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

High conductivity n-Al0.6Ga0.4N by ammonia-assisted molecular beam epitaxy for buried tunnel junctions in UV emitters

Open Access Open Access

Abstract

Highly doped n-Al0.6Ga0.4N can be used to form tunnel junctions (TJs) on deep ultraviolet (UVC) LEDs and markedly increase the light extraction efficiency (LEE) compared to the use of p-GaN/p-AlGaN. High quality Al0.6Ga0.4N was grown by NH3-assisted molecular beam epitaxy (NH3 MBE) on top of AlN on SiC substrate. The films were crack free under scanning electron microscope (SEM) for the thickness investigated (up to 1 µm). X-ray diffraction reciprocal space map scan was used to determine the Al composition and the result is in close agreement with atom probe tomography (APT) measurements. By varying the growth parameters including growth rate, and Si cell temperature, n-Al0.6Ga0.4N with an electron density of 4×1019 /cm3 and a resistivity of 3 mΩ·cm was achieved. SIMS measurement shows that a high Si doping level up to 2×1020 /cm3 can be realized using a Si cell temperature of 1450 °C and a growth rate of 210 nm/hr. Using a vanadium-based annealed contact, ohmic contact with a specific resistance of 10−6 Ω·cm2 was achieved as determined by circular transmission line measurement (CTLM). Finally, the n-type AlGaN regrowth was done on MOCVD grown UVC LEDs to form UVC TJ LED. The sample was processed into thin film flip chip (TFFC) configuration. The emission wavelength is around 278 nm and the excess voltage of processed UV LED is around 4.1 V.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

AlGaN alloys have a wide range of applications in optical and electronic devices due to their attractive physical properties such as tunable wide bandgap, strong piezoelectricity, and a high breakdown field. Recently, deep ultraviolet (UVC) LEDs based on AlGaN materials have attracted significant interest for its disinfection applications such as water purification [14] and sterilization of critical care patients rooms [5,6].

Realizing highly conductive n-AlGaN with controlled doping is essential to achieve high performance devices. UV LEDs, for example, can benefit from the use of n-AlGaN based tunnel junctions (TJs) [7]. This is because the efficiency of UV LEDs is primarily limited by the light extraction efficiency which is usually below 30% [817]. TJs can potentially double the LEE by enabling the use of highly reflective MgF2-based mirrors in thin-film flip-chip UVC LEDs [17,1821] and thus eliminate the need to use absorptive p-GaN [22,23]. TJs also facilitate current spreading and hole injection in UV LEDs [7,16,17].

Si-doped AlGaN has several challenges in comparison to than Si:GaN for several reasons. First, the ionization energy of Si dopant in AlGaN is higher than in GaN and increases with increased AlN content [23,2434]. Ionization energy of Si in AlxGa1-xN increases from ∼15 meV for x = 0 to 60-280 meV for x = 1, with a value for 25-30 meV for x = 0.65 [24,35]. Second, Si also shows self-compensation at high doping level for AlxGa1-xN with x > 0.5, leading to reduced electron concentration and increased resistivity as Si doping level increases above a certain value (also known as the “knee behavior”) [23,24,2628,3640]. At high Al composition, compensating occurs, either through DX center formation for the Si dopants [23,28,35,4050] or through group III (i.e., Al or Ga) vacancy complexes [2628,49,5156,5662]. Finally, unlike GaN, the electron mobility in AlGaN is impacted by alloy scattering.

Most of the studies on Si-doped AlGaN are done by metal organic chemical vapor deposition (MOCVD) [2427,30,3537,52,6371], with some studies by plasma-assisted molecular beam epitaxy (MBE) [23,28,31,4143], and few by NH3-assisted MBE (NH3 MBE) [29,38]. NH3 MBE has several advantages such as avoiding repassivation of p-type layers, the capacity to grow n-GaN with excellent transport properties [72], incorporating high levels of Si without compromising the material quality [73], sharp p+/n+ interface [74], less memory effect compared to MOCVD especially for Mg [75], and achieving high-performance TJ devices [74,7679]. It is thus of our interest to investigate Si-doping of AlGaN by NH3 MBE.

In this study, we investigated n-type doping of AlxGa1-xN by Si with an Al fraction (x) around 60% by NH3 MBE. This Al composition was of interest to this study since negligible absorption of UVC wavelength occurs at this composition [23]. Besides, the maximum realizable carrier concentration can decreases with increased Al content x [23,24,26,28,30,36,42,43,52,63,65]. The relationship between n-AlGaN quality and growth parameters including growth rate and Si cell temperature was investigated. Using the optimized n-AlGaN condition, we demonstrate low n-AlGaN contact resistance and UVC TJ LEDs (around 278 nm).

2. Experimental

The templates used for this study were AlN grown by MOCVD on Si-face c-plane 6H-silicon carbide (SiC) substrates. The templates were crack-free, with a threading dislocation density (TDD) around 109 /cm2. Details on the low TDD AlN growth can be found elsewhere [80,81]. Before growth, the templates were coated with 500 nm of Ti on the back side by electron beam evaporation to assist the heat spreading and pyrometer temperature reading. Then, they were diced into 1 cm × 1 cm squares, solvent-cleaned, and indium-bonded to silicon wafers for transfer into the MBE. The samples were baked for 1 hour at 400 °C in vacuum before growth. The growth was done by Veeco 930 NH3-assisted MBE with NH3 as the group V source and solid effusion cells for Al, Ga, and Si. The flux for each group III elemental beam was performed using an ion gauge under NH3-free conditions. Temperature was measured using a pyrometer calibrated using the melting point of Al. The sample morphology was monitored in real time using reflection high-energy electron diffraction (RHEED). The growth temperature was set at 780 °C with a NH3 flow of 200 sccm. The group III fluxes were adjusted to realize a growth rate of ∼100-400 nm/hr and an Al content around 60%. The silicon doping level was controlled by growth parameters such as the Si cell temperature.

The composition and relaxation of the AlGaN films were evaluated on the PANalynatic Pixcel Diffractometer using reciprocal spacing mapping technique (RSM). The measurement was taken in an asymmetric scattering geometry with Cu K-α1 X-ray source, 2-bounce Ge (220) monochromator with a PIXcel3D detector. The lattice parameters of AlGaN was determined from the separation of its (1 0 $\bar{1}$ 5) peak and SiC (1 0 $\bar{1}$ 15) in RSM assuming the SiC is completely relaxed. Based on the Vegard’s law for the AlGaN lattice parameters and elastic constants, its alloy composition and in-plane strain can be therefore calculated. The thickness of the AlGaN films was determined by the spacing of fringes of the on-axis ω-2θ (0002) scan. Optical microscopy, atomic force microscopy (AFM), and scanning electron microscope (SEM) were used to characterize the morphology of samples. Atom probe tomography (APT) with a Cameca local electrode atom probe (LEAP) 3000X HR was used to evaluate the three-dimensional distribution of Al and Ga atoms and the alloy distribution [82].

The n-type doping optimization was carried out by growing two series of n-AlGaN samples with various Si cell temperature. Series 1 samples were grown at 109 nm/hr while Series 2 samples were grown at 210 nm/hr. The n-type doping of the AlGaN samples was evaluated by room temperature Hall measurements in a Van der Pauw geometry. The AlN on SiC templates underneath the n-AlGaN layer were confirmed to be insulating. To rule out the effect of interface charge, we also confirmed that a sample with undoped AlGaN grown on the AlN/SiC template is highly insulating. The Si doping level, as well as the C, H, and O impurity levels were measured by Cameca IMS 7f Auto secondary ion mass microscopy (SIMS). The morphology of n-AlGaN samples were evaluated by atomic force microscopy (AFM). Contacts to the n-AlGaN were formed using a V/Al/V/Au (10/150/20/300 nm, respectively) metal stack. The samples were annealed in N2 at 720 °C for 30 s to achieve an Ohmic contact. Circular transmission line measurement (CTLM) pattern was formed on the n-AlGaN with the annealed contact to measure the contact resistance.

Finally, the optimized n-AlGaN condition was used to form a UVC hybrid tunnel junction (TJ) LED. Mesa-defined MOCVD-grown UVC LED with 0.5 nm p-GaN on top was patterned with SiO2 to allow selective area growth of MBE-grown n-Al0.59Ga0.41N on top of both the 0.5 nm MOCVD-grown p-GaN and n-AlGaN region. The Mg concentration in the 0.5 nm p-GaN and underlying 50 nm of AlGaN:Mg layers was ∼1×1020 cm−3 and ∼2-3×1018 cm−3, respectively. Only 0.5 nm of p-GaN was used to minimize the absorption by GaN. The lower 4 nm n++ layer forms the n-side of the TJ and was grown with high Si cell temperature (1450 °C) to achieve high donor concentration ND. The 500 nm n+ layer was grown with the optimized condition with Si cell temperature at 1425 °C for the highest n-type carrier concentration and the lowest resistivity to facilitate current spreading while maintaining low resistivity. The upper 4 nm n++ layer serves as the contact layer and was grown with the same condition as the lower n++ layer. The growth rate used for the regrowth was 426 nm/hr. The sample was then processed into thin film flip-chip (TFFC) LEDs using the vanadium-based contact as both n and p-side contact. Details of the MOCVD epitaxial structure and the TFFC UV LED fabrication techniques can be found elsewhere [18,82,83]. The electroluminescence (EL) spectrum of the UV TJ LED was acquired after regrowth and before further processing. The on-wafer testing was done using a fiber optic and UV-Vis spectrometer (Ocean Optics) at 10 mA with an integration time of 50 ms. The fiber was placed above the LED sample, in proximity to the device. The IV characteristics of the processed UV TJ LED was measured and compared with the reference UV LEDs without the TJ regrowth and with 5nm p-GaN.

3. Results

3.1 Epitaxial growth of AlGaN

The epitaxial growth of undoped AlGaN was carried out to map out the growth condition to achieve the desired crystal quality and alloy composition (∼60%). Figure 1(a) shows the SEM image of a sample with a 110 nm thick AlxGa1-xN layer, x = 0.55, and a growth rate of 218 nm/hr. The sample was crack free and no pits were observed. No cracks were observed for the grown undoped AlGaN and n-AlGaN up to a thickness of 1 µm (the highest thickness investigated) as shown in the optical microscope (OM) image in Fig. 1(b). Figure 1(c) shows the AFM image of MOCVD-grown AlN on SiC template. The surface morphology was smooth with a root mean square (RMS) roughness of 0.3 nm. Figure 1(d) shows the AFM image of 100 nm AlxGa1-xN layer grown by MBE on the template with x=0.65 and a growth rate of 216 nm/hr. Very few (∼107 /cm2) pits were observed on the sample surface. The granular features in the AFM image of AlGaN are commonly observed in NH3 MBE films [72]. The RMS roughness was 0.48 nm. The comparison of the sample’s AFM images before and after the AlGaN growth suggests that the MBE growth did not degrade the surface roughness.

 figure: Fig. 1.

Fig. 1. (a). SEM image of a crack-free undoped AlGaN layer (110 nm), with Al=55.5%, and a growth rate of 218 nm/hr. (b). Optical microscope image of a crack-free Al0.59Ga0.61N layer (1 µm) grown at 210 nm/hr with both doped and undoped sub-layers showing no cracks. (c). AFM image of the AlN on SiC template, which shows step flow growth. The RMS roughness is 0.3 nm (d). AFM image of 100 nm undoped AlGaN layer grown by MBE on the template with Al%=65.5% and a growth rate of 216 nm/hr. The surface was smooth and the RMS roughness was 0.48 nm.

Download Full Size | PDF

Figure 2(a) shows the ω-2θ scan of 100 nm Al0.60Ga0.40N film grown at 218 nm/hr. The clear thickness fringes indicate high crystal quality and the spacing of the fringes was used to calculate the film thickness and growth rate. Figure 2(b) shows an RSM scan of 400 nm Al0.60Ga0.40N film grown with the same condition. The AlN layer is fully relaxed [80] while the AlGaN film for this sample is coherent with the AlN layer as determined by the identical Qx values for the off axis (10$\bar{1}$5) AlGaN (bottom) and AlN (top) peaks. The separation between the SiC and the AlGaN peak was used to calculate the Al% composition.

 figure: Fig. 2.

Fig. 2. (a). ω-2θ scan of 100 nm Al0.60Ga0.40N film grown at 218 nm/hr with clear thickness fringes in log scale. The spacings between the fringes were used to calculate the film thickness and growth rate. (b). RSM scan of 400 nm fully strained Al0.60Ga0.40N film grown at the same condition as in 2 (a). The separation between the SiC (in the middle) and the AlGaN peak (bottom) was used to calculate the Al% composition and relaxation ratio. The same Qx between the AlN (top) and AlGaN peak shows that the AlGaN film is fully strained.

Download Full Size | PDF

At fixed group III fluxes ΦIIIAl: 1.35×10−7 torr; ΦGa: 1.19×10−7 torr) and an NH3 flow rate of 200 sccm, the growth temperature was varied in the range of 750-800 °C. Figure 3 shows that the Al% composition and growth rate were rather stable in this temperature range. The Al% changed within 3.5% and the growth rate varied within a 10% range.

 figure: Fig. 3.

Fig. 3. Alloy composition and growth rate dependence with the growth temperature.

Download Full Size | PDF

The relationship between alloy composition and group III fluxes were also investigated. At a NH3 flow rate of 200 sccm and a growth temperature of 800 °C, the Al flux ΦAl was varied while keeping the Ga flux fixed at 6.08×10−8 torr. The measured Al% values are shown in Table 1.

Tables Icon

Table 1. The relationship between group III fluxes and Al%. $S = \frac{{{S_{Al}}}}{{{S_{Ga}}}}$ is the ratio of incorporation factor between the two group III elements.

When the other growth parameters were kept the same, the Al content was determined by the group III fluxes. An incorporation factor (often referred to as a sticking coefficient) s can be defined as:

$$s = \frac{N}{\Phi }$$
where N is the flux of atoms incorporated into the AlGaN layer, and $\mathrm{\Phi }$ was the supplied flux for this element. The Al% composition is thus determined by:
$$Al\%= \frac{{{N_{Al}}}}{{{N_{Al}} + {N_{Ga}}}} = \frac{{{\Phi _{Al}} \times {s_{Al}}}}{{{\Phi _{Al}} \times {s_{Al}} + {\Phi _{Ga}} \times {s_{Ga}}}} = \frac{{{\Phi _{Al}} \times S}}{{{\Phi _{Al}} \times S + {\Phi _{Ga}}}}$$
where $S = \frac{{{s_{Al}}}}{{{s_{Ga}}}}$

The value of S was quantified for a given growth temperature and used to calculate the Al content for different Al and Ga fluxes when the change in group III flux were small. The calculated Al content value agreed well with the measured value for all samples investigated in this work. Typical range of S is 1.15-2.20, suggesting that Al atoms more readily incorporate into the AlGaN film than the Ga atoms which is consistent with the higher Al-N bond energy (2.2 eV) compared to the Ga-N bond (1.93 eV) [84].

APT measurements of the AlGaN layer are shown in Fig. 4. The Al% composition measured by APT is 59% and is in close agreement with the 58% Al composition calculated by XRD. The Al content was constant in the growth direction for the thickness studied (60 nm). The alloy was found to be random, and no Al rich clusters were observed [82].

 figure: Fig. 4.

Fig. 4. 3D APT reconstruction of 60 nm of Al0.59Ga0.41N showing Al and Ga atoms. (b) 2D distribution of the Aluminum fraction measured from the dashed rectangle shown in (a).

Download Full Size | PDF

3.2 Doping optimization for Si-doped AlGaN

The doping optimization of the n-AlGaN was done by varying the growth rate and Si cell temperature while keeping the Al content and other growth parameters fixed. Figure 5(a) shows the n-type carrier concentration, resistivity, and mobility of n-AlGaN (Series 1) with an Al content of 61.6% and a growth rate of 109 nm/hr. The Si cell temperature was varied from 1375 to 1450 °C with 25 °C steps. The Si-doped AlGaN grown with a Si cell temperature of 1450 °C was insulating and no valid Hall data was acquired. Using a Si cell temperature of 1400 °C, both the lowest resistivity (10.5 mΩ·cm) and the highest n-type carrier concentration (1.2×1019 /cm3) was achieved. Similar doping series (Series 2) was done at a higher growth rate of 210 nm/hr for AlGaN with the Al content fixed at 59.4% as shown in Fig. 5(b). At this growth rate, the optimum Si cell temperature was at 1425 °C. The carrier concentration was 4×1019 /cm3, the mobility was 48 cm2·V·s−1, and the resistivity was 3 mΩ·cm. The morphology for this sample was measured by AFM and shown in Fig. 6. No pits were observed on the surface. The RMS roughness was 0.6 nm.

 figure: Fig. 5.

Fig. 5. Doping optimization for n-AlxGa1-xN by varying the Si cell temperature using a growth rate of (a) 109 nm/hr at x = 0.62 (series 1) and (b) 210 nm/hr at x 0.59 (series 2). The n-AlGaN grown at 109 nm/hr with the Si cell at 1450 °C was resistive and thus not shown in (a). The star shape denotes the optimum growth condition for the 210 nm/hr growth rate.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. AFM image of the optimum n-AlGaN growth condition (star sign in Fig. 5). The RMS roughness is 0.6 nm. There were no pits on the surface.

Download Full Size | PDF

Increasing the AlGaN:Si growth rate resulted in a wider optimal doping window, higher carrier concentration, and lower resistivities. Both doping profiles showed the “knee behavior” commonly observed for Si:AlGaN [23,24,2628,3640]. As the Si cell temperature increased, the resistivity decreased while the carrier concentration increased at the low doping region and the opposite happened at the high doping region. The optimum growth condition was at the Si cell temperature of 1400 °C and 1425 °C for Series 1 and 2, respectively (refer to Fig. 5).

The Si, C, H, and O concentration (Fig. 7) were measured by SIMS for Series 2. The Si concentration increases with increased Si cell temperature. At a Si cell temperature of 1425 and 1450 °C, Si concentration of 1×1020 and 2×1020 /cm3 were achieved. The oxygen and carbon level were at 4-7×1017 and 2×1016 /cm3, respectively, and were stable at all Si cell temperatures. When the Si cell temperature increased from 1400 °C to 1425 °C and then 1450 °C, the hydrogen incorporation increased from 1.5×1017 /cm3 to 7.5×1017 /cm3 and then to 1.8×1018 /cm3. Suggesting that there could be increased H or H-related compensation at high doping levels.

 figure: Fig. 7.

Fig. 7. Si, C, H, and O concentration in n-AlGaN series 2 samples measured by secondary ion mass microscopy (SIMS). The n-type carrier concentration is included for comparison. The star shape denotes the optimum doping condition.

Download Full Size | PDF

CTLM measurements were done on a V/Al/V/Au contact deposited on a Si-doped Al0.59Ga0.41N. The growth of the AlGaN layers was done at the same time as the TJ regrowth and have the same epi structure as the MBE part in Fig. 9(b). The carrier concentration and resistivity for the n+ layer was 1.9×1019 /cm3 and 12 mΩ·cm respectively. Figure 8 shows the IV curve for the 10 nm gap CTLM pattern. The contact changed from Schottky to Ohmic after annealing in N2 gas at 720 °C for 30 s. The resistance of the ohmic contact was 7.9 Ω as measured by 2-point probe. CTLM measured on the sample showed that the specific contact resistance was 5.3×10−6-8.9×10−6 Ω·cm2. The transfer length was 2.3 µm. The sheet resistance was around 125 Ω/□. The contact resistance is low enough that it does not introduce any appreciable voltage penalty in the IV characteristics of the different LEDs in the later part of the study.

 figure: Fig. 8.

Fig. 8. IV curve for CTLM pattern with 10 nm gap on n-AlGaN with vanadium-based contact before and after annealing in N2 gas at 720 °C for 30 s.

Download Full Size | PDF

 figure: Fig. 9.

Fig. 9. (a). Schematic of the UV TJ LED. The upper image shows the top view of the LED. Both the n and p side epitaxial structure were terminated with MBE grown n-AlGaN. SiO2 was used for both the mask for selective area growth and isolation layer for the LED. The lower image shows the cross section of the LED. Selective area MBE n-AlGaN regrowth was done on the on top of the n-AlGaN and p-GaN region of the MOCVD grown UVC LED sample to form hybrid TJ UVC LED. The SiO2 layer served as the mask for the selective area growth. (b). Schematic of the epitaxial structure for the p-side of the UVC tunnel junction (TJ) LED.

Download Full Size | PDF

3.3 UVC TJ LEDs

Schematic of the top and cross-section view of the UVC TJ LED after regrowth is shown in Fig. 9(a). Figure 9(b) shows the schematic of the TJ UV LED epitaxial structure.

Figure 10(a) shows the EL spectrum of the TJ LED wafer before LED processing, where the EL was obtained from the UV light emitted below indium dots and its surrounding perimeter (100 µm in diameter). The EL emission wavelength (around 278 nm) confirmed hole injection through the TJ structure into the LED active region. The EL intensity was low due to the low extraction efficiency and substrate absorption.

 figure: Fig. 10.

Fig. 10. (a) Electroluminescence spectra of the UVC LED with TJ (after regrowth and before further processing). Note the low intensity is likely due to the absorption by SiC. (b) I-V characteristics for the processed TJ LED and an identical LED without regrowth but with 5 nm of p-GaN on top of the MOCVD structure.

Download Full Size | PDF

Figure 10(b) shows the IV characteristics of the processed TJ LED as compared to an identical LED without regrowth but with 5 nm of p-GaN on top of the MOCVD structure, which was reported elsewhere [81]. The TJ LED turned on, and had an excess voltage of 4.1 V at 20 A/cm2 compared to the LED with 5 nm of p-GaN. The excess voltage is high compared to blue TJ LEDs grown with similar approach (<1 V) [74,85,86]. This is partially due to the wider band gap of AlGaN, the higher activation energy of carriers, and the lower doping concentration (NA and ND) in AlGaN material. More optimizations are needed to reduce the excess voltage.

4. Conclusion

In this work, we demonstrated high carrier concentration, low resistivity n-AlGaN by NH3 MBE with the Al% around 60%. The films had smooth morphology and were crack free for the thickness investigated (up to 1 µm). XRD RSM scan was used to determine the alloy composition and the result agrees well with the APT result. The higher growth rate had a wider optimum doping window when varying the Si cell temperature. At a growth rate of 210 nm/hr, with a Si cell temperature of 1425 °C, n-AlGaN with a carrier concentration of 4×1019 /cm3 and a resistivity of 3 mΩ·cm was achieved. SIMS measurements shows that the Si incorporation increases with increased Si cell temperature and the hydrogen incorporation increased at high cell temperature. Ohmic contacts were achieved on the n-AlGaN using annealed vanadium-based contact stack. The optimized n-AlGaN was employed to show that a transparent tunnel junction can significantly reduce the thickness of p-GaN in UVC LEDs, however, further research is needed to reduce the excess voltage (4.1 V).

Funding

National Science Foundation (DMR 1720256, ECS-03357650); KACST-KAUST-UCSB Solid State Lighting Program SSLP; Solid State Lighting and Energy Electronics Center, University of California Santa Barbara.

Acknowledgements

The authors would like to thank Dr. Thomas Mates for help performing SIMS measurements. The authors would also like to thank Dr. Youli Li for insightful conversations about XRD. A portion of this work was carried out in the UCSB nanofabrication facility, with support from the NSF NNIN network, as well as the UCSB Materials Research Laboratory (MRL), which is supported by the MRSEC Program of the NSF under Award No. DMR 1720256; a member of the NSF-funded Materials Research Facilities Network (www.mrfn.org).

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. G. Y. Lui, D. Roser, R. Corkish, N. J. Ashbolt, and R. Stuetz, “Point-of-use water disinfection using ultraviolet and visible light-emitting diodes,” Sci. Total Environ. 553, 626–635 (2016). [CrossRef]  

2. K. Song, M. Mohseni, and F. Taghipour, “Application of ultraviolet light-emitting diodes (UV-LEDs) for water disinfection: A review,” Water Res. 94, 341–349 (2016). [CrossRef]  

3. A. Kheyrandish, M. Mohseni, and F. Taghipour, “Development of a method for the characterization and operation of UV-LED for water treatment,” Water Res. 122, 570–579 (2017). [CrossRef]  

4. S. Vilhunen, H. Särkkä, and M. Sillanpää, “Ultraviolet light-emitting diodes in water disinfection,” Environ. Sci. Pollut. Res. 16(4), 439–442 (2009). [CrossRef]  

5. P. Murphy, L. Kang, M. Fleming, C. Atkinson, R. Pryor, K. Cooper, E. Godbout, M. P. Stevens, M. Doll, and G. Bearman, “Effect of ultraviolet-C light disinfection at terminal patient discharge on hospital-acquired infections in bone marrow transplant and oncology units,” Am. J. Infect. Control. 48(6), 705–707 (2020). [CrossRef]  

6. “Coming Of Age: UV-C LED Technology Update,” https://www.wateronline.com/doc/coming-of-age-uv-c-led-technology-update-0001.

7. Y. Zhang, Z. Jamal-Eddine, and S. Rajan, “Recent progress of tunnel junction-based ultra-violet light emitting diodes,” Jpn. J. Appl. Phys. 58(SC), SC0805 (2019). [CrossRef]  

8. T. Takano, T. Mino, J. Sakai, N. Noguchi, K. Tsubaki, and H. Hirayama, “Deep-ultraviolet light-emitting diodes with external quantum efficiency higher than 20% at 275 nm achieved by improving light-extraction efficiency,” Appl. Phys. Express 10(3), 031002 (2017). [CrossRef]  

9. Y. Guo, Y. Zhang, J. Yan, H. Xie, L. Liu, X. Chen, M. Hou, Z. Qin, J. Wang, and J. Li, “Light extraction enhancement of AlGaN-based ultraviolet light-emitting diodes by substrate sidewall roughening,” Appl. Phys. Lett. 111(1), 011102 (2017). [CrossRef]  

10. J. Yun and H. Hirayama, “Investigation of the light-extraction efficiency in 280 nm AlGaN-based light-emitting diodes having a highly transparent p-AlGaN layer,” J. Appl. Phys. 121(1), 013105 (2017). [CrossRef]  

11. H.-Y. Ryu, I.-G. Choi, H.-S. Choi, and J.-I. Shim, “Investigation of Light Extraction Efficiency in AlGaN Deep-Ultraviolet Light-Emitting Diodes,” Appl. Phys. Express 6(6), 062101 (2013). [CrossRef]  

12. J. Rass and N. Lobo-Ploch, “Optical Polarization and Light Extraction from UV LEDs,” in III-Nitride Ultraviolet Emitters: Technology and Applications, M. Kneissl and J. Rass, eds., Springer Series in Materials Science (Springer International Publishing, 2016), pp. 137–170.

13. J. R. Grandusky, J. Chen, S. R. Gibb, M. C. Mendrick, C. G. Moe, L. Rodak, G. A. Garrett, M. Wraback, and L. J. Schowalter, “270 nm Pseudomorphic Ultraviolet Light-Emitting Diodes with Over 60 mW Continuous Wave Output Power,” Appl. Phys. Express 6(3), 032101 (2013). [CrossRef]  

14. T. Inazu, S. Fukahori, C. Pernot, M. H. Kim, T. Fujita, Y. Nagasawa, A. Hirano, M. Ippommatsu, M. Iwaya, T. Takeuchi, S. Kamiyama, M. Yamaguchi, Y. Honda, H. Amano, and I. Akasaki, “Improvement of Light Extraction Efficiency for AlGaN-Based Deep Ultraviolet Light-Emitting Diodes,” Jpn. J. Appl. Phys. 50, 122101 (2011). [CrossRef]  

15. B. T. Tran and H. Hirayama, “Growth and Fabrication of High External Quantum Efficiency AlGaN-Based Deep Ultraviolet Light-Emitting Diode Grown on Pattern Si Substrate,” Sci. Rep. 7(1), 12176 (2017). [CrossRef]  

16. Y. Zhang, A. A. Allerman, S. Krishnamoorthy, F. Akyol, M. W. Moseley, A. M. Armstrong, and S. Rajan, “Enhanced light extraction in tunnel junction-enabled top emitting UV LEDs,” Appl. Phys. Express 9(5), 052102 (2016). [CrossRef]  

17. Y. Zhang, S. Krishnamoorthy, F. Akyol, S. Bajaj, A. A. Allerman, M. W. Moseley, A. M. Armstrong, and S. Rajan, “Tunnel-injected sub-260 nm ultraviolet light emitting diodes,” Appl. Phys. Lett. 110(20), 201102 (2017). [CrossRef]  

18. B. K. SaifAddin, A. Almogbel, C. J. Zollner, H. Foronda, A. Alyamani, A. Albadri, M. Iza, S. Nakamura, S. P. DenBaars, and J. S. Speck, “Fabrication technology for high light-extraction ultraviolet thin-film flip-chip (UV TFFC) LEDs grown on SiC,” Semicond. Sci. Technol. 34(3), 035007 (2019). [CrossRef]  

19. B. K. SaifAddin (2018), “Development of Deep Ultraviolet (UV-C) Thin-Film Light-Emitting Diodes Grown on SiC,” [Doctoral dissertation, University of California, Santa Barbara]. ProQuest Dissertations Publishing.

20. J. K. Kim, J. W. Lee, D.-Y. Kim, J. H. Park, E. F. Schubert, J. Kim, and Y.-I. Kim, “An elegant route to overcome fundamentally-limited light extraction in AlGaN deep-ultraviolet light-emitting diodes: preferential outcoupling of strong in-plane emission (Conference Presentation),” in UV and Higher Energy Photonics: From Materials to Applications (International Society for Optics and Photonics, 2016), Vol. 9926, p. 99260F.

21. H. Kim, S.-N. Lee, Y. Park, K.-K. Kim, J. S. Kwak, and T.-Y. Seong, “Light extraction enhancement of GaN-based light emitting diodes using MgF2/Al omnidirectional reflectors,” J. Appl. Phys. 104(5), 053111 (2008). [CrossRef]  

22. J. F. Muth, J. H. Lee, I. K. Shmagin, R. M. Kolbas, H. C. Casey, B. P. Keller, U. K. Mishra, and S. P. DenBaars, “Absorption coefficient, energy gap, exciton binding energy, and recombination lifetime of GaN obtained from transmission measurements,” Appl. Phys. Lett. 71(18), 2572–2574 (1997). [CrossRef]  

23. J. Hwang, W. J. Schaff, L. F. Eastman, S. T. Bradley, L. J. Brillson, D. C. Look, J. Wu, W. Walukiewicz, M. Furis, and A. N. Cartwright, “Si doping of high-Al-mole fraction AlxGa1−xN alloys with rf plasma-induced molecular-beam-epitaxy,” Appl. Phys. Lett. 81(27), 5192–5194 (2002). [CrossRef]  

24. Y. Taniyasu, M. Kasu, and N. Kobayashi, “Intentional control of n-type conduction for Si-doped AlN and AlxGa1−xN (0.42⩽x&lt;1),” Appl. Phys. Lett. 81(7), 1255–1257 (2002). [CrossRef]  

25. R. Collazo, S. Mita, J. Xie, A. Rice, J. Tweedie, R. Dalmau, and Z. Sitar, “Progress on n-type doping of AlGaN alloys on AlN single crystal substrates for UV optoelectronic applications,” Phys. Status Solidi C 8(7-8), 2031–2033 (2011). [CrossRef]  

26. M. L. Nakarmi, K. H. Kim, K. Zhu, J. Y. Lin, and H. X. Jiang, “Transport properties of highly conductive n-type Al-rich AlxGa1−xN(x>0.7),” Appl. Phys. Lett. 85(17), 3769–3771 (2004). [CrossRef]  

27. Y. Taniyasu, M. Kasu, and T. Makimoto, “Electrical conduction properties of n-type Si-doped AlN with high electron mobility (>100cm2V−1s−1),” Appl. Phys. Lett. 85(20), 4672–4674 (2004). [CrossRef]  

28. T. Ive, O. Brandt, H. Kostial, K. J. Friedland, L. Däweritz, and K. H. Ploog, “Controlled n-type doping of AlN:Si films grown on 6H-SiC(0001) by plasma-assisted molecular beam epitaxy,” Appl. Phys. Lett. 86(2), 024106 (2005). [CrossRef]  

29. M. Ahoujja, J. L. McFall, Y. K. Yeo, R. L. Hengehold, and J. E. Van Nostrand, “Electrical and optical investigation of MBE grown Si-doped AlxGa1−xN as a function of Al mole fraction up to 0.5,” Mater. Sci. Eng., B 91-92, 285–289 (2002). [CrossRef]  

30. A. Y. Polyakov, N. B. Smirnov, A. V. Govorkov, M. G. Mil’vidskii, J. M. Redwing, M. Shin, M. Skowronski, D. W. Greve, and R. G. Wilson, “Properties of Si donors and persistent photoconductivity in AlGaN,” Solid-State Electron. 42(4), 627–635 (1998). [CrossRef]  

31. M. Stutzmann, O. Ambacher, A. Cros, M. S. Brandt, H. Angerer, R. Dimitrov, N. Reinacher, T. Metzger, R. Höpler, D. Brunner, F. Freudenberg, R. Handschuh, and C. Deger, “Properties and applications of MBE grown AlGaN,” Mater. Sci. Eng., B 50(1-3), 212–218 (1997). [CrossRef]  

32. S. N. Mohammad and H. Morkoç, “Progress and prospects of group-III nitride semiconductors,” Prog. Quantum Electron. 20(5-6), 361–525 (1996). [CrossRef]  

33. V. W. L. Chin, T. L. Tansley, and T. Osotchan, “Electron mobilities in gallium, indium, and aluminum nitrides,” J. Appl. Phys. 75(11), 7365–7372 (1994). [CrossRef]  

34. Y.-H. Liang and E. Towe, “Progress in efficient doping of high aluminum-containing group III-nitrides,” Appl. Phys. Rev. 5(1), 011107 (2018). [CrossRef]  

35. P. Pampili and P. J. Parbrook, “Doping of III-nitride materials,” Mater. Sci. Semicond. Process. 62, 180–191 (2017). [CrossRef]  

36. M. C. Wagener, G. R. James, and F. Omnès, “Intrinsic compensation of silicon-doped AlGaN,” Appl. Phys. Lett. 83(20), 4193–4195 (2003). [CrossRef]  

37. F. Mehnke, X. T. Trinh, H. Pingel, T. Wernicke, E. Janzén, N. T. Son, and M. Kneissl, “Electronic properties of Si-doped AlxGa1−xN with aluminum mole fractions above 80%,” J. Appl. Phys. 120(14), 145702 (2016). [CrossRef]  

38. B. Borisov, V. Kuryatkov, Y. Kudryavtsev, R. Asomoza, S. Nikishin, D. Y. Song, M. Holtz, and H. Temkin, “Si-doped AlxGa1−xN(0.56⩽x⩽1) layers grown by molecular beam epitaxy with ammonia,” Appl. Phys. Lett. 87(13), 132106 (2005). [CrossRef]  

39. J. Pyeon, J. Kim, M. Jeon, K. Ko, E. Shin, and O. Nam, “Self-compensation effect in Si-doped Al0.55Ga0.45N layers for deep ultraviolet applications,” Jpn. J. Appl. Phys. 54(5), 051002 (2015). [CrossRef]  

40. I. Bryan, Z. Bryan, S. Washiyama, P. Reddy, B. Gaddy, B. Sarkar, M. H. Breckenridge, Q. Guo, M. Bobea, J. Tweedie, S. Mita, D. Irving, R. Collazo, and Z. Sitar, “Doping and compensation in Al-rich AlGaN grown on single crystal AlN and sapphire by MOCVD,” Appl. Phys. Lett. 112(6), 062102 (2018). [CrossRef]  

41. R. Zeisel, M. W. Bayerl, S. T. B. Goennenwein, R. Dimitrov, O. Ambacher, M. S. Brandt, and M. Stutzmann, “DX-behavior of Si in AlN,” Phys. Rev. B 61(24), R16283 (2000). [CrossRef]  

42. D. Korakakis, H. M. Ng, K. F. Ludwig, and T. D. Moustakas, “Doping Studies of n- and p-Type AlxGa1-xN Grown by ECR-Assisted MBE,” Mater. Res. Soc. Symp. Proc. 449, 233–238 (1996). [CrossRef]  

43. M. Hermann, F. Furtmayr, A. Bergmaier, G. Dollinger, M. Stutzmann, and M. Eickhoff, “Highly Si-doped AlN grown by plasma-assisted molecular-beam epitaxy,” Appl. Phys. Lett. 86(19), 192108 (2005). [CrossRef]  

44. X. Thang Trinh, D. Nilsson, I. G. Ivanov, E. Janzén, A. Kakanakova-Georgieva, and N. Tien Son, “Negative-U behavior of the Si donor in Al0.77Ga0.23N,” Appl. Phys. Lett. 103(4), 042101 (2013). [CrossRef]  

45. C. Skierbiszewski, T. Suski, M. Leszczynski, M. Shin, M. Skowronski, M. D. Bremser, and R. F. Davis, “Evidence for localized Si-donor state and its metastable properties in AlGaN,” Appl. Phys. Lett. 74(25), 3833–3835 (1999). [CrossRef]  

46. C. G. Van de Walle, “DX-center formation in wurtzite and zinc-blende AlxGa1-xN,” Phys. Rev. B 57(4), R2033–R2036 (1998). [CrossRef]  

47. L. Gordon, J. L. Lyons, A. Janotti, and C. G. Van de Walle, “Hybrid functional calculations of DX centers in AlN and GaN,” Phys. Rev. B 89(8), 085204 (2014). [CrossRef]  

48. N. T. Son, M. Bickermann, and E. Janzén, “Shallow donor and DX states of Si in AlN,” Appl. Phys. Lett. 98(9), 092104 (2011). [CrossRef]  

49. Q. Yan, A. Janotti, M. Scheffler, and C. G. Van de Walle, “Origins of optical absorption and emission lines in AlN,” Appl. Phys. Lett. 105(11), 111104 (2014). [CrossRef]  

50. J. S. Harris, B. E. Gaddy, R. Collazo, Z. Sitar, and D. L. Irving, “Oxygen and silicon point defects in Al0.65Ga0.35N,” Phys. Rev. Mater. 3(5), 054604 (2019). [CrossRef]  

51. J. S. Harris, J. N. Baker, B. E. Gaddy, I. Bryan, Z. Bryan, K. J. Mirrielees, P. Reddy, R. Collazo, Z. Sitar, and D. L. Irving, “On compensation in Si-doped AlN,” Appl. Phys. Lett. 112(15), 152101 (2018). [CrossRef]  

52. M. D. McCluskey, N. M. Johnson, C. G. Van de Walle, D. P. Bour, M. Kneissl, and W. Walukiewicz, “Metastability of Oxygen Donors in AlGaN,” Phys. Rev. Lett. 80(18), 4008–4011 (1998). [CrossRef]  

53. T. Mattila and R. M. Nieminen, “Point-defect complexes and broadband luminescence in GaN and AlN,” Phys. Rev. B 55(15), 9571–9576 (1997). [CrossRef]  

54. C. Stampfl and C. G. Van de Walle, “Theoretical investigation of native defects, impurities, and complexes in aluminum nitride,” Phys. Rev. B 65(15), 155212 (2002). [CrossRef]  

55. I. Gorczyca, N. E. Christensen, and A. Svane, “Influence of hydrostatic pressure on cation vacancies in GaN, AlN, and GaAs,” Phys. Rev. B 66(7), 075210 (2002). [CrossRef]  

56. J. Slotte, F. Tuomisto, K. Saarinen, C. G. Moe, S. Keller, and S. P. DenBaars, “Influence of silicon doping on vacancies and optical properties of AlxGa1−xN thin films,” Appl. Phys. Lett. 90(15), 151908 (2007). [CrossRef]  

57. C. Stampfl and C. G. Van de Walle, “Doping of Al¬xGa1−xN,” Appl. Phys. Lett. 72(4), 459–461 (1998). [CrossRef]  

58. W. Walukiewicz, “Intrinsic limitations to the doping of wide-gap semiconductors,” Phys. B 302-303, 123–134 (2001). [CrossRef]  

59. A. Uedono, K. Tenjinbayashi, T. Tsutsui, Y. Shimahara, H. Miyake, K. Hiramatsu, N. Oshima, R. Suzuki, and S. Ishibashi, “Native cation vacancies in Si-doped AlGaN studied by monoenergetic positron beams,” J. Appl. Phys. 111(1), 013512 (2012). [CrossRef]  

60. T. Onuma, S. F. Chichibu, A. Uedono, T. Sota, P. Cantu, T. M. Katona, J. F. Keading, S. Keller, U. K. Mishra, S. Nakamura, and S. P. DenBaars, “Radiative and nonradiative processes in strain-free AlxGa1−xN films studied by time-resolved photoluminescence and positron annihilation techniques,” J. Appl. Phys. 95(5), 2495–2504 (2004). [CrossRef]  

61. S. F. Chichibu, H. Miyake, Y. Ishikawa, M. Tashiro, T. Ohtomo, K. Furusawa, K. Hazu, K. Hiramatsu, and A. Uedono, “Impacts of Si-doping and resultant cation vacancy formation on the luminescence dynamics for the near-band-edge emission of Al0.6Ga0.4N films grown on AlN templates by metalorganic vapor phase epitaxy,” J. Appl. Phys. 113(21), 213506 (2013). [CrossRef]  

62. S. Washiyama, P. Reddy, B. Sarkar, M. H. Breckenridge, Q. Guo, P. Bagheri, A. Klump, R. Kirste, J. Tweedie, S. Mita, Z. Sitar, and R. Collazo, “The role of chemical potential in compensation control in Si:AlGaN,” J. Appl. Phys. 127(10), 105702 (2020). [CrossRef]  

63. K. B. Nam, J. Li, M. L. Nakarmi, J. Y. Lin, and H. X. Jiang, “Achieving highly conductive AlGaN alloys with high Al contents,” Appl. Phys. Lett. 81(6), 1038–1040 (2002). [CrossRef]  

64. P. Cantu, S. Keller, U. K. Mishra, and S. P. DenBaars, “Metalorganic chemical vapor deposition of highly conductive Al0.65Ga0.35N films,” Appl. Phys. Lett. 82(21), 3683–3685 (2003). [CrossRef]  

65. M. D. Bremser, W. G. Perry, T. Zheleva, N. V. Edwards, O. H. Nam, N. Parikh, D. E. Aspnes, and R. F. Davis, “Growth, Doping and Characterization of AlxGa1−xN Thin Film Alloys on 6H-SiC (0001) Substrates,” MRS Internet J. Nitride Semicond. Res. 1, 8 (1996). [CrossRef]  

66. T. M. Al tahtamouni, A. Sedhain, J. Y. Lin, and H. X. Jiang, “Si-doped high Al-content AlGaN epilayers with improved quality and conductivity using indium as a surfactant,” Appl. Phys. Lett. 92(9), 092105 (2008). [CrossRef]  

67. K. Zhu, M. L. Nakarmi, K. H. Kim, J. Y. Lin, and H. X. Jiang, “Silicon doping dependence of highly conductive n-type Al0.7Ga0.3N,” Appl. Phys. Lett. 85(20), 4669–4671 (2004). [CrossRef]  

68. A. Kakanakova-Georgieva, D. Nilsson, X. T. Trinh, U. Forsberg, N. T. Son, and E. Janzén, “The complex impact of silicon and oxygen on the n-type conductivity of high-Al-content AlGaN,” Appl. Phys. Lett. 102(13), 132113 (2013). [CrossRef]  

69. Y. Shimahara, H. Miyake, K. Hiramatsu, F. Fukuyo, T. Okada, H. Takaoka, and H. Yoshida, “Growth of High-Quality Si-Doped AlGaN by Low-Pressure Metalorganic Vapor Phase Epitaxy,” Jpn. J. Appl. Phys. 50(9), 095502 (2011). [CrossRef]  

70. S. Keller, P. Cantu, C. Moe, Y. Wu, S. Keller, U. K. Mishra, J. S. Speck, and S. P. DenBaars, “Metalorganic Chemical Vapor Deposition Conditions for Efficient Silicon Doping in High Al-Composition AlGaN Films,” Jpn. J. Appl. Phys. 44(10), 7227–7233 (2005). [CrossRef]  

71. A. S. Almogbel, C. J. Zollner, B. K. Saifaddin, M. Iza, J. Wang, Y. Yao, M. Wang, H. Foronda, I. Prozheev, F. Tuomisto, and A. Albadri, “Growth of highly conductive Al-rich AlGaN: Si with low group-III vacancy concentration,” AIP Adv. 11(9), 095119 (2021). [CrossRef]  

72. E. C. H. Kyle, S. W. Kaun, P. G. Burke, F. Wu, Y.-R. Wu, and J. S. Speck, “High-electron-mobility GaN grown on free-standing GaN templates by ammonia-based molecular beam epitaxy,” J. Appl. Phys. 115(19), 193702 (2014). [CrossRef]  

73. L. Lugani, M. Malinverni, S. Tirelli, D. Marti, E. Giraud, J.-F. Carlin, C. R. Bolognesi, and N. Grandjean, “n+-GaN grown by ammonia molecular beam epitaxy: Application to regrown contacts,” Appl. Phys. Lett. 105(20), 202113 (2014). [CrossRef]  

74. E. C. Young, B. P. Yonkee, F. Wu, S. H. Oh, S. P. DenBaars, S. Nakamura, and J. S. Speck, “Hybrid tunnel junction contacts to III–nitride light-emitting diodes,” Appl. Phys. Express 9(2), 022102 (2016). [CrossRef]  

75. Y. Ohba and A. Hatano, “A study on strong memory effects for Mg doping in GaN metalorganic chemical vapor deposition,” J. Cryst. Growth 145(1-4), 214–218 (1994). [CrossRef]  

76. B. P. Yonkee, E. C. Young, C. Lee, J. T. Leonard, S. P. DenBaars, J. S. Speck, and S. Nakamura, “Demonstration of a III-nitride edge-emitting laser diode utilizing a GaN tunnel junction contact,” Opt. Express 24(7), 7816–7822 (2016). [CrossRef]  

77. C. A. Forman, S. Lee, E. C. Young, J. A. Kearns, D. A. Cohen, J. T. Leonard, T. Margalith, S. P. DenBaars, and S. Nakamura, “Continuous-wave operation of m-plane GaN-based vertical-cavity surface-emitting lasers with a tunnel junction intracavity contact,” Appl. Phys. Lett. 112(11), 111106 (2018). [CrossRef]  

78. M. Malinverni, D. Martin, and N. Grandjean, “InGaN based micro light emitting diodes featuring a buried GaN tunnel junction,” Appl. Phys. Lett. 107(5), 051107 (2015). [CrossRef]  

79. A. J. Mughal, E. C. Young, A. I. Alhassan, J. Back, S. Nakamura, J. S. Speck, and S. P. DenBaars, “Polarization-enhanced InGaN/GaN-based hybrid tunnel junction contacts to GaN p–n diodes and InGaN LEDs,” Appl. Phys. Express 10(12), 121006 (2017). [CrossRef]  

80. C. J. Zollner, A. Almogbel, Y. Yao, B. K. SaifAddin, F. Wu, M. Iza, S. P. DenBaars, J. S. Speck, and S. Nakamura, “Reduced dislocation density and residual tension in AlN grown on SiC by metalorganic chemical vapor deposition,” Appl. Phys. Lett. 115(16), 161101 (2019). [CrossRef]  

81. H. M. Foronda, F. Wu, C. Zollner, M. E. Alif, B. Saifaddin, A. Almogbel, M. Iza, S. Nakamura, S. P. DenBaars, and J. S. Speck, “Low threading dislocation density aluminum nitride on silicon carbide through the use of reduced temperature interlayers,” J. Cryst. Growth 483, 134–139 (2018). [CrossRef]  

82. B. K. SaifAddin, A. S. Almogbel, C. J. Zollner, F. Wu, B. Bonef, M. Iza, S. Nakamura, S. P. DenBaars, and J. S. Speck, “AlGaN Deep-Ultraviolet Light-Emitting Diodes Grown on SiC Substrates,” ACS Photonics 7(3), 554–561 (2020). [CrossRef]  

83. B. K. Saifaddin, M. Iza, H. Foronda, A. Almogbel, C. J. Zollner, F. Wu, A. Alyamani, A. Albadri, S. Nakamura, S. P. DenBaars, and J. S. Speck, “Impact of roughening density on the light extraction efficiency of thin-film flip-chip ultraviolet LEDs grown on SiC,” Opt. Express 27(16), A1074–A1083 (2019). [CrossRef]  

84. T. Kuech, Handbook of Crystal Growth: Thin Films and Epitaxy (Elsevier, 2014).

85. B. P. Yonkee, E. C. Young, J. T. Leonard, C. Lee, S. H. Oh, S. P. DenBaars, J. S. Speck, and S. Nakamura, “Hybrid MOCVD/MBE GaN tunnel junction LEDs with greater than 70% wall plug efficiency,” in 2016 Compound Semiconductor Week (CSW) [Includes 28th International Conference on Indium Phosphide Related Materials (IPRM) 43rd International Symposium on Compound Semiconductors (ISCS) (2016), p. 1.

86. J. Wang, E. C. Young, W. Y. Ho, B. Bonef, T. Margalith, and J. S. Speck, “III-nitride blue light-emitting diodes utilizing hybrid tunnel junction with low excess voltage,” Semicond. Sci. Technol. 35(12), 125026 (2020). [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 (10)

Fig. 1.
Fig. 1. (a). SEM image of a crack-free undoped AlGaN layer (110 nm), with Al=55.5%, and a growth rate of 218 nm/hr. (b). Optical microscope image of a crack-free Al0.59Ga0.61N layer (1 µm) grown at 210 nm/hr with both doped and undoped sub-layers showing no cracks. (c). AFM image of the AlN on SiC template, which shows step flow growth. The RMS roughness is 0.3 nm (d). AFM image of 100 nm undoped AlGaN layer grown by MBE on the template with Al%=65.5% and a growth rate of 216 nm/hr. The surface was smooth and the RMS roughness was 0.48 nm.
Fig. 2.
Fig. 2. (a). ω-2θ scan of 100 nm Al0.60Ga0.40N film grown at 218 nm/hr with clear thickness fringes in log scale. The spacings between the fringes were used to calculate the film thickness and growth rate. (b). RSM scan of 400 nm fully strained Al0.60Ga0.40N film grown at the same condition as in 2 (a). The separation between the SiC (in the middle) and the AlGaN peak (bottom) was used to calculate the Al% composition and relaxation ratio. The same Qx between the AlN (top) and AlGaN peak shows that the AlGaN film is fully strained.
Fig. 3.
Fig. 3. Alloy composition and growth rate dependence with the growth temperature.
Fig. 4.
Fig. 4. 3D APT reconstruction of 60 nm of Al0.59Ga0.41N showing Al and Ga atoms. (b) 2D distribution of the Aluminum fraction measured from the dashed rectangle shown in (a).
Fig. 5.
Fig. 5. Doping optimization for n-AlxGa1-xN by varying the Si cell temperature using a growth rate of (a) 109 nm/hr at x = 0.62 (series 1) and (b) 210 nm/hr at x 0.59 (series 2). The n-AlGaN grown at 109 nm/hr with the Si cell at 1450 °C was resistive and thus not shown in (a). The star shape denotes the optimum growth condition for the 210 nm/hr growth rate.
Fig. 6.
Fig. 6. AFM image of the optimum n-AlGaN growth condition (star sign in Fig. 5). The RMS roughness is 0.6 nm. There were no pits on the surface.
Fig. 7.
Fig. 7. Si, C, H, and O concentration in n-AlGaN series 2 samples measured by secondary ion mass microscopy (SIMS). The n-type carrier concentration is included for comparison. The star shape denotes the optimum doping condition.
Fig. 8.
Fig. 8. IV curve for CTLM pattern with 10 nm gap on n-AlGaN with vanadium-based contact before and after annealing in N2 gas at 720 °C for 30 s.
Fig. 9.
Fig. 9. (a). Schematic of the UV TJ LED. The upper image shows the top view of the LED. Both the n and p side epitaxial structure were terminated with MBE grown n-AlGaN. SiO2 was used for both the mask for selective area growth and isolation layer for the LED. The lower image shows the cross section of the LED. Selective area MBE n-AlGaN regrowth was done on the on top of the n-AlGaN and p-GaN region of the MOCVD grown UVC LED sample to form hybrid TJ UVC LED. The SiO2 layer served as the mask for the selective area growth. (b). Schematic of the epitaxial structure for the p-side of the UVC tunnel junction (TJ) LED.
Fig. 10.
Fig. 10. (a) Electroluminescence spectra of the UVC LED with TJ (after regrowth and before further processing). Note the low intensity is likely due to the absorption by SiC. (b) I-V characteristics for the processed TJ LED and an identical LED without regrowth but with 5 nm of p-GaN on top of the MOCVD structure.

Tables (1)

Tables Icon

Table 1. The relationship between group III fluxes and Al%. S = S A l S G a is the ratio of incorporation factor between the two group III elements.

Equations (2)

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

s = N Φ
A l % = N A l N A l + N G a = Φ A l × s A l Φ A l × s A l + Φ G a × s G a = Φ A l × S Φ A l × S + Φ G a
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.