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

Explanation of low efficiency droop in semipolar ( 2 0 2 ¯ 1 ¯ ) InGaN/GaN LEDs through evaluation of carrier recombination coefficients

Open Access Open Access

Abstract

We report the carrier dynamics and recombination coefficients in single-quantum-well semipolar (202¯1¯) InGaN/GaN light-emitting diodes emitting at 440 nm with 93% peak internal quantum efficiency. The differential carrier lifetime is analyzed for various injection current densities from 5 A/cm2 to 10 kA/cm2, and the corresponding carrier densities are obtained. The coupling of internal quantum efficiency and differential carrier lifetime vs injected carrier density (n) enables the separation of the radiative and nonradiative recombination lifetimes and the extraction of the Shockley-Read-Hall (SRH) nonradiative (A), radiative (B), and Auger (C) recombination coefficients and their n-dependency considering the saturation of the SRH recombination rate and phase-space filling. The results indicate a three to four-fold higher A and a nearly two-fold higher B0 for this semipolar orientation compared to that of c-plane reported using a similar approach [A. David and M. J. Grundmann, Appl. Phys. Lett. 96, 103504 (2010)]. In addition, the carrier density in semipolar (202¯1¯) is found to be lower than the carrier density in c-plane for a given current density, which is important for suppressing efficiency droop. The semipolar LED also shows a two-fold lower C0 compared to c-plane, which is consistent with the lower relative efficiency droop for the semipolar LED (57% vs. 69%). The lower carrier density, higher B0 coefficient, and lower C0 (Auger) coefficient are directly responsible for the high efficiency and low efficiency droop reported in semipolar (202¯1¯) LEDs.

© 2017 Optical Society of America

1. Introduction

III-nitride light-emitting diodes (LEDs) are key elements in solid-state lighting and visible-light communication systems. The semipolar (202¯1¯) orientation of GaN is of particular interest due to its substantially reduced polarization-related electric fields, narrow emission linewidth, robust temperature dependence, high indium incorporation efficiency, large optical polarization ratio, and low efficiency droop [1–14]. Recently, LEDs with very high output power (~1 Watt) and low efficiency droop were demonstrated using 12 to 14-nm-thick single-quantum-well (SQW) active regions on this orientation [2,15]. Theoretical investigations suggest that semipolar (202¯1¯) LEDs have a ~3X larger electron-hole wavefunction overlap [3], which should predict larger ABC recombination coefficients [16,17] and peak external quantum efficiencies that occur at higher current densities [18] compared to c-plane LEDs. However, despite the promising properties and excellent device results for semipolar (202¯1¯) LEDs, the reason behind the low efficiency droop is not well understood. The carrier dynamics have only been studied by time-resolved photoluminescence [19] and the ABC recombination coefficients have not been experimentally investigated at all. An investigation of the carrier dynamics and ABC recombination coefficients under electrical injection is useful to help explain the droop behavior, confirm previous theoretical investigations, and guide future LED design on semipolar (202¯1¯).

In this work, we determine the differential carrier lifetime (DLT) and ABC recombination coefficients for semipolar (202¯1¯) LEDs under electrical injection using methods similar to previously developed small-signal techniques [20–22]. Here we measure the small-signal frequency response to obtain the modulation bandwidth and extract the DLT. The DLT results are used in conjunction with internal-quantum efficiency (IQE) measurements to separate the radiative and nonradiative components of the lifetime. The recombination rates vs carrier density (n) are then computed in the context of an ABC model that includes phase-space filling effects. Fitting of the recombination rates vs carrier density then enables extraction of the ABC coefficients. The results indicate a three to four-fold higher A, a two-fold higher B0, and a two-fold lower C0 for this semipolar orientation compared to that of c-plane. To the best of our knowledge, this is the first systematic evaluation of the DLT and ABC parameters on a nonpolar or semipolar orientation using electrical injection.

2. Experimental details

The investigated sample is a semipolar (202¯1¯) LED grown using metal-organic chemical vapor deposition (MOCVD) on a free-standing (202¯1¯) GaN substrate from Mitsubishi Chemical Corporation. The active region of the LED consists of a single 12-nm-thick InGaN/GaN double heterostructure (DH) emitting around 440 nm. The LEDs consist of circular mesas with 60 µm diameter, indium-tin-oxide (ITO) p-contacts, and ground-signal-ground RF electrodes. Micro-LEDs are used to minimize the effects of the RC time constant and carrier transport in the evaluation of the DLTs [23–26]. Details of the epitaxial structure and device fabrication are described in [19] and [26].

The optical frequency response was measured using a microwave probe station and network analyzer. A small RF signal (100 mV) was combined with a DC bias using a bias tee. The light was collected from the top of the device into an optical fiber and focused on a silicon photodetector with a 2 GHz bandwidth. The received electrical signal was amplified using a 30 dB low-noise electrical amplifier. The amplified signal was then received by the network analyzer to determine the optical frequency response and 3-dB modulation bandwidth (f3dB). The DLT was extracted from the 3-dB modulation bandwidth using τΔn=1/(2πf3dB), where τΔn is the DLT. RC transport effects are neglected in this study based on their small contribution in micro-LEDs in previous studies [26]. However, in the future, RC transport effects can be included using more advanced rate equations and circuit modeling techniques [25]. Here we measure the optical frequency response under continuous-wave electrical injection since the temperature dependence of the optical frequency response and DLT are negligible based on the thermal study performed by David et al [27].

The shape of the IQE vs current density curve was determined by measuring the relative external-quantum efficiency (EQE) under pulsed (500 ns pulse-width, 10 kHz repetition rate, and 0.5% duty cycle) electrical injection to avoid self-heating effects. To adjust the relative IQE curve to an absolute IQE curve, we set the peak value to 93%, which was previously obtained for this sample using room-temperature/low-temperature photoluminescence (PL) [19]. We also assume the light-extraction efficiency is independent of current density.

3. Results and discussions

Figure 1 shows the DLT and IQE as a function of current density ranging from 5 A/cm2 to 10 kA/cm2. The DLT decreases with increasing current density and shows signs of saturation at high current densities (>8 kA/cm2) (Fig. 1(a)). The DLT ranges from 8.8 ns at 5 A/cm2 to 0.42 ns at 10 kA/cm2, which are lower than the reported values for the case of c-plane (~12 ns at 5 A/cm2 to 0.5 ns at 10 kA/cm2) [21]. The IQE (ηIQE) increases with current density to a maximum of 93% at ~165 A/cm2, after which the device experiences efficiency droop, reaching 40% at 10 kA/cm2 (Fig. 1(b)).

 figure: Fig. 1

Fig. 1 (a) Differential carrier lifetime and (b) internal-quantum efficiency as a function of current density for the semipolar (202¯1¯) LED.

Download Full Size | PDF

Using the carrier lifetime as a function of current density, we then calculate the corresponding injected carrier density. The carrier density (n) is related to the DLT (τΔn) and the generation rate (G) by the following [21,28]:

n=0GτΔndG.
For electrical injection, the generation rate G is
G=Je×d
where J, e, and d are the current density, electron charge, and active region thickness, respectively. The lower limit of the integral corresponds to a generation rate of zero, while the measureable minimum for the carrier lifetime corresponds to a non-zero generation rate. The carrier lifetime at zero generation rate is the Shockley-Read-Hall lifetime (τSRH), which is typically obtained by extrapolating the lifetime at low J to the y-axis. Here, τSRH is ~10 ns. Using (1) and (2), the carrier density corresponding to a given current density can be calculated with the knowledge of. τΔn Fig. 2 shows this dependency for the semipolar (202¯1¯) LED compared to the c-plane LED from [21]. To reasonably compare the two orientations, we scale the carrier densities for the c-plane case up by 1.25 to account for the different active region thicknesses (12 nm semipolar vs 15 nm c-plane). Figure 2 shows that the carrier density for a given current density is always lower in the semipolar case, which is due to the higher electron-hole wavefunction overlap in the semipolar case [3]. The difference in carrier density is particularly noticeable at low current densities before screening of the polarization field occurs. At higher current densities, the carrier densities become more similar.

 figure: Fig. 2

Fig. 2 Calculated carrier density (n) as a function of injected current density (J) for semipolar (black circles) and c-plane (blue circles) LEDs. The c-plane data is obtained from [21] and is adjusted to account for the difference in active layer thicknesses (12 nm semipolar vs 15 nm c-plane).

Download Full Size | PDF

Having n as a function of J, we then plot ηIQE and τΔn as a function of n as shown in Fig. 3(a). The peak IQE occurs at a carrier density of 3 × 1018 cm−3. Knowing τΔn, n, and ηIQE, we find the radiative (τΔnR) and nonradiative (τΔnNR) recombination lifetimes separately using [21]

τΔnR1=dGRdn=ηIQEτΔn1+G×dηIQEdn
and
τΔnNR1=dGNRdn=(1ηIQE)τΔn1G×dηIQEdn.
The resulting radiative (τΔnR), nonradiative (τΔnNR), and total differential carrier lifetimes (τΔn) as a function of carrier density are plotted in Fig. 3(b). At low n, τΔnNR approaches the SRH lifetime of τSRH ~ 10 ns that was used to determine the lower limit of the integral in (1). Interestingly, the radiative (τnR), nonradiative (τnNR), and total PL lifetimes (τn) obtained from previous optical measurements (see [19] for details of the measurement and calculations) follow very similar trends (Fig. 3(c)) to those obtained using the DLT approach (Fig. 3(b)). The DLTs are generally slightly lower than the PL lifetimes, as expected from the definition of DLT [29]. In addition, the optical excitation range was much lower for the PL case due to limitations in the setup, which results in an incomplete picture for the PL case (Fig. 3(c)).

 figure: Fig. 3

Fig. 3 (a) IQE and DLT vs. carrier density, (b) radiative, nonradiative, and total DLT vs. carrier density, and (c) radiative, nonradiative, and total PL lifetime vs carrier density.

Download Full Size | PDF

Now following the work of David et al. [21] and considering an ABC model with a phase space-filling effects [20,30], the radiative (GR) and nonradiative recombination rates (GNR) are defined as:

GR=GηIQE=B0(1+(nn*))n2
GNR=G(1ηIQE)=A(n)n+C0(1+(nn*))n3
where A(n), B0, and C0 denote the Shockley-Read-Hall nonradiative, radiative, and Auger recombination coefficients, respectively. Also, n* is a fitting parameter accounting for the Fermi-Dirac carrier statistics in phase-space filling. To obtain the coefficients, we compute and plot three different parameters: GNR/n (Fig. 4(a)), GR/n2 (Fig. 4(b)), and GNR/n3 (Fig. 4(c)). The first parameter, at low carrier density, represents A(n). The second parameter gives B(n), from which B0 and n* are obtained using fitting. The third parameter, at high carrier density, gives C(n), from which C0 is obtained knowing n* and A(n). All of the recombination coefficients can be derived from the fittings shown in Fig. 4. It is noteworthy that, unlike in the c-plane case [21], the A coefficient is a function of n, reaching a minimum at n ~ 2×1018cm3 (Fig. 4(a)). The value of n at the minimum corresponds to the peak of τΔnNR in Fig. 3(b). This behavior is explained by the saturation of SRH recombination centers at high carrier density [19]. To obtain clean fittings to the data at low carrier densities, we use the original SRH model [31,32] described in [33] to describe the saturation. The SRH nonradiative lifetime is τ(SRH)= τn0+τp0(nn+NA), where τn0 and τp0 are the electron and hole minority carrier lifetimes (which are defined as the inverse of the electron and hole capture coefficients [31]) and NA is the density of recombination centers. The choice of model for the saturation of SRH recombination is for fitting purposes only and does not affect the obtained values of the ABC coefficients since we extract (and compare with c-plane [21]) the A coefficient at the lowest carrier density (~3.5 × 1017 cm−3).

 figure: Fig. 4

Fig. 4 (a),  GNR/n, (b), GR/n2 and (c)  GNR/n3 as a function of carrier density calculated from the n-dependent radiative and nonradiative lifetimes. The dashed lines indicate the fitting by the ABC model.

Download Full Size | PDF

A summary of the recombination coefficients, peak IQE, relative droop, and current density at the peak IQE is shown in Table 1. For the sake of comparison, Table 1 also shows reference data reported using a similar technique for a c-plane LED with a single-quantum-well active region [21]. The semipolar LED shows some notable differences compared to the c-plane LED. First, the A and B0 coefficients for the semipolar LED are higher by a factor of approximately four and two, respectively. The higher A and B coefficients are consistent with theoretical work by Kioupakis et al [16,17] predicting that the recombination coefficients are roughly proportional to the square of the electron-hole wavefunction overlap, which is about 2.5 to 3 times higher in semipolar (202¯1¯) than c-plane for current densities below 500 A/cm2 [3]. The larger A coefficient in the semipolar LED is not surprising based on the nature of the quantum well profile at low bias. The semipolar quantum well is nearly flat banded at low bias due to the cancellation of the built-in p-n junction field and the polarization-related field [3]. The flat band profile leads to a large wavefunction overlap at low bias, which allows the first injected carriers to easily recombine through SRH centers at low carrier densities when the A(n)n term is dominant. In c-plane, the first injected carriers at low bias have a poor wavefunction overlap, which should suppress SRH recombination and reduce the A coefficient.

Tables Icon

Table 1. Recombination parameters for semipolar and polar InGaN/GaN LEDs

Correcting for the difference in quantum well thicknesses between the two orientations, Fig. 2 shows that the semipolar orientation has a lower carrier density for a given current density, especially at low carrier densities where the internal electric field is not yet screened. This is a direct result of the larger A and B0 coefficients in the semipolar LED due to the reduced internal electric fields. A related effect was reported by David et al. [34] where the A and B coefficients were found to be significantly reduced in longer wavelength LEDs with larger internal electric fields. The lower carrier density in the semipolar LED is also responsible for increasing the current density corresponding to the peak IQE (Jpeak). We observe a more than three-fold larger Jpeak in the semipolar LED despite the thinner active region (see Table 1), which effectively delays the onset of droop effects to higher current densities.

In contrast to A and B0, the C0 coefficient is lower in the semipolar LED compared to the c-plane LED (Table 1). While the lower C0 coefficient for the semipolar LED is not directly expected based on wavefunction overlap arguments alone, it is consistent with the lower relative efficiency droop in Table 1 and the higher value of GR/n2GNR/n3 shown in Fig. 5, which does not result from an ABC model. Therefore, in addition to the wavefunction overlap, three additional factors that affect the Auger rate are briefly considered. First, at the high current densities relevant to Auger recombination, significant coulomb screening is expected in the active region and the difference in wavefunction overlap between semipolar and c-plane becomes smaller [35]. Thus, the difference in wavefunction overlap should have a smaller effect on the Auger rate than on the radiative rate. The convergence of the carrier density vs. current density curves at high current density in Fig. 2 supports this conclusion. Second, the Auger rate depends upon the detailed conduction and valence band structures, which constrain the allowed Auger transitions based on the requirements of energy and momentum conservation. Based on anisotropic strain, the band structures for semipolar and nonpolar differ significantly from those of c-plane [36–39], which may affect the Auger rate. On the other hand, indirect recombination based on alloy and phonon scattering was identified as the primary contributor to the Auger rate in InGaN [40] and was shown to be relatively insensitive to anisotropic strain [41]. Thus, the lower effective mass in the semipolar orientation should result in a higher B coefficient but not significantly affect the C coefficient if indirect Auger is the dominant process. Finally, localization as a result of indium alloy fluctuations has been shown to significantly affect the recombination coefficients, with C being more strongly affected than B [42,43]. Some studies have shown low indium fluctuations in semipolar (202¯1¯) [6,44,45], which may result in the lower C coefficient and higher B coefficient observed in the semipolar case. Further experiments and first principles calculations specific to semipolar orientations are required to fully explain the origin of the lower Auger coefficient in semipolar (202¯1¯).

 figure: Fig. 5

Fig. 5 The ratio GR/n2GNR/n3 as a function of carrier density for semipolar (black circles) and c-plane (blue circles) from [21]. The dashed lines indicate the extracted B0C0 values from the fittings.

Download Full Size | PDF

The radiative and nonradiative rates can also be calculated without assuming an ABC model by using GR=GIQE and GNR=G(1IQE). Figure 5 shows the ratio GR/n2GNR/n3 as a function of carrier density. This ratio is essentially B/C at high carrier densities and is approximately three times larger for the semipolar LED. The higher GR/n2GNR/n3 in the semipolar case is important for realizing high-efficiency LEDs with low efficiency droop. Figure 5 also shows the B0/C0 ratio calculated using the coefficients obtained from our fittings. As expected, the ratio GR/n2GNR/n3 converges to B0/C0 at high carrier densities, further validating the fittings. The relatively constant value of GR/n2GNR/n3 at high carrier densities for the c-plane case is consistent with a phase-space filling model where n* is the same for both B(n) and C(n). In the semipolar case, the slight negative slope in GR/n2GNR/n3 at high carrier densities suggests the phase-space filling effects for C(n) are slightly weaker than those for B(n).

4. Summary and conclusions

In summary, we report, for the first time, the recombination dynamics and ABC coefficients in electrically injected semipolar (202¯1¯) InGaN/GaN single-quantum-well LEDs using a coupled differential carrier lifetime and internal-quantum-efficiency analysis. The radiative and nonradiative contributions to the differential lifetime as a function of carrier density were separated. In contrast to the nonradiative lifetime, the radiative lifetime initially decreases with n until 1.3 × 1019 cm−3 and slowly increases afterwards. The behavior of radiative and nonradiative differential lifetime obtained by electrical injection agrees well with the corresponding lifetimes obtained from PL analysis. In the context of an ABC model with phase-space filling, the A and B0 coefficients are found to be higher by a factor of about four and two than those reported for a similar c-plane LED using the same technique. This is consistent with the lower carrier density for the same injected current density at low current density in the semipolar structure. The C0 coefficient is lower by a factor of two in the semipolar LED, which is opposite to the first-order theoretical prediction based on wavefunction overlap arguments alone [16,17]. However, secondary effects such as coulomb screening at high current densities, band structure modifications, and the degree of carrier localization are also expected to affect the Auger coefficient and require further study in semipolar (202¯1¯) quantum wells to fully explain the lower C0 coefficient. The lower carrier density, higher B0 coefficient, and lower C0 (Auger) coefficient are directly responsible for the high efficiency and low efficiency droop reported in semipolar (202¯1¯) LEDs.

Funding

Department of Defense (DOD) (W911NF-15-1-0428); National Science Foundation DMR (NSF-DMR) (1121053).

Acknowledgments

This work was supported by Department of Defense award number W911NF-15-1-0428. The authors acknowledge the support from the Solid State Lighting and Energy Electronics Center (SSLEEC) at UCSB. The research carried out at UCSB made use of shared experimental facilities of the Materials Research Laboratory (MRL): an NSF MRSEC, supported by NSF DMR 1121053. The MRL is a member of the NSF-supported Materials Research Facilities Network (www.mrfn.org).

References and links

1. Y. Zhao, S. Tanaka, C.-C. Pan, K. Fujito, D. Feezell, J. S. Speck, S. P. DenBaars, and S. Nakamura, “High-Power Blue-Violet Semipolar(202¯1¯), ” Appl. Phys. Express 4(8), 082104 (2011). [CrossRef]  

2. C.-C. Pan, S. Tanaka, F. Wu, Y. Zhao, J. S. Speck, S. Nakamura, S. P. DenBaars, and D. Feezell, “High-Power, Low-Efficiency-Droop Semipolar(202¯1¯), ” Appl. Phys. Express 5(6), 062103 (2012). [CrossRef]  

3. D. F. Feezell, J. S. Speck, S. P. DenBaars, and S. Nakamura, “Semipolar InGaN/GaN Light-Emitting Diodes for High-Efficiency Solid-State Lighting,” J. Disp. Technol. 9, 190–198 (2013). [CrossRef]  

4. Y. Kawaguchi, C.-Y. Huang, Y.-R. Wu, Q. Yan, C.-C. Pan, Y. Zhao, S. Tanaka, K. Fujito, D. Feezell, C. G. V. de Walle, S. P. DenBaars, and S. Nakamaura, “Influence of polarity on carrier transport in semipolar(202¯1¯) and(202¯1¯) multiple-quantum-well light-emitting diodes, ” Appl. Phys. Lett. 100(23), 231110 (2012). [CrossRef]  

5. S. Marcinkevičius, R. Ivanov, Y. Zhao, S. Nakamaura, S. P. DenBaars, and J. S. Speck, “Highly polarized photoluminescence and its dynamics in semipolar(202¯1¯) InGaN/GaN quantum well, ” Appl. Phys. Lett. 104(11), 111113 (2014). [CrossRef]  

6. S. Marcinkevičius, Y. Zhao, K. M. Kelchner, S. Nakamaura, S. P. DenBaars, and J. S. Speck, “Near-field investigation of spatial variations of(202¯1¯) InGaN quantum well emission spectra, ” Appl. Phys. Lett. 103(13), 131116 (2013). [CrossRef]  

7. S.-H. Park and D. Ahn, “High optical polarization ratio of semipolar(202¯1¯) -oriented InGaN/GaN quantum wells and comparison with experiment, ” J. Appl. Phys. 112(9), 093106 (2012). [CrossRef]  

8. Y. Zhao, S. Tanaka, Q. Yan, C.-Y. Huang, R. B. Chung, C.-C. Pan, K. Fujito, D. Feezell, C. G. Van de Walle, J. S. Speck, S. P. DenBaars, and S. Nakamura, “High optical polarization ratio from semipolar(202¯1¯) bluegreen InGaN/GaN light-emitting diodes, ” Appl. Phys. Lett. 99(5), 051109 (2011). [CrossRef]  

9. C.-Y. Huang, M. T. Hardy, K. Fujito, D. Feezell, J. S. Speck, S. P. DenBaars, and S. Nakamaura, “Demonstration of 505 nm laser diodes using wavelength-stable semipolar(202¯1¯) InGaN/GaN quantum wells, ” Appl. Phys. Lett. 99(24), 241115 (2011). [CrossRef]  

10. Y. Zhao, Q. Yan, C.-Y. Huang, S.-C. Huang, P. S. Hsu, S. Tanaka, C.-C. Pan, Y. Kawaguchi, K. Fujito, C. G. Van de Walle, J. S. Speck, S. P. DenBaars, S. Nakamaura, and D. Feezell, “Indium incorporation and emission properties of nonpolar and semipolar InGaN quantum wells,” Appl. Phys. Lett. 100(20), 201108 (2012). [CrossRef]  

11. C.-C. Pan, T. Gilbert, N. Pfaff, S. Tanaka, Y. Zhao, D. Feezell, J. S. Speck, S. Nakamura, and S. P. DenBaars, “Reduction in Thermal Droop Using Thick Single-Quantum-Well Structure in Semipolar(202¯1¯) Blue Light- Emitting Diodes, ” Appl. Phys. Express 5(10), 102103 (2012). [CrossRef]  

12. Y. Zhao, S. H. Oh, F. Wu, Y. Kawaguchi, S. Tanaka, K. Fujito, J. S. Speck, S. P. DenBaars, and S. Nakamura, “Green Semipolar(202¯1¯) InGaN Light-Emitting Diodes with Small Wavelength Shift and Narrow Spectral Linewidth, ” Appl. Phys. Express 6(6), 062102 (2013). [CrossRef]  

13. S. Marcinkevičius, R. Ivanov, Y. Zhao, S. Nakamura, and S. P. DenBaars, “Highly polarized photoluminescence and its dynamics in semipolar(202¯1¯) InGaN/GaN quantum well, ” Appl. Phys. Lett. 104, 111113 (2014). [CrossRef]  

14. F. Wu, Y. Zhao, A. E. Romanov, S. P. DenBaars, S. Nakamura, and J. S. Speck, “Stacking faults and interface roughening in semipolar(202¯1¯) single InGaN quantum wells for long wavelength emission, ” Appl. Phys. Lett. 104(15), 151901 (2014). [CrossRef]  

15. S. H. Oh, B. P. Yonkee, M. Cantore, R. M. Farrell, J. S. Speck, S. Nakamura, and S. P. DenBaars, “Semipolar III–nitride light-emitting diodes with negligible efficiency droop up to ∼1 W,” Appl. Phys. Express 9(10), 102102 (2016). [CrossRef]  

16. E. Kioupakis, Q. Yan, D. Steiauf, and C. G. V. de Walle, “Temperature and carrier-density dependence of Auger and radiative recombination in nitride optoelectronic devices,” New J. Phys. 15(12), 125006 (2013). [CrossRef]  

17. E. Kioupakis, Q. Yan, and C. G. V. de Walle, “Interplay of polarization fields and Auger recombination in the efficiency droop of nitride light-emitting diodes,” Appl. Phys. Lett. 101(23), 231107 (2012). [CrossRef]  

18. T. Melo, Y.-L. Hu, C. Weisbuch, M. C. Schmidt, A. David, B. Ellis, C. Poblenz, Y.-D. Lin, M. R. Krames, and J. W. Raring, “Gain comparison in polar and nonpolar/semipolar gallium-nitride-based laser diodes,” Semicond. Sci. Technol. 27(2), 024015 (2012). [CrossRef]  

19. S. Okur, M. Nami, A. K. Rishinaramangalam, S. H. Oh, S. P. DenBaars, S. Liu, I. Brener, and D. F. Feezell, “Internal quantum efficiency and carrier dynamics in semipolar(202¯1¯) InGaN/GaN light-emitting diodes, ” Opt. Express 25(3), 2178–2186 (2017). [CrossRef]  

20. P. G. Eliseev, M. Osinski, H. Li, and I. V. Akimova, “Recombination balance in green-light-emitting GaN/InGaN/AlGaN quantum wells,” Appl. Phys. Lett. 75(24), 3838–3840 (1999). [CrossRef]  

21. A. David and M. J. Grundmann, “Droop in InGaN light-emitting diodes: A differential carrier lifetime analysis,” Appl. Phys. Lett. 96(10), 103504 (2010). [CrossRef]  

22. D.-P. Han, J.-I. Shim, and D.-S. Shin, “Analysis of carrier recombination dynamics in InGaN-based light-emitting diodes by differential carrier lifetime measurement,” Appl. Phys. Express 10(5), 052101 (2017). [CrossRef]  

23. S. Rajbhandari, J. J. D. McKendry, J. Herrnsdorf, H. Chun, G. Faulkner, H. Haas, I. M. Watson, D. O’Brien, and M. D. Dawson, “A review of gallium nitride LEDs for multi-gigabit-per-second visible light data communications,” Semicond. Sci. Technol. 32(2), 023001 (2017). [CrossRef]  

24. R. X. G. Ferreira, E. Xie, J. J. D. McKendry, S. Rajbhandari, H. Chun, G. Faulkner, S. Watson, A. E. Kelly, E. Gu, R. V. Penty, I. H. White, D. C. O’Brien, and M. D. Dawson, “High Bandwidth GaN-Based Micro-LEDs for Multi-Gb/s Visible Light Communications,” IEEE Photonics Technol. Lett. 28(19), 2023–2026 (2016). [CrossRef]  

25. A. Rashidi, M. Nami, M. Monavarian, A. Aragon, K. DaVico, F. Ayoub, S. Mishkat-Ul-Masabih, A. K. Rishinaramangalam, and D. Feezell, “Differential carrier lifetime and transport effects in electrically injected III-nitride light-emitting diodes,” J. Appl. Phys. 122(3), 035706 (2017). [CrossRef]  

26. A. Rashidi, M. Monavarian, A. Aragon, S. Okur, M. Nami, A. Rishinaramangalam, S. Mishkat-Ul-Masabih, and D. Feezell, “High-Speed Nonpolar InGaN/GaN LEDs for Visible-Light Communication,” IEEE Photonics Technol. Lett. 29(4), 381–384 (2017). [CrossRef]  

27. A. David, C. A. Hurni, N. G. Young, and M. D. Craven, “Carrier dynamics and Coulomb-enhanced capture in III-nitride quantum heterostructures,” Appl. Phys. Lett. 109(3), 033504 (2016). [CrossRef]  

28. D. Yevick and W. Streifer, “Radiative and nonradiative recombination law in lightly doped InGaAsP lasers,” Electron. Lett. 19(24), 1012–1014 (1983). [CrossRef]  

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

30. D. Huang, J.-I. Chyi, and H. Morkoç, “Carrier effects on the excitonic absorption in GaAs quantum-well structures: Phase-space filling,” Phys. Rev. B Condens. Matter 42(8), 5147–5153 (1990). [CrossRef]   [PubMed]  

31. W. Shockley and W. T. Read, “Statistics of the Recombinations of Holes and Electrons,” Phys. Rev. 87(5), 835–842 (1952). [CrossRef]  

32. R. N. Hall, “Electron-Hole Recombination in Germanium,” Phys. Rev. 87(2), 387 (1952). [CrossRef]  

33. A. Cuevas and D. Macdonald, “Measuring and interpreting the lifetime of silicon wafers,” Sol. Energy 76(1-3), 255–262 (2004). [CrossRef]  

34. A. David and M. J. Grundmann, “Influence of polarization fields on carrier lifetime and recombination rates in InGaN-based light-emitting diodes,” Appl. Phys. Lett. 97(3), 033501 (2010). [CrossRef]  

35. E. Kuokstis, C. Q. Chen, M. E. Gaevski, W. H. Sun, J. W. Yang, G. Simin, M. Asif Khan, H. P. Maruska, D. W. Hill, M. C. Chou, J. J. Gallagher, and B. Chai, “Polarization effects in photoluminescence of c- and m-plane GaN/AlGaN multiple quantum wells,” Appl. Phys. Lett. 81(22), 4130–4132 (2002). [CrossRef]  

36. Y.-R. Wu, C.-Y. Huang, Y. Zhao, and J. S. Speck, “8 - Nonpolar and semipolar LEDs,” in Nitride Semiconductor Light-Emitting Diodes (LEDs) (Woodhead Publishing, 2014), pp. 250–275.

37. H.-H. Huang and Y.-R. Wu, “Light emission polarization properties of semipolar InGaN/GaN quantum well,” J. Appl. Phys. 107(5), 053112 (2010). [CrossRef]  

38. I. L. Koslow, M. T. Hardy, P. Shan Hsu, P.-Y. Dang, F. Wu, A. Romanov, Y.-R. Wu, E. C. Young, S. Nakamura, J. S. Speck, and S. P. DenBaars, “Performance and polarization effects in (112¯2) long wavelength light emitting diodes grown on stress relaxed InGaN buffer layers,” Appl. Phys. Lett. 101(12), 121106 (2012). [CrossRef]  

39. S. Khatsevich and D. H. Rich, “The effects of crystallographic orientation and strain on the properties of excitonic emission from wurtzite InGaN/GaN quantum wells,” J. Phys. Condens. Matter 20(21), 215223 (2008). [CrossRef]  

40. E. Kioupakis, P. Rinke, K. T. Delaney, and C. G. Van de Walle, “Indirect Auger recombination as a cause of efficiency droop in nitride light-emitting diodes,” Appl. Phys. Lett. 98(16), 161107 (2011). [CrossRef]  

41. E. Kioupakis, D. Steiauf, P. Rinke, K. T. Delaney, and C. G. Van de Walle, “First-principles calculations of indirect Auger recombination in nitride semiconductors,” Phys. Rev. B 92(3), 035207 (2015). [CrossRef]  

42. C. Jones, C.-H. Teng, Q. Yan, P.-C. Ku, and E. Kioupakis, “Impact of Anderson localization on carrier recombination in InGaN quantum wells and the efficiency of nitride light-emitting diodes,” ArXiv170206009 Cond-Mat (2017).

43. T.-J. Yang, R. Shivaraman, J. S. Speck, and Y.-R. Wu, “The influence of random indium alloy fluctuations in indium gallium nitride quantum wells on the device behavior,” J. Appl. Phys. 116(11), 113104 (2014). [CrossRef]  

44. R. Shivaraman, Y. Kawaguchi, S. Tanaka, S. P. DenBaars, S. Nakamura, and J. S. Speck, “Comparative analysis of(202¯1¯)and(202¯1¯) semipolar GaN light emitting diodes using atom probe tomography, ” Appl. Phys. Lett. 102(25), 251104 (2013). [CrossRef]  

45. Y. Zhao, F. Wu, T.-J. Yang, Y.-R. Wu, S. Nakamura, and J. S. Speck, “Atomic-scale nanofacet structure in semipolar and InGaN single quantum wells,” Appl. Phys. Express 7, 025503 (2014). [CrossRef]  

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 (a) Differential carrier lifetime and (b) internal-quantum efficiency as a function of current density for the semipolar ( 2 0 2 ¯ 1 ¯ ) LED.
Fig. 2
Fig. 2 Calculated carrier density ( n ) as a function of injected current density ( J ) for semipolar (black circles) and c-plane (blue circles) LEDs. The c-plane data is obtained from [21] and is adjusted to account for the difference in active layer thicknesses (12 nm semipolar vs 15 nm c-plane).
Fig. 3
Fig. 3 (a) IQE and DLT vs. carrier density, (b) radiative, nonradiative, and total DLT vs. carrier density, and (c) radiative, nonradiative, and total PL lifetime vs carrier density.
Fig. 4
Fig. 4 (a),   G N R / n , (b),   G R / n 2 and (c)   G N R / n 3 as a function of carrier density calculated from the n-dependent radiative and nonradiative lifetimes. The dashed lines indicate the fitting by the ABC model.
Fig. 5
Fig. 5 The ratio G R / n 2 G N R / n 3 as a function of carrier density for semipolar (black circles) and c-plane (blue circles) from [21]. The dashed lines indicate the extracted B 0 C 0 values from the fittings.

Tables (1)

Tables Icon

Table 1 Recombination parameters for semipolar and polar InGaN/GaN LEDs

Equations (6)

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

n = 0 G τ Δ n d G .
G = J e × d
τ Δ n R 1 = d G R d n = η I Q E τ Δ n 1 + G × d η I Q E d n
τ Δ n N R 1 = d G N R d n = ( 1 η I Q E ) τ Δ n 1 G × d η I Q E d n .
G R = G η I Q E = B 0 ( 1 + ( n n * ) ) n 2
G N R = G ( 1 η I Q E ) = A ( n ) n + C 0 ( 1 + ( n n * ) ) n 3
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.