Next Article in Journal
Assessing of the Road Pavement Roughness by Means of LiDAR Technology
Next Article in Special Issue
Enhanced Methanol Oxidation Activity of PtRu/C100−xMWCNTsx (x = 0–100 wt.%) by Controlling the Composition of C-MWCNTs Support
Previous Article in Journal
Numerical Modeling of a Short-Dwell Coater for Bio-Based Coating Applications
Previous Article in Special Issue
Improved Mechanical Properties and Corrosion Resistance of Mg-Based Bulk Metallic Glass Composite by Coating with Zr-Based Metallic Glass Thin Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving Transport Properties of GaN-Based HEMT on Si (111) by Controlling SiH4 Flow Rate of the SiNx Nano-Mask

1
Department of Electrophysics, National Chiao Tung University, Hsinchu 30010, Taiwan
2
Technology Development Division, Episil-Precision Inc., Hsinchu 30010, Taiwan
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(1), 16; https://doi.org/10.3390/coatings11010016
Submission received: 30 November 2020 / Revised: 18 December 2020 / Accepted: 22 December 2020 / Published: 25 December 2020
(This article belongs to the Special Issue Recent Advances in the Growth and Characterizations of Thin Films)

Abstract

:
The AlGaN/AlN/GaN high electron mobility transistor structures were grown on a Si (111) substrate by metalorganic chemical vapor deposition in combination with the insertion of a SiNx nano-mask into the low-temperature GaN buffer layer. Herein, the impact of SiH4 flow rate on two-dimensional electron gas (2DEG) properties was comprehensively investigated, where an increase in SiH4 flow rate resulted in a decrease in edge-type threading dislocation density during coalescence process and an improvement of 2DEG electronic properties. The study also reveals that controlling the SiH4 flow rate of the SiNx nano-mask grown at low temperatures in a short time is an effective strategy to overcome the surface desorption issue that causes surface roughness degradation. The highest electron mobility of 1970 cm2/V·s and sheet carrier concentration of 6.42 × 1012 cm−2 can be achieved via an optimized SiH4 flow rate of 50 sccm.

1. Introduction

GaN-based high electron mobility transistors (HEMTs) have attracted much intense research because of its high-voltage operation, high-frequency switching, and high thermal conductivity for high-power and high-frequency device applications [1,2,3]. However, the growth of GaN-based HEMTs faces the problem of non-commercial nitride substrates, where sapphire (Al2O3) and silicon carbide (SiC) are two commonly used substrates. In recent years, to realize a compromise between high performance and wafer-size scalability, low cost, and integrated compatibility with the conventional Si-based technology, many research groups have grown GaN-based HEMT on a Si (111) substrate. Even so, the large lattice mismatch of 17% along with a thermal expansion coefficient mismatch of 54% between the Si (111) substrate and GaN layer is the most challenging of any epitaxial growth methods to achieve high-quality GaN layers. High-density dislocation (~1010 cm−2) and surface cracks due to thermal expansion during the cooling process are generally observed in the GaN/Si (111) layers, contributing to the degradation of the GaN-based HEMT performance. In general, the mainstream of research has focused on reducing the current collapse, which is attributed to both the interface trapping of AlGaN barrier/passivation layers and the bulk trapping of defects/dislocations [4,5,6,7,8,9,10]. To mitigate the bulk crystalline defects/dislocations, several buffer structural designs have been employed such as the graded/stepped AlGaN layer configurations [11,12,13], AlN/GaN or AlGaN/GaN superlattices (SL) [14,15,16], epitaxial lateral overgrowth (ELOG) technique [17], and the three-dimensional (3D)-to-two-dimensional (2D) growth mode [18,19]. Especially, inserting a SiNx nano-mask into the matrix of GaN layer to promote the growth mode transition in GaN from 3D island-like overgrowth to 2D coalescence has been becoming a widely used technique [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33], in which the threading dislocations (TDs) are efficiently bent or even annihilate each other. There are several approaches to investigate the effect of SiNx nano-masks on the TD reduction such as varying either the SiNx deposition temperature or time to modulate the SiNx nano-mask configuration [19,20,21,22], using different growth conditions of the (Al)GaN layer overgrown on the SiNx nano-mask [23,24,25], or inserting multi-SiNx nano-mask into the (Al)GaN buffer layers [26,27,28]. Among these approaches, configuring a single SiNx nano-mask is considered the simplest strategy. By adjusting the growth time and temperature, the grain size and distribution of the SiNx nano-mask are modulated in such a way that the 3D-to-2D growth mode transformation is promoted [20,32]. However, any excess of the growth temperature and growth time will enhance the surface desorption of the GaN underlayer, leading to the aggregation of Ga atoms [33]. As a result, the threading dislocation density (TDD), etch pit density (EPD), and surface roughness of the GaN layers would increase, causing detrimental effects on the HEMT device performance. On the other hand, if the SiNx nano-mask is grown at too low temperature or short duration, its grain size would be tiny and ineffective to generate the transition of 3D-to-2D growth mode as well as to suppress the TDD. It is therefore essential to study an alternative method to solve the surface desorption issue, while retaining the efficiency of the inserted-SiNx nano-mask, in terms of suppressing TDs and improving two-dimensional electron gas (2DEG) properties of the GaN HEMT structure.
For this conversation, we present inserting a SiNx nano-mask into the low-temperature GaN (LT-GaN) layer grown by metalorganic chemical vapor deposition (MOCVD), where the nano-mask formation was controlled via the SiH4 flow rate variation instead of either the growth temperature or the growth time. Thus, this novel method allows easily controlling the SiNx configuration, enhancing the annihilation of TDs. Importantly, its low temperature and short-time growth were expected to highly limit the surface desorption of GaN underlayer. The results in this study show that the crystalline quality and the transport properties of GaN-based HEMT structure were strongly governed by varying the SiH4 flow rate.

2. Experimental

The full structures of GaN-based HEMT were grown on 6” Si (111) substrates by the G4 MOCVD system (Aixtron, Herzogenrath, Germany). The conventional precursors including TMGa, TMAl, NH3, and silane (1% diluted-SiH4) were used for the Ga, Al, N, and Si source, respectively. In order to grow the lattice-matched GaN layer on the Si (111) substrate, an AlN nucleation layer of about 300 nm was initially grown to prevent Ga–Si melt-back etching, followed by a transition layer of ~1.8 µm. This is to effectively modulate stress and prevent misalignment. A semi-insulating LT-GaN buffer layer with a thickness of 1.8 µm was then deposited at 990 °C, followed by a 2DEG heterostructure. This 2DEG structure was composed of a 300 nm HT-GaN (1020 °C) channel layer, 0.5 nm AlN spacer layer, and 10 nm Al0.25Ga0.75N barrier layer as shown in Figure 1. Finally, the structure was capped by a 4 nm thin GaN layer to protect the wafer surface from oxidation. To study the effect of the SiNx nano-mask on reducing the stress and TDD, the SiNx nano-mask was inserted at 300 nm above the LT-GaN/transition layer interface as can be seen in Figure 1. The growth time of the nano-mask was 1 min, while the growth temperature and the ammonia gas flow were kept at the same with that of the LT-GaN layer. Herein, a highly concentrated 1%-SiH4 source was employed to conduct a high deposition rate and low surface desorption. Five samples with different SiH4 source flow rates of 0, 25, 50, 75, and 100 sccm were used and denoted as samples A, B, C, D, and E, respectively. Sample A is a control sample, where no SiNx nano-mask was inserted. The in situ reflectivity and curvature measurement were monitored simultaneously by the LayTec EpiCurve-TT (LayTec, Berlin, Germany) unit integrated inside the MOCVD chamber. The 633 nm light source and 650 nm laser beam were used for reflectivity and curvature characterization, respectively. The experimental etching pit density (EPD) was performed in molten-KOH (85%-KOH:H2O = 1:5) in 6 min at 220 °C. The pit density was then calculated from atomic force microscopes (AFM) (NT-MDT Spectrum Instruments, Moscow, Russia) imaging on scan areas of (5 × 5) μm2. The surface morphology of the samples before and after the etching process was carefully observed by JSM7001F (JEOL, Tokyo, Japan) scanning electron microscope (SEM) and AFM. The TDD was evaluated from the full width at half maximum (FWHM) XRD-PANalytical X’Pert PRO MRD (Malvern Panalytical, Almelo, The Netherlands), scanned on (002) and (102) planes, the etch pit density (EPD) technique, and G2 F-20 STEM (FEI TecnaiTM, Hillsboro, OR, USA). The wafer bowing measurement was also carried out after the growth by ADE 9700 equipment. To investigate the 2DEG properties, the standard Hall effect measurement (BioRad HL5500, Hercules, CA, USA) was conducted at room temperature (RT) under a magnetic field of 0.57 T.

3. Results and Discussion

The in situ monitoring of the reflectivity and surface curvature of all five samples is shown in Figure 2. The growth rate and surface morphology of the samples during the epitaxial growth could be evaluated from the reflectivity measurement (Figure 2a). Before the SiNx nano-mask growth, the growth rate of ~4.2 μm/h of the 300 nm thick LT-GaN initial layer of all samples was indicated from the dependence of reflectance oscillation on the growth time. During the SiNx nano-mask insertion, the reflectance intensity of the samples was extinguished and flattened because of depositing 3D SiNx nanocrystals. Then, the maximum intensity of the oscillation peaks was gradually restored during the 1.5 μm LT-GaN overlayer growth, where the required time to recover the reflectance peak intensity to maximum, as before the SiNx deposition, was different between the inserted SiNx samples. Obviously, the recovering time, marked as the length of colored arrows in Figure 2a, increased with increasing the SiH4 flow rate. In other words, the time required for the growth mode transition from 3D-island to 2D-coalescence of the LT-GaN overlayer strongly depends on the SiH4 flux, in which transition time increases with the amount of SiH4 flux used. It is also emphasized that this transition could not complete if the SiH4 flow rate exceeds a threshold of 75 sccm, as in the case of sample E. Indeed, the reflectivity at the end of the LT-GaN overlayer of this sample only equals ~40% as compared to that at end of the LT-GaN underlayer. However, the following high-temperature GaN (HT-GaN) channel layer growth of this sample exhibited a fast restoration in the reflectivity only after three oscillation cycles (see blue-dash arrow in Figure 2a, sample E). This result suggests that the SiH4 flux modulation during inserting SiNx interlayer is a considerable approach in terms of recovering the 2D growth mode of the HT-GaN layer even if a SiH4 flux as high as 100 sccm is used.
To study the surface morphology with varying SiH4 flow rate, scanning electron microscopy (SEM) and atomic force microscopy (AFM) were employed as shown in Figure 3a–j. No V-defects were observed on the surfaces of samples B and C, while a few defects were visible on both sample D and sample E. The SEM image (Figure 3d) of sample D exposes several small pits. These pits are well known as stacking mismatch boundary (SMB) defects, attributed to the atom alignment faults during the coalescence process. Meanwhile, the surface of sample E displays the additional appearance of hexagonal micro-pit (HMP) defects, which are caused by the incomplete coalescence process as mentioned in the reflectivity discussion (Figure 2a) [34]. The presence of SMB and HMP defects on these samples is expected to decline the transport properties of 2DEG, which is discussed later. On the other hand, the surface morphologies of the inserted SiNx samples were preserved as the control sample with a root mean square (RMS) roughness of around 0.22 nm as shown in Figure 3f–j. This result indicates clearly that the surface desorption of LT-GaN underlayer in our study was restrained efficiently during SiNx interlayer growth.
Figure 2b shows that all five samples were suffering compressive strains, where the strain in the transition layer and 300 nm LT-GaN underlayer increased slowly with the growth time. After inserting the SiNx nano-mask, the compressive strain in the LT-GaN overlayer of the SiNx samples (samples B–D) increased in the same fashion, quickly approaching the strain magnitude of sample A. This is in agreement with the reflectivity results, where a transition from 3D island-like to 2D growth mode in the following LT-GaN layers was taking place. Furthermore, note that the strain increasing rate of sample E was much lower than that of other samples as can be seen in Figure 2b. Thus, the compressive strain of sample E (highest SiH4 flow rate of 100 sccm) was relieved more effectively than in other samples. That is because the transformation from 3D to 2D growth mode of the LT-GaN overlayer is not completed in this sample, as the reflectivity results indicated (Figure 2a). Indeed, it is directly related to the morphology as revealed by SEM (Figure 3e), where the long 3D growth mode duration of the LT-GaN overlayer and the incomplete coalescence of the HT-GaN channel layer resulted in the formation of HMP defects. Moreover, measurements of curvature showed that strain of layers is strongly affected by SiH4 flow rate that is also in agreement with the ex situ wafer bow measurements, in which the positive warp of sample A, B, C, D, and E were 22.0, 15.7, 15.7, 14.8, and 9.6 μm, respectively. Obviously, the warp of 9.6 µm presented in sample E is much smaller than that of other samples.
In order to investigate the relationship between the TDD and the SiH4 flow rate, the EPD and X-ray diffraction (XRD) techniques were employed, and the results are summarized in Table 1. The omega XRD scan of the sample near (002) and (102) plane could be found in Figure S1. The EPD slightly decreased as the SiH4 flow rate was increased from 25 to 75 sccm. However, it rapidly reduced to 3.24 × 108 cm−2 when the SiH4 flow rate increased further to 100 sccm (sample E). Of course, the EPD directly relates to the TDD in the epilayers. Herein, the TD in GaN growth could be categorized into two types as screw-type TD and edge-type TD. The TDD of each type presented in the GaN layer can be identified from the FWHM of (002) and (102) X-ray rocking curves (i.e., β(002) and β(102)) via the following formulas [11].
D s c r e w = β ( 002 ) 2 4.35 × b s c r e w 2
D e d g e = β ( 102 ) 2 β ( 002 ) 2 4.35 × b e d g e 2
The screw-type TDD (Dscrew) and edge-type TDD (Dedge) are computed from β(002), β(102), and burger vector length (bscrew = 0.5185 nm and bedge = 0.3189 nm). The extracted TDDs of all samples are displayed in Table 1. Interestingly, the edge-type TDD decreased considerably with the increasing SiH4 flow rate, whereas the screw-type TDD showed a slight drop as SiH4 flow is more than 75 sccm. Thus, the increasing SiH4 flow rate during SiNx nano-mask forming has a dominant effect on the decrease in edge-type TDD rather than screw-type TDD. We notice that this result plays an important role in improving the 2DEG properties of the AlGaN/GaN HEMT structure.
To further understand the mechanisms of the decreasing edge-type TDD, the cross-sectional bright-field scanning transmission electron microscopy (STEM) micrograph of sample E was taken as can be seen in Figure 4. The edge dislocations, screw dislocations, and mixed dislocations (edge and screw dislocations) are labeled as E, S, and M, respectively. By the comparison between two STEM images observed along g = [ 0002 ] and g = [ 11 ¯ 20 ] zone axis, the edge dislocation density is distinctly reduced. Apparently, the propagation of most of TDs was terminated or bent at the SiNx interlayer and further annihilated during the 2D lateral overgrowth on 3D island-like LT-GaN overlayer, which is marked as white-dash lines in Figure 4. Three kinds of interaction between edge-type TDs are visible: (1) Type I is parallel, two TDs passed through the SiNx nano-mask with the same Burgers vector. (2) Type II is fusion, two TDs combined to form a new TD where its Burgers vector equals to the sum of two componential Burgers vectors. (3) Type III is annihilation, two TDs with the opposite Burgers vectors react against each other [33,35]. We notice that only type II and III interactions benefit the annihilation of TDs.
The effect of the SiH4 flow rate on the transport properties of 2DEG HEMT structures was carried out by van der Pauw–Hall measurements at room temperature (RT) and the results are shown in Figure 5. Below a SiH4 flow rate of 50 sccm, both the mobility and the sheet carrier concentration (NS) of 2DEG structure enhanced simultaneously with increasing the SiH4 flow rate. This behavior could be explained by the decrease in the edge-type TDs in the structure as observed from XRD results [32,36]. The edge-type TDs are usually accompanied by the dangling bonds. These dangling bonds generate an acceptor-like trap level in the band structure, capturing free charges and then forming negatively charged Coulombic scattering centers. Thus, any reduction in edge-type TDD is related to an improvement of the 2DEG mobility [37]. Meanwhile, the NS was also influenced by the edge-type TDD. The increase of NS with decreasing edge-type TDs is a result of the restrained charges at the acceptor-like traps [36,38]. As the SiH4 flow rate was increased further to 75 sccm (sample D), both the mobility and NS significantly degraded. It could be due to the formation of the nonuniform distribution of SMB defects as can be seen in Figure 3d. The appearance of SMB defects reflects the interface roughness and electrical field fluctuation at the 2DEG structure, leading to a decrease in mobility and NS [39,40,41]. A further decrease in 2DEG mobility of sample E is understandable, ascribed to the formation of not only SMB but also HMP defects as can be observed in Figure 3e. Interestingly, the NS of this sample increased as compared to sample D. This may be explained by the active diffusion of Si adatoms from the SiNx nanostructures. Under a SiH4 flow rate as high as 100 sccm, the excess Si adatoms from the SiNx nano-mask could diffuse and precipitate on the sidewalls of the HMP defects via the TDs [26,33]. In addition, we notice that this diffusion process could be even boosted under a 3D growth mode of GaN channel layer as observed in our case (see sample E in Figure 2a). The Si adatoms preferentially sticking on the sidewalls of the HMP defects play a role as donors and releases electrons to the GaN channel layer. As a result, it causes an increase in the NS of the 2DEG structure. Besides that, the Si diffusion and incorporation into TDs could be a pathway causing unexpected leakage currents. Similarly, this effect was also observed in the Mg-doped GaN film. The segregated Mg propagates through TDs and incorporates at the boundaries of pyramidal inversion domains (PIDs) to form an Mg-rich area, degrading device performance [42,43,44]. Up to now, the Si diffusion originated from the SiNx nano-mask, and how it involves in the 2DEG characteristics as well as GaN-based HEMT device performance have not been clearly understood yet, and needs to be explored further. Herein, we for the first time demonstrate that the V-defects existing in 2DEG structure can cause the reversion of its NS as a SiH4 overflow rate (>75 sccm) used during the SiNx nano-mask growth.

4. Conclusions

In conclusion, the fabrication of AlGaN/AlN/GaN HEMT structures with high mobility and NS has been demonstrated via modulating the SiH4 flow rate of the SiNx nano-mask. The surface roughness of the samples was maintained at 0.22 nm, proving an effective elimination of surface desorption issue from the LT-GaN underlayer during inserting the SiNx nano-mask. More importantly, the SiNx nano-mask effectively contributed stress relaxation via promoting 3D-to-2D growth mode transformation of the LT-GaN overlayer and bending or annihilating TDs. Thus, it helped to reduce the edge-type TDD and EPD of the GaN-based HEMT structure as low as 2.25 × 109 and 3.24 × 108 cm−2, respectively, as observed in the sample grown using 100 sccm SiH4 flow rate. However, when the SiH4 flow rate was larger than 50 sccm, an oversized SiNx nano-mask could be presented, resulting in the development of V-defects as SMB and HMP defects. These defects would be located at or near the 2DEG structure, causing degradation of the 2DEG mobility and sheet carrier concentration. Consequently, the highest mobility and NS of the sample can be achieved to be ~1970 cm2/V·s and 6.42 × 1012 cm−2, respectively, under an optimized SiH4 flow rate of 50 sccm. Interestingly, the NS of sample E exhibited a reverse trend, which may be attributed to the accumulation of the diffused Si adatoms from the SiNx nano-mask on the sidewalls of HMP defects as the SiH4 flux was used as high as 100 sccm.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-6412/11/1/16/s1, Figure S1: Normalized XRD rocking curves of GaN, (a) (002) and (b) (102) planes. (c) FWHM as a function of SiH4 flow rate of (002) and (102) planes.

Author Contributions

J.-J.D. conceived and designed the experiments, performed the epitaxial growth, and wrote the manuscript. C.-W.H. and R.X. supported the epitaxial growth. C.-W.L., S.-K.W., J.-G.J., and S.-A.Y. supported structural, optical, and electrical characterizations. S.-H.H. performed revising the manuscript and analyzing data. T.T.M. and H.-C.W. provided suggestions and discussion. W.-C.C. is the advisor who supervised the experiments. All authors contributed to discuss the results and comment on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Council, Taiwan, R.O.C., grant No. MOST 108-2112-M-009-005.

Data Availability Statement

Data is contained within the article or supplementary material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Amano, H.; Baines, Y.; Beam, E.; Borga, M.; Bouchet, T.; Chalker, P.R.; Charles, M.; Chen, K.J.; Chowdhury, N.; Chu, R.; et al. The 2018 GaN power electronics roadmap. J. Phys. D Appl. Phys. 2018, 51, 163001. [Google Scholar] [CrossRef]
  2. Chen, K.J.; Haberlen, O.; Lidow, A.; Tsai, C.L.; Ueda, T.; Uemoto, Y.; Wu, Y. GaN-on-Si power technology: Devices and applications. IEEE Trans. Electron Devices 2017, 64, 779–795. [Google Scholar] [CrossRef]
  3. Axelsson, O.; Gustafsson, S.; Hjelmgren, H.; Rorsman, N.; Blanck, H.; Splettstoesser, J.; Thorpe, J.; Roedle, T.; Thorsell, M. Application relevant evaluation of trapping effects in AlGaN/GaN HEMTs with Fe-doped buffer. IEEE Trans. Electron Devices 2016, 63, 326–332. [Google Scholar] [CrossRef]
  4. Hu, A.; Yang, X.; Cheng, J.; Song, C.; Zhang, J.; Feng, Y.; Ji, P.; Xu, F.; Zhang, Y.; Yang, Z.; et al. Vertical leakage induced current degradation and relevant traps with large lattice relaxation in AlGaN/GaN heterostructures on Si. Appl. Phys. Lett. 2018, 112, 032104. [Google Scholar] [CrossRef] [Green Version]
  5. Wespel, M.; Polyakov, V.M.; Dammann, M.; Reiner, R.; Waltereit, P.; Quay, R.; Mikulla, M.; Ambacher, O. Trapping effects at the drain edge in 600 V GaN-on-Si HEMTs. IEEE Trans. Electron Devices 2015, 63, 598–605. [Google Scholar] [CrossRef]
  6. Yang, S.; Zhou, C.; Han, S.; Wei, J.; Sheng, K.; Chen, K.J. Impact of substrate bias polarity on buffer-related current collapse in AlGaN/GaN-on-Si power devices. IEEE Trans. Electron Devices 2017, 64, 5048–5056. [Google Scholar] [CrossRef]
  7. Meneghini, M.; Rossetto, I.; Bisi, D.; Stocco, A.; Chini, A.; Pantellini, A.; Lanzieri, C.; Nanni, A.; Meneghesso, G.; Zanoni, E. Buffer traps in Fe-doped AlGaN/GaN HEMTs: Investigation of the physical properties based on pulsed and transient measurements. IEEE Trans. Electron Devices 2014, 61, 4070–4077. [Google Scholar] [CrossRef]
  8. Liang, Y.; Jia, L.; He, Z.; Fan, Z.; Zhang, Y.; Yang, F. The study of the contribution of the surface and bulk traps to the dynamic Rdson in AlGaN/GaN HEMT by light illumination. Appl. Phys. Lett. 2016, 109, 182103. [Google Scholar] [CrossRef]
  9. Liu, S.; Yang, S.; Tang, Z.; Jiang, Q.; Liu, C.; Wang, M.; Shen, B.; Chen, K.J. Interface/border trap characterization of Al2O3/AlN/GaN metal-oxide-semiconductor structures with an AlN interfacial layer. Appl. Phys. Lett. 2015, 106, 051605. [Google Scholar] [CrossRef]
  10. Hua, M.; Lu, Y.; Liu, S.; Liu, C.; Fu, K.; Cai, Y.; Zhang, B.; Chen, K.J. Compatibility of AlN/SiNx passivation with LPCVD-SiNx gate dielectric in GaN-based MIS-HEMT. IEEE Electron Device Lett. 2016, 37, 265–268. [Google Scholar] [CrossRef]
  11. Lee, H.-P.; Perozek, J.; Rosario, L.D.; Bayram, C. Investigation of AlGaN/GaN high electron mobility transistor structures on 200-mm silicon (111) substrates employing different buffer layer configurations. Sci. Rep. 2016, 6, 37588. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, M.-H.; Do, Y.-G.; Kang, H.C.; Noh, D.Y.; Park, S.-J. Effects of step-graded AlxGa1−xN interlayer on properties of GaN grown on Si(111) using ultrahigh vacuum chemical vapor deposition. Appl. Phys. Lett. 2001, 79, 2713–2715. [Google Scholar] [CrossRef]
  13. Cheng, J.; Yang, X.; Sang, L.; Guo, L.; Hu, A.; Xu, F.; Tang, N.; Wang, X.; Shen, B. High mobility AlGaN/GaN heterostructures grown on Si substrates using a large lattice-mismatch induced stress control technology. Appl. Phys. Lett. 2015, 106, 142106. [Google Scholar] [CrossRef]
  14. Feltin, E.; Beaumont, B.; Laügt, M.; De Mierry, P.; Vennéguès, P.; Lahrèche, H.; Leroux, M.; Gibart, P. Stress control in GaN grown on silicon (111) by metalorganic vapor phase epitaxy. Appl. Phys. Lett. 2001, 79, 3230–3232. [Google Scholar] [CrossRef]
  15. Selvaraj, S.L.; Suzue, T.; Egawa, T. Breakdown enhancement of AlGaN/GaN HEMTs on 4-in silicon by improving the GaN quality on thick buffer layers. IEEE Electron Device Lett. 2009, 30, 587–589. [Google Scholar] [CrossRef]
  16. Sugawara, Y.; Ishikawa, Y.; Watanabe, A.; Miyoshi, M.; Egawa, T. Characterization of dislocations in GaN layer grown on 4-inch Si (111) with AlGaN/AlN strained layer superlattices. Jpn. J. Appl. Phys. 2016, 55, 05FB08. [Google Scholar] [CrossRef]
  17. Tanaka, S.; Honda, Y.; Sawaki, N.; Hibino, M. Structural characterization of GaN laterally overgrown on a (111)Si substrate. Appl. Phys. Lett. 2001, 79, 955–957. [Google Scholar] [CrossRef]
  18. Chang, S.; Wei, L.L.; Luong, T.T.; Chang, C.; Chang, L. Threading dislocation reduction in three-dimensionally grown GaN islands on Si (111) substrate with AlN/AlGaN buffer layers. J. Appl. Phys. 2017, 122, 105306. [Google Scholar] [CrossRef]
  19. Lee, K.J.; Shin, E.H.; Lim, K.Y. Reduction of dislocations in GaN epilayers grown on Si(111) substrate using SixNy inserting layer. Appl. Phys. Lett. 2004, 85, 1502. [Google Scholar] [CrossRef]
  20. Hertkorn, J.; Lipski, F.; Brückner, P.; Wunderer, T.; Thapa, S.B.; Scholz, F.; Chuvilin, A.; Kaiser, U.; Beer, M.; Zweck, J. Process optimization for the effective reduction of threading dislocations in MOVPE grown GaN using in situ deposited masks. J. Cryst. Growth 2008, 310, 4867–4870. [Google Scholar] [CrossRef]
  21. Dadgar, A.; Poschenrieder, M.; Reiher, A.; Bläsing, J.; Christen, J.; Krtschil, A.; Finger, T.; Hempel, T.; Diez, A.; Krost, A. Reduction of stress at the initial stages of GaN growth on Si(111). Appl. Phys. Lett. 2003, 82, 28–30. [Google Scholar] [CrossRef]
  22. Cheng, K.; Leys, M.; DeGroote, S.; Germain, M.; Borghs, G. High quality GaN grown on silicon(111) using a SixNy interlayer by metal-organic vapor phase epitaxy. Appl. Phys. Lett. 2008, 92, 192111. [Google Scholar] [CrossRef]
  23. Scholz, F.; Forghani, K.; Klein, M.; Klein, O.; Kaiser, U.; Neuschl, B.; Tischer, I.; Feneberg, M.; Thonke, K.; Lazarev, S.; et al. Studies on defect reduction in AlGaN heterostructures by integrating an in-situ SiN interlayer. Jpn. J. Appl. Phys. 2013, 52, 08JJ07. [Google Scholar] [CrossRef]
  24. Forghani, K.; Klein, M.; Lipski, F.; Schwaiger, S.; Hertkorn, J.; Leute, R.A.R.; Scholz, F.; Feneberg, M.; Neuschl, B.; Thonke, K.; et al. High quality AlGaN epilayers grown on sapphire using SiNx interlayers. J. Cryst. Growth 2011, 315, 216–219. [Google Scholar] [CrossRef]
  25. Neuschl, B.; Fujan, K.J.; Feneberg, M.; Tischer, I.; Thonke, K.; Forghani, K.; Klein, M.; Scholz, F. Cathodoluminescence and photoluminescence study on AlGaN layers grown with SiNx interlayers. Appl. Phys. Lett. 2010, 97, 192108. [Google Scholar] [CrossRef]
  26. Zang, K.Y.; Wang, Y.; Wang, L.S.; Chow, S.Y.; Chua, S.J. Defect reduction by periodic SiNx interlayers in gallium nitride grown on Si (111). J. Appl. Phys. 2007, 101, 093502. [Google Scholar] [CrossRef]
  27. Wang, T.-Y.; Ou, S.-L.; Horng, R.-H.; Wuu, D.-S. Improved GaN-on-Si epitaxial quality by incorporating various SixNy interlayer structures. J. Cryst. Growth 2014, 399, 27–32. [Google Scholar] [CrossRef]
  28. Xie, J.; Chevtchenko, S.A.; Özgür, Ü.; Morkoç, H. Defect reduction in GaN epilayers grown by metal-organic chemical vapor deposition with in situ SiNx nanonetwork. Appl. Phys. Lett. 2007, 90, 262112. [Google Scholar] [CrossRef] [Green Version]
  29. Klein, O.; Biskupek, J.; Forghani, K.; Scholz, F.; Kaiser, U. TEM investigations on growth interrupted samples for the correlation of the dislocation propagation and growth mode variations in AlGaN deposited on SiNx interlayers. J. Cryst. Growth 2011, 324, 63–72. [Google Scholar] [CrossRef]
  30. Kappers, M.J.; Datta, R.; Oliver, R.A.; Rayment, F.D.G.; Vickers, M.E.; Humphreys, C.J. Threading dislocation reduction in (0001) GaN thin films using SiNx interlayers. J. Cryst. Growth 2007, 300, 70–74. [Google Scholar] [CrossRef]
  31. Riemann, T.; Hempel, T.; Christen, J.; Veit, P.; Clos, R.; Dadgar, A.; Krost, A.; Haboeck, U.; Hoffmann, A. Optical and structural microanalysis of GaN grown on SiN submonolayers. J. Appl. Phys. 2006, 99, 123518. [Google Scholar] [CrossRef] [Green Version]
  32. Lee, Y.S.; Chung, S.J.; Suh, E.-K. Effects of dislocations on the carrier transport and optical properties of GaN films grown with an in-situ SiNx insertion layer. Electron. Mater. Lett. 2012, 8, 141–146. [Google Scholar] [CrossRef]
  33. Wang, T.Y.; Ou, S.L.; Horng, R.-H.; Wuu, D.-S. Growth evolution of SixNy on the GaN underlayer and its effects on GaN-on-Si (111) heteroepitaxial quality. CrystEngComm 2014, 16, 5724–5731. [Google Scholar] [CrossRef] [Green Version]
  34. Szymański, T.; Wośko, M.; Wzorek, M.; Paszkiewicz, B.; Paszkiewicz, R. Origin of surface defects and influence of an in situ deposited SiN nanomask on the properties of strained AlGaN/GaN heterostructures grown on Si (111) using metal–organic vapour phase epitaxy. CrystEngComm 2016, 18, 8747–8755. [Google Scholar] [CrossRef]
  35. Contreras, O.; Ponce, F.A.; Christen, J.; Dadgar, A.; Krost, A. Dislocation annihilation by silicon delta-doping in GaN epitaxy on Si. Appl. Phys. Lett. 2002, 81, 4712–4714. [Google Scholar] [CrossRef]
  36. Weimann, N.G.; Eastman, L.F.; Doppalapudi, D.; Ng, H.M.; Moustakas, T.D. Scattering of electrons at threading dislocations in GaN. J. Appl. Phys. 1998, 83, 3656–3659. [Google Scholar] [CrossRef]
  37. Chen, K.X.; Dai, Q.; Lee, W.; Kim, J.K.; Schubert, E.F.; Grandusky, J.; Mendrick, M.; Li, X.; Smart, J.A. Effect of dislocations on electrical and optical properties of n-type Al0.34Ga0.66N. Appl. Phys. Lett. 2008, 93, 192108. [Google Scholar] [CrossRef]
  38. Krtschil, A.; Dadgar, A.; Krost, A. Decoration effects as origin of dislocation-related charges in gallium nitride layers investigated by scanning surface potential microscopy. Appl. Phys. Lett. 2003, 82, 2263–2265. [Google Scholar] [CrossRef]
  39. Asgari, A.; Babanejad, S.; Faraone, L. Electron mobility, Hall scattering factor, and sheet conductivity in AlGaN/AlN/GaN heterostructures. J. Appl. Phys. 2011, 110, 113713. [Google Scholar] [CrossRef]
  40. Xu, X.; Liu, X.; Han, X.; Yuan, H.; Wang, J.; Guo, Y.; Song, H.; Zheng, G.; Wei, H.; Yang, S.; et al. Dislocation scattering in AlxGa1−xN/GaN heterostructures. Appl. Phys. Lett. 2008, 93, 182111. [Google Scholar] [CrossRef]
  41. Li, H.; Liu, G.; Wei, H.; Jiao, C.; Wang, J.; Zhang, H.; Jin, D.-D.; Feng, Y.; Yang, S.; Wang, L.; et al. Scattering due to Schottky barrier height spatial fluctuation on two dimensional electron gas in AlGaN/GaN high electron mobility transistors. Appl. Phys. Lett. 2013, 103, 232109. [Google Scholar] [CrossRef]
  42. Iwata, K.; Narita, T.; Nagao, M.; Tomita, K.; Kataoka, K.; Kachi, T.; Ikarashi, N. Atomic resolution structural analysis of magnesium segregation at a pyramidal inversion domain in a GaN epitaxial layer. Appl. Phys. Express 2019, 12, 031004. [Google Scholar] [CrossRef]
  43. Duguay, S.; Echeverri, A.; Castro, C.; Latry, O. Evidence of Mg segregation to threading dislocation in normally-off GaN-HEMT. IEEE Trans. Nanotechnol. 2019, 18, 995–998. [Google Scholar] [CrossRef]
  44. Usami, S.; Mayama, N.; Toda, K.; Tanaka, A.; Deki, M.; Nitta, S.; Honda, Y.; Amano, H. Direct evidence of Mg diffusion through threading mixed dislocations in GaN p–n diodes and its effect on reverse leakage current. Appl. Phys. Lett. 2019, 114, 232105. [Google Scholar] [CrossRef]
Figure 1. Schematic structure of GaN-based HEMT with inserted SiNx nano-masks grown at different SiH4 flow rates.
Figure 1. Schematic structure of GaN-based HEMT with inserted SiNx nano-masks grown at different SiH4 flow rates.
Coatings 11 00016 g001
Figure 2. In situ (a) reflectivity and (b) curvature monitoring of sample surfaces.
Figure 2. In situ (a) reflectivity and (b) curvature monitoring of sample surfaces.
Coatings 11 00016 g002
Figure 3. (ae) SEM and (fj) AFM images of the inserted SiNx nano-mask samples with various SiH4 flow rates.
Figure 3. (ae) SEM and (fj) AFM images of the inserted SiNx nano-mask samples with various SiH4 flow rates.
Coatings 11 00016 g003
Figure 4. (a) Bright-field cross-sectional STEM micrograph of GaN-based HEMT structure on Si (111) with the SiNx nano-mask grown under 100 sccm SiH4 flow rate and (b) Selective area electron diffraction pattern of the HT-GaN top layer taken along the [ 1 1 ¯ 00 ] zone axis.
Figure 4. (a) Bright-field cross-sectional STEM micrograph of GaN-based HEMT structure on Si (111) with the SiNx nano-mask grown under 100 sccm SiH4 flow rate and (b) Selective area electron diffraction pattern of the HT-GaN top layer taken along the [ 1 1 ¯ 00 ] zone axis.
Coatings 11 00016 g004
Figure 5. (a) Mobility and NS of the 2DEG structure as a function of the SiH4 flow rate. The inset (b) shows the dependence of sheet resistance (RS) on the SiH4 flow rate.
Figure 5. (a) Mobility and NS of the 2DEG structure as a function of the SiH4 flow rate. The inset (b) shows the dependence of sheet resistance (RS) on the SiH4 flow rate.
Coatings 11 00016 g005
Table 1. Dependence of dislocation density on the SiH4 flow rate by XRD and EPD analysis.
Table 1. Dependence of dislocation density on the SiH4 flow rate by XRD and EPD analysis.
Sample IDSiH4 Source
(sccm)
FWHM
(arcsec)
TDD
(cm−2)
EPD
(cm−2)
Wafer Bowing (µm)
(002)(102)Screw-TypeEdge-Type
A0466 ± 5945 ± 5(4.36 ± 0.09) × 108(3.59 ± 0.02) × 1096.52 × 10822.0
B25468 ± 5920 ± 5(4.40 ± 0.09) × 108(3.33 ± 0.02) × 1095.76 × 10815.7
C50466 ± 5908 ± 5(4.36 ± 0.09) × 108(3.23 ± 0.02) × 1095.52 × 10815.7
D75467 ± 5872 ± 5(4.38 ± 0.09) × 108(2.88 ± 0.02) × 1095.34 × 10814.8
E100441 ± 5786 ± 5(3.91 ± 0.09) × 108(2.25 ± 0.02) × 1093.24 × 1089.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dai, J.-J.; Liu, C.-W.; Wu, S.-K.; Huynh, S.-H.; Jiang, J.-G.; Yen, S.-A.; Mai, T.T.; Wen, H.-C.; Chou, W.-C.; Hu, C.-W.; et al. Improving Transport Properties of GaN-Based HEMT on Si (111) by Controlling SiH4 Flow Rate of the SiNx Nano-Mask. Coatings 2021, 11, 16. https://doi.org/10.3390/coatings11010016

AMA Style

Dai J-J, Liu C-W, Wu S-K, Huynh S-H, Jiang J-G, Yen S-A, Mai TT, Wen H-C, Chou W-C, Hu C-W, et al. Improving Transport Properties of GaN-Based HEMT on Si (111) by Controlling SiH4 Flow Rate of the SiNx Nano-Mask. Coatings. 2021; 11(1):16. https://doi.org/10.3390/coatings11010016

Chicago/Turabian Style

Dai, Jin-Ji, Cheng-Wei Liu, Ssu-Kuan Wu, Sa-Hoang Huynh, Jhen-Gang Jiang, Sui-An Yen, Thi Thu Mai, Hua-Chiang Wen, Wu-Ching Chou, Chih-Wei Hu, and et al. 2021. "Improving Transport Properties of GaN-Based HEMT on Si (111) by Controlling SiH4 Flow Rate of the SiNx Nano-Mask" Coatings 11, no. 1: 16. https://doi.org/10.3390/coatings11010016

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