Brought to you by:
Letter

Magnetothermoelectric transport properties in graphene superlattices with one-dimensional periodic potentials

and

Published 20 January 2015 Copyright © EPLA, 2015
, , Citation R. Ma and L. Sheng 2015 EPL 109 17004 DOI 10.1209/0295-5075/109/17004

0295-5075/109/1/17004

Abstract

We numerically study the thermoelectric transport in graphene superlattice in the presence of a strong magnetic field and disorder, with a one-dimensional periodic potential barrier oriented along the armchair direction of graphene. The thermoelectric coefficients exhibit novel properties as functions of the superlattice period length. When the period length is taken to be the smallest, we find that the thermoelectric transport exhibits properties similar to those in the absence of the superlattice potential. The thermoelectric conductivities display different asymptotic behaviors, depending on the ratio between the temperature and the width of the disorder-broadened Landau levels (LLs). In the high-temperature regime, the transverse thermoelectric conductivity $\alpha_{xy}$ saturates to a universal value $2.27 k_B e/h$ at the center of each LL, and displays a linear temperature dependence at low temperatures. We attribute this unique behavior to the coexistence of particle and hole LLs. Both the Nernst signal and the thermopower show a large peak around the central LL. However, with increasing the superlattice period length, it is found that the thermoelectric transport properties are consistent with the behavior of a band insulator. $\alpha_{xy}$ displays a pronounced valley at low temperatures. This behavior can be understood as due to the split of the valley degeneracy in the central LL. The obtained results demonstrate the sensitivity of the thermoelectric conductivity to the superlattice period length.

Export citation and abstract BibTeX RIS

Introduction

Recently, the thermoelectric transport properties of graphene systems have attracted much experimental [14] and theoretical [512] attention. The thermopower (the longitudinal thermoelectric response) and the Nernst signal (the transverse response) of graphene in the presence of a strong magnetic field are found experimentally to be large, reaching the order of the quantum limit $k_B/e$ , where kB and e are the Boltzmann constant and the electron charge, respectively [13]. Theoretical calculations from the tight-binding model for graphene [11,12] are in agreement with the experimental observation. This makes the graphene-related materials more promising candidates for future thermoelectric applications.

More recently, much attention has been turned toward the transport properties of graphene superlattices with periodic potential [1321]. In a graphene-based superlattice, a one-dimensional periodic potential may result in a strong anisotropy for the group velocities of the low-energy charge carriers, which are reduced to zero in one direction but are unchanged in another [16]. Interestingly, Brey et al. [17] have shown that such behavior of the anisotropy is a precursor to the formation of further Dirac points in the electronic band structures, and new zero-energy states are controlled by the potential amplitude and the superlattice period. Furthermore, the extra Dirac points and their associated zero-energy modes [17] drastically affect the transport properties [1720] of the system and also the Landau level (LLs) sequence [21], and hence the plateaus in the quantum Hall conductivity when a magnetic field is applied. Owing to these interesting electronic behavior, graphene superlattices are expected to exhibit novel thermoelectric transport properties. However, theoretical studies of the thermoelectric transport properties of graphene superlattices have not been carried out so far. Such theoretical studies are highly desirable as they can provide theoretical understanding and guidance to the experimental research of the thermoelectric transport in such systems.

In this paper, we perform a numerical study of the thermoelectric transport properties in graphene systems in the presence of a one-dimensional superlattice potential. The effects of disorder and thermal activation on the broadening of the LLs and the corresponding thermoelectric transport properties are investigated. When the superlattice period length is the smallest, the thermoelectric coefficients exhibit unique behaviors at the central LL due to the coexistence of particle and hole LLs. Both the longitudinal and transverse thermoelectric conductivities are universal functions of the temperature and the disorder-induced LL width, and display different asymptotic behaviors in different temperature regimes. The Nernst signal displays a peak with a height of the order of $k_B/e$ , and the thermopower changes sign at the central LL. With the increase of the superlattice period length, quite different behaviors near the central LL are observed. Around the Dirac point, the transverse thermoelectric conductivity exhibits a pronounced valley at low temperatures. This is attributed to the splitting of the central LL near zero energy, and hence the graphene superlattices behave as conventional insulators. We further examine the validity of the semiclassical Mott relation in the graphene superlattice systems.

This paper is organized as follows. In the next section, we introduce the model Hamiltonian. In the third section, numerical results based on exact diagonalization and thermoelectric transport calculations are presented for graphene superlattice systems. The final section contains a summary.

Model and methods

We assume that a graphene sheet has Ly zigzag chains in total, with Lx atomic sites on each chain [22]. We consider a one-dimensional periodic Kronig-Penney type of electrostatic potential with alternating potential barriers and wells along the armchair direction of the graphene sheet, as shown in fig. 1, which can be described by

Equation (1)

where $L=2n\sqrt{3}a_0$ is the period of the potential. In the presence of an applied magnetic field perpendicular to the plane of graphene, the lattice model in real space can be written in the tight-binding form

Equation (2)

Here, $c_{i}^{\dagger}\ (c_{i})$ is the fermion creation (annihilation) operator at site i, t is the hopping integral between the nearest-neighbor (next-nearest-neighbor) sites i and j, and Vi is the superlattice potential given by eq. (1). The spin degrees of freedom have been omitted due to degeneracy. Under the applied magnetic field B, the vector potential is introduced into eq. (2) via an additional phase factor aij, which is determined by the magnetic flux per hexagon $\varphi =\sum_{{\small \unicode{x2B21}}}a_{ij}=\frac{2\pi}{M}$ . M is an integer proportional to the strength of the applied magnetic field B and the lattice constant is taken to be unity. We model charged impurities in substrate, randomly located in a plane at a distance d, either above or below the graphene sheet with a long-range Coulomb scattering potential [2326]. For charged impurities, $w_i=-\frac{Ze^2}{\epsilon}\sum_{\alpha}1/\sqrt{(\textbf{r}_i- \textbf{R}_{\alpha})^2+d^2}$ , where Ze is the charge carried by an impurity, epsilon is the effective background lattice dielectric constant, and $\textbf{r}_i$ and $\textbf{R}_{\alpha}$ are the planar positions of site i and impurity α, respectively. All the properties of the substrate can be absorbed into a dimensionless parameter $r_s= Ze^2/(\epsilon\hbar v_F)$ , where vF is the Fermi velocity of the electrons. $\hbar v_F=\frac{3}{2}ta$ , where a is lattice constant [24]. For simplicity, in the following calculation, we fix the distance $d=1\ \text{nm}$ and impurity density as 1% of the total sites, and tune rs to control the impurity scattering strength. The characteristic features of the calculated transport coefficients are insensitive to the details of the impurity scattering and choices of the parameters.

Fig. 1:

Fig. 1: (Color online) Schematic diagram of a Kronig-Penney superlattice potential applied to graphene along the armchair direction, where Nac indicates the number of dimer lines, and L is the period of the potential.

Standard image

In the linear response regime, the charge current in response to an electric field or a temperature gradient can be written as $\textbf{J} = {\hat\sigma}\textbf{E} + {\hat\alpha} (-\nabla T)$ , where ${\hat\sigma}$ and ${\hat\alpha}$ are the electrical and thermoelectric conductivity tensors, respectively. The Hall conductivity $\sigma_{xy}$ can be calculated by Kubo formula and the longitudinal conductivity $\sigma_{xx}$ can be obtained based on the calculation of the Thouless number (see appendix) [27]. We exactly diagonalize the model Hamiltonian in the presence of disorder [28], and obtain the transport coefficients by using the energy spectra and wave functions. In practice, we can first calculate the electrical conductivity $\sigma_{ji}(E_F)$ at zero temperature, and then use the relation [29]

Equation (3)

to obtain the electrical and thermoelectric conductivity at finite temperatures. Here, $f(x)=1/[e^{(x-E_F)/k_B T}+1]$ is the Fermi distribution function. At low temperatures, the second equation can be approximated as

Equation (4)

which is the semiclassical Mott relation [29,30]. The validity of this relation will be examined for the present graphene superlattice system. The thermopower and Nernst signal can be calculated subsequently from [31,32]

Equation (5)

Thermoelectric transport of graphene superlattices

We first show the Hall conductivity $\sigma_{xy}$ as functions of electron Fermi energy EF at zero temperature in fig. 2. For comparison, we also show the results of the graphene system in the absence of the superlattice potential. As shown in fig. 2(a), when no superlattice is present, the Hall conductivity exhibits a sequence of plateaus at $\sigma _{xy}=\nu e^2/h$ , where $\nu =kg_{s}$ with k a nonzero integer, and $g_{s}=4$ due to the fourfold degeneracy (double-spin and double-valley) [3336]. The transition from the $\nu =-2$ plateau to $\nu=2$ plateau is continuous without a $\nu=0$ plateau appearing in between, so that a step of height $4e^2/h$ occurs at the neutrality point. Moreover, the central LL shows a peak around the Dirac point, and the width of the peak increases with the increase of the disorder strength. However, when a superlattice potential is applied to graphene sheet, the Hall conductivity exhibits different quantization rules depending on the superlattice period length L. In fig. 2(b), we first show the result for $L=2\sqrt{3}a_0\ (n=1)$ . As we can see, there is a splitting of the LL around $E_F=\pm 0.45t$ , yielding a new plateau with ${\sigma_{xy}}=\pm 4,\pm 8$ , in addition to the original integer plateaus. When the superlattice period length increases to $L=24\sqrt{3}a_0\ (n=12)$ , there is a splitting of the central $(n = 0)$ LL, yielding a tiny new plateau with ${\sigma_{xy}}=0$ , as shown in fig. 2(c). Due to this large difference in the Hall conductivity and the LLs in the two cases, the thermoelectric transport properties of graphene superlattices are expected to exhibit interesting characteristics.

Fig. 2:

Fig. 2: (Color online) Calculated Hall conductivity $\sigma_{xy}$ in units of $e^2/h$ as a function of the Fermi energy at zero temperature for (a) graphene, and (b), (c) graphene superlattice. The periods of the superlattice potential in (b) and (c) are taken to be $L=2\sqrt{3}a_0\ (n=1)$ , and $L=24\sqrt{3}a_0\ (n=12)$ , respectively, and the superlattice potential strength is set to be $V_0=0.05t$ . The other parameters are taken to be $N=96\times48$ , magnetic flux $\varphi=2\pi/48$ , and disorder strength $r_s=0.3$ .

Standard image

We now study the thermoelectric conductivities at finite temperatures. In fig. 3 and fig. 4, we find that the transverse thermoelectric conductivity $\alpha_{xy}$ shows different behavior depending on the superlattice period length. In fig. 3(a) and (b), we first plot the calculated $\alpha_{xy}$ for $L=2\sqrt{3}a_0\ (n=1)$ . As shown in fig. 3(a) and (b), $\alpha_{xy}$ displays a series of peaks, while the longitudinal thermoelectric conductivity ${\alpha_{xx}}$ oscillates and changes sign at the center of each LL. At low temperatures, the peak of $\alpha_{xy}$ at the central LL is higher and narrower than others, which indicates that the impurity scattering has less effect on the central LL. These results are qualitatively similar to those found in monolayer graphene [11], but some obvious differences exist. More oscillations are observed in higher LLs than in the graphene case, consistently with the further lifting of the LL degeneracy in graphene superlattices. ${\alpha_{xx}}$ oscillates and changes sign around the center of each split LL. At low temperatures, $\alpha_{xy}$ splits around $E_F=\pm 0.45t$ , which can be understood as due to the presence of $\nu=\pm 4$ Hall plateau by lifting the subband degeneracy. In fig. 3(c), we find that $\alpha_{xy}$ shows different behavior depending on the relative strength of temperature kBT and the width of the central LL WL (WL is determined by the full width at the half-maximum of the $\sigma_{xx}$ peak). When $k_{B}T \ll W_L$ and $E_F \ll W_L$ , $\alpha_{xy}$ shows linear temperature dependence, indicating that there is a small energy range where extended states dominate, and the transport falls into the semiclassical Drude-Zener regime. When EF is shifted away from the Dirac point, the low-energy electron excitation is gapped due to Anderson localization. When kBT becomes comparable to or greater than WL, $\alpha_{xy}$ for all LLs saturates to a constant value $2.77 k_B e/h$ . This matches exactly the universal value $(\ln 2) k_B e/h$ predicted for the conventional integer quantum Hall effect (IQHE) systems in the case where thermal activation dominates [29,30], with an additional degeneracy factor 4. The saturated value of $\alpha_{xy}$ is in accordance with the fourfold degeneracy at zero energy.

Fig. 3:

Fig. 3: (Color online) Thermoelectric conductivities at finite temperatures of graphene superlattices with the superlattice period length $L=2\sqrt{3}a_0\ (n=1)$ . (a), (b): $\alpha_{xy}(E_F,T)$ and $\alpha_{xx}(E_F,T)$ as functions of the Fermi energy EF at different temperature T. Panel (c) shows the temperature dependence of $\alpha_{xy}(E_F,T)$ for the graphene superlattices. (d) Comparison of the results from numerical calculations and from the generalized Mott relation at two characteristic temperatures, $k_{B}T/W_L=0.1$ and $k_BT/W_L=2.0$ . Here, the width of the central LL is set to be $W_L/t=0.02$ . The other parameters are chosen to be the same as in fig. 2.

Standard image
Fig. 4:

Fig. 4: (Color online) Thermoelectric conductivities at finite temperatures of graphene superlattices with the superlattice period length $L=24\sqrt{3}a_0\ (n=12)$ . (a), (b): $\alpha_{xy}(E_F,T)$ and $\alpha_{xx}(E_F,T)$ for $W_L/t=0.018$ as functions of Fermi energy at different temperatures. The other parameters are chosen to be the same as in fig. 2.

Standard image

To examine the validity of the semiclassical Mott relation, we compare the above results with those calculated from eq. (4), as shown in fig. 3(d). The Mott relation is a low-temperature approximation and predicts that the thermoelectric conductivities have linear temperature dependence. This is in agreement with our low-temperature results, which proves that the semiclassical Mott relation is asymptotically valid in the Landau-quantized systems, as suggested in ref. [29].

When the superlattice period length increases to $L=24 \sqrt{3} a_0\ (n=12)$ , the calculated $\alpha_{xy}$ and $\alpha_{xx}$ at finite temperatures are shown in fig. 4. As seen from fig. 4(a), $\alpha_{xy}$ displays a pronounced valley at low temperatures, in striking contrast to the $L=2 \sqrt{3}a_0\ (n=1)$ case with a peak at $E_F=0$ . This behavior can be understood as due to the split of the valley degeneracy in the central LL. $\alpha_{xx}$ oscillates and changes sign around the center of each split LL. In fig. 4(c), we also compare the above results with those calculated from the semiclassical Mott relation using eq. (4). The Mott relation is found to remain valid at low temperatures.

We further calculate the thermopower Sxx and the Nernst signal Sxy using eq. (5), which can be directly determined in experiments by measuring the responsive electric fields. In fig. 5(a), (b), we first show the calculated Sxx and Sxy for $L=2 \sqrt{3} a_0\ (n=1)$ . As we can see, $S_{xy}\ (S_{xx})$ has a peak at the central LL (the other LLs) and changes sign near the other LLs (the central LL). This oscillatory feature is qualitatively similar to those found in monolayer graphene case [11], which has been observed experimentally [3]. At zero energy, both $\rho_{xy}$ and $\alpha_{xx}$ vanish, leading to a vanishing Sxx. Around zero energy, because $\rho_{xx}\alpha_{xx}$ and $\rho_{xy}\alpha_{xy}$ have opposite signs, depending on their relative magnitudes, Sxx could either increases or decreases when EF is increased passing the Dirac point. In the graphene superlattice case, we find that Sxx is always dominated by $\rho_{xy} \alpha_{xy}$ . Consequently, Sxx increases to a positive value as EF passes zero. This is quite different from the graphene case. At low temperatures, the peak value of Sxx near zero energy is $\pm 1.14\,k_B/e\ (\pm 98.23\ \mu \text{V/K})$ at $k_BT=0.5W_L$ . With the increase of temperature, the peak height increases to $\pm 3.97\,k_B/e\ (\pm 342.09\ \mu \text{V/K})$ at $k_BT=1.5W_L$ . On the other hand, Sxy has a strong peak structure around zero energy, which is dominated by $\rho_{xx}\alpha_{xy}$ . The peak height is $8.22\,k_B/e\ (708.32\,\mu V/K)$ at $k_BT=1.5W_L$ .

Fig. 5:

Fig. 5: (Color online) The thermopower Sxx and the Nernst signal Sxy as functions of the Fermi energy for the superlattice period lengths (a), (b) $L=2\sqrt{3}a_0\ (n=1)$ , and (c), (d) $L=24\sqrt{3}a_0\ (n=12)$ at different temperatures. The other parameters are chosen to be the same as in fig. 2, respectively.

Standard image

In fig. 5(c), (d), we show the calculated Sxx and Sxy for $L=24 \sqrt{3} a_0\ (n=12)$ . As we can see, $S_{xy}\ (S_{xx})$ has a peak around zero energy (the other LLs), and changes sign near the other LLs (zero energy). In our calculation, Sxx is dominated by $\rho_{xy}\alpha_{xy}$ . The peak value of Sxx near zero energy is $\pm 1.59\,k_B/e\ (\pm 137.01\,\mu V/K)$ at $k_BT=1.5W_L$ . On the other hand, Sxy has a valley structure around zero energy at low temperatures, which is dominated by $\rho_{xx}\alpha_{xy}$ . With the increase of temperature, we find Sxy exhibits a peak. The peak height is $3.68\,k_B/e\ (317.11\,\mu V/K)$ at $k_BT=1.5W_L$ .

Summary

In summary, we have numerically investigated the thermoelectric transport properties of graphene superlattice systems in the presence of both disorder and a strong magnetic field. We find that the thermoelectric coefficients strongly depend on the superlattice period length. When the superlattice period length is taken to be the smallest, the thermoelectric conductivities display different asymptotic behavior depending on the ratio between the temperature and the width of the disorder-broadened LLs. At low temperatures, $\alpha_{xy}$ splits around $E_F=\pm 0.45t$ , which can be understood as due to the presence of $\nu=\pm 4$ Hall plateau by lifting the subband degeneracy. In the high-temperature regime, the transverse thermoelectric conductivity $\alpha_{xy}$ saturates to a universal value $2.27\,k_B e/h$ at the center of each LL, and displays a linear temperature dependence at low temperatures. Both thermopower and Nernst signal display large peaks at high temperatures around the central LL. The validity of the semiclassical Mott relation between the thermoelectric and electrical transport coefficients is verified to be satisfied only at very low temperatures. However, when the superlattice period length is increased to $L=24\sqrt{3}a_0\ (n=12)$ , the thermoelectric coefficients display quite distinct behaviors. Around the Dirac point, the transverse thermoelectric conductivity $\alpha_{xy}$ and the Nernst signal Sxy exhibit a pronounced valley at low temperatures, in striking contrast to the graphene case, where $\alpha_{xy}$ and Sxy display a peak. These are consistent with the splitting of the central LL in the graphene superlattice case.

Acknowledgments

This work was supported by the General Financial Grant from the China Postdoctoral Science Foundation (Grant No. 2014M551546), and the NSFC Grant Nos. 11104146 (RM) and 11225420 (LS). This work was also supported by the State Key Program for Basic Researches of China under Grant Nos. 2015CB921202 and 2014CB921103 (LS).

Appendix:

The Hall conductivity $\sigma_{xy}$ at zero temperature can be calculated by using the Kubo formula [37]

Equation (A.1)

Here, $\epsilon_\alpha$ , $\epsilon_\beta$ are the eigenenergies corresponding to the eigenstates ∣α〉, ∣β〉 of the system, which can be obtained through exact diagonalization of the Haldane model Hamiltonian. S is the area of the sample, Vx and Vy are the velocity operators. With tuning chemical potential EF, a series of integer-quantized plateaus of $\sigma_{xy}$ appear, each one corresponding to EF moving in the gaps between two neighboring Landau Levels (LLs).

The longitudinal conductivity $\sigma_{xx}$ at zero temperature can be obtained based on the calculation of the Thouless number. The Thouless number g is calculated by using the following formula [38]:

Equation (A.2)

Here, $\Delta E$ is the geometric mean of the shift in the energy levels of the system caused by replacing periodic by antiperiodic boundary conditions, and $\text{d}E/\text{d}N$ is the mean spacing of the energy levels. The Thouless number g is proportional to the longitudinal conductivity $\sigma_{xx}$ .

Please wait… references are loading.
10.1209/0295-5075/109/17004