Passivation effects of phosphorus on 4H-SiC (0001) Si dangling bonds: A first-principles study
Li Wenbo1, Li Ling2, Wang Fangfang2, Zheng Liu2, Xia Jinghua2, Qin Fuwen3, Wang Xiaolin1, Li Yongping2, Liu Rui2, Wang Dejun1, Pan Yan2, ‡, Yang Fei2, 4, §
School of Electronic Science and Technology, Faculty of Electronic Information and Electrical Engineering, Dalian University of Technology, Dalian 116024, China
Global Energy Interconnection Research Institute, Beijing 102211, China
State Key Laboratory of Materials Modification by Laser, Ion and Electron Beams (Ministry of Education), Dalian University of Technology, Dalian 116024, China
State key Laboratory of Advanced Power Transmission Technology, Beijing 102211, China

 

† Corresponding author. E-mail: panyan@geiri.sgcc.com.cn yangsenji@163.com

Abstract

The effect of phosphorus passivation on 4H-SiC(0001) silicon (Si) dangling bonds is investigated using ab initio atomistic thermodynamic calculations. Phosphorus passivation commences with chemisorption of phosphorus atoms at high-symmetry coordinated sites. To determine the most stable structure during the passivation process of phosphorus, a surface phase diagram of phosphorus adsorption on SiC (0001) surface is constructed over a coverage range of 1/9–1 monolayer (ML). The calculated results indicate that the 1/3 ML configuration is most energetically favorable in a reasonable environment. At this coverage, the total electron density of states demonstrates that phosphorus may effectively reduce the interface state density near the conduction band by removing 4H-SiC (0001) Si dangling bonds. It provides an atomic level insight into how phosphorus is able to reduce the near interface traps.

1. Introduction

Silicon carbide (SiC) has a wide application in extremely harsh environments. Besides, it is also a promising wide-band-gap semiconductor material for fabrication of high-temperature, high-frequency, and high-power metal–oxide–semiconductor (MOS) devices due to its excellent physical properties as well as native oxide (SiO2).[14] Among the commercially available SiC polytypes, 4H-SiC is most attractive for high-power electronic devices such as MOS field effect transistors (MOSFETs) owing to its higher bulk electron mobility and smaller anisotropy.[5,6] However, the high density of interface traps (Dit) near the conduction band edge at the SiO2/SiC (0001) interface gives rise to the low channel mobility of 4H-SiC MOSFETs.[710] It is therefore important to passivate the interface to improve the channel mobility of SiC MOSFETs.

A large number of process methods have been tried to reduce the defect density to an acceptable level. It has been demonstrated experimentally that the interface state densities can be significantly reduced by annealing SiO2/SiC interfaces with nitrogen (N).[1121] Unfortunately, the obtained Dit ( 2 × 1011 cm−2·eV−1) after nitridation treatment is still too high for SiC MOS capacitors.[13] It is necessary to explore new passivation processes to achieve higher mobility for 4H-SiC MOSFETs. Okamoto et al. recently attempted to reduce the Dit by post-oxidation annealing in phosphoryl chloride (POCl3).[2225] Their results showed that the Dit was effectively reduced by phosphorus (P) atoms and the peak field-effect mobility of 4H-SiC MOSFETs could be increased to 89 cm2/V·s. Their subsequent studies suggested that the positive effects of P atoms might be attributed to the passivation of NITs, in particular, the elimination of Si–Si bonds.[26] However, atomic scale understanding of phosphorus at the interface has remained unclear.

Experiments have reported many kinds of defects exist at the SiO2/SiC interface. The dangling bond defect, as a simple interface defect, is also found at the SiO2/SiC interface,[27] similar to the SiO2/Si interface.[28] Pennington and Ashman recently used density functional theory (DFT) calculations to investigate nitrogen passivation of 4H-SiC (0001) Si-face dangling bonds.[1921] That work showed that the Si dangling bonds may create states below the conduction band edge and the dangling bond defect states may be removed by nitrogen. However, little is still known about the effects of phosphorus passivation on the dangling bonds. In this work, we study the effect of phosphorus passivation 4H-SiC (0001) Si dangling bonds by using the first-principles approach, in order to better understand the removal mechanism of the defect states.

2. Computation methods

DFT calculation is performed using Cambridge serial total energy package (CASTEP) code. The generalized gradient approximation (GGA) function proposed by Perdew and Wang (PW91)[29] and a plane-wave cutoff of 300 eV are employed after a full convergence test. The Vanderbilt ultrasoft pseudopotentials[30] are used to describe the ion–electron interaction. The equilibrium atomic positions are determined when the maximum Hellmann–Feynman force and the change of the total energy are below 0.03 eV/Å and 0.01 meV, respectively. The optimized lattice constants are a = 3.078 Å and c = 10.046 Å using a 9 × 9 × 2 Monkhorst–Pack grid, agreeing well with the experimental values (a = 3.073 Å and c = 10.053 Å).

The 3× 3 × 1 4H-SiC (0001) supercell consists of four bilayers of SiC with a vacuum space of 10 Å in dimension. The carbon dangling bonds at the bottom layer are terminated by hydrogen atoms and the bottom three layers are fixed to simulate the infinitely large bulk solid. A converged 3 × 3 × 1 Monkhorst–Pack k-point mesh[31] is used during the geometry optimizations of various surface models. On the SiC (0001) surface, P adatoms are initially placed on four high-symmetric adsorption sites as shown in Fig. 1. The adsorption energy per P adatom is calculated as follows:

(1)
where NP is the number of P atoms adsorbed on the 4H-SiC (0001) surface, and EP, Eslab, and EP/slab refer to the total energies of the isolated P atom, and the surfaces with and without the P adatoms, respectively.

Fig. 1. (color online) (a) Side and (b) top views of 4H-SiC (0001) surface and (c), (d) four adsorption sites on this surface. The on-top site (OT) is located on the top of a surface Si atom, the bridge site (BR) is a two-fold coordinated site between two surface Si atoms, the three-fold hollow site (H3) is above a carbon of the third bilayer, and the four-fold hollow site (T4) is above a carbon of the first bilayer. The orange, grey, magenta, and cyan spheres denote silicon, carbon, phosphorus, and hydrogen atoms, respectively.

The stability of each structure is assessed by calculating the Gibbs free energy of formation ΔGform in the grand canonical ensemble

(2)
where µP is the chemical potential of P atom. Combining Eqs. (1) and (2), one can obtain
(3)
with ΔµP = µP − EP.

3. Results and discussion

We start our analysis by studying P atom adsorption on the 4H-SiC (0001) surface. For a single P atom adsorption, corresponding to 1/9 monolayer (ML) coverage in our supercell model, only the H3 and T4 binding configurations are stable. According to the calculated result, the H3 site is more favorable since its adsorption energy is 0.19 eV higher than that of the configuration with a P atom at the T4 site. For the 2/9 ML coverage surface, five possible stable configurations are obtained after optimizing 37 initial structures. Unlike the nitrogen adsorption on 4H-SiC,[18] the most energetically favorable configuration is the one with two P atoms at different T4 sites rather than different H3 sites. It is interesting that the most stable location for P atoms shifts from the H3 site to T4 site. This might be attributed to the strong interaction between two P atoms which makes the P adatoms deviate from the H3 sites. When three P atoms are deposited on the supercell of 4H-SiC surface (i.e., 1/3 ML coverage), these P atoms still prefer separated T4 sites. It is also found that this coverage has a R30° pattern. Above 4/9 ML coverage, the P adatoms rearrange their bonding on the surface and form P2 dimer or P3 ring. As the coverage reaches 1 ML, three P3 rings appear symmetrically on the SiC surface. In this configuration, each P3 ring is located at the T4 site with three P–P bonds of about 2.215 Å.

The dependence of the adsorption energy per P atom on the phosphorus coverage is illustrated in Fig. 2. The positive adsorption energies suggest that the reactions between the P atoms and the surface atoms are exothermic. For the 1/9 ML configuration, the corresponding adsorption energy per P atom is 2.68 eV. With the increase of the coverage, the adsorption energy reaches maximum at 1/3 ML coverage, and then significantly decreases from 2.79 eV at 1/3 ML to 1.80 eV at 2/3 ML coverage. When the coverage exceeds 2/3 ML, the adsorption energy becomes relatively stable.

Fig. 2. The Gibbs free energy of formation as a function of phosphorus coverage.
Fig. 3. (color online) Adsorption energy per phosphorus adatom as a function of phosphorus chemical potential.

To determine the stability of various P/slab configurations, we plot the Gibbs free energy of formation as a function of phosphorus chemical potential using Eq. (3). This surface diagram describes the most stable structure at different chemical potentials related to a realistic gas environment. The formation energy of bare SiC surface (ΔGform = 0) is marked by the black horizontal line as a reference. When ΔµP < −2.79 eV, the bare 4H-SiC (0001) surface is the most stable structure. As the phosphorus chemical potential increases from −2.79 eV to −1.20 eV, the 1/3 ML configuration becomes most energetically favorable. When ΔµP > −1.20 eV, the 1 ML configuration is stable. Considering the fact that the P atom is difficult to diffuse into the SiC surface, only the 1/3 ML configuration may be observed as a stable structure. That the 1/3 ML adsorption configuration is energetically favorable may be simply explained as follows. On the 4HSiC surface, each surface Si atom has one unsaturated electron (i.e., a dangling bond). This gives rise to the surface with threefold symmetry adsorption sites, which are considered more favorable than the others. One P atom with three valence electrons and two lone-pair electrons is able to form a covalent bond with each of the three surface Si atoms. Thus, the surface dangling bonds could be fully saturated at 1/3 ML coverage. This picture can be further confirmed by the analysis of electron densities.

Figure 4 depicts the electron densities of the most stable P/4H-SiC (0001) configurations, in which the green and the blue parts represent the electron accumulation and depletion regions, respectively. Clearly, electrons transfer from the surface Si atoms to the P adatoms, which indicates the formation of P–Si bonds. At the 1/3 ML coverage, three P atoms can form nine P–Si bonds; thus the 4H-SiC (0001) surface can be completely saturated through post-oxidation annealing in POCl3.

Fig. 4. (color online) Different electron densities of the stable configurations for the coverages of (a) 1/9 ML, (b) 2/9 ML, (c) 1/3 ML, (d) 4/9 ML, (e) 5/9 ML, (f) 2/3 ML, (g) 7/9 ML, (h) 8/9 ML, (i) 1 ML. The violet spheres denote phosphorus atoms.

To further elucidate the phosphorus passivation effects on the density of interface traps, we present the electron densities of states (DOS) for the clean surface, 1/9 ML, 1/3 ML, and 4/9 ML coverage as shown in Fig. 5. For the clean surface, the calculated DOS near the conduction band is high enough in the band gap, which agrees with the experimental observation.[14] Comparing the DOS for different coverage, we find that the DOS near the conduction band decreases with the coverage increasing from 1/9 to 1/3 ML, and then increases again when the coverage is higher than 1/3 ML. Here we are most interested in the 1/3 ML configuration since it is most energetically favorable in a reasonable physical condition. At 1/3 ML, the DOS inside the band gap is almost completely eliminated. In other words, the P adatoms play an effective passivation role for 4H-SiC (0001) Si dangling bonds.

Fig. 5. (color online) Total densities of states for the clean surface and surfaces with 1/9 ML, 1/3 ML, and 4/9 ML phosphorus coverages.
4. Conclusion

We investigate adsorption behaviors of P atoms on 4HSiC (0001) and effects of P passivation on surface Si dangling bonds by using density functional theory methods. The surface diagram result indicates that the 1/3 ML configuration is most energetically favorable (i.e., a reasonable phosphorus concentration). At this coverage, the density of surface states within the band gap of bare surface is found to be decreased. It also shows that P passivation removes states from the conduction band gap. This is similar to the previous experimental observation[22,26] even though a complicated oxide during P passivation is not theoretically considered. Therefore, the physical mechanisms of P passivation of the bare SiC surface may also play a role during P passivation of the SiO2/SiC interface. The presence of P atoms at the SiO2/SiC interface region can effectively eliminate Si dangling bonds and decrease the interface state density.

Reference
[1] Ruff M Mitlehner H Helbig R 1994 IEEE Trans. Electron Devices 41 1040
[2] Nakamura D Gunjishima I Yamaguchi S Ito T Okamoto A Kondo H Onda S Takatori K 2004 Nature 430 1009
[3] Melinon P Masenelli B Tournus F Perez A 2007 Nat. Mater. 6 479
[4] Presser V Nickel K G 2008 Crit. Rev. Solid State Mater. Sci. 33 1
[5] Cooper J A 1997 Phys. Status Solidi A 162 305
[6] Ventra M D Pantelides S T 1999 Phys. Rev. Lett. 83 1624
[7] Afanasév V V Bassler M Pensl G Schultz M 1997 Phys. Status Solidi A 162 321
[8] Afanasév V V Stesmans A 1997 Phys. Rev. Lett. 78 2437
[9] Knaup J M Deák P Frauenheim T Gali A Hajnal Z Choyke W J 2005 Phys. Rev. B 71 235321
[10] Zhu Q Huang L Li W Li S Wang D 2011 Appl. Phys. Lett. 99 082102
[11] Jia Y Lv H Niu Y Li L Song Q Tang X Li C Zhao Y Xiao L Wang L Tang G Zhang Y Zhang Y 2016 Phys. B 25 097101
[12] Li H Dimitrijev S Harrison H B Sweatman D 1997 Appl. Phys. Lett. 70 2028
[13] Dhar S Wang S Williams J R Pantelides S T Feldman L C 2005 MRS Bull. 30 288
[14] Chung G Y Tin C C Williams J R McDonald K Di Ventra M Pantelides S T Feldman L C Weller R A 2000 Appl. Phys. Lett. 76 1713
[15] Yano H Hirao T Kimoto T Matsunami H 2001 Appl. Phys. Lett. 78 374
[16] Wang S Dhar S Wang S Ahyi A C Franceschetti A Williams J R Feldman L C Pantelides S T 2007 Phys. Rev. Lett. 98 026101
[17] Deák P Knaup J M Hornos T Thill C Gali A Frauenheim T 2007 J. Phys. D: Appl. Phys. 40 6242
[18] Ashman C R Först C J Schwarz K Blöchl P E 2004 Phys. Rev. B 69 075309
[19] Pennington G Ashman C R 2007 Appl. Phys. Lett. 91 072106
[20] Ashman C R Pennington G 2008 Surf. Sci. 602 3617
[21] Ashman C R Pennington G 2009 Phys. Rev. B 80 085318
[22] Okamoto D Yano H Hatayama T Fuyuki T 2010 Appl. Phys. Lett. 96 203508
[23] Okamoto D Yano H Hirata K Hatayama T Fuyuki T 2010 IEEE Electron Device Lett. 31 710
[24] Okamoto D Yano H Kotake S Hirata K Hatayama T Fuyuki T 2010 MRS Symp. Proc. 1246-B06-06
[25] Okamoto D Yano H Hatayama T Fuyuki T 2012 Materials Science Forum 717–720 733
[26] Okamoto D Yano H Kotake S Hirata K Hatayama T Fuyuki T 2011 Materials Science Forum 679–680 338
[27] Cantin J L Von Bardeleben H J Shishkin Y Ke Y Devaty R P Choyke W J 2004 Phys. Rev. Lett. 92 015502
[28] Lenahan P M Conley J F 1998 J. Vac. Sci. Technol. B 16 2134
[29] Perdew J P Wang Y 1992 Phys. Rev. B 45 13244
[30] Vanderbilt D 1990 Phys. Rev. B 41 7892
[31] Monkhorst H J Pack J D 1976 Phys. Rev. B 13 5188