A detailed study comparing defect incorporation between laser-assisted metal-organic chemical vapor deposition (MOCVD)-grown GaN and conventional low- and high-growth-rate MOCVD GaN was conducted. Using deep-level transient and optical spectroscopy, traps throughout the bandgap were characterized where traps were found at EC-0.25 eV, EC-0.57 eV, EC-0.72 eV, EC-0.9 eV, EC-1.35 eV, EC-2.6 eV, and EC-3.28 eV in all three samples. This indicates no new traps were observed in the laser-assisted MOCVD GaN sample. Overall, the trap concentrations in the laser-assisted MOCVD sample were ∼2× higher than the optimal low-growth-rate sample, but this is primarily due to the increase in gallium vacancy EC-2.6 eV and carbon-related EC-3.28 eV trap concentrations. The EC-0.9 eV trap concentration was ∼2× higher in the laser-assisted sample, so proton irradiation experiments were conducted to identify the physical source of this level. The results indicated this was a native point defect likely related to gallium interstitials. Overall, this study shows that the laser-assisted MOCVD growth method is promising for future thick, high-quality GaN epilayers after further growth optimizations.

Utilizing a CO2 laser to more efficiently decompose the ammonia precursor in metal organic chemical vapor deposition (MOCVD) growth has demonstrated high growth rates and promises lower manufacturing costs for vertical GaN transistors and diodes for multi-kV applications.1–5 Generally, vertical diode devices targeting a breakdown voltage greater than 4 kV require 1015–1016 cm−3 net n-type doping with thick drift layers to support the high electric fields.6–9 A high growth rate is desired to achieve thick drift layers, but typical MOCVD growth rates are only 2–3 μm/h.10–12 Laser-assisted MOCVD (LA-MOCVD) growth has achieved rates up to 84 μm/h and continues to improve as the optics and growth conditions are optimized.4,5 Controllable low n-type net doping (ND–NA) requires n-type doping (ND) and compensating trap concentrations (NA) as low as possible, so the net doping is easily targeted and less sensitive to variations in ND or NA. Additionally, compensating acceptors and traps can change occupancy as the Fermi level is swept above and below the trap levels due to biasing, which leads to time-dependent transistor instabilities including dynamic on-resistance, threshold voltage instability, and current collapse.13–16 Hence, identifying the traps in these materials, quantifying the trap concentrations, and identifying mechanisms to mitigate their formation or incorporation is critical to enable low-doped drift layers for vertical GaN transistors.

Trap characterization and growth studies of LA-MOCVD-grown GaN are very limited. Notably, Golgir et al. achieved a 25.8 μm/h growth rate and measured 5.9 × 108 cm−2 dislocation density and 369 cm2 V−1 s−1 mobility.5 Zhang et al. recently achieved GaN growth rates up to 5.2 μm/h, characterized impurity incorporation, and achieved a 604 cm2 V−1 s−1 mobility under the highest growth rate.1 

A higher growth rate with LA-MOCVD growth is possible with both pyrolytic and photolytic approaches.2 In the pyrolytic approach, the laser is used to heat the substrate surface and aid the chemical reaction locally or globally. On the other hand, the photolytic approach uses the laser to crack the precursors, specifically ammonia, in the gas stream to provide higher concentrations of elemental nitrogen to the growth surface.17 Using optical emission spectroscopy, it was shown that higher concentrations of NH2 and other NH3-related radicals are observed using a 9.219 μm CO2 laser beam in air, indicating enhanced cracking efficiency.4 The photolytic approach was chosen for the LA-MOCVD growth in this work as it allows higher growth rates while maintaining an independent control of growth conditions such as substrate temperature, growth pressure, etc.

In this work, an LA-MOCVD GaN sample utilizing the photolytic approach is compared with standard MOCVD samples with high and low growth rates to understand the effects of growth rate separately from that of the laser-assisted growth on the incorporation of deep levels. Trap energies and concentrations in these samples are characterized with deep-level transient and optical spectroscopy (DLTS/DLOS). Proton irradiation combined with DLTS characterization are employed to discern the intrinsic or extrinsic nature of the defects.

Three samples were grown by MOCVD and LA-MOCVD for this study. The first two were grown conventionally without laser assistance and were labeled low-growth-rate (LGR) and high-growth-rate (HGR) with growth rates of 2.0 and 5.2 μm/h, respectively. A third sample was grown with the laser at a growth rate of 2.8 μm/h and was labeled laser-assisted (LA). Growths for all three samples were done in a rotating-disk MOCVD reactor (Agnitron Technology Inc., Agilis MOCVD R&D System) at a growth pressure of 200 Torr and a substrate temperature of ∼960 °C, monitored by a thermocouple. The TMGa flow for the LGR sample was 68.6 μmol/min, and a higher TMGa flow of 203.9 μmol/min was applied for HGR and LA samples with an NH3 flow rate of 4 slm for all three samples. The silane dopant flow rate was controlled at 8.9, 17.7, and 14.7 pmol/min for LGR, LA, and HGR samples, respectively, to keep the doping concentration consistent among all samples. The LA sample was grown with a PRC 1500 W tunable CO2 laser system tuned to a wavelength of 9.219 μm and a power of 200 W for optimal ammonia cracking. The laser beam's shape was round with a diameter of 14 mm and was configured to be parallel to the substrate and wafer carrier to crack ammonia in the gas flow without impinging on the growth surface. The LA sample showed a significant reduction in growth rate compared to the HGR sample despite their same precursor flow rates, which resulted from significant gas phase parasitic reactions due to the round laser beam creating N radicals far from the growth surface. Since these growths, the laser beam has been shaped by focal lenses to have less vertical width so gas phase reactions would be suppressed,18 and defect studies of the latest growth method are ongoing. All three samples were grown on 2 in. unintentionally doped (UID) GaN-on-sapphire templates with similar growth structures, as shown in Fig. 1. A heavily doped n-type ([Si] ∼ 2 × 1018 cm−3) layer was grown on the substrate which enabled efficient lateral conductance and functioned as a base for good Ohmic contact. A lightly doped n-type ([Si] ∼ high-1016 cm−3) test layer was probed during the defect spectroscopy measurements with a thickness of 0.5 μm in the LGR sample and 2 μm in the HGR and LA samples. Since the defect spectroscopy measurements only measured the top ∼0.13 μm of the test layer, negligible influence of the test layer thickness has been observed in similar samples. For the contacts, 80 Å of Ni was deposited on the lightly doped test layer to form a 290 × 290 μm2 semitransparent Schottky contact to enable DLOS measurements. For the Ohmic contact, reactive ion etching with Cl chemistry was used to etch down to the n+ GaN layer where Ti/Al/Ni/Au (100/2000/200/2000 Å) were deposited.

FIG. 1.

Schottky diode structure for DLTS and DLOS characterization for the three samples. The semitransparent Schottky contact allows light to pass through during DLOS measurement. The test layer thickness “X” is 0.5 μm in the LGR sample and 2 μm in the HGR and LA samples.

FIG. 1.

Schottky diode structure for DLTS and DLOS characterization for the three samples. The semitransparent Schottky contact allows light to pass through during DLOS measurement. The test layer thickness “X” is 0.5 μm in the LGR sample and 2 μm in the HGR and LA samples.

Close modal

DLTS and DLOS were conducted to characterize traps throughout the GaN bandgap. DLTS was used to detect traps within ∼1 eV of the conduction band, and DLOS was used to detect deeper traps. For DLTS, the fill pulse was 0 V for 10 ms, and trap emission was measured at −0.5 V. DLTS was analyzed using the double boxcar method, where the majority traps cause positive peaks in a DLTS spectrum. For DLOS, the fill pulse was 0 V for 10 s, and then the device was biased to −0.5 V for the measurement phase with the shutter open for 300 s after a 100 s delay to allow thermal transients to subside. More details of the experimental apparatus and measurements are available in Refs. 19 and 20. Secondary ion mass spectrometry (SIMS) was conducted to detect extrinsic impurity density and aid the trap spectroscopy measurements.

To further understand the possible physical sources of the observed levels, proton irradiation was utilized to create native point defects. The LA sample was subject to 1.8 MeV proton irradiation with a 2 × 1013 cm−2 fluence at Sandia National Laboratories. DLTS was conducted before and after the proton irradiation to explore the effect of radiation on trap concentrations and gain insights as to whether the as-grown traps are intrinsic or extrinsic.

DLTS was conducted on each sample and is shown in Fig. 2. In a DLTS spectrum, each trap is represented by a peak of which the location shifts with rate windows. Each rate window characterizes the trap emission rate (en) that is the reciprocal of the lifetime (τn). From the peak location of each rate window, the Arrhenius plot is generated and used to calculate the trap energy (ET) and cross section (σn) based on the equation21 
1 τ n T 2 = ( v n T 1 / 2 ) ( N C T 3 / 2 ) σ n exp ( E C E T k T ) ,
(1)
where T is the temperature, EC–ET is the trap energy, k is Boltzmann's constant, vn is the electron thermal velocity, and NC is the conduction band effective density of states.
FIG. 2.

(a) DLTS spectra of LGR, LA, and HGR samples showing the peaks of EC-0.25 eV, EC-0.57 eV, EC-0.72 eV, and EC-0.9 eV traps. (b) Arrhenius plot shows that the traps were similar to some of the traps discovered in previous works.2,16,18,25,26 The EC-0.25 eV and EC-0.57 eV traps were in GaN bulk materials and Schottky devices, whereas EC-0.72 eV and EC-0.9 eV traps were seen in HEMT devices. (c) An example of fitting of LA to separate the EC-0.72 eV peak by Gaussian peak fitting. (d) DLTS of the LA sample after proton irradiation. Compared to Fig. 2(c), significant increase in the concentration of EC-0.72 eV and EC-0.9 eV traps indicates that these two traps may be related to intrinsic point defects.

FIG. 2.

(a) DLTS spectra of LGR, LA, and HGR samples showing the peaks of EC-0.25 eV, EC-0.57 eV, EC-0.72 eV, and EC-0.9 eV traps. (b) Arrhenius plot shows that the traps were similar to some of the traps discovered in previous works.2,16,18,25,26 The EC-0.25 eV and EC-0.57 eV traps were in GaN bulk materials and Schottky devices, whereas EC-0.72 eV and EC-0.9 eV traps were seen in HEMT devices. (c) An example of fitting of LA to separate the EC-0.72 eV peak by Gaussian peak fitting. (d) DLTS of the LA sample after proton irradiation. Compared to Fig. 2(c), significant increase in the concentration of EC-0.72 eV and EC-0.9 eV traps indicates that these two traps may be related to intrinsic point defects.

Close modal

Figure 2(a) shows three peaks in all samples. The high-temperature peak is asymmetric with a small shoulder near 350 K in the 4 s−1 rate window indicating the existence of a fourth level. The DLTS peaks have some overlap, so peak fitting was used for each DLTS rate window in each sample to extract more precise trap energies and concentrations, where an example is shown in Fig. 2(c). From the fitting results, Arrhenius plot is plotted in Fig. 2(b) where the extracted trap energies are EC-0.25 eV, EC-0.57 eV, EC-0.72 eV, and EC-0.9 eV. The clustering of traps on the Arrhenius plot suggest they are the same among all growth conditions. Therefore, the increased growth rate and the presence of the laser-assisted growth did not result in the formation of new or different traps. Additionally, the traps in the LA-MOCVD sample are all consistent with previously reported traps plotted in Fig. 2(b). The EC-0.25 eV and EC-0.57 eV traps are nearly ubiquitous in GaN and have been attributed to a VN-related defect and FeGa, respectively.22,23 Previous studies have shown that Fe and Fe-related trap incorporation can be minimized through improved wafer carrier cleaning and sample mounting, which has been used here to minimize Fe incorporation.24 The other two traps at EC-0.72 eV and EC-0.9 eV have been previously observed, but the physical sources of these traps need to be better understood.25–27 

Here, the trap concentrations are proportional to the DLTS peak height. Still, more accurate trap concentration extraction requires accounting for the volume where the traps are modulated, which differs from the depletion region volume and is referred to as Lambda effect correction.21 Depending on the trap level, doping, and measurement biases, the Lambda effect multiplier usually ranges from 1 to 10, and the concentration of deeper traps is more underestimated without this correction. The Lambda-corrected concentration of each trap is listed in Table I. The EC-0.25 eV and EC-0.57 eV traps had relatively low concentrations in the mid-1013 to low-1014 cm−3 range and varied less than 50% among all three samples. This indicated that the growth rate and laser-assisted growth had a minor impact on the incorporation of these traps. The EC-0.9 eV trap concentration in the LA sample was approximately 2× higher than the LGR sample but approximately 2× lower than the HGR sample. This trap may be more sensitive to certain variations in growth conditions. Overall, the trap concentrations, excluding the EC-0.9 eV trap, range from 100× lower to ∼10× lower than the necessary doping for vertical GaN devices (∼1 × 1016 cm−3), indicating these traps will have minimal impact on net doping/compensation and likely minimal impact on transistor stability. Because the EC-0.9 eV trap concentration is relatively high, proton irradiation experiments were conducted to help elucidate its physical source.

TABLE I.

Summary of each sample's DLTS and DLOS trap concentrations.

Trap energy EC-0.25 eV EC-0.57 eV EC-0.72 eV EC-0.9 eV EC-1.35 eV EC-2.6 eV EC-3.28 eV
LA (cm−3 5.5 × 1013  1.1 × 1014  1.8 × 1014  6.5 × 1014  3.4 × 1014  1.9 × 1015  2.0 × 1015 
HGR (cm−3 4.7 × 1013  1.4 × 1014  1.5 × 1014  1.4 × 1015  3.9 × 1014  2.5 × 1015  2.4 × 1015 
LGR (cm−3 6.9 × 1013  1.2 × 1014  5.7 × 1013  2.8 × 1014  2.6 × 1014  8.6 × 1014  8.8 × 1014 
Trap energy EC-0.25 eV EC-0.57 eV EC-0.72 eV EC-0.9 eV EC-1.35 eV EC-2.6 eV EC-3.28 eV
LA (cm−3 5.5 × 1013  1.1 × 1014  1.8 × 1014  6.5 × 1014  3.4 × 1014  1.9 × 1015  2.0 × 1015 
HGR (cm−3 4.7 × 1013  1.4 × 1014  1.5 × 1014  1.4 × 1015  3.9 × 1014  2.5 × 1015  2.4 × 1015 
LGR (cm−3 6.9 × 1013  1.2 × 1014  5.7 × 1013  2.8 × 1014  2.6 × 1014  8.6 × 1014  8.8 × 1014 

DLTS measurements were conducted on the LA sample before and after 1.8 MeV proton irradiation with a fluence of 2 × 1013 cm−2. Figure 2(d) shows the post-irradiation DLTS results where trap energies and concentrations were extracted by fitting. Radiation did not lead to any changes in trap energies as expected, and the trap concentrations were calculated again using the Lambda effect correction, which are summarized in Table II. Since the impact of irradiation on the EC-0.25 eV trap has already been investigated,28 the DLTS scan range was reduced starting at 200 K. From Table II, it is clear the EC-0.57 eV, EC-0.72 eV, and EC-0.9 eV trap concentrations increased by 1 × 1014, 3.2 × 1014, and 8.5 × 1014 cm−3, respectively. As the EC-0.57 eV level is likely an FeGa defect, it is not expected to have radiation dependence though similar dependence has previously been observed.28 Additionally, the EC-0.57 eV trap concentration was also observed in transistors to increase after simultaneous temperature and bias stressing.29,30 Others have suggested that FeGa-H defects exist in the as-grown material, and that H can be disassociated by heat and electric fields.31,32 This suggests local heating during the irradiation, and, likely to a lesser effect, knock-on displacement of hydrogens from the defects result in additional FeGa defects and higher measured EC-0.57 eV trap concentration. On the other hand, the increase in the concentrations of EC-0.72 eV and EC-0.9 eV traps is more significant. Their radiation sensitivity could indicate they are related to simple native point defects. Density functional theory (DFT) suggested that GaI, NI, GaN, and NGa point defects would form trap levels near EC-0.72 eV and EC-0.9 eV.26 Moreover, optical detection of magnetic resonance (ODMR) measurements detected an EC-0.9 eV level and was associated with GaI,33 which is consistent with the EC-0.9 eV trap concentration increasing with proton irradiation. Though MOCVD growth will not result in theoretical equilibrium trap concentrations, these interstitial and antisite defects have high predicted formation energies (up to 9 eV), so there remains whether these are the correct potential sources.26 Nonetheless, the strong radiation response indicates both the EC-0.72 eV and EC-0.90 eV traps are most likely native point defects.

TABLE II.

DLTS trap concentrations of LA before and after proton irradiation.

Trap energy EC-0.57 eV EC-0.72 eV EC-0.9 eV
Pre-irradiation conc. (cm−3 1.1 × 1014  1.8 × 1014  6.5 × 1014 
Post-irradiation conc. (cm−3 2.1 × 1014  5.0 × 1014  1.5 × 1015 
Trap energy EC-0.57 eV EC-0.72 eV EC-0.9 eV
Pre-irradiation conc. (cm−3 1.1 × 1014  1.8 × 1014  6.5 × 1014 
Post-irradiation conc. (cm−3 2.1 × 1014  5.0 × 1014  1.5 × 1015 

To explore beyond the ∼1 eV limitation of DLTS, DLOS measurements were conducted. From Fig. 3(a), the steady-state photo-capacitance (SSPC) shows three onsets in each sample at similar energies, indicating that the same traps exist in all three samples. These traps are all commonly observed in n-type GaN. The EC-1.35 eV trap has previously been attributed to CI.19 There are two candidates for the 3.28 eV onset, namely EC-3.22 eV and EC-3.28 eV levels. The EC-3.22 eV trap has been attributed to MgGa.34 The EC-3.28 eV trap features a negligible Franck–Condon energy (dFC), and its concentration is correlated with carbon concentration, which suggested it was due to CN.19 In our case, the EC-3.28 eV trap has a dFC of 0.03 eV, so it is likely related to carbon. The EC-2.6 eV trap also has two possible sources, which differ in dFC. The carbon-related EC-2.6 eV trap has a large dFC (∼0.4 eV), while the VGa-related EC-2.6 eV trap has a low dFC of ∼0.1 eV.19,25 Using Pässler's optical cross section model and fitting the extracted DLOS optical cross section in Fig. 3(b),35 the extracted dFC is 0.1 eV. Thus, the EC-2.6 eV level is likely primarily a VGa-related trap that is a compensating center.28 

FIG. 3.

(a) Steady-state photocapacitance of the LGR, LA, and HGR samples. (b) The traps in the DLOS spectra were fit using the Pässler model, and the Franck–Condon energies for the EC-2.6 eV and EC-3.28 eV traps were 0.1 eV and 0.03 eV, respectively.

FIG. 3.

(a) Steady-state photocapacitance of the LGR, LA, and HGR samples. (b) The traps in the DLOS spectra were fit using the Pässler model, and the Franck–Condon energies for the EC-2.6 eV and EC-3.28 eV traps were 0.1 eV and 0.03 eV, respectively.

Close modal

The trap concentrations are summarized in Table I. In the LGR sample, these trap concentrations are comparable to the best-reported trap concentrations.36–40 In the LA sample, highest concentrations are observed from the EC-0.9 eV, EC-1.35 eV, EC-2.6 eV, and EC-3.28 eV traps. Compared with the LGR sample, the EC-2.6 eV and EC-3.28 eV trap concentrations both increased ∼2× to ∼2 × 1015 cm−3, and the EC-0.9 eV trap increased ∼2× to 8.6 × 1014 cm−3. These trap concentrations in the LA sample are lower than the HGR sample suggesting that the laser-assisted growth may help reduce the trap concentrations, but the HGR sample also has a significantly higher growth rate. Thus, it is difficult to separate the effects of growth rate and laser for these traps, especially for the carbon-related traps, since the carbon concentration has been shown to depend on the growth rate, laser power, and laser beam shape.18 SIMS was conducted on similar samples under the same growth conditions as these three to determine the carbon incorporation dependence. SIMS indicates that the atomic carbon concentrations are 7 × 1015, 1.9 × 1016, and 2.6 × 1016 cm−3 for the LGR, LA, and HGR growth conditions, respectively, which is consistent with the DLOS trends. These numbers are higher than the carbon-related trap concentrations detected by DLOS, which is reasonable because carbon forms multiple levels and DLOS can underestimate the trap concentration when the trap energy is greater than half the bandgap.41,42 Overall, EC-1.35 eV, EC-2.6 eV, and EC-3.28 eV traps in the LA sample are comparable to the targeted doping (∼1016 cm−3), so work is needed to reduce these concentrations to allow more controllable doping through reduced carbon and gallium vacancy incorporation. Recently, it has been demonstrated that reduced gas phase reactions and lower carbon incorporation are possible through improved laser beam shaping, which enables both higher growth rates and suppressed carbon incorporation, so LA-MOCVD is continuing to improve.18 

In summary, we characterized traps in laser-assisted MOCVD-grown GaN and compared them with conventional MOCVD-grown GaN. No new traps were introduced compared with the baseline MOCVD samples. Trap concentrations were higher in the laser-assisted sample primarily due to a gallium vacancy-related defect (EC-2.6 eV), carbon defects (EC-1.35 eV and EC-3.28 eV), and a native point defect (EC-0.9 eV) with trap concentrations up to 2 × 1015 cm−3. However, the laser-assisted sample's total trap concentration is only ∼2× higher than the baseline sample, which was comparable with the best previously reported trap concentrations. Proton irradiation and previous ODMR and DFT studies indicate that the EC-0.9 eV trap is a native point defect likely related to GaI. Laser-assisted MOCVD growth is promising for high-growth-rate GaN applications such as high-voltage vertical GaN devices, but trap concentrations still need to be reduced for ideal device behavior.

The information, data, or work presented herein were funded in part by the Advanced Research Projects Agency-Energy (ARPA-E), U.S. Department of Energy (DOE), under Award No. DE-AR0001036, and the U.S. Department of Energy's Office of Energy Efficiency and Renewable Energy (EERE) under the Advanced Manufacturing Office, FY18/FY19 Lab Call, and the Office of Naval Research under Award No. N00014-20-1-2663. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. The authors also acknowledge the assistance of Ed Bielejec and Andrew Armstrong for the proton irradiation of the sample. Proton beam implantation was performed at Ion Beam Laboratory (IBL) at Sandia National Laboratories. Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and 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 No. DE-NA0003525. This paper describes objective technical results and analysis. Any subjective views or opinions that might be expressed in the paper do not necessarily represent the views of the DOE or the U.S. government.

The authors have no conflicts to disclose.

Wenbo Li: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Yuxuan Zhang: Resources (supporting); Writing – review & editing (supporting). Zhaoying Chen: Resources (supporting). Hongping Zhao: Conceptualization (supporting); Formal analysis (supporting); Funding acquisition (equal); Resources (equal); Writing – review & editing (equal). Steven A. Ringel: Funding acquisition (equal); Resources (equal); Supervision (equal); Writing – review & editing (equal). Aaron R. Arehart: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Resources (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
Y.
Zhang
,
Z.
Chen
,
W.
Li
,
A. R.
Arehart
,
S. A.
Ringel
, and
H.
Zhao
, “
Metalorganic chemical vapor deposition gallium nitride with fast growth rate for vertical power device applications
,”
Phys. Status Solidi A
218
(
6
),
2000469
(
2021
).
2.
Y.
Zhang
,
Z.
Chen
,
K.
Zhang
,
Z.
Feng
, and
H.
Zhao
, “
Laser-assisted metal–organic chemical vapor deposition of gallium nitride
,”
Phys. Status Solidi RRL
15
(
6
),
2100202
(
2021
).
3.
H.
Rabiee Golgir
,
Y.
Gao
,
Y. S.
Zhou
,
L.
Fan
,
P.
Thirugnanam
,
K.
Keramatnejad
,
L.
Jiang
,
J.-F.
Silvain
, and
Y. F.
Lu
, “
Low-temperature growth of crystalline gallium nitride films using vibrational excitation of ammonia molecules in laser-assisted metalorganic chemical vapor deposition
,”
Cryst. Growth Des.
14
(
12
),
6248
6253
(
2014
).
4.
H. R.
Golgir
,
Y. S.
Zhou
,
D.
Li
,
K.
Keramatnejad
,
W.
Xiong
,
M.
Wang
,
L. J.
Jiang
,
X.
Huang
,
L.
Jiang
,
J. F.
Silvain
, and
Y. F.
Lu
, “
Resonant and nonresonant vibrational excitation of ammonia molecules in the growth of gallium nitride using laser-assisted metal organic chemical vapour deposition
,”
J. Appl. Phys.
120
(
10
),
105303
(
2016
).
5.
H.
Rabiee Golgir
,
D. W.
Li
,
K.
Keramatnejad
,
Q. M.
Zou
,
J.
Xiao
,
F.
Wang
,
L.
Jiang
,
J.-F.
Silvain
, and
Y. F.
Lu
, “
Fast growth of GaN epilayers via laser-assisted metal–organic chemical vapor deposition for ultraviolet photodetector applications
,”
ACS Appl. Mater. Interfaces
9
(
25
),
21539
21547
(
2017
).
6.
H.
Ohta
,
N.
Asai
,
F.
Horikiri
,
Y.
Narita
,
T.
Yoshida
, and
T.
Mishima
, “
4.9 kV breakdown voltage vertical GaN p–n junction diodes with high avalanche capability
,”
Jpn. J. Appl. Phys., Part 1
58
,
SCCD03
(
2019
).
7.
H.
Ohta
,
K.
Hayashi
,
F.
Horikiri
,
M.
Yoshino
,
T.
Nakamura
, and
T.
Mishima
, “
5.0 kV breakdown-voltage vertical GaN p–n junction diodes
,”
Jpn. J. Appl. Phys., Part 1
57
,
04FG09
(
2018
).
8.
H.
Ohta
,
N.
Kaneda
,
F.
Horikiri
,
Y.
Narita
,
T.
Yoshida
,
T.
Mishima
, and
T.
Nakamura
, “
Vertical GaN p-n junction diodes with high breakdown voltages over 4 kV
,”
IEEE Electron Device Lett.
36
(
11
),
1180
1182
(
2015
).
9.
I. C.
Kizilyalli
,
T.
Prunty
, and
O.
Aktas
, “
4-kV and 2.8-mΩ-cm2 vertical GaN p-n diodes with low leakage currents
,”
IEEE Electron Device Lett.
36
(
10
),
1073
1075
(
2015
).
10.
G.
Piao
,
K.
Ikenaga
,
Y.
Yano
,
H.
Tokunaga
,
A.
Mishima
,
Y.
Ban
,
T.
Tabuchi
, and
K.
Matsumoto
, “
Study of carbon concentration in GaN grown by metalorganic chemical vapor deposition
,”
J. Cryst. Growth
456
,
137
139
(
2016
).
11.
S.
Hu
,
S.
Liu
,
Z.
Zhang
,
H.
Yan
,
Z.
Gan
, and
H.
Fang
, “
A novel MOCVD reactor for growth of high-quality GaN-related LED layers
,”
J. Cryst. Growth
415
,
72
77
(
2015
).
12.
Z.
Zhang
,
H.
Fang
,
Q.
Yao
,
H.
Yan
, and
Z.
Gan
, “
Species transport and chemical reaction in a MOCVD reactor and their influence on the GaN growth uniformity
,”
J. Cryst. Growth
454
,
87
95
(
2016
).
13.
W.
Sun
,
J.
Joh
,
S.
Krishnan
,
S.
Pendharkar
,
C. M.
Jackson
,
S. A.
Ringel
, and
A. R.
Arehart
, “
Investigation of trap-induced threshold voltage instability in GaN-on-Si MISHEMTs
,”
IEEE Trans. Electron Devices
66
(
2
),
890
895
(
2019
).
14.
A.
Sasikumar
,
A. R.
Arehart
,
D. W.
Cardwell
,
C. M.
Jackson
,
W.
Sun
,
Z.
Zhang
, and
S. A.
Ringel
, “
Deep trap-induced dynamic on-resistance degradation in GaN-on-Si power MISHEMTs
,”
Microelectron. Rel.
56
,
37
44
(
2016
).
15.
A.
Sasikumar
,
Z.
Zhang
,
P.
Kumar
,
E. X.
Zhang
,
D. M.
Fleetwood
,
R. D.
Schrimpf
,
P.
Saunier
,
C.
Lee
,
S. A.
Ringel
, and
A. R.
Arehart
, in
IEEE International Reliability Physics Symposium
(
IEEE
,
2015
), p.
2E.3.1
2E.3.6
.
16.
Z.
Zhang
,
D.
Cardwell
,
A.
Sasikumar
,
E. C. H.
Kyle
,
J.
Chen
,
E. X.
Zhang
,
D. M.
Fleetwood
,
R. D.
Schrimpf
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Correlation of proton irradiation induced threshold voltage shifts to deep level traps in AlGaN/GaN heterostructures
,”
J. Appl. Phys.
119
(
16
),
165704
(
2016
).
17.
J.
Mazumder
and
A.
Kar
,
Theory and Application of Laser Chemical Vapor Deposition
(
Springer Science & Business Media
,
2013
).
18.
Y.
Zhang
,
V. G.
Thirupakuzi Vangipuram
,
K.
Zhang
, and
H.
Zhao
, “
Investigation of carbon incorporation in laser-assisted MOCVD of GaN
,”
Appl. Phys. Lett.
122
(
16
),
162101
(
2023
).
19.
A.
Armstrong
,
A. R.
Arehart
,
D.
Green
,
U. K.
Mishra
,
J. S.
Speck
, and
S. A.
Ringel
, “
Impact of deep levels on the electrical conductivity and luminescence of gallium nitride codoped with carbon and silicon
,”
J. Appl. Phys.
98
(
5
),
053704
(
2005
).
20.
A.
Hierro
,
D.
Kwon
,
S. A.
Ringel
,
M.
Hansen
,
J. S.
Speck
,
U. K.
Mishra
, and
S. P.
DenBaars
, “
Optically and thermally detected deep levels in n-type Schottky and p+-n GaN diodes
,”
Appl. Phys. Lett.
76
(
21
),
3064
3066
(
2000
).
21.
P.
Blood
and
J. W.
Orton
,
The Electrical Characterization of Semiconductors: Majority Carriers and Electron States
(
Academic Press
,
London
,
1992
).
22.
A. R.
Arehart
,
A.
Corrion
,
C.
Poblenz
,
J. S.
Speck
,
U. K.
Mishra
, and
S. A.
Ringel
, “
Deep level optical and thermal spectroscopy of traps in n-GaN grown by ammonia molecular beam epitaxy
,”
Appl. Phys. Lett.
93
(
11
),
112101
(
2008
).
23.
K.
Galiano
,
J. I.
Deitz
,
S. D.
Carnevale
,
D. A.
Gleason
,
P. K.
Paul
,
Z.
Zhang
,
B. M.
McSkimming
,
J. S.
Speck
,
S. A.
Ringel
,
T. J.
Grassman
,
A. R.
Arehart
, and
J. P.
Pelz
, “
Spatial correlation of the EC-0.57 eV trap state with edge dislocations in epitaxial n-type gallium nitride
,”
J. Appl. Phys.
123
(
22
),
224504
(
2018
).
24.
Y.
Zhang
,
Z.
Chen
,
W.
Li
,
H.
Lee
,
M. R.
Karim
,
A. R.
Arehart
,
S. A.
Ringel
,
S.
Rajan
, and
H.
Zhao
, “
Probing unintentional Fe impurity incorporation in MOCVD homoepitaxy GaN: Toward GaN vertical power devices
,”
J. Appl. Phys.
127
(
21
),
215707
(
2020
).
25.
Z.
Zhang
,
E.
Farzana
,
W. Y.
Sun
,
J.
Chen
,
E. X.
Zhang
,
D. M.
Fleetwood
,
R. D.
Schrimpf
,
B.
McSkimming
,
E. C. H.
Kyle
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Thermal stability of deep level defects induced by high energy proton irradiation in n-type GaN
,”
J. Appl. Phys.
118
(
15
),
155701
(
2015
).
26.
J.
Neugebauer
and
C. G.
Van de Walle
, “
Atomic geometry and electronic structure of native defects in GaN
,”
Phys. Rev. B
50
(
11
),
8067
8070
(
1994
).
27.
F. D.
Auret
,
S. A.
Goodman
,
F. K.
Koschnick
,
J.-M.
Spaeth
,
B.
Beaumont
, and
P.
Gibart
, “
Electrical characterization of two deep electron traps introduced in epitaxially grown n-GaN during He-ion irradiation
,”
Appl. Phys. Lett.
73
(
25
),
3745
3747
(
1998
).
28.
Z.
Zhang
,
A. R.
Arehart
,
E.
Cinkilic
,
J.
Chen
,
E. X.
Zhang
,
D. M.
Fleetwood
,
R. D.
Schrimpf
,
B.
McSkimming
,
J. S.
Speck
, and
S. A.
Ringel
, “
Impact of proton irradiation on deep level states in n-GaN
,”
Appl. Phys. Lett.
103
(
4
),
042102
(
2013
).
29.
W.
Sun
,
C.
Lee
,
P.
Saunier
,
S. A.
Ringel
, and
A. R.
Arehart
, in
IEEE International Reliability Physics Symposium (IRPS)
(
IEEE
,
2016
), pp.
CD-3-1
CD-3-6
.
30.
A.
Sasikumar
,
A.
Arehart
,
S.
Kolluri
,
M. H.
Wong
,
S.
Keller
,
S. P.
DenBaars
,
J. S.
Speck
,
U. K.
Mishra
, and
S. A.
Ringel
, “
Access-region defect spectroscopy of DC-stressed N-polar GaN MIS-HEMTs
,”
IEEE Electron Device Lett.
33
(
5
),
658
660
(
2012
).
31.
A.
Sasikumar
,
D. W.
Cardwell
,
A. R.
Arehart
,
J.
Lu
,
S. W.
Kaun
,
S.
Keller
,
U. K.
Mishra
,
J. S.
Speck
,
J. P.
Pelz
, and
S. A.
Ringel
, in
IEEE International Reliability Physics Symposium
(
IEEE
,
2014
), pp.
2C.1.1
2C.1.6
.
32.
S.
Mukherjee
,
Y.
Puzyrev
,
J.
Chen
,
D. M.
Fleetwood
,
R. D.
Schrimpf
, and
S. T.
Pantelides
, “
Hot-carrier degradation in GaN HEMTs due to substitutional iron and its complexes
,”
IEEE Trans. Electron Devices
63
(
4
),
1486
1494
(
2016
).
33.
M.
Linde
,
S. J.
Uftring
,
G. D.
Watkins
,
V.
Härle
, and
F.
Scholz
, “
Optical detection of magnetic resonance in electron-irradiated GaN
,”
Phys. Rev. B
55
(
16
),
R10177
R10180
(
1997
).
34.
A.
Hierro
,
A. R.
Arehart
,
B.
Heying
,
M.
Hansen
,
U. K.
Mishra
,
S. P.
DenBaars
,
J. S.
Speck
, and
S. A.
Ringel
, “
Impact of Ga/N flux ratio on trap states in n-GaN grown by plasma-assisted molecular-beam epitaxy
,”
Appl. Phys. Lett.
80
(
5
),
805
807
(
2002
).
35.
R.
Pässler
, “
Photoionization cross-section analysis for a deep trap contributing to current collapse in GaN field-effect transistors
,”
J. Appl. Phys.
96
(
1
),
715
722
(
2004
).
36.
J.
Plesiewicz
et al, “
Deep-level traps in as-grown and electron-irradiated homo-epitaxial n-GaN layers grown by MOVPE
,”
Microelectron. Eng.
274
,
111977
(
2023
).
37.
T.
Tanaka
,
K.
Shiojima
,
T.
Mishima
, and
Y.
Tokuda
, “
Deep-level transient spectroscopy of low-free-carrier-concentration n-GaN layers grown on freestanding GaN substrates: Dependence on carbon compensation ratio
,”
Jpn. J. Appl. Phys., Part 1
55
(
6
),
061101
(
2016
).
38.
V. P.
Markevich
,
M. P.
Halsall
,
L.
Sun
,
I. F.
Crowe
,
A. R.
Peaker
,
P.
Kruszewski
,
J.
Plesiewicz
,
P.
Prystawko
,
S.
Bulka
, and
R.
Jakiela
, “
Electric-field enhancement of electron emission rates for deep-level traps in n-type GaN
,”
Phys. Status Solidi B
260
,
2200545
(
2023
).
39.
H.
Huang
,
X.
Yang
,
S.
Wu
,
J.
Shen
,
X.
He
,
L.
Wei
,
D.
Liu
,
F.
Xu
,
N.
Tang
,
X.
Wang
,
W.
Ge
, and
B.
Shen
, “
Investigation of carrier compensation traps in n-GaN drift layer by high-temperature deep-level transient spectroscopy
,”
Appl. Phys. Lett.
117
(
11
),
112103
(
2020
).
40.
M.
Horita
,
T.
Narita
,
T.
Kachi
, and
J.
Suda
, “
Identification of origin of EC −0.6 eV electron trap level by correlation with iron concentration in n-type GaN grown on GaN freestanding substrate by metalorganic vapor phase epitaxy
,”
Appl. Phys. Express
13
(
7
),
071007
(
2020
).
41.
E.
Farzana
,
A.
Mauze
,
J. B.
Varley
,
T. E.
Blue
,
J. S.
Speck
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Influence of neutron irradiation on deep levels in Ge-doped (010) β-Ga2O3 layers grown by plasma-assisted molecular beam epitaxy
,”
APL Mater.
7
(
12
),
121102
(
2019
).
42.
E.
Farzana
,
M. F.
Chaiken
,
T. E.
Blue
,
A. R.
Arehart
, and
S. A.
Ringel
, “
Impact of deep level defects induced by high energy neutron radiation in β-Ga2O3
,”
APL Mater.
7
(
2
),
022502
(
2019
).