Tuning transport coefficients of monolayer with biaxial strain
Abstract
Experimentally synthesized (\textcolor[rgb]0.00,0.00,1.00Science 369, 670-674 (2020)) is a piezoelectric semiconductor. Here, we systematically study the large biaxial (isotropic) strain effects (0.90 to 1.10) on electronic structures and transport coefficients of monolayer by density functional theory (DFT). With from 0.90 to 1.10, the energy band gap firstly increases, and then decreases, which is due to transformation of conduction band minimum (CBM). Calculated results show that the monolayer is mechanically stable in considered strain range. It is found that the spin-orbital coupling (SOC) effects on Seebeck coefficient depend on the strain. In unstrained , the SOC has neglected influence on Seebeck coefficient. However, the SOC can produce important influence on Seebeck coefficient, when the strain is applied, for example 0.96 strain. The compressive strain can change relative position and numbers of conduction band extrema (CBE), and then the strength of conduction bands convergence can be enhanced, to the benefit of n-type . Only about 0.96 strain can effectively improve n-type . Our works imply that strain can effectively tune the electronic structures and transport coefficients of monolayer , and can motivate farther experimental exploration.
pacs:
71.20.-b, 72.15.Jf Email:[email protected]I Introduction
The successful exfoliation of grapheneq6 induces increasing attention on two-dimensional (2D) materials. Many of them have semiconducting behaviour, which has various potential application in electronics, optoelectronics and piezoelectronicsm4-1 ; m5-1 ; xzq ; m2-1 . Their electronic structures, heat transport and piezoelectric properties have been widely investigatedm1 ; m2 ; m3 ; m4 ; m5 ; m6 ; m7 ; m8 ; m9 ; m10 ; m11 . It has been proved that the strain can effectively tune electronic structures, transport and piezoelectric properties of 2D materialsm12 ; m10 ; m11 ; m13 ; m14 ; m15 ; m16 ; m17 ; m18 , which shows great potential for better use in the nanoelectronic, thermoelectric and piezoelectric applications. For example, both compressive and tensile strain can induce the semiconductor to metal transition in monolayer m12 . In many transition metal dichalchogenides (TMD) monolayers, the power factor can be enhanced by strain due to bands convergem10 ; m11 ; m13 . With increased tensile strain, the lattice thermal conductivity shows monotonous decrease, up-and-down and jump behavior with similar penta-structuresm14 . Strain can also improve the piezoelectric strain coefficient by tuning the elastic and piezoelectric stress coefficientsm15 ; m16 ; m17 ; m18 .

Recently, the layered 2D and have been experimentally achieved by chemical vapor deposition (CVD)m19 . The septuple-atomic-layer monolayers with twelve different structures are constructed by intercalating -type monolayer into InSe-type monolayerm20 . The 66 thermodynamically and dynamically stable are predicted by the first principle calculations. They can be common semiconductor, half-metal ferromagnetism or spin-gapless semiconductor (SGS), Ising superconductor and topological insulator, which depends on the number of valence electronsm20 . We predict intrinsic piezoelectricity in monolayer m21 , which means that family may have potential application in piezoelectric field. Structure effect on intrinsic piezoelectricity in monolayer (M=Mo and W) has also been reported by the first principle calculationsm21-1 . By applied strain, the monolayer undergoes ferromagnetic metal (FMM) to SGS to ferromagnetic semiconductor (FMS) to SGS to ferromagnetic half-metal (FMHM) with increasing strainm22 . Some materials of lack inversion symmetry with a strong SOC effect, which are expected to exhibit rich spin-valley physicsm20 . The valley-dependent properties of monolayer , and have been predicted by the first-principles calculationsm20 ; m23 ; m24 . Recently, Janus 2D monolayer in the new septuple-atomic-layer 2D family has been achievedm25 , which shows Rashba spin splitting and out-of-plane piezoelectric polarizations.


In nanoscale devices, the residual strain usually exists in real applicationsl111 . In our previous work, the small strain effects (0.96 to 1.04) on piezoelectric coefficients of monolayer have been investigatedm21 . In this work, the large (0.90 to 1.10) biaxial strain-tuned electronic structures and transport coefficients of monolayer are studied by the first principle calculations. With from 0.90 to 1.10, the energy band gap of monolayer firstly increases, and then decreases. In n-type doping, the Seebeck coefficient S can be effectively enhanced by applying compressive strain, and then the can be improved. The tensile strain can induce flat valence bands around the point near the Fermi level, producing large p-type S. Therefore, our works give an experimental proposal to improve transport coefficients of monolayer .
The rest of the paper is organized as follows. In the next section, we shall give our computational details and methods about transport coefficients. In the third and fourth sections, we will present main results of monolayer about strain-tuned electronic structures and transport coefficients. Finally, we shall give our conclusions in the sixth section.
II Computational detail
To avoid interactions between two neighboring images, a vacuum spacing of more than 32 along the z direction is added to construct monolayer . The elastic stiffness tensor are calculated by using strain-stress relationship (SSR), which are performed by using the VASP codepv1 ; pv2 ; pv3 within the framework of DFT1 . A kinetic cutoff energy of 500 eV is adopted, and we use the popular generalized gradient approximation of Perdew, Burke and Ernzerh of (GGA-PBE)pbe as the exchange-correlation potential to calculate elastic and electronic properties. The total energy convergence criterion is set to eV, and the Hellmann-Feynman forces on each atom are less than 0.0001 . The Brillouin zone (BZ) sampling is done using a Monkhorst-Pack mesh of 15151 for elastic constants . The 2D elastic coefficients have been renormalized by the the length of unit cell along z direction (): = .
The electronic transport coefficients of monolayer are calculated through solving Boltzmann transport equations within the constant scattering time approximation (CSTA), which is performed by BoltzTrapb code. To include the SOC, a full-potential linearized augmented-plane-waves method is used to calculate the energy bands of monolayer, as implemented in the WIEN2k package2 . To attain accurate transport coefficients, a 35 35 1 k-point meshes is used in the first BZ for the energy band calculation, make harmonic expansion up to in each of the atomic spheres, and set .
III Electronic structures
The monolayer can be considered as the insertion of the 2H -type monolayer into the -InSe-type , and the side and top views of the structure of the monolayer are plotted in Figure 1. The structure breaks the inversion symmetry, but preserves a horizontal mirror corresponding to the plane of the Mo layer. This leads to that monolayer only has in-plane piezoelectric response, and has not out-of-plane piezoelectric polarizations. Using optimized lattice constantsm21 , the energy bands of monolayer using GGA and GGA+SOC are shown in Figure 2, and exhibit both the indirect band gaps with valence band maximum (VBM) at point and CBM at K point. Due to lacking inversion symmetry and containing the heavy element Mo, there exists a SOC induced spin splitting of about 0.13 eV near the Fermi level in the valence bands at K point. This may provide a platform for spin-valley physicsm20 ; m23 ; m24 , but the VBM is not at K point, which can be tuned by strain. According to orbital projected band structure, it is found that the states near the Fermi level are dominated by the Mo- orbitals. More specifically, the states around both CBM and VBM are dominated by the Mo orbital.

It is proved that the electronic structures, topological properties, transport and piezoelectric properties of 2D materials can be effectively tuned by strainm12 ; m10 ; m11 ; m13 ; m14 ; m15 ; m16 ; m17 ; m18 ; t9 . The biaxial strain can be simulated by or , where and are the strained and unstrained lattice constant, respectively. The 1 or 0 means compressive strain, while 1 or 0 implies tensile strain. With from 0.90 to 1.10, the energy band structures are plotted in Figure 2, and the energy band gap and spin-orbit splitting value at K point are shown in Figure 3.
It is found that the energy band gap firstly increases (0.90 to 0.96), and then decreases (0.96 to 1.10), which is due to transformation of CBM. Similar phenomenon can be observed in many TMD and Janus TMD monolayersm11 ; m11-1 . With strain from compressive one to tensile one, the has a rapid increase, and then a slight decrease. With increasing compressive strain (1.00 to 0.90), the position of CBM (VBM) changes from K () point to one point along the K- direction (K point), when the compressive strain reaches about 0.94 (0.96). The compressive strain can also tune the numbers and relative positions of valence band extrema (VBE) or CBE. For example, at 0.96, the four CBE can be observed, and they energies are very close, which has very important effects on transport properties. To explore orbital contribution to the conduction bands in the case of 0.96 strain, we project the states to atomic orbitals at 0.96 strained and unstrained conditions, which are shown in Figure 4. At 0.96 strain, the composition of the low-energy states has little change with respect to unstrained one. At 0.98, the energy of two VBE are nearly the same. The compressive strain can make K point with spin splitting become VBM, which is very useful to allow spin manipulation for spin-valley physics. For example, at 0.94 strain, the VBM at K point is 0.49 eV higher than that at point. It is clearly seen that the increasing tensile strain can make valence band around the point near the Fermi level more flat.

Finally, the elastic constants are calculated as a function of to study the mechanical stability of monolayer with strain. For for 2D hexagonal crystals, the Born criteria of mechanical stability ela ( and ) should be satisfied. The calculated and as a function of strain are plotted in Figure 5, and it is clearly seen that the monolayer in considered strain range is mechanically stable, which is very important for farther experimental exploration.


IV Electronic transport property
Proposed by Hicks and Dresselhaus in 1993q2 ; q3 , the potential thermoelectric materials can be achieved in the low-dimensional systems or nanostructures. The dimensionless figure of merit, , can be used to measure the efficiency of thermoelectric conversion of a thermoelectric material, where S, , T, and are the Seebeck coefficient, electrical conductivity, working temperature, electronic and lattice thermal conductivities, respectively. It is noted that, for the 2D material, the calculated , and depend on (here, =40 ), and the S and is independent of . For 2D materials, we use electrons or holes per unit cell instead of doping concentration, which is described by N, and the N () 0 mean n- (p-) type doping. It is proved that the SOC has important effects on transport coefficients of TMD and Janus TMD monolayersm11 ; m13 ; m11-1 . However, the SOC has neglectful influences on transport properties of unstrained monolayer, which can be observed from typical Seebeck coefficient S in Figure 6. This is because the energy bands near the Fermi level between GGA and GGA+SOC is nearly the same. However, the SOC has an important effect on p-type transport coefficients with the condition of compressive strain. For example at 0.96 strain, a detrimental effect on Seebeck coefficient S can be observed, when including SOC (See Figure 6). This is because the SOC can remove the band degeneracy near the VBM. So, the SOC is included to investigate the biaxial strain effects on transport coefficients of monolayer.
Using GGA+SOC, the room temperature S, and of monolayer under different strain (0.90 to 1.10) are shown in Figure 7. It is clearly seen that the compressive strain has important effects on S, especially for n-type doping. However, the tensile strain produces small influences on S, especially for n-type S. These can be explained by strain-induced energy bands. When the strain is less than or equal to about 0.98, the n-type S (absolute value) can be observably improved, which is due to compressive strain-driven accidental conduction band degeneracies, namely bands convergence. With expanding compressive strain, in the low doping, the p-type S firstly increases, and has almost no change. This is because the valence bands convergence can be observed at about 0.98, and then is removed (At 0.98, the energy of two VBE are nearly the same, and only one VBE near the Fermi level can be observed with compressive strain from 0.96 to 0.90.). Foe considered tensile strain, the conduction bands near the Fermi level have little change, which leads to almost unchanged n-type S. When the strain changes from 1.00 to 1.10, the p-type S increases, which is due to tensile strain-induced more flat valence bands around point near the Fermi level. This can be understood by , in which , T and is the effective mass of the carrier, temperature and carrier concentration, respectively. The flat bands can produce very large effective mass of the carrier, which will lead to improved S. It is found that the strain has nearly the opposite effects on with respect to S. It is found that the compressive strain can dramatically improve due to the strain-enhanced S.
An upper limit of can be measured by , neglecting the . The room temperature of monolayer under different strain as a function of doping level are also shown in Figure 7. Calculated results show that the dependence of is very similar to one of S (absolute value), which can be explained by the Wiedemann-Franz law: ( is the Lorenz number). And then the can be reformulated by . Thus, the strain-induced bands convergence improves S, which is beneficial to better .
V Conclusion
In summary, we investigate the biaxial strain (0.90 to 1.10) effects on electronic structures and transport coefficients of monolayer by the reliable first-principles calculations. With the strain from 0.90 to 1.10, the energy band gap of monolayer shows a nonmonotonic behavior. It is found that the SOC has little effects on transport coefficients of unstrained in considered doping range due to the hardly changed dispersion of bands near the Fermi level. However, the SOC has very important influences on transport properties of strained , for example 0.96 strain, which is due to the position change of VBM. Calculated results show that compressive strain can tune the numbers and relative positions of CBE, which can lead to enhanced n-type S, and then better n-type . Our works may provide an idea to optimize the electronic structures and transport properties of monolayer .
VI Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Acknowledgements.
This work is supported by the Natural Science Foundation of Shaanxi Provincial Department of Education (19JK0809). We are grateful to the Advanced Analysis and Computation Center of China University of Mining and Technology (CUMT) for the award of CPU hours and WIEN2k/VASP software to accomplish this work.References
- (1) Novoselov K S et al. 2004 Science 306 666
- (2) Mak K F and Shan J 2016 Nature Photonics 10 216
- (3) Blonsky M N, Zhuang H L, Singh A K and Hennig R G 2015 ACS Nano 9 9885
- (4) Hui Z Q, Xu W X, Li X H et al. 2019 Nanoscale 11 6045
- (5) Wu W and Wang Z L 2016 Nat. Rev. Mater. 1 16031
- (6) Chhowalla M, Shin H S, Eda G, Li L J, Loh K P and Zhang H 2013 Nature Chemistry 5 263
- (7) Lu A Y, Zhu H Y, Xiao J et al. 2017 Nature Nanotechnology 12 744
- (8) Fei R X, Li W B, Li J and Yang L 2015 Appl. Phys. Lett. 107 173104
- (9) Zhang S L et al. 2016 Angew. Chem. 128 1698
- (10) Blonsky M N, Zhuang H L, Singh A K and Hennig R G 2015 ACS Nano 9 9885
- (11) Fei R X, Li W B, Li J and Yang L 2015 Appl. Phys. Lett. 107 173104
- (12) Duerloo K N, Ong M T and Reed E J 2012 J. Phys. Chem. Lett. 3 2871
- (13) Chen Y, Liu J Y, Yu J B, Guo Y G and Sun Q 2019 Phys. Chem. Chem. Phys. 21 1207
- (14) Ji J P et al. 2016 Nat. Commun. 7 13352
- (15) Lv H Y, Lu W J, Shao D F, Lub H Y and Sun Y P 2016 J. Mater. Chem. C 4 4538
- (16) Guo S D 2016 J. Mater. Chem. C 4 9366
- (17) Scalise E, Houssa M, Pourtois G, Afanas’ev V and Stesmans A 2012 Nano Res. 5 43
- (18) Guo S D 2016 Comp. Mater. Sci. 123 8
- (19) Liu H K, Qin G Z, Lin Y and Hu M 2016 Nano Lett. 16 3831
- (20) Jena N, Dimple, Behere S D and Sarkar A D 2017 J. Phys. Chem. C 121 9181
- (21) Guo S D, Guo X S, Zhang Y Y and Luo K 2020 J. Alloy. Compd. 822 153577
- (22) Dimple, Jena N, Rawat A, Ahammed R, Mohanta M K and Sarkar A D 2018 J. Mater. Chem. A 6 24885
- (23) Guo S D, Mu W Q and Zhu Y T 2021 J. Phys. Chem. Solids 151 109896
- (24) Hong Y L, Liu Z B, Wang L et al. 2020 Science 369 670
- (25) Wang L, Shi Y P, Liu M F et al. 2020 arXiv:2008.02981
- (26) Guo S D, Zhu Y T, Mu W Q and Ren W C 2020 EPL 132 57002
- (27) Guo S D, Zhu Y T, Mu W Q, Wang L and Chen X Q 2021 Comp. Mater. Sci. 188 110223
- (28) Guo S D, Mu W Q, Zhu Y T and Chen X Q 2020 Phys. Chem. Chem. Phys. 22 28359
- (29) Li S, Wu W K, Feng X L et al. 2020 arXiv:2009.13253
- (30) Yang C, Song Z G, Sun X T and Lu J 2020 arXiv:2010.10764
- (31) Guo S D, Mu W Q, Zhu Y T, Han R Y and Ren W C 2021 J. Mater. Chem. C DOI: 10.1039/D0TC05649A (arXiv:2011.13523)
- (32) Bhowmick S and Shenoy V B 2006 J. Chem. Phys. 125 164513
- (33) Kresse G 1995 J. Non-Cryst. Solids 193 222
- (34) Kresse G and Furthmller J 1996 Comput. Mater. Sci. 6 15
- (35) Kresse G and Joubert D 1999 Phys. Rev. B 59 1758
- (36) Hohenberg P and Kohn W 1964 Phys. Rev. 136 B864; Kohn W and Sham L J 1965 Phys. Rev. 140 A1133
- (37) Perdew J P, Burke K and Ernzerhof M 1996 Phys. Rev. Lett. 77 3865
- (38) Madsen G K H and Singh D J 2006 Comput. Phys. Commun. 175 67
- (39) Blaha P, Schwarz K, Madsen G K H, Kvasnicka D and Luitz J 2001 WIEN2k, an Augmented Plane Wave + Local Orbitals Program for Calculating Crystal Properties (Karlheinz Schwarz Technische Universityt Wien, Austria)
- (40) Zhang S L, Xie M Q, Cai B et al. 2016 Phys. Rev. B 93 245303
- (41) Hicks L D and Dresselhaus M S 1993 Phys. Rev. B 47 12727
- (42) Hicks L D and Dresselhaus M S 1993 Phys. Rev. B 47 16631(R)
- (43) Guo S D and Dong J 2018 Semicond. Sci. Tech. 33 085003
- (44) Andrew R C, Mapasha R E, Ukpong A M and Chetty N 2012 Phys. Rev. B 85 125428