1. Introduction
Three-dimensional topological insulators (TIs) are promising materials which have been studied due to their potential for use in a range of applications. There is a host of materials with different topological properties, identified by their transport properties or other qualities, such as their Berry phase or winding number, among others [
1,
2,
3].
The experimental manifestation of topological properties leads to the appearance of plateaus in resistivity measurements, phase transitions and other phenomena [
4,
5,
6,
7,
8]. The properties that manifest from the topological character are robust under perturbations, which is one of the main reasons to study topological regimes. As the topological character can protect determined quantum states against the decoherence effect of the environment, such quantum states should be useful for transporting and storing quantum information.
Heterostructures involving topological insulators are of great interest [
9,
10,
11,
12,
13], due to the possibility of modulating or altering electronic properties. Topological superconductors are heterostructures constructed using the proximity effect between a topological insulator and a superconductor, although this combination is not the only one that has been suggested for this purpose: both the combination of magnetic and topological materials and the doping of the topological materials with appropriate impurities are also under consideration [
10].
In the case of topological superconductors, assessing
a priori which materials could produce certain effects is so complicated that some authors have proposed numerical ab initio methods for predicting novel combinations of materials [
14], substituting the effective, simpler models used traditionally.
Experimental investigations have confirmed the presence of a single Dirac cone on the surface of topological insulators such as
,
,
, which enables the precise characterization of protected surface states, facilitates experimental validation of theoretical models and highlights this family of materials for their potential applications in spintronics devices [
15].
To enable theoretical analysis and further exploration of the low-energy electronic properties of TIs, several effective Hamiltonian models have been developed. The four-band
model stands out as a particularly effective framework, balancing computational tractability with the ability to capture the essential physics of surface states [
16,
17,
18]. Recent research on TI nanostructures, including nanowires and nanotubes, has shown that they exhibit novel tunable properties due to the relationship between the topological character and system geometry [
19,
20]. For instance, cylindrical TI nanowires have been studied analytically using hard-wall boundary conditions and 1/R expansions to understand the effects of curvature and radius on surface states and optical responses [
21].
Heterostructures combining distinct topological materials or integrating TIs with conventional semiconductors offer enhanced flexibility in tailoring electronic properties. The interfaces between these structures can induce novel topological phases as well as different surface states. Heterostructured nanodevices, in particular, exhibit complex geometries that introduce curvature and quantum confinement effects, profoundly influencing transport, spin characteristics or optical responses [
22,
23]. Previous studies have highlighted the critical role of curvature in shaping the electronic structure of cylindrical systems [
20,
21]. On the other hand, the behavior of confined and resonant states remains crucial for technological innovation [
24], especially in systems where external fields or structural variations can induce transitions between localized and extended states [
25]. Moreover, beyond topological protection, the electronic band structure itself can be exploited in applications such as photothermal conversion in core–shell heterostructures, as demonstrated in water dispersed
quantum dots [
26].
Additionally, multiple studies on TIs have used the entanglement entropy as a tool to identify non-trivial topological order [
27]; for instance, bipartite von Neumann entropies [
28,
29] computed for cylindrical nanowires provide a complementary characterization of topological phases, revealing the roles of finite-size effects and geometry in the localization and structure of edge states [
30]. This approach is particularly convenient in systems where conventional topological invariants are difficult to compute or interpret, and it complements energy spectrum analysis in identifying transitions and phase boundaries.
In this study, we perform a systematic analysis of cylindrical heterostructures comprising two TIs configured in a core and shell structure. We also study hollow nanowires made of a single TI. In both cases, we employ a low-energy approach based on the four-band model. Our methodology enables a detailed investigation of how confinement, curvature and material interfaces affect band structures, edge states and quantum spin properties. In particular, we study the appearance of more topological states on hollow nanowires compared to their whole counterparts, and their behavior in the limit where the width of the shell is small compared with the internal and external radii of the hollow nanowire. Our study uses some well-known quantities, such as the energy spectrum near the bulk gap, expectation values of spin operators, the energy gap between the topological states at the center of the band, localization color maps of the quantum states, the inverse participation ratio, the fidelity between quantum states, finite-size scaling analysis and quantum entropies.
The confinement of electrons in finite geometries, such as rings, disks, billiards, thin films and surfaces, profoundly alters their electronic and magnetic properties because of boundary-induced effects, even for non-topological materials. Confinement effects lead to a rich spectrum of phenomena, including the emergence of persistent currents which are sensitive to the electron density [
31], geometry-dominated orbital magnetic responses [
32], the formation of magnetic surface levels observable in impedance experiments [
33] and the evolution of persistent currents from mesoscopic rings to macroscopic Corbino disks, where the interplay between curvature and edge states determines the total magnetic response [
32,
34,
35].
We organized the remainder of this paper as follows. In
Section 2, we present a model for core–shell nanowires and its corresponding Hamiltonian. This section also contains a brief description of the Rayleigh–Ritz method, which allows us to calculate accurate approximations of the spectrum and corresponding eigenstates. In
Section 3.1, we present results for nanowires made of a single TI, which could be
,
or
. Although the behavior of the spectrum and the appearance of topological states whose energies lie on the bulk gap is well-known, particularly for nanowires made of
, we include the results in
Section 3.1 for comparison purposes.
Section 3.2 presents results for core–shell nanowires made with different combinations of the three TIs mentioned in the paragraph above. We consider the spectra of nanowires with six combinations of materials and distinct core and shell radius values, while studying the number of localized states dwelling in the gap between the conduction and valence bands, as well as the expectation values of the spin operators.
Section 4 presents the results obtained for hollow wires (or empty cores), enabling calculation of the inverse participation ratio and the fidelities between different quantum states and quantum entropies of normal and topological states to those quantities calculated in
Section 3.2. Finally, we present our conclusions in
Section 5. For completeness, we include some technical details in
Appendix A,
Appendix B,
Appendix C and
Appendix D.
2. Model and Hamiltonian
We consider cylindrical nanowires composed of strong three-dimensional topological insulators—specifically, , and —which are well known for hosting spin-momentum locked surface states. The nanowires are assumed to have a uniform cross-section with radius on the order of tens of nanometers, and to be infinitely extended along the axial (z) direction.
To describe the low-energy electronic properties of these nanowires, we adopt the effective
Hamiltonian introduced by Liu et al. [
16]. This model captures the essential features of both bulk and surface states near the
point of the Brillouin zone, and has been proven to be accurate in reproducing the band structure of the
material family. In Ref. [
16], Liu and coworkers showed that the eigenvalues of the four-band Hamiltonian fit the band energy dispersions for small enough values of the wave vector modulus.
The Hamiltonian is a four-band model incorporating band inversion and strong spin–orbit coupling—two fundamental ingredients to describe topological insulating behaviors. It is constructed using symmetry arguments and perturbative
theory, starting from an atomic basis formed by
orbitals of the
family of atoms [
15,
19,
20,
36,
37,
38]. The low-energy subspace is spanned by the following basis of spinor states:
where
and
refer to bonding and antibonding combinations of
orbitals with even and odd parity, respectively.
The effective bulk Hamiltonian takes the form
where
,
,
and the functions are defined as:
,
,
,
.
The material-dependent parameters
,
,
and
are taken from Ref. [
18] and are summarized in
Table 1.
Notwithstanding our choice of parameter set, the parameter set that defines the Hamiltonian differs depending on the set of ab initio energies used to fit their values. The eigenvalues of the Hamiltonian must fit the band energies obtained from ab initio calculations, which are more accurate over larger portions of the Brillouin zone. Ref. [
18] presented a detailed study of the different sets obtained by the authors and previous works, and showed that the
relaxed set of parameters offers the best fit [
18].
To obtain the band structure and the corresponding eigenvectors, we transform the Hamiltonian in Equation (
2) to the coordinate representation by transforming the wave vector into a differential operator. This procedure has been described in numerous references [
19,
24,
25,
39]. We briefly present some details of this procedure in
Appendix B. The coordinate representation of the Hamiltonian depends on the cylindrical coordinates
,
and
z, where the
z-axis direction coincides with the axis of the cylinder. Note that, due to the symmetry of the problem under consideration—that is, cylindrical nanowires with a finite radius and azimuthal symmetry—the band structure depends only on the wave vector in the
z direction,
.
Using the coordinate representation of the Hamiltonian, we compute approximate energy eigenvalues and eigenfunctions by means of the Rayleigh–Ritz variational method. This approach transforms the differential eigenvalue problem
into a finite-dimensional algebraic problem through calculation of the expectation value of
in a suitable basis of trial wavefunctions. This method has been widely employed to compute approximate eigenvalues and eigenfunctions in nanostructured systems [
24,
25,
40,
41,
42].
Due to the cylindrical symmetry of the system, the total angular momentum operator
commutes with the Hamiltonian:
[
21]. This implies that the eigenvalues of
are good quantum numbers, which motivates the construction of spinor-like trial functions which are simultaneous eigenstates of
and
H. We use the following ansatz for the four-component trial functions:
where
,
,
and
are complex variational coefficients;
labels the radial basis functions; and
is the orbital angular momentum quantum number associated with the azimuthal dependence. The total angular momentum quantum number is given by
, consistent with the spinor structure. The functions
form an orthonormal basis satisfying the radial boundary conditions, while the longitudinal momentum
is a good quantum number due to translational invariance along the wire axis. As we are looking for border states, we impose homogeneous boundary conditions on the radial functions
; for instance, for solid cylindrical nanowires composed of a single material with exterior radius
, we impose
.
The Rayleigh–Ritz variational method [
43,
44] provides an approximate solution to the eigenvalue problem by restricting the Hilbert space to a finite-dimensional subspace spanned by the chosen basis functions. In our case, the energy is approximated by minimizing the expectation value
with respect to the expansion coefficients defining the trial wavefunction
. Here,
is the coordinate representation of
, which results from transforming
in Equation (
2) into differential operators.
Appendix B provides further details about these transformations.
Minimization of the expectation value in Equation (
6) results in a matricial eigenvalue problem [
24,
25,
39,
41,
45,
46,
47]; see
Appendix A for more details. By fixing the values of the quantum numbers
L and
in Equation (
5), we get a
matricial eigenvalue problem, where
N denotes the number of radial functions used in the expansion. We solve the numerical eigenvalue problem using well-known numerical packages, obtaining an approximate spectrum
. By changing the values of
L and
, we can obtain the band structure of a given nanowire.
According to the variational principle, this approximation satisfies
where
E is the exact state energy of the full Hamiltonian
H. The variational estimate
thus serves as an upper bound to the true eigenvalue and, as the size of the basis set increases, this bound becomes progressively closer to the exact value. In practice, the procedure leads to a generalized matrix eigenvalue problem, whose solutions approximate the low-energy spectrum of the system within the chosen truncated basis.
This method is particularly effective for confined topological systems, for which it provides accurate approximations to both bulk-like and surface-localized states [
25,
30,
39,
45,
46].
In this work, we study the band structure and the properties of particular sets of eigenstates. We calculate expectation values of pure states, quantum entropies of reduced density matrices and the fidelity between different pure quantum states. Thus, we do not study the effect of temperature or transport properties. As topological states are protected against small disturbances, to some extent, the expectation values and other quantities obtained for topological states should be robust under thermal perturbation [
48,
49].
There is a growing body of literature dealing with heterostructures composed of layers of materials with different topological properties. The band parameters of the heterostructure can, in principle, be quite different from those of the bulk materials. Different crystalline structures and the geometry of the heterostructure are two factors that can modify the properties observed in a given heterostructure. The theoretical description of the electronic properties of a heterostructure using the model Hamiltonian—or any other form for the Hamiltonian—requires several simplifications concerning the parameters determining the Hamiltonian.
The hard-wall mass approximation implies that ideal surfaces separate the different materials composing the heterostructure, and that the Hamiltonian parameters change their values step-wise at both sides of a given surface. In some works, the parameter values at the sides of the surfaces are the bulk values. In other studies, the values depend on the particular pair of materials separated by the surface, and the band parameter values differ from those of the bulk ones. Overall, in order to proceed with calculation of the band structure and states, it is necessary to know the whole set of parameters for all the materials in the heterostructure. For instance, Ref. [
22] considered core–shell nanoparticles made of
and
, where the first compound was the topological insulator and the second acted as a non-topological insulator. The authors assumed that some of the parameters have the same value for both materials, and that the lattice parameter mismatch of both semiconductors was negligible enough to neglect the elastic strain in the nanostructure. The topological behavior of crystalline
was studied in Ref. [
50].
The absence of an appropriate set of parameters to properly determine the electron Hamiltonian on the whole nanostructure leads to the assumption of homogeneous boundary conditions for free-standing nanostructures, as in Ref. [
51]. In any case, if there is an interface between a TI and a non-topological insulator, there are matching conditions for the wave function and its derivative in the perpendicular direction to the interface surface.
4. Wires with an Empty Core
In this section, we consider a different kind of heterostructure than in the previous one. We aim to study the topological states that appear in a hollow wire, such as the one depicted in
Figure 10. In other words, we aim to study the problem of a wire with homogeneous boundary conditions for the state vector at interior and exterior radii.
To calculate the spectra of distinct cylinders, we employ the variational method again, although using a different basis set. The trial function now reads:
It is clear that the spinor in Equation (
18) satisfies homogeneous boundary conditions when
and
,
. The dependence on the quantum numbers
L and
is the same as that used to calculate approximate eigenvalues for nanowires made of a single or two materials. The trial function expectation values depend on integrals of sine functions, which have exact analytical expressions that are simpler to manage than those depending on Bessel functions. We provide some technical details in
Appendix B.
Another interesting property of the trial function in Equation (
18) is that it is not necessary to assume that there will be a topological state localized at each surface where there is a homogeneous boundary condition, as in Ref. [
20]; and that, for small values of the width
W, hybridization between the states localized at each surface will occur. Nevertheless, we aim to analyze the behaviors of several quantities to determine the physical changes produced when
. Besides analyzing the localization, we study an adapted inverse participation ratio (IPR) [
53,
54,
55,
56], the expectation values of the spin operators and the fidelity between quantum states calculated for different widths.
When the width of the shell of the cylinder filled with material is large enough for each value of the quantum number
L, there are two localized states—one on each surface. In this case, the localization length associated with each quantum state is smaller than the width of the material.
Figure 11 shows the behavior of the energy of both topological states, each of which has a well-defined cone when plotted as a function of
. In the following, we focus on hollow wires made of
, as this material allows for a better appreciation of the phenomena associated with studying the limit
.
Slightly changing the notation used in Equation (
18) and writing the
component of the
k-th variational eigenfunction as
where
with
and
depending on
or 4, we get that the probability density associated with the
k-th eigenstate
is expressed as
Figure 12 shows the probability densities for the border states for three different inner radii. In the leftmost column of panels, we plot the probability density for the state localized near the inner radius. In the rightmost column, we plot the probability density for the state localized near the exterior radius. The shell with a smaller width has
nm. Although to the naked eye it seems that the probability density of both states in panels (e) and (f) do not overlap, we will see that, in order to assess the states localized at the borders of the shell, we need more nuanced quantities.
A hybridization regime for very thin films has been identified in Refs. [
20,
22,
49,
51,
57]. In this regime, the helical states are not well-localized at the borders of the films. In the following, we study the behaviors of several quantities as functions of the shell width. For nanowires with an external radius
nm and shell widths
nm, the probability density function is appreciable over a region
nm. In the hybrid regime, the probability density function is appreciable over the whole shell’s thickness; see
Figure A1 in
Appendix C.
In disordered systems—for instance, a finite disordered chain of spins—it is customary to define the
inverse participation ratio [
53,
54,
55,
56]. Given a quantum state
in the site basis, the inverse participation ratio of the state is calculated as
where the coefficients
satisfy
and
is the site basis set.
Of course, there is no obvious site basis set to use in a continuous problem. Therefore, instead of using a site basis, we study how some states spread over the function basis set.
To study the spatial localization of the eigenstates as a function of the inner radius of the cylinder and the width occupied by the material, we compute the inverse participation ratio (IPR) for each edge state
k, defined as
where
denotes the expansion coefficient of the
k-th eigenstate in the variational basis, with
labeling the spinor component and
n labeling the radial mode index. The IPR serves as a measure of localization: larger values of
indicate that the state is more spatially localized, while lower values correspond to more de-localized states.
Figure 13 shows the behaviors of two different quantities: the inverse participation ratio
in Equation (
24), and the coefficient
of Ref. [
20] (see Figure 3 of the reference). The solid points correspond to the values of
, calculated for wires with different inner radius, as a function of the width occupied by the material. The color map corresponds to the coefficient
. In Ref. [
20], the authors derived an effective Hamiltonian for a hollow cylinder made of a topological insulator, starting from the Liu four-band Hamiltonian, Equation (
2). In doing so, they introduced several width-dependent coefficients, including
. It is appreciable that
presents the same simple behavior for all the cases shown for those values of the width
W where the coefficient
changes appreciably.
Although the changes shown by are not abrupt, this does not mean that physical quantities cannot exhibit noticeable changes. One of the advantages of using variational methods over an effective Hamiltonian is that we can calculate the expectation values of the spin operators and study the fidelity between quantum states.
Figure 14 shows the calculated expectation values of the spin operators
and
for cylinders whose shell width is in the range where the changes in the inverse participation ratio are appreciable. For
, the changes in the expectation value of
are barely noticeable; in contrast, for
, the polarization in the
z-direction grows noticeably at the expense of the helicity.
To further study the physical behaviors in the range
, we employ the
fidelity between pure quantum states. The fidelity and its derivatives, sometimes called fidelity susceptibility, allow for the detection of quantum phase transitions and
crossovers [
47,
58,
59,
60].
In terms of quantum fidelities, a crossover can be identified as a smooth but rapid variation in the fidelity as the system transitions between different physical regimes. Given a control parameter
, we define the quantum fidelity associated with the
k-th eigenstate as
which quantifies the overlap between the same state computed at slightly different parameter values.
To enhance sensitivity to small deviations near
, we introduce an auxiliary function
defined as:
which diverges as
, allowing for a sharper resolution of crossover behavior.
Figure 15b shows the behavior of
computed for several values of
. It can be observed that, for each choice of
, the function
differs only by a constant vertical shift. This suggests that
can be decomposed as:
where
is a function of the system parameter
W (but constant in
), and
is a constant with respect to
W that depends only on the value of the fidelity increment
.
To obtain
Figure 15c, the curve for
was taken as a reference, and each of the remaining curves—corresponding to larger values of
—were shifted by a constant offset. This procedure revealed that all curves collapse onto a single one.
To perform the alignment, the shift
was computed as the difference between each curve and the reference curve at a fixed, arbitrarily chosen value of the parameter
, according to the following equation:
This confirms that the dependence of
on
is purely additive, supporting the decomposition given in Equation (
27).
The crossover behavior shown by the fidelity reveals the hybridization regime. Other quantities should also reveal this regime; in particular, the expectation values of the spin operator components. In
Appendix D, we include
Figure A2 showing the derivatives of two spin operator components
The derivatives of the expectation values also signal the presence of the hybridization regime, showing a markedly different behavior in the region
, which is consistent with the range where the fidelity indicates a crossover behavior.
There are several quantum information quantities whose behavior separates topological from non-topological states, including the Kitaev–Preskill topological entropy [
27], the entanglement spectrum [
61] and the real-space quantum entropy [
29]. Not all of these quantities are computable for every case, depending on symmetry factors, the availability of the many-particle quantum states and so on. In a previous paper, we used the fact that realistic border states in strong topological insulators have spinors with four components to calculate the Kitaev–Preskill entropy of border states in cylindrical nanowires composed of a single topological insulator. The topological entropy should be a constant for topological states, where the value of the constant depends on the topological properties.
To calculate the topological entropy
, Kitaev and Preskill constructed a disk separated into three equal triangular sectors, where each one subtends an angle of
. By labeling the sectors with the letters
A,
B and
C and tracing out the spatial region outside a given sector, the topological entropy reads as
where
S is the von Neumann entropy of the reduced quantum state given by
D is a spatial region taken from the set
or
,
is the region of the space outside a sector
D, and
is the variational eigenstate that is an approximation for a topological or normal state; for more details, see Refs. [
27,
30].
Figure 16a shows the behavior of the topological entropy as a function of the wavevector
for the inner and outer topological states, using the same color code as in
Figure 14 and
Figure 15a. Except for a dip at
, it is clear that the topological entropy is
for the outer states and
for the inner states. The value of
for the outer states is compatible with the result found in Ref. [
30], while the value obtained for the inner states is not. We believe that this difference is due not to a difference in the topological properties of both sets of states but to the conditions required by the Kitaev–Preskill construction, which imposes that the spatial sectors
and
R must dwell outside the region where the topological state extends. For the outer states, it is a straightforward task to construct the triangular sectors fulfilling this condition. For the inner states, there is no way to do so—at least, if the triangular sectors have the same symmetry axis as the nanowire. Despite this fact, the topological entropy for the topological states is independent of
, while
for the non-topological states depends on
; see the corresponding black curves in
Figure 16a.
The results shown in
Figure 16b support the hypothesis that the inner and outer topological states share the same properties. In Ref. [
30], we proposed a
mode-dependent entropy, which distinguishes between topological and non-topological even better than the Kitaev–Preskill one as its value is far larger for topological states than for non-topological ones.
Following the notation of Ref. [
30], we write the variational state as
and define the mode-dependent reduced density matrix
, with matrix elements
where
is a normalization constant such that
,
is the ring outside
and
is a
matrix. The mode-dependent entropy is the von Neumann entropy of the mode-dependent reduced matrix,
Figure 16b shows the behavior of the mode-dependent entropy,
, as a function of the wavevector
. It is appreciable that the curves corresponding to the inner and outer states collapse into a single curve. The two colored curves now correspond to the upper and lower branches of the topological spectrum, which are distinguishable because they show abrupt decays at the values of
where the energy of the topological states meets with the band of normal states; compare both panels of
Figure 11 and
Figure 16.
5. Discussion and Conclusions
In this work, we do not consider the inclusion of an external magnetic field for the sake of brevity. Nevertheless, the inclusion of a magnetic field will undoubtedly enrich the phenomenology observed in the band structure and the expectation values of spin operator components [
19,
38,
48,
49,
62]. The addition of a uniform and time-independent magnetic field is straightforward through the Peierls substitution. Refs. [
24,
25,
39,
45] have discussed the procedure to include an magnetic field and its treatment using the Rayleigh–Ritz variational method. In the same sense, including electron–electron interactions is a different matter. The application of the Rayleigh–Ritz variational method to few-body problems is cumbersome, but doable; see, for instance, Refs. [
63,
64].
Our results reveal that the outer material in core–shell configurations predominantly governs the features of the low-energy edge states. In contrast, the bulk conduction and valence bands reflect a more intricate interplay between core and shell materials. The topological surface states retain their localization near the outer interface and display spin-momentum locking consistent with the underlying topology, even in the presence of radial discontinuities in material parameters.
Other core–shell structures can also be considered, such as a heterostructure composed of a topological insulator and a non–topological insulator. This would modify the boundary condition at the interface, as well as the specific values of the minigap and the expectation values of the spin operator components. The topological states are expected to remain robust against changes in boundary conditions: they would continue to localize near the interface, decaying strongly into the non–topological region. The situation in which the states and their expectation values are least affected likely appears when the core is made of a non–topological material and the shell is made of a topological one. It is expected that the topological states belonging to the Dirac cone will be only weakly affected; although additional surface states may emerge, as discussed in Ref. [
57].
If the modulation of the spectral properties of a given TI is the purpose of constructing core–shell nanowires, our results draw a mixed scenario. It is clear that the effective azimuthal Fermi velocity,
, is determined by the shell material. Of course, this is only a quantity related to what happens to the spectrum near
. Nevertheless, wires with appropriate material combinations and inner and outer radii show remarkably flat portions on their spectra, as can be seen in
Figure 6c. On the other hand, constructing core–shell nanowires using materials whose spectra have different concavities near
could alter the global properties of the spectra, resulting in overlapping conduction and valence bands.
The presented numerical evidence indicates that energy gaps between topological states are weakly dependent on shell thickness beyond a threshold width, confirming the robustness of surface states against moderate geometric perturbations. Additionally, we showed that the inverse participation ratios, spin expectation values and quantum fidelity measures track and quantify the emergence and stability of edge-localized states in hollow TI cylinders. These tools allow for the identification of crossover regimes between hybridized and uncoupled surface modes as the shell width decreases. For uncoupled surface modes, we understand the usual picture employed to derive effective Hamiltonians; i.e., a surface state that decays exponentially away from the surface to which it is attached. For hybridized modes, we understand that the probability density at points near the middle of the shell width is comparable with that close to the borders of the shell, and is not negligible.
The understanding of the hybrid regime is still an open question from both theoretical and experimental points of view. Ref. [
49] claimed that the hybrid regime would allow for measurement of the Quantum Spin Hall effect in scalable and controllable platforms. The authors found that the minigap energy presents an oscillating behavior as a function of the thickness of doped
and
thin films. In any case, interference and other phenomena appear when a magnetic field is applied and in the transport regime [
48], both of which are beyond the scope of the present work. Although we did not explore the behavior of the minigap for hollow core wires, the comparison with the behavior observed on films would not be direct.
The variational approach implemented here proved to be flexible and accurate in capturing features of topological states under complex boundary and interface conditions. Unlike effective surface models, our method naturally incorporates bulk–boundary interplay, allowing for the direct computation of observables such as spin currents and entanglement measures. In particular, our results show that the spin current components depend on the shell width, opening the possibility of exploiting this dependence to tune the spin orientation of surface states. For example, in the region , one can selectively enhance either the azimuthal component or the axial component at a fixed momentum by adjusting the shell geometry.
The conversion between spin and charge currents in topological insulators has been identified as the key mechanism enabling their use in spintronic applications [
65]. It is likely that favoring
over
modifies the spin–charge conversion process, due to changes in the probability density function of the edge states in the hybridized regime. Experimental evidence indicates that, in a thin film composed of a topological insulator, the torque per unit charge is larger than that associated with other studied mechanisms [
62]. To analyze this possibility quantitatively, it is necessary to investigate the non–equilibrium regime and explicitly consider the interface between the topological insulator and a conductor, as discussed in Ref. [
65].
The scaling behavior revealed by the fidelity is remarkable in its own way. Other quantities also indicate the hybridization regime responsible for the crossover behavior, as shown by the first derivatives of the expectation values of the spin operator components. At present, we lack the physical insight to formulate a quantitative relationship between the scaling behavior revealed by the fidelity and expectation values of physical quantities.
The topological states in hollow nanowires should have the same topological properties. Remarkably, the confirmation of this fact comes from results obtained for the mode-dependent entropy, which is independent of the basis set employed in the variational calculations; see Ref. [
30]. Moreover, the mode-dependent entropy is more flexible to implement than the Kitaev–Preskill entropy in finite domains with topological states distributed on different regions of the domain.