Next Article in Journal
Hyperbolic Behavior and Antiferromagnetic Order in Rare-Earth Tellurides
Previous Article in Journal
First-Principles Calculations of the Phonon, Mechanical and Thermoelectric Properties of Half-Heusler Alloy VIrSi Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Observation of Carrier Transportation between Localized States in InGaN Quantum Wells

1
College of Semiconductors (College of Integrated Circuits), Hunan University, Changsha 410082, China
2
School of Physics and Electronics, Hunan University, Changsha 410082, China
3
Hunan San’an Semiconductor Co., Ltd., Changsha 410000, China
4
Key Laboratory for Renewable Energy, Beijing Key Laboratory for New Energy Materials and Devices, Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
5
Songshan Lake Materials Laboratory, Dongguan 523808, China
*
Authors to whom correspondence should be addressed.
Submission received: 28 November 2022 / Revised: 12 December 2022 / Accepted: 14 December 2022 / Published: 16 December 2022
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
Despite the large misfit dislocation densities, indium gallium nitride (InGaN) demonstrates high luminous efficiency both for electroluminescence and photoluminescence. The mechanism behind it has been interpreted as the existence of potential minima (i.e., localized states), which will screen the non-radiative recombination centers to avoid carriers being trapped by the defects. The existence of localized states has been testified by many experiments. However, almost all of the observations are indirect observations, and some experiments, such as those focused on whether the indium clusters observed by transmission electron microscopy are localized states, still remain controversial. Here, we report the direct observation of carrier transportation between localized states driven by temperature-dependent photoluminescence (TDPL) and excitation power-dependent PL in InGaN quantum wells. This enriches the experimental evidence on the existence of localized states.

1. Introduction

The transcendental optoelectronic properties of indium gallium nitride (InGaN) enable the prevailing application and development of InGaN-based devices in solid-state lighting [1,2,3,4,5,6], micro light-emitting diode (μLED) displays [7,8], light communication [9,10], etc. Since the early research stages for InGaN, the high luminous efficiency accompanied by high misfit defects (e.g., the threading dislocation density in the magnitude order of ~108–109 cm−2) has attracted much attention from researchers, seeking to interpret such a contradictory phenomenon. Chichibu et al. suggest the existence of potential minima in the InGaN materials, which will localize the carriers spatially, thus avoiding carrier consumption via non-radiative recombination centers caused by defects [11]. The main experimental evidence includes a large Stokes shift between the photoluminescence (PL) peak and the absorption peak given by the photoluminescence excitation (PLE), and also the energy deviation between the electroluminescence (EL) peak and the modulated electroabsorption (EA) spectra [12]. In addition, the blue shift of the EL peak under a high injection current also indicates the band filling of the localized states. A long decay time (~ns) obtained from the time-resolved PL (TRPL) measurements also implies the carrier’s transportation towards localized states. Moreover, the decay time elongates at the lower-energy side of the emission peak, in agreement with the localized state-induced band-tail distribution of carriers [13]. According to the cathodoluminescence (CL) results, a small-area scan demonstrates sharp peaks with narrower full-width at half-maximum (FWHM) values, which also reveals spatial potential fluctuations [12]. However, due to the limited spatial resolution of CL measurement, the size of localized states could only be estimated to be smaller than 60 nm [14]. Eliseev et al. proposed a model to quantitatively evaluate the degree of localized states and suggested that an S-shaped curve, which is demonstrated by the evolution of the PL peak with increasing temperature, should be considered evidence of the carrier’s transportation between deep and shallow localized states [15]. As the possible mechanisms causing localized states include thickness undulation and composition fluctuation (most probably caused by phase separation) [16], transmission electron microscopy (TEM) is widely used to provide direct evidence of the localized states [11]. A number of researchers have observed indium clusters with a lateral size in several to tens of nanometers in the InGaN quantum well (QW), and such a morphology, with a higher indium composition than that of the adjoining sections, is usually interpreted as a localized state [2,11,17]. However, other researchers suggest that the indium clusters should not be localized states. Lefebvre et al. found that the S-shaped curve even exists in the InGaN quantum box; then, they concluded that the spatial extension of localized states should be smaller than the size of the quantum box (5–10 nm) [18]. Graham et al. suggested that the in-plane length of localized states should be ~2 nm according to the analyses of the longitudinal-optical (LO) phonon replicas [19]. Moreover, they attributed the formation of localized states to the well-width fluctuations. Smeeton et al. pointed out that the indium clusters were most probably false images, as they will emerge with a prolonged electron irradiation time or when exposed to a high-voltage electron beam [20]. The possible mechanisms may be attributable to radiolysis and knock-on damage, which cause the lattice distortion of InGaN, resulting in an electron-beam-induced strain [21]. Other researchers used the three-dimensional atom probe (3DAP) technique to analyze the atom distribution of InGaN multiple-quantum wells (MQWs) and found that the indium distribution coincides with the binomial distribution that occurs in random ternary alloys [22,23]. They stated that an inhomogeneous indium distribution was not observed and suggested that the localized states should be a result of well-width fluctuations [22]. Galtrey et al. further explained that localized states are a result of QW width fluctuations, which are caused by a surface-segregated indium layer [24,25]. As a result, the observation of localized states from TEM becomes a less convincing method. Thereafter, the localized states were further attributed to the atomic condensates of In–N via analyses of implanted positron annihilation experimental results [26].
In 2019, we reported the experimental results of carrier transportation between different localized states as the temperature was ramped up from 10 K to 300 K in a single yellow-green QW [27]. Such a study offers more concrete evidence for the existence of localized states. However, as the energy difference between the two localized states was ~70 meV, it was difficult to distinguish the integrated intensity for each peak exactly. In this study, we fabricated three InGaN MQWs, which were subjected to an in-situ post-annealing process after the growth of each QW to establish quantum-dot-like states. From the temperature-dependent PL (TDPL) and excitation power-dependent PL measurements, we directly observed the carrier’s transport from the deep localized states to the relatively shallow localized states as the temperature or the excitation power increased. The results provide a vivid picture of the existence of localized states.

2. Materials and Methods

Three MQWs were grown on a 350-nm-thick GaN template following an aluminum nitride (AlN) buffer layer composed of a 15-nm low-temperature AlN and a 300-nm high-temperature AlN on a 2-inch silicon (111) substrate by metal organic chemical vapor deposition (MOCVD). The detailed growth conditions for the GaN template on the silicon substrate can be found in our previous study [28]. To assess the presence of quantum dots (QDs) in the QW, we exploited the in-situ post-annealing method, which is also called the growth interruption method, to evaporate the indium clusters surrounding defects and facilitate the formation of quantum-dot-like localized states after the ending of each QW [29,30,31]. The three QWs underwent the same growth conditions. Lastly, a 50-nm GaN was deposited as the capping layer. The schematic structure is shown in Figure 1. The TDPL were taken by a 325-nm He-Cd continuous-wavelength laser with the sample mounted on the cold finger of a closed-cycle helium compressor. The temperature could be changed from ~15 K to 300 K consecutively. The excitation power-dependent PL was carried out with the same setup, with the excitation power adjusted by a neutral optical attenuator at different temperatures (viz. 15 K, 70 K, 150 K, and 300 K) [32]. The TEM images were obtained via a JEM 2100Plus microscope, where the lamella was mechanically thinned, followed by further thinning by argon ions with a Gatan precision ion polishing system (PIPS). The voltage of TEM was kept at 200 kV.

3. Results and Discussion

An example of the PL spectra obtained from 15 K to 300 K is presented in Figure 2a. At lower temperatures (15 K–~60 K), four emission peaks are obviously distinguishable, which are attributable to the emissions from deep localized states (PL), shallow localized states (PH), the wetting layer (PW), and GaN (PGaN), respectively. With increasing temperature, the PGaN quenches rapidly and becomes indistinguishable at 300 K. The PW also decays rapidly and disappears at the relatively lower excitation power at 300 K. A noticeable emission intensity change from PL to PH can be easily observed, as shown in Figure 2a. For a more detailed study, the spectra at each temperature are fitted by a multiple-peak curve fit with a Gaussian function. The exemplary fitted curves, as well as the experimentally obtained spectra at 15 K and 300 K, are depicted in Figure 2b, where the black lines are the measured data, the green lines are the fitted lines for each emission peak, and the red lines are the sums of all the green lines. The extracted relative PL-integrated intensities of the three emission peaks are delineated in Figure 2c. The relative PL-integrated intensity for a certain peak at each temperature is obtained by calculating the proportion of the absolute PL-integrated intensity of this peak in the sum of the absolute PL-integrated intensity for the three peaks. According to Figure 2c, the relative PL-integrated intensity of PH remains on an increasing trajectory along with the increase in temperature, while those of PL and PW are liable to a monotonous decrease as the temperature ramps up. According to the changing tendency of the relative PL-integrated intensity, we infer that as the temperature increases, the carriers will gain enough kinetic energy to overcome the restriction of the deep localized states, thus causing a weakened PL emission of PL. The consecutive increase in PH emission among the three peaks implies that the thermally activated carriers tend to accumulate in the shallow localized state (i.e., PH). It is worth noting that the absolute PL-integrated intensity of all the peaks decreases with the temperature increase. This could be due to the thermally activated non-radiative recombination centers, which consume the carriers, thus weakening the emission efficiency [4,30]. However, as the growth conditions are the same for the three QWs, the deep and shallow localized states should be randomly distributed in the MQWs; then, the effects of the non-radiative recombination centers on both peaks should be equal at each temperature. In this case, the increase in the relative PL-integrated intensity in PH is mainly due to the carrier’s transportation from deep localized state (PL). As the relative emission intensity of the wetting layer also demonstrates a decrease with temperature, the carrier transportation from PW to PH also cannot be excluded.
For a further study of the carrier transportation behavior, and to evaluate the non-radiative recombination centers in each peak at different temperatures, the excitation power-dependent PL was carried out at 15 K, 70 K, 150 K, and 300 K, respectively, with the excitation power being increased from 0.2 mW to 8 mW. Figure 3a,d,g,j present the spectral evolution under different excitation powers at 15 K, 70 K, 150 K, and 300 K, respectively. All the four peaks could clearly be distinguished due to the high radiative recombination efficiency at lower temperature, viz. 15 K. The peaks of PL, PH, and PW are fitted by a Gaussian function, similarly to the treatment of the TDPL spectra; thus, the integrated intensities are extracted. As the PW is submerged into the background noise in 300 K, only PL and PH could be well fitted through the excitation power range; thus, the data of PW are absent at 300 K. The PL-integrated intensity obeys a power-law evolution with the excitation power, viz. I P m , where I represents the PL-integrated intensity, P denotes the excitation power, and m is the index. If radiative recombination dominates the recombination process, the index m is expected to be close to 1. If non-radiative recombination plays an important role in the recombination process, then the index m will be larger than 1, as the non-radiative recombination centers will not be effectively saturated by the excited carriers under relatively low excitation powers [33,34]. The experimental data in Figure 3b,e,h,k are offset for clarity; therefore, they are not equal to the absolute PL-integrated intensities for PL, PH, and PW. The fitted index m values for the three peaks (only two peaks at 300 K) are all close to 1 from 15 K to 300 K, and the values of m for each peak remain a small deviation at different temperatures. Such a phenomenon indicates that the recombination mechanism is analogous at different temperatures for each peak, confirming the assumption that the effects of the non-radiative recombination centers on all peaks are equivalent at each temperature. In addition, as the values of m are all close to 1, radiative recombination should dominate the recombination process in all three peaks. Therefore, the evolution of the PL-integrated intensity with temperature is mainly caused by the carrier transportation rather than the different non-radiative recombination processes in each peak. A more obvious carrier transportation between different localized states could be found in the evolution of the relative PL-integrated intensities with excitation power. The relative PL-integrated intensities of PH and PW maintain an increasing trend, whereas that of PL diminishes monotonously, as shown in Figure 3c,f,i,j. Moreover, from Figure 3c,f, we find that, at lower excitation powers (0.2–~1 mW), the relative PL-integrated intensity of PL is the highest. This indicates that, at lower excitation powers, due to the deficiency of activated carriers, the photo-activated carriers are likely to fill the deep localized states as a first priority, thus generating the strongest emission from the deep localized states. When the excitation power increases to a certain level, the spillover carriers will fill the shallow localized states and the wetting layer, resulting in an enhancement in the emission intensities of PH and PW. As the peaks are somewhat weak under 0.2 mW excitation at 150 K, the PL-integrated intensity at 0.2 mW for PL and PW may suffer from an inaccurate evaluation because the peaks are difficult to be extracted from the background noise. Therefore, the data at 0.2 mW in Figure 3i show a deviation from the tendency in Figure 3c,f. However, as the PL signals are strong enough at higher excitation powers, the other data in Figure 3i are in agreement with those at other temperatures. As the Pw peak is severely weakened at 300 K, especially under lower excitation powers, only PL and PH are investigated at this temperature. From Figure 3k, it is verified that both PL and PH are dominated by the radiative recombination process. The monotonous decrease in the relative PL-integrated intensity of PL and the concomitant increase in PH in Figure 3l are indicators of the carrier transportation between different localized states. As the temperature of 300 K is relatively high, the carrier transportation behavior becomes more prominent.
The TEM images of the as-grown sample are presented in Figure 4. Figure 4a illustrates the stacking structure of the sample. Three MQWs with non-consecutive regions are sandwiched between the GaN template and the capping layer. In Figure 4b, one can clearly observe the morphology of the unambiguous quantum-dot-like blotches in the QWs. As the acceleration voltage was not high (200 kV), an aperture slot was added before the objective lens, and the electron spot was dispersed to a rather weak extent, we tend to accept such features to be the real characteristics of the material. In addition, we captured the images less than 10 s after exposure, and we also could not observe a morphology change as the exposure time elapsed. A further magnified image of the MQW region is shown in Figure 4c, where the quantum-dot-like speckles were measured to be 2.5–4.2 nm in width and 2.8–5 nm in height. Such quantum-dot-like structures may contribute to the formation of localized states.

4. Conclusions

In conclusion, the carrier transportation between deep and shallow localized states is directly observed by TDPL and excitation power-dependent PL. The transportation behavior is further verified by the excitation power-dependent PL at different temperatures, where all the emission peaks are confirmed to be dominated by the radiative recombination process. The TEM images demonstrate a large number of quantum-dot-like blotches, which may contribute to the formation of localized states. This study enriches the evidence from experiments supporting the localized state assumption in InGaN QWs.

Author Contributions

Y.L. (Yangfeng Li) proposed and conducted the overall project. Y.L. (Yixiao Li), Y.W. and T.L. contributed to the methodology. Y.L. (Yixiao Li), Y.W. and T.L. contributed to the validation. J.Z., Y.J., H.J., W.W., R.Y. and H.C. supervised. J.Z., Y.J., H.J., W.W., R.Y. and H.C. obtained the funding support. Y.L. (Yangfeng Li) provided the crystal growth; performed the TDPL and excitation-power dependent PL measurements, TEM sample preparation, TEM measurements, and data analysis; and composed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation of China (NSFC) (grant numbers 62122084, 12074412, 62150027, and 61991441), the Strategic Priority Research Program of Chinese Academy of Sciences (grant number XDB33000000), the Fundamental Research Funds for the Central Universities (grant number 531118010804), and the Education Department of Hunan Province (grant number 22B0020).

Data Availability Statement

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

Acknowledgments

We thank Ziguang Ma from the Institute of Physics, Chinese Academy of Sciences, for the discussion of the experimental results. We are indebted to Professor Aizi Jin from the Institute of Physics, Chinese Academy of Sciences, for providing the facility for TEM sample fabrication, and Professor Ying Li, Huanfang Tian, and Xin’an Yang from the Institute of Physics, Chinese Academy of Sciences, for the TEM test.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schubert, E.F.; Kim, J.K. Solid-State Light Sources Getting Smart. Science 2005, 308, 5726. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, L.; Wang, L.; Yu, J.; Hao, Z.; Luo, Y.; Sun, C.; Han, Y.; Xiong, B.; Wang, J.; Li, H. Abnormal Stranski-Krastanov Mode Growth of Green InGaN Quantum Dots: Morphology, Optical Properties, and Applications in Light-Emitting Devices. ACS Appl. Mater. Interfaces 2019, 11, 1228–1238. [Google Scholar] [CrossRef] [PubMed]
  3. Iida, D.; Zhuang, Z.; Kirilenko, P.; Velazquez-Rizo, M.; Najmi, M.A.; Ohkawa, K. 633-nm InGaN-based red LEDs grown on thick underlying GaN layers with reduced in-plane residual stress. Appl. Phys. Lett. 2020, 116, 162101. [Google Scholar] [CrossRef]
  4. Li, Y.; Liu, C.; Zhang, Y.; Jiang, Y.; Hu, X.; Song, Y.; Su, Z.; Jia, H.; Wang, W.; Chen, H. Realizing Single Chip White Light InGaN LED via Dual-Wavelength Multiple Quantum Wells. Materials 2022, 15, 3998. [Google Scholar] [CrossRef]
  5. Li, Y.; Jiang, Y.; Jia, H.; Wang, W.; Yang, R.; Chen, H. Superior Optoelectronic Performance of N-Polar GaN LED to Ga-Polar Counterpart in the "Green Gap" Range. IEEE Access 2022, 10, 95565. [Google Scholar] [CrossRef]
  6. Jiang, Y.; Li, Y.; Li, Y.; Deng, Z.; Lu, T.; Ma, Z.; Zuo, P.; Dai, L.; Wang, L.; Jia, H.; et al. Realization of high-luminous-efficiency InGaN light-emitting diodes in the green gap range. Sci. Rep. 2015, 5, 10883. [Google Scholar] [CrossRef] [Green Version]
  7. Meng, W.; Xu, F.; Yu, Z.; Tao, T.; Shao, L.; Liu, L.; Li, T.; Wen, K.; Wang, J.; He, L.; et al. Three-dimensional monolithic micro-LED display driven by atomically thin transistor matrix. Nat. Nanotechnol. 2021, 16, 1231–1236. [Google Scholar] [CrossRef]
  8. Hwangbo, S.; Hu, L.; Hoang, A.T.; Choi, J.Y.; Ahn, J.H. Wafer-scale monolithic integration of full-colour micro-LED display using MoS(2) transistor. Nat. Nanotechnol. 2022, 17, 500–506. [Google Scholar] [CrossRef]
  9. Wang, L.; Wei, Z.; Chen, C.-J.; Wang, L.; Fu, H.Y.; Zhang, L.; Chen, K.-C.; Wu, M.-C.; Dong, Y.; Hao, Z.; et al. 1.3  GHz E-O bandwidth GaN-based micro-LED for multi-gigabit visible light communication. Photonics Res. 2021, 9, 792. [Google Scholar] [CrossRef]
  10. Li, G.; Hu, F.; Zou, P.; Wang, C.; Lin, G.R.; Chi, N. Advanced Modulation Format of Probabilistic Shaping Bit Loading for 450-nm GaN Laser Diode based Visible Light Communication. Sensors 2020, 20, 6143. [Google Scholar] [CrossRef]
  11. Chichibu, S.; Azuhata, T.; Sota, T.; Nakamura, S. Spontaneous emission of localized excitons in InGaN single and multiquantum well structures. Appl. Phys. Lett. 1996, 69, 4188. [Google Scholar] [CrossRef]
  12. Chichibu, S.; Sota, T.; Wada, K.; Nakamura, S. Exciton localization in InGaN quantum well devices. J. Vac. Sci. Technol. B 1998, 16, 2204. [Google Scholar] [CrossRef]
  13. Narukawa, Y.; Kawakami, Y.; Fujita, S.; Fujita, S.; Nakamura, S. Recombination dynamics of localized excitons in In0.20Ga0.80N-In0.05Ga0.95 multiple quantum wells. Phys. Rev. B 1997, 55, R1938–R1941. [Google Scholar] [CrossRef]
  14. Chichibu, S.; Wada, K.; Nakamura, S. Spatially resolved cathodoluminescence spectra of InGaN quantum wells. Appl. Phys. Lett. 1997, 71, 2346. [Google Scholar] [CrossRef]
  15. Eliseev, P.G.; Perlin, P.; Lee, J.; Osiński, M. “Blue” temperature-induced shift and band-tail emission in InGaN-based light sources. Appl. Phys. Lett. 1997, 71, 569–571. [Google Scholar] [CrossRef]
  16. Ho, I.-h.; Stringfellow, G.B. Solid phase immiscibility in GaInN. Appl. Phys. Lett. 1996, 69, 2701. [Google Scholar] [CrossRef]
  17. Narukawa, Y.; Kawakami, Y.; Funato, M.; Fujita, S.; Fujita, S.; Nakamura, S. Role of self-formed InGaN quantum dots for exciton localization in the purple laser diode emitting at 420 nm. Appl. Phys. Lett. 1997, 70, 981–983. [Google Scholar] [CrossRef]
  18. Lefebvre, P.; Taliercio, T.; Morel, A.; Allègre, J.; Gallart, M.; Gil, B.; Mathieu, H.; Damilano, B.; Grandjean, N.; Massies, J. Effects of GaAlN barriers and of dimensionality on optical recombination processes in InGaN quantum wells and quantum boxes. Appl. Phys. Lett. 2001, 78, 1538–1540. [Google Scholar] [CrossRef]
  19. Graham, D.M.; Soltani-Vala, A.; Dawson, P.; Godfrey, M.J.; Smeeton, T.M.; Barnard, J.S.; Kappers, M.J.; Humphreys, C.J.; Thrush, E.J. Optical and microstructural studies of InGaN/GaN single-quantum-well structures. J. Appl. Phys. 2005, 97, 103508. [Google Scholar] [CrossRef]
  20. Smeeton, T.M.; Kappers, M.J.; Barnard, J.S.; Vickers, M.E.; Humphreys, C.J. Electron-beam-induced strain within InGaN quantum wells: False indium "cluster" detection in the transmission electron microscope. Appl. Phys. Lett. 2003, 83, 5419. [Google Scholar] [CrossRef]
  21. Smeeton, T.M.; Humphreys, C.J.; Barnard, J.S.; Kappers, M.J. The impact of electron beam damage on the detection of indium-rich localisation centres in InGaN quantum wells using transmission electron microscopy. J. Mater. Sci. 2006, 41, 2729–2737. [Google Scholar] [CrossRef]
  22. Galtrey, M.J.; Oliver, R.A.; Kappers, M.J.; Humphreys, C.J.; Stokes, D.J.; Clifton, P.H.; Cerezo, A. Three-dimensional atom probe studies of an InxGa1-xN / GaN multiple quantum well structure: Assessment of possible indium clustering. Appl. Phys. Lett. 2007, 90, 061903. [Google Scholar] [CrossRef]
  23. Bennett, S.E.; Saxey, D.W.; Kappers, M.J.; Barnard, J.S.; Humphreys, C.J.; Smith, G.D.; Oliver, R.A. Atom probe tomography assessment of the impact of electron beam exposure on InxGa1-xN / GaN quantum wells. Appl. Phys. Lett. 2011, 99, 021906. [Google Scholar] [CrossRef]
  24. Galtrey, M.J.; Oliver, R.A.; Kappers, M.J.; Humphreys, C.J.; Clifton, P.H.; Larson, D.; Saxey, D.W.; Cerezo, A. Three-dimensional atom probe analysis of green- and blue-emitting InxGa1-xN / GaN multiple quantum well structures. J. Appl. Phys. 2008, 104, 013524. [Google Scholar] [CrossRef]
  25. Northrup, J.E.; Walle, C.G.V.d. Indium versus hydrogen-terminated GaN(0001) surfaces: Surfactant effect of indium in a chemical vapor deposition environment. Appl. Phys. Lett. 2004, 84, 4322. [Google Scholar] [CrossRef]
  26. Chichibu, S.F.; Uedono, A.; Onuma, T.; Haskell, B.A.; Chakraborty, A.; Koyama, T.; Fini, P.T.; Keller, S.; Denbaars, S.P.; Speck, J.S.; et al. Origin of defect-insensitive emission probability in In-containing (Al,In,Ga)N alloy semiconductors. Nat. Mater. 2006, 5, 810–816. [Google Scholar] [CrossRef]
  27. Li, Y.; Deng, Z.; Ma, Z.; Wang, L.; Jia, H.; Wang, W.; Jiang, Y.; Chen, H. Visualizing carrier transitions between localization states in a InGaN yellow–green light-emitting-diode structure. J. Appl. Phys. 2019, 126, 095705. [Google Scholar] [CrossRef]
  28. Li, Y.; Tang, C.W.; Lau, K.M. Comparison of the AlN and GaN crystalline quality on 2-inch silicon substrate via two growth methods. J. Cryst. Growth 2020, 535, 125545. [Google Scholar] [CrossRef]
  29. Li, Y.; Jin, Z.; Han, Y.; Zhao, C.; Huang, J.; Tang, C.W.; Wang, J.; Lau, K.M. Surface morphology and optical properties of InGaN quantum dots with varying growth interruption time. Mater. Res. Express 2020, 7, 015903. [Google Scholar] [CrossRef]
  30. Zhu, Y.; Lu, T.; Zhou, X.; Zhao, G.; Dong, H.; Jia, Z.; Liu, X.; Xu, B. Origin of huge photoluminescence efficiency improvement in InGaN GaN multiple quantum wells with low-temperature GaN cap layer grown in N-2/H-2 mixture gas. Appl. phys. Express 2017, 10, 061004. [Google Scholar] [CrossRef]
  31. Woolf, A.; Puchtler, T.; Aharonovich, I.; Zhu, T.; Niu, N.; Wang, D.; Oliver, R.; Hu, E.L. Distinctive signature of indium gallium nitride quantum dot lasing in microdisk cavities. Proc. Natl. Acad. Sci. USA 2014, 111, 14042–14046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, Y.; Hu, X.; Song, Y.; Su, Z.; Jia, H.; Wang, W.; Jiang, Y.; Chen, H. The influence of temperature of nitridation and AlN buffer layer on N-polar GaN. Mater. Sci. Semicond. Process. 2022, 141, 106423. [Google Scholar] [CrossRef]
  33. Lee, Y.-J.; Chiu, C.-H.; Ke, C.C.; Lin, P.C.; Lu, T.-C.; Kuo, H.-C.; Wang, S.-C. Study of the Excitation Power Dependent Internal Quantum Efficiency in InGaN GaN LEDs Grown on Patterned Sapphire Substrate. IEEE J. Sel. Top. Quantum Electron. 2009, 15, 1137–1143. [Google Scholar] [CrossRef]
  34. Wang, H.; Ji, Z.; Qu, S.; Wang, G.; Jiang, Y.; Liu, B.; Xu, X.; Mino, H. Influence of excitation power and temperature on photoluminescence in InGaN GaN multiple quantum wells. Opt. Express 2012, 20, 3932–3940. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The schematic illustration of the InGaN MQW structure.
Figure 1. The schematic illustration of the InGaN MQW structure.
Crystals 12 01837 g001
Figure 2. (a) The TDPL spectra from 15 K to 300 K. PL, PH, PW, and PGaN denote the emission peaks of the deep localized states (lower energy), shallow localized states (higher energy), wetting layer, and GaN, respectively. (b) The spectra as well as the fitting curves at 15 K and 300 K. The green lines are the fitted lines for each peak, while the red lines are the gross fitted lines for the whole spectrum. (c) The relative PL intensity for PL, PH, and PW at different temperatures.
Figure 2. (a) The TDPL spectra from 15 K to 300 K. PL, PH, PW, and PGaN denote the emission peaks of the deep localized states (lower energy), shallow localized states (higher energy), wetting layer, and GaN, respectively. (b) The spectra as well as the fitting curves at 15 K and 300 K. The green lines are the fitted lines for each peak, while the red lines are the gross fitted lines for the whole spectrum. (c) The relative PL intensity for PL, PH, and PW at different temperatures.
Crystals 12 01837 g002
Figure 3. (a,d,g,j) The excitation power-dependent PL spectra at 15 K, 70 K, 150 K, and 300 K, respectively. (b,e,h,k) The PL-integrated intensity versus the excitation power at 15 K, 70 K, 150 K, and 300 K, respectively. The red lines are the fitted lines obeying a power law. The PL-integrated intensities are offset for clarity. (c,f,i,l) The relative PL-integrated intensity of the emission peaks under different excitation powers at 15 K, 70 K, 150 K, and 300 K, respectively.
Figure 3. (a,d,g,j) The excitation power-dependent PL spectra at 15 K, 70 K, 150 K, and 300 K, respectively. (b,e,h,k) The PL-integrated intensity versus the excitation power at 15 K, 70 K, 150 K, and 300 K, respectively. The red lines are the fitted lines obeying a power law. The PL-integrated intensities are offset for clarity. (c,f,i,l) The relative PL-integrated intensity of the emission peaks under different excitation powers at 15 K, 70 K, 150 K, and 300 K, respectively.
Crystals 12 01837 g003aCrystals 12 01837 g003b
Figure 4. (a) The TEM image of the as-grown sample. Three MQWs with non-consecutive regions are noticeable. (b) The zoomed-in picture in the MQW region. Quantum-dot-like blotches are distinguishable. (c) A further magnified image of the MQW region. The quantum-dot-like speckles were measured to be 2.5–4.2 nm in width and 2.8–5 nm in height.
Figure 4. (a) The TEM image of the as-grown sample. Three MQWs with non-consecutive regions are noticeable. (b) The zoomed-in picture in the MQW region. Quantum-dot-like blotches are distinguishable. (c) A further magnified image of the MQW region. The quantum-dot-like speckles were measured to be 2.5–4.2 nm in width and 2.8–5 nm in height.
Crystals 12 01837 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Li, Y.; Zhang, J.; Wang, Y.; Li, T.; Jiang, Y.; Jia, H.; Wang, W.; Yang, R.; Chen, H. Direct Observation of Carrier Transportation between Localized States in InGaN Quantum Wells. Crystals 2022, 12, 1837. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12121837

AMA Style

Li Y, Li Y, Zhang J, Wang Y, Li T, Jiang Y, Jia H, Wang W, Yang R, Chen H. Direct Observation of Carrier Transportation between Localized States in InGaN Quantum Wells. Crystals. 2022; 12(12):1837. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12121837

Chicago/Turabian Style

Li, Yangfeng, Yixiao Li, Jie Zhang, Yi Wang, Tong Li, Yang Jiang, Haiqiang Jia, Wenxin Wang, Rong Yang, and Hong Chen. 2022. "Direct Observation of Carrier Transportation between Localized States in InGaN Quantum Wells" Crystals 12, no. 12: 1837. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12121837

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

Article Metrics

Back to TopTop