Olga
Arroyo-Gascón
ab,
Yuriko
Baba
b,
Jorge I.
Cerdá†
a,
Oscar
de Abril
c,
Ruth
Martínez
b,
Francisco
Domínguez-Adame
b and
Leonor
Chico
*ab
aInstituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Científicas, C/Sor Juana Inés de la Cruz 3, E-28049 Madrid, Spain. E-mail: leochico@ucm.es
bGISC, Departamento de Física de Materiales, Universidad Complutense, E-28040 Madrid, Spain
cDepartamento de Estructuras y Física de Edificación, Universidad Politécnica de Madrid, E-28040 Madrid, Spain
First published on 5th April 2022
We investigate the effect of a non-magnetic donor impurity located at the surface of the SnTe topological crystalline insulator. In particular, the changes on the surface states due to a Sb impurity atom are analyzed by means of ab initio simulations of pristine and impurity-doped SnTe. Both semi-infinite and slab geometries are considered within the first-principles approach. Furthermore, minimal and Green's function continuum models are proposed with the same goal. We find that the Dirac cones are shifted down in energy upon doping; this shift strongly depends on the position of the impurity with respect to the surface. In addition, we observe that the width of the impurity band presents an even–odd behavior by varying the position of the impurity. This behavior is related to the position of the nodes of the wave function with respect to the surface, and hence it is a manifestation of confinement effects. We compare slab and semi-infinite geometries within the ab initio approach, demonstrating that the surface states remain gapless and their spin textures are unaltered in the doped semi-infinite system. In the slab geometry, a gap opens due to hybridization of the states localized at opposite surfaces. Finally, by means of a continuum model, we extrapolate our results to arbitrary positions of the impurity, clearly showing a non-monotonic behavior of the Dirac cone.
One of the families within the broad class of symmetry-protected matter are topological crystalline insulators (TCIs). Unlike 2 insulators, TCIs are protected by crystal symmetries and present an even number of Dirac cones in their surfaces, so that their
2 index is zero. These symmetries include rotations, reflections, or even glide planes. The first TCIs discovered were IV–VI semiconductors, such as SnTe and Pb1−xSnxTe ternary alloys, which are protected by mirror symmetries.6–9 When a mirror symmetry is responsible for the topological protection, the relevant topological invariant is the mirror Chern number nM. There are two types of surface states in TCIs, depending on the location of the Dirac points, with different properties. Type-I surface states have their Dirac points at time-reversal invariant momenta, whereas in type-II TCIs the Dirac cones are displaced from such time-reversal invariant k-points.10
SnTe has a rocksalt crystalline structure with small bandgaps at the L points of the Brillouin zone (BZ), where the conduction and valence bands are inverted.6,11 The mirror plane that protects these topological states is (110) [see Fig. 1(c)], so that an even number of robust Dirac cones appear at the (001), (110) and (111) crystal surfaces, which are symmetric with respect to the (110) mirror reflection. For the (111) termination, surface states are of type-I, whereas (001) and (110) surface states are identified as type-II. For the [001] surface termination, which is of type II, two equivalent L points of the BZ are projected in the ΓL1L2 plane to the same point, as depicted in Fig. 1(b). Hence, two parent Dirac cones coexist at
, and their interaction gives rise to two pairs of massless Dirac fermions deviated from
, while the high-symmetry point is found to be gapped10 [see Fig. 1(d)]. SnTe undergoes a crystalline phase transition at low temperatures (below 100 K) to a rhombohedral lattice. Such distortion has a small effect in the bulk bands, but destroys the surface Dirac cones, as any other perturbation that breaks the mirror (110) symmetry would do.12 Interestingly, within this rhombohedral phase, SnTe has been recently shown to be a higher-order TI, showing hinge states along specific directions.13
On the other hand, PbTe, which also has a rocksalt structure, is not a TCI and its surface states are gapped. Notwithstanding, a topological phase transition can be induced either by pressure or alloying. In particular, as anticipated above, Pb1−xSnxTe is a topologically non-trivial material (nM = −2) for x ⪆ 1/3.7,9 Actually, a band inversion as a function of the alloy composition was reported long time ago.14
Thin films of these TCIs also show nontrivial topological behavior. However, hybridization between the surface states opens a gap that decreases for increasing layer thickness.15,16 In fact, a non-monotonic damped oscillation has been found, that it is ultimately related to the alternation between nonsymmorphic and symmorphic symmetry with the number of atomic layers.15,17 Specifically, if a [001]-oriented SnTe slab has an even number of layers, it possesses nonsymmorphic symmetry. This reduces the hybridization between surface states, whereas for an odd number of layers the symmetry is symmorphic, and more sizeable band gaps appear.
Additionally, symmetry-protected surface states in TCIs can be tuned by introducing impurities. The effects of doping on the electronic structure of bulk SnTe have been addressed in ref. 18 and 19. As for the (001) surface, In doping,20 hydrogen adsorption21 and diluted impurities22 have been considered. Doped SnTe, specially by In impurities, is an ongoing subject of discussion regarding topological superconductivity.23
Impurities in TCI thin films may open a gap or modify the surface states, allowing for the control of topological phase transitions in these materials and constituting an additional tool to modulate their properties.24,25 Indeed, tuning the electronic properties by doping or by applied external fields opens a way for the application of these materials in nanoscale devices.16 Such modification by doping may be also relevant for thermoelectric applications, another arena in which topological materials have been proven to be of interest.18,19
In this work, the effect of non-magnetic impurities in the topological crystalline insulator SnTe is studied. Specifically, we consider Sn substitution by a Sb atom, which acts as a donor impurity. In fact, Sb-doped SnTe has been found to be a superconductor.26 Here, we focus on the changes in the TCI dispersion relations induced by the presence of the Sb impurity. As mentioned above, doping may alter or modify the surface states by hybridization or just by opening a gap, therefore changing the electronic properties of the system. This is the main issue we address herein. We explore both semi-infinite and slab geometries, with a twofold goal. Firstly, ultrathin layers of TCIs are currently grown and incorporated in quasi-2D materials and devices as building blocks or integrating parts, so the thin slab geometry has an interest in itself. Secondly, in most first-principles calculations, a finite slab geometry is employed to model surfaces, so it is equally important to evaluate the differences between finite-size and semi-infinite calculations aimed at modeling the same system. To this end, DFT-based ab initio simulations are used to characterize SnTe and the effect of the non-magnetic donor impurity near the crystal surface. We analyze the hybridization between surface and impurity states as well as the effect of the impurity position on the electronic structure and crystal symmetries. Furthermore, both perturbation theory and continuum models are proposed. The former is used to predict the behavior of the Dirac cone upon doping, while the latter describes the impurity states.
Our main results are summarized as follows: (i) the Dirac cones appearing at the TCI surfaces are shifted down upon the n-doping due to Sb substitutional impurities. (ii) The shift of the Dirac cones is strongly dependent on the position of the impurity with respect to the surface. (iii) The width of the impurity band presents an even–odd behavior with the layer position of the dopant, both in the semi-infinite and the slab geometries; we relate it to the location of the nodes of the wave function with respect to the surface. (iv) In the slab system, a gap opens in the Dirac points due to coupling of the states at opposite surfaces, even though they are almost 50 Å apart. We verify that the surface states remain gapless in the semi-infinite geometry, with unaltered spin textures. (v) Finally, we derive continuum models in order to extrapolate the behavior of this doped system for larger sizes, unreachable with more accurate, ab initio simulations.
Our work provides a complete picture of the interplay between impurity and surface states in TCIs, highlighting the differences between finite-sized slabs and single surfaces, and additionally including simpler continuum models which yield a better interpretation of the numerical first-principles results.
A double-ζ singly-polarized basis set was chosen for all the atoms involved, with a radial extension of the strictly localized orbitals determined from a confinement energy (energy shift) of 100 meV. The resolution of the real space grids was set to around 0.06 Å3 (equivalent to a mesh cutoff of 600 Ry). For the integration over the BZ we employed k-supercells of size around (17 × 17) relative to the SnTe-(1 × 1) bulk unit cell, with a temperature in the Fermi–Dirac occupation function of 100 meV.
A lattice parameter of 6.39 Å was used following ref. 6 and 18, which yields a satisfactory agreement to experimental results and a successful magnitude of the band gap.14 This allows us to model rocksalt room-temperature SnTe, avoiding the aforementioned rhombohedral phase transition.9 In order to tackle the robustness of the surface states against film thickness, several slab calculations were performed using different number of layers (not shown here). We found an optimal compromise between accuracy and our computational capability for a nonsymmorphic 16 atomic-layer SnTe slab. Therefore, all the slab calculations presented in this work have been performed with this thickness.
Substitutional donor impurities have been considered by replacing a Sn atom by Sb within a c(4 × 4) supercell—that is, in a defective layer one out of eight Sn atoms is substituted by a Sb atom. We have considered up to four different locations of the Sb: at the top surface layer (referred to as Sb1 hereafter), the second layer (Sb2), the third layer (Sb3) and the fourth layer (Sb4). Furthermore, in order to avoid net dipoles in the slab, a second Sb substitutional impurity was always placed at an equivalent position, but with respect to the bottom layer [see Fig. 1(a)]. In this way, both impurities are located at the same distance from the corresponding nearest surface. Thus, the simulated impurity doping in the slabs corresponds to a small value of x = 0.016.
Atomic relaxation was performed for each structure. For the Sb1 slab, the two top and bottom layers were relaxed while for the Sb2, Sb3 and Sb4 slabs, the 3, 4 and 5 layers counted from the top and bottom surfaces were included in the relaxations until forces on the atoms were below 0.02 eV Å−1. The top and bottom layers of the final equilibrium structure are the most affected by the relaxation. To be specific, some atoms of the Sb1 and Sb3 slabs are displaced up to 0.17 Å compared to 0.14 Å for the Sb2 and Sb4 slabs.
Although minimal, the hybridization of opposite surface states still has an effect on the band structure. Hence, the electronic and spin structure of the Sb-doped SnTe(001) surface will be mainly presented as calculated under a semi-infinite geometry via projected density of states and magnetization maps, PDOS(k,E) and Mα(k,E) respectively, with α = x,y,z indicating the direction.32,33 To this end, we have employed surface Green's function matching techniques in order to couple the Hamiltonian matrix elements of the top half of the slab with those of a bulk SnTe calculation. Self-consistency was preserved in the matching process, as reflected by the fact that maximum deviations were always smaller than 10 meV between the on-site energies of the Sn and Te central layers of the slab and those in the SnTe bulk. PDOS(k,E) and Mα(k,E) maps were computed with a resolution of 0.003 Å−1 in reciprocal space and 5 meV in energy. Accordingly, the self-energy that determines the spectral width of any surface states (imaginary part of the energy entering the Green's function) was set to 5 meV. For zoom-ins of the Dirac points we computed additional maps with an increased resolution (0.0016 Å−1 and 3 meV) around the -point and over a small energy window. In the semi-infinite calculations, a single impurity is used. For topologically-protected surface states, we expect them to remain unaltered, since the placement of a single impurity does not change the (110) mirror symmetry, as outlined in Fig. 1(c). On the contrary, the 16 atomic layer slabs do break this symmetry when taking into account two impurities, as explained before. Since two-dimensional (few-layer) topological crystalline insulators are also the subject of great attention, we consider this comparison to be relevant.
The bottom panel of Fig. 2 displays the equivalent electronic structures to the top panel but in the form of PDOS(k,E) maps, since they have been computed under a semi-infinite geometry (see previous section). The Sb-doped surfaces, overall, show a similar behavior to that found for the 16-layer slabs, namely, the appearance of Sb impurity bands across the gap showing an even–odd effect in their dispersion, a clear n-type doping and the presence of Dirac cones slightly away from . However, these Dirac cones remarkably persist in all defective cases of the semi-infinite structure. The splitting of the impurity bands induced by the spin–orbit coupling remains small until it approaches the
point, where the two branches start to deviate from each other by more than 100 meV.
Focusing now on the electronic structure around the point, the most relevant feature is the survival of a gapless Dirac point in all structures, which is consistent with the fact that the (110) mirror plane still holds in the doped structures and, hence, the Dirac points remain symmetry-protected. However, there are qualitative differences with respect to the pristine surface which are best seen in Fig. 3, where zoom-ins with an enhanced resolution are presented. First of all, and in accordance with the n-type doping mentioned above, the Dirac points consistently shift by around 150–200 meV towards lower energies. Precisely at the
point the interpretation of the electronic structure becomes more complex due to the profusion of extra bands arising from both the Sb impurity and the band folding.
All relevant bands within the gap are better resolved in Fig. 3(a), where we present enlarged views of the PDOS(k,E) maps. The projections have been taken over the surface SnTe layer, superimposing the Sb contributions in blue.
For the clean surface case, the branches forming the Dirac cone can be clearly identified, as well as the Sb-split impurity bands, denoted as a and b. There is a further band inside the gap, labeled by f, corresponding to the back-folded topological surface state along –
[see Fig. 1(d)]. Since the supercell size is much smaller than the effective Bohr radius, impurities cannot be regarded as truly isolated and the impurity bands become dispersive. In addition, further broadening of the bands arises from the hybridization of impurity states with Bloch states of the conduction band of SnTe. In fact, the impurity states are mainly composed of Sb and Sn contributions, as depicted in the Appendix A for both slab and bulk band structures. Band inversion between Sn and Te orbitals also persists in the doped material, as discussed and shown in Appendix A.
In the lower panels (b) and (c) of Fig. 3, magnetization maps along the x- and y-axis, Mx(k,E) and My(k,E), respectively, are shown. We omit the out-of-plane component Mz(k,E) since it is always negligible except for Sb1. As it could be expected, the in-plane spin orientations of the surface band remain unaltered with respect to the pristine case. Indeed, this is best seen in the PDOS(kx,ky) maps shown in the lower panels, computed at energies 20 meV below [panel (d)] and above [panel (e)] the location of the Dirac point.
The spin texture for all structures is shown in Fig. 4 in the form of Mx/y(k,E) maps. Note that pristine SnTe only shows in-plane spin polarization due to time-reversal and mirror symmetries.34 The effect of folding on the spin texture of pristine SnTe is explicitly detailed in the Appendix A, showing additional Mx/y(kx,ky) maps near the Dirac point for all structures. A close inspection of the spin textures reveals that the spin orientation of the topological surface states always remains unchanged with respect to the pristine surface. More surprising is the behavior of the Sb bands: close to they behave as standard Rashba-split states, with a reduced separation in energy between a and b of only 10–20 meV while they present opposite magnetizations. In fact, this is the only region where out-of-plane magnetization, Mz(k,E), is not negligible. However, as they disperse towards
, band a keeps its spin orientation but b undergoes an abrupt spin inversion and ends up with the same orientation as a. Note that the spin texture of the impurity states is not as robust as that of the Dirac cone, since it varies along the high-symmetry path and is dependent on the choice of atom projections.
Given such ultrathin layers, the energy shift of the Dirac point can be related to quantum size effects. In order to clarify this point, we study the spatial distribution of the wave function for the impurity band, depicted in Fig. 5. A series of nodes appear right after the impurity and approximately each two atomic layers in the z-direction. When the impurity is located at an odd atomic layer, there is an even number of layers above it, so the wave function amplitude is almost zero at the surface layer. In contrast, when the impurity is placed in an even layer, there is an odd number of layers above it and the wave function is nonzero at the surface, therefore favoring hybridization with the surface states and a larger band gap. Since the derivative of the wave function is related to the kinetic energy operator, it is also expected for these impurity bands to move upwards, as shown in both slab and semi-infinite calculations in Fig. 2. Thus, we come to the conclusion that this energy shift can be also seen as a size or quantum confinement effect.
![]() | (1) |
Modifications of this model, including anisotropy and on-site momentum dependent terms, have been successfully proposed to study the Bi2Se3 family of materials.36,38 In fact, the model has also been applied to describe the (001) surface states of SnTe including just the leading terms linear in momentum.6,10,15,17,34,39 Since our first goal within the continuum approach is to analyze the effect of an impurity in close proximity of an isolated surface, quadratic terms are indispensable to have a nontrivial topology in a vacuum–SnTe interface.35 We expect to find the same shifting behavior as predicted by Araújo et al.17 and our DFT calculations. Thus, we perform first-order perturbation theory (PT) on the surface states obtained within this formalism with the aim of elucidating the effect of the impurity on the gapless surface states. Secondly, in order to describe the energy dependence of the impurity states with respect to the impurity position, we solve a simplified version of the Hamiltonian presented in eqn (1), summing up all terms in the perturbation series using a Green's function method. These two approaches allow us to investigate complementary aspects of the problem.
ψ± = As(e−λ1z − e−λ2z)exp[i(kxx + kyy)]Φ0±. | (2) |
In the former expression, the inverses of the decay lengths are given by λ1 and λ2, which are functions of the in-plane momenta and As is a normalization factor (see Appendix B). The dispersion of the states is linear in the in-plane momenta, yielding the Dirac cones of the surface states inside the bulk gap E± = ε ± vk∥sign(B), where .
For the purpose of obtaining the corrections on the energy, we perform first-order PT, modeling the impurity with a Coulomb-like interaction
![]() | (3) |
The first-order correction in PT is given by
![]() | (4) |
Fig. 6 shows the displacement of the Dirac point as a function of the impurity depth z0. Similarly to Fig. 2, a shift of the Dirac point as the impurity is placed farther from the surface is observed, in opposition to hole doping.20 While DFT calculations are difficult to address due to their computational cost when the impurity is far from the surface, this minimal model allows us to predict a return to the pristine case for large z0. Despite the aforementioned limitations, the model describes qualitatively well the phenomenon and it can be used as a simple approach to the effect of the impurity on the surface bands.
![]() | ||
Fig. 6 Displacement of the Dirac point ΔE(1) as a function of the impurity depth z0, obtained from PT calculations. |
The envelope functions of the electron states near the band extrema are then determined from the following Dirac-like Hamiltonian
![]() | (5) |
The presence of the impurity, located at r0 = (0,0,z0) without loss of generality, breaks the translational symmetry in the surface and the in-plane momentum is no longer conserved. Expressing distance in units of d and energy in units of m, the Dirac-like equation for the envelope function can be cast in the form
![]() | (6) |
The actual interaction potential V(r − r0) will be replaced by a non-local separable pseudo-potential of the form
![]() | (7) |
The Green's function for the unperturbed problem (λ = 0) and the Green's function for the total Hamiltonian in terms of the transition operator are derived in Appendix B. Poles in the complex plane of the transition matrix, E = Er − iΓ, where Er and Γ are real magnitudes, yield the energies of the impurity states. Resonances within the gap are given by the condition |Er| < 1. According to the model presented, an impurity embedded in a bulk semiconductor can support a truly bound state. Lengthy but straightforward calculations lead to the following energy for the bound state
![]() | (8) |
Fig. 7 shows the real and imaginary parts of the poles of the transition matrix as a function of the position of the impurity close to the heterojunction, when kc = 100 and λ = −0.15. The real part Er approaches the value in the bulk material Eb when increasing z0, as expected. These impurity states move to higher energies as the distance from the surface increases, as in ref. 49. The alternating behavior found in DFT calculations, with different energy shifts of the impurity bands for even (odd) slabs, is not reflected in this model since simpler, exponentially-decaying wave functions are used here.
![]() | ||
Fig. 7 Real and imaginary parts of the complex poles E = Er − iΓ of the transition matrix as a function of the impurity position. The parameters of the interaction potential are λ = −0.15 and kc = 100. Dashed line indicates the energy of the impurity bound state in the bulk semiconductor [see eqn (8)]. |
A continuum Hamiltonian is also used to describe the effect of doping on the Dirac cone, which shifts downwards. This result coincides qualitatively with our DFT calculations and allows us to predict the behavior of the system when the impurity is placed deeper in the slab. As for the hybridization between the topologically-protected surface states and the impurity states, a simplified version of the aforementioned Hamiltonian is used, showing a diminishing resonance as the impurity moves inwards.
The n-doped character of SnTe:Sb could make difficult to separate the contribution to the electronic properties of surfaces and bulk states. However, angle-resolved photoemission spectroscopy (ARPES) enables revealing the signatures of surface states (i.e. Dirac cone) even if the Fermi level lies well above the Dirac point, as was demonstrated by Chen et al. in the Bi2Te3 samples.50 We are then confident that our predictions could be observed in ARPES experiments.
In conclusion, SnTe, the first discovered TCI, is an excellent platform to explore the interplay of symmetry and topology in this class of Dirac matter. A simple perturbation such as a substitutional impurity is shown to give rise to non-trivial spin textures. Besides its fundamental interest, tuning the Dirac cones of topological insulators can be of interest for transport and spintronic applications.
Fig. 9 depicts PDOS(kx,ky,E) and spin texture maps of SnTe above the Dirac point. Two pairs of Dirac cones appear, which are folded into a ×-shaped fourfold pattern at the Γ point for the c(4 × 4) supercell. The Mz(k,E) component is omitted due to the absence of out-of-plane spin texture.
Fig. 10 shows the persistence of the pristine SnTe spin texture when impurities are included. The helicity is unaltered independently of the impurity position, as mentioned in Fig. 3. Additionally, it can be seen in Fig. 4 that the spin texture changes sign upon moving above or below the Dirac point.
Due to the breaking of translational symmetry, we solve the Hamiltonian (3) with kz → −i∂z, using the ansatz of the form
ψ ∼ e−λz![]() | (9) |
We found that two decay lengths satisfy the eigenvalue equation
![]() | (10) |
By combining solutions of the form (9) in order to fulfill the hard-wall boundary conditions at z = 0, we obtain two surface states with linearly dispersive energy given by E± = ε ± vk∥sign(B) with wave functions
ψ+ = As(e−λ1z − e−λ2z)exp[i(kxx + kyy)]Φ0+, | (11) |
ψ− = As(e−λ1z − e−λ2z)exp[i(kxx + kyy)]Φ0−, | (12) |
![]() | (13) |
The numerical values of the parameters have been obtained by fitting to DFT data and are reported in Table 1.
SnTe |
---|
ε = −0.065 eV |
m = 0.07 eV |
v = 1.591 eV Å |
B = 2.74 eV Å2 |
In eqn (3), replacing the electron free mass by the effective mass and the vacuum dielectric constant by that of the medium, the following relation for the effective Rydberg and Bohr radius can be easily obtained
![]() | (14) |
The Green's function for the unperturbed problem (λ = 0) satisfies the following equation
![]() | (15) |
The Green's function G0(r,r′;E) = 〈r|Ĝ0(E)|r′〉 is the matrix element of the unperturbed resolvent Ĝ0(E) in coordinate representation. Let Ĝ(E) be the resolvent of the Hamiltonian , where
is the operator associated to the non-local separable potential (7). Ĝ(E) and Ĝ0(E) are related by the well-known equation
(see ref. 54), where
![]() | (16) |
![]() | (17) |
![]() | (18) |
![]() | (19) |
The proper Green's function of the unperturbed band-inverted heterojunction can be obtained as follows. We introduce (18) in eqn (15) to obtain
![]() | (20) |
The resolvent of ĥ0 is Ĝ0(E) = (E14 − ĥ0)−1. We now define the auxiliary resolvent ĝ0(E) = (E214 − ĥ02)−1 so that Ĝ0(E) = (E14 + ĥ0)ĝ0(E). Therefore, once ĝ0(E) is known, we can easily obtain Ĝ0(E). In coordinate representation
![]() | (21) |
This equation can be easily solved regarding the term −2iβαxδ(x) as an interaction potential and then using the Dyson equation. In doing so we obtain
![]() | (22) |
Finally, recalling that Ĝ0(E) = (E + ĥ0)ĝ0(E), we get eqn (19).
Once the basic equations are presented, we now need to take a particular shape function to perform the calculation. In what follows we consider
ω(r − r0) = f(z − z0)δ(r∥), | (23) |
det[14 − M(E)] = 0, | (24a) |
![]() | (24b) |
We have numerically solved eqn (24a) considering a finite bandwidth. This is equivalent to introducing an upper cutoff for the in-plane momentum, kc. In this way, we prevent divergences when taking the limit L → 0, i.e. f(z − z0) → δ(z − z0).
Footnote |
† Deceased. |
This journal is © The Royal Society of Chemistry 2022 |