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

III-nitride photonic crystal emitters by selective photoelectrochemical etching of heterogeneous quantum well structures

Open Access Open Access

Abstract

We demonstrate a top-down fabrication strategy for creating a III-nitride hole array photonic crystal (PhC) with embedded quantum wells (QWs). Our photoelectrochemical (PEC) etching technique is highly bandgap selective, permitting the removal of QWs with well-defined indium (In) concentration. Room-temperature micro-photoluminescence (μ-PL) measurements confirm the removal of one multiple quantum well (MQW) while preserving a QW of differing In concentration. Moreover, PhC cavity resonances, wholly unobservable before, are present following PEC etching. Our results indicate an interesting route for creating III-nitride membranes with tailorable emission wavelengths. Our top-down fabrication approach offers exciting opportunities for III-nitride based light emitters.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

III-nitride materials have become important as efficient light sources for generating visible light, as indicated by their pervasive presence in solid state lighting [2,3]. These materials are also becoming increasingly important for coherent [4] and quantum light sources [5–7] operating at room temperature. Recent demonstration of topological lasing [8] and unidirectional single photon emission [9] in III-V systems that achieve high slope efficiency while operating in single mode also suggests such possibilities at visible and ultra-violet wavelength using III-nitrides. In such light sources, high emission rates and control of emission directionality are extremely desirable. Increased emission rates reduce lasing thresholds (and enhances photon counts for quantum light sources) while control of emission directionality improves light collection efficiency.

Photonic crystals (PhC) offer both of those desired properties and have been extensively used to modify light emission. For instance, PhCs have been broadly demonstrated in III-V semiconductor systems as lasers and single photon emitters. In III-nitride material systems there have also been demonstrations of incoherent light emission and lasing in PhC geometries featuring nanowire arrays [10–16]. Complementary structures fabricated from III-nitrides, namely two-dimensional hole array PhCs [17–27], have also been demonstrated. Compared to nanowire array PhCs, hole array PhCs are more suitable for making electrical contacts, thus lending themselves more easily to electrical injection. Furthermore, hole array PhCs can enable excellent light confinement within a thin active layer through optical isolation within a high refractive index substrate. Optical isolation can be achieved by having a low index cladding layer separating the substrate from the active layer or by completely removing the cladding to create an active layer membrane. Such optical isolation substantially enhances quality factors (Q-factors) and Purcell enhancement, thereby maximizing the benefits of photonic nanostructuring. PhC membranes have been extensively demonstrated in photonic devices made from narrower gap infrared III-V materials like GaAs and InP [28–35]. In such material systems, the active layers have high refractive index (n ~3.4) enabling excellent light confinement. Furthermore, sacrificial layers like AlxGa1-xAs can be grown to desired thickness which can be subsequently oxidized to a low refractive index AlxOy or etched away entirely, thereby providing isolation from the underlying substrate. Alternatively, easier substrate removal (e.g. InP) with wet etching enables flip-chip bonding to low-index substrates.

Unlike infrared III-V materials, achieving optical isolation in III-nitride hole array PhCs remains challenging. III-nitrides typically possess a lower refractive index (n ~2.4) resulting in reduced light confinement and thereby requiring larger separation from an underlying substrate for isolation. Moreover, a sacrificial material analogous to AlxGa1-xAs that can be grown to sufficient thickness and subsequently either etched away (or chemically modified) is also unavailable for the III-nitrides. Due to these challenges, a few different approaches have been explored to address the issue of optical isolation. Such approaches include growing III-nitride material on Si substrates coupled with a substrate dry etch [36], focused ion milling [37], or flip-chip bonding techniques. However, all such approaches pose different challenges and have their own limitations.

Another approach that has been studied for achieving optical isolation of active regions in III-nitrides is the photoelectrochemical etching (PEC) technique [18–20]. PEC etching is a bandgap selective wet-etch [1,38–40], wherein absorption of light with energy above the material electronic bandgap generates carriers aiding in the etch process. If a sacrificial material of smaller bandgap absorbs light of lower energies, it can be preferentially etched away while preserving any structures with larger bandgap. In previous approaches [18–20,40], optical stimulation was provided by high powered Xe lamps using unintentionally doped GaN film as a filter to etch sacrificial InxGa1-x N superlattice layers with low indium content (x~8%). However, a GaN filter still transmits some amount of higher energy light causing a non-trivial amount of etching within any higher bandgap materials.

In this work we demonstrate laser-assisted PEC etching of a sacrificial multi-quantum well (MQW) layer containing In0.16Ga0.84N for a PhC cavity structure using laser illumination. The In content in the sacrificial MQW InGaN layer used for PEC etching is substantially higher than previously reported. Moreover, we utilize a tunable picosecond (ps) laser source for illumination as reported in our previous works [1,38,39], instead of a filtered broadband source. Our tunable laser source allows for substantially better control of the PEC etch process due to the narrow linewidth of the laser whereby there is negligible illumination at higher energies. This fact helps minimize unintentional etching of higher bandgap material as described previously [40]. Use of a high intensity tunable laser also enables incorporation of higher In concentration in the sacrificial InxGa1-xN layers which consequently permits higher In concentration in the active QW regions. The higher active layer In content enables PhC devices with longer emission wavelengths.

2. Nanofabrication

Here, the epitaxial structure is grown using metal-organic chemical vapor deposition (MOCVD) (Fig. 1). The wafer consists of a nominally 4.7 μm thick n-GaN buffer layer grown on a sapphire template. The buffer layer growth is followed by a 150nm thick In0.03Ga0.97N underlayer and proceeds with a fifteen pair MQW structure, featuring 2.7 nm thick In0.16Ga0.84N QW layers and 8 nm thick n-GaN barriers, resulting in a total thickness of approximately 160nm. A MQW structure was chosen for the sacrificial region instead of a thick InGaN layer in order to enable higher In content while retaining good epitaxial morphology for the single quantum well (SQW) active layer. Peak emission associated with the MQW is centered near a wavelength of 460nm. The MQW growth is followed with a 65nm thick n-GaN layer. Next, a single 2.5 nm thick In0.10Ga0.90N QW layer is grown andcapped with 65 nm of n-GaN. Peak emission associated with the single quantum well (SQW) is centered near a wavelength of 410 nm. In future work, the SQW can be PEC etched to create a single quantum dot. With an appropriately designed topological PhC [41–43] in-plane unidirectional single photon emission is also possible [9].

 figure: Fig. 1

Fig. 1 Schematic showing the MOCVD epitaxial growth structure. The inset shows the single quantum well (SQW) active layer and the multiple quantum well (MQW) sacrificial layer.

Download Full Size | PDF

To fabricate the photonic crystal device (Fig. 2(a)-2(e)) we begin with a plasma-enhanced deposition of SiO2 to be used as a hard mask. Next, we spin polymethylmethacrylate (PMMA) resist onto our sample and pattern a series of holes using electron beam lithography (EBL). Following development, a F3- based reacive ion etch (RIE) transfers the pattern from the PMMA to the underlying SiO2. Next, an inductively coupled plasma (ICP) Cl2/BCl3-based etch transfers the photonic crystal pattern to the III-nitride sample. The dry etch step is followed by a laser-assisted PEC wet etch. Figure 2(f) shows a schematic of the PEC etch setup used in this step. During the PEC etch process, the III-nitride sample is partially submerged in a solution of 0.2 M H2SO4. An In contact is bonded to the III-nitride sample while a Pt electrode is suspended in the electrolyte solution. A tunable, frequency doubled ps Ti:Sapphire laser with a relatively narrow-bandwidth (~1 nm) is used as the photoexcitation source. During the PEC wet-etch, H2SO4 accesses the MQW sacrificial region through holes in the patterned substrate (Fig. 2(e)), and then proceeds to etch laterally through the MQW layer. Since the PEC wet-etch is a bandgap selective etching mechanism, the laser wavelength was set at 450nm. The resulting laser energy is above the absorption edge of the In0.16Ga0.84N layer while still below that of the SQW (In0.10Ga0.90N) layer. Further, the narrow linewidth of the laser enables high PEC etch selectivity between the MQW and the SQW regions. Laser light absorption results in the photoexcitation of carriers needed for the wet chemical etching. A voltage is applied across the In contact and Pt electrode to attract photoexcited carriers towards the surface of the sample where holes promote surface oxidation required for etching.

 figure: Fig. 2

Fig. 2 Device fabrication: (a) PECVD of SiO2 (b) EBL patterning and development (c) Flourine based dry etch (d) Chlorine based dry etch and (e) PEC etching. Light green indicates GaN, dark green indicates SiO2, blue indicates PMMA, a single black line represents a SQW structure and a series of black lines represents a MQW structure. (f) Schematic of the QSC-PEC etching setup. A tunable picosecond laser (blue) excites carriers in a III-nitride sample (green) that is partially submerged in an electrolyte solution (clear).(figure adapted from Fig. 1 of Ref [1])

Download Full Size | PDF

The PhC cavity pattern consists of a modified L1 cavity design (Fig. 3(a)) with lattice constants varying between a = 130nm-160nm. The cavity was designed using finite-difference time-domain method (Lumerical FDTD). The parameters associated with this cavity are r/a = 0.3 and h/a = 0.93, where r refers to the radius of the holes, h is the height of the slab and a refers to the lattice constant of the hexagonal lattice. The six holes surrounding the center defect are smaller (r/a = 0.25). The four holes along ‘y’ direction were shifted outward from their nominal lattice position, Δy = 0.15a while the two holes along x where not shifted (Δx = 0). This design exhibited a reasonably high Q factor (~3000) while providing a high electric field intensity distribution in the defect region (Fig. 3(b)) at a frequency of ~0.36 a/λ. Figure 3(c) shows an SEM image of structure fabricated in III-nitride using EBL pattering approach described in Fig. 2.

 figure: Fig. 3

Fig. 3 (a) Schematic illustration of the shifted-hole L1 PhC cavity indicating vertical shifts (Δy) of top and bottom holes of the cavity and horizontal shifts (Δx) of the two side holes. (b) Simulated E-filed intensity profile of the cavity mode. (c) SEM of a fabricated device. The black scale bar represents a physical length of 500 nm.

Download Full Size | PDF

3. Optical characterization

Fabricated devices were optically characterized at room-temperature using an ultraviolet micro-photoluminescence (µ-PL) setup. Our excitation source was a quadrupled Nd:YAGlaser with a peak emission wavelength of 266 nm, a 10 kHz repetition rate and pulse lengths of ~500 ps. The laser’s peak power density was adjusted using a tunable neutral density filter. A 50X Mitutoyo deep-UV objective focused the laser to a spot size of nearly 5 µm onto the sample. The device was imaged using the same objective and directed into a 300 mm focal length spectrometer with an attached liquid nitrogen cooled CCD camera.

Figure 4 shows room-temperature photoluminescence (PL) taken from an unpatterned, unetched region (Fig. 4(a)) and an unpatterned, PEC-etched region (Fig. 4(b)) of the sample. The PL was collected after various PEC etch durations. As shown in Fig. 4(a), the unetched area produces broad emission peaks associated with the SQW and MQW structures (center emission wavelengths near 410 nm and 460 nm, respectively). As the PEC etch duration lengthens, no significant differences in emitted PL are observed in the unetched reference region. However, (Fig. 4(b)), illustrates a different result for the unpatterned, PEC-etched region. Before any etching takes place (t = 0 min.), two broad peaks are again observed near wavelengths of 410 nm and 460 nm. However, as the etch duration lengthens, the broad peak associated with the broad MQW emission reduces; the reduction in PL intensity results from the MQW structure being selectively etched away. Following 150 minutes of etching, emission from the MQW structure is almost unobservable, indicating the material has been removed. Despite significant etching of the MQW structure, emission associated with the SQW is uninhibited by the PEC etch or its corresponding durations. This result is a consequence of selecting the appropriate etch wavelength: carriers can be excited and contribute to surface oxidation in the MQW layers but not within the SQW. This preferential behavior is enabled by the narrow laser linewidth. Consequently, a membrane-like structure has been selectively formed through PEC etching. These results suggest an interesting procedure for selectively fabricating membrane structures in pre-fabricated III-nitride PhCs while preserving active layers of lower In composition(s). Also, by using higher In content in the sacrificial region, the active region can also contain higher In concentration, thereby enabling longer wavelength emission.

 figure: Fig. 4

Fig. 4 Room-temperature PL of an (a) unetched reference region and (b) PEC etched reference region after various etch durations. The dashed line in each figure coincides with the wavelength of the laser excitation source.

Download Full Size | PDF

Figure 5 shows room-temperature μ-PL measurements taken from various sample regions following the complete PEC etch duration. PL is collected from an unetched reference region (red), etched reference region (purple) and etched PhC device region (black). The upper inset in the figure shows a bright-field optical micro-graph (OM) of the sample. Each measured region is indicated with a dot of the corresponding color. For the unetched reference region, we again observe broad PL centered near wavelengths of 410 nm and 460 nm, associated with peak SQW and MQW emission, respectively. However, for the etched reference region, the MQW peak is unobservable due to selective etching of the MQW structure. Finally, PL collected from a PhC device region reveals the presence of a strong resonance near a wavelength of 385 nm. This resonant peak is a cavity mode of the L1 PhC and was unobservable preceding PEC etching. Further, the broad emission associated with the embedded SQW is clearly visible in the spectrum. Thus, our fabrication strategy enables wide tuning of cavity resonances matched to InGaN quantum dot emitters, spanning from the near-UV into the far-blue regions of the visible spectrum. Such wide-range tunability would be wholly unobtainable in a membrane composed entirely of GaN. However, it should be noted that the associated cavity resonance is detuned from peak SQW emission, and thus detunedform our theoretical calculations. This discrepancy is likely due to fabrication imperfections resulting from hard mask degradation, as evidenced from asymmetric as well as reduced diameter holes in the fabricated structure (Fig. 3(c)). In the future, a thicker SiO2 hard mask with better proximity correction in EBL pattering should help improve the spectral positioning of our cavity mode.

 figure: Fig. 5

Fig. 5 (a) schematic of PL measurement of the PhC cavity (b) Room-temperature PL emission from various regions on a sample. Red represents an unetched reference region, purple represents an etched reference region and black repre-sents an etched device region. The inset shows an optical microscope image of the sample. Markers of corresponding color indicate each region. Scale bar in inset is 5 μm.

Download Full Size | PDF

4. Summary

We have demonstrated a top-down fabrication strategy for preferential membrane formation in a pre-patterned III-nitride photonic crystal (PhC) cavity. Making use of a bandgap selective PEC wet-etching technique, we were able to selectively remove a multiple quantum well (MQW) heterostructure without affecting a single quantum well (SQW) of lower In concentration. Room-temperature micro-photoluminescence (μ-PL) measurements confirmed the selective removal of the MQW structure and preservation of an embedded SQW. Moreover, PhC cavity resonances, unobserved before PEC etching, were present in the final fabricated devices. Our results indicate an interesting method for creating III-nitride membranes of specified thickness that contain QWs with large emission bandwidths. Further, the embedded QW could be used for subsequent quantum-size-controlled photoelectrochemical (QSC-PEC) etching, which has been shown to create quantum dots (QDs) of controlled size and placement [1,39]. Thus, by spectrally tuning the second PEC etch wavelength to an observed cavity resonance, it is possible to incorporate QDs into a pre-fabricated PhC cavity. We believe that our top-down fabrication approach offers much opportunity for the seamless, spectral and spatial coupling of QDs to high-Q photonic crystal cavities for the III-nitride material system.

Acknowledgements

We note co-author Dr. Daniel D. Koleske who was responsible for the III-nitride epitaxial growth of the material studied in this work sadly passed away during the preparation of this manuscript.

We thank Dr. Andrew Allerman and Prof. Michelle Povinelli for useful discussions. This work was primarily supported by Sandia’s Laboratory Directed Research and Development Program. Portions of this work were performed at the Center for Integrated Nanotechnologies (CINT), a U.S. Department of Energy, Office of Basic Energy Sciences user facility. Sandia National Laboratories is a multimission laboratory managed and operated by National Technology & Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell International Inc., for the U.S. Department of Energy’s National Nuclear Security Administration under contract DE-NA0003525. The views expressed in the article do not necessarily represent the views of the U.S. Department of Energy or the United States Government.

References

1. A. J. Fischer, P. D. Anderson, D. D. Koleske, and G. Subramania, “Deterministic Placement of Quantum-Size Controlled Quantum Dots for Seamless Top-Down Integration,” ACS Photonics 4(9), 2165–2170 (2017). [CrossRef]  

2. J. Y. Tsao, M. H. Crawford, M. E. Coltrin, A. J. Fischer, D. D. Koleske, G. S. Subramania, G. T. Wang, J. J. Wierer, and R. F. Karlicek Jr., “Toward Smart and Ultra-efficient Solid-State Lighting,” Adv. Opt. Mater. 2(9), 809–836 (2014). [CrossRef]  

3. M. E. Coltrin, A. M. Armstrong, I. Brener, W. W. Chow, M. H. Crawford, A. J. Fischer, D. F. Kelley, D. D. Koleske, L. J. Lauhon, J. E. Martin, M. Nyman, E. F. Schubert, L. E. Shea-Rohwer, G. Subramania, J. Y. Tsao, G. T. Wang, J. J. Wierer Jr., and J. B. Wright, “Energy Frontier Research Center for Solid-State Lighting Science: Exploring New Materials Architectures and Light Emission Phenomena,” J. Phys. Chem. C 118(25), 13330–13345 (2014). [CrossRef]  

4. Q. Li, J. B. Wright, W. W. Chow, T. S. Luk, I. Brener, L. F. Lester, and G. T. Wang, “Single-mode GaN nanowire lasers,” Opt. Express 20(16), 17873–17879 (2012). [CrossRef]   [PubMed]  

5. L. Zhang, C.-H. Teng, P.-C. Ku, and H. Deng, “Site-controlled InGaN/GaN single-photon-emitting diode,” Appl. Phys. Lett. 108(15), 153102 (2016). [CrossRef]  

6. S. Kako, M. Holmes, S. Sergent, M. Bürger, D. J. As, and Y. Arakawa, “Single-photon emission from cubic GaN quantum dots,” Appl. Phys. Lett. 104(1), 011101 (2014). [CrossRef]  

7. S. Deshpande, J. Heo, A. Das, and P. Bhattacharya, “Electrically driven polarized single-photon emission from an InGaN quantum dot in a GaN nanowire,” Nat. Commun. 4(1), 1675 (2013). [CrossRef]   [PubMed]  

8. M. A. Bandres, S. Wittek, G. Harari, M. Parto, J. Ren, M. Segev, D. N. Christodoulides, and M. Khajavikhan, “Topological insulator laser: Experiments,” Science 359(6381), eaar4005 (2018). [CrossRef]   [PubMed]  

9. S. Barik, A. Karasahin, C. Flower, T. Cai, H. Miyake, W. DeGottardi, M. Hafezi, and E. Waks, “A topological quantum optics interface,” Science 359(6376), 666–668 (2018). [CrossRef]   [PubMed]  

10. J. B. Wright, S. Liu, G. T. Wang, Q. Li, A. Benz, D. D. Koleske, P. Lu, H. Xu, L. Lester, T. S. Luk, I. Brener, and G. Subramania, “Multi-Colour Nanowire Photonic Crystal Laser Pixels,” Sci. Rep. 3(1), 2982 (2013). [CrossRef]   [PubMed]  

11. I. Shusuke, K. Katsumi, A. Ryuichi, K. Akihiko, and S. Shuichi, “Optically Pumped Green (530–560 nm) Stimulated Emissions from InGaN/GaN Multiple-Quantum-Well Triangular-Lattice Nanocolumn Arrays,” Appl. Phys. Express 4(5), 055001 (2011). [CrossRef]  

12. T. Kouno, K. Kishino, M. Sakai, Y. Inose, A. Kikuchi, and K. Ema, “Stimulated emission on two-dimensional distributed feedback scheme in triangular GaN nanocolumn arrays,” Electron. Lett. 46(9), 644–645 (2010). [CrossRef]  

13. P. D. Anderson, D. D. Koleske, M. L. Povinelli, and G. Subramania, “Improving emission uniformity and linearizing band dispersion in nanowire arrays using quasi-aperiodicity,” Opt. Mater. Express 7(10), 3634–3642 (2017). [CrossRef]  

14. Q. Li, K. R. Westlake, M. H. Crawford, S. R. Lee, D. D. Koleske, J. J. Figiel, K. C. Cross, S. Fathololoumi, Z. Mi, and G. T. Wang, “Optical performance of top-down fabricated InGaN/GaN nanorod light emitting diode arrays,” Opt. Express 19(25), 25528–25534 (2011). [CrossRef]   [PubMed]  

15. W. Y. Fu, K. K.-Y. Wong, and H. W. Choi, “Room temperature photonic crystal band-edge lasing from nanopillar array on GaN patterned by nanosphere lithography,” J. Appl. Phys. 107(6), 063104 (2010). [CrossRef]  

16. T. Kouno, K. Kishino, K. Yamano, and A. Kikuchi, “Two-dimensional light confinement in periodic InGaN/GaN nanocolumn arrays and optically pumped blue stimulated emission,” Opt. Express 17(22), 20440–20447 (2009). [CrossRef]   [PubMed]  

17. F. La China, F. Intonti, N. Caselli, F. Lotti, A. Vinattieri, N. Vico Triviño, J. F. Carlin, R. Butté, N. Grandjean, and M. Gurioli, “Vectorial near-field imaging of a GaN based photonic crystal cavity,” Appl. Phys. Lett. 107(10), 101110 (2015). [CrossRef]  

18. C.-H. Lin, J.-Y. Wang, C.-Y. Chen, K.-C. Shen, D.-M. Yeh, Y.-W. Kiang, and C. C. Yang, “A GaN photonic crystal membrane laser,” Nanotechnology 22(2), 025201 (2011). [CrossRef]   [PubMed]  

19. L. Chang, C. Hou, Y. Ting, C. Chen, C. Hsu, J. Chang, C. Lee, G. Chen, and J. Chyi, “Laser emission from GaN photonic crystals,” Appl. Phys. Lett. 89(7), 071116 (2006). [CrossRef]  

20. Y. S. Choi, K. Hennessy, R. Sharma, E. Haberer, Y. Gao, S. P. DenBaars, S. Nakamura, E. L. Hu, and C. Meier, “GaN blue photonic crystal membrane nanocavities,” Appl. Phys. Lett. 87(24), 243101 (2005). [CrossRef]  

21. P.-H. Weng, T.-T. Wu, T.-C. Lu, and S.-C. Wang, “Threshold gain analysis in GaN-based photonic crystal surface emitting lasers,” Opt. Lett. 36(10), 1908–1910 (2011). [CrossRef]   [PubMed]  

22. J.-Y. Kim, M.-K. Kwon, S.-J. Park, S. H. Kim, and K.-D. Lee, “Enhancement of light extraction from GaN-based green light-emitting diodes using selective area photonic crystal,” Appl. Phys. Lett. 96(25), 251103 (2010). [CrossRef]  

23. S. Kawashima, T. Kawashima, Y. Nagatomo, Y. Hori, H. Iwase, T. Uchida, K. Hoshino, A. Numata, and M. Uchida, “GaN-based surface-emitting laser with two-dimensional photonic crystal acting as distributed-feedback grating and optical cladding,” Appl. Phys. Lett. 97(25), 251112 (2010). [CrossRef]  

24. J. J. Wierer, A. David, and M. M. Megens, “III-nitride photonic-crystal light-emitting diodes with high extraction efficiency,” Nat. Photonics 3(3), 163–169 (2009). [CrossRef]  

25. T. A. Truong, L. M. Campos, E. Matioli, I. Meinel, C. J. Hawker, C. Weisbuch, and P. M. Petroff, “Light extraction from GaN-based light emitting diode structures with a noninvasive two-dimensional photonic crystal,” Appl. Phys. Lett. 94(2), 023101 (2009). [CrossRef]  

26. H. Matsubara, S. Yoshimoto, H. Saito, Y. Jianglin, Y. Tanaka, and S. Noda, “GaN Photonic-Crystal Surface-Emitting Laser at Blue-Violet Wavelengths,” Science 319(5862), 445–447 (2008). [CrossRef]   [PubMed]  

27. A. Fischer, G. Subramania, Y.-J. Lee, T. Luk, J. Wendt, and D. Koleske, “Improved InGaN LED Efficiency using Photonic Crystal Patterning and Surface Plasmon Enhanced Emission,” ECS Trans. 13, 5–11 (2008).

28. K. Nozaki, A. Shinya, S. Matsuo, Y. Suzaki, T. Segawa, T. Sato, Y. Kawaguchi, R. Takahashi, and M. Notomi, “Ultralow-power all-optical RAM based on nanocavities,” Nat. Photonics 6(4), 248–252 (2012). [CrossRef]  

29. D. A. Fuhrmann, S. M. Thon, H. Kim, D. Bouwmeester, P. M. Petroff, A. Wixforth, and H. J. Krenner, “Dynamic modulation of photonic crystal nanocavities using gigahertz acoustic phonons,” Nat. Photonics 5(10), 605–609 (2011). [CrossRef]  

30. M. Toishi, D. Englund, A. Faraon, and J. Vucković, “High-brightness single photon source from a quantum dot in a directional-emission nanocavity,” Opt. Express 17(17), 14618–14626 (2009). [CrossRef]   [PubMed]  

31. K. Rivoire, A. Faraon, and J. Vuckovic, “Gallium phosphide photonic crystal nanocavities in the visible,” Appl. Phys. Lett. 93(6), 063103 (2008). [CrossRef]  

32. K. Nozaki, S. Kita, and T. Baba, “Room temperature continuous wave operation and controlled spontaneous emission in ultrasmall photonic crystal nanolaser,” Opt. Express 15(12), 7506–7514 (2007). [CrossRef]   [PubMed]  

33. G. Subramania, S. Y. Lin, J. R. Wendt, and J. M. Rivera, “Tuning the microcavity resonant wavelength in a two-dimensional photonic crystal by modifying the cavity geometry,” Appl. Phys. Lett. 83, 4491–4493 (2003). [CrossRef]  

34. Y. Akahane, T. Asano, B.-S. Song, and S. Noda, “High-Q photonic nanocavity in a two-dimensional photonic crystal,” Nature 425(6961), 944–947 (2003). [CrossRef]   [PubMed]  

35. A. A. Erchak, D. J. Ripin, S. Fan, P. Rakich, J. D. Joannopoulos, E. P. Ippen, G. S. Petrich, and L. A. Kolodziejski, “Enhanced coupling to vertical radiation using a two-dimensional photonic crystal in a semiconductor light-emitting diode,” Appl. Phys. Lett. 78(5), 563–565 (2001). [CrossRef]  

36. N. Vico Triviño, G. Rossbach, U. Dharanipathy, J. Levrat, A. Castiglia, J. F. Carlin, K. A. Atlasov, R. Butté, R. Houdré, and N. Grandjean, “High quality factor two dimensional GaN photonic crystal cavity membranes grown on silicon substrate,” Appl. Phys. Lett. 100(7), 071103 (2012). [CrossRef]  

37. T. S. Kao, T. T. Wu, C. W. Tsao, J. H. Lin, D. W. Lin, S. J. Huang, T. C. Lu, H. C. Kuo, S. C. Wang, and Y. K. Su, “Light Emission Characteristics of Nonpolar $a$ -Plane GaN-Based Photonic Crystal Defect Cavities,” IEEE J. Quantum Electron. 52, 1–7 (2016). [CrossRef]  

38. X. Xiao, A. J. Fischer, M. E. Coltrin, P. Lu, D. D. Koleske, G. T. Wang, R. Polsky, and J. Y. Tsao, “Photoelectrochemical etching of epitaxial InGaN thin films: self-limited kinetics and nanostructuring,” Electrochim. Acta 162, 163–168 (2015). [CrossRef]  

39. X. Xiao, A. J. Fischer, G. T. Wang, P. Lu, D. D. Koleske, M. E. Coltrin, J. B. Wright, S. Liu, I. Brener, G. S. Subramania, and J. Y. Tsao, “Quantum-Size-Controlled Photoelectrochemical Fabrication of Epitaxial InGaN Quantum Dots,” Nano Lett. 14(10), 5616–5620 (2014). [CrossRef]   [PubMed]  

40. E. D. Haberer, R. Sharma, A. R. Stonas, S. Nakamura, S. P. DenBaars, and E. L. Hu, “Removal of thick (>100nm) InGaN layers for optical devices using band-gap-selective photoelectrochemical etching,” Appl. Phys. Lett. 85(5), 762–764 (2004). [CrossRef]  

41. P. D. Anderson and G. Subramania, “Unidirectional edge states in topological honeycomb-lattice membrane photonic crystals,” Opt. Express 25(19), 23293–23301 (2017). [CrossRef]   [PubMed]  

42. S. Barik, H. Miyake, W. DeGottardi, E. Waks, and M. Hafezi, “Two-dimensionally confined topological edge states in photonic crystals,” New J. Phys. 18(11), 113013 (2016). [CrossRef]  

43. L.-H. Wu and X. Hu, “Scheme for Achieving a Topological Photonic Crystal by Using Dielectric Material,” Phys. Rev. Lett. 114(22), 223901 (2015). [CrossRef]   [PubMed]  

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 (5)

Fig. 1
Fig. 1 Schematic showing the MOCVD epitaxial growth structure. The inset shows the single quantum well (SQW) active layer and the multiple quantum well (MQW) sacrificial layer.
Fig. 2
Fig. 2 Device fabrication: (a) PECVD of SiO2 (b) EBL patterning and development (c) Flourine based dry etch (d) Chlorine based dry etch and (e) PEC etching. Light green indicates GaN, dark green indicates SiO2, blue indicates PMMA, a single black line represents a SQW structure and a series of black lines represents a MQW structure. (f) Schematic of the QSC-PEC etching setup. A tunable picosecond laser (blue) excites carriers in a III-nitride sample (green) that is partially submerged in an electrolyte solution (clear).(figure adapted from Fig. 1 of Ref [1])
Fig. 3
Fig. 3 (a) Schematic illustration of the shifted-hole L1 PhC cavity indicating vertical shifts (Δy) of top and bottom holes of the cavity and horizontal shifts (Δx) of the two side holes. (b) Simulated E-filed intensity profile of the cavity mode. (c) SEM of a fabricated device. The black scale bar represents a physical length of 500 nm.
Fig. 4
Fig. 4 Room-temperature PL of an (a) unetched reference region and (b) PEC etched reference region after various etch durations. The dashed line in each figure coincides with the wavelength of the laser excitation source.
Fig. 5
Fig. 5 (a) schematic of PL measurement of the PhC cavity (b) Room-temperature PL emission from various regions on a sample. Red represents an unetched reference region, purple represents an etched reference region and black repre-sents an etched device region. The inset shows an optical microscope image of the sample. Markers of corresponding color indicate each region. Scale bar in inset is 5 μm.
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.