1 INTRODUCTION
Graphene is a two-dimensional material that displays unusual electronic properties due to the conical shape of the energy bands in the proximity of the so-called Dirac points. For this reason, charge transport in graphene attracts a lot of interest from the scientific community. In the literature, there are several papers describing electron hydrodynamics with semiclassical or quantum models [1-7]. In particular, the viscous hydrodynamic regime, which is only possible in a time-scale where electron–electron collisions are dominant, has been deeply investigated both theoretically [3, 5, 6] and experimentally [8]. In refs. [3-6], the author reports a derivation of the viscous hydrodynamic equations based on a semiclassical Boltzmann equation, that is, a kinetic description that is classical except for some key elements such as the conical energy bands and the use of Fermi statistics. On the other hand, quantum-corrected hydrodynamic equations (obtained from a quantum kinetic description based on the Wigner equation) have been obtained only in the case of an inviscid regime and assuming regularized energy bands [2]. This is because the singular behavior of the bands at the conical intersection causes the appearance of divergent integrals when computing quantum corrections to the transport coefficients.
The aim of the present work is to explore the possibility of deriving fully-quantum and quantum-corrected hydrodynamic equations in both the inviscid and viscous regimes. Hence, we are implicitly assuming a time scale where the momentum conserving electron-electron collisions are dominant with respect to the electron-phonon (or electron-impurity) collisions [3]. For the sake of simplicity, we only consider the isothermal case, since the nonisothermal case would only bring more calculation but no methodological differences. We also work with regularized bands, by introducing a regularization parameter (with the physical meaning of an energy gap); however, the conical limit is considered whenever possible.
Quantum hydrodynamic models date back to the early period of quantum mechanics. The Schrödinger equation can be formulated in a hydrodynamic form, the so-called Madelung equations [9], which represent an isothermal Euler system with an additional term, called Bohm potential [10, 11]. Madelung's approach is restricted to pure states, and so it is not suitable to introduce statistical concepts, such as equilibrium states, which are needed in the derivation of more general quantum fluid models. A more systematic way to derive quantum hydrodynamic models was proposed by Degond, Ringhofer, and Méhats [12, 13] by introducing the quantum maximum entropy principle (QMEP) to close the systems of the macroscopic moments. In refs. [12, 13], the authors, similarly to the classical maximum entropy principle [14, 15] and to Levermore's method [16], close the moment system by means of a quantum distribution function which maximizes the entropy with given moments. This approach is fairly general and can be applied to different systems and different regimes.
Our derivation starts from the one-particle Wigner equation [17, 18] with a BGK collisional term, which describes the relaxation of the system to the local equilibrium. Here, we are interested in the isothermal fluid equations for the particle density and for the momentum density . Following the lines of Ringhofer–Degond–Méhats' theory, we obtain the local equilibrium Wigner function by using the QMEP, which determines the local equilibrium by maximizing an entropy functional with the constraints given by the macroscopic moments. The equilibrium Wigner function can be formally derived and it contains an implicit dependence upon the moments and through the dependence on a set of three Lagrange multipliers, and . The quantum Navier–Stokes equations for graphene are then obtained by performing the limit of the Wigner equation as the mean free path converges to zero, by using the Chapman–Enskog expansion.
At order 0 in the collision-time parameter, we obtain the quantum analog of the classical Euler equations, while at order 1 the viscous corrections to the Euler model are obtained. Finally, we determine the expression of macroscopic moments as functions of the Lagrange multipliers in the semiclassical case (namely, at second-order in ) and then compute the approximated quantum Navier–Stokes equations. We also discuss the case of nonregularized (conical) energy bands, letting the regularization parameter go to 0, and show that, remarkably, the Euler equations remain nonsingular in this limit.
We remark that our approach is analogous to the one used in ref. [19], also based on the QMEP, where the authors derive for the first time quantum isothermal Navier–Stokes equations for a standard particle (i.e., with a quadratic energy band). In that paper, the authors compute, in a purely formal way, the viscous correction terms for their model. Inspired by this idea, our work positions itself in the framework of mathematical models for the inclusion of viscous effects in graphene, the existing works being focused on the Euler equations. The novelty of our paper is the theoretical calculation of the quantum corrections to the viscous terms for a fluid of electrons in graphene, with the aim of a mathematical study of their form and possibly their numerical estimation (left for future work).
The outline of the paper is the following. In Section 2 we introduce the model recalling the form of the energy bands for graphene and their regularization and write down the Wigner equation with a BGK collisional operator. In Section 3 we perform the Chapman–Enskog expansion and derive quantum Euler equations and quantum Navier–Stokes equations. In Section 4 we compute the semiclassical expansion up to for the equilibrium distribution and consequently for the Navier–Stokes equations. Section 5 contains our conclusions.
2 GRAPHENE ENERGY BANDS AND KINETIC EQUATIONS
Graphene is a monolayer crystal consisting of a honeycomb lattice of carbon atoms. In the tight-binding approximation, the valence and conduction bands have conical intersections at the vertices of the Brillouin zone (the so-called Dirac points), where they touch each other with a null gap [
20]. This leads to consider the carriers as “massless fermions.” In this work, we restrict our model to the conduction band, for which the explicit expression
(1) holds for
in a neighborhood of a given Dirac point. Here,
is the Fermi speed and
is the reduced Planck constant. It is customary in the models involving the graphene bands to place the origin of the Brillouin zone at the Dirac point, so the crystal vector
in Equation (
1) is measured from the Dirac point. The band profile given by Equation (
1) presents a singular point at the origin, which causes certain integrals appearing in the kinetic formulations to diverge. However, one can introduce a very small gap related to the first and second neighbor hopping energy; this assumption provides a smooth version of the energy band [
2, 20] given by
(2) where
is a small parameter, physically related to the nearest-neighbor and next nearest-neighbor hopping energies.
Let
be the single electron Wigner function [
18, 21-23], which is the Wigner transform of the electron density matrix
(3) We suppose that the carrier population lives mainly near the Dirac point of the Brillouin zone, which we have chosen as our origin of the
space. This allows to effectively enlarge the Brillouin zone to the whole of
when integrating over
, and thus to identify the crystal momentum
with the Wigner canonical momentum
so that the Wigner function is viewed as a function of
(instead of
), that is,
.
Let us introduce reference length
, time
, energy
, temperature
and density
. From now on, we shall switch to dimensionless variables, by making the following substitutions:
where, for the sake of simplicity, the new dimensionless quantities are denoted by the same symbols as the old ones. Moreover, we assume that the following relations hold
(4) where
is the Boltzmann constant. Consequently, the dimensionless energy band reads as follows:
(5) where
is the dimensionless momentum and
A crucial quantity associated with the energy band is the semiclassical velocity, which is defined according to the classical Hamiltonian picture:
(6) For the sake of simplicity, in this paper we consider an isothermal electron gas, so we only need the first two hydrodynamic moments, namely the charge density
and the momentum density
:
(7)
(8) With a more compact notation, we define the monomials
(9) and the corresponding moments
(10) so that
(11) In the Wigner (phase-space) formalism, the operator product translates into the Moyal product between two phase space functions
,
:
(12) where
(13) and
(14) is the dimensionless Planck constant. It is important to note that
(15) The time evolution of a single particle is governed by the Wigner equation with collisions [
2]
(16) The left-hand side of this equation accounts for the Hamiltonian dynamics of the electrons and is equivalent to the von Neumann equation for the density matrix. Here,
and
are pseudo-differential operators describing the effect of the crystal lattice and the (dimensionless) external potential
. In terms of the Moyal product we have [
24]:
(17)
(18) (note that our sign convention for
is opposite to that of ref. [
2]). In the following, we will also encounter the symmetric version of
, namely
(19) Definitions (
17), (
18), and (
19) in terms the Fourier integral representation become:
(20)
(21)
(22) where
is the Fourier transform of
with respect to
and
is the Fourier transform of
with respect to to
, with the symbols
for
,
,
.
The right-hand side of Equation (16) accounts for the effect of collisions. According to Ringhofer–Degond–Méhats' theory [12, 13, 19], collisions in a quantum system can be represented by means of an operator of BGK type [25], describing the relaxation of the system toward a local equilibrium in a characteristic (dimensionless) time . The classical BGK operators share some aspects with the Boltzmann collision operator, like for example local conservation of mass, momentum, and energy. What changes in the quantum case is that the local equilibrium state is represented by a suitable Wigner function that maximizes quantum entropy. Since we are considering isothermal hydrodynamics, only mass and momentum conservation will be imposed. The local equilibrium Wigner function will be described in detail in Section 3.2.
3 THE CHAPMAN–ENSKOG EXPANSION
In this section, we derive formally the hydrodynamic equations in a fully quantum picture by using the Chapman–Enskog procedure (see ref. [26] for the classical case and, for example, see ref. [13] for the quantum case).
3.1 Chapman–Enskog expansion and moments
The starting point of the Chapman–Enskog method is the formal expansion of the Wigner function in a series of powers of the collision time
:
(23) As we said before, we are supposed to work in the isothermal case, and that the moments
(i.e.,
and
) are locally conserved hydrodynamic quantities, so we assume that
(24) (a condition that guarantees that the BGK operator conserves
and
).
The second step of the Chapman–Enskog procedure is to expand the time-derivative of the macroscopic moments (not the macroscopic moments themselves) in powers of
:
(25) By substituting the expansion (
23) in the Wigner equation (
16), we obtain
(26)
(27) where
and
stands for the time-derivatives of
approximated at order 0, as it results from the expansion (
25). Indeed, the function
depends on time only through the dependence on the macroscopic quantities
and
, so that
(28) Note that we used the symbol of functional derivative because, as we shall see, the dependence of
on the moments
is deeply nonlocal. Also note that Equations (
24) and (
26) entail
Multiplying the Wigner equation (
16) by
and integrating over
we obtain
(29) Then, by using expression (
27), we obtain up to
,
(30) We have therefore identified
(31) The
equations,
(32) are the equations of the inviscid hydrodynamics and are the quantum analog of the Euler equations of classical hydrodynamics, while the
Equations (
30) give the quantum analog of the classical Navier–Stokes equations. The meaning of
is that the time-derivatives of moments are given by Euler equations (
32).
3.2 Equilibrium
In order to proceed with the Chapman–Enskog expansion, the equilibrium distribution
must be made explicit. As already mentioned in Section
2,
is a Wigner function that satisfies the QMEP. More precisely, according to the QMEP,
is the (unique) maximizer of the von Neumann entropy subject to the constraint that the observed macroscopic moments
and
are determined by Equation (
10). We do not repeat here the derivation of the form of
(which can be found e.g., in refs. [
12, 13, 24]) and we limit ourselves to state the result that one obtains assuming that the entropy is compatible with Fermi–Dirac statistics:
(33) where we have taken into account Equation (
4). Here,
represents the Weyl quantization and is defined, for a phase space function
, by
(34) We recall [
18, 22, 23] that the inverse transform
coincides with the Wigner transform up to the identification of the operator
with its integral kernel
. To this extent, the Wigner transform (
3) and the Weyl transform (
34) are the inverse of each other.
The definition of is completed by imposing the constraints (24). In (33) the scalar function and the vector function , both real-valued, are the Lagrange multipliers that provide the necessary degrees of freedom for to fulfill such constraints.
3.3 Quantum Euler equations
We now express the quantum Euler equations (32) in terms of the Lagrange multipliers.
For the operator
we have
where we used the fact that
, since
is a function of (and therefore commutes with)
(see (
33)). Moreover (using the convention of summing over repeated indices
),
In the last expression, only the terms with
and
are nonzero in the factor
and we obtain:
Comparing with (
18) and (
19) yields
and, therefore,
Then,
(35) Here and in the following section we shall use the identities
(36) Thanks to (
36) we obtain from (
35):
(37) and, for
,
(38) Hence, Equation (
32) reads as follows:
(39) As discussed in Section
3, Equations (
39) are the quantum Euler equations for electrons in graphene. The form of Equation (
39) is similar to that of the quantum Euler equations found in other contexts [
24, 27].
3.4 Quantum Navier–Stokes equations
The next order in the Chapman–Enskog expansion,
, provides the quantum Navier–Stokes equations which include viscosity terms. From Equation (
30) we see that these terms are given by
, for
. We have
where
. Then, by using (
36), we obtain
(40) and
(41) Unfortunately, no simplifying identities like (
36) are available for
. Moreover, the nonpolynomial form of the energy band
makes the Navier–Stokes terms (
40) and (
41) much less treatable than the corresponding terms for quadratic bands [
19].
Turning to the expression
, we recall that
means that the time-derivatives of the moments
are approximated by the Euler equations (
39). The term
cannot be made more explicit, given the lack of a general explicit form for the dependence of the Lagrange multipliers on the moments
. This can only be done in the semiclassical approximation that will be discussed in Section
4. For the moment, we shall just improve a little bit the expression by using Equations (
37) and (
38):
(42) and, for
,
(43) In conclusion we get:
(44) and, for
,
(45) which are the quantum Navier–Stokes equations for electrons in graphene. Here,
means that every time-derivative of the moments
,
, and
must be computed by using
(46) Equations (
44) and (
45) are still in implicit form, since we need to express the Lagrange multipliers
and
in terms of the densities
,
, and
. We shall address this point in the next section, by adopting a semiclassical approach. This will allow us to obtain a closed-form set for Equations (
44) and (
45). We also note that, once this issue has been solved, the equations can directly be used for a numerical approach and thus for applications to real systems.
4 SEMICLASSICAL EXPANSION
In this section we introduce the semiclassical approximation, by expanding all quantities and equations to second-order in
. To this aim, we introduce the phase space function
given by
(47) where the energy band
is given by (
2);
and
are the Lagrange multipliers. For the sake of simplicity, let us assume that the Fermi-Dirac distribution can be approximated with the Maxwell–Boltzmann one. The Fermi–Dirac is certainly more appropriate to the physical regime under consideration but we preferred to avoid technical complications in order to better illustrate the method. Equation (
33) then becomes
(48) which is the so-called
quantum exponential of
[
12, 13, 24, 27]. The quantum exponential admits the semiclassical expansion
(49) where
(50) is the classical exponential and [
13, 22, 24, 27]
(51) where summation convention over repeated indices is adopted. In our case, recalling (
6), we have
Substituting into (
51) yields
(52)
4.1 Semiclassical expansion of the constraints
In order to proceed with the semiclassical approximation, we need to expand the constraints (
24) up to order
, that is,
(53) We begin with the leading order,
. By using the polar representations
we obtain
where
is the zeroth-order modified Bessel function of the first kind. Since
for
(see Remark
1 and e.g., ref. [
28]), we obtain
(54) Now,
and so we have
Moreover,
(55) It is convenient to define
such that
(56) and to introduce the semiclassical current
as
(57) Hence, from the preceding computations, we see that the following relations between the Lagrange multipliers
and
and the hydrodynamic moments
,
and
hold true at leading-order:
(58) where, as usual,
.
Remark 1.Since is a nonconstant probability distribution in , and since (for ), then from Jensen's inequality we have
with the strict inequality sign, which ensures that the consistency relation
is fulfilled (also for
).
Remark 2.For we obtain the relations
(59) which are known to hold true for the conical band (with Maxwell–Boltzmann statistics), see for example, ref. [
6].
To write down the constraints (
53) at second-order, we need to compute
and
, that is,
for
. From (
52) we have:
(60) for
.
The moments (i.e., the expressions in angular brackets) in Equation (60), can be (in principle) computed explicitly as functions of and by using techniques similar to those employed to obtain the relations (58), but the resulting expressions are much more cumbersome.
Note that all the moments appearing in (60) are finite also for (because the only singularity comes from , which behaves like , and thus integrable in dimension two). This implies that the semiclassical expansion of the Lagrange multipliers has no singularities, at least to second-order. On the other hand, the quantum Euler equations (39) have been proven to depend only on the Lagrange multipliers. Hence, we obtain the remarkable result that, in our approach, the second-order semiclassical Euler equations contain no singularities, even without band regularization.
4.2 Inversion of the constraint system
Equations (58) and (60) are the expressions of the hydrodynamic moments as functions of the Lagrange multipliers to second-order (first-order in ). However, in order to get explicit semiclassical Navier–Stokes equations, it is necessary to express the Lagrange multipliers in terms of the hydrodynamic moments in the quantum Navier–Stokes equations (44) and (45). Then, our next task is to invert the constraint system (53).
To better understand this point, let us schematically indicate by
the vector of Lagrange multipliers, and by
the vector of hydrodynamic moments. In the previous section, we have expressed
as a function of
up to order
:
(61) where
and
are given by Equations (
58) and (
60), respectively. Inverting these relations at order
means writing
(62) To determine the functions
,
, and
, we insert (
62) into (
61), which yields (by Taylor expansion)
(63) At leading-order we obtain
(64) corresponding to the inversion of (
58), that is,
(65) where the link with
is given by the leading-order relation
(66) From the first-order equation
(67) we immediately obtain
, so that the second-order inversion reduces to the linear system
(68) for the unknown
(see also ref. [
24]), where
is the Jacobian matrix of the mapping
evaluated at
,
(given by (
65)), and
is given by (
60), also evaluated at
,
. More explicitly, using (
54) and (
55),
(69) where
is the
matrix
(70) Moreover,
where
and
are given by the right-hand side of (
60), for
, with
,
substituted by their leading-order expressions (
65), and
is still given by the leading-order relation (
66). Hence, the second-order corrections to the relations (
65) are obtained by solving the linear system
(71) (together with the relation (
66) between
and
).
4.3 Semiclassical expansion of the quantum Navier–Stokes equations
In the semiclassical approximation, the hydrodynamic equations for electrons in graphene can be summarized in the following theorem.
Theorem 1.The quantum corrected semiclassical model for a regularized energy band, in the isothermal case, reads as
(72) where the Euler terms are
(73)
(74) where
,
,
are given by (
65), (
71), and are functions of the macroscopic moments
and
, while the viscosity terms have the expressions
(75)
(76) (for
).
(77) and, for
,
(78)
Proof.The derivation of the model is reported in Appendix.
Note that and provide the leading-order and second-order approximations of the quantum Euler equations, while and are the analogous approximation of the “viscous” terms. The quantum corrections and depend on the second-order Lagrange multipliers and , but the dependence of the latter on the hydrodynamic unknowns is complicated and it is not explicitly written down here: it can be deduced starting from Equation (71).
The high complexity of the model and the cumbersome equations require a numerical implementation to better understand the role of the terms involved. In ref. [29] a calculation of the bulk and shear viscosity of the electron liquid in a doped graphene sheet is reported. In ref. [30] the authors show a quantitative calculation of the electronic shear and Hall viscosities in graphene. In ref. [5] the authors determine the ratio of the shear viscosity to the entropy density, also discussing the possible consequences of the low viscosity. In this framework, as a natural continuation of our work, it could be desirable to perform a numerical estimation of the viscosity terms derived in the paper and consequently discuss a comparison with the quantities mentioned above.
4.4 The small current regime
A simplified model can be obtained by assuming a regime of small current, namely,
We recall that
and (see (
58))
(79) where
. As in Section
4.2, let
Since now
following the procedure of Section
4.2, the leading-order Equation (
64) yields
(80) In particular,
(81) The first-order Equation (
67) has now a nonvanishing
at the right-hand side and the Jacobian matrix is given by
(82) By evaluating such matrix (
82) at
and
, we see that the first-order equation is
(83) where
This leads to
(84) The second-order inversion equation, given by (
63), requires the evaluation of
at
and
, which leads to the much simpler expression
(85) Hence, by using (
80) and obvious symmetry reasons,
(86) where
(87) The last ingredient we need to invert in (
63) at second-order are the Hessian matrices of
and
evaluated at
and
. Exploiting symmetries, we readily see that the first one is given by
and the second one vanishes. Hence, according to (
63),
and
are given by
that is,
(88) A further simplification is given by the fact that the terms
are of third-order with respect to the small parameters
and
and so one can decide to neglect them in a suitable regime. All such simplifications result in a system of Navier–Stokes-like equations of the form
(89) where the semiclassical Euler terms
are given by (
73), the semiclassical viscosity terms
are given by (
75) and (
76) and the quantum corrections
are given by
(90)
being given by (
81). Note that
is the Bohm potential for our system [
10, 11, 22, 24]. We remark that in the case of nonsmoothed cones, that is for
, the model is nonsingular, since the coefficients
and
remain finite. This was expected, as we already remarked that the singular terms are all in
.
5 CONCLUSIONS
In this work, we have obtained the hydrodynamic equations for a population of conduction electron carriers in graphene, by using a regularized form for the energy band and by assuming a isothermal regime. The macroscopic equation include the inviscid Euler equations and the viscosity terms. The underlying kinetic model is given by the Wigner-function approach, in which the effect of the periodic potential is accounted for by a pseudo-differential operator with the symbol given by the band shape. We use the Chapman–Enskog procedure accompanied by the quantum maximum energy principle, which provides the Wigner equilibrium function in terms of a set of Lagrange multipliers. The relationship between the hydrodynamic quantities and the Lagrange multipliers cannot be inverted in general, thus we obtain hydrodynamic equations which, in their general quantum form, are left implicit. Subsequently, we expand the hydrodynamic equations together with all relevant quantities to second-order in , thus obtaining the semiclassical approximation for the inviscid and viscous hydrodynamic equations. It is worth to remark that, in contrast with previous studies, the semiclassical Euler equations that we obtain by applying the QMEP are nonsingular in the zero-gap limit .
ACKNOWLEDGMENTS
The authors have nothing to report.
Open access publishing facilitated by Universita degli Studi di Firenze, as part of the Wiley - CRUI-CARE agreement.