This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Transport properties of multilayer graphene

Glenn Wagner Rudolf Peierls Centre for Theoretical Physics, Parks Road, Oxford, OX1 3PU, UK    Dung X. Nguyen Brown Theoretical Physics Center and Department of Physics, Brown University, 182 Hope Street, Providence, RI 02912, USA    Steven H. Simon Rudolf Peierls Centre for Theoretical Physics, Parks Road, Oxford, OX1 3PU, UK
Abstract

We apply the semi-classical quantum Boltzmann formalism for the computation of transport properties to multilayer graphene. We compute the electrical conductivity as well as the thermal conductivity and thermopower for Bernal-stacked multilayers with an even number of layers. We show that the window for hydrodynamic transport in multilayer graphene is similar to the case of bilayer graphene. We introduce a simple hydrodynamic model which we dub the multi-fluid model and which can be used to reproduce the results for the electrical conductivity and thermopower from the quantum Boltzmann equation.

I Introduction

Ultra-clean materials such as graphene offer a new perspective on electronic transport. At low enough temperatures, momentum-relaxing scattering of the electrons such as the scattering off phonons or impurities is sub-dominant and the dominant source of collisions are the collisions of the electrons with themselves. This is the realm of electron hydrodynamics Lucas and Fong (2018). In the hydrodynamic regime the electron-electron scattering rate τee1\tau_{ee}^{-1} is larger than the electron-phonon scattering rate τep1\tau_{ep}^{-1}. Both monolayer and bilayer graphene have garnered much attention in recent years for their supposed hydrodynamic transport Bandurin et al. (2016); Levitov and Falkovich (2016); Sulpizio et al. (2019); Svintsov et al. (2012); Pellegrino et al. (2017); Bistritzer and MacDonald (2009); Müller et al. (2009); Andreev et al. (2011); Narozhny et al. (2015). In the present paper, we focus on a related material: multilayer graphene.

The Boltzmann equation is an equation of motion for the distribution function of particles and has traditionally been applied to the classical kinetic theory of gases. Extending this approach to the study of electron gases in graphene and related materials leads to the celebrated quantum Boltzmann equation (QBE) Greenwood (1958); Fritz et al. (2008); Lux and Fritz (2013, 2012); Dumitrescu (2015); Müller and Sachdev (2008); Zarenia et al. (2019a, b).

In recent work, the present authors presented the QBE formalism for bilayer graphene Nguyen et al. (2020). It was then shown in Ref. Wagner et al. (2020) that the QBE results agree well with experimental measurements of the electrical conductivity of suspended bilayer graphene in Nam et al. (2017). Despite its success, the QBE is a heavy-handed approach and this led to the development of the two-fluid model. In bilayer graphene, the low-energy bandstructure consists of two gapless quadratic bands which can be populated with electrons and holes. The dynamics of the electron and hole fluids can be captured accurately from simple hydrodynamic equations, at least for the calculation of the electrical conductivity and thermopower Wagner et al. (2020).

In this work, we generalize the formalism we developed for BLG to multilayer graphene (MLG), in particular, we focus on Bernal (AB) stacked multilayers. We consider the special case of an even number of layers NN, to avoid the additional complication of the linear band that arises for odd NN. We study the regime near charge neutrality, i.e. βμ1\beta\mu\lesssim 1, where β\beta is inverse temperature and μ\mu is the chemical potential. In fact, in this regime, we expect the behaviour for even NN and odd N+1N+1 to be very similar, since the density of states is dominated by the quadratic bands. We use the QBE to compute the electrical conductivity, the thermal conductivity and thermopower for multilayers with N=2N=2 to N=8N=8 layers. We discuss how the transport properties evolve, as the number NN of layers is increased. In particular, in previous work Wagner et al. (2020); Nguyen et al. (2020), the present authors discussed two signatures of the hydrodynamic regime: The Wiedemann-Franz law violation and the fast increase of the electrical conductivity away from charge neutrality. We will show that both of the signatures remain, as we increase the number of layers in our graphene multilayer. We then develop a hydrodynamic approach in terms of a multicomponent fluid and show that it accurately matches the QBE predictions.

There has been previous theoretical work on transport in multilayer graphene using the Kubo formula Nakamura and Hirasawa (2008). Ref. Min et al. (2011) does study MLG using the Boltzmann formalism, however, they focus on the case where impurities are the main source of scattering. Ref. Lindsay et al. (2011) calculates the thermal conductivity due to phonons, however they do not explore the electronic contribution to the thermal conductivity.

The electrical conductivity of multilayer graphene has been measured experimentally for a range of temperatures and densities Nam et al. (2016, 2018). In Ref. Wang et al. (2008), measurements on the minimum of the electrical conductivity for different numbers of layers are reported. Further experiments by a different experimental group have been reported Ye et al. (2011), however, they consider the high-density regime which is the opposite limit to the one we will consider in this work.

The structure of our paper is as follows. We start by introducing the tight-binding model for MLG and we discuss the screening of the Coulomb interaction in MLG. We then introduce the QBE formalism and present the numerical results for different numbers of layers. Finally, we discuss how many of the salient features of our numerics can be captured by a simple hydrodynamic model.

II Bandstructure and interactions

For the Bernal (AB) stacking of NN graphene multilayers shown in Fig. 1, the tight-binding (Bloch) Hamiltonian expanded near the KK and KK^{\prime} valleys is the 2N×2N2N\times 2N matrix Min and MacDonald (2008); Nakamura and Hirasawa (2008); Koshino and Ando (2007); Latil and Henrard (2006); Guinea et al. (2006); Ohta et al. (2007); Nilsson et al. (2008); Koshino and McCann (2010)

Refer to caption
Figure 1: AB stacked multilayer with NN=4 layers. We use a tight-binding model with nearest neighbour intralayer (t0t_{0}) and interlayer (tt_{\perp}) hopping.
H=(0vπ000000vπ0t000000t0vπ0t0000vπ0000000000vπ0000t0vπ0t000000t0vπ000000vπ0).H=\left(\begin{array}[]{ccccccccc}0&v\pi^{\dagger}&0&0&0&0&0&0&\\ v\pi&0&t_{\perp}&0&0&0&0&0&\\ 0&t_{\perp}&0&v\pi^{\dagger}&0&t_{\perp}&0&0&\\ 0&0&v\pi&0&0&0&0&0&\\ 0&0&0&0&0&v\pi^{\dagger}&0&0&\\ 0&0&t_{\perp}&0&v\pi&0&t_{\perp}&0&\\ 0&0&0&0&0&t_{\perp}&0&v\pi^{\dagger}&\\ 0&0&0&0&0&0&v\pi&0&\\ &&&&&&&&\ddots\\ \end{array}\right).

Here π=ξpx+ipy\pi=\xi p_{x}+ip_{y}, with valley index ξ=±\xi=\pm. The Fermi velocity is v=32at0v=\frac{\sqrt{3}}{2}at_{0}, where t0t_{0} is the intralayer hopping parameter and aa is the lattice constant. tt_{\perp} is the interlayer hopping. The wavefunction is ψ=(φA1,φB1,φA2,φB2,,φAN,φBN)\psi=(\varphi_{A_{1}},\varphi_{B_{1}},\varphi_{A_{2}},\varphi_{B_{2}},\cdots,\varphi_{A_{N}},\varphi_{B_{N}}), where φAi(Bi)\varphi_{A_{i}(B_{i})} is the wave function of an electron at site Ai(Bi)A_{i}(B_{i}).

At low energies, we can focus on the gapless bands. We write down a low energy N×NN\times N Hamiltonian for these NN quadratic bands, which we label with r=(R,σ)r=(R,\sigma) where R=1,,N/2R=1,\dots,N/2 and σ=±1\sigma=\pm 1. The energies are

εRσ(p)=σp22mR,\varepsilon_{R\sigma}(p)=\sigma\frac{p^{2}}{2m_{R}}, (1)

where the mass is

mR=2mcos(RπN+1),m_{R}=2m^{*}\ \textrm{cos}\left(\frac{R\pi}{N+1}\right), (2)

with m=t/2v2m^{*}=t_{\perp}/2v^{2}. Therefore the bands appear in pairs labelled by the same RR which have the same mass. For a fixed RR, we have the same bandstructure as in BLG. We call the corresponding Bloch wavefunctions |ψRσ(𝐩)|\psi_{R\sigma}(\mathbf{p})\rangle. The matrix elements for the Bloch functions are

MRσ,Rσ(𝐩,𝐩)\displaystyle M_{R\sigma,R^{\prime}\sigma^{\prime}}(\mathbf{p},\mathbf{p}^{\prime}) ψRσ(𝐩)|ψRσ(𝐩)\displaystyle\equiv\langle\psi_{R^{\prime}\sigma^{\prime}}(\mathbf{p}^{\prime})|\psi_{R\sigma}(\mathbf{p})\rangle (3)
=δRR2(1+σσe2i(θpθp)),\displaystyle=\frac{\delta_{RR^{\prime}}}{2}(1+\sigma\sigma^{\prime}e^{-2i(\theta_{p}-\theta_{p^{\prime}})}),

where cosθp=𝐩𝐱^\cos\theta_{p}=\mathbf{p}\cdot\hat{\mathbf{x}}. The derivations of the effective mass mRm_{R} as well as the matrix elements (42) are left for Appendix A. There will be no vertex coupling electrons with different RR in the Coulomb interaction. For a pair of bands with the same value of RR, the matrix elements are the same as for BLG. Using this result, one can perform the classic Lindhart calculation for the polarization Π0(q,ω)\Pi^{0}(\vec{q},\omega) in the limit βμ1\beta\mu\lesssim 1 and βq2/m1\beta q^{2}/m\lesssim 1. This is a calculation analogous to Refs. Lv and Wan (2010); Min et al. (2012) and the details are in Appendix B. We focus on the static polarization, which is valid at low enough temperatures. The result for the polarization is

Π0(q,0)=Nfm2π(1sin(π2(N+1))1),\Pi^{0}(\vec{q},0)=-\frac{N_{f}m^{*}}{2\pi}\bigg{(}\frac{1}{\sin(\frac{\pi}{2(N+1)})}-1\bigg{)}, (4)

where Nf=2×2N_{f}=2\times 2 accounts for the spin and valley degrees of freedom. In the screening calculation we have assumed that the screening due to the phonons is negligible. The Thomas-Fermi screening wavevector is then given by

qTF(q)=Π0(𝐪,0)2πα,q_{TF}(q)=-\Pi^{0}(\mathbf{q},0)2\pi\alpha, (5)

where α\alpha is the electromagnetic fine-structure constant. We will use the fully screened Coulomb interaction V(q)=2πα/qTF(q)V(q)=2\pi\alpha/q_{TF}(q), which is a good approximation at low temperatures, where the typical momentum of electrons is much smaller than qTFq_{TF}. Now we approximately have the behaviour qTFNq_{TF}\propto N. Since each electron can now scatter off NN species of electrons, the electron-electron scattering rate will be τee1N/qTF21/N\tau_{ee}^{-1}\propto N/q_{TF}^{2}\propto 1/N.

III Quantum Boltzmann equation

Away from charge neutrality, one needs to include momentum-relaxing scattering in order to obtain a well-defined conductivity. Based on the results in bilayer graphene, we expect electron-phonon collisions to be the dominant source of momentum-relaxing scattering and hence this is the only momentum-relaxing mechanism that we include in our calculations Wagner et al. (2020). Depending on the experimental conditions, we may envisage electron-impurity and electron-boundary scattering as well and this would be a simple extension of the present calculation. The phonon scattering is proportional to band mass and we extract the proportionality constant by comparing the BLG results to available experimental data Wagner et al. (2020). The QBE is an evolution equation for the distribution function fr(𝐤,𝐱,t)f_{r}(\mathbf{k},\mathbf{x},t) of the particles of species rr of the form

(t+𝐯r(𝐤)𝐱+e𝐄𝐤)fr(𝐤,𝐱,t)=Ir[{fri}](𝐤,𝐱,t),\left(\frac{\partial}{\partial t}+\mathbf{v}_{r}(\mathbf{k})\cdot\frac{\partial}{\partial\mathbf{x}}+e\mathbf{E}\cdot\frac{\partial}{\partial\mathbf{k}}\right)f_{r}(\mathbf{k},\mathbf{x},t)\\ =-I_{r}[\{f_{r_{i}}\}](\mathbf{k},\mathbf{x},t), (6)

where 𝐯r(𝐤)=𝐤ϵr(𝐤)\mathbf{v}_{r}(\mathbf{k})=\partial_{\mathbf{k}}\epsilon_{r}(\mathbf{k}), e<0e<0 is the electron charge and the collision integral on the RHS includes electron-electron and electron-phonon collisions. The electron-electron collision integral can be derived from the Kadanoff-Baym equations Kadanoff and Baym (1962) using the Born approximation. The derivation is identical to the BLG case in Ref. Nguyen et al. (2020). The electron-phonon collision integral uses the simple relaxation-time approximation with scattering rate

τep,r1=D2mrkBT2ρ3c2,\tau_{ep,r}^{-1}=\frac{D^{2}m_{r}k_{B}T}{2\rho\hbar^{3}c^{2}}, (7)

where DD is the deformation potential, ρ\rho is the mass density of multilayer graphene and cc is the speed of sound. We also define the corresponding dimensionless parameter αep=βτep=βm/mrτep,r1\alpha_{ep}=\beta\tau_{ep}=\beta m^{*}/m_{r}\tau_{ep,r}^{-1}. The full details of the QBE are shown in appendix C. We note that ρN\rho\propto N and c=const.c=\textrm{const.}, so assuming that DD only depends weakly on NN, we have αep1/N\alpha_{ep}\propto 1/N. We now see that both the electron-electron and the electron-phonon scattering rates behave like 1/N1/N, although the reasons behind this scaling are very different for the two scattering mechanisms. Based on this simple scaling, it stands to reason that the hydrodynamic window in multilayer graphene is similar to that of BLG: τee1/τep1\tau_{ee}^{-1}/\tau_{ep}^{-1} is only weakly NN-dependent. Since we successfully applied a hydrodynamic model to BLG Wagner et al. (2020), we expect this to work for MLG as well.

In order to solve the QBE, we expand the distribution function in terms of 4N4N basis functions. Based on our previous work Wagner et al. (2020) this is a sufficient number of basis functions to obtain a convergent result. The QBE then turns into an equation for the expansion coefficients in front of the basis functions. Once we know the perturbation of the distribution function due to an applied thermal gradient T\nabla T or electric field 𝐄\mathbf{E}, we can compute the electrical current

𝐉=Nferd2k(2π)2σ𝐤mrfr(𝐤),\mathbf{J}=N_{f}e\sum_{r}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{\sigma\mathbf{k}}{m_{r}}f_{r}(\mathbf{k}), (8)

and heat current

𝐉Q=Nfrd2k(2π)2σ𝐤mr(ϵr(𝐤)μ)fr(𝐤).\mathbf{J}^{Q}=N_{f}\sum_{r}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{\sigma\mathbf{k}}{m_{r}}(\epsilon_{r}(\mathbf{k})-\mu)f_{r}(\mathbf{k}). (9)

We define the electrical conductivity σ\sigma, the thermal conductivity KK and the thermopower Θ\Theta by

(𝐉𝐉Q)=(σΘTΘK)(𝐄T).\begin{pmatrix}\mathbf{J}\\ \mathbf{J}^{Q}\end{pmatrix}=\begin{pmatrix}\sigma&\Theta\\ T\Theta&K\end{pmatrix}\begin{pmatrix}\mathbf{E}\\ -\mathbf{\nabla}T\end{pmatrix}. (10)

Note that the open circuit thermal conductivity κ\kappa measuring the heat current in the absence of electrical current is given by κ=KTΘσ1Θ\kappa=K-T\Theta\sigma^{-1}\Theta. In the absence of a magnetic field, the transport coefficients are diagonal. In Fig. 2 we plot the dimensionless transport coefficients

σ~xx2Nfe2σxx,\tilde{\sigma}_{xx}\equiv\frac{2}{N_{f}e^{2}}\sigma_{xx}, (11)
Θ~xx2NfekBΘxx,\tilde{\Theta}_{xx}\equiv\frac{2}{N_{f}ek_{B}}\Theta_{xx}, (12)
K~xx2NfkB2TKxx\tilde{K}_{xx}\equiv\frac{2}{N_{f}k_{B}^{2}T}K_{xx} (13)

for different values of NN.

Refer to caption
Figure 2: Results from the QBE calculation for different even values of NN. We plot the normalized electrical conductivity σ~xx(βμ)/σ~xx(βμ=0)\tilde{\sigma}_{xx}(\beta\mu)/\tilde{\sigma}_{xx}(\beta\mu=0), thermal conductivity K~xx(βμ)/K~xx(βμ=0)\tilde{K}_{xx}(\beta\mu)/\tilde{K}_{xx}(\beta\mu=0) and thermopower Θ~xx(βμ)/Θ~xx(βμ=1)\tilde{\Theta}_{xx}(\beta\mu)/\tilde{\Theta}_{xx}(\beta\mu=1). We have set αep=0.1/N\alpha_{ep}=0.1/N as in our previous work on BLG.
Refer to caption
Figure 3: Results for the thermal conductivity K~(μ=0)/N2\tilde{K}(\mu=0)/N^{2} from the QBE calculation for different even values of NN. We have set αep=0.1/N\alpha_{ep}=0.1/N as in our previous work on BLG.

If we could treat the NN-layer multilayer as N/2N/2 independent bilayers, then we would expect the transport coefficients to increase proportionally to NN. However, this is not what happens. Indeed we find approximately K(βμ=0)N2K(\beta\mu=0)\propto N^{2} in Fig. 3. The reason for this behaviour is that KK at charge neutrality is limited by collisions with phonons – it would diverge in the absence of phonon scattering since Coulomb scattering does not relax the mode where all carriers move at the same velocity. Recall the formula from basic kinetic theory KΛkB2nτepT/mK\sim\Lambda k_{B}^{2}n\tau_{ep}T/m, where ΛkB\Lambda k_{B} is the heat capacity per particle and nn is the number density. Now τ1/N\tau\propto 1/N as explained in the previous section. We also have nNn\propto N. This explains the observed K(βμ=0)N2K(\beta\mu=0)\propto N^{2} behaviour.

We plot the results for the electrical conductivity at charge neutrality as a function of NN in Fig. 4. We observe that the conductivity scales with NN approximately as σ(βμ=0)N2\sigma(\beta\mu=0)\propto N^{2}. Recall the Drude formula σe2nτ/m\sigma\sim e^{2}n\tau/m, where τ\tau is the collision time for current-relaxing collisions. For different species the individual conductivities will add up, ie σre2nrτr/mr\sigma\sim\sum_{r}e^{2}n_{r}\tau_{r}/m_{r}. σ(βμ=0)\sigma(\beta\mu=0) is well-defined even in the absence of phonons and indeed electron-electron collisions will dominate the current relaxation. As we increase NN, we increase the density of states and hence the screening wavevector scales approximately as qTFNq_{TF}\propto N and hence the potential scales as V1/NV\sim 1/N. Therefore Fermi’s Golden rule for scattering of particles of species rr off particles of species rr^{\prime} yields τrr1|V|21/N2\tau_{rr^{\prime}}^{-1}\propto|V|^{2}\propto 1/N^{2}. The scattering time for species rr then roughly scales as τr1rτrr11/N\tau_{r}^{-1}\equiv\sum_{r^{\prime}}\tau_{rr^{\prime}}^{-1}\propto 1/N. So τrN\tau_{r}\propto N and with nrmrn_{r}\propto m_{r}, we find σre2nrτr/mrrτrN2\sigma\sim\sum_{r}e^{2}n_{r}\tau_{r}/m_{r}\propto\sum_{r}\tau_{r}\propto N^{2}, as in the numerical results.

Refer to caption
Figure 4: Results for the electrical conductvity σ~(μ=0)/N2\tilde{\sigma}(\mu=0)/N^{2} from the QBE calculation for different even values of NN. We have set αep=0.1/N\alpha_{ep}=0.1/N as in our previous work on BLG.

Let us discuss two signatures of the hydrodynamic regime: (i) the ratio σ(βμ=1)/σ(βμ=0)\sigma(\beta\mu=1)/\sigma(\beta\mu=0) and (ii) the Wiedemann-Franz law violation. The ratio σ(1)/σ(0)\sigma(1)/\sigma(0) stays relatively constant as NN is increased. Recall that for bilayer graphene, the reason for the large value of σ(1)/σ(0)\sigma(1)/\sigma(0) is that σ(βμ=0)\sigma(\beta\mu=0) is limited by electron-electron collisions, which operate on a time-scale τee\tau_{ee}. On the other hand, away from CN the momentum mode carries charge and this momentum mode is relaxed on a much longer time-scale τep\tau_{ep}. Since τepτee\tau_{ep}\gg\tau_{ee} in the hydrodynamic regime, charge transport is greatly enhanced away from CN. Since both τee\tau_{ee} and and τep\tau_{ep} scale proportional to NN, σ(1)/σ(0)\sigma(1)/\sigma(0) does not vary significantly with NN. In the hydrodynamic regime, the Lorenz number is much larger than predicted by the Wiedemann-Franz law. Since σ\sigma increases as fast as KK with NN, the violation of the Wiedemann-Franz law at charge neutrality will also remain relatively constant as a function of NN. We show plots of σ(1)/σ(0)\sigma(1)/\sigma(0) and the Lorenz number in Appendix E.

In Fig. 5 we show the results for the three transport properties considered for the representative case of N=8N=8. Other even values of NN yield similar results. As NN\to\infty and we approach the graphite limit, our numerics become unmanageable and a full 3d theory becomes necessary, where the bands are dispersive along kzk_{z}, instead of having large number NN of 2d bands as in our 2d model. In fact, due to the approximations we have made, the low energy theory we have derived is valid in the limit N100N\ll 100 111In (21), the gap Δr=εr+(p=0)εr(p=0)=2γcos(rπ/(N+1))\Delta_{r}=\varepsilon_{r}^{+}(p=0)-\varepsilon_{r}^{-}(p=0)=2\gamma\cos(r\pi/(N+1)). The smallest gap therefore occurs when r=N/2N+1r=\frac{N/2}{N+1} and is Δmin2γsin(π/N)2γπ/N\Delta_{\textrm{min}}\approx 2\gamma\sin(\pi/N)\approx 2\gamma\pi/N. We want the gap to be so large that we can neglect the gapped bands, i.e. Δmin1β\Delta_{\textrm{min}}\gg\frac{1}{\beta}. Using γ0.3\gamma\approx 0.3eV McCann and Koshino (2013) and for T=40KT=40K this leads to N100N\ll 100. .

IV multi-fluid model

Following the usual procedure for deriving hydrodynamic equations from kinetic theory, we can obtain the fluid equations from the full QBE. We have rr species of fermions in the low energy theory, and in the hydrodynamic description, we can associate each fluid species with a mean velocity 𝐮r\mathbf{u}^{r}. The equation of motion that follows from the QBE for 𝐮r\mathbf{u}^{r} under an applied electric field 𝐄\mathbf{E} and thermal gradient T\mathbf{\nabla}T is

mrt𝐮r=rmrτrr(𝐮r𝐮r)mr𝐮rτep,r+σe𝐄ΛrkBT,m_{r}\partial_{t}\mathbf{u}^{r}=-\sum_{r^{\prime}}\frac{m_{r}}{\tau_{rr^{\prime}}}(\mathbf{u}^{r}-\mathbf{u}^{r^{\prime}})-\frac{m_{r}\mathbf{u}^{r}}{\tau_{ep,r}}\\ +\sigma e\mathbf{E}-\Lambda^{r}k_{B}\nabla T, (14)

where τrr\tau_{rr^{\prime}} is the effective scattering time of particles of species rr off particles of species rr^{\prime} due to Coulomb interactions, τep,r\tau_{ep,r} is the effective electron-phonon scattering time for species rr and Λr\Lambda^{r} is the entropy per particle:

Λ(R,+)=d2𝐩(2π)2[(1fr0(𝐩))ln(1fr0(𝐩))+fr0(𝐩)lnfr0(𝐩)]d2𝐩(2π)2fr0(𝐩),\Lambda^{(R,+)}=-\frac{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\left[\left(1-f^{0}_{r}(\mathbf{p})\right)\ln\left(1-f^{0}_{r}(\mathbf{p})\right)+f^{0}_{r}(\mathbf{p})\ln f^{0}_{r}(\mathbf{p})\right]}{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}f^{0}_{r}(\mathbf{p})}, (15)
Λ(R,)=d2𝐩(2π)2[(1fr0(𝐩))ln(1fr0(𝐩))+fr0(𝐩)lnfr0(𝐩)]d2𝐩(2π)2[1fr0(𝐩)].\Lambda^{(R,-)}=-\frac{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\left[\left(1-f^{0}_{r}(\mathbf{p})\right)\ln\left(1-f^{0}_{r}(\mathbf{p})\right)+f^{0}_{r}(\mathbf{p})\ln f^{0}_{r}(\mathbf{p})\right]}{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}[1-f^{0}_{r}(\mathbf{p})]}. (16)

Solving the fluid equations for a steady state flow yields an expression for 𝐮r\mathbf{u}^{r}. The electrical current is then given by

𝐉=erσnr𝐮r\mathbf{J}=e\sum_{r}\sigma n^{r}\mathbf{u}^{r} (17)

and the heat current by

𝐐=kBTrΛrnr𝐮r.\mathbf{Q}=k_{B}T\sum_{r}\Lambda^{r}n^{r}\mathbf{u}^{r}. (18)

The detailed derivation is in Appendix D.

Refer to caption
Figure 5: Dimensionless transport coefficient calculated for the representative case N=8N=8. Comparison of the QBE results with the multi-fluid model introduced in the main text. The value for the Coulomb scattering strength α0\alpha_{0} to use in the multi-fluid model has been extracted from the QBE results, such that σ~xx(βμ=0)\tilde{\sigma}_{xx}(\beta\mu=0) is the same for both plots (ie we have one fitting parameter). The multi-fluid model performs well for the electrical conductivity and the thermopower, but a bit less well for the thermal conductivity.

In Fig. 5 we compare the results from the multi-fluid model and the QBE and find that for the electrical conductivity the agreement is excellent, whereas for the thermal conductivity, the qualitative behaviour is correct but the quantitative agreement is off by around 20%. The reason for this is that the multi-fluid model is equivalent to solving the QBE by using only NN basis functions in the expansion of the distribution function. These basis functions correspond to uniform motion with velocity 𝐮r\mathbf{u}^{r} of the fermions of species rr. For an applied electric field, these modes capture the charge transport accurately, as exemplified by the good overlap in Fig. 5. On the other hand, for an applied thermal gradient, we do not accurately capture the heat transport with those modes. We found the same situation in Ref. Wagner et al. (2020) for BLG.

The success of the multi-fluid model as well as the two fluid model in our previous work on BLG Wagner et al. (2020) suggests that the hydrodynamic description of electrons in bilayer and multi-layer graphene is accurate. This once again confirms the idea that electrons in strongly interacting systems can be considered as (multi-component) fluids Lucas and Fong (2018).

V summary

We have applied the quantum Boltzmann formalism to study the transport properties of multilayer graphene. We find results very similar to bilayer graphene. We introduce a hydrodynamic model which agrees accurately with the QBE results for the electrical conductivity and thermopower. We hope that future experiments on transport in multilayer graphene will reveal whether the QBE formalism performs as well for multilayer graphene as it does for BLG, although we see no apparent reason why it should not.

We have only studied even NN in this paper. For odd NN, the low energy theory consists of N1N-1 parabolic bands and one Dirac cone. However, in the regime βμ1\beta\mu\lesssim 1 the density of states will be dominated by the quadratic bands. Therefore the results for odd NN are expected to be similar to the results for even NN, as long as one accounts for the different values of the band masses.

The behaviour of the transport properties as the number NN of layers is varied shows some interesting features. Firstly, the thermal conductivity at charge neutrality (CN) K(βμ=0)K(\beta\mu=0) is approximately proportional to N2N^{2}. This is due to the fact that the thermal conductivity at CN is limited by phonons and the phonon scattering time is proportional to NN, so K(βμ=0)nτN2K(\beta\mu=0)\propto n\tau\propto N^{2}. The electrical conductivity at CN σ(βμ=0)N2\sigma(\beta\mu=0)\propto N^{2} as well, but for a different reason. In contrast to K(βμ=0)K(\beta\mu=0), σ(βμ=0)\sigma(\beta\mu=0) is limited by electron-electron collisions. As NN is increased, the screening increases and so the electron scattering time τN\tau\propto N, leading to σ(βμ=0)nτN2\sigma(\beta\mu=0)\propto n\tau\propto N^{2}. Put together, this implies that the violation of the Wiedemann-Franz law stays constant as NN is increased. Finally, σ(βμ=1)/σ(βμ=0)\sigma(\beta\mu=1)/\sigma(\beta\mu=0), which is another measure of the relative size of the electron-electron and the electron-phonon scattering times. is relatively flat as a function of NN.

In future work, we plan to compute the viscosity for MLG. Adding the viscosity to the multi-fluid model will give us the Navier-Stokes equations, which can then be used to simulate the electron fluid in MLG for realistic geometries. We expect those simulations to yield interesting results such as the vortices which have been predicted for single-layer graphene Levitov and Falkovich (2016) and negative resistivity, which has been seen in experiments in single-layer graphene Bandurin et al. (2016). One can go even further and consider spin-transport by applying a weak magnetic field. We then have a very interesting multi-component fluid which carries charge, heat and spin.

VI Acknowledgements

This work was supported by EPSRC grants EP/N01930X/1 and EP/S020527/1. DXN was supported partially by Brown Theoretical Physics Center.

Appendix A Derivation of effective mass mrm_{r} and matrix element Mrr(𝐩,𝐩)M_{rr^{\prime}}(\mathbf{p},\mathbf{p}^{\prime})

A.1 Low-energy band theory and matrix elements

We use the effective Hamiltonian from Ref. Min and MacDonald (2008). For Bernal (AB) stacking of NN graphene multilayers, the tight-binding Hamiltonian is the 2N×2N2N\times 2N matrix

H=(0vπ000000vπ0γ000000γ0vπ0γ0000vπ0000000000vπ0000γ0vπ0γ000000γ0vπ000000vπ0).H=\left(\begin{array}[]{ccccccccc}0&v\pi^{\dagger}&0&0&0&0&0&0&\\ v\pi&0&\gamma&0&0&0&0&0&\\ 0&\gamma&0&v\pi^{\dagger}&0&\gamma&0&0&\\ 0&0&v\pi&0&0&0&0&0&\\ 0&0&0&0&0&v\pi^{\dagger}&0&0&\\ 0&0&\gamma&0&v\pi&0&\gamma&0&\\ 0&0&0&0&0&\gamma&0&v\pi^{\dagger}&\\ 0&0&0&0&0&0&v\pi&0&\\ &&&&&&&&\ddots\\ \end{array}\right). (19)

Here π=ξpx+ipy\pi=\xi p_{x}+ip_{y}, with valley index ξ=±\xi=\pm. The Fermi velocity is v=32at0v=\frac{\sqrt{3}}{2}at_{0}, where t0t_{0} is the intralayer hopping parameter and aa is the lattice constant. γ=t\gamma=t_{\perp} is the interlayer hopping. The Schrödinger equation becomes H|ψr±=εr±|ψr±H|\psi_{r}^{\pm}\rangle=\varepsilon_{r}^{\pm}|\psi_{r}^{\pm}\rangle where the 2N2N eigenfunctions are

|ψr±=(φA1φB1φA2φB2φANφBN)|\psi_{r}^{\pm}\rangle=\begin{pmatrix}\varphi_{A_{1}}\\ \varphi_{B_{1}}\\ \varphi_{A_{2}}\\ \varphi_{B_{2}}\\ \vdots\\ \varphi_{A_{N}}\\ \varphi_{B_{N}}\end{pmatrix} (20)

and the eigenenergies are

εr±=γcos(rπN+1missing)±(vp)2+γ2cos(rπN+1missing)2\varepsilon_{r}^{\pm}=\gamma\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})\pm\sqrt{(vp)^{2}+\gamma^{2}\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})^{2}} (21)

for r=1Nr=1...N. Let us focus on NN even for now, in which case there are 2N2N quadratic bands, of which NN bands, |ψr|\psi_{r}^{-}\rangle, are at low energies (gapless). For odd NN there is also a Dirac cone and we will avoid the complications coming from that situation. The low energy bands are

εr(p)={p22mrif cos(rπN+1missing)<0p22mrif cos(rπN+1missing)>0\varepsilon_{r}^{-}(p)=\left\{\begin{array}[]{ll}\frac{p^{2}}{2m_{r}}&\textrm{if }\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})<0\\ -\frac{p^{2}}{2m_{r}}&\textrm{if }\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})>0\end{array}\right. (22)

where

mr=|γcos(rπN+1missing)|v2=2m|cos(rπN+1missing)|m_{r}=\frac{\bigg{|}\gamma\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})\bigg{|}}{v^{2}}=2m^{*}\bigg{|}\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})\bigg{|} (23)

where m=γ/2v2m^{*}=\gamma/2v^{2}. So the bands come in pairs with the same effective mass mrm_{r}, the bands related by r+r=N+1r+r^{\prime}=N+1 are such pairs. Let us call them conjugate bands.

A.2 Low energy effective theory

In the low energy limit εrpv\varepsilon_{r}\ll pv we can also write down a low-energy effective Hamiltonian. The Schrödinger equation of the full Hamiltonian is

vπφB2n1=εφA2n1v\pi^{\dagger}\varphi_{B_{2n-1}}=\varepsilon\varphi_{A_{2n-1}} (24)
γ(φA2n2+φA2n)+vπφA2n1=εφB2n1\gamma(\varphi_{A_{2n-2}}+\varphi_{A_{2n}})+v\pi\varphi_{A_{2n-1}}=\varepsilon\varphi_{B_{2n-1}} (25)
vπφA2n=εφB2nv\pi\varphi_{A_{2n}}=\varepsilon\varphi_{B_{2n}} (26)
γ(φB2n1+φB2n+1)+vπφB2n=εφA2n\gamma(\varphi_{B_{2n-1}}+\varphi_{B_{2n+1}})+v\pi^{\dagger}\varphi_{B_{2n}}=\varepsilon\varphi_{A_{2n}} (27)

We can now eliminate φA2n\varphi_{A_{2n}} and φB2n1\varphi_{B_{2n-1}} from these equations and use επv\varepsilon\ll\pi v. We can then write these equations as the Schrödinger equation for the simpler effective Hamiltonian

Heff=h+hH_{\textrm{eff}}=h+h^{\dagger} (28)

where

h=12m(000000π200000000000π20π2000000000π20π20π20)h=-\frac{1}{2m^{*}}\begin{pmatrix}0&0&0&0&0&0&\dots\\ \pi^{2}&0&0&0&0&0&\dots\\ 0&0&0&0&0&0&\dots\\ -\pi^{2}&0&\pi^{2}&0&0&0&\dots\\ 0&0&0&0&0&0&\dots\\ \pi^{2}&0&-\pi^{2}&0&\pi^{2}&0&\dots\\ \vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\ddots\end{pmatrix} (29)

To solve this, we note that φA0=φAN+1=0\varphi_{A_{0}}=\varphi_{A_{N+1}}=0 and φB0=φBN+1=0\varphi_{B_{0}}=\varphi_{B_{N+1}}=0 so we try the ansatz

φAn=Asin(nrπN+1missing),φBn=Bsin(nrπN+1missing)\varphi_{A_{n}}=A\sin\bigg(\frac{nr\pi}{N+1}\bigg{missing}),\qquad\varphi_{B_{n}}=B\sin\bigg(\frac{nr\pi}{N+1}\bigg{missing}) (30)

This ansatz works and the reduced (Bloch) wavefunction is then (for the KK valley)

|ψr(𝐩)=(φA1φB1φA2φB2φANφBN)=2N+1(e2iθpsinrπ/(N+1)sin2rπ/(N+1)e2iθpsin3rπ/(N+1))|\psi_{r}(\mathbf{p})\rangle=\begin{pmatrix}\varphi_{A_{1}}\\ \varphi_{B_{1}}\\ \varphi_{A_{2}}\\ \varphi_{B_{2}}\\ \vdots\\ \varphi_{A_{N}}\\ \varphi_{B_{N}}\end{pmatrix}=\sqrt{\frac{2}{N+1}}\begin{pmatrix}e^{-2i\theta_{p}}\sin r\pi/(N+1)\\ \sin 2r\pi/(N+1)\\ e^{-2i\theta_{p}}\sin 3r\pi/(N+1)\\ \vdots\end{pmatrix} (31)

with eigenvalues εr(𝐩)\varepsilon_{r}^{-}(\mathbf{p}). We can easily see that for N=2N=2 we obtain the same results as previously for BLG.

A.3 Matrix elements

We define the matrix elements Mrr(𝐩,𝐩)M_{rr^{\prime}}(\mathbf{p},\mathbf{p^{\prime}}) as

Mrr(𝐩,𝐩)\displaystyle M_{rr^{\prime}}(\mathbf{p},\mathbf{p^{\prime}}) ψr(𝐩)|ψr(𝐩)\displaystyle\equiv\langle\psi_{r^{\prime}}(\mathbf{p}^{\prime})|\psi_{r}(\mathbf{p})\rangle (32)
=2N+1[e2i(θpθp)n oddsin(nrπN+1missing)sin(nrπN+1missing)\displaystyle=\frac{2}{N+1}\bigg{[}e^{-2i(\theta_{p}-\theta_{p^{\prime}})}\sum_{\textrm{n odd}}\sin\bigg(\frac{nr\pi}{N+1}\bigg{missing})\sin\bigg(\frac{nr^{\prime}\pi}{N+1}\bigg{missing}) (33)
+n evensin(nrπN+1missing)sin(nrπN+1missing)]\displaystyle+\sum_{\textrm{n even}}\sin\bigg(\frac{nr\pi}{N+1}\bigg{missing})\sin\bigg(\frac{nr^{\prime}\pi}{N+1}\bigg{missing})\bigg{]} (34)

where 1nN1\leq n\leq N. Using the trigonometric identities

n oddsin(nrπN+1missing)sin(nrπN+1missing)={N+14if r=rN+14if r+r=N+10otherwise\sum_{\textrm{n odd}}\sin\bigg(\frac{nr\pi}{N+1}\bigg{missing})\sin\bigg(\frac{nr^{\prime}\pi}{N+1}\bigg{missing})=\left\{\begin{array}[]{ll}\frac{N+1}{4}&\textrm{if }r=r^{\prime}\\ \frac{N+1}{4}&\textrm{if }r+r^{\prime}=N+1\\ 0&\textrm{otherwise}\end{array}\right. (35)

and

n evensin(nrπN+1missing)sin(nrπN+1missing)={N+14if r=rN+14if r+r=N+10otherwise,\sum_{\textrm{n even}}\sin\bigg(\frac{nr\pi}{N+1}\bigg{missing})\sin\bigg(\frac{nr^{\prime}\pi}{N+1}\bigg{missing})=\left\{\begin{array}[]{ll}\frac{N+1}{4}&\textrm{if }r=r^{\prime}\\ -\frac{N+1}{4}&\textrm{if }r+r^{\prime}=N+1\\ 0&\textrm{otherwise}\end{array},\right. (36)

we obtain

Mrr(𝐩,𝐩)={12(1+e2i(θpθp))if r=r12(1+e2i(θpθp))if r+r=N+10otherwiseM_{rr^{\prime}}(\mathbf{p},\mathbf{p^{\prime}})=\left\{\begin{array}[]{ll}\frac{1}{2}(1+e^{-2i(\theta_{p}-\theta_{p^{\prime}})})&\textrm{if }r=r^{\prime}\\ \frac{1}{2}(-1+e^{-2i(\theta_{p}-\theta_{p^{\prime}})})&\textrm{if }r+r^{\prime}=N+1\\ 0&\textrm{otherwise}\end{array}\right. (37)

We will find it useful to introduce a more appropriate notation for the even NN case, namely (r,N+1r)(R,σ)(r,N+1-r)\to(R,\sigma) where R=rmod(N2+1)=1N/2R=r\ \textrm{mod}(\frac{N}{2}+1)=1...N/2 and σ=+,\sigma=+,-, where

σ={+if cos(rπN+1missing)<0i.e.r>N/2if cos(rπN+1missing)>0i.e.rN/2\sigma=\left\{\begin{array}[]{ll}+&\textrm{if }\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})<0\qquad\textrm{i.e.}\ r>N/2\\ -&\textrm{if }\cos\bigg(\frac{r\pi}{N+1}\bigg{missing})>0\qquad\textrm{i.e.}\ r\leq N/2\end{array}\right. (38)

In this notation we have paired up conjugate bands and hence

εRσ(p)=σp22mR\varepsilon_{R\sigma}(p)=\sigma\frac{p^{2}}{2m_{R}} (39)

and

MRσ,Rσ(𝐩,𝐩)=δRR12(σσ+e2i(θpθp))M_{R\sigma,R^{\prime}\sigma^{\prime}}(\mathbf{p},\mathbf{p}^{\prime})=\delta_{RR^{\prime}}\frac{1}{2}(\sigma\sigma^{\prime}+e^{-2i(\theta_{p}-\theta_{p^{\prime}})}) (40)

Now just make a slight redefinition of our wavefunctions

ψRσ(𝐩)=σψr(𝐩)\psi_{R\sigma}(\mathbf{p})=\sigma\psi_{r}(\mathbf{p}) (41)

and with this additional sign

MRσ,Rσ(𝐩,𝐩)=δRR12(1+σσe2i(θpθp))M_{R\sigma,R^{\prime}\sigma^{\prime}}(\mathbf{p},\mathbf{p}^{\prime})=\delta_{RR^{\prime}}\frac{1}{2}(1+\sigma\sigma^{\prime}e^{-2i(\theta_{p}-\theta_{p^{\prime}})}) (42)

so we just have N/2N/2 copies of the BLG matrix elements, labelled by RR and where we denote particle-hole index by σ\sigma.

We have π=px+ipy\pi=p_{x}+ip_{y} in the KK band and π=px+ipy\pi=-p_{x}+ip_{y} in the KK^{\prime} band. So to treat the KK^{\prime} band we need to replace πK=πK\pi_{K^{\prime}}=-\pi_{K}^{\dagger}. Since only π2\pi^{2} appears in the Hamiltonian, we obtain the KK^{\prime} wavefunctions from the KK wavefunctions by simple complex conjugation. So the matrix elements will also be complex conjugates of each other. However, we have a freedom to choose the overall phase of our wavefunctions, and this allows us to redefine our wavefunctions to cancel off this complex conjugation and we end up with the same matrix elements as for the KK valley. Therefore, the valley degeneracy can be taken into account simply by including a factor of Nf=2×2N_{f}=2\times 2 for the number of fermion species in the calculation (the additional factor of 22 comes from spin degeneracy).

The charge density operator can be derived in the same manner as in Ref. Nguyen et al. (2020), and we obtain the result

ρ(𝐪)=fRRσσd2𝐤(2π)2cRσf(𝐤)cRσf(𝐤+𝐪)MRσ,Rσ(𝐤,𝐤+𝐪),\rho(\mathbf{q})=\sum_{f}\sum_{RR^{\prime}}\sum_{\sigma\sigma^{\prime}}\int\frac{d^{2}\mathbf{k}}{(2\pi)^{2}}c^{\dagger}_{R\sigma f}(\mathbf{k})c_{R^{\prime}\sigma^{\prime}f}(\mathbf{k+q})M_{R\sigma,R^{\prime}\sigma^{\prime}}(\mathbf{k},\mathbf{k+q}), (43)

where cRσf(cRσf)c^{\dagger}_{R\sigma f}(c_{R^{\prime}\sigma^{\prime}f}) is the creation (annihilation) operator of an electron. The result shows that the Coulomb vertex will not allow transitions between bands with different masses and different flavors due to the explicit form of MRσ,Rσ(𝐤,𝐤+𝐪)M_{R\sigma,R^{\prime}\sigma^{\prime}}(\mathbf{k},\mathbf{k+q}) in Eq. (42).

Appendix B RPA screening calculation

In this section, we calculate the screened Coulomb potential in the random phase approximation (RPA). We use the explicit form of the density operator (43) and consider the RPA diagram Fig. 6.

Refer to caption
Figure 6: Feynman diagram for polarization. Note that due to the form of the vertex, the two electrons have the same value of RR and ff.

One can calculate the RPA polarizability and obtain

Π0(q,0)=Nfσ,σR,Rd2k(2π)2fRσ(𝐤+q)fRσ(𝐤)εRσ(𝐤+q)εRσ(𝐤)|MRσ,Rσ(𝐤,𝐤+𝐪)|2\Pi^{0}(\vec{q},0)=-N_{f}\sum_{\sigma,\sigma^{\prime}}\sum_{R,R^{\prime}}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{f^{R^{\prime}\sigma^{\prime}}(\mathbf{k}+\vec{q})-f^{R\sigma}(\mathbf{k})}{\varepsilon_{R^{\prime}\sigma^{\prime}}(\mathbf{k}+\vec{q})-\varepsilon_{R\sigma}(\mathbf{k})}|M_{R\sigma,R^{\prime}\sigma^{\prime}}(\mathbf{k},\mathbf{k+q})|^{2} (44)

and using the δRR\delta_{RR^{\prime}} in the matrix elements

Π0(q,0)=NfRσ,σd2k(2π)2fRσ(𝐤+q)fRσ(𝐤)εRσ(𝐤+q)εRσ(𝐤)|MRσ,Rσ(𝐤,𝐤+𝐪)|2\Pi^{0}(\vec{q},0)=-N_{f}\sum_{R}\sum_{\sigma,\sigma^{\prime}}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{f^{R\sigma^{\prime}}(\mathbf{k}+\vec{q})-f^{R\sigma}(\mathbf{k})}{\varepsilon_{R\sigma^{\prime}}(\mathbf{k}+\vec{q})-\varepsilon_{R\sigma}(\mathbf{k})}|M_{R\sigma,R\sigma^{\prime}}(\mathbf{k},\mathbf{k+q})|^{2} (45)

But now for each RR, the calculation is identical with the BLG case, so in the limit βμ1\beta\mu\ll 1

Π0(q,0)=RΠR0(q,0)=Nf2πRmR=Nfm2πR=1N/22cos(πRN+1missing)=Nfm2π(1sin(π2(N+1))1)\Pi^{0}(\vec{q},0)=\sum_{R}\Pi^{0}_{R}(\vec{q},0)=\frac{N_{f}}{2\pi}\sum_{R}m_{R}=\frac{N_{f}m^{*}}{2\pi}\sum_{R=1}^{N/2}2\cos\bigg(\frac{\pi R}{N+1}\bigg{missing})=\frac{N_{f}m^{*}}{2\pi}\bigg{(}\frac{1}{\sin(\frac{\pi}{2(N+1)})}-1\bigg{)} (46)

Appendix C Details of quantum Boltzmann equation

In this section we follow Ref. Nguyen et al. (2020). In the low-energy bandstructure of multilayer graphene with an even number NN of layers, there are NN quadratic bands, which we label by r=(R,σ)r=(R,\sigma), where R=1,2,N/2R=1,2,\cdots N/2 and σ=±\sigma=\pm. The band energy is

εRσ(p)=σp22mR\varepsilon_{R\sigma}(p)=\sigma\frac{p^{2}}{2m_{R}} (47)

The band mass is mR=2m|cos((Rπ)/(N+1))|m_{R}=2m^{*}|\textrm{cos}((R\pi)/(N+1))|, where m=0.033mem^{*}=0.033m_{e}. The equilibrium distribution of the electrons in band r=(R,σ)r=(R,\sigma) is given by the Fermi distribution

fr(𝐩)=f0(ϵr(𝐩))=11+eβ(ϵr(𝐩)μ).f_{r}(\mathbf{p})=f^{0}(\epsilon_{r}(\mathbf{p}))=\frac{1}{1+e^{\beta(\epsilon_{r}(\mathbf{p})-\mu)}}. (48)

We write the deviation from the equilibrium distribution as

fr(𝐩)=f0(ϵr(𝐩))+f0(ϵr(𝐩))[1f0(ϵr(𝐩))]hr(𝐩)f_{r}(\mathbf{p})=f^{0}(\epsilon_{r}(\mathbf{p}))+f^{0}(\epsilon_{r}(\mathbf{p}))[1-f^{0}(\epsilon_{r}(\mathbf{p}))]h_{r}(\mathbf{p}) (49)

and expand the Boltzmann equation up to first order in hr(𝐩)h_{r}(\mathbf{p}). The Boltzmann equation is now a set of NN equations

2πσβmrfr0(𝐩)[1fr0(𝐩)](e𝐄𝐩1TT𝐩(ϵr(𝐩)μ))=Irtot[hri(𝐤i)](𝐩)\frac{2\pi\sigma\beta}{m_{r}}f^{0}_{r}(\mathbf{p})[1-f^{0}_{r}(\mathbf{p})]\bigg{(}e\mathbf{E}\cdot\mathbf{p}-\frac{1}{T}\nabla T\cdot\mathbf{p}(\epsilon_{r}(\mathbf{p})-\mu)\bigg{)}=I_{r}^{\textrm{tot}}[h_{r_{i}}(\mathbf{k}_{i})](\mathbf{p}) (50)

The LHS of the QBE includes the driving force due to the electric field 𝐄\mathbf{E} and the thermal gradient T\mathbf{\nabla}T. The collision integral is

Irtot[hri(𝐤i)](𝐩)=Ir,Coul(1)[hri(𝐤i)](𝐩)\displaystyle I_{r}^{\textrm{tot}}[h_{r_{i}}(\mathbf{k}_{i})](\mathbf{p})=I_{r,\textrm{Coul}}^{(1)}[h_{r_{i}}(\mathbf{k}_{i})](\mathbf{p}) 1τep,rfr0(p)[1fr0(p)]hr(𝐩)\displaystyle-\frac{1}{\tau_{ep,r}}f_{r}^{0}(p)[1-f_{r}^{0}(p)]h_{r}(\mathbf{p}) (51)

The second term on the RHS of Eq. (51) is the contribution to the collision integral coming from electron-phonon collisions, for which the scattering rate is

τep,r1=D2mrkBT2ρ3c2,\tau_{ep,r}^{-1}=\frac{D^{2}m_{r}k_{B}T}{2\rho\hbar^{3}c^{2}}, (52)

where DD is the deformation potential, ρ\rho is the mass density of multilayer graphene and cc is the speed of sound. Let us define the corresponding dimensionless number

αep=βτep1=D2m2ρ3c2\alpha_{ep}=\beta\tau_{ep}^{-1}=\frac{D^{2}m^{*}}{2\rho\hbar^{3}c^{2}} (53)

The first term on the RHS of Eq. (51) is the linearized collision integral for scattering between electrons which is

Ir,Coul(1)[hri(𝐤i)](𝐩)\displaystyle I_{r,\textrm{Coul}}^{(1)}[h_{r_{i}}(\mathbf{k}_{i})](\mathbf{p}) =(2π)r1r2r3d2𝐤(2π)2d2𝐪(2π)2δ(ϵr(𝐩)+ϵr1(𝐤)ϵr2(𝐩+𝐪)ϵr3(𝐤𝐪))\displaystyle=-(2\pi)\sum_{r_{1}r_{2}r_{3}}\int\frac{\mathop{}\!\mathrm{d}^{2}\mathbf{k}}{(2\pi)^{2}}\frac{\mathop{}\!\mathrm{d}^{2}\mathbf{q}}{(2\pi)^{2}}\delta(\epsilon_{r}(\mathbf{p})+\epsilon_{r_{1}}(\mathbf{k})-\epsilon_{r_{2}}(\mathbf{p+q})-\epsilon_{r_{3}}(\mathbf{k-q}))
×[Nf|Trr1r3r2(𝐩,𝐤,𝐪)|2Trr1r3r2(𝐩,𝐤,𝐪)Trr1r2r3(𝐩,𝐤,𝐤𝐩𝐪)]\displaystyle\times\bigg{[}N_{f}|T_{rr_{1}r_{3}r_{2}}(\mathbf{p},\mathbf{k},\mathbf{q})|^{2}-T_{rr_{1}r_{3}r_{2}}(\mathbf{p},\mathbf{k},\mathbf{q})T_{rr_{1}r_{2}r_{3}}^{*}(\mathbf{p},\mathbf{k},\mathbf{k-p-q})\bigg{]} (54)
×[[1fr0(𝐩)][1fr10(𝐤)]fr20(𝐩+𝐪)fr30(𝐤𝐪)]\displaystyle\times\bigg{[}[1-f_{r}^{0}(\mathbf{p})][1-f_{r_{1}}^{0}(\mathbf{k})]f_{r_{2}}^{0}(\mathbf{p+q})f_{r_{3}}^{0}(\mathbf{k-q})\bigg{]}
×[hr(𝐩)hr1(𝐤)+hr2(𝐩+𝐪)+hr3(𝐤𝐪)]\displaystyle\times\bigg{[}-h_{r}(\mathbf{p})-h_{r_{1}}(\mathbf{k})+h_{r_{2}}(\mathbf{p+q})+h_{r_{3}}(\mathbf{k-q})\bigg{]}

The matrix elements in (C) are

Tr1r2r3r4(k,k,q)=V(q)Mr1r4(k+q,k)Mr2r3(kq,k)T_{r_{1}r_{2}r_{3}r_{4}}(\vec{k},\vec{k^{\prime}},\vec{q})=V(-\vec{q})M_{r_{1}r_{4}}(\vec{k}+\vec{q},\vec{k})M_{r_{2}r_{3}}(\vec{k^{\prime}-q},\vec{k^{\prime}}) (55)

with

Mr,r(𝐩,𝐩)=δRR12(1+σσe2i(θpθp))M_{r,r^{\prime}}(\mathbf{p},\mathbf{p}^{\prime})=\delta_{RR^{\prime}}\frac{1}{2}(1+\sigma\sigma^{\prime}e^{-2i(\theta_{p}-\theta_{p^{\prime}})}) (56)

and with screened Coulomb potential

V(𝐪)=2πNfm(1sin(π2(N+1))1)1V(\mathbf{q})=\frac{2\pi}{N_{f}m^{*}}\bigg{(}\frac{1}{\sin(\frac{\pi}{2(N+1)})}-1\bigg{)}^{-1} (57)

The equations for the charge current and heat current are

𝐉=Nferd2k(2π)2σ𝐤mrfr(𝐤),\mathbf{J}=N_{f}e\sum_{r}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{\sigma\mathbf{k}}{m_{r}}f_{r}(\mathbf{k}), (58)
𝐉Q=Nfrd2k(2π)2σ𝐤mr(ϵr(𝐤)μ)fr(𝐤).\mathbf{J}^{Q}=N_{f}\sum_{r}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{\sigma\mathbf{k}}{m_{r}}(\epsilon_{r}(\mathbf{k})-\mu)f_{r}(\mathbf{k}). (59)

In the case where we only have an applied electric field 𝐄\mathbf{E}, the suggested ansatz to solve the QBE (50) is Wagner et al. (2020); Nguyen et al. (2020); Lux and Fritz (2012)

hr(𝐩)=βe𝐄m𝐩χr(p).h_{r}(\mathbf{p})=\beta\frac{e\mathbf{E}}{m^{*}}\cdot\mathbf{p}\chi_{r}(p). (60)

We expand (60) in terms of basis functions

χr(k)=βnangn(r,k)\chi_{r}(k)=\beta\sum_{n}a_{n}g_{n}(r,k) (61)

such that the ana_{n} are dimensionless. Here the basis functions are taken to be

gn(r,k)=δr=1,δr=2,δr=N,δr=1K,δr=2K,δr=NK,g_{n}(r,k)=\delta_{r=1},\delta_{r=2},...\delta_{r=N},\delta_{r=1}K,\delta_{r=2}K,...\delta_{r=N}K,... (62)

where K=β/mkK=\sqrt{\beta/m}k is the dimensionless momentum. For all powers n>2n>2 we multiply by an exponential factor so the basis function is KneK/2K^{n}e^{-K/2}. We expand in up to 4N4N basis functions. Increasing the number of basis function changes the results only marginally. We use the fact that this must be valid for all 𝐄\mathbf{E}, sum over rr, multiply separately by 𝐩^gm(r,p)\mathbf{\hat{p}}g_{m}(r,p) and integrate over 𝐩\mathbf{p}. This yields an equation that can be summarized in matrix form as

Mmnan=FmM_{mn}a_{n}=F_{m} (63)

where we defined the dimensionless matrices

Mmn=β(βm)3/2rd2𝐩(2π)2gm(r,p)Ir[{𝐩^𝐤ign(ri,ki)}](𝐩)M_{mn}=\beta\bigg{(}\frac{\beta}{m}\bigg{)}^{3/2}\sum_{r}\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}g_{m}(r,p)I_{r}\bigg{[}\bigg{\{}\mathbf{\hat{p}}\cdot\mathbf{k}_{i}g_{n}(r_{i},k_{i})\bigg{\}}\bigg{]}(\mathbf{p}) (64)

and the dimensionless vector

Fm=β(βm)1/2rd2𝐩(2π)2σpmrfr0(p)[1fr0(p)]gm(r,p)F_{m}=\beta\bigg{(}\frac{\beta}{m}\bigg{)}^{1/2}\sum_{r}\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\frac{\sigma p}{m_{r}}f^{0}_{r}(p)[1-f^{0}_{r}(p)]g_{m}(r,p) (65)

(63) can be inverted to yield 𝐚\mathbf{a}. The charge current is

𝐉\displaystyle\mathbf{J} =Nfremrd2𝐩(2π)2σ𝐩fr(𝐩)\displaystyle=N_{f}\sum_{r}\frac{e}{m_{r}}\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\sigma\mathbf{p}f_{r}(\mathbf{p}) (66)
=βNfrd2𝐩(2π)2σ𝐩fr0(𝐩)[1fr0(𝐩)]e2𝐄mmr𝐩χr(p).\displaystyle=\beta N_{f}\sum_{r}\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\sigma\mathbf{p}f^{0}_{r}(\mathbf{p})[1-f^{0}_{r}(\mathbf{p})]\frac{e^{2}\mathbf{E}}{m^{*}m_{r}}\cdot\mathbf{p}\chi_{r}(p).

The DC conductivity is read off as

σxx=βNfrd2𝐩(2π)2σfr0(p)[1fr0(p)]e2px2mrmχr(p)=Nfe22𝐆M1𝐅,\sigma_{xx}=\beta N_{f}\sum_{r}\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\sigma f^{0}_{r}(p)[1-f^{0}_{r}(p)]\frac{e^{2}p_{x}^{2}}{m_{r}m^{*}}\chi_{r}(p)=\frac{N_{f}e^{2}}{2\hbar}\mathbf{G}\cdot M^{-1}\mathbf{F}, (67)

where we have exceptionally restored \hbar and where the dimensionless vector

Gm=β(βm)rd2𝐩(2π)2σp2mrfr0(p)[1fr0(p)]gm(r,p).G_{m}=\beta\bigg{(}\frac{\beta}{m}\bigg{)}\sum_{r}\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\frac{\sigma p^{2}}{m_{r}}f^{0}_{r}(p)[1-f^{0}_{r}(p)]g_{m}(r,p). (68)

The thermal conductivity and thermopower can be calculated completely analogously.

Appendix D Multi-fluid model

We can now derive a multi-fluid model. We assume that the electrons/holes in band (R,+/)(R,+/-) have mean velocity 𝐮(R,+/)\mathbf{u}^{(R,+/-)}. The corresponding ansatz for the perturbation of the distribution function is hr(𝐤)=β𝐤𝐮rh_{r}(\mathbf{k})=\beta\mathbf{k}\cdot\mathbf{u}^{r}. We obtain the fluid equations by multiplying the QBE (50) by 𝐩\mathbf{p} and integrating over 𝐩\mathbf{p}. We then divide by the number density nrn^{r} to obtain the coupled set of equations

mrt𝐮r=rmrτrr(𝐮r𝐮r)mr𝐮rτep,r+σe𝐄ΛrkBT.m_{r}\partial_{t}\mathbf{u}^{r}=-\sum_{r^{\prime}}\frac{m_{r}}{\tau_{rr^{\prime}}}(\mathbf{u}^{r}-\mathbf{u}^{r^{\prime}})-\frac{m_{r}\mathbf{u}^{r}}{\tau_{ep,r}}+\sigma e\mathbf{E}-\Lambda^{r}k_{B}\nabla T. (69)

Remember that we have defined e<0e<0 as the electron charge. We define Λr\Lambda^{r} through the integrals

Λ(R,+)=d2𝐩(2π)2βp2mrβ(ϵr(p)μ)fr0(𝐩)[1fr0(𝐩)]d2𝐩(2π)2fr0(𝐩)\Lambda^{(R,+)}=\frac{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\frac{\beta p^{2}}{m_{r}}\beta(\epsilon_{r}(p)-\mu)f^{0}_{r}(\mathbf{p})[1-f^{0}_{r}(\mathbf{p})]}{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}f^{0}_{r}(\mathbf{p})} (70)
Λ(R,)=d2𝐩(2π)2βp2mrβ(ϵr(p)+μ)fr0(𝐩)[1fr0(𝐩)]d2𝐩(2π)2(1fr0(𝐩)).\Lambda^{(R,-)}=\frac{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\frac{\beta p^{2}}{m_{r}}\beta(-\epsilon_{r}(p)+\mu)f^{0}_{r}(\mathbf{p})[1-f^{0}_{r}(\mathbf{p})]}{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}(1-f^{0}_{r}(\mathbf{p}))}. (71)

Note that Λr=Λ(R,σ)\Lambda^{r}=\Lambda^{(R,\sigma)} only depends on σ\sigma, not rr itself, since the species mass mrm_{r} drops out when we de-dimensionalize. So this is in fact exactly the same expression as in BLG. These integrals are in fact the entropy per particle

Λ(R,+)=d2𝐩(2π)2[(1fr0(𝐩))ln(1fr0(𝐩))+fr0(𝐩)lnfr0(𝐩)]d2𝐩(2π)2fr0(𝐩),\Lambda^{(R,+)}=-\frac{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\left[\left(1-f^{0}_{r}(\mathbf{p})\right)\ln\left(1-f^{0}_{r}(\mathbf{p})\right)+f^{0}_{r}(\mathbf{p})\ln f^{0}_{r}(\mathbf{p})\right]}{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}f^{0}_{r}(\mathbf{p})}, (72)
Λ(R,)=d2𝐩(2π)2[(1fr0(𝐩))ln(1fr0(𝐩))+fr0(𝐩)lnfr0(𝐩)]d2𝐩(2π)2[1fr0(𝐩)],\Lambda^{(R,-)}=-\frac{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\left[\left(1-f^{0}_{r}(\mathbf{p})\right)\ln\left(1-f^{0}_{r}(\mathbf{p})\right)+f^{0}_{r}(\mathbf{p})\ln f^{0}_{r}(\mathbf{p})\right]}{\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}[1-f^{0}_{r}(\mathbf{p})]}, (73)

The number density of species rr is

nr=Nfmr2πβln(1+eσβμ)n^{r}=\frac{N_{f}m_{r}}{2\pi\beta}\ln(1+e^{\sigma\beta\mu}) (74)

where σ=+/\sigma=+/- depending on whether we are dealing with a particle or a hole band. In order to obtain the scattering time between species rr and rr^{\prime} due to Coulomb interactions, we need to solve the equation

d2𝐩(2π)2𝐩Ir,Coul(1)[hri(𝐤ri)=β𝐤ri𝐮ri](𝐩)=rnrmrτrr(𝐮r𝐮r).\int\frac{d^{2}\mathbf{p}}{(2\pi)^{2}}\mathbf{p}I_{r,\textrm{Coul}}^{(1)}\bigg{[}h_{r_{i}}(\mathbf{k}_{r_{i}})=\beta\mathbf{k}_{r_{i}}\cdot\mathbf{u}^{r_{i}}\bigg{]}(\mathbf{p})=-\sum_{r^{\prime}}\frac{n^{r}m_{r}}{\tau_{rr^{\prime}}}(\mathbf{u}^{r}-\mathbf{u}^{r^{\prime}}). (75)

Instead of explicitly computing this collision integral for all βμ\beta\mu and all rr, a reasonable first guess (that needs to be checked against the QBE results) would be

τrr1=nrr′′nr′′(mrr′′mr′′)1/2τee1\tau_{rr^{\prime}}^{-1}=\frac{n^{r^{\prime}}}{\sum_{r^{\prime\prime}}n^{r^{\prime\prime}}}\bigg{(}\frac{m_{r}}{\sum_{r^{\prime\prime}}m_{r^{\prime\prime}}}\bigg{)}^{-1/2}\tau_{ee}^{-1} (76)

such that we only need to evaluate τee\tau_{ee}. To see where this guess comes from, remember that basic kinetic theory yields τrr1nrΣvr\tau_{rr^{\prime}}^{-1}\sim n^{r^{\prime}}\Sigma\langle v_{r}\rangle (Σ\Sigma is the collision cross-section) and vrkBT/mr\langle v_{r}\rangle\sim\sqrt{k_{B}T/m_{r}}. We then plug the steady-state solution of the fluid equations (69) into the formula for the electrical current

𝐉=erσnr𝐮r\mathbf{J}=e\sum_{r}\sigma n^{r}\mathbf{u}^{r} (77)

and heat current

𝐐=kBTrΛrnr𝐮r.\mathbf{Q}=k_{B}T\sum_{r}\Lambda^{r}n^{r}\mathbf{u}^{r}. (78)

We can change this into a 1d problem in the absence of a magnetic field. To this end consider the steady-state form of equation (69)

0=rmrτrr(𝐮r𝐮r)mr𝐮rτep,r+σe𝐄ΛrkBT0=-\sum_{r^{\prime}}\frac{m_{r}}{\tau_{rr^{\prime}}}(\mathbf{u}^{r}-\mathbf{u}^{r^{\prime}})-\frac{m_{r}\mathbf{u}^{r}}{\tau_{ep,r}}+\sigma e\mathbf{E}-\Lambda^{r}k_{B}\nabla T (79)

We can turn this into a matrix equation by defining the scattering rate between species as

Γrr=δrr(r′′τrr′′1+τep,r1)τrr1\Gamma_{rr^{\prime}}=\delta_{rr^{\prime}}\bigg{(}\sum_{r^{\prime\prime}}\tau_{rr^{\prime\prime}}^{-1}+\tau_{ep,r}^{-1}\bigg{)}-\tau_{rr^{\prime}}^{-1} (80)

such that (79) becomes

mrrΓrrur=σeEΛrkBT.m_{r}\sum_{r^{\prime}}\Gamma_{rr^{\prime}}u^{r^{\prime}}=\sigma eE-\Lambda^{r}k_{B}\nabla T. (81)

The solution is obtained by taking the matrix inverse

ur=r1mrΓrr1(σeEΛrkBT)u^{r}=\sum_{r^{\prime}}\frac{1}{m_{r}}\Gamma_{rr^{\prime}}^{-1}(\sigma^{\prime}eE-\Lambda^{r^{\prime}}k_{B}\nabla T) (82)

Plugging (82) into the electrical current (77) one finds

𝐉=erσnrr1mrΓrr1(σeEΛrkBT)\mathbf{J}=e\sum_{r}\sigma n^{r}\sum_{r^{\prime}}\frac{1}{m_{r}}\Gamma_{rr^{\prime}}^{-1}(\sigma^{\prime}eE-\Lambda^{r^{\prime}}k_{B}\nabla T) (83)

and similarly for the heat current (78) one obtains

𝐐=kBTrΛrnrr1mrΓrr1(σeEΛrkBT).\mathbf{Q}=k_{B}T\sum_{r}\Lambda^{r}n^{r}\sum_{r^{\prime}}\frac{1}{m_{r}}\Gamma_{rr^{\prime}}^{-1}(\sigma^{\prime}eE-\Lambda^{r^{\prime}}k_{B}\nabla T). (84)

Finally, this leads to the expressions for the electrical conductivity

σxx=e2rrnrmrσσΓrr1=Nfe2rrσσn~rαrr1Nfe22σ~xx\sigma_{xx}=e^{2}\sum_{rr^{\prime}}\frac{n^{r}}{m_{r}}\sigma\sigma^{\prime}\Gamma_{rr^{\prime}}^{-1}=N_{f}e^{2}\sum_{rr^{\prime}}\sigma\sigma^{\prime}\tilde{n}^{r}\alpha_{rr^{\prime}}^{-1}\equiv\frac{N_{f}e^{2}}{2}\tilde{\sigma}_{xx} (85)

the thermopower

Θxx=NfekBrrnrmrΓrr1Λrσ=NfekBrrσn~rαrr1ΛrNfekB2Θ~xx\Theta_{xx}=N_{f}ek_{B}\sum_{rr^{\prime}}\frac{n^{r}}{m_{r}}\Gamma_{rr^{\prime}}^{-1}\Lambda^{r^{\prime}}\sigma=N_{f}ek_{B}\sum_{rr^{\prime}}\sigma\tilde{n}^{r}\alpha_{rr^{\prime}}^{-1}\Lambda^{r^{\prime}}\equiv\frac{N_{f}ek_{B}}{2}\tilde{\Theta}_{xx} (86)

and the thermal conductivity (modulo the usual caveat about the open-circuit thermal conductivity)

Kxx=kB2TrrnrmrΛrΓrr1Λr=NfkB2Trrn~rΛrαrr1ΛrNfkB2T2K~xxK_{xx}=k_{B}^{2}T\sum_{rr^{\prime}}\frac{n^{r}}{m_{r}}\Lambda^{r}\Gamma_{rr^{\prime}}^{-1}\Lambda^{r^{\prime}}=N_{f}k_{B}^{2}T\sum_{rr^{\prime}}\tilde{n}^{r}\Lambda^{r}\alpha_{rr^{\prime}}^{-1}\Lambda^{r^{\prime}}\equiv\frac{N_{f}k_{B}^{2}T}{2}\tilde{K}_{xx} (87)

where we have de-dimensionalized by defining

n~r=βnrNfmr=ln(1+eσβμ)2π\tilde{n}^{r}=\beta\frac{n^{r}}{N_{f}m_{r}}=\frac{\ln(1+e^{\sigma\beta\mu})}{2\pi} (88)

and

αrr=βΓrr.\alpha_{rr^{\prime}}=\beta\Gamma_{rr^{\prime}}. (89)

In analogy with BLG we define dimensionless tildered quantities, which for conciseness we reproduce here again

σ~xx=2rrσσn~rαrr1\tilde{\sigma}_{xx}=2\sum_{rr^{\prime}}\sigma\sigma^{\prime}\tilde{n}^{r}\alpha_{rr^{\prime}}^{-1} (90)
Θ~xx=2rrσn~rαrr1Λr\tilde{\Theta}_{xx}=2\sum_{rr^{\prime}}\sigma\tilde{n}^{r}\alpha_{rr^{\prime}}^{-1}\Lambda^{r^{\prime}} (91)
K~xx=2rrn~rΛrαrr1Λr.\tilde{K}_{xx}=2\sum_{rr^{\prime}}\tilde{n}^{r}\Lambda^{r}\alpha_{rr^{\prime}}^{-1}\Lambda^{r^{\prime}}. (92)

Appendix E Supplementary figures

Refer to caption
Figure 7: Results for (μ=0)/WF\mathcal{L}(\mu=0)/\mathcal{L}_{WF}, where =σ/κT\mathcal{L}=\sigma/\kappa T is the Lorenz number and WF=π2/3(kB2/e2)\mathcal{L}_{WF}=\pi^{2}/3(k_{B}^{2}/e^{2}) is the Wiedemann-Franz result. We present the QBE results for different even values of NN. We have set αep=0.1/N\alpha_{ep}=0.1/N as in our previous work on BLG.
Refer to caption
Figure 8: Results for σ~(βμ=1)/σ~(βμ=0)\tilde{\sigma}(\beta\mu=1)/\tilde{\sigma}(\beta\mu=0) from the QBE calculation for different even values of NN. We have set αep=0.1/N\alpha_{ep}=0.1/N as in our previous work on BLG.

References