First-principles study of strain effect on the formation and electronic structures of oxygen vacancy in SrFeO2
Zhang Wei1, †, , Huang Jie2
Physics Group of Department of Criminal Science and Technology, Nanjing Forest Police College, Nanjing 210023, China
Department of Physics, Nanjing Normal University, Nanjing 210023, China

 

† Corresponding author. E-mail: zhangw@nfpc.edu.cn

Project supported by the Creative Plan Project of Nanjing Forest Police College, China (Grant Nos. 201512213045xy and 201512213007x).

Abstract
Abstract

Motivated by recent experimental observations of metallic conduction in the quasi-two-dimensional SrFeO2, we study the epitaxial strain effect on the formation and electronic structures of oxygen vacancy (Vo) by first-principles calculations. The bulk SrFeO2 is found to have the G-type antiferromagnetic ordering (G-AFM) at zero strain, which agrees with the experiment. Under compressive strain the bulk SrFeO2 keeps the G-AFM and has the trend of Mott insulator-metal transition. Different from most of the previous similar work about the strain effect on Vo, both the tensile strain and the compressive strain enhance the Vo formation. It is found that the competitions between the band energies and the electrostatic interactions are the dominant mechanisms in determining the Vo formation. We confirm that the Vo in SrFeO2 would induce the n-type conductivity where the donor levels are occupied by the delocalized dx2y2 electrons. It is suggested that the vanishing of n-type conductivity observed by the Hall measurement on the strained films are caused by the shift of donor levels into the conduction band. These results would provide insightful information for the realization of metallic conduction in SrFeO2.

1. Introduction

Transition-metal oxides (TMOs) have been widely studied because of the peculiar physics and the novel functions found in these materials. Among these TMOs, the Fe related oxides are one of the most extensively studied systems. Here, we focus on the newly synthesized compound SrFeO2, which is the first ternary earth alkaline oxoferrate.[1,2] Unlike the conventional ABO3 perovskite structure, the lack of apical oxygen made SrFeO2 have the infinite-layer structure with a square-planar coordination (Fig. 1). This layered structure was isotypic with SrCuO2, the parent compound for the high temperature (high-Tc) superconductor. In spite of this quasi-two-dimensional structure, SrFeO2 showed the three-dimensional G-type antiferromagnetic ordering (G-AFM) with a very high Neel temperature (TN = 473 K). According to the crystal field theory,[3] the down-spin electron of high-spin Fe2+(d6) was expected to occupy the degenerate (dxz, dyz) orbitals, and SrFeO2 should be subject to Jahn–Teller distortion when the temperature was lowered. However, in experiment, SrFeO2 was found to maintain the P4/mmm symmetry down to 4.2 K,[1] which showed a nostructural instability. Besides, the TN often decreased sharply when the dimensionality decreased. Such high TN in a two-dimensional system was unexpected, which attracted the attention of researchers. Xiang et al.[4] regarded that the orbital was the lowest energy orbital and the last down-spin electron occupied the orbital. Pruneda et al.[5] also found the double occupation of the orbital of Fe2+, but they attributed the occupations to the reduction of spin splitting of orbitals. Kawakami et al.[6] reported the high pressure induced spin transition in SrFeO2, where the high-spin Fe2+ (S = 2) changed to the intermediate-spin state (S = 1) at ∼33 GPa. This spin transition was simultaneously accompanied by an AFM–FM transition and a Mott insulator–metal transition. Ju et al.[7] found that SrFeO2 showed giant optical anisotropy. It was found[8] that the substitution of Fe by Co or Mn can keep the tetragonal structure unchanged up to the solubility limit of x = 0.3. The neutron diffraction[9] at 5 K revealed that the long-range magnetic order disappeared in SrFe0.5Ru0.5O2. Horigane et al.[10] found the suppression of magnetic coupling by in-plane buckling. It was also reported that the FeO2-terminated (001) surface underwent a magnetic reconstruction consisting of a spin-flip process.[11] Despite extensive research on this layered TMO, previous work mostly focused on the bulk SrFeO2, and very little was known about the defect SrFeO2. So far there has been no study of the oxygen vacancy (Vo) in SrFeO2. Vo is one of the fundamental and intrinsic defects in TMOs, and it often has a crucial impact on the properties of TMOs. For example, the Vo can enable ionic conductivity in the solid oxide fuel cell (SOFC) cathodes, oxygen sensors, and mixed conducting membranes.[12,13] Therefore, the lack of knowledge of Vo in SrFeO2 would inhibit the technological applications and fundamental understanding of this quasi-two-dimensional material.

Fig. 1. (a) The four-atom primitive cell of tetragonal SrFeO2. (b) The supercell of SrFeO2 with G-AFM. The arrows denote the up-spin and down-spin. (c) Local coordinate of Fe–O bonds.

On the other hand, the functional behavior of TMOs can be improved when they are prepared as thin films. The enhanced behavior is usually attributed to the strain introduced by the substrate with different lattice constants. The epitaxial strain would affect the bond lengths and/or the bond angles, which finally changes the electronic structures, the magnetic properties and so on. Importantly, as SrFeO2 was isotypic with SrCuO2 where the high-Tc could be induced by carrier doping,[14] the researchers concentrated on the realization of metallic conduction inSrFeO2 by chemical substitution.[1518] However, the obtained compounds were still insulators. Recently, the single-crystalline epitaxial SrFeO2 thin films were grown on the perovskite substrates with various lattice constants.[19] Metallic conduction was found on the lattice-matched KTaO3 (001) substrate, and Hall measurements showed that the conduction carriers are n-type. Katayama et al. confirmed that the metallic conduction was an intrinsic property of SrFeO2 film and speculated the carriers were supplied by Vo as there were Vo’s in their films. However, they did not observe n-type conductivity on the lattice-mismatched substrates. Thus, it is worth examining whether the Vo of SrFeO2 can induce the n-type conductivity and how the behavior of Vo is affected by the epitaxial strain. In this work, we provide a first-principles survey of the following questions. (i) Do the bulk SrFeO2 have a magnetic transition under strain? (ii) How difficult is it to form a Vo in SrFeO2? (iii) Does Vo act as electron donors to mobile conducting states at the Fermi level in SrFeO2? (iv) If the Vo in SrFeO2 induces the n-type conductivity, how does the strain affect the performance of the n-type conductivity?

The remainder of this paper is organized as follows. In Section 2 we describe the calculation method used throughout this work. In Section 3 we first study the properties of bulk SrFeO2 under strain, then discuss the strain induced changes of formation energy of Vo. After that we present our results for the electronic structures of Vo under strain. Finally, the conclusion is given in Section 4.

2. Computational method

First-principles calculations are performed using the projector augmented wave (PAW) method,[20] as implemented in the Vienna ab initio simulation package (VASP).[21,22] The exchange–correlation functional is treated by the generalized gradient approximation (GGA).[23] The Hubbard U correction[24] is applied to Fe 3d electrons. We use Ueff = 4 eV[5] to reproduce the experimentally determined electronic and magnetic properties of bulk SrFeO2. The valence electron configurations are Sr (4s24p65s2), Fe (3p63d63s2), and O (2s2p4). To model the Vo, we construct a (∼11.4 Å×11.4 Å×14 Å) supercell with 128 atoms. A plane-wave cutoff of 600 eV is used throughout. A 2×2×2 and a 4×4×4 k-mesh[25] are used for the structural relaxations and density of states (DOS) calculations, respectively. The crystal structures are relaxed until the inter-atomic forces are smaller than 0.01 eV/Å and the convergence criteria for energy is 10−5 eV. To simulate the epitaxial strain in a thin film constrained on a cubic (001)-oriented substrate, we vary the in-plane lattice constants (a = b) of the tetragonal cell, and relax the out-of-plane lattice constant c and all internal coordinates. The strained structures are relaxed until the stresses along the out-of-plane direction are less than 0.1 kbar. The strain is defined as ɛ = a/a0 − 1, where a0 corresponds to the optimized lattice constant without strain.

The formation energy (Eform) of Vo is calculated according to the following formula: Eform(ɛ, Vo) = Etot,VoEtot,bulk+μo, where μo is the chemical potential of oxygen. The Etot,bulk and Etot,Vo denote the total energy of the bulk simulation cell and that containing Vo, respectively.

3. Results
3.1. Bulk SrFeO2

We first reproduce the experimentally determined properties of bulk SrFeO2. As shown in Tables 1 and 2, the calculated lattice constants, band gap, and magnetic moment overall agree with the experiment and previous theoretical work.[5] Then we study the strain effect on the bulk SrFeO2. The SrFeO2 thin film may be grown on various substrates. Except for the lattice-matched KTaO3 substrate,[1,2,19] the practical substrate lattice constants are smaller than 3.99 Å, such as 3.91 Å (SrTiO3),[19,28] 3.79 Å (LaAlO3),[29] 3.94 Å (DyScO3),[19,30] etc. Therefore, the compressive strains are introduced up to −0.06. The tensile strains are also introduced for comparison. To determine the magnetic orderings under strain, we consider the A-AFM, C-AFM, G-AFM, and FM. As shown in Fig. 2, for the G-AFM, both the intralayer nearest neighbor (NN) interactions and the interlayer NN interactions are AFM. For the A-AFM, the intralayer NN interactions are FM, while the interlayer NN interactions are AFM. For the C-AFM, the intralayer NN interactions are AFM while the interlayer NN interactions are FM. From Tables 1 and 2 we can see that the G-AFM has the lowest total energy, indicating that the G-AFM is the magnetic ground state in the strain range we studied. Table 1 also shows the bulk SrFeO2 has the G-AFM at zero strain, which agrees with the experiment.[1] From Fig. 3 we can see both the magnetic moment and the volume decrease with increasing strain. The change of volume indicates that SrFeO2 does not have a perfect Poisson ratio, as a material with an ideal Poisson ratio has a constant volume with strain. This is also reflected in the lattice constants. Due to the high symmetric positions of Fe and O, the Fe–O bonds synchronously expand (reduce) with increasing (decreasing) strain, while the O–Fe–O bond angles keep constant (90°) on the strain range. Consistent with previous work,[4,5] the density of states (DOS) in Fig. 4(b) shows that the bulk SrFeO2 displays a Mott–Hubbard band gap, where the down Fe states locate right below the gap and other down 3d orbitals locate above it. The electronic configurations (see Fig. 4(b-2)) are (dxy)1(dyzdxz)2(dx2y2)1(dz2)1 for the up-spin Fe2+(d6) ion while the sixth down-spin electron occupies the orbital. As shown in Figs. 4(a) and 4(c), when the tensile strain is imposed on SrFeO2, the degenerate dyzdxz states in the conduction band (CB) are pushed into the higher energy levels, thus the insulator character is enhanced with increasing tensile strain.

Table 1.

The calculated properties of bulk SrFeO2 without strain. For comparison, the experimental values and previous theoretical values are also provided. The units of a(b) and c are Å. The units of MFe and Egap are μB and eV, respectively.

.
Table 2.

The calculated total energies (meV/f.u.) of bulk SrFeO2 at different strains. EG, EA, EC, and EF denote the total energies of G-AFM, A-AFM, C-AFM, and FM, respectively.

.
Fig. 2. Schematic view of different magnetic orderings. For clarity the supercell of SrFeO2 is used.
Fig. 3. Evolution of structural and magnetic properties with strain in bulk SrFeO2. (a) The relaxed volume (vol) and magnetic moment. (b) Lattice constants. The dashed line shows the c-axis lattice constant (cst) for a material with an ideal Poisson ratio.
Fig. 4. (a-1), (b-1), (c-1), and (d-1) Total DOS and (a-2), (b-2), (c-2), and (d-2) orbital-resolved DOS of bulk SrFeO2.

On the contrary, the compressive strain makes the dyzdxz states in the CB come nearer with respect to the valence band maximum (VBM), thus the band gap decreases with increasing compressive strain. At the same time, the separation between the dx2y2 states and the down dz2 states becomes larger around the Fermi level. When the compressive strain increases up to −0.06, the band gap nearly comes to zero (not shown), the dx2y2 states also come nearer with respect to the Fermi level. A Mott-type insulator–metal transition (IMT), where the dx2y2 orbital would cross the Fermi level, is expected if we continue to increase the compressive strain, which is similar to the case of IMT at hydrostatic pressure.[6,31] Note that tests are performed at −0.08 and −0.1 strain, and we find the occupations of dx2y2 states at the Fermi level, but the large strain values which may not be very practical are beyond the scope of the present work. Figure 5 shows the evolution of exchange interactions obtained using the classic Heisenberg spin model: H = −J1Si SjJ2Si SjJ3Si Sj, Si(j) = ± 2.

Fig. 5. Exchange interactions with respect to the strain. The data (zero strain) of Ref. [5] are provided for comparison.

We consider the intralayer NN exchange interactions J1, the interlayer NN exchange interactions J2, and the interlayer next nearest neighbor (NNN) exchange interactions J3. The calculated values at zero strain agree well with previous work.[5] The exchange interactions show the anisotropic behavior where J1 is about three times larger than J2, indicating the dominant role of intralayer superexchange interactions. It is interesting to note that the strength (absolute value) of J1 shows a decrease trend at −0.06 strain. This may imply the instability of G-AFM at very large strain as the AFM–FM transition has been found at high pressure.[6,31]

3.2. Defect SrFeO2: Formation energy of Vo

In the tetragonal SrFeO2 structure (see Fig. 1), the oxygen positions are equivalent to each other. Therefore, an arbitrary O atom is removed from the supercell to calculate the Eform of Vo. In extreme oxidation conditions (oxygen rich), the μo is subject to an upper bound given by the energy of O in an O2 molecule, i.e., μo = (1/2)Eo2. In extreme reduction conditions (oxygen deficient), μSr and μFe are subject to an upper bound given by the energy of Sr and Fe in the bulk phases (μSr = ESr,bulk, μFe = EFe,bulk). Thus in this condition, μo = [Etot,bulk(f.u.)μSrμFe]/2. The calculated Vo formation energy is shown in Fig. 6. We can see the Eform is relatively high. Thus, the Vo concentration is expected to be low in equilibrium condition. Although the Eform is positive (e.g., 5.1 eV, zero strain, oxygen rich), it is comparable with the Eform (zero strain, oxygen rich) of some typical TMOs, such as HfO2[32] (5.93 eV), ZnO[33] (5.37 eV), ZrO2[34] (5.97 eV), etc. The Vo is often unintentionally introduced during the growth or annealing process of these TMOs.[35,36] Thus, the Vo of SrFeO2 should be stably formed under certain growth conditions.

Fig. 6. Formation energy of Vo as a function of strain.

Generally speaking, the increase in bond lengths and volumes associated with the tensile strain is likely to decrease the formation energies of Vo due to the decreased Coulomb interactions between ions.[37] It is also believed that the remaining electrons caused by Vo would strengthen the pressure of the electron gas and thus induce an intrinsic compressive stress in the unrelaxed lattice.[38] Therefore, the formation energy is expected to increase in the compressive strain range, and the tensile strain would relieve the intrinsic compressive stress and promote to create the Vo, which has been widely reported in previous similar work.[39,40] As expected, the Vo formation energy decreases with increasing tensile strain, as shown in Fig. 6. However, we can see from Fig. 6 that the compressive strain also decreases the formation energy. Especially in the compressive strain range the formation energy shows a sharper decrease than that of the tensile strain range. To understand this phenomenon, we first examine the structural relaxations around Vo at each strain. As shown in Table 3 and Fig. 7(a), the removal of oxygen would unscreen the positive Fe ions around Vo by the Coulomb interactions, which results in significant outward relaxations for the two NN Fe ions. For the four NNN Fe ions, the distances between them (∼9 Å) are far larger than that of the two NN Fe ions, and the non-screen effect is smaller, thus the four NNN Fe ions show a slight inward relaxation. Although the relaxation patterns are similar for the compressive strain and the tensile strain, we note the outward relaxations are larger in the tensile strain range when the tensile strain is larger than 0.02, which may indicate the dominant role of the rapidly decreased Coulomb interactions in the tensile strain range. Noting that the decreased bond lengths and volumes caused by compressive strain would make the electronic structures more complex due to the enhanced overlap between orbitals, and thus change the electronic energies which may compensate the increased Coulomb interactions. To confirm this assumption, we decompose the Vo formation energy into its components in Fig. 8. We can see that in the compressive strain range the rapidly decreased band energies can overcompensate the increased remaining energies (mainly electrostatic), which induces the decrease of formation energy. On the contrary, in the tensile strain range the rapidly decreased remaining energies overcompensate the increased band energies when the tensile strain is larger than 0.02, indicating the dominant role of the electrostatic interactions in the tensile strain range. Therefore, the competitions between the band energies and the electrostatic interactions are the dominant mechanisms in determining the Vo formation.

Fig. 7. (a) Local structure around Vo. b1 and b2 are the bond lengths (Å). (b) Partial charge density of the defect levels shown in Fig. 9(b-2). The isosurface value is 0.015 e/Bohr3.
Table 3.

Structural relaxations around Vo. Δ denotes the changes of b1(b2) before and after relaxations. The units of b1(b2) are Å.

.
Fig. 8. Decomposition of Eform into its contributions.
Fig. 9. (a-1), (b-1), (c-1), and (d-1) Total DOS and (a-2), (b-2), (c-2), and (d-2) orbital-resolved DOS of defect SrFeO2 with Vo.
3.3. Defect SrFeO2: electronic structures of Vo

As shown in Fig. 9, the Vo formation produces a large shift of the Fermi level. According to the crystal field theory, the removal of oxygen will make the Fe 3d orbitals with the y component fall in energy level. The dx2y2 orbitals fall the most, as the electrons of dx2y2 orbitals are concentrated in lobes along the directions of Fe–O bonds. Thus, the remaining electrons of Vo mainly occupy the empty down-spin dx2y2 orbitals (see Figs. 7(b) and 9(b-2)) and the resulting donor levels locate right below the CBM, indicating the character of n-type conductivity. Comparing with the DOS (see Fig. 4(b-2)) of bulk SrFeO2, the dx2y2 electrons become delocalized, as can be seen from the wide distribution of more peaks in the energy range. Therefore, the mobility of the conduction electrons is expected to be large. With the increase of compressive strain, the intralayer superexchange interactions between dx2y2 states through O pσ increase, which lifts the energy levels of the dx2y2 orbitals. Thus, the occupations of the conduction electrons rapidly decrease when the compressive strain is imposed (Fig. 9(c)). When the compressive strain increases to −0.06, the Fermi level is occupied mainly by the Sr d electrons while the down-spin dx2y2 states are fully pushed into the CB. The tensile strain (Fig. 9(a)) displays the behavior which favors the occupations of dx2y2 orbitals for the conduction electrons. This is because the decreased intralayer superexchange interactions further decrease the energy level of dx2y2 orbitals. Even so, the separation between the donor level and the CBM becomes larger and the donor level approximately shifts to the middle of the enlarged gap at 0.06 strain. Thus, the n-type conductivity becomes poor under tensile strain. Finally, based on the above discussion, we propose a possible explanation for the observations of the Hall experiment. As reported in Fig. 6, the Vo concentration is expected to be low in SrFeO2. In Ref. [19], Katayama et al. found the n-type conductivity on the lattice-matched KTaO3 substrate which corresponds to the case of zero strain. The DyScO3 substrate (3.94 Å) and the SrTiO3 substrate (3.91 Å) would introduce −0.01 and −0.04 compressive strain. As discussed in Fig. 9, the delocalized dx2y2 electrons are rapidly pushed into the CB once the compressive strain is applied. Although the DOS at −0.06 strain displays the metallic character, the Fermi level is mainly occupied by the localized Sr d electrons. Therefore, no n-type conductivity is found in the strained films.

4. Conclusion and perspectives

In summary, we study the epitaxial strain effect on the formation and electronic structures of Vo in SrFeO2 by first-principles calculations. It is found that the bulk SrFeO2 has the G-AFM at zero strain, which agrees with the experiment. Under compressive strain the bulk SrFeO2 keeps the G-AFM and has the trend of Mott insulator–metal transition. We find both the tensile strain and the compressive strain enhance the Vo formation which is different from most of the previous similar work about the strain effect on Vo. The competition between the band energies and the electrostatic interactions is found to be the dominant mechanism in determining the formation of Vo. The Vo in SrFeO2 is confirmed to induce the n-type conductivity where the donor levels are occupied by the delocalized dx2y2 electrons. Our results suggest that the vanishing of n-type conductivity observed by the Hall measurement on the strained films is caused by the shift of donor levels into the conduction band. The calculations would provide insightful information for the realization of metallic conduction in SrFeO2.

Reference
1Tsujimoto YTassel CHayashi NWatanabe TKageyama HYoshimura KTakano MCeretti MRitter CPaulus W 2007 Nature 450 1062
2Inoue SKawai MShimakawa YMizumaki MKawamura NWatanabe TTsujimoto YKageyama HYoshimura K 2008 Appl. Phys. Lett. 92 161911
3Wells A F1975Structural Inorganic ChemistryOxfordOxford University Press
4Xiang H JWei S HWhangbo M H 2008 Phys. Rev. Lett. 100 167207
5Pruneda J MÍñiguez JCanadell EKageyama HTakano M 2008 Phys. Rev. 78 115101
6Kawakami TTsujimoto YKageyama HChen X QFu C LTassel CKitada ASuto SHirama KSekiya YMakino YOkada TYagi THayashi NYoshimura KNasu SPodloucky RTakano M 2009 Nat. Chem. 1 371
7Ju SCai T Y 2009 Appl. Phys. Lett. 94 061902
8Seinberg LYamamoto TTassel CKobayashi YHayashi NKitada ASumida YWatanabe TNishi MOhoyama KYoshimura KTakano MPaulus WKageyama H 2011 Inorg. Chem. 50 3988
9Romero F DBurr S JMcGrady J EGianolio DCibin GHayward M A 2013 J. Am. Chem. Soc. 135 1838
10Horigane KLlobet ALouca D 2014 Phys. Rev. Lett. 112 097001
11Lu H SCai T YJu SGong C D 2015 J. Phys. Chem. 119 17673
12Bouwmeester H J M 2003 Catal. Today 82 141
13Fleig J 2003 Annu. Rev. Mater. Res. 33 361
14Smith M GManthiram AZhou JGoodenough J BMarkert J T 1991 Nature 351 549
15Tamura RKawashima NYamamoto TTassel CKageyama H 2011 Phys. Rev. 84 214408
16Yamamoto TKobayashi YHayashi NTassel CSaito TYamanaka STakano MOhoyama KShimakawa YYoshimura KKageyama H 2012 J. Am. Chem. Soc. 134 11444
17Retuerto MJiménz-Villacorta FMartínez-LopeM JFernández-Díaz M TAlonso J A 2011 Inorg. Chem. 50 10929
18Matsuyama TChikamatsu AHirose YFukumura THasegawa T 2011 Appl. Phys. Express 4 013001
19Katayama TChikamatsu AHirose YKumigashira HFukumura THasegawa T 2014 J. Phys. D: Appl. Phys. 47 135304
20Blöchl P E 1994 Phys. Rev. 50 17953
21Kresse GFurthmüller J 1996 Phys. Rev. 54 11169
22Kresse GJoubert D1999Phys. Rev. B591758
23Perdew J PBurke KErnzerhof M1997Phys. Rev. Lett.781396
24Dudarev S LBotton G ASavrasov S YHumphreys C JSutton A P 1998 Phys. Rev. 57 1505
25Monkhorst H JPack J D 1976 Phys. Rev. 13 5188
26de Walle C G VNeugebauer J 2004 J. Appl. Phys. 95 3851
27Chikamatsu AMatsuyama THirose YKumigashira HOshima MHasegawa T 2012 J. Electron Spectrosc. Relat. Phenom. 184 547
28Béa HBibes MBarthélémy ABouzehouane KJacquet EKhodan AContour J PFusil SWyczisk FForget ALebeugle DColson DViret M 2005 Appl. Phys. Lett. 87 072508
29Martin L WZhan QSuzuki YRamesh RChi M FBrowning NMizoguchi TKreisel J 2007 Appl. Phys. Lett. 90 062903
30Wordenweber R K RHollmann ESchubert J 2007 J. Appl. Phys. 102 044119
31Rahman MNie Y ZGuo G H 2013 Inorg. Chem. 52 12529
32Chen WSun Q QDing S JZhang D WWang L K 2006 Appl. Phys. Lett. 89 152904
33Lu Y BDai YWei WZhu Y THuang B B 2013 Chem. Phys. Chem. 14 3916
34Zhao S JXue J MWang Y GYan S 2012 J. Appl. Phys. 111 043514
35Freysoldt CGrabowski BHickel TNeugebauer JKresse GJanotti Avan de Walle C G 2014 Rev. Mod. Phys. 86 253
36van de Walle C GJanotti A 2011 Phys. Status Solidi 248 19
37Aschauer UPfenninger RSelbach S MGrande TSpaldin N A 2013 Phys. Rev. 88 054111
38Zhu JLiu FStringfellow G BWei S H 2010 Phys. Rev. Lett. 105 195503
39Yang QCao J XMa YZhou Y CJiang L MZhong X L 2013 J. Appl. Phys. 113 184110
40Ma D WLu Z STang Y NLi T XTang Z JYang Z X 2014 Phys. Lett. 378 2570