To begin with, (
10) is transformed into center of mass coordinates in compliance with the transformation rules to define the locations of the density matrix elements in position space
$$\begin{aligned} \chi =\frac{(n+m)}{2} \quad \textrm{and}\quad \xi =n-m. \end{aligned}$$
(13)
Then, the Von Neumann equation becomes a two-staggered grid formulation which differ by the positions of the density matrix elements for each grid. Both sub-grids are offset from each other by half of the discretization width
\(\Delta \chi =a_0/2\) in
\(\chi\)-direction and by the width
\(\Delta \xi =2a_0\) in
\(\xi\)-direction. Here, for example, the matrix elements of the Tight-Binding Hamiltonian can be formed according to the following relations assuming the next nearest neighbor approximation:
$$\begin{aligned} \langle l|{\hat{H}}|k\rangle =\left\{ \begin{array}{ll} \epsilon _{\chi ,\xi }, &{} l=k \\ \gamma _{\chi ,\xi }, &{} l=k\pm 1 \\ 0, &{} \mathrm {otherwise\,.} \end{array}\right. \end{aligned}$$
(14)
These relations can be extended along with the density operator
\(\hat{\rho }\), implying that the Hamiltonian operator will have elements significantly different from zero between atoms which are second- or third-nearest neighbors [
1]. For a selected point (
\(\chi _0, \xi _0\)) the transport equations are set up exemplarily based on (
14):
$$\begin{aligned} \begin{aligned} i\hbar \frac{\partial }{\partial t}{\hat{C}}_{\chi _0,\xi _0}&= \gamma _{\chi _0-\frac{1}{2},\xi _0+1}{\hat{C}}_{\chi _0-\frac{1}{2},\xi _0+1} +\gamma _{\chi _0-\frac{1}{2},\xi _0-1}{\hat{C}}_{\chi _0-\frac{1}{2},\xi _0-1}\\&- \gamma _{\chi _0+\frac{1}{2},\xi _0+1}{\hat{C}}_{\chi _0+\frac{1}{2},\xi _0+1} -\gamma _{\chi _0+\frac{1}{2},\xi _0-1} {\hat{C}}_{\chi _0+\frac{1}{2},\xi _0-1}\\&-{\mathcal {E}}_{\chi _0,\xi _0}\\\ \end{aligned} \end{aligned}$$
(15)
$$\begin{aligned} \begin{aligned} i\hbar \frac{\partial }{\partial t}{\hat{C}}_{\chi ^\textrm{mid}_0,\xi ^\textrm{mid}_0}&= \gamma _{\chi _0,\xi _0}{\hat{C}}_{\chi _0,\xi _0}+\gamma _{\chi _0,\xi _0-2}{\hat{C}}_{\chi _0,\xi _0-2}\\&- \gamma _{\chi _0+1,\xi _0}{\hat{C}}_{\chi _0+1,\xi _0}-\gamma _{\chi _0+1,\xi _0-2}{\hat{C}}_{\chi _0+1,\xi _0-2} \\&-{\mathcal {E}}_{\chi ^\textrm{mid}_0,\xi ^\textrm{mid}_0}\\ \end{aligned} \end{aligned}$$
(16)
It should be noticed that the index "mid" refers to the grid points of the sub-grid g which lie halfway between the points of grid f indicated by Fig.
2. Applying (
15) and (
16) to all points of the density operator, this results in a matrix vector expression with the matrices
\({\varvec{D}}_f\) and
\(\varvec{D}_g\) containing the hopping terms
\(\gamma\) and vectors containing
\({\hat{C}}_{\chi _0,\xi _0}\) and
\({\hat{C}}_{\chi ^\textrm{mid}_0,\xi ^\textrm{mid}_0}\) which are shown below.
$$\begin{aligned} {\varvec{D}}_f = \begin{pmatrix} -\gamma &{} \gamma &{} &{}0 \\ &{} -\gamma &{} \gamma \\ &{} &{} \ddots \\ 0 &{} &{} -\gamma &{} \gamma \\ \end{pmatrix} \end{aligned}$$
(17)
$$\begin{aligned} {\varvec{D}}_g = \begin{pmatrix} \gamma &{} &{} &{}0\\ -\gamma &{} \gamma &{} \\ &{} &{} \ddots \\ &{} &{} -\gamma &{} \gamma \\ 0 &{} &{} &{} -\gamma \\ \end{pmatrix} \end{aligned}$$
(18)
The positions (indices) of the hopping terms are left out since all the terms have the same value. The matrix
\({\varvec{V}}_f\) containing the potential for one of the grids is shown as an example.
$$\begin{aligned} {\varvec{V}}_{f} = \begin{pmatrix} {\mathcal {E}}_{\chi _0,\xi _0} &{} &{} 0 \\ &{} {\mathcal {E}}_{\chi _1,\xi _1} &{} \\ &{} \hspace{5.0pt}\hspace{5.0pt}\hspace{5.0pt}\ddots &{} \\ 0&{} &{} {\mathcal {E}}_{\chi _n,\xi _n} \\ \end{pmatrix} \end{aligned}$$
(19)
Analogously, the matrix
\({\varvec{V}}_g\) can be constructed similarly with its potential values on the main diagonal. Finally, the two general matrix–vector equations result:
$$\begin{aligned} i\hbar \frac{\partial }{\partial t}\hat{{\varvec{C}}}_{\chi ,\xi } = {\varvec{D}}_f\cdot \hat{{\varvec{C}}}_{\chi ,\xi } -{\varvec{V}}_f \cdot \hat{{\varvec{C}}}_{\chi _\textrm{mid},\xi _\textrm{mid}} \end{aligned}$$
(20)
$$\begin{aligned} i\hbar \frac{\partial }{\partial t}\hat{\varvec{C}}_{\chi _\textrm{mid},\xi _\textrm{mid}} = {\varvec{D}}_g\cdot \hat{\varvec{C}}_{\chi _\textrm{mid},\xi _\textrm{mid}}-{\varvec{V}}_g\cdot \hat{\varvec{C}}_{\chi ,\xi } \end{aligned}$$
(21)
The matrices,
\({\varvec{D}}_f\) and
\({\varvec{D}}_g\), which include all the elements of the density matrix related to the hopping between the atoms, are of the size [
\(N_{\xi }-1\times N_\xi\)] and [
\(N_{\xi }\times N_{\xi }-1\)], respectively. The matrix
\({\varvec{V}}_f\) has the dimension [
\(N_{\xi }\times N_\xi\)] and
\({\varvec{V}}_g\) the dimension [
\(N_{\xi }-1\times N_{\xi }-1\)].
\(N_{\xi }\) are the number of elements in
\(\xi -\)direction.
The discrete density matrix consists of two sub-grids offset to each other. One sub-grid is located on even grid points; while, the other contains grid points for odd numbers in
\(\chi\)-direction. The density operator
\(\hat{\rho }\) is now mapped to the corresponding functions
\(f(\chi ,k)\) and
\(g(\chi _\textrm{mid},k)\) with the help of a Wigner–Weyl Transformation introducing a plane wave basis
\(\varvec{\Phi }_f\) and
\(\varvec{\Phi }_g\). In this respect, the Weyl transform according to
$$\begin{aligned} W(\chi ,k)= \int d\xi \exp (-ik\xi )\quad \langle \chi +\xi /2 | \hat{\rho } |\chi -\xi /2 \rangle \end{aligned}$$
(22)
is applied, in which the density operator
\(\hat{\rho }\) in center of mass coordinates is defined by
$$\begin{aligned} \hat{\rho }=|\Psi (\chi +\xi /2)\rangle \langle \Psi (\chi -\xi /2)|. \end{aligned}$$
(23)
The Weyl transform is approximated by using the values at discrete locations
\(\xi _p\), where the density matrix is defined. Hence, we have
$$\begin{aligned} W(\chi ,k)= \sum _p \Delta \xi \exp (-ik\xi _p)\, \langle \chi +\xi _p/2 | \hat{\rho } |\chi -\xi _p/2 \rangle . \end{aligned}$$
(24)
Introducing the expansions
$$\begin{aligned} |\Psi (\chi +\xi /2)\rangle = \sum _p {\hat{c}}_{\chi +\xi _p/2} |n(\chi +\xi _p/2)\rangle \end{aligned}$$
(25)
and
$$\begin{aligned} \langle \Psi (\chi -\xi /2)|= \sum _p {\hat{c}}_{\chi -\xi _p/2}^\dagger \langle n(\chi -\xi _p/2)| \end{aligned}$$
(26)
dependent on the orbital functions
\(|n\rangle\), the Wigner function result
$$\begin{aligned} W(\chi ,k)= \sum _p \Delta \xi \exp (-ik\xi _p)\hspace{5.0pt}{\hat{c}}_{\chi +\xi _p/2} \hspace{5.0pt}{\hat{c}}_{\chi -\xi _p/2}^{\dagger } \end{aligned}$$
(27)
Accordingly, the inverse transform for both grids can be expressed by the use of different expansion coefficients
\(f_{\chi ,k}\) and
\(g_{\chi _\textrm{mid},k}\) as
$$\begin{aligned} {\hat{C}}_{\chi ,\xi } =\frac{1}{2\pi \hbar }\sum _{k_l} \Delta k\exp {(-ik_l\xi )}f_{\chi ,k}. \end{aligned}$$
(28)
and
$$\begin{aligned} {\hat{C}}_{\chi _\textrm{mid},\xi _\textrm{mid}} = \frac{1}{2\pi \hbar }\sum _{k_l}\Delta k \exp {(-ik_l\xi _g) g_{\chi _\textrm{mid},k}} \hspace{5.0pt}, \end{aligned}$$
(29)
where the wave number
\(k\) is defined by
$$\begin{aligned} k = [-L_k/2...,k_l,...L_k/2]\,\Delta k\,, \end{aligned}$$
(30)
with
\(L_k\) as the number of
\(k\)-values, the reduced Planck constant
\(\hbar\) and
\(\Delta k=\pi /(L_k\Delta x)\). The resulting plane wave basis for
\(\varvec{\Phi }_f\) is of the size [
\(N_{\xi }\times N_{L_k}\)] and [
\(N_{\xi -1}\times N_{L_k}\)] for
\(\varvec{\Phi }_g\). In order to arrive at the final Quantum Liouville type equations, the relations (
28) and (
29) with their basis matrices
\(\varvec{\Phi }_{f,g}\) and their Hermitian conjugate
\(\varvec{\Phi }^\dagger _{f,g}\) are applied onto the matrices of Eqs. (
15) and (
16):
$$\begin{aligned}{} & {} i\hbar \frac{\partial }{\partial t}f_{\chi ,k} = (\varvec{\Phi }_g^\dagger {\varvec{D}}_f \varvec{\Phi }_f)g_{\chi ,k}-(\varvec{\Phi }_f^\dagger \varvec{V}_f \varvec{\Phi }_f)f_{\chi ,k}. \end{aligned}$$
(31)
$$\begin{aligned}{} & {} i\hbar \frac{\partial }{\partial t}g_{\chi ,k} = (\varvec{\Phi }_f^\dagger {\varvec{D}}_g \varvec{\Phi }_g)f_{\chi ,k} -(\varvec{\Phi }_g^\dagger \varvec{V}_g \varvec{\Phi }_g)g_{\chi ,k}. \end{aligned}$$
(32)
In these equations
\(\varvec{\Phi }_{f,g}\) denote the basis of the plane waves,
\({\varvec{D}}_{f,g}\) include the kinetic energy operator and
\({\varvec{V}}_{f,g}\) depend on the Hartree–Fock potential for each grid
\(f\) and
\(g\), respectively. The resulting transformation matrices are
\(\varvec{F}=\varvec{\Phi }_g^\dagger {\varvec{D}}_f \varvec{\Phi }_f\) and
\({\varvec{G}}=\varvec{\Phi }_f^\dagger {\varvec{D}}_g \varvec{\Phi }_g\) such as
\(\varvec{V_F}=\varvec{\Phi }_f^\dagger {\varvec{V}}_f \varvec{\Phi }_f\) and
\(\varvec{V_G}=\varvec{\Phi }_g^\dagger {\varvec{V}}_g \varvec{\Phi }_g\). This scheme is related to a formalism presented by Mains and Haddad [
8] which is based on a finite difference discretization of the discrete Schrodinger equation to determine the Wigner functions. Thus, as an alternative, a continuum Hamiltonian can also be discretized using finite differences to produce a Tight-Binding Hamiltonian [
12]. The basic idea is that by using two discrete Wigner functions,
\(f(x_j,k_m)\) and
\(g(x_{j}^\textrm{mid}, k_m^g)\), which are defined on meshpoints and midway between meshpoints, the inconsistencies between the density operator and the Wigner function as a result of the transformation are resolved. These inconsistencies can be described as a loss of information in the density matrix occurring during the Wigner–Weyl transformation and as an incomplete translation from the conventional Wigner function’s domain into the density matrix domain [
7]. According to the statistical ensemble of the density matrix all the information can be contained the way presented here and by Mains and Haddad, what should lead to a more accurate description of the quantum system. However, in contrast to the method used in [
8], here the same basis with respect to
\(k\) is applied for both grids and distribution functions (
31) and (
32), respectively. The wave vector
\(k\) with the discretization width
\(\Delta k\) take the same values for both functions. Another difference to Haddad’s method is that a CAP is used as a supplementary boundary condition in
\(\xi\)-direction; whereas, Haddad’s method uses artificial scattering terms, which is associated with challenges like stability issues when solving time-dependent problems as described in [
8]. The occurrence of negative charge carrier densities is the result, which could be successfully prevented in this work.
3.1 Boundary conditions
First, boundary conditions are specified in the
\(\chi\)-domain with regard to the inflow and outflow concept [
7]. The approach taken here consists of the implementation of the inflow boundary conditions splitting the wave vector (
30) into a positive (incoming waves) and a negative part (outgoing waves). Hence, the distribution functions
\(f\) and
\(g\) can be divided according to their propagation properties. The first
\(N_{L_k}/2\) values of
\(k\) of
\(f\) and
\(g\) are correlated with the forward components
\(f^+\),
\(g^+\) and the last
\(N_{L_k}/2\) values with the backward components
\(f^-\),
\(g^-\), so we have
$$\begin{aligned} f(\chi )= \left( \begin{array}{c} f^+(\chi ) \\ f^-(\chi )\end{array}\right) \,,\,g(\chi )= \left( \begin{array}{c} g^+(\chi ) \\ g^-(\chi )\end{array}\right) \end{aligned}$$
(33)
Now the following auxiliary equations are set up for the f-grid and g-grid at the left and right contact:
$$\begin{aligned}{}[IL]\,f^-(\chi _1)=f^L_{BC}\,,\, [IR]\,f^+(\chi _{N_{\chi _f}})=f^R_{BC}. \end{aligned}$$
(34a)
$$\begin{aligned} \,g^-(\chi _{1.5})=g^L_{BC}\,,\, [IR]\,g^+(\chi _{N_{\chi _g}})=g^R_{BC}. \end{aligned}$$
(34b)
The auxiliary Eq. (
34) are now embedded as Dirichlet boundary conditions within the system matrix. By introducing and embedding the unity matrices
\([IL]\) and
\([IR]\) of dimension [
\(N_{L_k} \times N_{L_k}\)] the system matrix then results as indicated in Fig.
3. The block matrix
\([IL]\) is composed of
\([IL] = [[Id], [0]]\) at the left edge for the electron transport. The unit matrix
\([Id]\) and the zero valued matrix
\([0]\) form a dimension of [
\(N_{L_k}/2 \times N_{L_k}/2\)] each. Analogously, the block matrix [IR] indicates the electron transport at the right edge
\([IR]=[[0], [Id]]\). The inflow of electrons is determined by the Fermi–Dirac statistics at the contacts. For one-dimensional transport problems, the two-dimensional Fermi gas
\(f^{FD}_{2D}(E_c(k)-\mu _c)\) is defined according to
$$\begin{aligned} f^{FD}_{2D}(E_c(k)-\mu _c) = \frac{mk_BT}{2\pi \hbar ^2}\cdot \log \bigg (1+\exp {\bigg (\frac{E_c(k)-\mu _c}{k_BT}}\bigg )\bigg ) \end{aligned}$$
(35)
with energy
\(E_c(k)\) in the system at temperature
\(T\) and the chemical potential
\(\mu _c\). The Boltzmann constant
\(k_B\), the effective mass
\(m\) and the reduced Planck constant
\(\hbar\) are part of the equation as well. Furthermore, the boundary conditions are supplemented by a CAP in
\(\xi\)-direction to ensure that outgoing wave functions decay exponentially, preventing reflections at the boundaries of the
\(\xi\)-domain. The basic concept behind a CAP is to add a complex-valued potential to the system’s Hamiltonian before the basis transformation is carried out. Consequently, the matrices related to the Hartree potential are extended by the CAP:
$$\begin{aligned}{} & {} V_f(\chi ,\xi ) = V_f(\chi ,\xi )-i{\mathcal {C}}_f(\xi ). \end{aligned}$$
(36a)
$$\begin{aligned}{} & {} V_g(\chi ,\xi ) = V_g(\chi ,\xi )-i{\mathcal {C}}_g(\xi ). \end{aligned}$$
(36b)
This complex potential causes the statistical density matrix within the layers to decay, reducing reflections at the computational domain’s edges considerably. There are many different types of CAPs that have been developed over the years, each with its own strengths and weaknesses [
13]. Choosing the right CAP for a particular application will require careful consideration [
14]. The complex absorbing potential can be efficiently constructed within the proposed formalism by a
\(\xi\)-dependent monomial basis of the form:
$$\begin{aligned} {\mathcal {C}}(\xi ):= {\left\{ \begin{array}{ll} \beta \cdot (|\xi |-(\frac{L_\xi }{2}-\epsilon ))^n,&{} \text {if } \frac{L_\xi }{2}-\epsilon< |\xi |<\frac{L_\xi }{2}\\ 0, &{} \text {elsewhere} \end{array}\right. } \end{aligned}$$
(37)
The width of the area containing the CAP is given by
\(\epsilon\) and the coefficient
\(\beta\) factorizes the amplitude of the complex potential. The exponent
\(n\) is the third adjustable parameter and defines the monomial order of the CAP.
Finally, the global system matrix can be set up. Putting the block matrices together and adding the boundary conditions, the structure is given by
$$\begin{aligned} \begin{pmatrix} f_{BC}^L \\ g_{BC}^L \\ 0\\ 0 \\ \vdots \\ 0 \\ 0 \\ g_{BC}^R \\ f_{BC}^R \end{pmatrix} = \begin{pmatrix} [IL] &{} [IL] \\ [F] &{} [V_F] &{} [F]\\ &{} [G] &{} [V_G] &{} [G] \\ &{} \ddots &{} \ddots &{} \ddots \\ &{} &{} [F] &{} [V_F] &{} [F]\\ &{} &{} &{} [IR] &{} [IR]\\ \end{pmatrix} \begin{pmatrix} f(\chi _1) \\ g(\chi _{1.5}) \\ f(\chi _2) \\ g(\chi _{2.5}) \\ \vdots \\ g(\chi _{N_{\chi g}})\\ f(\chi _{N_{\chi f}}) \end{pmatrix} \end{aligned}$$
(38)
The system matrix has a quadratic dimension and adding the inflow boundary conditions yields an unambiguous solution. This completes the formulation of the transport equations and the density matrix from the Von Neumann equation is expressed in terms of the staggered phase space distribution functions f and g. The next section contains the evaluation and the validation of this method by applying it on a resonant tunneling diode (RTD). The RTD is a simple quantum device often used as a standard model to evaluate methods for electron transport, both steady-state and transient. Here, it is used to validate the derived model.