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

High-speed GaN-based laser diode with modulation bandwidth exceeding 5 GHz for 20 Gbps visible light communication

Open Access Open Access

Abstract

Visible light communication (VLC) based on laser diodes demonstrates great potential for high data rate maritime, terrestrial, and aerial wireless data links. Here, we design and fabricate high-speed blue laser diodes (LDs) grown on c-plane gallium nitride (GaN) substrate. This was achieved through active region design and miniaturization toward a narrow ridge waveguide, short cavity length, and single longitudinal mode Fabry–Perot laser diode. The fabricated mini-LD has a low threshold current of 31 mA and slope efficiency of 1.02 W/A. A record modulation bandwidth of 5.9 GHz (3dB) was measured from the mini-LD. Using the developed mini-LD as a transmitter, the VLC link exhibits a high data transmission rate of 20.06 Gbps adopting the bit and power loading discrete multitone (DMT) modulation technique. The corresponding bit error rate is 0.003, satisfying the forward error correction standard. The demonstrated GaN-based mini-LD has significantly enhanced data transmission rates, paving the path for energy-efficient VLC systems and integrated photonics in the visible regime.

© 2024 Chinese Laser Press

1. INTRODUCTION

Visible light communication (VLC) is an optical wireless communication technology whereby data is transmitted using visible light with wavelengths between 375 and 780 nm [1,2]. The benefits of the plentiful spectrum (400–800 THz) have attracted increasing interest in VLC research worldwide [15]. VLC can be considered as an essential component of next-generation 6G wireless networks [6]. VLC technology is characterized by electromagnetic immunity and high security and is license-free [7,8]. These characteristics are ideal for certain settings, including stadiums, hospitals, airplanes, and laboratories [7,9]. Therefore, VLC plays an important role in local networks where it is used to address the foreseeable limited radio frequency (RF) spectrum resources [10]. Due to its unique light penetration characteristics underwater, VLC not only facilitates underwater communication but also integrates seamlessly with wireless communication systems in space, enabling a unified communication network across both aquatic and aerial environments. To date, light-emitting diode (LED)-based VLC systems have demonstrated significant potential and benefits in several applications [1115]. For instance, a record 7.91 Gbps data transmission rate based on a micro-LED was accomplished [16]. However, the rapid increase in the requirement for high data transmission rates has resulted in a greater demand for VLC systems [17]. Because the carrier lifetime of an LED is of the order of nanoseconds, the bandwidth of the LED always limits the data rate of an LED-based VLC system [18,19]. Compared with LED, the modulation response of a laser diode (LD) is controlled by the photon lifetime, which is at the picosecond level. Consequently, the LD exhibits a higher modulation bandwidth, which is more suitable for high-speed VLC [2023]. Moreover, the LD exhibits improved optical directionality and optical spectrum purity, both of which help to extend the transmission distance in VLC systems.

High-speed gallium nitride (GaN)-based blue LDs are key devices for VLC. A GaN-based LD was developed after Nakamura demonstrated the first violet laser in 1996 [24]. With the development of the GaN growth technique and improvement of the p-doping level, GaN-based LDs have demonstrated suitability for applications including optical storage [25], bio-medical treatment [26], and high brightness lighting [27,28]. However, the design and fabrication of high-speed LDs for VLC are yet to be thoroughly investigated, and the modulation bandwidth for visible LDs is still far below that of InP-based and GaAs-based NIR LDs [29,30]. The progress of research into high-speed GaN-based LDs is presented in Table 1. The modulation bandwidth of the c-plane LD was still limited to 3 GHz, which did not satisfy the fast-growing demands for higher communication data rates. Low slope efficiency and high threshold current density remain issues that are to be resolved. Although an LD grown on a semipolar GaN substrate has a modulation bandwidth of over 5 GHz [32], problems such as costly substrate and immature material growth process remain. Therefore, it is critical to design and fabricate high-speed GaN-based blue LDs on c-plane substrates.

Tables Icon

Table 1. Summary of the Key Parameters for High-Speed Blue GaN LDsa

Various approaches can be used to achieve high-speed LDs. Both active region and device structure design are critical for the improvement of modulation bandwidth in GaN-based LDs. For example, adopting quantum-dots (QDs) can enhance the electron–photon interactions as well as the coupling between the active region and the optical field, resulting in a reduced threshold current and increased differential gain [35,36]. Those features contribute to a higher resonance frequency. However, the low conversion efficiency and limited output optical power at high current densities hinder the application of GaN-based QD lasers [37]. The introduction of a Bragg grating along the ridge waveguide to develop a single-mode GaN-based distributed-feedback (DFB) LD helps improve mode stability and modulation response [38]. However, the complicated fabrication process hinders the realization of high-performance DFB gratings at blue color regime. Adopting semipolar GaN quantum wells (QWs) can effectively suppress the quantum confinement Stark effect (QCSE) and improve the recombination efficiency, but the growth of such high-quality epitaxial layers on c-plane substrates remains challenging [39]. Since the miniaturization of transistors and LEDs has yielded a significant improvement in device performance, the study of small form-factor GaN-based LDs has become another promising approach toward high-speed operation. Studies reveal that the structure of the ridge width and cavity length affect the dynamic properties and high-speed performance of NIR LDs [40]. Therefore, the dimensional scaling down in GaN-based LDs becomes an important topic to investigate.

In this study, we design and fabricate a blue GaN mini-LD with reduced cavity length from millimeter level to sub-millimeter (500 μm) and a narrow ridge width of 1.8 μm. The design principles of GaN mini-LD epi-structures concerning the modulation response are also investigated. The use of multilayer QW/quantum barriers (QBs) can improve the relaxation resonance frequency by increasing differential gain yet increasing the threshold current [32]. Our mini-LD features a double-QW active region with optimized QW/QB thickness. The total effective active region of the fabricated mini-LD is less than 0.001mm2. The threshold current and the corresponding current density are 31 mA and 3.4kA/cm2, respectively. The slope efficiency of the single-mode mini-LD reaches 1.02 W/A. The 3dB modulation bandwidth is measured to be 5.9 GHz. Adopting this mini-LD in the VLC system allowed us to obtain a data transmission rate exceeding 20 Gbps. To the best of our knowledge, this is the highest data rate reported for a blue LD-based VLC system.

2. DEVICE DESIGN AND FABRICATION

To design a high-speed GaN-based LD, a dynamic response model is established to analyze the effect of the active region on the bandwidth. The transfer function can be extracted from the carrier-photon rate formula [41], which is expressed as [42]

H(f)=fR2fR2f2+jγf/2π.

Here, fR is the relaxation resonance frequency, γ=KfR2+γ0 is the damping coefficient, and K is a constant whose value indicates the damping effect. Parameter fR mainly depends on the differential gain, optical field volume, and injection efficiency at a given injection current level.

The optical and electronic properties of the LDs with different quantum well/barrier thicknesses were simulated using PICS3D. In the simulation, the spontaneous emission coefficient is set as 1×104, and the screening coefficient is 0.35. The additional nonlinear gain suppression is not considered. Figure 1(a) shows LD with the 5 nm barrier is featured with the highest fR response and low γ0 level when the quantum well thickness is 2 nm. Meanwhile, low-threshold currents appeared in structures with smaller barrier thicknesses, leading to a higher fR. Figure 1(b) shows that the device with a 3 nm/5 nm active region exhibited a higher differential gain and 3dB bandwidth. Note that the comparison at a relatively low current is fair, avoiding the self-heating and saturation of fR. Figure 1(c) suggests that the LD has an optical confinement factor of 0.0152. The simulated 3dB bandwidth can reach 6GHz at 140 mA, as shown in Fig. 1(d).

 figure: Fig. 1.

Fig. 1. Design and simulation of the high-speed GaN mini-LD (with a ridge width of 1.8 μm and cavity length 500 μm). (a) Simulated dynamic parameters dfR/d(IIth)1/2 curve slope efficiency (red) and optical confinement factor (blue) of the devices with various quantum barrier thickness and 2 nm quantum well. (b) Histogram of simulated differential gain (red) and −3 dB bandwidth for an injection current of 140 mA, various quantum well thicknesses, and a 5 nm quantum barrier. (c) Simulated light field profile of quasi-TE fundamental mode in LD with 3 nm/5 nm active region. A ridge of 1.8 μm width is visualized in the white frame. Γ is the optical confinement factor. (d) Simulated frequency response under injection currents ranging from 110 to 140 mA of a 3 nm/5 nm active region structure. The dashed line in the figure is the 3dB line related to the measurement start point (100 MHz). Damping behavior at the low frequency marked by the blue dashed circle is related to the RC (resistance-capacitance) roll-off.

Download Full Size | PDF

Based on the simulation results, we design and fabricate the III-nitride mini-LDs. The devices were grown on a c-plane (0001) GaN substrate using the metal-organic chemical vapor deposition (MOCVD) technique. The epitaxial structure consists of a 3 μm Si-doped n-GaN contact layer ([Si]=2×1018cm3), a 1000 nm n-Al0.07Ga0.93N cladding layer ([Si]=3×1018cm3), and a 150 nm n-In0.03Ga0.97N waveguide layer ([Si]=2×1018cm3). The active region consists of two 3 nm In0.21Ga0.79N QWs separated by three 5 nm GaN quantum barriers, followed by a p-In0.03Ga0.97N waveguide layer. Then, an 8 nm Al0.18Ga0.82N Mg-doped electron blocking layer (EBL) is grown on the topside ([Mg]=5×1018cm3) followed by a 300 nm p-Al0.07Ga0.93N cladding layer. This is then followed by a 100 nm highly doped p+-GaN contact layer with a doping concentration of 1×1019cm3 optimized to reduce the series resistance ([Mg]=1×1019cm3). Transmission electron microscopy (TEM) images of the active region and energy-dispersive X-ray spectroscopy (EDS) mapping of the active region, including In and Al, are shown in Figs. 2(f) and 2(g), respectively. The epitaxial layer was etched into a ridge waveguide structure using lithography and an inductively coupled plasma (ICP) process. To achieve passivation and electrical isolation, 200 nm SiO2 was sputtered on the wafer. The epitaxial layers between the n-contact and p-contact layers were covered by the SiO2 layer from bottom to top. A self-alignment process was adopted to selectively remove the SiO2 layer from the ridge. Pd/Ti/Pt/Ti/Au and Ti/Pt/Ti/Au were deposited as the p- and n-GaN metal electrodes, respectively. The LD resonance cavity was formed by mechanical cleavage. Facet coatings with pairs of SiO2/TiO2 were deposited on the front and rear facets to achieve reflectivities of 70% and 99%, respectively.

 figure: Fig. 2.

Fig. 2. Macroscopic and microscopic structures of the laser. (a) 3D illustration of the fabricated laser. The annotation on the right shows the epitaxial layer structure of the active region from top to bottom. (b) Far-field emission pattern of the laser. (c) Optical microscopy image of the fabricated laser. The n- and p-electrodes are marked in the picture. (d) Scanning electron microscope (SEM) image of cross-sectional view. The ridge waveguide width is 1.8μm. (e) STEM image of the active region. From top to bottom are the p-cladding layer, electron blocking layer (EBL), upper waveguide, and MQWs. (f) Indium mapping of the active region. The two high-brightness lines (marked with the yellow triangle) are quantum wells separated by a quantum barrier. (g) Aluminum mapping of active region. The line with high brightness (marked with the yellow triangle) is the EBL (electron blocking layer).

Download Full Size | PDF

3. EXPERIMENTAL RESULTS AND DISCUSSIONS

The light–current–voltage (L–I–V) characteristics of the fabricated mini-LDs are shown in Fig. 3(a). The threshold current density of the LD was 3.4kA/cm2 and the slope efficiency was 1.02 W/A. Continuous (CW) measurements were performed using a Keithley 2450 source measure unit (SMU) and a Newport 2936-R with a calibrated 818-SL Si detector as the power meter. Compared to other reports on GaN laser diodes for VLC, our mini-LD shows a lower threshold current density and higher slope efficiency [17,31,34,43]. The electrical luminescence spectra of the blue LD were collected using a high-resolution optical spectrum analyzer (Advantest Q8347), and the results are shown in Fig. 3(b). The measurement was performed at room temperature with a pulsed injection current (PL202 sourcemeter). The pulse width is 200 μs and the duty cycle is set at 1% to eliminate the effects of self-heating. The peak wavelength was 451 nm for injection currents in the range of 60–140 mA. The peak wavelength of the chip remained constant during the testing and a single longitudinal mode was observed from the fabricated mini-LD.

 figure: Fig. 3.

Fig. 3. (a) Light–current–voltage (L–I–V) characteristic of the laser under the condition of continuous wave (CW) injection. (b) Spectra of the mini-LD for injection currents ranging from 60 to 140 mA at room temperature.

Download Full Size | PDF

The small-signal modulation characteristics of mini-LDs were investigated experimentally. Prior to the S21 test, we adopted the short-load-open-through (SLOT) method using an MPI AC2 calibration board to calibrate the entire experimental setup to the tips of the GS probe (MPI TITAN-T26-GS150). Sinusoidal small-signal modulation was performed using an Agilent N5230C PNA-L network analyzer. A 10 GHz photodetector (PD, Newport 818-BB-45A) was used as the receiver.

The measured small-signal modulation response of the entire system S21M can be expressed as

S21M=S21LR.

Here, S21L and R refer to the electrical-optical and optical-electrical modulation responses of the laser and PD, respectively. Since the 3dB modulation bandwidth of the PD was 10 GHz and above, we considered R to be constant. Therefore, S21M revealed the modulation characteristics of the laser when S21M<10GHz. The S21L for different injection currents are shown in Fig. 4(a). To extract the intrinsic S21 response from the original response, we established an equivalent circuit model of laser diodes as follows:

η=Zr2R(1+S22P)S21PS21M.

η is the current modulation response coefficient of the intrinsic laser, Zr is the specific impedance, and S22P and S21P are the scattering parameters of parasitic part in cascaded networks. The component values of the circuit can be obtained from the S11 results, as shown in Fig. 4(b). The intrinsic modulation characteristics of the LD through the de-embedding of the parasitic network response are shown in Fig. 4(c). The results showed that the 3dB bandwidth of the laser chips increased with the injection current from 100 to 140 mA. Further increasing the injection current led to a significant degradation in the modulation characteristics, which was caused by a strong nonlinear gain with a self-heating effect. The highest modulation bandwidth achieved was 5.90 GHz. To the best of our knowledge, this is the highest 3dB modulation bandwidth for c-GaN laser diodes. The plot of extracted 3 and 10dB modulation bandwidth versus injection current is shown in Fig. 4(d). The results revealed that the nonlinear effect was not significant at 140 mA.

 figure: Fig. 4.

Fig. 4. (a) Measured frequency response of the fabricated mini-LD for injection currents ranging from 100 to 140 mA. (b) S11 response for injection currents ranging from 100 to 140 mA. (c) Extracted intrinsic S21 response for injection currents ranging from 100 to 140 mA. The 3 and 10dB are labelled in the figure using dashed lines. (d) Extracted intrinsic modulation bandwidth (3 and 10dB) versus injection currents ranging from 70 to 140 mA. (e) Relationship between resonance frequency (fR) and (IIth)1/2, where I is the injection current and Ith is the threshold current. The line in this figure is the fitting curve. (f) Relationship between square of resonance frequency (fR) and the damping factor. The line in this figure is the fitting curve. K is the slope of the curve, and γ0 is the intercept on the Y axis.

Download Full Size | PDF

We further explore the relationship between the resonance frequency and the injection current:

fR=12π[ΓυgdgdnqVηi(IIth)]1/2.

To characterize the differential gain of the laser, we investigated the relationship between the resonance frequency and (IIth)1/2. In Eq. (4), fR is the resonance frequency, I is the injection current, Ith is the threshold current, Γ is the confinement factor, υg is the group velocity, dgdn is the differential gain, q is the elemental charge, V is the volume of the active region, and ηi is the injection efficiency. The square root of the driving current above the threshold is linearly proportional to the resonance frequency (fR). For the fabricated LD, the dfR/d(IIth)1/2 is obtained as 0.32GHz/mA1/2 as seen in Fig. 4(e), which was relatively high and close to the simulation results. The extracted damping factors of the LD are shown in Fig. 4(f).

The relationship between the damping factor and the resonance frequency can be expressed as

γ=KfR2+γ0,
where γ is the damping factor corresponding to the injection current, fR is the resonance frequency, and γ0 is the damping factor offset. According to Eq. (5), the relationship between the square of fR and γ is linear. By adopting a linear fitting of the curve, the K factor was determined to be 0.395 ns. The intercept γ0 is 0.20 GHz. Compared to other reports, the K factor of the fabricated mini-LD was comparable to that of the NIR laser. The damping factor offset γ0 is far below the GaN LD grown on a semipolar substrate and is equivalent to that of the NIR laser [44]. The high-speed blue LD has a small damping factor for currents below 140 mA, which indicates excellent performance in dynamic modulation.

To characterize the communication performance of the LD, an experimental VLC system was established, as shown in Fig. 5. Figure 5(a) shows a schematic of the transmitter end, where discrete multitone (DMT) modulation with a bit-power loading scheme is employed [45,46]. The generated modulated signal [x(t)] can be expressed as

x(t)=k=0N11NP(k)X(k)ej2πNkt,
where N is the product of the number of subcarriers and the upsampling factor, P(k) is a coefficient related to the allocated power, and X(k) represents the data adopted for the kth carrier. To satisfy the conjugate symmetry to achieve Hermitian symmetry, there are a total of 1024 subcarriers including 512 subcarriers containing the signal. To avoid the influence of the bias tee, the first eight subcarriers are set as zero padding. The signal-to-noise ratio (SNR) of each subcarrier is calculated using quadratic phase-shift keying (QPSK). The allocated power ratio is determined using the measured SNR. To balance the frequency-domain response, a digital pre-equalization process is adopted in the system. This is expressed as follows:
x(t)=F1{F[x(t)]·Hpre},
where x(t) is the signal after pre-equalization, and F and F1 are fast Fourier transform (FFT) and inverse fast Fourier transform (IFFT), respectively.
 figure: Fig. 5.

Fig. 5. (a) Transmitter in the system using the discrete multitone (DMT) bit-power loading modulation method. The digital signal is processed through bit-power loading, DMT modulation, and digital pre-equalization. AWG is the arbitrary waveform generator. EA is the electronic amplifier. The generated signal is combined with the direct current (D.C.) of the bias and then transmitted to the mini-LD. (b) Experimental setup of the high-bandwidth mini-LD-based high-speed VLC system with DMT bit-power loading modulation. The mini-LD is equipped with a heat sink with the TEC to ensure constant temperature (20°C) and the transmitted signal is collected by the PD. (c) The receiver in the system which contains the post-processing of the signal. PD is the photodetector, OSC is the oscilloscope, LMS Volterra is the least mean square Volterra nonlinear filter, and ZF Equ. is the zero-forcing equalizer. After the post-processing, we obtained the transmitted signal.

Download Full Size | PDF

The transmitted signal is generated using an arbitrary waveform generator. The signal is then amplified by an electronic amplifier (EA) and sent to the bias tee, where direct current is coupled with the amplified signal. Finally, the merged signal is directly imposed on the laser. Figure 5(b) shows an image of the communication test system. The laser chip is fixed to a heat sink equipped with a thermoelectric cooler (TEC). Light is collected through the objective and delivered to the PD. The lower-right corner of Fig. 5(c) shows the system receiver. The transmitted light signal is converted to an electrical signal by the PD and transmitted to an oscilloscope (OSC). The bit and power loading signals received from the oscilloscope are processed by post-equalization and demodulation of the DMT model. A 4 GHz bandwidth is used to realize a high data rate of 20.06 Gbps. The measured SNR and allocated bits in each subcarrier channel are shown in Fig. 6(a). The obtained constellation diagram in Fig. 6(b) shows a clear profile. The transmitted and received partial symbols exhibit a high degree of overlap. The measured bit error ratio (BER) is 0.0030, which satisfies the threshold for forward error correction (FEC) coding (0.0038), as illustrated in Fig. 6(c).

 figure: Fig. 6.

Fig. 6. Experimental results of the VLC system using fabricated mini-LD. (a) Bit-power loading scheme of each subcarrier according to measured channel SNR. The effective carrier number is 504. (b) Received constellation diagrams of the system, including 128 QAM, 64 QAM, 32 QAM, 16 QAM, 8 QAM, 4 QAM, and 2 QAM. (c) Partial T/R symbol of the mini-LD-based VLC data links. The tested BER is 0.0030.

Download Full Size | PDF

4. CONCLUSIONS

In this study, we demonstrated a c-plane GaN-based blue mini-LD with a high 3dB modulation bandwidth of 5.9 GHz. The fabricated laser chip is characterized by a relatively low threshold current density of 3.4kA/cm2 at room temperature and a relatively high slope efficiency of 1.02 W/A. The device shows a K factor of 0.395 ns and γ0 of 0.20 GHz. Using this mini-LD as the transmitter, we developed a blue-laser-based VLC system that demonstrated a record data rate of 20.06 Gbps. The system currently utilizes only 4 GHz of the mini-LD’s bandwidth owing to the hardware limitations of our testing system. Exploring avenues for achieving higher data rates remains an area for further work. Nonetheless, our work reveals that this high-performance GaN blue mini-LD has great potential for applications in next-generation 6G wireless optical communication systems as well as photonic integrated chips in the visible color regime.

Funding

National Key Research and Development Program of China (2022YFB2802803); National Natural Science Foundation of China (61925104, 62031011, 62274042); Natural Science Foundation of Shanghai Municipality (21ZR1406200); China Mobile Research Institute X-NET; Jiangsu Provincial Key Research and Development Program (BE2021008-5).

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. N. Chi, H. Haas, M. Kavhrad, et al., “Visible light communications: demand factors, benefits and opportunities [Guest Editorial],” IEEE Wireless Commun. 22, 5–7 (2015). [CrossRef]  

2. L. E. M. Matheus, A. B. Vieira, L. F. M. Vieira, et al., “Visible light communication: concepts, applications and challenges,” IEEE Commun. Surv. Tutorials 21, 3204–3237 (2019). [CrossRef]  

3. P. H. Pathak, X. Feng, P. Hu, et al., “Visible light communication, networking, and sensing: a survey, potential and challenges,” IEEE Commun. Surv. Tutorials 17, 2047–2077 (2015). [CrossRef]  

4. A. Jovicic, J. Li, and T. Richardson, “Visible light communication: opportunities, challenges and the path to market,” IEEE Commun. Mag. 51(12), 26–32 (2013). [CrossRef]  

5. L. Grobe, A. Paraskevopolous, J. Hilt, et al., “High-speed visible light communication systems,” IEEE Commun. Mag. 51, 60–66 (2013). [CrossRef]  

6. N. Chi, Y. Zhou, Y. Wei, et al., “Visible light communication in 6G: advances, challenges, and prospects,” IEEE Veh. Technol. Mag. 15(4), 93–102 (2020). [CrossRef]  

7. Y.-F. Huang, Y.-C. Chi, H.-Y. Kao, et al., “Blue laser diode based free-space optical data transmission elevated to 18 Gbps over 16 m,” Sci. Rep. 7, 10478 (2017). [CrossRef]  

8. Y. Zhou, X. Zhu, F. Hu, et al., “Common-anode LED on a Si substrate for beyond 15 Gbit/s underwater visible light communication,” Photonics Res. 7, 1019–1029 (2019). [CrossRef]  

9. Y.-C. Chi, D.-H. Hsieh, C.-Y. Lin, et al., “Phosphorous diffuser diverged blue laser diode for indoor lighting and communication,” Sci. Rep. 5, 18690 (2015). [CrossRef]  

10. J. Hu, F. Hu, J. Jia, et al., “46.4 Gbps visible light communication system utilizing a compact tricolor laser transmitter,” Opt. Express 30, 4365–4373 (2022). [CrossRef]  

11. G. Cossu, A. M. Khalid, P. Choudhury, et al., “3.4 Gbit/s visible optical wireless transmission based on RGB LED,” Opt. Express 20, B501–B506 (2012). [CrossRef]  

12. H. Le Minh, D. O’Brien, G. Faulkner, et al., “100-Mb/s NRZ visible light communications using a postequalized white LED,” IEEE Photonics Technol. Lett. 21, 1063–1065 (2009). [CrossRef]  

13. Z. Wang, C. Yu, W.-D. Zhong, et al., “Performance of a novel LED lamp arrangement to reduce SNR fluctuation for multi-user visible light communication systems,” Opt. Express 20, 4564–4573 (2012). [CrossRef]  

14. J. Wang, C. Ma, D. Li, et al., “Ultrafast and high-power green micro-LED for visible light communications,” in Conference on Lasers and Electro-Optics/Pacific Rim (Optica Publishing Group, 2022), paper CTuP11E_02.

15. L. Wang, Z. Wei, C.-J. Chen, et al., “1.3 GHz EO bandwidth GaN-based micro-LED for multi-gigabit visible light communication,” Photonics Res. 9, 792–802 (2021). [CrossRef]  

16. M. S. Islim, R. X. Ferreira, X. He, et al., “Towards 10 Gb/s orthogonal frequency division multiplexing-based visible light communication using a GaN violet micro-LED,” Photonics Res. 5, A35–A43 (2017). [CrossRef]  

17. Y.-C. Chi, D.-H. Hsieh, C.-T. Tsai, et al., “450-nm GaN laser diode enables high-speed visible light communication with 9-Gbps QAM-OFDM,” Opt. Express 23, 13051–13059 (2015). [CrossRef]  

18. J. J. McKendry, R. P. Green, A. E. Kelly, et al., “High-speed visible light communications using individual pixels in a micro light-emitting diode array,” IEEE Photonics Technol. Lett. 22, 1346–1348 (2010). [CrossRef]  

19. S. Zhang, S. Watson, J. J. D. McKendry, et al., “1.5 Gbit/s multi-channel visible light communications using CMOS-controlled GaN-based LEDs,” J. Lightwave Technol. 31, 1211–1216 (2013). [CrossRef]  

20. D. Tsonev, S. Videv, and H. Haas, “Towards a 100 Gb/s visible light wireless access network,” Opt. Express 23, 1627–1637 (2015). [CrossRef]  

21. C. Lee, C. Shen, C. Cozzan, et al., “Gigabit-per-second white light-based visible light communication using near-ultraviolet laser diode and red-, green-, and blue-emitting phosphors,” Opt. Express 25, 17480–17487 (2017). [CrossRef]  

22. C. Shen, O. Alkhazragi, X. Sun, et al., “Laser-based visible light communications and underwater wireless optical communications: a device perspective,” Proc. SPIE 10939, 109390E (2019). [CrossRef]  

23. L.-Y. Wei, C.-W. Hsu, C.-W. Chow, et al., “20.231 Gbit/s tricolor red/green/blue laser diode based bidirectional signal remodulation visible-light communication system,” Photonics Res. 6, 422–426 (2018). [CrossRef]  

24. S. Nakamura, “First III–V-nitride-based violet laser diodes,” J. Cryst. Growth 170, 11–15 (1997). [CrossRef]  

25. S. Nakamura, “The roles of structural imperfections in InGaN-based blue light-emitting diodes and laser diodes,” Science 281, 956–961 (1998). [CrossRef]  

26. J. Kang, H. Wenzel, E. Freier, et al., “Continuous-wave operation of 405 nm distributed Bragg reflector laser diodes based on GaN using 10th-order surface gratings,” Photonics Res. 10, 1157–1161 (2022). [CrossRef]  

27. S. P. DenBaars, D. Feezell, K. Kelchner, et al., “Development of gallium-nitride-based light-emitting diodes (LEDs) and laser diodes for energy-efficient lighting and displays,” Acta Mater. 61, 945–951 (2013). [CrossRef]  

28. L. Hu, X. Ren, J. Liu, et al., “High-power hybrid GaN-based green laser diodes with ITO cladding layer,” Photonics Res. 8, 279–285 (2020). [CrossRef]  

29. S. Yamaoka, N.-P. Diamantopoulos, H. Nishi, et al., “Directly modulated membrane lasers with 108 GHz bandwidth on a high-thermal-conductivity silicon carbide substrate,” Nat. Photonics 15, 28–35 (2021). [CrossRef]  

30. A. Dunn, C. Poyser, P. Dean, et al., “High-speed modulation of a terahertz quantum cascade laser by coherent acoustic phonon pulses,” Nat. Commun. 11, 835 (2020). [CrossRef]  

31. S. Watson, M. Tan, S. P. Najda, et al., “Visible light communications using a directly modulated 422 nm GaN laser diode,” Opt. Lett. 38, 3792–3794 (2013). [CrossRef]  

32. C. Lee, C. Zhang, D. L. Becerra, et al., “Dynamic characteristics of 410 nm semipolar (2021) III-nitride laser diodes with a modulation bandwidth of over 5 GHz,” Appl. Phys. Lett. 109, 101104 (2016). [CrossRef]  

33. S. Watson, S. Viola, G. Giuliano, et al., “High speed visible light communication using blue GaN laser diodes,” Proc. SPIE 9991, 99910A (2016). [CrossRef]  

34. S. P. Najda, P. Perlin, T. Suski, et al., “GaN laser diode technology for visible-light communications,” Electronics 11, 1430 (2022). [CrossRef]  

35. Y. Mei, G.-E. Weng, B.-P. Zhang, et al., “Quantum dot vertical-cavity surface-emitting lasers covering the ‘green gap’,” Light Sci. Appl. 6, e16199 (2017). [CrossRef]  

36. L. Wang, L. Wang, C. J. Chen, et al., “Green InGaN quantum dots breaking through efficiency and bandwidth bottlenecks of micro-LEDs,” Laser Photonics Rev. 15, 2000406 (2021). [CrossRef]  

37. M. Zhang, A. Bannerjee, C. S. Lee, et al., “A InGaN/GaN quantum dot green (λ = 524 nm) laser,” Appl. Phys. Lett. 98, 221104 (2011). [CrossRef]  

38. K. Lau, C. M. Gee, T. R. Chen, et al., “Signal-induced noise in fiber-optic links using directly modulated Fabry-Perot and distributed-feedback laser diodes,” J. Lightwave Technol. 11, 1216–1225 (1993). [CrossRef]  

39. T. Wang, “Topical Review: Development of overgrown semi-polar GaN for high efficiency green/yellow emission,” Semicond. Sci. Technol. 31, 093003 (2016). [CrossRef]  

40. P. Gourley, “Microstructured semiconductor lasers for high-speed information processing,” Nature 371, 571–577 (1994). [CrossRef]  

41. L. A. Coldren, “Diode lasers and photonic integrated circuits,” Opt. Eng. 36, 616 (1997). [CrossRef]  

42. L. A. Coldren, S. W. Corzine, and M. L. Mashanovitch, Diode Lasers and Photonic Integrated Circuits (Wiley, 2012).

43. Y. Sun, K. Zhou, M. Feng, et al., “Room-temperature continuous-wave electrically injected InGaN-based laser directly grown on Si,” Nat. Photonics 10, 595–599 (2016). [CrossRef]  

44. J. D. Ralston, S. Weisser, I. Esquivias, et al., “Control of differential gain, nonlinear gain and damping factor for high-speed application of GaAs-based MQW lasers,” IEEE J. Quantum Electron. 29, 1648–1659 (1993). [CrossRef]  

45. J. Shi, Z. Xu, W. Niu, et al., “Si-substrate vertical-structure InGaN/GaN micro-LED-based photodetector for beyond 10 Gbps visible light communication,” Photonics Res. 10, 2394–2404 (2022). [CrossRef]  

46. W. Niu, Z. Xu, Y. Liu, et al., “Key technologies for high-speed Si-substrate LED based visible light communication,” J. Lightwave Technol. 41, 3316–3331 (2023). [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 (6)

Fig. 1.
Fig. 1. Design and simulation of the high-speed GaN mini-LD (with a ridge width of 1.8 μm and cavity length 500 μm). (a) Simulated dynamic parameters dfR/d(IIth)1/2 curve slope efficiency (red) and optical confinement factor (blue) of the devices with various quantum barrier thickness and 2 nm quantum well. (b) Histogram of simulated differential gain (red) and −3 dB bandwidth for an injection current of 140 mA, various quantum well thicknesses, and a 5 nm quantum barrier. (c) Simulated light field profile of quasi-TE fundamental mode in LD with 3 nm/5 nm active region. A ridge of 1.8 μm width is visualized in the white frame. Γ is the optical confinement factor. (d) Simulated frequency response under injection currents ranging from 110 to 140 mA of a 3 nm/5 nm active region structure. The dashed line in the figure is the 3dB line related to the measurement start point (100 MHz). Damping behavior at the low frequency marked by the blue dashed circle is related to the RC (resistance-capacitance) roll-off.
Fig. 2.
Fig. 2. Macroscopic and microscopic structures of the laser. (a) 3D illustration of the fabricated laser. The annotation on the right shows the epitaxial layer structure of the active region from top to bottom. (b) Far-field emission pattern of the laser. (c) Optical microscopy image of the fabricated laser. The n- and p-electrodes are marked in the picture. (d) Scanning electron microscope (SEM) image of cross-sectional view. The ridge waveguide width is 1.8μm. (e) STEM image of the active region. From top to bottom are the p-cladding layer, electron blocking layer (EBL), upper waveguide, and MQWs. (f) Indium mapping of the active region. The two high-brightness lines (marked with the yellow triangle) are quantum wells separated by a quantum barrier. (g) Aluminum mapping of active region. The line with high brightness (marked with the yellow triangle) is the EBL (electron blocking layer).
Fig. 3.
Fig. 3. (a) Light–current–voltage (L–I–V) characteristic of the laser under the condition of continuous wave (CW) injection. (b) Spectra of the mini-LD for injection currents ranging from 60 to 140 mA at room temperature.
Fig. 4.
Fig. 4. (a) Measured frequency response of the fabricated mini-LD for injection currents ranging from 100 to 140 mA. (b) S11 response for injection currents ranging from 100 to 140 mA. (c) Extracted intrinsic S21 response for injection currents ranging from 100 to 140 mA. The 3 and 10dB are labelled in the figure using dashed lines. (d) Extracted intrinsic modulation bandwidth (3 and 10dB) versus injection currents ranging from 70 to 140 mA. (e) Relationship between resonance frequency (fR) and (IIth)1/2, where I is the injection current and Ith is the threshold current. The line in this figure is the fitting curve. (f) Relationship between square of resonance frequency (fR) and the damping factor. The line in this figure is the fitting curve. K is the slope of the curve, and γ0 is the intercept on the Y axis.
Fig. 5.
Fig. 5. (a) Transmitter in the system using the discrete multitone (DMT) bit-power loading modulation method. The digital signal is processed through bit-power loading, DMT modulation, and digital pre-equalization. AWG is the arbitrary waveform generator. EA is the electronic amplifier. The generated signal is combined with the direct current (D.C.) of the bias and then transmitted to the mini-LD. (b) Experimental setup of the high-bandwidth mini-LD-based high-speed VLC system with DMT bit-power loading modulation. The mini-LD is equipped with a heat sink with the TEC to ensure constant temperature (20°C) and the transmitted signal is collected by the PD. (c) The receiver in the system which contains the post-processing of the signal. PD is the photodetector, OSC is the oscilloscope, LMS Volterra is the least mean square Volterra nonlinear filter, and ZF Equ. is the zero-forcing equalizer. After the post-processing, we obtained the transmitted signal.
Fig. 6.
Fig. 6. Experimental results of the VLC system using fabricated mini-LD. (a) Bit-power loading scheme of each subcarrier according to measured channel SNR. The effective carrier number is 504. (b) Received constellation diagrams of the system, including 128 QAM, 64 QAM, 32 QAM, 16 QAM, 8 QAM, 4 QAM, and 2 QAM. (c) Partial T/R symbol of the mini-LD-based VLC data links. The tested BER is 0.0030.

Tables (1)

Tables Icon

Table 1. Summary of the Key Parameters for High-Speed Blue GaN LDsa

Equations (7)

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

H(f)=fR2fR2f2+jγf/2π.
S21M=S21LR.
η=Zr2R(1+S22P)S21PS21M.
fR=12π[ΓυgdgdnqVηi(IIth)]1/2.
γ=KfR2+γ0,
x(t)=k=0N11NP(k)X(k)ej2πNkt,
x(t)=F1{F[x(t)]·Hpre},
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.