Get 20M+ Full-Text Papers For Less Than $1.50/day. Start a 14-Day Trial for You or Your Team.

Learn More →

Strain-induced Weyl and Dirac states and direct-indirect gap transitions in group-V materials

Strain-induced Weyl and Dirac states and direct-indirect gap transitions in group-V materials We perform comprehensive density-functional theory calculations on strained two-dimensional phosphorus (P), arsenic (As) and antimony (Sb) in the monolayer, bilayer, and bulk -phase, from which we compute the key mechanical and electronic properties of these materials. Speci cally, we compute their electronic band structures, band gaps, and charge-carrier e ective masses, and identify the qualitative electronic and structural transitions that may occur. Moreover, we compute the elastic properties such as the Young's modulus Y ; shear modulus G; bulk modulus B; and Poisson ratio  and present their isotropic averages of as well as their dependence on the in-plane orientation, for which the relevant expressions are derived. We predict strain-induced Dirac states in the monolayers of As and Sb and the bilayers of P, As, and Sb, as well as the possible existence of Weyl states in the bulk phases of P and As. These phases are predicted to support charge velocities 6 1 up to 10 ms and, in some highly anisotropic cases, permit one-dimensional ballistic conductivity in the puckered direction. We also predict numerous band gap transitions for moderate in-plane stresses. Our results contribute to the mounting evidence for the utility of these materials, made possible by their broad range in tuneable properties, and facilitate the directed exploration of their potential application in next-generation electronics. PACS numbers: 62.25.-g, 73.61.-r, 81.07.-b omuinneg@tcd.ie arXiv:1801.08233v1 [cond-mat.mtrl-sci] 24 Jan 2018 I. BACKGROUND Two-dimensional black phosphorus (BP), or phosphorene, is one of several predicted sta- ble allotropes of few-layer phosphorus [1{4], and it has attracted considerable attention since its recent successful synthesis [5{9] that is now possible with liquid phase exfoliation [10, 11]. The excitement behind BP is driven by its growing list of technologically relevant anisotropic mechanical and electronic properties. The theoretically predicted properties include a tune- able band-gap [8, 12{17], a negative Poisson's ratio [18], anisotropic conduction [19, 20], and linear dichroism [21, 22]. The properties that have been experimentally veri ed so far in- clude a high hole-mobility between 300 1000 cm /Vs [7, 8, 21, 23], considerable mechanical exibility [24], and a layer-dependent band gap [22, 25, 26] ranging from 0.3 eV in bulk to 2.0 eV in the monolayer. The anisotropic crystal structure of BP is responsible for its unusual electro-mechanical properties, which are predicted to be strongly directional-dependent and highly responsive to mechanically strain [11, 19]. With this renewed interest in BP, focus has quickly turned to few-layer phases of the other pnictogens, namely arsenic [28] (As), antimony [29] (Sb), bismuth [30, 31] (Bi), and their alloys [32{36], which are attracting steadily increasing attention. Moreover, the search for a bulk counterpart to graphene, capable of supporting ballistic electron transport, has recently driven the search for new Weyl and Dirac semi-metals [37{47]. In spite of the diculty in attaining Dirac points in two-dimensional materials [46, 47], the rst experimental Dirac semi-metal in few-layer BP was observed by Kim et al. [48], while As [49], Sb [50{52], Bi [51], and P [51{54] are also predicted to be potential candidates. Indeed, many of the predicted strain-induced properties of these materials, such as direct-indirect band gap transitions [13, 28, 29, 55, 56], a negative Poisson's ratio [18, 57], as well as electronic [2, 55, 58], structural [59], and topological [51, 60{62] transitions, are already spurring their incorporation in emergent technologies such as eld-e ect transistors [8, 20], gas-sensors [63, 64], optical switches [65, 66], solar-cells [34], next-generation batteries [67{69], reinforcing llers [11, 70], and topological insulators [50, 51, 60{62]. In this work, our aim is to provide a comprehensive analysis of the monolayer, bilayer and bulk phases of orthorhombic P, As and Sb, in order to identify and compare the quali- tative strain-related properties of each structure from a consistent set of calculations, thus treating each material on the same footing. Our ndings provide new insights into their 2 (b) (a) (c) FIG. 1: (Color online) a) Top and side view of the 2D orthorhombic puckered structure (generated in VESTA [27]) with the primary vectors along the zigzag (~x) and puckered (~y) directions shown. The unit cell, given by the shaded region, is described by the lattice parameters a and b with the in-plane angle  also de ned. b) 3D Brillouin zone with high-symmetry points , X , S, Y , and Z . c) Side-view of the buckled Sb state at " = 4% compressive strain. yy electro-mechanical properties, especially regarding arsenic and antimony, which have been relatively unexplored to date. Speci cally, we identify qualitative transitions in band gaps, e ective masses, structure, and topology that occur at various strains, and compute the elastic properties that determine the required stresses to attain these electronic states. We predict the existence of strain-induced Dirac states in monolayer As and Sb, bilayer P, As and Sb, as well as possible Weyl states in bulk P and As, at moderate stress values. Our ndings show that all of the predicted Dirac and Weyl points are indeed linear, at 3 least in the Y direction, i.e. the puckered direction. Thus, following the convention of terminology found in the Refs. [38, 48, 71], and other sources, we classify Dirac or Weyl states as those associated with regions of sustained linear dispersion in the band structure, at or near the Fermi level, in at least one direction. We predict these states to support ballistic conduction and are una ected by the spin-orbit coupling (SOC). In particular, few- layer P and As exhibit a strong indication of anisotropic conduction, dominated by ballistic conductivity along Y . The outline of this article is as follows: In Section II we review the details of our calcula- tions and discuss the Voigt-Reuss-Hill averaging scheme used to compute the experimentally- relevant elastic properties. In Section III we present the results of our calculations including the lattice constants, strain dependence of electronic properties, in particular the potential Dirac and Weyl states identi ed, and computed elastic properties. We conclude in Section IV with a discussion of the implications of our key results. II. METHODOLOGY A. Calculation details The calculations were performed with the QuantumEspresso package [72] using the Perdew-Burke-Ernzerhof (PBE) form of the generalized-gradient approximation (GGA) exchange-correlation functional [73]. An ultrasoft pseudopotential [74] from the SSSP Li- brary [75] (with 5 valence electrons) was used to represent the core electrons. Non-SOC calculations were initially performed and those that exhibited potential Dirac or Weyl states were reassessed including non-perturbative SOC. In all calculations, van der Waals (vdW) in- teractions were incorporated using the B97-D empirical dispersion correction functional [76]. In order to achieve an energy convergence of at least 1 meV/atom and force convergence of at least 1:3  10 eV/a , we found it sucient to use a common plane-wave energy cuto of 1100 eV with `cold' smearing [77] of 10 K for all elements. To achieve the same convergence, the Brillouin zone sampling for bulk systems was 15 15 15, and 15 15 1 for monolayers and bilayers. Uniaxial and shear strains between 5% were applied in incre- ments of 1% to the unit cell with internal relaxation subject to the same force convergence criterion as above. Electronic band structures were calculated along the high-symmetry 4 points of the Brillouin zone f, X , S, Y ,Zg (Fig. 1b) for each value of in-plane strain (Figs. S1 - S9 in the Supplemental Material). For shear strains the Brillouin zone deforms into an asymmetric honeycomb, yet we continued to sample along the original path since the deformation up to 5% strain is negligible and the e ective masses are all calculated at the - point. We determine the Kohn-Sham band gap from the band structures and charge-carrier ? 2 2 e ective masses according to the nearly-free electron model m = ~ (@ E=@k @k ) using i j ij a cubic spline t to 9 data points about the -point. The charge velocities were similarly determined according to the dispersion relation v = ~ dE(k)=dk from the linear t to the Dirac or Weyl lines. The elements of the sti ness matrix C were derived from the gradients of the resultant stress-strain pro les c = @ =@" , from which all elastic properties were ij i j derived. In practice, however, the computed sti ness tensors are not exactly symmetric due to numerical noise but we make them so by taking the average of C and its transpose C as the e ective sti ness tensor. We begin our discussion with a brief overview of the the Voigt-Reuss-Hill scheme, which is a popular model used for computing e ective isotropic elastic properties. B. The Voigt-Reuss-Hill scheme In order to e ectively preserve, study and strain-engineer few-layer nano-structures, such as BP [78], graphene [79], or molybdenum disul de [80] (MoS ), the nano- akes are typically deposited onto a suitable substrate. The cumulative contribution of dispersed nano- akes distributed on or within a bulk medium results in the macroscopic elastic properties that are measured by experiments. The theoretical calculation of these elastic properties requires an appropriate mixture model (such as the rule-of-mixtures [11] (ROM) or the Halpin-Tsai [81] models) that require the (typically averaged) elastic properties of the interstitial nano- akes. In the theory of e ective media, isotropic bulk properties are computed by averaging the sti ness tensor C over all possible rotated reference frames [82{84], as outlined in the Supplemental Material. The result is called the Voigt average [85, 86] and it gives isotropic averages for the bulk Young's modulus Y , and shear modulus G , given in Eq. S7. The same V V scheme applied to the compliance tensor S results in the corresponding Reuss averages [87], Y and G , given in Eq. S9. The Voigt scheme assumes that the material undergoes constant R R strain and it returns over-estimated elastic constants. Conversely, the Reuss scheme assumes 5 constant stress and it tends to under-estimate the elastic constants. The Hill averages [88] Y + Y G + G V R V R Y = ; and G = ; (1) H H 2 2 from which the isotropic Poisson's ratio  and bulk modulus B are expressed as H H Y Y G H H H = 1; and B = : (2) H H 2G 3(3G Y ) H H H These are widely considered as reliable estimates of the actual physical values [84]. The Voigt-Reuss-Hill approach described above is used in the present work to determine the isotropic averages of the Young's, shear and bulk moduli, and the Poisson ratio of the bulk P, As, and Sb structures using the elements of the elastic tensors. Let us now discuss how the above approach may be adapted to derive the relevant equations for the speci c case of two-dimensional materials. C. In-plane Voigt-Reuss-Hill average If the interstitial nano- akes in a bulk medium form high-quality planar sediments [11, 89{ 95], the random orientation occurs instead in the plane of the akes and we must calculate isotropic-averages in-plane. Due to the weak vdW bonds between layers, strains related to out-of-plane directions can be ignored resulting in the reduced-sti ness tensor (Eq. S3). In the Supplemental Material, we re-derive the angular dependence of the rotated tensor- elements C () and S () about the ~z-axis (Eq. S6) as a function of the elements in the ij ij original reference frame, similar to the general Voigt-Reuss scheme. The angular-dependence of the in-plane elastic constants are then expressed as 2 2 C C C 11 12 Y () = ; G () = C ;  () = ; V V 66 V C C 11 11 (3) 1 1 S Y () = ; G () = ;  () = ; R R R S S S 11 66 11 with the Hill-average taken as in Eq. 1. By integrating the elastic tensors C () and S () ij ij over 2, the in-plane averages are then computed analogously. 6 III. RESULTS A. Lattice Constants The lattice constants a, b, c of the fully-relaxed structures are presented in Table I, where, in the monolayer and bilayer cases, we quote the layer thickness c instead of the unit cell height c. Our computed lattice parameters compare well with other recent theoretically predicted values [28, 29, 51, 59, 96{99] and the available experimental data [100, 101]. For a a (A) b (A) c (A) P 4.57 3.31 2.11 mono P 4.51 3.31 7.34 bi a a a P 4.43 (4.37 ) 3.32 (3.31 ) 10.47 (10.47 ) bulk As 4.70 3.67 2.39 mono As 4.64 3.69 7.86 bi b b b As 4.56 (4.47 ) 3.71 (3.65 ) 10.94 (11.0 ) bulk Sb 5.02 4.23 2.79 mono Sb 4.88 4.26 8.83 bi c c Sb 4.73 4.29 (4.3 ) 12.09 (11.2 ) bulk Ref. [100] Ref. [101] Ref. [102] TABLE I: Lattice parameters (A) for monolayer, bilayer and bulk structures of P, As, Sb compared to experimental data [100{102] quoted in parentheses. For the monolayers and bilayers the layers thickness c is given. given element, we nd that the lattice parameter along the puckered direction, `a', shortens as the number of layers increases, which agrees with observations in other studies. This is attributed to the increased vdW forces between layers, which leads to increased binding primarily in the softer puckered direction. 7 B. Electronic properties All of the band structures pertaining to the following analysis are presented in Figs. S1 - S9 of the Supplemental Material. Where we identify possible Dirac or Weyl states, high resolution, three-dimensional band structures with SOC at representative strains are recal- culated. To con rm the existence of linear-dispersion, we also plot lines along the surface of the Dirac and Weyl points at 0 , 30 , 60 and 90 with respect to the X line. A representative sample of these results are presented in Figs. 5a - 5g, while the rest can be found in Figs. S10 - S13 of the Supplemental Material. In general we nd the band gap to be very responsive to uniaxial in-plane strain but signi cantly less so with respect to shear strain. We identify several direct-indirect band gap transitions as well as the opening and closing of band gaps, summarized in Table III. We nd the charge-carrier e ective masses vary approximately linearly with respect to the uniaxial strain in general but with notable exceptions that will be discussed. This section is divided into three parts discussing each of the species - P, As and Sb - for which we review the qualitative calculation results including band gap transitions, e ective mass behavior and linearly-dispersive bands. 1. Phosphorus As shown in Fig. 2a, our calculations reproduce the direct band gap of 0.88 eV at the - point in the relaxed P monolayer, which falls within range of the reported gap between 0.7 eV (DFT-PBEsol [12]) and 1.0 eV (DFT-HSE06 [8]). On the other hand, quasi-particle calcu- lations predict a larger 2 eV band gap [22] with signi cant exciton binding [103] (between 0.4-0.83 eV). However, it is well understood that approximate semi-local functionals such as PBE su er from a systematic band gap problem [104] that may also adversely a ect the metal-insulator critical strains. Nevertheless, it is important to emphasize that band align- ments and rates of change are quite often reliably reproduced [105], as are the direct-indirect transitions [13] in two-dimensional materials. While absolute band gaps are therefore not expected to be exactly reproduced, we can expect reasonable agreement with trends in elec- tronic and mechanical behavior [106, 107]. The application of uniaxial in-plane strain is found to open the band gap for tensile strain and diminish it for compressive one, while 8 shear strain has a negligible e ect. The electron and hole e ective masses (Figs. 2b & 2c), compare well to the gures computed in Ref. [108], where, at " = +5% tensile strain, xx the electron and hole e ective masses coincide at 0.9 m as higher energy bands fall below the conduction band. For compressive strains, the hole e ective mass along X rises signi cantly as the valence band attens. Bilayer P is also found to have a direct band gap of 0.43 eV (Fig. 2d) in the relaxed state, and broadly the same behavior as the monolayer, in which case the band gap closes at around 5% uniaxial compressive strain with a predicted Dirac state at the -point. The e ect of SOC on the band structure (Fig. 5a) induces no qualitative di erence and the three-dimensional bands plotted about the Dirac points (Fig. 5b) con rm the linear- dispersion, albeit in only one direction. A linear t to the surface of the bands (Fig. 5c) 6 1 returns a maximum charge velocity of v = 3:80(1)  10 ms along Y , while, in the orthogonal direction, the bands are at with a a charge velocity that is relatively negligible. This high anisotropy in charge velocities, dominated by ballistic conduction along Y , is indicative of e ective one-dimensional conductivity and is further supported by the large disparity in e ective masses at " = 5%, evident in Figs. 2e & 2f. The same analysis for xx " = 5%, for which the Dirac states are due to band inversion and consequently occur o yy the Y symmetry line at a point X , can be found in Figs. 5a - 5c. Here the maximum 6 1 charge velocity is v = 3:22(1) 10 ms . These results are further supported by the work of Doh et al. [54], who demonstrated the e ect of strain on hopping parameters can lead to a Dirac semi-metallic state in bilayer P. Similarly, Baik et al. [71] found that the SOC did not induce a band gap in potassium-doped multi-layer P, but did lift the spin-degeneracy of the Dirac points. A direct-indirect band gap transition is also observed at +2% uniaxial tensile strain. The e ective masses (Figs. 2e & 2f), also exhibit broadly the same behavior as the monolayer, due to band- attening at X and the falling conduction bands along Y , which lead to the charge carrier e ective masses along X converging at +4% strain and an increasing hole e ective mass for compressive strains. In the bulk, however, we nd that the band gap is completely closed (Fig. 2g), i.e. that the material is metallic. After investigation, we concluded that this was an e ect of the smearing functionality [77] in the relaxation procedure and that it contradicts numerous experiments [109{112] that have measured a direct gap in the range of 0.31-0.36 eV. When relaxed under xed-occupancy conditions, instead, a band gap of 0.35 eV was produced. 9 While the PBE gap remains closed in the relaxed state, under uniaxial tensile strain it brie y becomes a single-point semi-metal at +2%. At such strains a possible Weyl state is observed, before a direct gap opens that subsequently transitions to an indirect gap at +3%. Shown in Fig. S10 in the Supplemental Material is the three-dimensional band structure with SOC in which a pair of potential Weyl points occur on an o -symmetry point X along X . Here, the SOC slightly reduces the band gap by  0:05 eV and does not qualitatively a ect the 6 1 overall results. The maximum charge velocity is v = 2:40(1) 10 ms for both " = 5% xx and " = 5% and occurs along a line parallel to Y . Under greater compression yy this band-inversion may also lead to further Weyl states, which have been experimentally observed at similar pressures [60{62]. Again, shear strain is seen to have a negligible e ect on the gap. The electron e ective masses are quite responsive to strain (Fig. 2h), where those along Y rise for both tensile strain along " , due to falling conduction bands, xx and compressive strain along " due to attening bands along Y . The e ective masses yy along X were necessarily not computed once the bands overlapped below +2% strain and the hole e ective masses are found to vary with respect to the strain to a slightly lesser extent (Fig. 2i). To summarize, we predict the onset of -point Dirac states in bilayer P at -5% uniaxial compressive strain, with e ective one-dimensional conductivity at " = 5%, and a direct- xx indirect band gap transition at +2% tensile strain. We also predict the existence of a possible Weyl states at +2% tensile strain in bulk P, followed by a direct-indirect band gap transition at +3%. Finally, e ective masses are found to be particularly responsive to " xx uniaxial strain. 2. Arsenic In contrast to P, we identify an indirect band gap of 0.15 eV along the Y direction in the relaxed As monolayer (Fig. 3a), which is signi cantly lower than the predicted DFT- HSE06 gap [98] of 0.83 eV. However, the relaxed band structure and band gap pro les closely resemble those in Refs. [28, 49]. The band gap diminishes for tensile strain along " xx and at +2% the material becomes semi-metallic with a Dirac state at the -point emerging at " = +5% accompanied by an electron pocket above the Fermi-level (Fig. S11 in the xx Supplemental Material), which is una ected by the SOC. The maximum charge velocity here 10 1.6 ■ 15 ■■ Γ-X Γ-X Γ-X ■ E ■ E ■ E ■■ xx yy xy ■ ■ ■ ■ ■ ■■■ ■■■ ■ ■ ■■ ■ ■■ ■■ m m m ■■ ■■ ■ ■ ■ ■■ xx yy xy 1.2 1.4 * * * E E E ▲ ▲ ▲ Γ-Y Γ-Y Γ-Y xx yy xy ▲ Γ-X Γ-X Γ-X 10 ▲ m ▲ m ▲ m 1.2 m m m xx yy xy ■ ■ ■ xx yy xy ■▲ ■▲ 0.8 ■▲ ■▲ ■▲ ■▲ ■▲ 1.0 ■▲ Γ-Y Γ-Y Γ-Y ■■▲▲ m m m ▲ ▲ ▲ ■ ■▲ ▲■ ▲■■■▲▲ ▲■ ▲■ xx yy xy ■▲ ■▲ ■▲ ■▲ ■▲ ■▲ ▲ 5 ■■▲▲ ■ 0.4 0.8 ■ ■ ■ ■ ■▲▲ ■ ■ ■ ■ ■ ■ ■ ■■■ ■■ ■ ■ ■ ■ ■ ■ ■■▲ ■ ■ ■ ▲ ■ ■ ▲ ■ ■ ■■▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ 0.6 ■▲ ▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ ▲ -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (a) Monolayer P band gap (eV) (b) Monolayer P m =m (c) Monolayer P m =m e 0 h 0 1.2 3.0 3.0 Γ-X Γ-X Γ-X Γ-X Γ-X Γ-X ■ E ■ E ■ E xx yy xy m m m m m m ■ ■ ■ ■ ■ ■ xx yy xy xx yy xy 1.0 2.5 2.5 * * * E E E Γ-Y Γ-Y Γ-Y ■ Γ-Y Γ-Y Γ-Y ▲ ▲ ▲ xx yy xy ▲ m ▲ m ▲ m ▲ m ▲ m ▲ m 0.8 2.0 ■ 2.0 xx yy xy xx yy xy ■ ■ ■ ▲ ■■ ■ ▲ ■▲▲ ■ ■ 0.6 ■ ▲ 1.5 ■ ■ ■ ■ ■ ■■ ■ ■ ■ ■ 1.5 ■ ■ ▲ ■■ ■■■ ■ ▲ ■■ ■ ■▲ ■■ ■■ ■ ■ ■■▲▲ ■■ ■ ■ ■ ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ ■ ■ ■▲ ■▲ ■▲ ■▲ ■▲ ■■■▲▲▲ ■▲ ■▲ ■▲ ■▲ ■▲ ■ ■ ■ ■ 0.4 1.0 ▲ ▲ ▲ 1.0 ■ ■■▲▲ ■ ■■▲▲ ■■▲▲ 0.2 0.5 ▲ 0.5 ■▲ ■▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲ ▲ ▲ ▲▲ ▲▲ ■▲ ▲ ▲ ▲ ▲ ▲▲ ▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲ ▲▲ ▲▲ ▲ ▲ ▲ ▲ ■▲ 0.0 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (d) Bilayer P band gap (eV) (e) Bilayer P m =m (f ) Bilayer P m =m e 0 h 0 3.0 1.2 0.3 ■ Γ-X Γ-X Γ-X Γ-X Γ-X Γ-X ■ m ■ m ■ m m m m xx yy xy ■ ■ ■ xx yy xy 2.5 1.0 * ▲ E E ■ xx ▲ ■ xx Γ-Y Γ-Y Γ-Y Γ-Y 0.2 2.0 m m 0.8 ■ ■ ▲ ▲ m m ▲ xx yy ▲ ▲ ■ xx yy ■ ■ ■ ■ E ■ yy ▲ E ■ yy 1.5 ■ 0.6 ■ ■■ ■■ * ■▲ 0.1 ■ xy ▲ E 1.0 0.4 ▲ xy ■ ■ ▲ ■ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ■ ■ ■ ■▲ ▲ ▲ ▲ ▲ ▲ ■▲▲ ■ ■▲ ■ ■▲ ■ ■ 0.5 ▲ ■ 0.2 ▲ ■ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ■▲ ■▲▲ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ▲ ■▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■▲ ■▲ ■▲ ■▲ 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (g) Bulk P band gap (eV) (h) Bulk P m =m (i) Bulk P m =m e 0 h 0 FIG. 2: (Color online) The relationships between the applied in-plane strains " (blue), xx " (red) and " (green) against [(a), (d), (g)] the direct E (solid squares) and indirect E yy xy (dashed triangles) band gaps (eV) ; [(b), (e), (h)] the e ective electron masses m =m along e 0 X (solid squares) and Y (dashed triangles) ; [(c), (f ), (i)] the e ective hole masses m =m along X (solid squares) and Y (dashed triangles); for each phase of P. h 0 6 1 is v = 3:01(1) 10 ms and lies along Y . In the orthogonal direction the bands are at, similarly to monolayer P, with a relatively small charge velocity. This high anisotropy in charge velocities, dominated by the ballistic conduction along Y , is again indicative of e ective one-dimensional conductivity and is further supported by the large disparity in e ective masses at " = 5%, shown in Figs. 3b & 3c. xx For compressive strain along " the indirect band gap closes along Y at " = 2%. xx xx Bandgap (eV) Bandgap (eV) Bandgap (eV) m /m m /m m /m e 0 e 0 e 0 m /m m /m m /m h 0 h 0 h 0 For tensile strain along " the indirect band gap opens where an indirect-direct transi- yy tion [28] occurs at " = 3%. Similar to monolayer phosphorus, there is no appreciable xx e ect due to shear-strain. Meanwhile, the charge-carrier e ective masses (Figs. 3b & 3c) re- spond linearly to uniaxial strain and compare well to other works [101], where, in particular, valence band broadening along X leads to an increasing hole e ective mass. For bilayer As, we identify a direct band gap of 0:45 eV (Fig. 3d), in contrast to the indirect band gap observed in the monolayer. Here, the band gap opens for uniaxial tensile strain and diminishes for compressive strain. The direct band gap transitions to an indirect gap at both " = 3% and " = +2%, while at " = +2% it also transitions to an indirect yy yy xx gap before resuming to a direct gap again at " = +3%. Moreover, we predict a Dirac state xx at the -point at a compressive strain of " = 4% (Fig. S11 in the Supplemental Material) xx 6 1 for which the maximum charge velocity is v = 2:62(2)10 ms along Y . Along X the bands are also at, similar to the monolayer, and have a relatively negligible charge velocity. The high anisotropy in charge velocities, is again indicative of e ective one-dimensional conductivity, dominated by the ballistic conduction along Y , and is further supported by the large disparity in e ective masses at " = 5%, shown in Figs. 3e & 3f. Here again, xx the SOC has no appreciable e ect on the bands. The electron and hole e ective masses (Figs. 3e & 3f) respond approximately linearly to the applied strain, where conduction band broadening leads to increased e ective electron masses, and valence band attening at leads to increasing hole e ective masses for strain along " . yy Finally, no band gap is determined in the relaxed bulk phase (Fig. 3g), again contrary to experiments [113], where a small direct band gap of  0:3 eV is observed. However, at " = +1% strain, a potential Weyl state is brie y observed on an o -symmetry point xx X (Fig. S12 in the Supplemental Material) before a direct gap opens that subsequently transitions to an indirect one at " = +3%, after which it reduces again. Another potential xx Weyl state around the same o -symmetry point X is also predicted to occur between +1% "  +2% after which a direct band gap also appears. The recalculated band structure with yy the SOC for " = +1% (Fig. S12) con rms the linear-dispersion. The maximum charge yy 6 1 velocity in both cases occurs along a line parallel to Y and is v = 1:38(1) 10 ms . Meanwhile, the electron and hole e ective masses along Y (Figs. 3h & 3i) increase rapidly for compressive strains as the band peaks rapidly atten at the -point. In summary, we predict -point Dirac states in the monolayer and bilayer of As, which 12 1.5 1.5 1.0 ■ E E E ■ xx ■ yy ■ xy ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■■ ■■■ ■■ ■■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ 0.8 ■ ■ ■ ■■ ■■■ ■ ■ ■ ■ ■ ■ ■ * * * ■ ■ E E E Γ-X Γ-X Γ-X ■ ▲ ▲ ▲ 1.0 1.0 xx yy xy ■ ■ m ■ m ■ m ■ ■ xx yy xy ■ ▲ Γ-X Γ-X Γ-X 0.6 ■ m ■ m ■ m xx yy xy Γ-Y Γ-Y Γ-Y ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ ■ ▲ ■ ■ ▲ m ▲ m ▲ m xx yy xy 0.4 ■ 0.5 0.5 Γ-Y Γ-Y Γ-Y ■▲ ▲ m ▲ m ▲ m ■ ▲ ■ ▲ ▲ xx yy xy ■▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲▲▲ ▲ ■ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.2 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲▲ ■ ▲ ▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ▲ 0.0 0.0 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (a) Monolayer As band gap (eV) (b) Monolayer As m =m (c) Monolayer As m =m e 0 h 0 1.2 ■ E ■ E ■ E 1.4 ■ ■ xx yy xy 1.2 ■ ■ ■■ ■ 1.0 ■■ 1.2 ■ ■ ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ * * * ■ ■■ ■ ■ ■ ■ ▲ E ▲ E ▲ E ■ 1.0 ■ ■ ■■ ■ ■ ■ ■ xx yy xy ■ ■ ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ ■ ■ ■ 0.8 1.0 Γ-X Γ-X Γ-X ■ ■ ▲ m m m 0.8 ■ ▲ ■ ■ ■ ■ ▲ xx yy xy ■ ■▲ 0.6 ■▲ ▲ 0.8 Γ-X Γ-X Γ-X ■▲ ■ ▲ ■ m ■ m ■ m xx yy xy ■▲ ■▲ ■▲ ■▲ ■▲■■▲▲ ■▲ ■▲ ■▲ ■▲ ■▲ ■▲ ■▲ 0.6 ■▲ Γ-Y Γ-Y Γ-Y 0.6 0.4 ■▲ ■▲ m m m ■ ▲ ▲ ▲ xx yy xy ■▲ Γ-Y Γ-Y Γ-Y 0.4 m m m 0.4 ▲ ▲ ▲ ▲ xx yy xy 0.2 ■ ■▲▲ ▲ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ■▲ ▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.2 ▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ 0.2 ▲ ▲ ▲ ■▲▲ ▲ ▲ ▲ ▲ 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (d) Bilayer As band gap (eV) (e) Bilayer As m =m (f ) Bilayer As m =m e 0 h 0 2.5 0.30 3.0 Γ-X Γ-Y Γ-Y m m m Γ-X Γ-Y Γ-Y ■ ■ ▲ ▲ xx xx xy ■ m ▲ m ▲ m xx xx xy ▲ ■ 0.25 2.5 2.0 * Γ-X Γ-Y ■ Γ-X Γ-Y ■ E ▲ E ■ m ▲ m xx xx yy yy ■ m ▲ m 0.20 2.0 yy yy ■▲ 1.5 ▲ ■ ■▲ E ▲ ■ ▲ E ▲ yy 0.15 yy 1.5 ■ ■ ■ ■■ ■ ■▲ ■■ 1.0 ▲ ▲ ■ * ■ ■ 0.10 ■ E 1.0 ■ ■ ■■ ■ xy E ▲ ▲ ■ ▲ ■ xy ▲ ▲ ▲ ▲▲ ▲ 0.5 ▲ ▲▲ ▲ ■▲ ▲ ▲ ▲▲ ▲ ▲▲ ▲ ▲ ▲ 0.05 0.5 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲▲▲ ▲ ▲▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■▲ ■▲ ■▲ ■▲ 0.00 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (g) Bulk As band gap (eV) (h) Bulk As m =m (i) Bulk As m =m e 0 h 0 FIG. 3: (Color online) The relationships between the applied in-plane strains " (blue), xx " (red) and " (green) against [(a), (d), (g)] the direct (solid squares) and indirect yy xy (dashed triangles) band gaps (eV) ; [(b), (e), (h)] the e ective electron masses m =m along e 0 X (solid squares) and Y (dashed triangles) ; [(c), (f ), (i)] the e ective hole masses m =m along X (solid squares) and Y (dashed triangles); for each phase of As. h 0 support one-dimensional ballistic conduction, as well as possible Weyl states on o -symmetry points in the bulk at moderate levels of in-plane stress that are una ected by the SOC. We also observe several band gap transitions, in particular in the monolayer phase, which also include semi-conducting-metallic transitions. Finally, the e ective masses respond approxi- mately linearly with respect to uniaxial strain, except in the bulk, which exhibits quadratic behavior. Bandgap (eV) Bandgap (eV) Bandgap (eV) m /m m /m m /m h 0 e 0 e 0 m /m m /m m /m h 0 h 0 h 0 3. Antimony The relaxed Sb monolayer is found to possess an indirect band gap of 0:21 eV along Y (Fig. 4a), which is reasonably comparable to other PBE values 0.28 [29]-0.37 [114] eV, although these have been obtained by including SOC. For tensile strain along " the band yy gap opens, suggesting an indirect-direct transition for strains above 67%, and it diminishes for compressive strains before nally closing at " = 2%, where the material becomes yy a semi-metal. Similarly, the indirect gap closes along X at a compressive strain of " = 2% at which monolayer Sb again becomes semi-metallic. The indirect gap transitions xx to a direct gap at " = +1% tensile strain and remains so until nally closing at " = +4%, xx xx at which point we predict a potential Dirac state along Y at an o -symmetry point [37] Y (Fig. 5g) that has also been predicted in Ref. [51]. Fig. 5g depicts the calculated band structure, in which it is shown that SOC preserves the Dirac state but not does not open the band gap. The three-dimensional band structure about the Dirac point is shown in Fig. 5h in 6 1 which the maximum charge velocity is v = 4:31(1)10 ms and occurs along a line parallel to the Y direction (Fig. 5i). Moreover, the valence band at the X -point undergoes a Rashba splitting [115] due to SOC, which is also predicted to occur in the monolayers of -P [116], and -Sb [52, 117]. Finally, the electron e ective masses experience a weak linear response to strain (Figs. 4b & 4c), while the hole e ective masses along X respond much more strongly to a rapid broadening or attening of the valence band. Furthermore, the relaxed bilayer phase is found to be semi-metallic where an indirect band gap opens at " = +3% tensile strain and for uniaxial strains < 1% band-inversion yy at the -point leads to to a fully-metallic state. In addition, a possible Dirac state emerges at a similar non-symmetry-point Y along Y for " = +2% tensile strain (Fig. S13 in xx the Supplemental Material) and remains in place up to at least +5% strain. The maximum 6 1 charge velocity v = 4:47(3) 10 ms is also along Y and is approximately the same as that of the monolayer. The e ective masses (Figs. 4e & 4f) experience mild linear- response to strains prior to the transition to full metallicity, at which point a rapid attening of the bands at the -point suggesting strong electron localization Beyond a compressive strain of " = 3%, however, at a stress of 0.3 GPa, the bilayer undergoes a structural yy transition and buckles in the puckered (~y) direction. This buckled structure has a total energy 1.7 meV/atom lower than that of the relaxed state of the unperturbed -bilayer and 14 1.0 1.4 1.2 E E E ■■ ■ ■ xx ■ yy ■ xy ■ ■■ ■■ ■ ■ ■ ■ ■ ■ ■■■ ■■■ ■■ ■ ■ ■ ■■ ■■ ■■ ■■ ■■ ■ ■ 1.2 ■ ■ 0.8 1.0 ■ ■ ■ * * * ■ ■ ■■■ ■ ■ E E E ■ ■ ■ ■ ■ ■ ▲ ▲ ▲ Γ-X Γ-X Γ-Y ■ xx yy xy 1.0 ■ ■ m ■ m ▲ m ■ ■ 0.8 xx xy yy ■ ■ ■ 0.6 ■ ■ ■ ▲ 0.8 Γ-X Γ-X Γ-X Γ-Y ■ ▲ 0.6 Γ-Y Γ-Y ■ m m m m ■ ▲ m ▲ m ■ ■ ▲ ■ ▲ yy xx xy yy ■ xx xy ■ 0.6 0.4 ■ ■ ■ ■■■ ■ ■▲ ■ ■ ■ ■ ■ 0.4 ▲ ▲ ▲ ▲ ▲ ■ ■ ▲ ■ ■▲ ▲ ▲ Γ-X ▲ ▲ ▲ ▲ Γ-Y Γ-Y ■ ▲ ▲ ▲ ▲ 0.4 m ▲ ▲ ▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ■ m m ■ ▲ ▲ ▲ ▲ ▲ yy ▲ ▲ ■ ▲ ▲▲▲ ▲ ▲ xx xy ▲ ▲ ■▲▲ ▲ 0.2 ▲ ▲ 0.2 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.2 ▲ ▲ ▲ ▲▲ ▲▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ▲▲ ▲▲ ▲▲ ▲▲ ■▲ 0.0 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (a) Monolayer Sb band gap (eV) (b) Monolayer Sb m =m (c) Monolayer Sb m =m e 0 h 0 3.0 0.4 Γ-X Γ-X Γ-X m m m ■ ■ ■ Γ-X Γ-X Γ-Y xx yy xy ■ m ■ m ▲ m ■ xx xy yy 2.5 0.3 E E ■ Γ-Y Γ-Y Γ-Y ■ ▲ Γ-X xx xx m m m Γ-Y Γ-Y ▲ ▲ ▲ m xx yy xy 2.0 ■ m m yy ▲ ▲ xx xy ■ yy ▲ E 2 0.2 yy ■ 1.5 ■ ■ ■ * ■ ■ ■ ▲ ■ E ■ ■■ ■ ■ ■ ■■ ■■■ ■■ ■■ ■■ ■ ■ xy ▲ E ▲ ■■ ■■ ■ 1.0 ■ ■ ■ ■ ■ ■ ■ ■ ■■ ■■■ ■■■ ■ ■ ■ ■ ■ ■ xy 0.1 1 ▲ ▲ ▲ ▲ ▲ 0.5 ▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲▲ ▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■■▲▲ ■■▲▲ ■■▲▲■▲ ■■▲▲■▲ ■■▲▲■▲ ■▲■■▲▲ ■▲■■▲▲ ■■▲▲■▲ ■■▲▲ ■■▲▲ ■■▲▲ 0.0 0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (d) Bilayer Sb band gap (eV) (e) Bilayer Sb m =m (f ) Bilayer Sb m =m e 0 h 0 FIG. 4: (Color online) The relationships between the applied in-plane strains " (blue), xx " (red) and " (green) against [(a), (d), (g)] the direct (solid squares) and indirect yy xy (dashed triangles) band gaps (eV) ; [(b), (e), (h)] the e ective electron masses m =m e 0 along X (solid squares) and Y (dashed triangles) ; [(c), (f ), (i)] the e ective hole masses m =m along X (solid squares) and Y (dashed triangles); for each h 0 monolayer and bilayer Sb. 3.0 meV/atom lower when allowed to fully-relax, as shown above in Fig. 1c. This suggests the possible existence of a new structure that is attainable via strain. Finally, bulk Sb is found to be completely metallic for all levels of strain explored in this work. However, shear strains in this case do appear to have a signi cant e ect on the bands despite not opening a gap. In summary, we predict possible non-symmetry-point Dirac states in the strained monolayer and bilayer of Sb, which are qualitatively una ected by SOC, as well as Rashba splitting at the X -point in the monolayer. We also predict indirect-direct and indirect-semi-metallic transitions in the monolayer phase and a band gap opening in the bilayer phase. Finally, we observe a buckled state induced in bilayer Sb at 4% compressive strain. Bulk Sb was found to be metallic at all levels of strain explored. In Table II below we present a summary of the calculated band gaps and e ective charge Bandgap (eV) Bandgap (eV) m /m m /m e 0 e 0 m /m m /m h 0 h 0 carrier masses for the relaxed phases of each structure, and in Table III we provide a synopsis of the band gap and phase transitions of interest. Finally, in Fig. 5, we present the band structures of bilayer P and monolayer Sb, that form a representative sample of the di erent 0 0 Dirac points predicted at , X and Y , as well the three-dimensional band structures about the region of the points. E (eV) Me Me Mh Mh g X Y X Y P 0.9 1.25(1) 0.16(1) 2.8(2) 0.14(1) mono As 0:15 1.16(1) 0.26(1) 1.09(1) 0.18(2) mono Sb 0:2 1.10(1) 0.28(1) 1.04(1) 0.19(2) mono P 0.4 1.41(1) 0.19(1) 1.21(3) 0.15(2) bi As 0.45 1.15(1) 0.24(1) 0.94(4) 0.17(2) bi Sb 0 1.16(1) 0.39(1) 0.99(4) 0.33(4) bi P 0 - 0.37(2) - 0.21(2) bulk As 0 - 0.36(1) - 0.30(3) bulk Sb 0 - - - - bulk TABLE II: (Color online) Kohn-Sham band gaps (eV), indicating the indirect semiconducting (?), semi-metallic (y) and metallic (z) states, as well as the charge-carrier e ective masses (m ) for each phase of P, As, Sb. C. Isotropic bulk properties In order to obtain these electronic states, and to ensure accurate strain-engineering, knowledge of the mechanical properties is paramount. Therefore, in this section we review the mechanical response of the few-layer and bulk phases in order to compute the elastic properties, both isotropically averaged and as a function of orientation of applied in-plane stress. The computed elements of the sti ness tensor C in GPa of each structure are given in Table S1 in the Supplemental Material, where those pertaining to bulk P compare well to experiments [118, 119] and similarly computed values [96, 108]. For the elements related to in-plane strains (c , c , c , c ) we observe an expected increase in sti ness as the layer 11 22 66 12 16 0.2 0° 30° 60° 90° 0.1 0.0 -1 -0.1 -2 -0.2 Γ X S Γ Y S -0.04 -0.02 Γ 0.02 0.04 (a) Bilayer P band structure " = 5% (b) (c) -point Dirac state xx 0.2 0° 30° 60° 90° 0.1 0.0 -1 -0.1 X' -2 Γ X S Γ Y S -0.2 -0.04 -0.02 X' 0.02 0.04 (d) Bilayer P band structure " = 5% (e) (f ) X -point Dirac state yy 0.15 0.10 0° 0.05 30° 60° 0.00 90° -1 -0.05 -2 Γ X S Γ Y' Y S -0.10 -0.02 -0.01 Y' 0.01 0.02 (g) Monolayer Sb band structure " = 3% (h) (i) Y -point Dirac state xx FIG. 5: (Color online) a) Band structure of bilayer P at " = 5% with SOC (thick xx lines), and without SOC (thin lines) b) three-dimensional bands about the predicted Dirac point at c) slices through the Dirac point at 0 , 30 , 60 and 90 relative to the X line that indicate highly anisotropic conduction. [5d,5e,5f] Illustrates the same for bilayer P at " = 5%, where the Dirac state occurs at the non-symmetry point X along X . yy [5g,5h,5i] Illustrates the same for monolayer Sb at " = 3%, where the possible Dirac state xx occurs at the non-symmetry point Y along Y . Energy (eV) Energy (eV) Energy (eV) Energy (eV) Energy (eV) Energy (eV) Transition Direction Strain (%) P D Gap ! SM XX,YY -5 bi D Gap ! ID Gap XX,YY +2 P SM ! D Gap ! ID Gap XX,YY +1!+3 bulk SM!SM XX,YY +2 As ID ! D Gap XX -3 mono ID Gap ! SM XX +2 ID Gap ! SM XX +2 SM ! SM XX +5 As D Gap ! ID Gap ! D Gap XX +2!+3 bi D Gap ! ID Gap YY -3,+2 As SM ! D Gap ! ID Gap XX 0!+3 bulk SM ! D Gap YY 0!+3 SM!SM XX,YY +1 Sb ID Gap ! SM XX -2 mono ID Gap ! D Gap ! SM XX +2!+4 ID Gap ! SM YY -2 Sb SM ! SM XX +4 bi Structural Transition YY -3 SM ! ID Gap YY +3 TABLE III: Summary of the band gap transitions including direct (D); indirect (ID); metallic (M); and semi-metallic (SM), in particular those that indicate potential Dirac states (4), Weyl states (O), and the structural phase transition. number increases and for increasing atomic number. However, for other elements relating to out-of-plane and shear stresses (c , c , c , c , c ) the sti ness actually increases in the 33 44 55 23 13 bulk phase from P to As to Sb. The Hill-averaged bulk properties are presented in Table IV, which compare well to other DFT values [96, 108], while our calculated bulk modulus for bulk P (37.2 GPa) is also within reasonable range of the experimental values (32.32 [100] - 36.02 GPa [120]). We also observe 18 Y (GPa) G (GPa)  B (GPa) H H H H a a a a b c P 61.1 (70.3 ) 24.9 (29.4 ) 0.23 (0.30 ) 37.2 (38.5 , 32.32 , 36.02 ) bulk As 60.0 23.8 0.26 41.4 bulk Sb 52.5 21.1 0.24 33.8 bulk Ref. [96] Ref. [100] Ref. [120] TABLE IV: The Hill-averaged Young's modulus Y , shear modulus G and bulk modulus H H B in GPa, and Poisson's ratio  for bulk P, As and Sb compared to similarly calculated DFT [96] values and available experimental data [100, 120]. that the bulk properties remain largely comparable for all the species, but generally decrease from P to As to Sb (except for the Poisson's ratio and bulk modulus, which are largest for As). We also note that while bulk P has the largest in-plane responses, As and Sb have larger out-of-plane and shear responses, which enable the net isotropic properties for all the species to remain comparable overall. D. In-plane elastic properties In section S1 of the Supplemental Material we re-derive the equations for the elastic properties as a function of the in-plane orientation angle , de ned in Fig. 1, as outlined in Ref. [121]. These functions are the plotted in Fig. 6 and include the Young's modulus Y () and it's average hY ()i, the shear modulus G (), and Poisson's ratio  (). The experimental Young's modulus of 130 GPa, determined in Ref. [11] via the ROM of a nano- ake-polymer composite, lies precisely between the average in-plane bulk Young's modulus 85.7 GPa and the elastic sti ness in the zigzag (~x) direction 187.9 GPa, thus tting the results of our model reasonably well. The anisotropy of the elastic properties is also apparent in the mechanical pro les, partic- ularly with regard to the Young's modulus, which has a 2-fold symmetry about the x-axis, in contrast to the shear modulus and Poisson's ratios, which display 4-fold symmetry about both the axes (except for the Poisson's ratio of P, which remains 2-fold symmetric). Indeed, 19 π π 2 2 2 3π π 3π π 3π π 4 4 4 4 4 4 π 0 π 0 π 0 0. 20. 40. 60. 0. 10. 20. 30. 40. 0. 10. 20. 30. 40. 5π 7π 5π 7π 5π 7π 4 4 4 4 4 4 3π 3π 3π 2 2 Y(θ) G(θ) 100×ν(θ) ⟨Y(θ)⟩ Y (θ) G (θ) 50×ν (θ) ⟨Y (θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ xy xy xy xy (a) Monolayer P (b) Monolayer As (c) Monolayer Sb π π π 2 2 2 3π π 3π π 3π π 4 4 4 4 4 4 π 0 π 0 π 0 0. 25. 50. 75. 0. 15. 30. 45. 0. 10. 20. 30. 5π 7π 5π 7π 5π 7π 4 4 4 4 4 4 3π 3π 3π 2 2 2 Y(θ) G(θ) 100×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ (d) Bilayer P (e) Bilayer As (f ) Bilayer Sb π π π 2 2 2 π π 3π 3π 3π π 4 4 4 4 4 4 π 0 π 0 π 0 0. 50. 100. 150. 0. 30. 60. 90. 0. 15. 30. 45. 60. 5π 7π 5π 7π 5π 7π 4 4 4 4 4 4 3π 3π 3π 2 2 Y(θ) G(θ) 100×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y⟩ (g) Bulk P (h) Bulk As (i) Bulk Sb FIG. 6: (Color online) In-plane functions for the Young's modulus Y () (blue) and its isotropic average hY ()i (red) in units of GPa; shear modulus G () in GPa (orange); and Poisson's ratio  () (green, scaled by 100 for P and 50 for As, Sb) for each of the structures. 20 Y  Y  hYi G  G  hGi     hi min max min max min max P 17.1 69 72.4 0 36.4 20.9 90 42.2 45 31.0 0.3 31 0.5 63 0.5 mono As -2.1 90 30.5 18 12.7 10.4 77 13.8 45 11.2 0.2 39 0.9 90 0.4 mono Sb 1.4 90 24.4 22 12.6 8.9 45 13.6 90 9.4 0.3 43 0.8 90 0.6 mono Pb 20.5 75 85.6 0 44.4 28.1 90 49.2 45 37.8 0.31 34 0.5 68 0.4 bi As 9.7 90 56.0 0 31.0 23.0 78 28.6 45 24.3 0.4 40 0.6 90 0.5 bi Sb 2.6 90 31.6 33 16.3 11.8 45 19.5 90 12.8 0.2 42 0.8 90 0.5 bi P 34.2 87 166.4 0 85.7 64.9 87 92.4 45 74.5 0.2 36 0.5 75 0.3 bulk As 10.9 90 105.1 20 55.0 44.4 73 51.1 45 66.7 0.3 40 0.7 90 0.4 bulk Sb 12.4 90 58.9 30 35.4 28.1 59 28.2 45 63.2 0.2 41 0.6 90 0.4 bulk TABLE V: Summary of the minima and maxima of the Hill-averaged in-plane Young's modulus Y () (GPa), shear modulus G () (GPa), and Poisson's ratio  () as well as the angle  with respect to the ~x-direction (zigzag) at which they occur in degrees, and their in-plane averages. for a given species, the general shape of each pro le is approximately preserved with respect to the number of layers, while the range of each property tends to increase. This discovery is advantageous in the strain-engineering of nano- akes since one may now forecast in advance the response of a material to in-plane strain, once the underlying pro le and number of layers are known. Another interesting feature is that the extrema of the elastic functions do not necessarily coincide with the coordinate-axes. For instance the Young's modulus maximum for mono- layer Sb occurs at 22 . Table V summarizes the global minima and maxima of each function and the angles at which they occur. While most of the function extrema occur expectedly at 0 , 45 or 90 , many are incident away from the coordinate-axes. This result lends further insight into the mechanical anisotropy of the orthorhombic group-V materials. The emergence of a negative Young's modulus of 2:1 GPa in monolayer As (Fig. 6b), at rst glance, may give cause for concern. It arises due to a negative Voigt estimate for the 2 2 Young's modulus at 90 where (c c )=c = 8:9 GPa (since c > c ), which is larger 22 12 22 22 12 in absolute magnitude than the Reuss estimate at 90 given by 1=s = 4:7 GPa and results 21 in a net negative Hill-average. In this instance, we surmise that either the assumptions of the Voigt model break down, or the Hill-method is not universally appropriate in arbitrary directions in-plane and a more robust averaging scheme must be employed. Nevertheless, the qualitative in-plane functions and their isotropic averages remain physically meaningful and in general can provide valuable physical insight. In contrast to the isotropic averages for the bulk properties in Table IV, which remained largely comparable for P, As and Sb, a much clearer trend across the species emerges once we have eliminated contributions from the out-of-plane and shear stresses. In this instance, P clearly possesses superior in-plane mechanical strength in both moduli, which decrease with the number of layers as expected As and Sb are largely similar in the monolayer, though less so in the bilayer and bulk phases, where they are stronger in the ~x-direction. In contrast, the Poisson's ratio tends to remain relatively stable aside from generally decreasing with increasing number of layers. In summary, there exists in the elastic properties a broad range of responses, pro le shapes and behavior that is re ective of the underlying anisotropic crystal structure, which also strongly depend on the number of layers with P typically the sti est and Sb the most exible. In general we nd the shapes of the in-plane response pro les to be conveniently consistent, which implies that, for a given number of layers, the in-plane elastic response of a nano- ake can be reliably estimated a priori. IV. CONCLUSION We have extensively explored the mechanical and electronic properties of P, As and Sb in their few-layer and bulk phases. We have identi ed several band gap transitions in almost all of the structures. The SOC tended to close the bands by  0:05 eV but did not alter any of our qualitative ndings. We also predict the existence of Dirac states in the strained phases of monolayer As and Sb, bilayer P, As and Sb as well as possible Weyl states in bulk P and As for moderate levels of strain. The linear-dispersion was observed along Y of each of the predicted Dirac or Weyl states, corresponding to the direction of softest mechanical response 6 1 in the puckered direction. The maximum charge velocity is calculated to be over 10 ms . In particular, for bilayer P and few-layer As we predict highly anisotropic conductivity dominated by ballistic transport along the puckered direction that is indicative of e ective, 22 one-dimensional conduction. We predict that an appropriate strain could yield these e ects in experiments. We also observe the existence of a notable buckled state of compressed bilayer Sb at 4% strain. Finally, the angular-resolved elastic properties as well as the stress-dependence of the Kohn-Shame band gaps and charge-carrier e ective masses revealed highly anisotropic behavior, spanning a broad range of values, and angular-dependent behavior that has be- come characteristic of these group-V layered structures. Moreover, the critical stresses at which these transitions occur are expected to be experimentally accessible and highly switch- able, paving the way for possible veri cation in the near future. Thus, the group-V layered materials are poised to become central to the next generation of electronic devices with po- tential novel applications in eld-e ect transistors; batteries; gas-sensors and opto-electronic devices. ACKNOWLEDGMENTS The authors would like to sincerely thank Damien Hanlon, Claudia Backes, Conor Boland, Jonathan Coleman, Beata Szydlowska, Gaozhong Wang, and Werner Blau for their helpful discussions relating to the work conducted for this Article. This work was enabled by Science Foundation Ireland (SFI) funded centre AMBER (SFI/12/RC/2278). All calculations were performed on the Kelvin cluster maintained by Trinity College Dublin Research IT and funded through grants from SFI. [1] M. Wu, H. Fu, L. Zhou, K. Yao, and X. C. Zeng, Nano Lett. 15, 3557 (2015). [2] J. Guan, Z. Zhu, and D. Tom anek, Phys. Rev. Lett. 113, 046804 (2014). [3] Z. Zhu and D. Tom anek, Phys. Rev. Lett. 112, 176802 (2014). [4] J. Guan, Z. Zhu, and D. Tomnek, ACS Nano 8, 12763 (2014). [5] L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen, and Y. Zhang, Nat. Nano 9, 372 (2014). [6] F. Xia, H. Wang, and Y. Jia, Nat. Commun. 5, 4458 EP (2014). [7] S. P. Koenig, R. A. Doganov, H. Schmidt, A. H. C. Neto, and B. Ozyilmaz, Appl. Phys. Lett. 104, 103106 (2014). 23 [8] H. Liu, A. T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tomnek, and P. D. Ye, ACS Nano 8, 4033 (2014). [9] X. Ling, H. Wang, S. Huang, F. Xia, and M. S. Dresselhaus, Proc. Natl. Acad. Sci. U.S.A. 112, 4523 (2015). [10] J. R. Brent, N. Savjani, E. A. Lewis, S. J. Haigh, D. J. Lewis, and P. O'Brien, Chem. Commun. 50, 13338 (2014). [11] D. Hanlon, C. Backes, E. Doherty, C. S. Cucinotta, N. C. Berner, C. Boland, K. Lee, A. Har- vey, P. Lynch, Z. Gholamvand, S. Zhang, K. Wang, G. Moynihan, A. Pokle, Q. M. Ramasse, N. McEvoy, W. J. Blau, J. Wang, G. Abellan, F. Hauke, A. Hirsch, S. Sanvito, D. D. O'Regan, G. S. Duesberg, V. Nicolosi, and J. N. Coleman, Nat. Commun. 6, 8563 (2015). [12] A. S. Rodin, A. Carvalho, and A. H. Castro Neto, Phys. Rev. Lett. 112, 176801 (2014). [13] X. Peng, Q. Wei, and A. Copple, Phys. Rev. B 90, 085402 (2014). [14] J.-W. Jiang and H. S. Park, Phys. Rev. B 91, 235118 (2015). [15] B. Sa, Y.-L. Li, J. Qi, R. Ahuja, and Z. Sun, J. Phys. Chem. C 118, 26560 (2014). [16] M. Elahi, K. Khaliji, S. M. Tabatabaei, M. Pourfath, and R. Asgari, Phys. Rev. B 91, 115412 (2015). [17] T. Hu, Y. Han, and J. Dong, Nanotechnology 25, 455703 (2014). [18] J.-W. Jiang and H. S. Park, Nat. Commun. 5, 4727 EP (2014). [19] R. Fei and L. Yang, Nano Lett. 14, 2884 (2014). [20] H. Liu, Y. Du, Y. Deng, and P. D. Ye, Chem. Soc. Rev. 44, 2732 (2015). [21] J. Qiao, X. Kong, Z.-X. Hu, F. Yang, and W. Ji, Nat. Commun. 5, 4475 EP (2014). [22] V. Tran, R. Soklaski, Y. Liang, and L. Yang, Phys. Rev. B 89, 235319 (2014). [23] R. W. Keyes, Phys. Rev. 92, 580 (1953). [24] Q. Wei and X. Peng, Appl. Phys. Lett. 104, 251915 (2014). [25] C. Yongqing, Z. Gang, and Z. Yong-Wei, Sci. Rep. , 1 (2014). [26] A. C.-G. Zant, L. Vicarelli, E. Prada, J. O. Island, K. L. Narasimha-Acharya, S. I. Blanter, D. J. Groenendijk, M. Buscema, G. A. Steele, J. V. Alvarez, H. W. Zandbergen, J. J. Palacios, and H. S. J. van der, 2D Mater. 1, 025001 (2014). [27] K. Momma and F. Izumi, J. Appl. Crystallogr. 44, 1272 (2011). [28] C. Kamal and M. Ezawa, Phys. Rev. B 91, 085423 (2015). [29] G. Wang, R. Pandey, and S. P. Karna, ACS Appl. Mater. Inter. 7, 11490 (2015). 24 [30] I. Kokubo, Y. Yoshiike, K. Nakatsuji, and H. Hirayama, Phys. Rev. B 91, 075429 (2015). [31] Y. Lu, W. Xu, M. Zeng, G. Yao, L. Shen, M. Yang, Z. Luo, F. Pan, K. Wu, T. Das, P. He, J. Jiang, J. Martin, Y. P. Feng, H. Lin, and X.-s. Wang, Nano Lett. 15, 80 (2015). [32] Z. Zhu, J. Guan, and D. Tomnek, Nano Lett. 15, 6042 (2015). [33] F. Shojaei and H. S. Kang, J. Phys. Chem. C 119, 20210 (2015). [34] M. Xie, S. Zhang, B. Cai, Y. Huang, Y. Zou, B. Guo, Y. Gu, and H. Zeng, Nano Energy 28, 433 (2016). [35] B. Liu, M. Kpf, A. N. Abbas, X. Wang, Q. Guo, Y. Jia, F. Xia, R. Weihrich, F. Bachhuber, F. Pielnhofer, H. Wang, R. Dhall, S. B. Cronin, M. Ge, X. Fang, T. Nilges, and C. Zhou, Adv. Mater 27, 4423 (2015). [36] L. Kou, Y. Ma, X. Tan, T. Frauenheim, A. Du, and S. Smith, J. Phys. Chem. C 119, 6918 (2015). [37] V. Pardo and W. E. Pickett, Phys. Rev. Lett. 102, 166803 (2009). [38] C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, I. Leermakers, U. Zeitler, Y. Sk- ourski, J. Wosnitza, Z. Liu, Y. Chen, W. Schnelle, H. Borrmann, Y. Grin, C. Felser, and B. Yan, Nat Phys 11, 645 (2015), letter. [39] X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, Phys. Rev. B 83, 205101 (2011). [40] H. Nielsen and M. Ninomiya, Phys. Lett. B 130, 389 (1983). [41] A. M. Turner, A. Vishwanath, and C. O. Head, Topological Insulators 6, 293 (2013). [42] S. A. Parameswaran, T. Grover, D. A. Abanin, D. A. Pesin, and A. Vishwanath, Phys. Rev. X 4, 031035 (2014). [43] S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J. Mele, and A. M. Rappe, Phys. Rev. Lett. 108, 140405 (2012). [44] Z. Wang, Y. Sun, X.-Q. Chen, C. Franchini, G. Xu, H. Weng, X. Dai, and Z. Fang, Phys. Rev. B 85, 195320 (2012). [45] Z. Wang, H. Weng, Q. Wu, X. Dai, and Z. Fang, Phys. Rev. B 88, 125427 (2013). [46] Z. Liu, J. Wang, and J. Li, Phys. Chem. Chem. Phys. 15, 18855 (2013). [47] J. Wang, S. Deng, Z. Liu, and Z. Liu, Nat. Sci. Rev. 2, 22 (2015). [48] J. Kim, S. S. Baik, S. H. Ryu, Y. Sohn, S. Park, B.-G. Park, J. Denlinger, Y. Yi, H. J. Choi, and K. S. Kim, Science 349, 723 (2015). [49] C. Wang, Q. Xia, Y. Nie, M. Rahman, and G. Guo, AIP Advances 6, 035204 (2016). 25 [50] G. Yao, Z. Luo, F. Pan, W. Xu, Y. P. Feng, and X.-s. Wang, Sci. Rep. 3, 2010 EP (2013). [51] Y. Lu, D. Zhou, G. Chang, S. Guan, W. Chen, Y. Jiang, J. Jiang, X.-s. Wang, S. A. Yang, Y. P. Feng, Y. Kawazoe, and H. Lin, NPJ Comput. Mater. 2, 16011 EP (2016). [52] M. Zhao, X. Zhang, and L. Li, Sci Rep 5, 16108 (2015). [53] S. Zhang, M. Xie, B. Cai, H. Zhang, Y. Ma, Z. Chen, Z. Zhu, Z. Hu, and H. Zeng, Phys. Rev. B 93, 245303 (2016). [54] H. Doh and H. J. Choi, 2D Mater. 4, 025071 (2017). [55] S. Zhang, Z. Yan, Y. Li, Z. Chen, and H. Zeng, Angew. Chem. Int. Ed. 54, 3112 (2015). [56] A. Manjanath, A. Samanta, T. Pandey, and A. K. Singh, Nanotechnology 26, 075701 (2015). [57] J. Han, J. Xie, Z. Zhang, D. Yang, M. Si, and D. Xue, Appl. Phys. Express 8, 041801 (2015). [58] Z. Zhu, J. Guan, and D. Tom anek, Phys. Rev. B 91, 161404 (2015). [59] T. Hu and J. Dong, Phys. Rev. B 92, 064114 (2015). [60] Z. J. Xiang, G. J. Ye, C. Shang, B. Lei, N. Z. Wang, K. S. Yang, D. Y. Liu, F. B. Meng, X. G. Luo, L. J. Zou, Z. Sun, Y. Zhang, and X. H. Chen, Phys. Rev. Lett. 115, 186403 (2015). [61] P.-L. Gong, D.-Y. Liu, K.-S. Yang, Z.-J. Xiang, X.-H. Chen, Z. Zeng, S.-Q. Shen, and L.-J. Zou, Phys. Rev. B 93, 195434 (2016). [62] R. Fei, V. Tran, and L. Yang, Phys. Rev. B 91, 195319 (2015). [63] A. N. Abbas, B. Liu, L. Chen, Y. Ma, S. Cong, N. Aroonyadet, M. Kpf, T. Nilges, and C. Zhou, ACS Nano 9, 5618 (2015). [64] L. Kou, T. Frauenheim, and C. Chen, J. Phys. Chem. Lett. 5, 2675 (2014). [65] T. Low, A. S. Rodin, A. Carvalho, Y. Jiang, H. Wang, F. Xia, and A. H. Castro Neto, Phys. Rev. B 90, 075434 (2014). [66] M. Baba, Y. Takeda, K. Shibata, T. Ikeda, and A. Morita, Jpn. J. Appl. Phys. 28, L2104 (1989). [67] C.-M. Park and H.-J. Sohn, Adv. Mater 19, 2465 (2007). [68] L.-Q. Sun, M.-J. Li, K. Sun, S.-H. Yu, R.-S. Wang, and H.-M. Xie, J. Phys. Chem. C 116, 14772 (2012). [69] J. Sun, H.-W. Lee, M. Pasta, H. Yuan, G. Zheng, Y. Sun, Y. Li, and Y. Cui, Nat. Nano 10, 980 (2015). 26 [70] J. Sun, G. Zheng, H.-W. Lee, N. Liu, H. Wang, H. Yao, W. Yang, and Y. Cui, Nano Lett. 14, 4573 (2014). [71] S. S. Baik, K. S. Kim, Y. Yi, and H. J. Choi, Nano Lett. 15, 7788 (2015). [72] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.: Condens. Mat. 21, 395502 (2009). [73] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996). [74] D. Vanderbilt, Phys. Rev. B 41, 7892 (1990). [75] E. Kucukbenli, M. Monni, B. Adetunji, X. Ge, G. Adebayo, N. Marzari, S. De Gironcoli, and A. D. Corso, arXiv preprint arXiv:1404.3015 (2014). [76] S. Grimme, J. Comput. Chem. 27, 1787 (2006). [77] N. Marzari, D. Vanderbilt, A. De Vita, and M. C. Payne, Phys. Rev. Lett. 82, 3296 (1999). [78] V. Iyer, P. Ye, and X. Xu, 2D Mater. 4, 021032 (2017). [79] C. R. Dean, F. A. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K. Watanabe, T. Taniguchi, P. Kim, K. L. Shepard, and J. Hone, Nat. Nano 5, 722 (2010). [80] K.-K. Liu, W. Zhang, Y.-H. Lee, Y.-C. Lin, M.-T. Chang, C.-Y. Su, C.-S. Chang, H. Li, Y. Shi, H. Zhang, C.-S. Lai, and L.-J. Li, Nano Lett. 12, 1538 (2012). [81] J. C. H. A dl and J. L. Kardos, Polym. Eng. Sci. 16, 344 (1976). [82] R. Mullen, R. Ballarini, Y. Yin, and A. Heuer, Acta. Mater. 45, 2247 (1997). [83] R. Hearmon, Physics of the Solid State , 401 (1969). [84] J. M. J. den Toonder, J. A. W. van Dommelen, and F. P. T. Baaijens, Model. Simul. Mater. Sc. 7, 909 (1999). [85] A. Reuss, ZAMM-Z Angew. Math. Mech. 9, 49 (1929). [86] R. D. Cook and W. C. Young, Advanced Mechanics of Materials (Pearson College Division, 1999). [87] W. Voigt, Ann. Phys. 274, 573 (1889). [88] R. Hill, P. Phys. Soc. 65, 349 (1952). 27 [89] U. Khan, P. May, A. O'Neill, A. P. Bell, E. Boussac, A. Martin, J. Semple, and J. N. Coleman, Nanoscale 5, 581587 (2013). [90] W. E. Mahmoud, Eur. Polym. J. 47, 1534 (2011). [91] J. Liang, Y. Huang, L. Zhang, Y. Wang, Y. Ma, T. Guo, and Y. Chen, Adv. Funct. Mater. 19, 2297 (2009). [92] J. A. King, D. R. Klimek, I. Miskioglu, and G. M. Odegard, J. Appl. Polym. Sci. 128, 4217 (2013). [93] R. J. Young, I. A. Kinloch, L. Gong, and K. S. Novoselov, Compos. Sci. Technol. 72, 1459 (2012). [94] Y. Yang, W. Rigdon, X. Huang, and X. Li, Sci. Rep. 3, 2086 EP (2013). [95] S.-K. Kim, J. J. Wie, Q. Mahmood, and H. S. Park, Nanoscale 6, 7430 (2014). [96] S. Appalakondaiah, G. Vaitheeswaran, S. Leb egue, N. E. Christensen, and A. Svane, Phys. Rev. B 86, 035105 (2012). [97] D. Kecik, E. Durgun, and S. Ciraci, Phys. Rev. B 94, 205410 (2016). [98] Z. Zhang, J. Xie, D. Yang, Y. Wang, M. Si, and D. Xue, Appl. Phys. Express 8, 055201 (2015). [99] Y. Xu, B. Peng, H. Zhang, H. Shao, R. Zhang, and H. Zhu, Ann. Phys. 529, 1600152 (2017). [100] L. Cartz, S. R. Srinivasa, R. J. Riedner, J. D. Jorgensen, and T. G. Worlton, J. Chem. Phys. 71, 1718 (1979). [101] P. M. Smith, A. J. Leadbetter, and A. J. Apling, Philos. Mag. 31, 57 (1975). [102] C. S. Barrett, P. Cucka, and K. Haefner, Acta Criystallogr. 16, 451 (1963). [103] D. C  ak r, H. Sahin, and F. m. c. M. Peeters, Phys. Rev. B 90, 205421 (2014). [104] A. J. Cohen, P. Mori-S anchez, and W. Yang, Phys. Rev. B 77, 115123 (2008). [105] Y. Hinuma, A. Gruneis,  G. Kresse, and F. Oba, Phys. Rev. B 90, 155405 (2014). [106] U. von Barth, Phys. Scripta 2004, 9 (2004). [107] A. J. Cohen, P. Mori-Snchez, and W. Yang, Chem. Rev. 112, 289 (2012). [108] Y. Wang and Y. Ding, Nanoscale Res. Lett. 10, 254 (2015). [109] A. Morita, Appl. Phys. A 39, 227 (1986). [110] D. Warschauer, J. Appl. Phys. 34, 1853 (1963). [111] S. Narita, Y. Akahama, Y. Tsukiyama, K. Muro, S. Mori, S. Endo, M. Taniguchi, M. Seki, S. Suga, A. Mikuni, and H. Kanzaki, Physica B&C 117, 422 (1983). 28 [112] Y. Maruyama, S. Suzuki, K. Kobayashi, and S. Tanuma, Physica B&C 105, 99 (1981). [113] G. Greaves, S. Elliott, and E. Davis, Adv. Phys. 28, 49 (1979). [114] O. U. Akturk,  V. O. Oz celik, and S. Ciraci, Phys. Rev. B 91, 235446 (2015). [115] G. Bihlmayer, O. Rader, and R. Winkler, New J. Phys. 17, 050202 (2015). [116] Z. S. Popovi c, J. M. Kurdestany, and S. Satpathy, Phys. Rev. B 92, 035135 (2015). [117] L. Yang, Y. Song, W. Mi, and X. Wang, RSC Adv. 6, 66140 (2016). [118] M. Yoshizawa, I. Shirotani, and T. Fujimura, J. Phys. Soc. Jpn. 55, 1196 (1986). [119] Y. Kzuki, Y. Hanayama, M. Kimura, T. Nishitake, and S. Endo, J. Phys. Soc. Jpn. 60, 1612 (1991). [120] T. Akai, S. Endo, Y. Akahama, K. Koto, and Y. Marljyama, High Pressure Res. 1, 115 (1989). [121] R. M. Jones, Mechanics of Composite Materials, Vol. 193 (Scripta Book Company Washing- ton, DC, 1975). http://www.deepdyve.com/assets/images/DeepDyve-Logo-lg.png Condensed Matter arXiv (Cornell University)

Strain-induced Weyl and Dirac states and direct-indirect gap transitions in group-V materials

Condensed Matter , Volume 2018 (1801) – Jan 24, 2018

Loading next page...
 
/lp/arxiv-cornell-university/strain-induced-weyl-and-dirac-states-and-direct-indirect-gap-E0Wo4gK0q4

References (131)

ISSN
2053-1583
eISSN
ARCH-3331
DOI
10.1088/2053-1583/aa89d2
Publisher site
See Article on Publisher Site

Abstract

We perform comprehensive density-functional theory calculations on strained two-dimensional phosphorus (P), arsenic (As) and antimony (Sb) in the monolayer, bilayer, and bulk -phase, from which we compute the key mechanical and electronic properties of these materials. Speci cally, we compute their electronic band structures, band gaps, and charge-carrier e ective masses, and identify the qualitative electronic and structural transitions that may occur. Moreover, we compute the elastic properties such as the Young's modulus Y ; shear modulus G; bulk modulus B; and Poisson ratio  and present their isotropic averages of as well as their dependence on the in-plane orientation, for which the relevant expressions are derived. We predict strain-induced Dirac states in the monolayers of As and Sb and the bilayers of P, As, and Sb, as well as the possible existence of Weyl states in the bulk phases of P and As. These phases are predicted to support charge velocities 6 1 up to 10 ms and, in some highly anisotropic cases, permit one-dimensional ballistic conductivity in the puckered direction. We also predict numerous band gap transitions for moderate in-plane stresses. Our results contribute to the mounting evidence for the utility of these materials, made possible by their broad range in tuneable properties, and facilitate the directed exploration of their potential application in next-generation electronics. PACS numbers: 62.25.-g, 73.61.-r, 81.07.-b omuinneg@tcd.ie arXiv:1801.08233v1 [cond-mat.mtrl-sci] 24 Jan 2018 I. BACKGROUND Two-dimensional black phosphorus (BP), or phosphorene, is one of several predicted sta- ble allotropes of few-layer phosphorus [1{4], and it has attracted considerable attention since its recent successful synthesis [5{9] that is now possible with liquid phase exfoliation [10, 11]. The excitement behind BP is driven by its growing list of technologically relevant anisotropic mechanical and electronic properties. The theoretically predicted properties include a tune- able band-gap [8, 12{17], a negative Poisson's ratio [18], anisotropic conduction [19, 20], and linear dichroism [21, 22]. The properties that have been experimentally veri ed so far in- clude a high hole-mobility between 300 1000 cm /Vs [7, 8, 21, 23], considerable mechanical exibility [24], and a layer-dependent band gap [22, 25, 26] ranging from 0.3 eV in bulk to 2.0 eV in the monolayer. The anisotropic crystal structure of BP is responsible for its unusual electro-mechanical properties, which are predicted to be strongly directional-dependent and highly responsive to mechanically strain [11, 19]. With this renewed interest in BP, focus has quickly turned to few-layer phases of the other pnictogens, namely arsenic [28] (As), antimony [29] (Sb), bismuth [30, 31] (Bi), and their alloys [32{36], which are attracting steadily increasing attention. Moreover, the search for a bulk counterpart to graphene, capable of supporting ballistic electron transport, has recently driven the search for new Weyl and Dirac semi-metals [37{47]. In spite of the diculty in attaining Dirac points in two-dimensional materials [46, 47], the rst experimental Dirac semi-metal in few-layer BP was observed by Kim et al. [48], while As [49], Sb [50{52], Bi [51], and P [51{54] are also predicted to be potential candidates. Indeed, many of the predicted strain-induced properties of these materials, such as direct-indirect band gap transitions [13, 28, 29, 55, 56], a negative Poisson's ratio [18, 57], as well as electronic [2, 55, 58], structural [59], and topological [51, 60{62] transitions, are already spurring their incorporation in emergent technologies such as eld-e ect transistors [8, 20], gas-sensors [63, 64], optical switches [65, 66], solar-cells [34], next-generation batteries [67{69], reinforcing llers [11, 70], and topological insulators [50, 51, 60{62]. In this work, our aim is to provide a comprehensive analysis of the monolayer, bilayer and bulk phases of orthorhombic P, As and Sb, in order to identify and compare the quali- tative strain-related properties of each structure from a consistent set of calculations, thus treating each material on the same footing. Our ndings provide new insights into their 2 (b) (a) (c) FIG. 1: (Color online) a) Top and side view of the 2D orthorhombic puckered structure (generated in VESTA [27]) with the primary vectors along the zigzag (~x) and puckered (~y) directions shown. The unit cell, given by the shaded region, is described by the lattice parameters a and b with the in-plane angle  also de ned. b) 3D Brillouin zone with high-symmetry points , X , S, Y , and Z . c) Side-view of the buckled Sb state at " = 4% compressive strain. yy electro-mechanical properties, especially regarding arsenic and antimony, which have been relatively unexplored to date. Speci cally, we identify qualitative transitions in band gaps, e ective masses, structure, and topology that occur at various strains, and compute the elastic properties that determine the required stresses to attain these electronic states. We predict the existence of strain-induced Dirac states in monolayer As and Sb, bilayer P, As and Sb, as well as possible Weyl states in bulk P and As, at moderate stress values. Our ndings show that all of the predicted Dirac and Weyl points are indeed linear, at 3 least in the Y direction, i.e. the puckered direction. Thus, following the convention of terminology found in the Refs. [38, 48, 71], and other sources, we classify Dirac or Weyl states as those associated with regions of sustained linear dispersion in the band structure, at or near the Fermi level, in at least one direction. We predict these states to support ballistic conduction and are una ected by the spin-orbit coupling (SOC). In particular, few- layer P and As exhibit a strong indication of anisotropic conduction, dominated by ballistic conductivity along Y . The outline of this article is as follows: In Section II we review the details of our calcula- tions and discuss the Voigt-Reuss-Hill averaging scheme used to compute the experimentally- relevant elastic properties. In Section III we present the results of our calculations including the lattice constants, strain dependence of electronic properties, in particular the potential Dirac and Weyl states identi ed, and computed elastic properties. We conclude in Section IV with a discussion of the implications of our key results. II. METHODOLOGY A. Calculation details The calculations were performed with the QuantumEspresso package [72] using the Perdew-Burke-Ernzerhof (PBE) form of the generalized-gradient approximation (GGA) exchange-correlation functional [73]. An ultrasoft pseudopotential [74] from the SSSP Li- brary [75] (with 5 valence electrons) was used to represent the core electrons. Non-SOC calculations were initially performed and those that exhibited potential Dirac or Weyl states were reassessed including non-perturbative SOC. In all calculations, van der Waals (vdW) in- teractions were incorporated using the B97-D empirical dispersion correction functional [76]. In order to achieve an energy convergence of at least 1 meV/atom and force convergence of at least 1:3  10 eV/a , we found it sucient to use a common plane-wave energy cuto of 1100 eV with `cold' smearing [77] of 10 K for all elements. To achieve the same convergence, the Brillouin zone sampling for bulk systems was 15 15 15, and 15 15 1 for monolayers and bilayers. Uniaxial and shear strains between 5% were applied in incre- ments of 1% to the unit cell with internal relaxation subject to the same force convergence criterion as above. Electronic band structures were calculated along the high-symmetry 4 points of the Brillouin zone f, X , S, Y ,Zg (Fig. 1b) for each value of in-plane strain (Figs. S1 - S9 in the Supplemental Material). For shear strains the Brillouin zone deforms into an asymmetric honeycomb, yet we continued to sample along the original path since the deformation up to 5% strain is negligible and the e ective masses are all calculated at the - point. We determine the Kohn-Sham band gap from the band structures and charge-carrier ? 2 2 e ective masses according to the nearly-free electron model m = ~ (@ E=@k @k ) using i j ij a cubic spline t to 9 data points about the -point. The charge velocities were similarly determined according to the dispersion relation v = ~ dE(k)=dk from the linear t to the Dirac or Weyl lines. The elements of the sti ness matrix C were derived from the gradients of the resultant stress-strain pro les c = @ =@" , from which all elastic properties were ij i j derived. In practice, however, the computed sti ness tensors are not exactly symmetric due to numerical noise but we make them so by taking the average of C and its transpose C as the e ective sti ness tensor. We begin our discussion with a brief overview of the the Voigt-Reuss-Hill scheme, which is a popular model used for computing e ective isotropic elastic properties. B. The Voigt-Reuss-Hill scheme In order to e ectively preserve, study and strain-engineer few-layer nano-structures, such as BP [78], graphene [79], or molybdenum disul de [80] (MoS ), the nano- akes are typically deposited onto a suitable substrate. The cumulative contribution of dispersed nano- akes distributed on or within a bulk medium results in the macroscopic elastic properties that are measured by experiments. The theoretical calculation of these elastic properties requires an appropriate mixture model (such as the rule-of-mixtures [11] (ROM) or the Halpin-Tsai [81] models) that require the (typically averaged) elastic properties of the interstitial nano- akes. In the theory of e ective media, isotropic bulk properties are computed by averaging the sti ness tensor C over all possible rotated reference frames [82{84], as outlined in the Supplemental Material. The result is called the Voigt average [85, 86] and it gives isotropic averages for the bulk Young's modulus Y , and shear modulus G , given in Eq. S7. The same V V scheme applied to the compliance tensor S results in the corresponding Reuss averages [87], Y and G , given in Eq. S9. The Voigt scheme assumes that the material undergoes constant R R strain and it returns over-estimated elastic constants. Conversely, the Reuss scheme assumes 5 constant stress and it tends to under-estimate the elastic constants. The Hill averages [88] Y + Y G + G V R V R Y = ; and G = ; (1) H H 2 2 from which the isotropic Poisson's ratio  and bulk modulus B are expressed as H H Y Y G H H H = 1; and B = : (2) H H 2G 3(3G Y ) H H H These are widely considered as reliable estimates of the actual physical values [84]. The Voigt-Reuss-Hill approach described above is used in the present work to determine the isotropic averages of the Young's, shear and bulk moduli, and the Poisson ratio of the bulk P, As, and Sb structures using the elements of the elastic tensors. Let us now discuss how the above approach may be adapted to derive the relevant equations for the speci c case of two-dimensional materials. C. In-plane Voigt-Reuss-Hill average If the interstitial nano- akes in a bulk medium form high-quality planar sediments [11, 89{ 95], the random orientation occurs instead in the plane of the akes and we must calculate isotropic-averages in-plane. Due to the weak vdW bonds between layers, strains related to out-of-plane directions can be ignored resulting in the reduced-sti ness tensor (Eq. S3). In the Supplemental Material, we re-derive the angular dependence of the rotated tensor- elements C () and S () about the ~z-axis (Eq. S6) as a function of the elements in the ij ij original reference frame, similar to the general Voigt-Reuss scheme. The angular-dependence of the in-plane elastic constants are then expressed as 2 2 C C C 11 12 Y () = ; G () = C ;  () = ; V V 66 V C C 11 11 (3) 1 1 S Y () = ; G () = ;  () = ; R R R S S S 11 66 11 with the Hill-average taken as in Eq. 1. By integrating the elastic tensors C () and S () ij ij over 2, the in-plane averages are then computed analogously. 6 III. RESULTS A. Lattice Constants The lattice constants a, b, c of the fully-relaxed structures are presented in Table I, where, in the monolayer and bilayer cases, we quote the layer thickness c instead of the unit cell height c. Our computed lattice parameters compare well with other recent theoretically predicted values [28, 29, 51, 59, 96{99] and the available experimental data [100, 101]. For a a (A) b (A) c (A) P 4.57 3.31 2.11 mono P 4.51 3.31 7.34 bi a a a P 4.43 (4.37 ) 3.32 (3.31 ) 10.47 (10.47 ) bulk As 4.70 3.67 2.39 mono As 4.64 3.69 7.86 bi b b b As 4.56 (4.47 ) 3.71 (3.65 ) 10.94 (11.0 ) bulk Sb 5.02 4.23 2.79 mono Sb 4.88 4.26 8.83 bi c c Sb 4.73 4.29 (4.3 ) 12.09 (11.2 ) bulk Ref. [100] Ref. [101] Ref. [102] TABLE I: Lattice parameters (A) for monolayer, bilayer and bulk structures of P, As, Sb compared to experimental data [100{102] quoted in parentheses. For the monolayers and bilayers the layers thickness c is given. given element, we nd that the lattice parameter along the puckered direction, `a', shortens as the number of layers increases, which agrees with observations in other studies. This is attributed to the increased vdW forces between layers, which leads to increased binding primarily in the softer puckered direction. 7 B. Electronic properties All of the band structures pertaining to the following analysis are presented in Figs. S1 - S9 of the Supplemental Material. Where we identify possible Dirac or Weyl states, high resolution, three-dimensional band structures with SOC at representative strains are recal- culated. To con rm the existence of linear-dispersion, we also plot lines along the surface of the Dirac and Weyl points at 0 , 30 , 60 and 90 with respect to the X line. A representative sample of these results are presented in Figs. 5a - 5g, while the rest can be found in Figs. S10 - S13 of the Supplemental Material. In general we nd the band gap to be very responsive to uniaxial in-plane strain but signi cantly less so with respect to shear strain. We identify several direct-indirect band gap transitions as well as the opening and closing of band gaps, summarized in Table III. We nd the charge-carrier e ective masses vary approximately linearly with respect to the uniaxial strain in general but with notable exceptions that will be discussed. This section is divided into three parts discussing each of the species - P, As and Sb - for which we review the qualitative calculation results including band gap transitions, e ective mass behavior and linearly-dispersive bands. 1. Phosphorus As shown in Fig. 2a, our calculations reproduce the direct band gap of 0.88 eV at the - point in the relaxed P monolayer, which falls within range of the reported gap between 0.7 eV (DFT-PBEsol [12]) and 1.0 eV (DFT-HSE06 [8]). On the other hand, quasi-particle calcu- lations predict a larger 2 eV band gap [22] with signi cant exciton binding [103] (between 0.4-0.83 eV). However, it is well understood that approximate semi-local functionals such as PBE su er from a systematic band gap problem [104] that may also adversely a ect the metal-insulator critical strains. Nevertheless, it is important to emphasize that band align- ments and rates of change are quite often reliably reproduced [105], as are the direct-indirect transitions [13] in two-dimensional materials. While absolute band gaps are therefore not expected to be exactly reproduced, we can expect reasonable agreement with trends in elec- tronic and mechanical behavior [106, 107]. The application of uniaxial in-plane strain is found to open the band gap for tensile strain and diminish it for compressive one, while 8 shear strain has a negligible e ect. The electron and hole e ective masses (Figs. 2b & 2c), compare well to the gures computed in Ref. [108], where, at " = +5% tensile strain, xx the electron and hole e ective masses coincide at 0.9 m as higher energy bands fall below the conduction band. For compressive strains, the hole e ective mass along X rises signi cantly as the valence band attens. Bilayer P is also found to have a direct band gap of 0.43 eV (Fig. 2d) in the relaxed state, and broadly the same behavior as the monolayer, in which case the band gap closes at around 5% uniaxial compressive strain with a predicted Dirac state at the -point. The e ect of SOC on the band structure (Fig. 5a) induces no qualitative di erence and the three-dimensional bands plotted about the Dirac points (Fig. 5b) con rm the linear- dispersion, albeit in only one direction. A linear t to the surface of the bands (Fig. 5c) 6 1 returns a maximum charge velocity of v = 3:80(1)  10 ms along Y , while, in the orthogonal direction, the bands are at with a a charge velocity that is relatively negligible. This high anisotropy in charge velocities, dominated by ballistic conduction along Y , is indicative of e ective one-dimensional conductivity and is further supported by the large disparity in e ective masses at " = 5%, evident in Figs. 2e & 2f. The same analysis for xx " = 5%, for which the Dirac states are due to band inversion and consequently occur o yy the Y symmetry line at a point X , can be found in Figs. 5a - 5c. Here the maximum 6 1 charge velocity is v = 3:22(1) 10 ms . These results are further supported by the work of Doh et al. [54], who demonstrated the e ect of strain on hopping parameters can lead to a Dirac semi-metallic state in bilayer P. Similarly, Baik et al. [71] found that the SOC did not induce a band gap in potassium-doped multi-layer P, but did lift the spin-degeneracy of the Dirac points. A direct-indirect band gap transition is also observed at +2% uniaxial tensile strain. The e ective masses (Figs. 2e & 2f), also exhibit broadly the same behavior as the monolayer, due to band- attening at X and the falling conduction bands along Y , which lead to the charge carrier e ective masses along X converging at +4% strain and an increasing hole e ective mass for compressive strains. In the bulk, however, we nd that the band gap is completely closed (Fig. 2g), i.e. that the material is metallic. After investigation, we concluded that this was an e ect of the smearing functionality [77] in the relaxation procedure and that it contradicts numerous experiments [109{112] that have measured a direct gap in the range of 0.31-0.36 eV. When relaxed under xed-occupancy conditions, instead, a band gap of 0.35 eV was produced. 9 While the PBE gap remains closed in the relaxed state, under uniaxial tensile strain it brie y becomes a single-point semi-metal at +2%. At such strains a possible Weyl state is observed, before a direct gap opens that subsequently transitions to an indirect gap at +3%. Shown in Fig. S10 in the Supplemental Material is the three-dimensional band structure with SOC in which a pair of potential Weyl points occur on an o -symmetry point X along X . Here, the SOC slightly reduces the band gap by  0:05 eV and does not qualitatively a ect the 6 1 overall results. The maximum charge velocity is v = 2:40(1) 10 ms for both " = 5% xx and " = 5% and occurs along a line parallel to Y . Under greater compression yy this band-inversion may also lead to further Weyl states, which have been experimentally observed at similar pressures [60{62]. Again, shear strain is seen to have a negligible e ect on the gap. The electron e ective masses are quite responsive to strain (Fig. 2h), where those along Y rise for both tensile strain along " , due to falling conduction bands, xx and compressive strain along " due to attening bands along Y . The e ective masses yy along X were necessarily not computed once the bands overlapped below +2% strain and the hole e ective masses are found to vary with respect to the strain to a slightly lesser extent (Fig. 2i). To summarize, we predict the onset of -point Dirac states in bilayer P at -5% uniaxial compressive strain, with e ective one-dimensional conductivity at " = 5%, and a direct- xx indirect band gap transition at +2% tensile strain. We also predict the existence of a possible Weyl states at +2% tensile strain in bulk P, followed by a direct-indirect band gap transition at +3%. Finally, e ective masses are found to be particularly responsive to " xx uniaxial strain. 2. Arsenic In contrast to P, we identify an indirect band gap of 0.15 eV along the Y direction in the relaxed As monolayer (Fig. 3a), which is signi cantly lower than the predicted DFT- HSE06 gap [98] of 0.83 eV. However, the relaxed band structure and band gap pro les closely resemble those in Refs. [28, 49]. The band gap diminishes for tensile strain along " xx and at +2% the material becomes semi-metallic with a Dirac state at the -point emerging at " = +5% accompanied by an electron pocket above the Fermi-level (Fig. S11 in the xx Supplemental Material), which is una ected by the SOC. The maximum charge velocity here 10 1.6 ■ 15 ■■ Γ-X Γ-X Γ-X ■ E ■ E ■ E ■■ xx yy xy ■ ■ ■ ■ ■ ■■■ ■■■ ■ ■ ■■ ■ ■■ ■■ m m m ■■ ■■ ■ ■ ■ ■■ xx yy xy 1.2 1.4 * * * E E E ▲ ▲ ▲ Γ-Y Γ-Y Γ-Y xx yy xy ▲ Γ-X Γ-X Γ-X 10 ▲ m ▲ m ▲ m 1.2 m m m xx yy xy ■ ■ ■ xx yy xy ■▲ ■▲ 0.8 ■▲ ■▲ ■▲ ■▲ ■▲ 1.0 ■▲ Γ-Y Γ-Y Γ-Y ■■▲▲ m m m ▲ ▲ ▲ ■ ■▲ ▲■ ▲■■■▲▲ ▲■ ▲■ xx yy xy ■▲ ■▲ ■▲ ■▲ ■▲ ■▲ ▲ 5 ■■▲▲ ■ 0.4 0.8 ■ ■ ■ ■ ■▲▲ ■ ■ ■ ■ ■ ■ ■ ■■■ ■■ ■ ■ ■ ■ ■ ■ ■■▲ ■ ■ ■ ▲ ■ ■ ▲ ■ ■ ■■▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ 0.6 ■▲ ▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ ▲ -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (a) Monolayer P band gap (eV) (b) Monolayer P m =m (c) Monolayer P m =m e 0 h 0 1.2 3.0 3.0 Γ-X Γ-X Γ-X Γ-X Γ-X Γ-X ■ E ■ E ■ E xx yy xy m m m m m m ■ ■ ■ ■ ■ ■ xx yy xy xx yy xy 1.0 2.5 2.5 * * * E E E Γ-Y Γ-Y Γ-Y ■ Γ-Y Γ-Y Γ-Y ▲ ▲ ▲ xx yy xy ▲ m ▲ m ▲ m ▲ m ▲ m ▲ m 0.8 2.0 ■ 2.0 xx yy xy xx yy xy ■ ■ ■ ▲ ■■ ■ ▲ ■▲▲ ■ ■ 0.6 ■ ▲ 1.5 ■ ■ ■ ■ ■ ■■ ■ ■ ■ ■ 1.5 ■ ■ ▲ ■■ ■■■ ■ ▲ ■■ ■ ■▲ ■■ ■■ ■ ■ ■■▲▲ ■■ ■ ■ ■ ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ ■ ■ ■▲ ■▲ ■▲ ■▲ ■▲ ■■■▲▲▲ ■▲ ■▲ ■▲ ■▲ ■▲ ■ ■ ■ ■ 0.4 1.0 ▲ ▲ ▲ 1.0 ■ ■■▲▲ ■ ■■▲▲ ■■▲▲ 0.2 0.5 ▲ 0.5 ■▲ ■▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲ ▲ ▲ ▲▲ ▲▲ ■▲ ▲ ▲ ▲ ▲ ▲▲ ▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲▲ ▲▲ ▲▲ ▲▲ ▲ ▲ ▲ ▲ ■▲ 0.0 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (d) Bilayer P band gap (eV) (e) Bilayer P m =m (f ) Bilayer P m =m e 0 h 0 3.0 1.2 0.3 ■ Γ-X Γ-X Γ-X Γ-X Γ-X Γ-X ■ m ■ m ■ m m m m xx yy xy ■ ■ ■ xx yy xy 2.5 1.0 * ▲ E E ■ xx ▲ ■ xx Γ-Y Γ-Y Γ-Y Γ-Y 0.2 2.0 m m 0.8 ■ ■ ▲ ▲ m m ▲ xx yy ▲ ▲ ■ xx yy ■ ■ ■ ■ E ■ yy ▲ E ■ yy 1.5 ■ 0.6 ■ ■■ ■■ * ■▲ 0.1 ■ xy ▲ E 1.0 0.4 ▲ xy ■ ■ ▲ ■ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ■ ■ ■ ■▲ ▲ ▲ ▲ ▲ ▲ ■▲▲ ■ ■▲ ■ ■▲ ■ ■ 0.5 ▲ ■ 0.2 ▲ ■ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ■▲ ■▲▲ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ▲ ■▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■▲ ■▲ ■▲ ■▲ 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (g) Bulk P band gap (eV) (h) Bulk P m =m (i) Bulk P m =m e 0 h 0 FIG. 2: (Color online) The relationships between the applied in-plane strains " (blue), xx " (red) and " (green) against [(a), (d), (g)] the direct E (solid squares) and indirect E yy xy (dashed triangles) band gaps (eV) ; [(b), (e), (h)] the e ective electron masses m =m along e 0 X (solid squares) and Y (dashed triangles) ; [(c), (f ), (i)] the e ective hole masses m =m along X (solid squares) and Y (dashed triangles); for each phase of P. h 0 6 1 is v = 3:01(1) 10 ms and lies along Y . In the orthogonal direction the bands are at, similarly to monolayer P, with a relatively small charge velocity. This high anisotropy in charge velocities, dominated by the ballistic conduction along Y , is again indicative of e ective one-dimensional conductivity and is further supported by the large disparity in e ective masses at " = 5%, shown in Figs. 3b & 3c. xx For compressive strain along " the indirect band gap closes along Y at " = 2%. xx xx Bandgap (eV) Bandgap (eV) Bandgap (eV) m /m m /m m /m e 0 e 0 e 0 m /m m /m m /m h 0 h 0 h 0 For tensile strain along " the indirect band gap opens where an indirect-direct transi- yy tion [28] occurs at " = 3%. Similar to monolayer phosphorus, there is no appreciable xx e ect due to shear-strain. Meanwhile, the charge-carrier e ective masses (Figs. 3b & 3c) re- spond linearly to uniaxial strain and compare well to other works [101], where, in particular, valence band broadening along X leads to an increasing hole e ective mass. For bilayer As, we identify a direct band gap of 0:45 eV (Fig. 3d), in contrast to the indirect band gap observed in the monolayer. Here, the band gap opens for uniaxial tensile strain and diminishes for compressive strain. The direct band gap transitions to an indirect gap at both " = 3% and " = +2%, while at " = +2% it also transitions to an indirect yy yy xx gap before resuming to a direct gap again at " = +3%. Moreover, we predict a Dirac state xx at the -point at a compressive strain of " = 4% (Fig. S11 in the Supplemental Material) xx 6 1 for which the maximum charge velocity is v = 2:62(2)10 ms along Y . Along X the bands are also at, similar to the monolayer, and have a relatively negligible charge velocity. The high anisotropy in charge velocities, is again indicative of e ective one-dimensional conductivity, dominated by the ballistic conduction along Y , and is further supported by the large disparity in e ective masses at " = 5%, shown in Figs. 3e & 3f. Here again, xx the SOC has no appreciable e ect on the bands. The electron and hole e ective masses (Figs. 3e & 3f) respond approximately linearly to the applied strain, where conduction band broadening leads to increased e ective electron masses, and valence band attening at leads to increasing hole e ective masses for strain along " . yy Finally, no band gap is determined in the relaxed bulk phase (Fig. 3g), again contrary to experiments [113], where a small direct band gap of  0:3 eV is observed. However, at " = +1% strain, a potential Weyl state is brie y observed on an o -symmetry point xx X (Fig. S12 in the Supplemental Material) before a direct gap opens that subsequently transitions to an indirect one at " = +3%, after which it reduces again. Another potential xx Weyl state around the same o -symmetry point X is also predicted to occur between +1% "  +2% after which a direct band gap also appears. The recalculated band structure with yy the SOC for " = +1% (Fig. S12) con rms the linear-dispersion. The maximum charge yy 6 1 velocity in both cases occurs along a line parallel to Y and is v = 1:38(1) 10 ms . Meanwhile, the electron and hole e ective masses along Y (Figs. 3h & 3i) increase rapidly for compressive strains as the band peaks rapidly atten at the -point. In summary, we predict -point Dirac states in the monolayer and bilayer of As, which 12 1.5 1.5 1.0 ■ E E E ■ xx ■ yy ■ xy ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■■ ■■■ ■■ ■■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ 0.8 ■ ■ ■ ■■ ■■■ ■ ■ ■ ■ ■ ■ ■ * * * ■ ■ E E E Γ-X Γ-X Γ-X ■ ▲ ▲ ▲ 1.0 1.0 xx yy xy ■ ■ m ■ m ■ m ■ ■ xx yy xy ■ ▲ Γ-X Γ-X Γ-X 0.6 ■ m ■ m ■ m xx yy xy Γ-Y Γ-Y Γ-Y ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ ■ ▲ ■ ■ ▲ m ▲ m ▲ m xx yy xy 0.4 ■ 0.5 0.5 Γ-Y Γ-Y Γ-Y ■▲ ▲ m ▲ m ▲ m ■ ▲ ■ ▲ ▲ xx yy xy ■▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲▲▲ ▲ ■ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.2 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲▲ ■ ▲ ▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ▲ 0.0 0.0 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (a) Monolayer As band gap (eV) (b) Monolayer As m =m (c) Monolayer As m =m e 0 h 0 1.2 ■ E ■ E ■ E 1.4 ■ ■ xx yy xy 1.2 ■ ■ ■■ ■ 1.0 ■■ 1.2 ■ ■ ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ * * * ■ ■■ ■ ■ ■ ■ ▲ E ▲ E ▲ E ■ 1.0 ■ ■ ■■ ■ ■ ■ ■ xx yy xy ■ ■ ■ ■ ■ ■ ■■■ ■ ■ ■ ■ ■ ■ ■ ■ 0.8 1.0 Γ-X Γ-X Γ-X ■ ■ ▲ m m m 0.8 ■ ▲ ■ ■ ■ ■ ▲ xx yy xy ■ ■▲ 0.6 ■▲ ▲ 0.8 Γ-X Γ-X Γ-X ■▲ ■ ▲ ■ m ■ m ■ m xx yy xy ■▲ ■▲ ■▲ ■▲ ■▲■■▲▲ ■▲ ■▲ ■▲ ■▲ ■▲ ■▲ ■▲ 0.6 ■▲ Γ-Y Γ-Y Γ-Y 0.6 0.4 ■▲ ■▲ m m m ■ ▲ ▲ ▲ xx yy xy ■▲ Γ-Y Γ-Y Γ-Y 0.4 m m m 0.4 ▲ ▲ ▲ ▲ xx yy xy 0.2 ■ ■▲▲ ▲ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ■ ■▲ ▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.2 ▲ ▲▲ ▲▲ ▲▲ ▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ 0.2 ▲ ▲ ▲ ■▲▲ ▲ ▲ ▲ ▲ 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (d) Bilayer As band gap (eV) (e) Bilayer As m =m (f ) Bilayer As m =m e 0 h 0 2.5 0.30 3.0 Γ-X Γ-Y Γ-Y m m m Γ-X Γ-Y Γ-Y ■ ■ ▲ ▲ xx xx xy ■ m ▲ m ▲ m xx xx xy ▲ ■ 0.25 2.5 2.0 * Γ-X Γ-Y ■ Γ-X Γ-Y ■ E ▲ E ■ m ▲ m xx xx yy yy ■ m ▲ m 0.20 2.0 yy yy ■▲ 1.5 ▲ ■ ■▲ E ▲ ■ ▲ E ▲ yy 0.15 yy 1.5 ■ ■ ■ ■■ ■ ■▲ ■■ 1.0 ▲ ▲ ■ * ■ ■ 0.10 ■ E 1.0 ■ ■ ■■ ■ xy E ▲ ▲ ■ ▲ ■ xy ▲ ▲ ▲ ▲▲ ▲ 0.5 ▲ ▲▲ ▲ ■▲ ▲ ▲ ▲▲ ▲ ▲▲ ▲ ▲ ▲ 0.05 0.5 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲▲▲ ▲ ▲▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■■■▲▲▲ ■▲ ■▲ ■▲ ■▲ 0.00 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (g) Bulk As band gap (eV) (h) Bulk As m =m (i) Bulk As m =m e 0 h 0 FIG. 3: (Color online) The relationships between the applied in-plane strains " (blue), xx " (red) and " (green) against [(a), (d), (g)] the direct (solid squares) and indirect yy xy (dashed triangles) band gaps (eV) ; [(b), (e), (h)] the e ective electron masses m =m along e 0 X (solid squares) and Y (dashed triangles) ; [(c), (f ), (i)] the e ective hole masses m =m along X (solid squares) and Y (dashed triangles); for each phase of As. h 0 support one-dimensional ballistic conduction, as well as possible Weyl states on o -symmetry points in the bulk at moderate levels of in-plane stress that are una ected by the SOC. We also observe several band gap transitions, in particular in the monolayer phase, which also include semi-conducting-metallic transitions. Finally, the e ective masses respond approxi- mately linearly with respect to uniaxial strain, except in the bulk, which exhibits quadratic behavior. Bandgap (eV) Bandgap (eV) Bandgap (eV) m /m m /m m /m h 0 e 0 e 0 m /m m /m m /m h 0 h 0 h 0 3. Antimony The relaxed Sb monolayer is found to possess an indirect band gap of 0:21 eV along Y (Fig. 4a), which is reasonably comparable to other PBE values 0.28 [29]-0.37 [114] eV, although these have been obtained by including SOC. For tensile strain along " the band yy gap opens, suggesting an indirect-direct transition for strains above 67%, and it diminishes for compressive strains before nally closing at " = 2%, where the material becomes yy a semi-metal. Similarly, the indirect gap closes along X at a compressive strain of " = 2% at which monolayer Sb again becomes semi-metallic. The indirect gap transitions xx to a direct gap at " = +1% tensile strain and remains so until nally closing at " = +4%, xx xx at which point we predict a potential Dirac state along Y at an o -symmetry point [37] Y (Fig. 5g) that has also been predicted in Ref. [51]. Fig. 5g depicts the calculated band structure, in which it is shown that SOC preserves the Dirac state but not does not open the band gap. The three-dimensional band structure about the Dirac point is shown in Fig. 5h in 6 1 which the maximum charge velocity is v = 4:31(1)10 ms and occurs along a line parallel to the Y direction (Fig. 5i). Moreover, the valence band at the X -point undergoes a Rashba splitting [115] due to SOC, which is also predicted to occur in the monolayers of -P [116], and -Sb [52, 117]. Finally, the electron e ective masses experience a weak linear response to strain (Figs. 4b & 4c), while the hole e ective masses along X respond much more strongly to a rapid broadening or attening of the valence band. Furthermore, the relaxed bilayer phase is found to be semi-metallic where an indirect band gap opens at " = +3% tensile strain and for uniaxial strains < 1% band-inversion yy at the -point leads to to a fully-metallic state. In addition, a possible Dirac state emerges at a similar non-symmetry-point Y along Y for " = +2% tensile strain (Fig. S13 in xx the Supplemental Material) and remains in place up to at least +5% strain. The maximum 6 1 charge velocity v = 4:47(3) 10 ms is also along Y and is approximately the same as that of the monolayer. The e ective masses (Figs. 4e & 4f) experience mild linear- response to strains prior to the transition to full metallicity, at which point a rapid attening of the bands at the -point suggesting strong electron localization Beyond a compressive strain of " = 3%, however, at a stress of 0.3 GPa, the bilayer undergoes a structural yy transition and buckles in the puckered (~y) direction. This buckled structure has a total energy 1.7 meV/atom lower than that of the relaxed state of the unperturbed -bilayer and 14 1.0 1.4 1.2 E E E ■■ ■ ■ xx ■ yy ■ xy ■ ■■ ■■ ■ ■ ■ ■ ■ ■ ■■■ ■■■ ■■ ■ ■ ■ ■■ ■■ ■■ ■■ ■■ ■ ■ 1.2 ■ ■ 0.8 1.0 ■ ■ ■ * * * ■ ■ ■■■ ■ ■ E E E ■ ■ ■ ■ ■ ■ ▲ ▲ ▲ Γ-X Γ-X Γ-Y ■ xx yy xy 1.0 ■ ■ m ■ m ▲ m ■ ■ 0.8 xx xy yy ■ ■ ■ 0.6 ■ ■ ■ ▲ 0.8 Γ-X Γ-X Γ-X Γ-Y ■ ▲ 0.6 Γ-Y Γ-Y ■ m m m m ■ ▲ m ▲ m ■ ■ ▲ ■ ▲ yy xx xy yy ■ xx xy ■ 0.6 0.4 ■ ■ ■ ■■■ ■ ■▲ ■ ■ ■ ■ ■ 0.4 ▲ ▲ ▲ ▲ ▲ ■ ■ ▲ ■ ■▲ ▲ ▲ Γ-X ▲ ▲ ▲ ▲ Γ-Y Γ-Y ■ ▲ ▲ ▲ ▲ 0.4 m ▲ ▲ ▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ■ m m ■ ▲ ▲ ▲ ▲ ▲ yy ▲ ▲ ■ ▲ ▲▲▲ ▲ ▲ xx xy ▲ ▲ ■▲▲ ▲ 0.2 ▲ ▲ 0.2 ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ 0.2 ▲ ▲ ▲ ▲▲ ▲▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■▲ ▲▲ ▲▲ ▲▲ ▲▲ ■▲ 0.0 0.0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (a) Monolayer Sb band gap (eV) (b) Monolayer Sb m =m (c) Monolayer Sb m =m e 0 h 0 3.0 0.4 Γ-X Γ-X Γ-X m m m ■ ■ ■ Γ-X Γ-X Γ-Y xx yy xy ■ m ■ m ▲ m ■ xx xy yy 2.5 0.3 E E ■ Γ-Y Γ-Y Γ-Y ■ ▲ Γ-X xx xx m m m Γ-Y Γ-Y ▲ ▲ ▲ m xx yy xy 2.0 ■ m m yy ▲ ▲ xx xy ■ yy ▲ E 2 0.2 yy ■ 1.5 ■ ■ ■ * ■ ■ ■ ▲ ■ E ■ ■■ ■ ■ ■ ■■ ■■■ ■■ ■■ ■■ ■ ■ xy ▲ E ▲ ■■ ■■ ■ 1.0 ■ ■ ■ ■ ■ ■ ■ ■ ■■ ■■■ ■■■ ■ ■ ■ ■ ■ ■ xy 0.1 1 ▲ ▲ ▲ ▲ ▲ 0.5 ▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲▲ ▲▲ ▲▲▲ ▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲▲ ▲▲ ▲▲▲ ▲ ▲ ▲ ▲ ▲ ▲ ▲ ■■▲▲ ■■▲▲ ■■▲▲■▲ ■■▲▲■▲ ■■▲▲■▲ ■▲■■▲▲ ■▲■■▲▲ ■■▲▲■▲ ■■▲▲ ■■▲▲ ■■▲▲ 0.0 0 0.0 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 Strain (%) Strain (%) Strain (%) (d) Bilayer Sb band gap (eV) (e) Bilayer Sb m =m (f ) Bilayer Sb m =m e 0 h 0 FIG. 4: (Color online) The relationships between the applied in-plane strains " (blue), xx " (red) and " (green) against [(a), (d), (g)] the direct (solid squares) and indirect yy xy (dashed triangles) band gaps (eV) ; [(b), (e), (h)] the e ective electron masses m =m e 0 along X (solid squares) and Y (dashed triangles) ; [(c), (f ), (i)] the e ective hole masses m =m along X (solid squares) and Y (dashed triangles); for each h 0 monolayer and bilayer Sb. 3.0 meV/atom lower when allowed to fully-relax, as shown above in Fig. 1c. This suggests the possible existence of a new structure that is attainable via strain. Finally, bulk Sb is found to be completely metallic for all levels of strain explored in this work. However, shear strains in this case do appear to have a signi cant e ect on the bands despite not opening a gap. In summary, we predict possible non-symmetry-point Dirac states in the strained monolayer and bilayer of Sb, which are qualitatively una ected by SOC, as well as Rashba splitting at the X -point in the monolayer. We also predict indirect-direct and indirect-semi-metallic transitions in the monolayer phase and a band gap opening in the bilayer phase. Finally, we observe a buckled state induced in bilayer Sb at 4% compressive strain. Bulk Sb was found to be metallic at all levels of strain explored. In Table II below we present a summary of the calculated band gaps and e ective charge Bandgap (eV) Bandgap (eV) m /m m /m e 0 e 0 m /m m /m h 0 h 0 carrier masses for the relaxed phases of each structure, and in Table III we provide a synopsis of the band gap and phase transitions of interest. Finally, in Fig. 5, we present the band structures of bilayer P and monolayer Sb, that form a representative sample of the di erent 0 0 Dirac points predicted at , X and Y , as well the three-dimensional band structures about the region of the points. E (eV) Me Me Mh Mh g X Y X Y P 0.9 1.25(1) 0.16(1) 2.8(2) 0.14(1) mono As 0:15 1.16(1) 0.26(1) 1.09(1) 0.18(2) mono Sb 0:2 1.10(1) 0.28(1) 1.04(1) 0.19(2) mono P 0.4 1.41(1) 0.19(1) 1.21(3) 0.15(2) bi As 0.45 1.15(1) 0.24(1) 0.94(4) 0.17(2) bi Sb 0 1.16(1) 0.39(1) 0.99(4) 0.33(4) bi P 0 - 0.37(2) - 0.21(2) bulk As 0 - 0.36(1) - 0.30(3) bulk Sb 0 - - - - bulk TABLE II: (Color online) Kohn-Sham band gaps (eV), indicating the indirect semiconducting (?), semi-metallic (y) and metallic (z) states, as well as the charge-carrier e ective masses (m ) for each phase of P, As, Sb. C. Isotropic bulk properties In order to obtain these electronic states, and to ensure accurate strain-engineering, knowledge of the mechanical properties is paramount. Therefore, in this section we review the mechanical response of the few-layer and bulk phases in order to compute the elastic properties, both isotropically averaged and as a function of orientation of applied in-plane stress. The computed elements of the sti ness tensor C in GPa of each structure are given in Table S1 in the Supplemental Material, where those pertaining to bulk P compare well to experiments [118, 119] and similarly computed values [96, 108]. For the elements related to in-plane strains (c , c , c , c ) we observe an expected increase in sti ness as the layer 11 22 66 12 16 0.2 0° 30° 60° 90° 0.1 0.0 -1 -0.1 -2 -0.2 Γ X S Γ Y S -0.04 -0.02 Γ 0.02 0.04 (a) Bilayer P band structure " = 5% (b) (c) -point Dirac state xx 0.2 0° 30° 60° 90° 0.1 0.0 -1 -0.1 X' -2 Γ X S Γ Y S -0.2 -0.04 -0.02 X' 0.02 0.04 (d) Bilayer P band structure " = 5% (e) (f ) X -point Dirac state yy 0.15 0.10 0° 0.05 30° 60° 0.00 90° -1 -0.05 -2 Γ X S Γ Y' Y S -0.10 -0.02 -0.01 Y' 0.01 0.02 (g) Monolayer Sb band structure " = 3% (h) (i) Y -point Dirac state xx FIG. 5: (Color online) a) Band structure of bilayer P at " = 5% with SOC (thick xx lines), and without SOC (thin lines) b) three-dimensional bands about the predicted Dirac point at c) slices through the Dirac point at 0 , 30 , 60 and 90 relative to the X line that indicate highly anisotropic conduction. [5d,5e,5f] Illustrates the same for bilayer P at " = 5%, where the Dirac state occurs at the non-symmetry point X along X . yy [5g,5h,5i] Illustrates the same for monolayer Sb at " = 3%, where the possible Dirac state xx occurs at the non-symmetry point Y along Y . Energy (eV) Energy (eV) Energy (eV) Energy (eV) Energy (eV) Energy (eV) Transition Direction Strain (%) P D Gap ! SM XX,YY -5 bi D Gap ! ID Gap XX,YY +2 P SM ! D Gap ! ID Gap XX,YY +1!+3 bulk SM!SM XX,YY +2 As ID ! D Gap XX -3 mono ID Gap ! SM XX +2 ID Gap ! SM XX +2 SM ! SM XX +5 As D Gap ! ID Gap ! D Gap XX +2!+3 bi D Gap ! ID Gap YY -3,+2 As SM ! D Gap ! ID Gap XX 0!+3 bulk SM ! D Gap YY 0!+3 SM!SM XX,YY +1 Sb ID Gap ! SM XX -2 mono ID Gap ! D Gap ! SM XX +2!+4 ID Gap ! SM YY -2 Sb SM ! SM XX +4 bi Structural Transition YY -3 SM ! ID Gap YY +3 TABLE III: Summary of the band gap transitions including direct (D); indirect (ID); metallic (M); and semi-metallic (SM), in particular those that indicate potential Dirac states (4), Weyl states (O), and the structural phase transition. number increases and for increasing atomic number. However, for other elements relating to out-of-plane and shear stresses (c , c , c , c , c ) the sti ness actually increases in the 33 44 55 23 13 bulk phase from P to As to Sb. The Hill-averaged bulk properties are presented in Table IV, which compare well to other DFT values [96, 108], while our calculated bulk modulus for bulk P (37.2 GPa) is also within reasonable range of the experimental values (32.32 [100] - 36.02 GPa [120]). We also observe 18 Y (GPa) G (GPa)  B (GPa) H H H H a a a a b c P 61.1 (70.3 ) 24.9 (29.4 ) 0.23 (0.30 ) 37.2 (38.5 , 32.32 , 36.02 ) bulk As 60.0 23.8 0.26 41.4 bulk Sb 52.5 21.1 0.24 33.8 bulk Ref. [96] Ref. [100] Ref. [120] TABLE IV: The Hill-averaged Young's modulus Y , shear modulus G and bulk modulus H H B in GPa, and Poisson's ratio  for bulk P, As and Sb compared to similarly calculated DFT [96] values and available experimental data [100, 120]. that the bulk properties remain largely comparable for all the species, but generally decrease from P to As to Sb (except for the Poisson's ratio and bulk modulus, which are largest for As). We also note that while bulk P has the largest in-plane responses, As and Sb have larger out-of-plane and shear responses, which enable the net isotropic properties for all the species to remain comparable overall. D. In-plane elastic properties In section S1 of the Supplemental Material we re-derive the equations for the elastic properties as a function of the in-plane orientation angle , de ned in Fig. 1, as outlined in Ref. [121]. These functions are the plotted in Fig. 6 and include the Young's modulus Y () and it's average hY ()i, the shear modulus G (), and Poisson's ratio  (). The experimental Young's modulus of 130 GPa, determined in Ref. [11] via the ROM of a nano- ake-polymer composite, lies precisely between the average in-plane bulk Young's modulus 85.7 GPa and the elastic sti ness in the zigzag (~x) direction 187.9 GPa, thus tting the results of our model reasonably well. The anisotropy of the elastic properties is also apparent in the mechanical pro les, partic- ularly with regard to the Young's modulus, which has a 2-fold symmetry about the x-axis, in contrast to the shear modulus and Poisson's ratios, which display 4-fold symmetry about both the axes (except for the Poisson's ratio of P, which remains 2-fold symmetric). Indeed, 19 π π 2 2 2 3π π 3π π 3π π 4 4 4 4 4 4 π 0 π 0 π 0 0. 20. 40. 60. 0. 10. 20. 30. 40. 0. 10. 20. 30. 40. 5π 7π 5π 7π 5π 7π 4 4 4 4 4 4 3π 3π 3π 2 2 Y(θ) G(θ) 100×ν(θ) ⟨Y(θ)⟩ Y (θ) G (θ) 50×ν (θ) ⟨Y (θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ xy xy xy xy (a) Monolayer P (b) Monolayer As (c) Monolayer Sb π π π 2 2 2 3π π 3π π 3π π 4 4 4 4 4 4 π 0 π 0 π 0 0. 25. 50. 75. 0. 15. 30. 45. 0. 10. 20. 30. 5π 7π 5π 7π 5π 7π 4 4 4 4 4 4 3π 3π 3π 2 2 2 Y(θ) G(θ) 100×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ (d) Bilayer P (e) Bilayer As (f ) Bilayer Sb π π π 2 2 2 π π 3π 3π 3π π 4 4 4 4 4 4 π 0 π 0 π 0 0. 50. 100. 150. 0. 30. 60. 90. 0. 15. 30. 45. 60. 5π 7π 5π 7π 5π 7π 4 4 4 4 4 4 3π 3π 3π 2 2 Y(θ) G(θ) 100×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y(θ)⟩ Y(θ) G(θ) 50×ν(θ) ⟨Y⟩ (g) Bulk P (h) Bulk As (i) Bulk Sb FIG. 6: (Color online) In-plane functions for the Young's modulus Y () (blue) and its isotropic average hY ()i (red) in units of GPa; shear modulus G () in GPa (orange); and Poisson's ratio  () (green, scaled by 100 for P and 50 for As, Sb) for each of the structures. 20 Y  Y  hYi G  G  hGi     hi min max min max min max P 17.1 69 72.4 0 36.4 20.9 90 42.2 45 31.0 0.3 31 0.5 63 0.5 mono As -2.1 90 30.5 18 12.7 10.4 77 13.8 45 11.2 0.2 39 0.9 90 0.4 mono Sb 1.4 90 24.4 22 12.6 8.9 45 13.6 90 9.4 0.3 43 0.8 90 0.6 mono Pb 20.5 75 85.6 0 44.4 28.1 90 49.2 45 37.8 0.31 34 0.5 68 0.4 bi As 9.7 90 56.0 0 31.0 23.0 78 28.6 45 24.3 0.4 40 0.6 90 0.5 bi Sb 2.6 90 31.6 33 16.3 11.8 45 19.5 90 12.8 0.2 42 0.8 90 0.5 bi P 34.2 87 166.4 0 85.7 64.9 87 92.4 45 74.5 0.2 36 0.5 75 0.3 bulk As 10.9 90 105.1 20 55.0 44.4 73 51.1 45 66.7 0.3 40 0.7 90 0.4 bulk Sb 12.4 90 58.9 30 35.4 28.1 59 28.2 45 63.2 0.2 41 0.6 90 0.4 bulk TABLE V: Summary of the minima and maxima of the Hill-averaged in-plane Young's modulus Y () (GPa), shear modulus G () (GPa), and Poisson's ratio  () as well as the angle  with respect to the ~x-direction (zigzag) at which they occur in degrees, and their in-plane averages. for a given species, the general shape of each pro le is approximately preserved with respect to the number of layers, while the range of each property tends to increase. This discovery is advantageous in the strain-engineering of nano- akes since one may now forecast in advance the response of a material to in-plane strain, once the underlying pro le and number of layers are known. Another interesting feature is that the extrema of the elastic functions do not necessarily coincide with the coordinate-axes. For instance the Young's modulus maximum for mono- layer Sb occurs at 22 . Table V summarizes the global minima and maxima of each function and the angles at which they occur. While most of the function extrema occur expectedly at 0 , 45 or 90 , many are incident away from the coordinate-axes. This result lends further insight into the mechanical anisotropy of the orthorhombic group-V materials. The emergence of a negative Young's modulus of 2:1 GPa in monolayer As (Fig. 6b), at rst glance, may give cause for concern. It arises due to a negative Voigt estimate for the 2 2 Young's modulus at 90 where (c c )=c = 8:9 GPa (since c > c ), which is larger 22 12 22 22 12 in absolute magnitude than the Reuss estimate at 90 given by 1=s = 4:7 GPa and results 21 in a net negative Hill-average. In this instance, we surmise that either the assumptions of the Voigt model break down, or the Hill-method is not universally appropriate in arbitrary directions in-plane and a more robust averaging scheme must be employed. Nevertheless, the qualitative in-plane functions and their isotropic averages remain physically meaningful and in general can provide valuable physical insight. In contrast to the isotropic averages for the bulk properties in Table IV, which remained largely comparable for P, As and Sb, a much clearer trend across the species emerges once we have eliminated contributions from the out-of-plane and shear stresses. In this instance, P clearly possesses superior in-plane mechanical strength in both moduli, which decrease with the number of layers as expected As and Sb are largely similar in the monolayer, though less so in the bilayer and bulk phases, where they are stronger in the ~x-direction. In contrast, the Poisson's ratio tends to remain relatively stable aside from generally decreasing with increasing number of layers. In summary, there exists in the elastic properties a broad range of responses, pro le shapes and behavior that is re ective of the underlying anisotropic crystal structure, which also strongly depend on the number of layers with P typically the sti est and Sb the most exible. In general we nd the shapes of the in-plane response pro les to be conveniently consistent, which implies that, for a given number of layers, the in-plane elastic response of a nano- ake can be reliably estimated a priori. IV. CONCLUSION We have extensively explored the mechanical and electronic properties of P, As and Sb in their few-layer and bulk phases. We have identi ed several band gap transitions in almost all of the structures. The SOC tended to close the bands by  0:05 eV but did not alter any of our qualitative ndings. We also predict the existence of Dirac states in the strained phases of monolayer As and Sb, bilayer P, As and Sb as well as possible Weyl states in bulk P and As for moderate levels of strain. The linear-dispersion was observed along Y of each of the predicted Dirac or Weyl states, corresponding to the direction of softest mechanical response 6 1 in the puckered direction. The maximum charge velocity is calculated to be over 10 ms . In particular, for bilayer P and few-layer As we predict highly anisotropic conductivity dominated by ballistic transport along the puckered direction that is indicative of e ective, 22 one-dimensional conduction. We predict that an appropriate strain could yield these e ects in experiments. We also observe the existence of a notable buckled state of compressed bilayer Sb at 4% strain. Finally, the angular-resolved elastic properties as well as the stress-dependence of the Kohn-Shame band gaps and charge-carrier e ective masses revealed highly anisotropic behavior, spanning a broad range of values, and angular-dependent behavior that has be- come characteristic of these group-V layered structures. Moreover, the critical stresses at which these transitions occur are expected to be experimentally accessible and highly switch- able, paving the way for possible veri cation in the near future. Thus, the group-V layered materials are poised to become central to the next generation of electronic devices with po- tential novel applications in eld-e ect transistors; batteries; gas-sensors and opto-electronic devices. ACKNOWLEDGMENTS The authors would like to sincerely thank Damien Hanlon, Claudia Backes, Conor Boland, Jonathan Coleman, Beata Szydlowska, Gaozhong Wang, and Werner Blau for their helpful discussions relating to the work conducted for this Article. This work was enabled by Science Foundation Ireland (SFI) funded centre AMBER (SFI/12/RC/2278). All calculations were performed on the Kelvin cluster maintained by Trinity College Dublin Research IT and funded through grants from SFI. [1] M. Wu, H. Fu, L. Zhou, K. Yao, and X. C. Zeng, Nano Lett. 15, 3557 (2015). [2] J. Guan, Z. Zhu, and D. Tom anek, Phys. Rev. Lett. 113, 046804 (2014). [3] Z. Zhu and D. Tom anek, Phys. Rev. Lett. 112, 176802 (2014). [4] J. Guan, Z. Zhu, and D. Tomnek, ACS Nano 8, 12763 (2014). [5] L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen, and Y. Zhang, Nat. Nano 9, 372 (2014). [6] F. Xia, H. Wang, and Y. Jia, Nat. Commun. 5, 4458 EP (2014). [7] S. P. Koenig, R. A. Doganov, H. Schmidt, A. H. C. Neto, and B. Ozyilmaz, Appl. Phys. Lett. 104, 103106 (2014). 23 [8] H. Liu, A. T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tomnek, and P. D. Ye, ACS Nano 8, 4033 (2014). [9] X. Ling, H. Wang, S. Huang, F. Xia, and M. S. Dresselhaus, Proc. Natl. Acad. Sci. U.S.A. 112, 4523 (2015). [10] J. R. Brent, N. Savjani, E. A. Lewis, S. J. Haigh, D. J. Lewis, and P. O'Brien, Chem. Commun. 50, 13338 (2014). [11] D. Hanlon, C. Backes, E. Doherty, C. S. Cucinotta, N. C. Berner, C. Boland, K. Lee, A. Har- vey, P. Lynch, Z. Gholamvand, S. Zhang, K. Wang, G. Moynihan, A. Pokle, Q. M. Ramasse, N. McEvoy, W. J. Blau, J. Wang, G. Abellan, F. Hauke, A. Hirsch, S. Sanvito, D. D. O'Regan, G. S. Duesberg, V. Nicolosi, and J. N. Coleman, Nat. Commun. 6, 8563 (2015). [12] A. S. Rodin, A. Carvalho, and A. H. Castro Neto, Phys. Rev. Lett. 112, 176801 (2014). [13] X. Peng, Q. Wei, and A. Copple, Phys. Rev. B 90, 085402 (2014). [14] J.-W. Jiang and H. S. Park, Phys. Rev. B 91, 235118 (2015). [15] B. Sa, Y.-L. Li, J. Qi, R. Ahuja, and Z. Sun, J. Phys. Chem. C 118, 26560 (2014). [16] M. Elahi, K. Khaliji, S. M. Tabatabaei, M. Pourfath, and R. Asgari, Phys. Rev. B 91, 115412 (2015). [17] T. Hu, Y. Han, and J. Dong, Nanotechnology 25, 455703 (2014). [18] J.-W. Jiang and H. S. Park, Nat. Commun. 5, 4727 EP (2014). [19] R. Fei and L. Yang, Nano Lett. 14, 2884 (2014). [20] H. Liu, Y. Du, Y. Deng, and P. D. Ye, Chem. Soc. Rev. 44, 2732 (2015). [21] J. Qiao, X. Kong, Z.-X. Hu, F. Yang, and W. Ji, Nat. Commun. 5, 4475 EP (2014). [22] V. Tran, R. Soklaski, Y. Liang, and L. Yang, Phys. Rev. B 89, 235319 (2014). [23] R. W. Keyes, Phys. Rev. 92, 580 (1953). [24] Q. Wei and X. Peng, Appl. Phys. Lett. 104, 251915 (2014). [25] C. Yongqing, Z. Gang, and Z. Yong-Wei, Sci. Rep. , 1 (2014). [26] A. C.-G. Zant, L. Vicarelli, E. Prada, J. O. Island, K. L. Narasimha-Acharya, S. I. Blanter, D. J. Groenendijk, M. Buscema, G. A. Steele, J. V. Alvarez, H. W. Zandbergen, J. J. Palacios, and H. S. J. van der, 2D Mater. 1, 025001 (2014). [27] K. Momma and F. Izumi, J. Appl. Crystallogr. 44, 1272 (2011). [28] C. Kamal and M. Ezawa, Phys. Rev. B 91, 085423 (2015). [29] G. Wang, R. Pandey, and S. P. Karna, ACS Appl. Mater. Inter. 7, 11490 (2015). 24 [30] I. Kokubo, Y. Yoshiike, K. Nakatsuji, and H. Hirayama, Phys. Rev. B 91, 075429 (2015). [31] Y. Lu, W. Xu, M. Zeng, G. Yao, L. Shen, M. Yang, Z. Luo, F. Pan, K. Wu, T. Das, P. He, J. Jiang, J. Martin, Y. P. Feng, H. Lin, and X.-s. Wang, Nano Lett. 15, 80 (2015). [32] Z. Zhu, J. Guan, and D. Tomnek, Nano Lett. 15, 6042 (2015). [33] F. Shojaei and H. S. Kang, J. Phys. Chem. C 119, 20210 (2015). [34] M. Xie, S. Zhang, B. Cai, Y. Huang, Y. Zou, B. Guo, Y. Gu, and H. Zeng, Nano Energy 28, 433 (2016). [35] B. Liu, M. Kpf, A. N. Abbas, X. Wang, Q. Guo, Y. Jia, F. Xia, R. Weihrich, F. Bachhuber, F. Pielnhofer, H. Wang, R. Dhall, S. B. Cronin, M. Ge, X. Fang, T. Nilges, and C. Zhou, Adv. Mater 27, 4423 (2015). [36] L. Kou, Y. Ma, X. Tan, T. Frauenheim, A. Du, and S. Smith, J. Phys. Chem. C 119, 6918 (2015). [37] V. Pardo and W. E. Pickett, Phys. Rev. Lett. 102, 166803 (2009). [38] C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, I. Leermakers, U. Zeitler, Y. Sk- ourski, J. Wosnitza, Z. Liu, Y. Chen, W. Schnelle, H. Borrmann, Y. Grin, C. Felser, and B. Yan, Nat Phys 11, 645 (2015), letter. [39] X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, Phys. Rev. B 83, 205101 (2011). [40] H. Nielsen and M. Ninomiya, Phys. Lett. B 130, 389 (1983). [41] A. M. Turner, A. Vishwanath, and C. O. Head, Topological Insulators 6, 293 (2013). [42] S. A. Parameswaran, T. Grover, D. A. Abanin, D. A. Pesin, and A. Vishwanath, Phys. Rev. X 4, 031035 (2014). [43] S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J. Mele, and A. M. Rappe, Phys. Rev. Lett. 108, 140405 (2012). [44] Z. Wang, Y. Sun, X.-Q. Chen, C. Franchini, G. Xu, H. Weng, X. Dai, and Z. Fang, Phys. Rev. B 85, 195320 (2012). [45] Z. Wang, H. Weng, Q. Wu, X. Dai, and Z. Fang, Phys. Rev. B 88, 125427 (2013). [46] Z. Liu, J. Wang, and J. Li, Phys. Chem. Chem. Phys. 15, 18855 (2013). [47] J. Wang, S. Deng, Z. Liu, and Z. Liu, Nat. Sci. Rev. 2, 22 (2015). [48] J. Kim, S. S. Baik, S. H. Ryu, Y. Sohn, S. Park, B.-G. Park, J. Denlinger, Y. Yi, H. J. Choi, and K. S. Kim, Science 349, 723 (2015). [49] C. Wang, Q. Xia, Y. Nie, M. Rahman, and G. Guo, AIP Advances 6, 035204 (2016). 25 [50] G. Yao, Z. Luo, F. Pan, W. Xu, Y. P. Feng, and X.-s. Wang, Sci. Rep. 3, 2010 EP (2013). [51] Y. Lu, D. Zhou, G. Chang, S. Guan, W. Chen, Y. Jiang, J. Jiang, X.-s. Wang, S. A. Yang, Y. P. Feng, Y. Kawazoe, and H. Lin, NPJ Comput. Mater. 2, 16011 EP (2016). [52] M. Zhao, X. Zhang, and L. Li, Sci Rep 5, 16108 (2015). [53] S. Zhang, M. Xie, B. Cai, H. Zhang, Y. Ma, Z. Chen, Z. Zhu, Z. Hu, and H. Zeng, Phys. Rev. B 93, 245303 (2016). [54] H. Doh and H. J. Choi, 2D Mater. 4, 025071 (2017). [55] S. Zhang, Z. Yan, Y. Li, Z. Chen, and H. Zeng, Angew. Chem. Int. Ed. 54, 3112 (2015). [56] A. Manjanath, A. Samanta, T. Pandey, and A. K. Singh, Nanotechnology 26, 075701 (2015). [57] J. Han, J. Xie, Z. Zhang, D. Yang, M. Si, and D. Xue, Appl. Phys. Express 8, 041801 (2015). [58] Z. Zhu, J. Guan, and D. Tom anek, Phys. Rev. B 91, 161404 (2015). [59] T. Hu and J. Dong, Phys. Rev. B 92, 064114 (2015). [60] Z. J. Xiang, G. J. Ye, C. Shang, B. Lei, N. Z. Wang, K. S. Yang, D. Y. Liu, F. B. Meng, X. G. Luo, L. J. Zou, Z. Sun, Y. Zhang, and X. H. Chen, Phys. Rev. Lett. 115, 186403 (2015). [61] P.-L. Gong, D.-Y. Liu, K.-S. Yang, Z.-J. Xiang, X.-H. Chen, Z. Zeng, S.-Q. Shen, and L.-J. Zou, Phys. Rev. B 93, 195434 (2016). [62] R. Fei, V. Tran, and L. Yang, Phys. Rev. B 91, 195319 (2015). [63] A. N. Abbas, B. Liu, L. Chen, Y. Ma, S. Cong, N. Aroonyadet, M. Kpf, T. Nilges, and C. Zhou, ACS Nano 9, 5618 (2015). [64] L. Kou, T. Frauenheim, and C. Chen, J. Phys. Chem. Lett. 5, 2675 (2014). [65] T. Low, A. S. Rodin, A. Carvalho, Y. Jiang, H. Wang, F. Xia, and A. H. Castro Neto, Phys. Rev. B 90, 075434 (2014). [66] M. Baba, Y. Takeda, K. Shibata, T. Ikeda, and A. Morita, Jpn. J. Appl. Phys. 28, L2104 (1989). [67] C.-M. Park and H.-J. Sohn, Adv. Mater 19, 2465 (2007). [68] L.-Q. Sun, M.-J. Li, K. Sun, S.-H. Yu, R.-S. Wang, and H.-M. Xie, J. Phys. Chem. C 116, 14772 (2012). [69] J. Sun, H.-W. Lee, M. Pasta, H. Yuan, G. Zheng, Y. Sun, Y. Li, and Y. Cui, Nat. Nano 10, 980 (2015). 26 [70] J. Sun, G. Zheng, H.-W. Lee, N. Liu, H. Wang, H. Yao, W. Yang, and Y. Cui, Nano Lett. 14, 4573 (2014). [71] S. S. Baik, K. S. Kim, Y. Yi, and H. J. Choi, Nano Lett. 15, 7788 (2015). [72] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.: Condens. Mat. 21, 395502 (2009). [73] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996). [74] D. Vanderbilt, Phys. Rev. B 41, 7892 (1990). [75] E. Kucukbenli, M. Monni, B. Adetunji, X. Ge, G. Adebayo, N. Marzari, S. De Gironcoli, and A. D. Corso, arXiv preprint arXiv:1404.3015 (2014). [76] S. Grimme, J. Comput. Chem. 27, 1787 (2006). [77] N. Marzari, D. Vanderbilt, A. De Vita, and M. C. Payne, Phys. Rev. Lett. 82, 3296 (1999). [78] V. Iyer, P. Ye, and X. Xu, 2D Mater. 4, 021032 (2017). [79] C. R. Dean, F. A. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K. Watanabe, T. Taniguchi, P. Kim, K. L. Shepard, and J. Hone, Nat. Nano 5, 722 (2010). [80] K.-K. Liu, W. Zhang, Y.-H. Lee, Y.-C. Lin, M.-T. Chang, C.-Y. Su, C.-S. Chang, H. Li, Y. Shi, H. Zhang, C.-S. Lai, and L.-J. Li, Nano Lett. 12, 1538 (2012). [81] J. C. H. A dl and J. L. Kardos, Polym. Eng. Sci. 16, 344 (1976). [82] R. Mullen, R. Ballarini, Y. Yin, and A. Heuer, Acta. Mater. 45, 2247 (1997). [83] R. Hearmon, Physics of the Solid State , 401 (1969). [84] J. M. J. den Toonder, J. A. W. van Dommelen, and F. P. T. Baaijens, Model. Simul. Mater. Sc. 7, 909 (1999). [85] A. Reuss, ZAMM-Z Angew. Math. Mech. 9, 49 (1929). [86] R. D. Cook and W. C. Young, Advanced Mechanics of Materials (Pearson College Division, 1999). [87] W. Voigt, Ann. Phys. 274, 573 (1889). [88] R. Hill, P. Phys. Soc. 65, 349 (1952). 27 [89] U. Khan, P. May, A. O'Neill, A. P. Bell, E. Boussac, A. Martin, J. Semple, and J. N. Coleman, Nanoscale 5, 581587 (2013). [90] W. E. Mahmoud, Eur. Polym. J. 47, 1534 (2011). [91] J. Liang, Y. Huang, L. Zhang, Y. Wang, Y. Ma, T. Guo, and Y. Chen, Adv. Funct. Mater. 19, 2297 (2009). [92] J. A. King, D. R. Klimek, I. Miskioglu, and G. M. Odegard, J. Appl. Polym. Sci. 128, 4217 (2013). [93] R. J. Young, I. A. Kinloch, L. Gong, and K. S. Novoselov, Compos. Sci. Technol. 72, 1459 (2012). [94] Y. Yang, W. Rigdon, X. Huang, and X. Li, Sci. Rep. 3, 2086 EP (2013). [95] S.-K. Kim, J. J. Wie, Q. Mahmood, and H. S. Park, Nanoscale 6, 7430 (2014). [96] S. Appalakondaiah, G. Vaitheeswaran, S. Leb egue, N. E. Christensen, and A. Svane, Phys. Rev. B 86, 035105 (2012). [97] D. Kecik, E. Durgun, and S. Ciraci, Phys. Rev. B 94, 205410 (2016). [98] Z. Zhang, J. Xie, D. Yang, Y. Wang, M. Si, and D. Xue, Appl. Phys. Express 8, 055201 (2015). [99] Y. Xu, B. Peng, H. Zhang, H. Shao, R. Zhang, and H. Zhu, Ann. Phys. 529, 1600152 (2017). [100] L. Cartz, S. R. Srinivasa, R. J. Riedner, J. D. Jorgensen, and T. G. Worlton, J. Chem. Phys. 71, 1718 (1979). [101] P. M. Smith, A. J. Leadbetter, and A. J. Apling, Philos. Mag. 31, 57 (1975). [102] C. S. Barrett, P. Cucka, and K. Haefner, Acta Criystallogr. 16, 451 (1963). [103] D. C  ak r, H. Sahin, and F. m. c. M. Peeters, Phys. Rev. B 90, 205421 (2014). [104] A. J. Cohen, P. Mori-S anchez, and W. Yang, Phys. Rev. B 77, 115123 (2008). [105] Y. Hinuma, A. Gruneis,  G. Kresse, and F. Oba, Phys. Rev. B 90, 155405 (2014). [106] U. von Barth, Phys. Scripta 2004, 9 (2004). [107] A. J. Cohen, P. Mori-Snchez, and W. Yang, Chem. Rev. 112, 289 (2012). [108] Y. Wang and Y. Ding, Nanoscale Res. Lett. 10, 254 (2015). [109] A. Morita, Appl. Phys. A 39, 227 (1986). [110] D. Warschauer, J. Appl. Phys. 34, 1853 (1963). [111] S. Narita, Y. Akahama, Y. Tsukiyama, K. Muro, S. Mori, S. Endo, M. Taniguchi, M. Seki, S. Suga, A. Mikuni, and H. Kanzaki, Physica B&C 117, 422 (1983). 28 [112] Y. Maruyama, S. Suzuki, K. Kobayashi, and S. Tanuma, Physica B&C 105, 99 (1981). [113] G. Greaves, S. Elliott, and E. Davis, Adv. Phys. 28, 49 (1979). [114] O. U. Akturk,  V. O. Oz celik, and S. Ciraci, Phys. Rev. B 91, 235446 (2015). [115] G. Bihlmayer, O. Rader, and R. Winkler, New J. Phys. 17, 050202 (2015). [116] Z. S. Popovi c, J. M. Kurdestany, and S. Satpathy, Phys. Rev. B 92, 035135 (2015). [117] L. Yang, Y. Song, W. Mi, and X. Wang, RSC Adv. 6, 66140 (2016). [118] M. Yoshizawa, I. Shirotani, and T. Fujimura, J. Phys. Soc. Jpn. 55, 1196 (1986). [119] Y. Kzuki, Y. Hanayama, M. Kimura, T. Nishitake, and S. Endo, J. Phys. Soc. Jpn. 60, 1612 (1991). [120] T. Akai, S. Endo, Y. Akahama, K. Koto, and Y. Marljyama, High Pressure Res. 1, 115 (1989). [121] R. M. Jones, Mechanics of Composite Materials, Vol. 193 (Scripta Book Company Washing- ton, DC, 1975).

Journal

Condensed MatterarXiv (Cornell University)

Published: Jan 24, 2018

There are no references for this article.