafer bonding-enabled heterogeneous integration, and is impossible to replicate on a hybrid integration platform. In addition to the excellent optical, mechanical, and thermal properties of Si, this novel heterogeneous MOS capacitor activates the desired and well-studied electrical properties of Si on this heterogeneous platform. A fully integrated DWDM transceiver on Si is being developed to include previously discussed QD comb lasers, MOS microring modulators and drop filters, and QD APDs. The straightforward integration of MOS capacitors in lasers, modulators, and filters can be expected to enable agile wavelength tuning and locking in each channel to correct the laser/microring resonance frequency mismatch due to fabrication imperfection and temperature fluctuation. A single QD epitaxial transfer and a concurrent process to simultaneously fabricate all these key components would also greatly simplify the design and fabrication in comparison with designs that use different III-V epitaxial stacks for different building blocks168.