Mexican-hat-shaped quartic dispersion manifests itself in certain families of single-layer two-dimensional hexagonal crystals such as compounds of groups III–VI and groups IV–V as well as elemental crystals of group V. A quartic band forms the valence band edge in various of these structures, and some of the experimentally confirmed structures are GaS, GaSe, InSe, SnSb, and blue phosphorene. Here, we numerically investigate strictly one-dimensional and quasi-one dimensional (Q1D) systems with quartic dispersion and systematically study the effects of Anderson disorder on their transport properties with the help of a minimal tight-binding model and Landauer formalism. We compare the analytical expression for the scaling function with simulation data to distinguish the domains of diffusion and localization regimes. In one dimension, it is shown that conductance drops dramatically at the quartic band edge compared to the quadratic case. As for the Q1D nanoribbons, a set of singularities emerge close to the band edge, suppressing conductance and leading to short mean-free-paths and localization lengths. Interestingly, wider nanoribbons can have shorter mean-free-paths because of denser singularities. However, the localization lengths sometimes follow different trends. Our results display the peculiar effects of quartic dispersion on transport in disordered systems.

Several families of two-dimensional (2D) materials have been identified since the first isolation of a graphene monolayer, and they have received considerable attention due to their unusual and unique properties. They are not only interesting for fundamental science but are also promising candidates for next-generation devices.1–12 Many of the novel properties are related to their electronic band structures, some of which are unprecedented. Mexican hat-shaped (MHS) quartic energy dispersion is one of them. Graphene like honeycomb lattices of group V elements and group III–VI and group IV–V compounds in hexagonal P 6 ¯ m 2 symmetry display MHS quartic dispersion in their valance bands (VBs). In some of these structures, the valence band maximum is formed by the MHS quartic band.13–46 Owing to their unusual band structure, these materials are reported to have peculiar magnetic and ferroelectric phases, very high thermoelectric efficiencies, and the potential to be useful for applications like water splitting, optoelectronics, photonics, and thermoelectrics.21,30–36,45,47–51 On the experimental side, single and few layers of quartic materials such as gallium based compounds (GaS, GaSe, GaTe),30,52–56 InSe,57–59 blue phosphorene,60 and SnSb61 have been synthesized and their distinctive properties have been reported. It is worth noting that similar band dispersion has been studied in the context of optical lattices, with implications about chiral spin liquid, and that the tight binding approach has been used to show the absence of Bose condensation on these lattices.62–64 

A key feature in the electronic structure of MHS quartic dispersion is the strong (inverse-square-root) Van Hove singularity with divergent DOS at the VB edge. Such strong singularity and divergent DOS is uncommon in two dimensions. In addition to the strong singularity, the pristine transmission spectrum displays a step-like behavior at the band edge. A stepwise transmission spectrum is a characteristic feature of 1D systems, and in two dimensions, it is observed only in quartic materials. These peculiarities of quartic dispersion, together with finite amounts of disorder, are potentially responsible for strong scatterings that could alter the electronic properties significantly. The effects of disorder on the transport properties of quartic materials have not been addressed in the literature before. Here, using nonequilibrium Green’s function (NEGF) methodology, we study quantum transport properties around the quartic VB edge. We consider large-scale quartic systems by taking advantage of a minimal tight-binding (TB) model,21,62,63 and thus, we are able to determine the transport length scales of these systems in the presence of the Anderson disorder. The effects of the Anderson disorder on conductance ( G) are numerically investigated in the case of strictly one-dimensional (1D) and quasi-one dimensional (Q1D) systems at the weak disorder limit. Our results show that there is a dramatic suppression of transmission at the quartic band edge, which cannot be understood using the conventional effective mass approaches or any multiple band models.

The rest of this paper is organized as follows. In Sec. II, the computational models of interest are introduced. The effects of disorder on conductance are separately discussed in Sec. III for both 1D and Q1D systems. Section IV summarizes our conclusions.

For two-dimensional hexagonal lattices of group-V elements, it was previously shown that a second nearest neighbor TB approximation is suitable to study the quartic dispersion21 
(1)
where c i ( c i ) annihilates (creates) an electron at lattice site i, t 1 and t 2 are the first and the second nearest neighbor hopping parameters, respectively, and accordingly the summations run over the first and the second nearest neighbors.
The solution of above Hamiltonian is familiar from the graphene band structure,65 which can be expressed as E ± ( k ) = ± t 1 3 + f ( k ) t 2 f ( k ), with f ( k ) = 2 cos ( 3 k y a ) + 4 cos ( 3 k y a / 2 ) cos ( 3 k x a / 2 ) , k and a being the wave vector and the lattice constant, respectively. Different from the half-filled bands of π-orbitals of graphene, we are interested in completely filled bands (e.g., replacing carbon with nitrogen, see Ref. 21). In this sense, valance band maximum (VBM) is set to that of the highest occupied electronic state by assuming two electrons per atom in this study. For ξ = t 2 / t 1 > 1 / 6, the above dispersion yields a quartic band around the VB edge, such that
(2)
We choose the TB parameters to reproduce the topmost VB of hexagonal nitrogene because it is well separated from the rest of the bands.16,21,27 That is, t 1 = 6.1 eV and t 2 = 1.27 eV are chosen, which corresponds to ξ = 0.21. In Fig. 1(a), the quartic band for nitrogene as obtained from our TB approximation is shown to have a good agreement with that obtained from density functional theory (DFT) calculations, especially for low energy holes.21 For the sake of simplicity, we refer the energy values ranging from the minimum of the Mexican hat at the center of the BZ to its maximum, namely, the VBM, as the MHS energy region. This is not meant to exclude the energies close to these values from being quartic in dispersion, but to emphasize that in this energy range, there are four solutions at a given direction in k-space for the 2D structure. In the case of hexagonal nitrogene, the MHS energy region lies within 0.47 eV < E E VBM < 0.
FIG. 1.

The top panel shows the fitting curve (red dots) of the topmost VB of DFT bands (black circles), which is defined by a ring of radius k 0 0.78 Å 1 and a well of E 0 0.47 eV at the Γ-point. The bandwidth of the upper VB nearly equals to 2.6 eV. The lower panel illustrates the top view of hexagonal lattice structure in which the unit cells, the lattice constant a, the basis vectors a 1 = a ( 3 / 2 , 1 / 2 ) and a 2 = a ( 3 / 2 , 1 / 2 ) are depicted. The widths of the NRs denoted by the number of N are separately shown for a ZNR with N = 8 and an ANR with N = 12. The device lengths are taken as L = M a for a ZNR and L = M a 3 for an ANR, respectively, where M is the number of unit cell in the Q1D system.

FIG. 1.

The top panel shows the fitting curve (red dots) of the topmost VB of DFT bands (black circles), which is defined by a ring of radius k 0 0.78 Å 1 and a well of E 0 0.47 eV at the Γ-point. The bandwidth of the upper VB nearly equals to 2.6 eV. The lower panel illustrates the top view of hexagonal lattice structure in which the unit cells, the lattice constant a, the basis vectors a 1 = a ( 3 / 2 , 1 / 2 ) and a 2 = a ( 3 / 2 , 1 / 2 ) are depicted. The widths of the NRs denoted by the number of N are separately shown for a ZNR with N = 8 and an ANR with N = 12. The device lengths are taken as L = M a for a ZNR and L = M a 3 for an ANR, respectively, where M is the number of unit cell in the Q1D system.

Close modal

In this study, starting from a toy model, namely, monatomic chain, we investigate the electronic and transport properties of the Q1D nitrogene nanoribbons (NRs) with zigzag and armchair edges (ZNR and ANR) in the presence of uncorrelated disorders. The hexagonal lattice structure and the unit cells of these Q1D systems are illustrated in Fig. 1(b). The puckered geometry of the lattice does not play a role for the purposes of this study, therefore, it is disregarded. As for short-range disorder, the Anderson disorder is introduced by adding the term H A = i ϵ i c i c i to the Hamiltonian. Here, ϵ i represents the on-site energy, which randomly fluctuates in the energy interval [ W/2, W/2] with a fixed disorder strength W = 25 and 250 meV for the chain and ribbon geometries, respectively. The overall contribution of the on-site potential energies is set to zero, that is, i ϵ i = 0. In all cases, Eq. (1) with additional on-site terms is numerically solved for various system lengths, L.

Conductance values are computed within the Landauer approach using NEGF formalism. The system is partitioned as the central device region ( C), and the left/right electrodes ( L / R). The electrodes have semi-infinite geometry and are made-up of the pristine form of the same material. The self-energies due to coupling to the electrodes are computed using the surface Green’s functions.66 Scattering events are allowed to take place only in the C region for which the retarded Green function can be expressed as G C C r ( E ) = [ ( E + i 0 + ) I H C C Σ L r Σ R r ] 1, with 0 + being an infinitesimal positive number, I the identity matrix, H C C is Hamiltonian of the C region, Σ L / R r = H C L / C R G L L / R R r , 0 H L C / R C are the self-energy matrices with the free Green’s functions G L L / R R r , 0 = [ ( E + i 0 + ) I H L L / R R ] 1 of the isolated L and R reservoirs. The level-broadening caused by system-electrode coupling is Γ L / R = i ( Σ L / R Σ L / R ), and the transmission amplitude is given by
(3)
At zero temperature, conductance is given by G ( E ) = G o T ( E ), where G o = 2 e 2 / h is the quantum of conductance, and the DOS is computed as ρ ( E ) = 1 / π Im [ Tr [ G r ( E ) ] ]. Further details of which can be found elsewhere.67,68

System sizes should be large enough for a reliable analysis of the localization regime. In order to be able to simulate large systems a recursion–decimation algorithm is implemented.69 As for the statistical average of conductance, simple G averaging over a set of samples may not converge toward any meaningful value since G does not follow Gaussian statistics in the localized regime.70–73 Therefore, a logarithmic average ln G over 100 samples at each system length L is considered in this study. In this way, statistical average of conductance is obtained via G av = exp ln G , which is expected to give a better statistical result.72,73

Mean-free path ( mfp) and localization length ( loc) are extracted by fitting simulation data to
(4)
where N ch being the number of channels, and α is a fitting parameter. In the diffusion regime, the conductance scales as G av 1 / L, obeying Ohm’s law. On the other hand, it decays exponentially with L in the Anderson localization regime.74 Interrelation between transport length scales in 1D systems is set by the Thouless relation based on the random matrix theory, which conditions that loc = ( η[ N ch-1]+2)/2 mfp,73 where η = 1 in the absence of external magnetic field, simplifying to
(5)
Consistency of numerical results is confirmed using Eq. (5).
Effects of disorder on the transport characteristics of structures with quartic dispersion have not been addressed in the literature before. Therefore, we first study these systems in their simplest realizations, namely, the strictly 1D case. For this purpose, we use a monatomic chain within the second nearest neighbor empirical TB model. The corresponding dispersion relation is E ( k ) = 2 t 1 cos ( k a ) + 2 t 2 cos ( 2 k a ), with k [ π / a , π / a ]. Such a dispersion relation can be expanded around k = 0 as E / t 1 2 ( 1 + ξ ) ( k a ) 2 ( 1 + 4 ξ ) + ( k a ) 4 ( 1 + 16 ξ ) / 12. As a special case, ξ = 1 / 4 leads to a purely quartic dispersion, i.e., E k 4. For ξ < 1 / 4, a quartic band with MHS emerges [Fig. 2(d)]. In this study, t 1 = 1 eV and t 2 = 1 / 3 eV, corresponding to ξ chain = 1 / 3, are utilized to represent the quartic dispersion, whereas for the quadratic chain, t 1 = 1 eV and t 2 = 0 eV, corresponding to ξ = 0, are set [Fig. 2(a)]. Band dispersion, density of states (DOS, ρ), and zero temperature ballistic conductance ( G ballistic) of the quadratic chain are shown in Figs. 2(a)2(c), respectively. The corresponding plots for the quartic dispersion are given in Figs. 2(d)2(f). Both numerical (black) and analytical results (red-dashed) for DOS are shown in agreement with each other, as shown in Figs. 2(b)2(e). The analytical form of the DOS is given with
(6)
for the quartic case, where ξ < 1 / 4, φ = E / 4 t 2, γ = 1 / 4 ξ, and χ = γ 2 + φ + 1 2. It is evident that the quartic dispersion gives rise to a much stronger Van Hove singularity at E VBM (see Fig. 2) together with an additional singularity at E E VBM = t 1 ( 1 + 4 ξ ) 2 / 4 ξ . Stepwise G ballistic emerges in both cases, which is the characteristic of 1D systems. G ballistic is doubled for the quartic chain at the MHS energy region, whose width depends on t 1 and t 2, and it is 0.083 t 1 in the present case.
FIG. 2.

In the first column, the band structure (a), DOS per atom (b), and G ballistic (c) for the quadratic chain within ( E E VBM ) / t 1 [ 0.5 , 0.1 ] are plotted. The second column shows the same plots for the quartic chain with the same scales as in the first column. The insets in each plot exhibit the full spectra of the chains.

FIG. 2.

In the first column, the band structure (a), DOS per atom (b), and G ballistic (c) for the quadratic chain within ( E E VBM ) / t 1 [ 0.5 , 0.1 ] are plotted. The second column shows the same plots for the quartic chain with the same scales as in the first column. The insets in each plot exhibit the full spectra of the chains.

Close modal

For investigating the effects of disorder, we restrict ourselves to the energies around the VB edge. In the chosen energy interval ( E E VBM ) / t 1 [ 0.25 , 0 ], G av is plotted as a function of E and L in Fig. 3(a) for the quadratic (top panel) and quartic (bottom panel) chains. G av is computed for lengths ranging from 0 up to 5 × 10 4 a. At short distances, G av decreases very fast at the quartic edge and the step exhibits a rounded shape. At energies away from the VBM, conductance decreases relatively slowly, in fact slower than the corresponding energies of the quadratic band. Simulation data of the quadratic and quartic chains are fitted with Eq. (4), and mfp and loc are shown in Figs. 3(b) and 3(c), respectively. It should be noted that the simulation data above G ballistic / 2 are used for the mfp fitting processes of both chains, which will be discussed in further detail in Sec. III D. As it is shown in Fig. 3(b), mfp and loc of quadratic chain are equal, in agreement with the Thouless relation, i.e., loc / mfp = 1 for N ch = 1. In a similar fashion, mfp and loc curves for the quartic chain are displayed in Fig. 3(c). As expected, mfp of the quartic chain equals to loc except for the quartic edge with N ch = 2, where the ratio turns out to be loc/ mfp 1.5 as suggested by the Thouless relation. Interestingly, mfp ( loc) of the quartic chain is approximately six (four) times smaller than that of the quadratic chain at the quartic edge [see insets of Figs. 3(b) and 3(c)]. In the MHS energy region with a disorder strength of W = 25 meV, we have mfp = 700 a and loc = 1010 a at E = E VBM 0.05 t 1 [see the inset of Fig. 3(c)]. In comparison, we compute transmission at E = E VBM t 1 with a much stronger disorder ( W = 250 meV) and find mfp = 1983 a and loc = 1996 a, which are several times larger than those in the MHS energy region with weaker disorder. mfp and loc are almost equal because N ch = 1 at these energies. This comparison in another illustration of the role of strong singularity in transport properties of quartic systems.

FIG. 3.

(a) G av as a function of E and L for quadratic (top panel) and quartic (bottom panel) chains with W = 25 meV. mfp (solid curves) and loc (dashed-dotted curves) are plotted for quadratic chain in (b) and quartic chain in (c). The zoom-ins in (b) and (c) show the transport length scales within ( E E VBM ) / t 1 [-0.05, 0]. In (b) and (c), the analytical solutions of mfp (gray solid curve) and loc (black dashed curve) are given for comparison. A pink solid curve with unfilled diamonds in (c) corresponds to the Thouless relation.

FIG. 3.

(a) G av as a function of E and L for quadratic (top panel) and quartic (bottom panel) chains with W = 25 meV. mfp (solid curves) and loc (dashed-dotted curves) are plotted for quadratic chain in (b) and quartic chain in (c). The zoom-ins in (b) and (c) show the transport length scales within ( E E VBM ) / t 1 [-0.05, 0]. In (b) and (c), the analytical solutions of mfp (gray solid curve) and loc (black dashed curve) are given for comparison. A pink solid curve with unfilled diamonds in (c) corresponds to the Thouless relation.

Close modal
It is also possible to derive analytical expressions for mfp and loc using Fermi’s golden rule and Thouless relations.75 Assuming that the Bloch wave has equal probability weight on each atom in the unit cell, one can write
(7)
where σ ϵ = W / 12 is the standard deviation of the uniform random distribution, uc is the width of the unit cell along the transport direction, N uc is the number of atoms in the unit cell, N ch ( E ) is the number of channels, and ρ uc ( E ) is the density of states per unit cell. For both the quadratic and quartic chains, analytical and numerical results are in agreement as shown in Figs. 3(b) and 3(c).

The electronic and transport properties of the Q1D quartic systems are investigated for ZNR and ANR of nitrogene with the TB parameters of t 1 = 6.1 eV and t 2 = 1.27 eV. Widths are chosen as N = 10, i.e., N atom = 20 in a unit cell for both cases. The full spectra of considered bands, ρ ( E ) per area, and G ballistic as a function of dimensionless energy ( E E VBM ) / t 1 are shown in the upper panels of Fig. 4. Valence band edges of the corresponding spectra are displayed in the lower panels. In the band structure of ZNR, a family of quartic bands appears around the VBM, where some of them exhibit crossings (see the bottom-left of the left panel of Fig. 4). In two dimensions, the critical wave-vectors that form the band edge are degenerate and nonisolated, which give rise to a strong Van Hove singularity. In Q1D NRs, the critical wave-vectors are isolated and nondegenerate, whereas Van Hove singularities are still strong because of the reduced dimension. In addition, there exists numerous strong singularities in the entire spectrum, which are characteristic in Q1D structures. A distinguishing character of ρ in Q1D quartic materials from those of other materials is the emergence of excessively dense singularities at the MHS energy region (see Fig. 4). Both ANR and ZNR display rapid changes in their G ballistic close to the VBM (Fig. 4). In ZNR, top bands are more dispersive than in ANR, which is because of the narrower Brillouin zone of ANR. As a result, G ballistic values are larger for ZNR in these energies. Interestingly, we observe formation of nearly flatbands in ZNRs, close to the band edge. These nearly flatbands are formed by hybridization of quartic bands due to the edges. Flattening gives rise to even stronger peaks in the DOS.

FIG. 4.

Energy band diagram, density of states, and pristine conductance for ZNR (left) and ANR (right). In the left group, energy bands E ( k ) (left), density of states per area ρ ( E ) (middle), and ballistic conductance G ballistic (right) are shown with their zoom-ins to the band edge for a ZNR with N = 10. In the right group, the same spectra with the same scales are plotted for an ANR with N = 10.

FIG. 4.

Energy band diagram, density of states, and pristine conductance for ZNR (left) and ANR (right). In the left group, energy bands E ( k ) (left), density of states per area ρ ( E ) (middle), and ballistic conductance G ballistic (right) are shown with their zoom-ins to the band edge for a ZNR with N = 10. In the right group, the same spectra with the same scales are plotted for an ANR with N = 10.

Close modal

It can be expected that the increase in the NR’s width causes both the DOS and G ballistic spectra to evolve into those of the 2D system. Figure 5 exhibits the evolution of ρ ( E ) per area and G ballistic for different system widths, N, in the case of pristine ZNRs. Here, the corresponding 2D spectra are given by black curves. It can be observed in Fig. 5-top that ρ ( E ) approaches to that of the 2D quartic system as N is systematically increased. Similarly, G ballistic of NRs approach to that of 2D quartic systems with increasing N, which is shown in the bottom panel of Fig. 5.

FIG. 5.

Dependence of pristine DOS and conductance on ribbon width. DOS per area ρ ( E ) (top) and G ballistic (bottom) for pristine ZNRs with N = 10 (blue), 30 (red), 50 (green), 70 (burgundy), and 90 (orange). For comparison, the corresponding 2D spectra are depicted with the black curves in each plot, and 2D G ballistic is given in arbitrary units.

FIG. 5.

Dependence of pristine DOS and conductance on ribbon width. DOS per area ρ ( E ) (top) and G ballistic (bottom) for pristine ZNRs with N = 10 (blue), 30 (red), 50 (green), 70 (burgundy), and 90 (orange). For comparison, the corresponding 2D spectra are depicted with the black curves in each plot, and 2D G ballistic is given in arbitrary units.

Close modal

The edge shape is a determinant factor for the electronic structure and hence for the transport regimes. The effects of short-range Anderson disorder are studied for the Q1D systems for which the numerical calculations are performed for the ribbon widths of N = 10 for both edge shapes. The evolution of conductance is examined for various device lengths at energies close to the VBM ( ( E E VBM ) / t 1 [ 0.12 , 0 ]) for a relatively weak realization of disorder ( W = 250 meV). Length dependent G av values of ZNR and ANR are plotted in Fig. 6(a). Although G ballistic values are close within this range, length dependent G av decays much faster for the MHS energy region. This behavior is the same for both ANR and ZNR. The origin of this behavior can be understood by examining the pristine DOS at these energies. As it is shown in Figs. 6(b) and 6(c), the pristine DOS (red curves) has multiple singularities at the energies, where G av drops suddenly. The computed mfp values are shown in the same plots (blue curves). We note that simulation data around G ballistic / 2 are used for fitting the diffusion formula, cf. Eq. (4). As it is expected, the dips in mfp correspond to the peaks in the DOS. It is much probable for the particle to scatter due to the higher DOS. In the MHS energy region, where the DOS values are high, mfp of ZNR (ANR) always remains below 300 a (400 a). In the first conduction step, it is lower than 10 a and at the band edge, mfp converges to zero as ( E E VBM ) / t 1 0. On the contrast, mfp 3200 a at energies outside the MHS energy region, namely, for ( E E VBM ) / t 1 0.1 for ANR.

FIG. 6.

Dependence of G av on system length is shown in (a) for zigzag (top panel) and armchair (bottom panel) edge shapes with N = 10. Disorder strength is W = 250 meV. Mean-free-path ( mfp) (blue solid curves) and pristine DOS per unit area (red solid curves) are shown for ZNR and ANR in (b) and (c), respectively. In both cases, a family of strong singularities due to quartic dispersion are observed for E > E VBM 0.07 t 1. The mean-free-path is much longer ( mfp 3200 a) at energies away from the singularities. In panel (d), it is observed that the fitted localization length ( loc, red solid curve) is in a good agreement with that obtained from the Thouless relation (dashed black curves).

FIG. 6.

Dependence of G av on system length is shown in (a) for zigzag (top panel) and armchair (bottom panel) edge shapes with N = 10. Disorder strength is W = 250 meV. Mean-free-path ( mfp) (blue solid curves) and pristine DOS per unit area (red solid curves) are shown for ZNR and ANR in (b) and (c), respectively. In both cases, a family of strong singularities due to quartic dispersion are observed for E > E VBM 0.07 t 1. The mean-free-path is much longer ( mfp 3200 a) at energies away from the singularities. In panel (d), it is observed that the fitted localization length ( loc, red solid curve) is in a good agreement with that obtained from the Thouless relation (dashed black curves).

Close modal

The loc values are obtained by fitting the G av at L mfp. The consistency of mfp and loc is checked by using the Thouless relation [Eq. (5)]. Within the MHS energy region, the localization lengths are short, and, therefore, they can be obtained within reasonable sizes of simulated devices. However, for lower energies, this is not the case. Figure 6(d) demonstrates loc for the ZNR (red solid curve) and the ANR (blue solid curve) in which loc approaches zero for both NRs as the energy goes to E VBM. It is clear from Fig. 6(d) that for both NRs, Thouless relations (black dashed curves) are in good agreement with the fitted data, especially within the MHS energy region.

In order to reveal the effects of device width N on the transport length scales, the Q1D systems with N = 10 are compared to those with N = 20 for both edge shapes. Figure 7(a) shows pristine DOS per area for the ZNRs with N = 10 (red solid curve) and N = 20 (blue solid curve) at the MHS energy region. The spectra explicitly display that N = 20 case has more singularities compared to the ZNR with N = 10 at the band edge. Multiple singularities exist around E VBM for both sizes. Similarly, DOS for the ANR with N = 10 (pink solid curve) and N = 20 (orange solid curve) are displayed in Fig. 7(d). It is observed that the density of singularities increase with width. On the other hand, when the number of atoms in the unit cells are equal, the number of singularities in ZNRs is larger than in ANRs simply because the number of bands within the MHS energy region is larger in ZNRs (see Fig. 4). Dense singularities in the DOS give rise to shorter mfp independent of the edge termination.

FIG. 7.

Width dependence of DOS and transport length scales for both edge shapes. Pristine DOS per area for ZNRs with N = 10 (red) and N = 20 (blue) are plotted in (a). Corresponding mean-free-path ( mfp) and localization lengths ( loc) is shown in (b) and (c), respectively. Black dashed curves exhibit the agreement in terms of the Thouless relation. At W = 250 meV, mfp and loc at the 2D limit are given with gray solid curves for comparison. The same quantities are plotted for ANR with N = 10 (pink) and N = 20 (orange) in (d), (e), and (f).

FIG. 7.

Width dependence of DOS and transport length scales for both edge shapes. Pristine DOS per area for ZNRs with N = 10 (red) and N = 20 (blue) are plotted in (a). Corresponding mean-free-path ( mfp) and localization lengths ( loc) is shown in (b) and (c), respectively. Black dashed curves exhibit the agreement in terms of the Thouless relation. At W = 250 meV, mfp and loc at the 2D limit are given with gray solid curves for comparison. The same quantities are plotted for ANR with N = 10 (pink) and N = 20 (orange) in (d), (e), and (f).

Close modal

In quasi-1D structures, the system width affects mfp by means of more than one mechanisms. First, considering the ribbon as a dimensionally reduced form of the two-dimensional structure, edges are sources of scatterings, hence mfp is expected to reduce with ribbon width. Indeed, it was reported that mfp increases with width in edge disordered graphene nanoribbons.76,77 Similarly, in graphene NRs with oxygen functionalization, mfp was shown to increase with the ribbon width.78 However, in quartic NRs, we do not observe such an increase in mfp with width. On the contrary, mfp can be greater for narrower ribbons at some energy values. This can be understood in terms of another factor that determines mfp, namely, the width dependence of DOS at the MHS energy region. Recalling Fermi’s golden rule, the scattering rate increases with the DOS, which is the reason of the dips in the computed mfp [see Figs. 6(b), 6(c), and 7(b)7(e)]. In quartic NRs, the number of Van Hove singularities depends on the ribbon width, namely, narrower the ribbon, less the number of singularities. As a result, mfp of wider ribbons are suppressed due to denser singularities. The trade-off between these two mechanisms is the main factor that determines width dependence of mfp.

The localization length also depends on these factors. Additionally, the number of channels has an important role as it was described in Eq. (5). The number of channels increases with width, therefore, unlike mfp, loc increases with width for both ANRs and ZNRs [Figs. 7(c)7(f)]. Comparing loc of ANRs and ZNRs, those of ZNRs are considerably longer, whereas their mfp are similar. This difference is also rooted in the difference between the number of channels. ANRs have wider unit cells, and narrower Brillouin zones; therefore, the quartic bands are less dispersive and the number of channels are less than ZNRs. Further details in the energy dependent loc of quartic NRs can be understood by investigating the details in the DOS and G ballistic.

An exploration of transport in the 2D limit is possible by assuming a tubular geometry with an arbitrarily large radius with the transport direction parallel to the tube’s axis. In this limit, the effects of curvature on the onsite and off diagonal terms of the Hamiltonian can be ignored. In the presence of the Anderson-type disorder, it is possible to find an exact expression for the mean-free-path in tubular geometries of hexagonal monolayers. Thanks to the symmetries in these structures, Bloch waves have equal weights on each atom, and one can use Eq. (7) for obtaining the mean-free-path.75 As tube radius increases, DOS converges to that of the 2D structure. In the MHS region, it is given as79 
(8)
where α = 1.155 eV Å 4 for nitrogene. The transmission spectrum can be approximated with a step function to write N ch , 2 D ( E ) = 2 k c a o / π, with k c = 0.798 Å 1,21 and mfp increases linearly with | E |. At the same disorder strength as those in quasi-1D structures, mfp of the 2D quartic system is plotted in Figs. 7(b)7(e) for comparison. Owing to the high DOS near E VBM, mfp of the 2D system mostly stays below those of NRs and goes to zero at the quartic band edge.
Having obtained the mean-free-path from Eq. (7), which is further compared with numerical results for tubular geometries with a very large radius ( N = 250), the 2D conductivity, σ 2 D = G o mfp N ch / w, can be used together with the 2D generalization of the Thouless relation, which reads80 
(9)
2D loc is found to be much larger compared to those of NRs due to the larger number of channels. The loc of the 2D quartic system suddenly diverges in the MHS energy region, but it goes to zero at the VB edge, as shown in Figs. 7(c)7(f). The crossover from quasi-one dimension to two dimensions demands further investigation because of the degeneracy at the 2D band edge; however, the obtained results suggest that a numerical examination of the 2D limit within the disorder strengths used in the present study is out of reach using the present computational approach, but order- N approaches like the Kubo–Greenwood formalism could be useful.81 
For analyzing the transport regimes in the limits of small and large L, the scaling function
(10)
of the one-parameter scaling theory82–84 is used. In the case of disordered systems, the metal-insulator transition (MIT) can be studied through β as a function of ln( g), where the dimensionless conductance is g = G av / G o. There is a fixed point above (below) which the sytem scales toward a metal (an insulator) as the system size is increased. However, the system dimension plays a crucial role in our conceptual understanding of MIT. The 2D system is the marginal case,83 but the systems with a dimension smaller than d < 2 undergo to the Anderson localization.82 In addition, the Q1D quartic systems have the multiple strong singularities at the quartic band edge, pointing to the strong localization regime. For these reasons, we make use of the β-function to reveal the points where the diffusion regime ends and the localization regime starts. The two different forms of the β-function given below render easy to distinguish the transport regimes.
After some algebra, the β-function can be expressed in terms of corresponding length scales as
(11)
The functions β dif and β loc depend on mfp and loc, which are to be extracted from fitting. It is also possible to investigate the scaling behavior directly from the simulation data without referring to any fitted parameter as85 
(12)
Comparing β [Eq. (11)] with β data [Eq. (12)] for the systems of interest, the transport regimes and their crossover are easy to distinguish. Moreover, the agreement between fitted transport length scales ( mfp and loc) and the simulation data is possible to test.

In Figs. 8(a) and 8(b), we compare β with β data for the strictly 1D systems, namely the quadratic and quartic chains, as functions of L for ( E E VBM ) / t 1 = 0.017. The solid curves represent the β function ( β dif and β loc), which are obtained using mfp and loc [which were obtained by fitting simulation data to Eq. (4)]. The data points represent β data ( β dif data and β loc data), which do not rely on any fittings but are obtained solely from the simulation data. At short lengths, there is a perfect agreement between β dif and β dif data [Figs. 8(a)8(b), red curves and red dots]. This also verifies that the mfp fit is satisfactory. The mfp values are indicated by vertical dashed curves ( mfp = 1903 a for quadratic chain and mfp = 308 a for quartic chain). For L > mfp, the agreement between β dif and β dif data is lost, namely, the simulation data do not obey the diffusion equation any more. This suggests that the fitting procedure for mfp should use only the data from short distances (where G av G ballistic / 2), as was done in this work. At long distances ( L mfp), we observe that the β loc data converges to β loc (blue curves and dots), which confirms that the fitted loc is correct. We should note that the β loc data converges to β loc at distances longer than loc. The intermediate distances where β data is not in agreement with β dif or β loc are considered the crossover region. The disagreement does not imply that mfp or loc are incorrect. The validity of loc is confirmed by the fact that the slope of β loc matches perfectly with β loc data at long distances.

FIG. 8.

Scaling function is used to distinguish transport regimes. In (a) and (b), the scaling functions are plotted for quadratic and quartic chains at ( E E VBM ) / t 1 = 0.017. They are compared with data [cf. Eq. (11) and Eq. (12)]. Similar analysis is shown in (c) and (d) for ZNR and ANR with N = 10, where energy is set to ( E E VBM ) / t 1 = 0.0023.

FIG. 8.

Scaling function is used to distinguish transport regimes. In (a) and (b), the scaling functions are plotted for quadratic and quartic chains at ( E E VBM ) / t 1 = 0.017. They are compared with data [cf. Eq. (11) and Eq. (12)]. Similar analysis is shown in (c) and (d) for ZNR and ANR with N = 10, where energy is set to ( E E VBM ) / t 1 = 0.0023.

Close modal

The same scaling analysis is performed for quartic NRs as well. In Figs. 8(c) and 8(d), β and β data are shown for ZNR( N = 10) and ANR( N = 10) at ( E E VBM ) / t 1 = 0.0023, respectively. Transport length scales are found to be mfp = 12 a ( 28.8 a) and loc = 842.26 a ( 51.43 a) for the ZNR (ANR). As it was the case in strictly 1D systems, β dif data is in very good agreement with β dif at short L up to mfp. At L mfp, β loc data converges to β loc, showing that obtained loc is consistent. We again observe that diffusion regime survives only at relatively short distances in quartic NRs. Although N ch can be large, loc is also relatively short because of multiple strong singularities close to the VBM. These findings are confirmed by the β-function analysis of simulation data as well.

In summary, a minimal TB model is combined with Landauer formalism, which is used for studying the transport properties of disordered quartic systems. The scaling theory is utilized in a way to distinguish the diffusion and localization regimes and to find the relevant length scales. A comparison of a strictly 1D quartic chain against its quadratic counterpart at a given disorder strength shows that mfp and loc of the quartic chain are much smaller in the MHS energy region. In quasi-1D quartic ribbons, mfp and loc are considerably short because of multiple strong Van Hove singularities, which are denser and stronger in quartic systems compared to quadratic ones. Interestingly, mfp can be longer in narrower NRs compared to the wider NRs, which is because the number of singularities increases with width in the MHS energy region. Exploration of the 2D limit suggests very long localization lengths and requires further numerical studies. Transport length scales of randomly distributed defects can also be studied using the same methodology. Dependence of conductivity on the sample quality and the doping rate could be interpreted using the results presented in this study.

This work was supported by The Scientific and Technological Research Council of Turkey (TÜBİTAK) under 1001 Grant Project No. 119F353. Support from the Air Force Office of Scientific Research (AFOSR, Award No. FA9550-21-1-0261) is also acknowledged.

The authors have no conflicts to disclose.

Mustafa Polat: Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Hazan Özkan: Investigation (supporting); Validation (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Hâldun Sevinçli: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article.

1.
K. S.
Burch
,
D.
Mandrus
, and
J.-G.
Park
,
Nature
563
,
47
(
2018
).
2.
S.
Barraza-Lopez
,
B. M.
Fregoso
,
J. W.
Villanova
,
S. S. P.
Parkin
, and
K.
Chang
,
Rev. Mod. Phys.
93
,
011001
(
2021
).
3.
S. Z.
Butler
,
S. M.
Hollen
,
L.
Cao
,
Y.
Cui
,
J. A.
Gupta
,
H. R.
Gutiérrez
,
T. F.
Heinz
,
S. S.
Hong
,
J.
Huang
,
A. F.
Ismach
,
E.
Johnston-Halperin
,
M.
Kuno
,
V. V.
Plashnitsa
,
R. D.
Robinson
,
R. S.
Ruoff
,
S.
Salahuddin
,
J.
Shan
,
L.
Shi
,
M. G.
Spencer
,
M.
Terrones
,
W.
Windl
, and
J. E.
Goldberger
,
ACS Nano
7
,
2898
(
2013
).
4.
G. R.
Bhimanapati
,
Z.
Lin
,
V.
Meunier
,
Y.
Jung
,
J.
Cha
,
S.
Das
,
D.
Xiao
,
Y.
Son
,
M. S.
Strano
,
V. R.
Cooper
,
L.
Liang
,
S. G.
Louie
,
E.
Ringe
,
W.
Zhou
,
S. S.
Kim
,
R. R.
Naik
,
B. G.
Sumpter
,
H.
Terrones
,
F.
Xia
,
Y.
Wang
,
J.
Zhu
,
D.
Akinwande
,
N.
Alem
,
J. A.
Schuller
,
R. E.
Schaak
,
M.
Terrones
, and
J. A.
Robinson
,
ACS Nano
9
,
11509
(
2015
).
5.
K.
Khan
,
A.
Khan Tareen
,
M.
Aslam
,
R.
Wang
,
Y.
Zhang
,
A.
Mahmood
,
Z.
Ouyang
,
H.
Zhang
, and
Z.
Guo
,
J. Mater. Chem. C
8
,
387
(
2020
).
6.
A.
Avsar
,
H.
Ochoa
,
F.
Guinea
,
B.
Ozyilmaz
,
B. J.
van Wees
, and
I. J.
Vera-Marun
,
Rev. Mod. Phys.
92
,
021003
(
2020
).
7.
P.
Miró
,
M.
Audiffred
, and
T.
Heine
,
Chem. Soc. Rev.
43
,
6537
(
2014
).
8.
M.
Xu
,
T.
Liang
,
M.
Shi
, and
H.
Chen
,
Chem. Rev.
113
,
3766
(
2013
).
9.
T.
Vincent
,
J.
Liang
,
S.
Singh
,
E. G.
Castanon
,
X.
Zhang
,
A.
McCreary
,
D.
Jariwala
,
O.
Kazakova
, and
Z. Y.
Al Balushi
,
Appl. Phys. Rev.
8
,
041320
(
2021
).
10.
A. K.
Geim
and
I. V.
Grigorieva
,
Nature
499
,
419
(
2013
).
11.
C.
Liu
,
H.
Chen
,
S.
Wang
,
Q.
Liu
,
Y.-G.
Jiang
,
D. W.
Zhang
,
M.
Liu
, and
P.
Zhou
,
Nat. Nanotechnol.
15
,
545–557
(
2020
).
12.
M.
Zeng
,
Y.
Xiao
,
J.
Liu
,
K.
Yang
, and
L.
Fu
,
Chem. Rev.
118
,
6236
(
2018
).
13.
Z.
Zhu
and
D.
Tománek
,
Phys. Rev. Lett.
112
,
176802
(
2014
).
14.
C.
Kamal
and
M.
Ezawa
,
Phys. Rev. B
91
,
085423
(
2015
).
15.
O.Ü.
Aktürk
,
V. O.
Özçelik
, and
S.
Ciraci
,
Phys. Rev. B
91
,
235446
(
2015
).
16.
V. O.
Özcelik
,
O.Ü.
Aktürk
,
E.
Durgun
, and
S.
Ciraci
,
Phys. Rev. B
92
,
125420
(
2015
).
17.
S.
Zhang
,
Z.
Yan
,
Y.
Li
,
Z.
Chen
, and
H.
Zeng
,
Angew. Chem.
127
,
3155
(
2015
).
18.
J.
Ji
,
X.
Song
,
J.
Liu
,
Z.
Yan
,
C.
Huo
,
S.
Zhang
,
M.
Su
,
L.
Liao
,
W.
Wang
,
Z.
Ni
, et al.,
Nat. Commun.
7
,
13352
(
2016
).
19.
S.
Zhang
,
M.
Xie
,
F.
Li
,
Z.
Yan
,
Y.
Li
,
E.
Kan
,
W.
Liu
,
Z.
Chen
, and
H.
Zeng
,
Angew. Chem.
128
,
1698
(
2016
).
20.
E.
Aktürk
,
O.Ü.
Aktürk
, and
S.
Ciraci
,
Phys. Rev. B
94
,
014115
(
2016
).
22.
S.
Zhang
,
S.
Guo
,
Z.
Chen
,
Y.
Wang
,
H.
Gao
,
J.
Gómez-Herrero
,
P.
Ares
,
F.
Zamora
,
Z.
Zhu
, and
H.
Zeng
,
Chem. Soc. Rev.
47
,
982
(
2018
).
23.
R.
Bhuvaneswari
,
V.
Nagarajan
, and
R.
Chandiramouli
,
Surf. Interfaces
28
,
101610
(
2022
).
24.
Y.
Nie
,
M.
Rahman
,
P.
Liu
,
A.
Sidike
,
Q.
Xia
, and
G.-h.
Guo
,
Phys. Rev. B
96
,
075401
(
2017
).
25.
J.
Zhao
,
Z.-H.
Qi
,
Y.
Xu
,
J.
Dai
,
X. C.
Zeng
,
W.
Guo
, and
J.
Ma
,
Wiley Interdiscipl. Rev.
9
,
e1387
(
2019
).
26.
S.
Guo
,
W.
Zhou
,
B.
Cai
,
K.
Zhang
,
S.
Zhang
, and
H.
Zeng
,
Nanoscale Horiz.
4
,
1145
(
2019
).
27.
A.
Majumdar
,
X.
Yang
,
W.
Luo
,
S.
Chowdhury
,
S.
Chakraborty
, and
R.
Ahuja
,
Appl. Surf. Sci.
529
,
146552
(
2020
).
28.
S.
Zhang
,
H.
Qu
,
J.
Cao
,
Y.
Wang
,
S. A.
Yang
,
W.
Zhou
, and
H.
Zeng
,
IEEE Trans. Electron Devices
69
,
857
(
2022
).
29.
H. L.
Zhuang
and
R. G.
Hennig
,
Chem. Mater.
25
,
3232
(
2013
).
30.
X.
Li
,
M.-W.
Lin
,
A. A.
Puretzky
,
J. C.
Idrobo
,
C.
Ma
,
M.
Chi
,
M.
Yoon
,
C. M.
Rouleau
,
I. I.
Kravchenko
,
D. B.
Geohegan
et al.,
Sci. Rep.
4
,
5497
(
2014
).
31.
M.-W.
Chen
,
H.
Kim
,
D.
Ovchinnikov
,
A.
Kuc
,
T.
Heine
,
O.
Renault
, and
A.
Kis
,
npj 2D Mater. Appl.
2
,
1
(
2018
).
32.
H.
Wang
,
G.
Qin
,
J.
Yang
,
Z.
Qin
,
Y.
Yao
,
Q.
Wang
, and
M.
Hu
,
J. Appl. Phys.
125
,
245104
(
2019
).
33.
K.
Xu
,
L.
Yin
,
Y.
Huang
,
T. A.
Shifa
,
J.
Chu
,
F.
Wang
,
R.
Cheng
,
Z.
Wang
, and
J.
He
,
Nanoscale
8
,
16802
(
2016
).
34.
P.
Das
,
D.
Wickramaratne
,
B.
Debnath
,
G.
Yin
, and
R. K.
Lake
,
Phys. Rev. B
99
,
085409
(
2019
).
35.
M. N.
Çınar
,
G.Ö
Sargın
,
K.
Sevim
,
B.
Özdamar
,
G.
Kurt
, and
H.
Sevinçli
,
Phys. Rev. B
103
,
165422
(
2021
).
36.
E. A.
Stepanov
,
V.
Harkov
,
M.
Rösner
,
A. I.
Lichtenstein
,
M. I.
Katsnelson
, and
A. N.
Rudenko
,
npj Comput. Mater.
8
,
1
(
2022
).
37.
M.
Ashton
,
S. B.
Sinnott
, and
R. G.
Hennig
,
Appl. Phys. Lett.
109
,
192103
(
2016
).
38.
C.
Barreteau
,
B.
Michon
,
C.
Besnard
, and
E.
Giannini
,
J. Cryst. Growth.
443
,
75
(
2016
).
39.
S.
Zhang
,
S.
Guo
,
Y.
Huang
,
Z.
Zhu
,
B.
Cai
,
M.
Xie
,
W.
Zhou
, and
H.
Zeng
,
2D Mater.
4
,
015030
(
2016
).
40.
P.
Wu
and
M.
Huang
,
Phys. Status Solidi B
253
,
862
(
2016
).
41.
F.
Shojaei
and
H. S.
Kang
,
J. Phys. Chem. C
120
,
23842
(
2016
).
42.
A.-Q.
Cheng
,
Z.
He
,
J.
Zhao
,
H.
Zeng
, and
R.-S.
Chen
,
ACS Appl. Mater. Interfaces
10
,
5133
(
2018
).
43.
L.
Zhou
,
Y.
Guo
, and
J.
Zhao
,
Phys. E
95
,
149
(
2018
).
44.
D.
Singh
,
S.
Kansara
,
S. K.
Gupta
, and
Y.
Sonvane
,
J. Mater. Sci.
53
,
8314
(
2018
).
45.
B.
Özdamar
,
G.
Özbal
,
M. N.
Çınar
,
K.
Sevim
,
G.
Kurt
,
B.
Kaya
, and
H.
Sevinçli
,
Phys. Rev. B
98
,
045431
(
2018
).
46.
S.
Demirci
,
N.
Avazlı
,
E.
Durgun
, and
S.
Cahangirov
,
Phys. Rev. B
95
,
115409
(
2017
).
47.
D.
Wickramaratne
,
F.
Zahid
, and
R. K.
Lake
,
J. Appl. Phys.
118
,
075101
(
2015
).
48.
A.
Kuc
,
E.
Cusati
,
T.a.
Dib
,
A. F.
Oliveira
,
A.
Fortunelli
,
G.
Iannaccone
,
T.
Heine
, and
G.
Fiori
,
Adv. Electron. Mater.
3
,
1600399
(
2017
).
49.
T.
Cao
,
Z.
Li
, and
S. G.
Louie
,
Phys. Rev. Lett.
114
,
236602
(
2015
).
50.
Y.
Ma
,
Y.
Dai
,
M.
Guo
,
L.
Yu
, and
B.
Huang
,
Phys. Chem. Chem. Phys.
15
,
7098
(
2013
).
51.
D. V.
Rybkovskiy
,
A. V.
Osadchy
, and
E. D.
Obraztsova
,
Phys. Rev. B
90
,
235302
(
2014
).
52.
X.
Wang
,
Y.
Sheng
,
R.-J.
Chang
,
J. K.
Lee
,
Y.
Zhou
,
S.
Li
,
T.
Chen
,
H.
Huang
,
B. F.
Porter
,
H.
Bhaskaran
, and
J. H.
Warner
,
ACS Omega
3
,
7897
(
2018
).
53.
S.
Lei
,
L.
Ge
,
Z.
Liu
,
S.
Najmaei
,
G.
Shi
,
G.
You
,
J.
Lou
,
R.
Vajtai
, and
P. M.
Ajayan
,
Nano Lett.
13
,
2777
(
2013
).
54.
C. S.
Jung
,
F.
Shojaei
,
K.
Park
,
J. Y.
Oh
,
H. S.
Im
,
D. M.
Jang
,
J.
Park
, and
H. S.
Kang
,
ACS Nano
9
,
9585
(
2015
).
55.
F.
Liu
,
H.
Shimotani
,
H.
Shang
,
T.
Kanagasekaran
,
V.
Zólyomi
,
N.
Drummond
,
V. I.
Fal’ko
, and
K.
Tanigaki
,
ACS Nano
8
,
752
(
2014
).
56.
P.
Hu
,
L.
Wang
,
M.
Yoon
,
J.
Zhang
,
W.
Feng
,
X.
Wang
,
Z.
Wen
,
J. C.
Idrobo
,
Y.
Miyamoto
,
D. B.
Geohegan
, et al.,
Nano Lett.
13
,
1649
(
2013
).
57.
S.
Lei
,
F.
Wen
,
L.
Ge
,
S.
Najmaei
,
A.
George
,
Y.
Gong
,
W.
Gao
,
Z.
Jin
,
B.
Li
,
J.
Lou
,
J.
Kono
,
R.
Vajtai
,
P.
Ajayan
, and
N. J.
Halas
,
Nano Lett.
15
,
3048
(
2015
).
58.
D. A.
Bandurin
,
A. V.
Tyurnina
,
G. L.
Yu
,
A.
Mishchenko
,
V.
Zólyomi
,
S. V.
Morozov
,
R. K.
Kumar
,
R. V.
Gorbachev
,
Z. R.
Kudrynskyi
,
S.
Pezzini
,
Z. D.
Kovalyuk
,
U.
Zeitler
,
K. S.
Novoselov
,
A.
Patanè
,
L.
Eaves
,
I. V.
Grigorieva
,
V. I.
Fal’ko
,
A. K.
Geim
, and
Y.
Cao
,
Nat. Nanotechnol.
12
,
223
(
2017
).
59.
J.
Zhou
,
J.
Shi
,
Q.
Zeng
,
Y.
Chen
,
L.
Niu
,
F.
Liu
,
T.
Yu
,
K.
Suenaga
,
X.
Liu
,
J.
Lin
, and
Z.
Liu
,
2D Mater.
5
,
025019
(
2018
).
60.
W.
Zhang
,
H.
Enriquez
,
Y.
Tong
,
A.
Bendounan
,
A.
Kara
,
A. P.
Seitsonen
,
A. J.
Mayne
,
G.
Dujardin
, and
H.
Oughaddou
,
Small
14
,
1804066
(
2018
).
61.
H.
Li
,
D.
Zhou
,
Q.
He
,
N.
Si
,
B.
Xin
,
S.
Bu
,
Q.
Ji
,
H.
Li
,
H.
Fuchs
, and
T.
Niu
,
ACS Nano
15
,
16335
(
2021
).
62.
T. A.
Sedrakyan
,
L. I.
Glazman
, and
A.
Kamenev
,
Phys. Rev. Lett.
114
,
037203
(
2015
).
63.
T. A.
Sedrakyan
,
V. M.
Galitski
, and
A.
Kamenev
,
Phys. Rev. Lett.
115
,
195301
(
2015
).
64.
T. A.
Sedrakyan
,
L. I.
Glazman
, and
A.
Kamenev
,
Phys. Rev. B
89
,
201112
(
2014
).
66.
M. L.
Sancho
,
J. L.
Sancho
,
J. L.
Sancho
, and
J.
Rubio
,
J. Phys. F: Met. Phys.
15
,
851
(
1985
).
67.
S.
Datta
,
Electronic Transport in Mesoscopic Systems
(
Cambridge University Press
,
1997
).
68.
D.
Ryndyk
,
Theory of Quantum Transport at Nanoscale: An Introduction
(
Springer
,
2015
).
69.
N.
Nemec
and
G.
Cuniberti
,
Phys. Rev. B
74
,
165411
(
2006
).
70.
P. W.
Anderson
,
D.
Thouless
,
E.
Abrahams
, and
D.
Fisher
,
Phys. Rev. B
22
,
3519
(
1980
).
71.
J.
Sak
and
B.
Kramer
,
Phys. Rev. B
24
,
1761
(
1981
).
72.
K.
Chase
and
A.
MacKinnon
,
J. Phys. C: Solid State. Phys.
20
,
6189
(
1987
).
73.
C. W.
Beenakker
,
Rev. Mod. Phys.
69
,
731
(
1997
).
74.
75.
N.
Nemec
,
K.
Richter
, and
G.
Cuniberti
,
New J. Phys.
10
,
065014
(
2008
).
76.
D. A.
Areshkin
,
D.
Gunlycke
, and
C. T.
White
,
Nano Lett.
7
,
204
(
2007
).
77.
A.
Lherbier
,
B.
Biel
,
Y.-M.
Niquet
, and
S.
Roche
,
Phys. Rev. Lett.
100
,
036803
(
2008
).
78.
A.
Cresti
,
A.
Lopez-Bezanilla
,
P.
Ordejón
, and
S.
Roche
,
ACS Nano
5
,
9271
(
2011
).
79.
A. U.
Canbolat
,
H.
Sevincli
, and
O.
Cakir
,
Phys. Rev. B
106
,
104409
(
2022
).
80.
N.
Leconte
,
A.
Lherbier
,
F.
Varchon
,
P.
Ordejon
,
S.
Roche
, and
J.-C.
Charlier
,
Phys. Rev. B
84
,
235420
(
2011
).
81.
Z.
Fan
,
J. H.
Garcia
,
A. W.
Cummings
,
J. E.
Barrios-Vargas
,
M.
Panhans
,
A.
Harju
,
F.
Ortmann
, and
S.
Roche
,
Phys. Rep.
903
,
1
(
2021
).
82.
E.
Abrahams
,
P.
Anderson
,
D.
Licciardello
, and
T.
Ramakrishnan
,
Phys. Rev. Lett.
42
,
673
(
1979
).
83.
P. A.
Lee
and
T.
Ramakrishnan
,
Rev. Mod. Phys.
57
,
287
(
1985
).
84.
Y.
Imry
,
Introduction to Mesoscopic Physics
(
Oxford University Press
,
2002
).
85.
M. N.
Çınar
,
A.
Antidormi
,
V.-H.
Nguyen
,
A.
Kovtun
,
S.
Lara-Avila
,
A.
Liscio
,
J.-C.
Charlier
,
S.
Roche
, and
H.
Sevinçli
,
Nano Lett.
22
,
2202
(
2022
).