Next Article in Journal
Achieving High Tensile Strength of Heat-Resistant Ni-Fe-Based Alloy by Controlling Microstructure Stability for Power Plant Application
Previous Article in Journal
Electro-Acoustic Properties of Scandium-Doped Aluminum Nitride (ScxAl1-xN) Material and its Application to Phononic Crystal-Coupled SAW Devices
Previous Article in Special Issue
Influence of Temperature Field Distribution on the Growth of Aluminum Nitride Crystal by Simulation Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Simple Method to Build High Power PCSEL Array with Isolation Pattern Design

1
Department of Photonics, College of Electrical Engineering, National Yang Ming Chiao Tung University, Hsinchu City 30010, Taiwan
2
Semiconductor Research Center, Hon Hai Research Institute, Taipei City 23678, Taiwan
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(10), 1432; https://doi.org/10.3390/cryst12101432
Submission received: 2 September 2022 / Revised: 6 October 2022 / Accepted: 9 October 2022 / Published: 10 October 2022
(This article belongs to the Special Issue Artificial Crystals)

Abstract

:
Photonic crystal surface-emitting lasers (PCSELs) hold promising properties of both edge emitting lasers (EELs) and vertical-cavity surface-emitting lasers (VCSELs). They possess high output power while radiating light vertically, being thought of as the next generation laser source. One of the main advantages of PCSELs is their scalability of size and power, which makes them applicable to high power applications or long-distance detection. However, due to problems such as current injection and mode competition, there are limits on their dimensions. To further increase the power, the capability of two-dimensional array integration paves the way. In this work, we demonstrate a new method to fabricate PCSEL arrays by defining an isolation pattern. We also investigate the influence of aperture size and array arrangement on lasing performance.

1. Introduction

The applications of three-dimensional sensing and light detection and ranging (LiDAR) technologies in autonomous vehicles and mobile devices are growing fast these years. To meet the needs, high power and compact laser systems are urgently needed [1,2]. Recently, photonic crystal surface-emitting lasers (PCSELs) have emerged as the most promising laser sources because of their unique properties. PCSELs are a new class of semiconductor lasers that incorporate a two-dimensional photonic crystal layer into the structure. Forming standing waves with zero group velocity at band edges in a photonic band, PCSELs oscillate in a horizontal direction and radiate power in a vertical direction [3,4,5,6,7,8]. Surface emitting means a low cost of testing and packaging, and in-plane oscillation means a high power due to a larger active region and external efficiency. They offer high power single mode emission beams with a narrow divergence angle, and enable on-chip beam steering [9,10,11,12,13,14,15]. It is believed that lens-free and adjustment-free LiDAR systems can be realized in the future [16].
Many efforts have been made to increase the output power of PCSELs. They include buried air hole photonic crystal structures, introducing DBR into a PCSEL structure, filling factor optimization, and nonsymmetric photonic crystal design [17,18,19,20,21,22]. As well as these, PCSELs are especially suitable for power scaling [23,24] because of the in-plane oscillation behavior and the absence of catastrophic damage. As one scales up the device in size, the power also scales up in proportional to its area. The main challenges encountered when scaling devices are the current injection and mode competition [25,26]. When the device size grows too large, the carrier cannot be injected uniformly into the active layer inside the aperture. On the other hand, a larger size means smaller in-plane loss, which also shrinks the gain margin and causes multi-mode operation. To ameliorate these problems, a p-side down structure and double-lattice photonic crystal structure were introduced [27,28,29]. With these efforts, a 10-W peak power has been realized at room temperature on a single emitter until now. In a recent publication, Noda et al. theoretically derived the conditions of a photonic crystal to maintain single mode operation as the device size increases above 3 mm [30]. However, other issues needed consideration in practice, such as nonuniform current distribution, self-heating, and stitching error from e-beam writing, which have not been addressed, and they may put some upper limit on device scaling. To further improve the output power, scaling with an array provides a new route.
An output power larger than 100 W has already been demonstrated with a VCSEL array and received applications in many fields [31]. However, the VCSEL array suffers problems such as a larger divergence angle and donut-shape beam. Unlike VCSELs, which oscillate and radiate in the same direction, PCSELs in an array can easily couple with others with an appropriate connection design. When the array oscillates coherently, it can be viewed as an ultra-large single device. It can possess an output power greater than 100 W and a divergence angle even smaller than the prevailing VCSELs while retaining high beam quality. The future of PCSEL arrays is promising. However, only a few researches have focused on it. In previous researches, the group of Hogg connected PCSELs through bridge waveguides. With current injection into the waveguides, a coherent PCSEL array was realized successfully, and the coherence was demonstrated with the interference of a far-field pattern [32,33]. As an alternative method, we propose here a simple way to build PCSEL arrays. By defining an isolation pattern on our epitaxy structure, we can control the injection region and split them into individual PCSELs with no current injection in-between. PCSEL arrays formed in this way were demonstrated, and we also investigated the influence of aperture size and arrangement.

2. Methods

The structure of our PCSEL is depicted in Figure 1a, which consists of cladding layers, active region, a photonic crystal (PC) layer, and a p-cladding layer on the top. We designed the lasing wavelength at around 940 nm by optimizing the lattice constant of the PC and geometries of the unit cell. The isosceles right triangular-shaped air hole as shown at Figure 1b was applied as the unit cell of photonic crystal for higher out coupled efficiency. We further deposited SiNx and indium tin oxide (ITO) layers, followed by fabricating metal electrodes, where the SiNx layer served as passivation layer to define the aperture of PCSELs. The region under passivation was estimated to have nearly zero injection and split those apertures into individual PCSELs. The highly conductive and transparent ITO laying on the top of it assisted current distribution evenly through the emission aperture defined by the SiNx. Finally, metal was deposited above ITO with the same pattern as SiNx. The laser emission direction was designed from the top surface due to its ease of fabrication. We demonstrated such a PCSEL device, and the process of output power optimization has been detailed in our previous reports [21,22].
To investigate the influence of aperture size and array arrangement, we designed four patterns with different aperture sizes and arrangements. The optical microscope (OM) images of the four designed patterns are shown in Figure 2. The four arrays had the same size of 1 mm. The aperture sizes we used were 60 μm and 100 μm in diameter. The current injection uniformity of p-side up PCSELs with these aperture sizes was demonstrated to be good enough. The interval of two adjacent apertures was fixed at 40 μm. PCSELs with each aperture size were arranged in square and triangular forms, respectively. Arrays with aperture size 100 μm had about 50 PCSEL emitters in total, while 100 PCSEL emitters for aperture size 60 μm. For convenience, the four fabricated samples were named as S100, T100, S60, T60, where T and S mean triangular and square arrangements, and the later numbers indicate aperture diameters.

3. Results and Discussion

Figure 3a,b represent the light–current-voltage (L-I-V) characteristics. The measurements were conducted at room temperature, with 1 μs pulse duration and 1 kHz repetition rate. The turn-on voltages were about the same for all the arrays; however, obvious differences in serial resistance were observed. The arrays with larger aperture size had higher serial resistance, this magnitude was closely related to the current spreading uniformity in structure. The thermal rollover phenomenon did not appear in the four samples until a 6A injection current, and the pumping current was limited by the diode driver. The threshold current density and slope efficiency of the four samples extracted from Figure 3a,b are shown in Figure 3c. The threshold current density was calculated by dividing the current by the total area of the aperture. The four samples sort by threshold current were T100 < S100 < T60 < S60. We can deduce from the measured results that arrays with a larger aperture size and triangular arrangement have smaller thresholds. The order of the slope efficiency of the four samples was T100 > S100 >T60 > S60. Again, arrays with a larger aperture size and triangular arrangement have better performance. Figure 3d shows the power conversion efficiency (PCE) of the four samples. They reach maximum PCE 2~3.5% at a current injection of about 5A. The relatively low PCE is due to the high internal loss and ohmic loss from the unoptimized design and fabrication process. Methods such as grading the hetero-interface to improve its electrical property, properly designing the doping profile to reduce free carrier absorption, optimizing the barrier or electron-blocking layer to reduce the current carrier leakage, optimizing the unit cell pattern to properly balance the threshold and slope efficiency, or adding DBR layers to reflect the downward propagating light upward can be used to improve the PCE of PCSELs. Array T100 had the highest PCE because of its obviously larger power, while array S100 had the lowest PCE because it had the largest serial resistance. Figure 4a shows the near field pattern of array S100 measured above the threshold. We can observe uniform emissions for the array emitters. Almost all the PCSELs in the array show similar emission near field patterns. However, as shown in Figure 4b, multiple peaks in the spectrum were observed in the array S100. Similar results were obtained from other samples, they all possessed multiple peaks, which imply the low coherence of the array. The measured far-field patterns of the four arrays are shown in Figure 4c. They possess single-lobe distribution with no obvious interference pattern. These results imply low coherence again, in agreement with the measured spectra. The far-field angles (1/e2) of them were obtained and were around 4 degrees, and no obvious differences were observed in the different arrangements or aperture sizes.
From the point of view of coupling, the arrays with square arrangement are easier to couple in both an upper–lower and left–right direction, and are thought to have stronger in-plane feedback. Stronger in-plane feedback implies a lower threshold and higher slope efficiency. However, the experiment shows contrasting results. To understand the underlying mechanism of how aperture size and arrangement influence the electrical and optical properties, we performed numerical simulation to derive the current and carrier distribution in the active layer at a current slightly below the threshold. Figure 5 shows the simulation results of the normalized current density distribution in the active layer. The dashed circles in the figures are positions where current was injected according to the experiment situation. The current density drops rapidly outside the apertures; here we tune the range of color to focus on the current distribution inside the apertures. About 10~15% decay of the current density from the dashed circle to the center of the array is observed. To make the comparison quantitative, we first calculated the averaged current density of each aperture, then normalized them to their own maximum in an array. Finally, we defined the current distribution nonuniformity of an array as the RMS of these values. The calculated nonuniformities of the four samples were 10.1%, 12.3%, 6.1%, and 7.2%, respectively. PCSEL arrays with a smaller aperture size show much better current distribution. Moreover, the calculation results reveal that the uniformity is weakly dependent on the array arrangement. This result is consistent with the measured serial resistance. We suggest that the coverage of p-metal is the key factor. Metal coverage ratios in the unit cell of the four samples were 59.9%, 53.7%, 71.7%, and 67.3%, respectively, which gave a good prediction of the current injection uniformity.
Figure 6 shows the corresponding carrier density in the active layer along the two dashed lines in Figure 5. The carrier density of the areas below the SiN pattern is much smaller than 1018 cm−3, which is the approximate value of the transparency condition of the InGaAs quantum well; so outside the aperture there exists strong band-to-band absorption. This can explain why PCSELs in our samples did not couple with each other. By comparing the carrier distribution along two lines, we observe a lower carrier density in the interval where the aperture spacing is larger. Thus, the triangular arrangement, which has closer packing is thought to have a lower threshold and higher slope efficiency due to the smaller absorption loss. On the other hand, it is possible to fabricate coherent PCSEL arrays as long as the interval of the apertures is narrow enough. It is worth noting that close packing or a narrower metal path result in worse current spreading. To achieve coherent lasing while retaining uniform current spreading, SiNx and metal patterns need further optimization and design. Similarly, PCSELs with a larger aperture size have smaller in-plane loss, resulting in a lower threshold and higher slope efficiency. These could well explain our experimental results. Further improvements include adding epitaxial mirrors below the n-cladding layer to boost the upward emission efficiency and fine-tuning the aperture size and spacing between apertures to balance requirements between current distribution, in-plane loss, out-coupling efficiency, and thermal accumulation.

4. Conclusions

In this report, we demonstrate PCSEL arrays that are simply formed with a silicon nitride isolation pattern. The aperture size and arrangement of the PCSEL array will directly affect the output power, with a larger aperture size and closer packing result and a smaller optical loss but with poor current spreading. Despite our PCSEL arrays not being coherent, they can still be used in three-dimensional sensing and ToF applications, which have low requirements of coherence. In future work, the emitters will be put closer together to improve the coherence in order to scale up the brightness; that is, we can reduce the width of SiNx. To maintain good spreading of the current at the same time, we can just reduce the width of SiNx while keeping the width of the metal the same or increasing the thickness of the metal. On the other hand, we can also reduce the metallization area while increasing the thickness of the p-cladding layer to compensate for the reduced contact area with the metal.

Author Contributions

Conceptualization, L.-R.C. and T.-C.L.; methodology, L.-R.C., W.-C.W., B.-H.C. and T.-C.L.; software, C.-J.C.,W.-C.W. and K.-B.H.; validation, C.-J.C. and K.-B.H. and T.-C.L.; formal analysis, C.-J.C.,W.-C.W., L.-R.C. and K.-B.H.; investigation, C.-J.C., L.-R.C. and K.-B.H.; writing—original draft preparation, C.-J.C.; writing—review and editing, K.-B.H. and T.-C.L.; supervision, H.-C.K. and T.-C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology in TAIWAN under Contract No. MOST 110-2221-E-A49-058-MY3, MOST 110-2622-8-A49-008-SB.

Data Availability Statement

Due to the nature of this research, participants of this study did not agree for their data to be shared publicly, so supporting data is not available.

Acknowledgments

The authors would like to acknowledge the LandMark Optoelectronics Corp., Taiwan, Nano Facility Center (NFC), Center for Nano Science and Technology (CNST) in NYCU, and also the Industrial Technology Research Institute (ITRI) for their technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schwarz, B. Mapping the world in 3D. Nat. Photonics 2010, 4, 429–430. [Google Scholar] [CrossRef]
  2. Poulton, C.V.; Yaacobi, A.; Cole, D.B.; Byrd, M.J.; Raval, M.; Vermeulen, D.; Watts, M.R. Coherent solid-state LIDAR with silicon photonic optical phased arrays. Opt. Lett. 2017, 42, 4091–4094. [Google Scholar] [CrossRef] [PubMed]
  3. Imada, M.; Noda, S.; Chutinan, A.; Tokuda, T.; Murata, M.; Sasaki, G. Coherent two-dimensional lasing action in surface-emitting laser with triangular-lattice photonic crystal structure. Appl. Phys. Lett. 1999, 75, 316–318. [Google Scholar] [CrossRef]
  4. Imada, M.; Chutinan, A.; Noda, S.; Mochizuki, M. Multidirectionally distributed feedback photonic crystal lasers. Phys. Rev. B 2002, 65, 195306. [Google Scholar] [CrossRef]
  5. Noda, S.; Yokoyama, M.; Imada, M.; Chutinan, A.; Mochizuki, M. Polarization mode control of two-dimensional photonic crystal laser by unit cell structure design. Science 2001, 293, 1123–1125. [Google Scholar] [CrossRef] [PubMed]
  6. Sakai, K.; Miyai, E.; Sakaguchi, T.; Ohnishi, D.; Okano, T.; Noda, S. Lasing band-edge identification for a surface-emitting photonic-crystal laser. IEEE J. Sel. Areas Commun. 2005, 23, 1335–1340. [Google Scholar] [CrossRef]
  7. Ryu, H.Y.; Kwon, S.H.; Lee, Y.J.; Lee, Y.H. Very-low-threshold photonic band-edge lasers from free-standing triangular photonic crystal slabs. Appl. Phys. Lett. 2002, 80, 3476–3478. [Google Scholar] [CrossRef]
  8. Kim, M.; Kim, C.S.; Bewley, W.W.; Lindle, J.R.; Canedy, C.L.; Vurgaftman, I.; Meyer, J.R. Surface-emitting photonic-crystal distributed-feedback laser for the midinfrared. Appl. Phys. Lett. 2006, 88, 191105. [Google Scholar] [CrossRef]
  9. Huang, S.-C.; Hong, K.-B.; Chiu, H.-L.; Lan, S.-W.; Chang, T.-C.; Li, H.; Lu, T.-C. Design of photonic crystal surface emitting lasers with indium-tin-oxide top claddings. Appl. Phys. Lett. 2018, 112, 061105. [Google Scholar] [CrossRef]
  10. McKenzie, A.F.; King, B.C.; Rae, K.J.; Thoms, S.; Gerrard, N.D.; Orchard, J.R.; Nishi, K.; Takemasa, K.; Sugawara, M.; Taylor, R.J.E.; et al. Void engineering in epitaxially regrown GaAs-based photonic crystal surface emitting lasers by grating profile design. Appl. Phys. Lett. 2021, 118, 021109. [Google Scholar] [CrossRef]
  11. Noda, S.; Kitamura, K.; Okino, T.; Yasuda, D.; Tanaka, Y. Photonic-Crystal Surface-Emitting Lasers: Review and Introduction of Modulated-Photonic Crystals. IEEE J. Sel. Top. Quantum Electron. 2017, 23, 1–7. [Google Scholar] [CrossRef]
  12. Chen, L.-R.; Chiu, H.-L.; Hong, K.-B.; Lu, T.-C. Low threshold current photonic crystal surface emitting lasers with beam modulation capability. In Proceedings of the 2019 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 5–10 May 2019. [Google Scholar]
  13. Chen, L.R.; Hong, K.B.; Chen, H.L.; Huang, K.C.; Lu, T.C. Vertically integrated diffractive gratings on photonic crystal surface emitting lasers. Sci. Rep. 2021, 11, 2427. [Google Scholar] [CrossRef]
  14. Heck, M.J.R. Highly integrated optical phased arrays: Photonic integrated circuits for optical beam shaping and beam steering. Nanophotonics 2017, 6, 93–107. [Google Scholar] [CrossRef]
  15. Chiu, H.-L.; Hong, K.-B.; Huang, K.-C.; Lu, T.-C. Photonic Crystal Surface Emitting Lasers with Naturally Formed Periodic ITO Structures. ACS Photonics 2019, 6, 684–690. [Google Scholar] [CrossRef]
  16. Hong, Y.-H.; Miao, W.-C.; Hsu, W.-C.; Hong, K.-B.; Lin, C.-L.; Lin, C.; Chen, S.-C.; Kuo, H.-C. Progress of Photonic-Crystal Surface-Emitting Lasers: A Paradigm Shift in LiDAR Application. Crystals 2022, 12, 800. [Google Scholar] [CrossRef]
  17. Hirose, K.; Liang, Y.; Kurosaka, Y.; Watanabe, A.; Sugiyama, T.; Noda, S. Watt-class high-power, high-beam-quality photonic-crystal lasers. Nat. Photonics 2014, 8, 406–411. [Google Scholar] [CrossRef]
  18. Williams, D.M.; Groom, K.M.; Stevens, B.J.; Childs, D.T.D.; Taylor, R.J.E.; Khamas, S.; Hogg, R.A.; Ikeda, N.; Sugimoto, Y. Epitaxially regrown GaAs-based photonic crystal surface-emitting laser. IEEE Photonics Technol. Lett. 2012, 24, 966–968. [Google Scholar] [CrossRef]
  19. Chen, L.-R.; Hong, K.-B.; Huang, K.-C.; Liu, C.-L.; Lin, W.; Lu, T.-C. Study of an Epitaxial Regrowth Process by MOCVD for Photonic-Crystal Surface-Emitting Lasers. Cryst. Growth Des. 2021, 21, 3521–3527. [Google Scholar] [CrossRef]
  20. Sakaguchi, T.; Kunishi, W.; Arimura, S.; Nagase, K.; Miyai, E.; Ohnishi, D.; Sakai, K.; Noda, S. Surface-emitting photonic-crystal laser with 35W peak power. In Proceedings of the Conference on Lasers and Electro-Optics, Munich, Germany, 14–19 June 2009; pp. 1–2. [Google Scholar]
  21. Hong, K.-B.; Chen, L.-R.; Huang, K.-C.; Yen, H.-T.; Weng, W.-C.; Chuang, B.-H.; Lu, T.-C. Impact of Air-Hole on the Optical Performances of Epitaxially Regrown P-Side Up Photonic Crystal Surface-Emitting Lasers. IEEE J. Sel. Top. Quantum Electron. 2022, 28, 1–7. [Google Scholar] [CrossRef]
  22. Chen, L.R.; Hong, K.B.; Huang, K.C.; Yen, H.T.; Lu, T.C. Improvement of output efficiency of p-face up photonic-crystal surface-emitting lasers. Opt. Express 2021, 29, 11293–11300. [Google Scholar] [CrossRef]
  23. Kalapala, A.; Song, A.; Pan, M.; Gautam, C.; Overman, L.; Reilly, K.; Rotter, T.; Balakrishnan, G.; Gibson, R.; Bedford, R.; et al. Scaling Challenges in High Power Photonic Crystal Surface-Emitting Lasers. IEEE J. Quantum Electron. 2022, 58, 1–9. [Google Scholar] [CrossRef]
  24. Liu, S.; Zhao, D.; Yang, H.; Reuterskiöld-Hedlund, C.; Hammar, M.; Fan, S.; Ma, Z.; Zhou, W. Lateral Size Scaling of Photonic Crystal Surface-Emitting Lasers on Si, in Conference on Lasers and Electro-Optics; OSA Technical Digest: Washington, DC, USA, 2017. [Google Scholar]
  25. Liang, Y.; Peng, C.; Sakai, K.; Iwahashi, S.; Noda, S. Three-dimensional coupled-wave model for square-lattice photonic crystal lasers with transverse electric polarization: A general approach. Phys. Rev. B 2011, 84, 195119. [Google Scholar] [CrossRef] [Green Version]
  26. Liang, Y.; Peng, C.; Sakai, K.; Iwahashi, S.; Noda, S. Three-dimensional coupled-wave analysis for square-lattice photonic-crystal lasers with transverse electric polarization: Finite-size effects. Opt. Express 2012, 20, 15945–15961. [Google Scholar] [CrossRef] [PubMed]
  27. Yoshida, M.; De Zoysa, M.; Ishizaki, K.; Tanaka, Y.; Kawasaki, M.; Hatsuda, R.; Song, B.; Gelleta, J.; Noda, S. Double-lattice photonic-crystal resonators enabling high-brightness semiconductor lasers with symmetric narrow-divergence beams. Nat. Mater. 2019, 18, 121–128. [Google Scholar] [CrossRef]
  28. Yoshida, M.; Kawasaki, M.; De Zoysa, M.; Ishizaki, K.; Inoue, T.; Tanaka, Y.; Hatsuda, R.; Noda, S. Experimental Investigation of Lasing Modes in Double-Lattice Photonic-Crystal Resonators and Introduction of In-Plane Heterostructures. Proc. IEEE 2020, 108, 819–826. [Google Scholar] [CrossRef]
  29. De Zoysa, M.; Yoshida, M.; Song, B.-S.; Ishizaki, K.; Inoue, T.; Katsuno, S.; Izumi, K.; Tanaka, Y.; Hatsuda, R.; Gelleta, J.; et al. Thermal management for CW operation of large-area double-lattice photonic-crystal lasers. J. Opt. Soc. Am. B 2020, 37, 3882–3887. [Google Scholar] [CrossRef]
  30. Inoue, T.; Yoshida, M.; Gelleta, J.; Izumi, K.; Yoshida, K.; Ishizaki, K.; De Zoysa, M.; Noda, S. General recipe to realize photonic-crystal surface-emitting lasers with 100-W-to-1-kW single-mode operation. Nat. Commun. 2022, 13, 3262. [Google Scholar] [CrossRef]
  31. Seurin, J.-F. et al. High-power high-efficiency 2D VCSEL arrays. In Vertical-Cavity Surface-Emitting Lasers XII; SPIE: Bellingham, WA, USA, 2008; Volume 6908. [Google Scholar]
  32. King, B.C.; Rae, K.J.; McKenzie, A.F.; Boldin, A.; Kim, D.; Gerrard, N.D.; Li, G.; Nishi, K.; Takemasa, K.; Sugawara, M.; et al. Coherent power scaling in photonic crystal surface emitting laser arrays. AIP Adv. 2021, 11, 015017. [Google Scholar] [CrossRef]
  33. Taylor, R.J.E.; Childs, D.T.D.; Ivanov, P.; Stevens, B.J.; Babazadeh, N.; Crombie, A.J.; Ternent, G.; Thoms, S.; Zhou, H.; Hogg, R.A. Electronic control of coherence in a two-dimensional array of photonic crystal surface emitting lasers. Sci. Rep. 2015, 5, 13203. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Schematic diagram of high power PCSEL array. (b) Top view SEM image of the photonic crystal.
Figure 1. (a) Schematic diagram of high power PCSEL array. (b) Top view SEM image of the photonic crystal.
Crystals 12 01432 g001
Figure 2. Optical microscope images of the four PCSEL array patterns. From left to right are PCSEL arrays S100, T100, S60, and T60, respectively.
Figure 2. Optical microscope images of the four PCSEL array patterns. From left to right are PCSEL arrays S100, T100, S60, and T60, respectively.
Crystals 12 01432 g002
Figure 3. (a,b) LIV characteristics of PCSEL arrays. (c) Extracted threshold current density and slope efficiency. (d) Power conversion efficiencies of the four samples.
Figure 3. (a,b) LIV characteristics of PCSEL arrays. (c) Extracted threshold current density and slope efficiency. (d) Power conversion efficiencies of the four samples.
Crystals 12 01432 g003
Figure 4. (a) Measured near field pattern of array S100 operated above threshold. (b) Measured spectrum of array S100. (c) Measured far-field patterns of the four arrays.
Figure 4. (a) Measured near field pattern of array S100 operated above threshold. (b) Measured spectrum of array S100. (c) Measured far-field patterns of the four arrays.
Crystals 12 01432 g004
Figure 5. Cross-section plots of normalized current density distributions located in the active layers below threshold. From left to right are PCSEL arrays S100, T100, S60, and T60, respectively. The current is injected from dashed circle on top of the device and is drained from bottom electrode.
Figure 5. Cross-section plots of normalized current density distributions located in the active layers below threshold. From left to right are PCSEL arrays S100, T100, S60, and T60, respectively. The current is injected from dashed circle on top of the device and is drained from bottom electrode.
Crystals 12 01432 g005
Figure 6. The estimated carrier density distribution along the two lines in Figure 5.
Figure 6. The estimated carrier density distribution along the two lines in Figure 5.
Crystals 12 01432 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chang, C.-J.; Chen, L.-R.; Hong, K.-B.; Weng, W.-C.; Chuang, B.-H.; Kuo, H.-C.; Lu, T.-C. A Simple Method to Build High Power PCSEL Array with Isolation Pattern Design. Crystals 2022, 12, 1432. https://doi.org/10.3390/cryst12101432

AMA Style

Chang C-J, Chen L-R, Hong K-B, Weng W-C, Chuang B-H, Kuo H-C, Lu T-C. A Simple Method to Build High Power PCSEL Array with Isolation Pattern Design. Crystals. 2022; 12(10):1432. https://doi.org/10.3390/cryst12101432

Chicago/Turabian Style

Chang, Chia-Jui, Lih-Ren Chen, Kuo-Bin Hong, Wei-Chih Weng, Bing-Hong Chuang, Hao-Chung Kuo, and Tien-Chang Lu. 2022. "A Simple Method to Build High Power PCSEL Array with Isolation Pattern Design" Crystals 12, no. 10: 1432. https://doi.org/10.3390/cryst12101432

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop