Abstract

A two-grid finite element method is proposed to solve the Schrödinger–Poisson system on a Lipschitz polyhedral domain. To reduce the computational complexity, we solve the Poisson equation of electric potential on a fine mesh |${\cal T}_{h}$| and the Schrödinger eigenvalue problem on a coarse mesh |${\cal T}_{H}$|⁠. The electron density |$n$| is defined by an infinite series whose terms depend on the eigenvalues and eigenfunctions of the Schrödinger equation. The discrete electron density |$n_{H}$| is defined by truncating the series into the sum of |$L_{H}$| terms, where |$H$| is the mesh size of |${\cal T}_{H}$| and |$L_{H}=O(|\!\ln H|^{3/2})$|⁠. Optimal |$H^{1}$|-error estimate is obtained for the discrete electric potential |$V_{h}$|⁠, provided that the mesh sizes satisfy |$h=O(H^{2})$|⁠. To solve the discrete problem, we propose two Newton-type methods: the Newton–Raphson method and an inexact Newton method, by seeking an explicit representation of the Fréchet derivative |$n^{\prime}_{H}$|⁠. We prove that the Newton–Raphson has second-order convergence. The inexact Newton method reduces the computational quantities from |$O\big (N_{h}^{3/2}|\!\ln H|^{3/2}\big )$| of the Newton–Raphson method to |$O(N_{h}|\!\ln H|^{3})$|⁠. Numerical experiments show that the inexact Newton method is very efficient and superior to fixed-point iterations.

1. Introduction

The Schrödinger–Poisson (SP) system describes the thermodynamical and electrostatic equilibrium of electrons confined in narrow quantum wells and incorporates all relevant quantum phenomena. It is widely used in the simulation of quantum devices Ando et al. (1982); Vinter (1984); Beaumont & Sotomayor-Torres (1990). Much attention has been paid to the study of one- and two-dimensional problems Stern (1972); Ohkura (1990); Wu et al. (1996); Spinelli et al. (1998). To avoid numerical challenges of solving coupled quantum systems, physicists and engineers came up with approximate models instead of solving the microscopic Schrödinger model in Ancona & Tiersten (1987); Ancona & Iafrate (1989); Ben Abdallah & Unterreiter (1998); Pinnau (2002); de Falco et al. (2005); Mu & Zheng (2023). However, for extremely small devices like nanowires, numerical simulations based on quantum models become more and more preferable to fully account for quantum effects in such systems in Laux & Frank (1986); Landauer (1988).

The coupled SP system bears to be a popular model for semiconductor devices in the quantum regime. Based on the assumption that the bounded domain |$\varOmega \in \mathbb{R}^{d}$|  |$(d\leq 3)$| is of |${\cal C}^{2}$|-class or a convex polyhedron, the existence and uniqueness of the solution to the SP model are established in Nier (1993); Kaiser & Rehberg (1997) by using traces of differential operators. In the numerical aspect, Ben Abdallah, Cáceres, Carrillo and Vecil proposed a Newton–Raphson method for solving the SP problem based on the facts that their Schrödinger equation is one-dimensional (1D) and all eigenvalues are single in Ben Abdallah et al. (2009). Their numerical experiments show that the Newton–Raphson iteration is much superior to the decoupled Gummal iterative method. In Zhang et al. (2014), Zhang, Cao and Luo proposed a multi-scale asymptotic method for the stationary SP model with rapidly oscillating discontinuous coefficients. The finite element approximation to the homogenized model is studied. They proved the convergence of the numerical solution as the mesh size tends to zero. However, the convergence rate of numerical solutions is not obtained.

The objectives of this work are twofold. The first one is to propose a two-grid finite element method for solving the SP model on a bounded Lipschitz polyhedral domain |$\varOmega \subset{\mathbb{R}}^{3}$|⁠. Unlike Nier (1993); Kaiser & Rehberg (1997) where the solution is assumed to be |$H^{2}$|-regular, the domain here is not necessarily convex and the solution of the Poisson equation has a lower regularity, that is, |$V\in H^{1+s}(\varOmega )$| for some |$1/2<s\le 1$|⁠. The two-grid method is to solve the nonlinear Poisson equation on a fine mesh |${\cal T}_{h}$| of |$\varOmega $| and solve the Schrödinger eigenvalue problem on a coarse mesh |${\cal T}_{H}$|⁠. Provided that |$H\gg h$|⁠, the computational complexity for solving the eigenvalue problem is greatly reduced. Another issue concerns the electron-density operator |$n[V]$| in the Poisson equation. It is defined by an infinite series whose terms depend on eigenvalues and eigenfunctions of the Schrödinger equation. In practice, we must truncate the series into the sum of finitely many terms. To further reduce the computational complexity, we define the discrete electron-density operator |$n_{H}$| by the sum of |$L_{H}$| terms, where |$H$| is the mesh size of |${\cal T}_{H}$| and |$L_{H}=O(|\!\ln H|^{3/2})$|⁠. Therefore, only |$L_{H}$| eigenvalues and eigenfunctions of the discrete Schrödinger equation are computed on the coarse mesh.

A priori error estimate is obtained for linear Lagrangian finite element approximations to both the Poisson equation on |${\cal T}_{h}$| and the Schrödinger equation on |${\cal T}_{H}$|⁠. Provided that the mesh sizes satisfy |$h=O(H^{2})$|⁠, the |$H^{1}$|-error between the discrete electric potential |$V_{h}$| and the exact electric potential |$V$| is bounded by

where the constant |$C$| depends only on |$\left \|{V}\right \|_{L^{2}(\varOmega )}$| and given information of the system. If |$V\in H^{2}(\varOmega )$|⁠, optimal error estimate is obtained in the |$H^{1}$|-norm. Here for the SP system, an interesting observation is that |$\left \|{V}\right \|_{H^{1+s}(\varOmega )}$| is bounded by a constant which depends only on |$\left \|{V}\right \|_{L^{2}(\varOmega )}$|⁠.

The second objective is to propose an efficient method for solving the discrete nonlinear SP system. In Ben Abdallah et al. (2009), the Newton–Raphson method for the SP system is derived upon the feature that all eigenvalues of the 1D Schrödinger equation are single. The argument does not apply to our case since the Schrödinger equation is three-dimensional and multiple eigenvalues are unavoidable. Fortunately, we can follow Kaiser and Rehberg in Kaiser & Rehberg (1997) to derive an explicit expression of the Fréchet derivative |$n_{H}^{\prime}$| of the discrete electron-density operator by using a trace formula. This enables us to design the Newton–Raphson method for the discrete SP problem. Moreover, we prove that the convergence of the Newton–Raphson method is of second-order.

The Newton–Raphson method requires to solve a linearized problem of the nonlinear Poisson equation on |${\cal T}_{h}$|⁠. Although computing |$n_{H}$| only needs |$L_{H}$| eigenvalues and eigenfunctions of the Schrödinger equation, its Fréchet derivative |$n^{\prime}_{H}$| depends on all |$N_{H}$| eigenvalues and eigenfunctions. Building the stiffness matrix corresponding to |$n^{\prime}_{H}$| needs |$O\big (N_{h}N_{H}|\!\ln H|^{3/2}\big )=O\big (N_{h}^{3/2}|\!\ln H|^{3/2}\big )$| computations, provided that |$h=O(H^{2})$|⁠, where |$N_{h}$| and |$N_{H}$| are the numbers of DOFs (degrees of freedom) on |${\cal T}_{h}$| and |${\cal T}_{H}$|⁠, respectively. To reduce the computational complexity further, we propose an inexact Newton method by replacing |$n^{\prime}_{H}$| with an approximate derivative |$\tilde n_{H}^{\prime}$| which depends only on |$L_{H}$| eigenvalues and eigenfunctions. Treating the term concerning |$\tilde n_{H}^{\prime}$| only needs |$O(N_{h}|\!\ln H|^{3})$| operations. Numerical experiments show that the inexact Newton method is very efficient and superior to fixed-point iterations.

1.1 Outline

The paper is organized as follows. In Section 2, we introduce the SP model and present some preliminary results. In Section 3, we present a two-grid finite element discrete for the SP system and prove the well-posedness of the discrete problem. In Section 4, we establish a priori error estimate of numerical solution in |$H^{1}$|-norm regarding the regularity of solution. In Section 5, we propose two Newton-type methods for solving the nonlinear discrete SP problem. Convergence rate of the Newton–Raphson method is obtained. In Section 6, we present two numerical experiments to verify the convergence rate of the two-grid finite element method and demonstrate the efficiency of the inexact Newton method.

1.2 Notation

We introduce the notation to be used in this paper. Let |$\varOmega \subset{\mathbb{R}}^{3}$| be a bounded Lipschitz polyhedral domain and |$\partial \varOmega $| its boundary.

  • The space of square integrable functions on |$\varOmega $| is denoted by formula and its inner product is denoted by |$(u,v)$|⁠.

  • |$H^{m}(\varOmega )$| denotes the Sobolev space whose functions have square-integrable partial derivatives up to order |$m\ge 0$|⁠, and |$H^{1}_{0}(\varOmega )= \{v\in H^{1}(\varOmega ): v|_{\partial \varOmega }=0\}$|⁠.

  • |$f$| denotes the Boltzmann or Fermi–Dirac distribution upon a given Fermi level |$\varepsilon _{F}$| (cf. Nier (1993))
    (1.1)
  • |$V_{0}\in L^{2}(\varOmega )$| denotes an external electric potential.

  • |$u\in L^{2}(\varOmega )$| denotes an arbitrary, but fixed function and |$\varTheta _{u}:=\left \|{u+V_{0}}\right \|_{L^{2}(\varOmega )}$|⁠.

  • |$s\in (1/2,1]$| is the regularity constant such that, for any formula, the solution of the elliptic boundary value problem satisfies |$z\in H^{1+s}(\varOmega )$| (cf. Amrouche et al. (1998) and (Monk, 2003, Theorem 3.18)), where
    (1.2)
  • Suppose |$H$| is a separable Hilbert space and |$\{\varphi _{n}\}^{\infty }_{n=1}$| is an orthonormal basis of |$H$|⁠. For any |${\cal A}\in{\cal L}(H)$|⁠, |$\mathrm{tr}\,{\cal A}$| denotes the trace of |${\cal A}$| and is defined by |$\mathrm{tr}\,{\cal A} =\sum \limits ^\infty _{n=1}(\varphi _{n},{\cal A}\varphi _{n})_{H}$|⁠.

  • An operator |${\cal A}\in{\cal L}(H)$| is of trace-class if and only if |$\mathrm{tr}\big ({\cal A}^{*}{\cal A}\big )^{1/2} <\infty $|⁠, where |${\cal A}^{*}$| denotes the adjoint of |${\cal A}$| with respect to |$(\cdot ,\cdot )_{H}$|⁠. The family of all trace-class operators is denoted by |$\mathscr{B}_{1}$|⁠.

  • Some properties of the trace are listed below (cf. (Reed & Simon, 1975, Section VI.6)).

    • If |${\cal A}\in \mathscr{B}_{1}$|⁠, then |$\mathrm{tr}\,{\cal A}$| is independent of the choose of orthonormal basis.

    • |$\mathrm{tr}$| is a linear mapping from |$\mathscr{B}_{1}$| to |${\mathbb{R}}$|⁠.

    • If |${\cal A}\in \mathscr{B}_{1}$| and |${\cal B}\in{\cal L}(H)$|⁠, then |${\cal A}{\cal B},{\cal B}{\cal A}\in \mathscr{B}_{1}$| and |$\mathrm{tr}({\cal A}{\cal B})=\mathrm{tr}({\cal B}{\cal A})$|⁠.

2. The SP system

The SP model is the coupling of the Poisson equation for the electric potential and the eigenvalue problem of the Hamitonian. After proper nondimensionalization, the Poisson equation has the form

(2.1)

where |$V$| is the electrostatic potential, |$n_{D}\in L^{2}(\varOmega )$| is the doping profile and formula is the electron-density operator defined by

(2.2)

Here |$(\varepsilon _{l},\psi _{l})$|⁠, |$l\ge 1$|⁠, denote the eigenpairs of the Hamiltonian |${\cal H}[u] = -\varDelta +u+V_{0}$|⁠, namely

(2.3)

We write |$\varepsilon _{l}[u]$| to specify the dependency of |$\varepsilon _{l}$| on |$u$| and order them monotonically |$\varepsilon _{1}[u]<\varepsilon _{2}[u]\le \cdots $| by counting multiplicities.

From (Kaiser & Rehberg, 2000, Theorem 5.13), problem (2.1) has a unique solution formula. Moreover, the electron-density operator formula is Fréchet-differentiable and monotone, namely

(2.4)

 

Lemma 2.1.
There exists two positive constants |$C_{0},C_{1}$| depending only on |$\varOmega $| such that
(2.5)

 

Proof.
First we define the bilinear form formula by
(2.6)
The weak formulation of (2.3) is to find formula such that
(2.7)
Using Poincaré’s inequality and the Gagliardo–Nirenberg inequality, we have
(2.8)
This shows the right inequality of (2.5).

Similarly, we have formula for a constant |$\hat{C}>0$|⁠. Poincaré’s inequality implies formula. This yields the left inequality of (2.5).

 

Remark 2.2.
Let |$0<\lambda _{1}\le \lambda _{2}\le \cdots \le \lambda _{l}\le \cdots $| be the eigenvalues of formula. From (Kaiser & Rehberg, 1997, equation (2.1)) we know that |$\lim \limits _{l\to \infty }(l^{-2/3}\lambda _{l}) = C_\varOmega>0$|⁠. By arguments similar to (2.8), we can find two positive constants |$C_{\min }$| and |$C_{\max }$| depending only on |$\varOmega $| such that
(2.9)

 

Theorem 2.3.
Suppose formula is the solution of problem (2.1). There exist two positive constants |$C$| and |$\alpha $| depending only on |$\varOmega $| such that

 

Proof.
Let |$\varepsilon _{l},\psi _{l}$| denote the eigenvalues and eigenfunctions of the Hamiltonian |${\cal H}[V]$|⁠. From Lemma 2.1, it is easy to see that
(2.10)
Using (1.2) and Remark 2.2, we get
for a constant |$\alpha>0$| depending only on |$\varOmega $|⁠. The proof is finished.

3. A two-grid finite element method

Let |${\cal T}_{H}$| be a coarse mesh of |$\varOmega $| and |${\cal T}_{h}$| a fine mesh obtained by refining |${\cal T}_{H}$|⁠. Both of them are quasi-uniform and consist of shape-regular tetrahedra. The mesh sizes of |${\cal T}_{H}$| and |${\cal T}_{h}$| are denoted by |$H$| and |$h$|⁠, respectively. The lowest-order continuous finite element spaces on |${\cal T}_{H}$| and |${\cal T}_{h}$| are defined by

where |$P_{1}(K)$| is the space of linear polynomials on |$K$|⁠. Clearly |$X_{H}\subset X_{h}$|⁠. Write |$N_{H}=\operatorname{dim} (X_{H})$| and |$N_{h}=\operatorname{dim} (X_{h})$| for convenience.

3.1 The discrete problem

To propose the discrete problem, we introduce a truncated distribution function |$f_{H}\in C^\infty ({\mathbb{R}})$| which satisfies |$f_{H}^{\prime}\le 0$| and

(3.1)

Since |$f$| deceases exponentially, we can assume that, with a constant |$C$| independent of |$H$|⁠,

(3.2)

The two-grid finite element method for solving (2.1) is to seek |$V_{h}\in X_{h}$| such that

(3.3)

where formula is the discrete electron-density operator defined as follows:

(3.4)

Here |$\varepsilon _{l,H}[u]\in{\mathbb{R}}$| and |$\psi _{l,H}\in X_{H}$| solve the discrete eigenvalue problem on the coarse mesh

(3.5)

The eigenvalues are ordered monotonically |$\varepsilon _{1,H}[u]\le \varepsilon _{2,H}[u] \le \cdots \le \varepsilon _{N_{H},H}[u]$| by counting multiplicities. The Galerkin approximation of the eigenvalue problem implies |$\varepsilon _{l}[u] \le \varepsilon _{l,H}[u]$|⁠.

 

Remark 3.1.

By (2.1) and (2.2), the right-hand side of the Poisson equation depends only on eigenvalues and eigenfunctions of the Hamiltonian |${\cal H}[V]$|⁠, not on the gradients of eigenfunctions. This motivates us to solve the eigenvalue problem on the coarse mesh |${\cal T}_{H}$| and the Poisson equation on the fine mesh |${\cal T}_{h}$|⁠, without causing the deterioration of the convergence rate of |$V_{h}$|⁠.

 

Lemma 3.2.
Let |$L=L[u]$| and |$L_{H}=L_{H}[u]$| be the maximal integers satisfying |$f_{H}(\varepsilon _{L}[u])\ne 0$| and |$f_{H}(\varepsilon _{L_{H},H}[u])\ne 0$|⁠, respectively. When |$H$| is small enough, there is a constant |$C$| independent of |$H$| and |$u$| such that
(3.6)

 

Proof.
Since |$\varepsilon _{L}[u]\le \varepsilon _{L,H}[u]$|⁠, from (3.1) we know that |$L_{H}\le L$| and
(3.7)
Moreover, (2.9) implies |$\lim \limits _{H\to 0}L_{H} =+\infty $|⁠. For |$H$| small enough, we have
The proof is finished.

3.2 Frechét derivative of |$n_{H}[u]$|

Following Kaiser and Rehberg Kaiser & Rehberg (1997), we define the iterated primitive functions of |$f_{H}$| by

(3.8)

Since |$f_{H}(t)=0$| for |$t\ge \varepsilon _{F}+2\mu ^{-1}|\!\ln H|+1$|⁠, we also have |$f_{H}^{(-j)}(t)=0$| for |$t\ge \varepsilon _{F}+2\mu ^{-1}|\!\ln H|+1$| for |$j=1,2$|⁠. For simplicity, we make the following assumptions which are rather mild for many practical distribution functions.

  • (A1) |$f_{H}$| is the trace of some holomorphic function |$f_{H}(z)$| in a bounded domain |${\mathbb{G}}\subset{\mathbb{C}}$| which covers the interval |$[\varepsilon _{1,h},\varepsilon _{N_{h},h}]$| and is symmetric about the real axis of |${\mathbb{C}}$|⁠.

  • (A2) There is a bounded contour |$\varGamma \subset{\mathbb{G}}$| and a constant |$\delta>0$| independent of |$H$| such that |$\varGamma $| encloses |$[\varepsilon _{1,h},\varepsilon _{N_{h},h}]$| and satisfies
    (3.9)
    Moreover, |$\varGamma $| is oriented such that the interval |$[\varepsilon _{1,H}, \varepsilon _{N_{H},H}]$| lies on its left-hand side.
  • (A3) |$f_{H}(z)$| has primitive functions |$f_{H}^{(-1)}(z)$| and |$f_{H}^{(-2)}(z)$| which are holomorphic in |${\mathbb{G}}$| and have real values on |${\mathbb{R}}$|⁠. There exists a constant |$C>0$| independent of |$H$| such that
    (3.10)

 

Remark 3.3.

The assumptions are mild in practice. Below we give some remarks on them.

  • Since the truncated distribution function |$f_{H}\in C^\infty ({\mathbb{R}})$|⁠, we can define |$f_{H}(z)$| by the analytical continuation of |$f_{H}(t), t\in [\varepsilon _{1,h},\varepsilon _{N_{h},h}]$|⁠, to a bounded domain |${\mathbb{G}}\subset{\mathbb{C}}$|⁠.

  • For any formula, we shall represent |$n_{H}[u]$| by the integral of |$f_{H}(z)$| on some contour |$\varGamma \subset{\mathbb{G}}$|⁠. This motivates us to assume |$\varGamma $| encloses the spectrum interval |$[\varepsilon _{1,h},\varepsilon _{N_{h},h}]$|⁠.

  • The estimates of |$n_{H}$| depend on |$\delta ^{-1}$|⁠. When |$\delta $| increases, |$\varGamma $| leaves away from |$[\varepsilon _{1,h},\varepsilon _{N_{h},h}]$| and |$f_{H}(z)$| is required to be analytical in a larger domain |${\mathbb{G}}$|⁠.

  • Since |$f_{H}$| is analytical in |${\mathbb{G}}$|⁠, it has primitive functions |$f_{H}^{(-1)}$| and |$f_{H}^{(-2)}$| in |${\mathbb{G}}$|⁠.

 

Lemma 3.4.
The discrete density operator |$n_{H}$| is Fréchet-differentiable at any |$u\in L^{2}(\varOmega )$| and the Fréchet derivative formula. Moreover, for any formula, we have
(3.11)
In case of |$\varepsilon _{i,H}=\varepsilon _{j,H}$|⁠, the coefficient |$\big [f_{H}(\varepsilon _{i,H})-f_{H}(\varepsilon _{j,H})\big ]/(\varepsilon _{i,H}-\varepsilon _{j,H})$| is replaced with |$f_{H}^{\prime}(\varepsilon _{i,H})$|⁠.

 

Proof.
First we define the discrete Schrödinger operator |${\cal H}_{H}[u]:X_{H}\to X_{H}$| as follows: for any |$\xi _{H}\in X_{H}$|⁠, |${\cal H}_{H}[u](\xi _{H})\in X_{H}$| is the unique solution to the discrete problem
(3.12)
Clearly the eigenpairs |$(\varepsilon _{l,H}, \psi _{l,H})$|⁠, |$1\le l\le N_{H}$|⁠, are also eigenpairs of |${\cal H}_{H}[u]$|⁠.
For any formula, let formula be the multiplying operator defined by
(3.13)
Since |$X_{h}\subset C(\overline \varOmega )$|⁠, the right-hand side is well-defined for any |$\phi _{H}\in X_{H}$|⁠. In fact, |${\cal M}_{H}[\eta ](v)$| is the formula-projection of |$\eta v$| onto |$X_{H}$|⁠.
For any |$z\in \varGamma $|⁠, we denote the resolvent of |${\cal H}_{H}[u]$| by |${\cal R}_{z}[u]=(z-{\cal H}_{H}[u])^{-1}$|⁠. For |$k=0,1,2$|⁠, we define the linear operators |${\cal U}_{H,k}:X_{H}\to X_{H}$| by
By the Residue theorem and the definition of trace in section 1.2, it is easy to see that
(3.14)
First we write |${\cal N}_{H}[u](\xi ):= \sum \limits ^{N_{H}}_{i,j=1}\frac{f_{H}(\varepsilon _{i,H})-f_{H}(\varepsilon _{j,H})}{\varepsilon _{i,H}-\varepsilon _{j,H}} \big (\xi ,\psi _{i,H}\psi _{j,H}\big ) \psi _{i,H}\psi _{j,H}$|⁠. Since formula, we have formula, and by arguments similar to (3.14), find that
It is left to show |$n^{\prime}_{H}[u] ={\cal N}_{H}[u]$|⁠.
In view of (3.14), it is easy to see that
(3.15)
Since |$\mathrm{dist}(\varepsilon _{l,H}[u+\xi ],\varGamma )\geq \delta $| by Assumption (A2), we choose formula small enough such that |$\mathrm{dist}(\varepsilon _{l,H}[u+\xi ],\varGamma )\geq \delta /2$| for |$1\le l\le N_{H}$|⁠. It follows from (3.9) that
(3.16)
Since norms are equivalent on |$X_{H}$|⁠, there is a generic constant |$C_{H}$|⁠, depending on |$H$| such that
It is easy to see that |${\cal D}_{z}[\xi ]:={\cal R}_{z}[u+\xi ]-{\cal R}_{z}[u]$| satisfies
By the identity |${\cal R}_{z}[u+\xi ]-{\cal R}_{z}[u]={\cal R}_{z}[u+\xi ]{\cal M}_{H}[\xi ]{\cal R}_{z}[u]$| and proper arrangements, we get
Inserting the above estimate into the right-hand side of (3.15), we end up with
This shows
Then |$n_{H}$| is Fréchet-differentiable at formula and |$n^{\prime}_{H}[u]={\cal N}[u]$|⁠.

 

Corollary 3.5.

formula for all formula.

 

Proof.

Since |$f_{H}$| is non-increasing, the conclusion comes directly from Lemma 3.4.

 

Remark 3.6.
By arguments similar to the proof of Lemma 3.4 and using the primitive function |$f_{H}^{(-2)}$|⁠, we can derive another form of |$n_{H}^{\prime}[u]$|⁠, given by

 

Lemma 3.7.
Suppose |$u,w,\xi ,\eta \in L^{2}(\varOmega )$|⁠. Then

 

Proof.
For any |$z\in \varGamma $| and |$k=1,2,3$|⁠, it is easy to see that
(3.17)
Then Remark 3.6 implies
The proof is finished by inserting (3.17) into the above equality.

3.3 Well-posedness of the discrete problem

We end this section by proving the well-posedness of problem (3.3).

 

Theorem 3.8.
The discrete problem (3.3) has a unique solution. Moreover, there is a constant |$C>0$| depending only on |$\varOmega $| such that, with |$\varTheta _{0}=\left \|{V_{0}}\right \|_{L^{2}(\varOmega )}$|⁠,
(3.18)

 

Proof.
Corollary 3.5 implies
(3.19)
Classical theories for monotone operators show that problem (3.3) has a unique solution (cf. (Zeider, 1990, Section 25.4, pp. 503–506)).
Next we consider the discrete eigenvalue problem: find |$(\lambda _{l,H},\phi _{l,H})\in{\mathbb{R}}\times X_{H}$| which satisfy
Taking |$v_{H}=\phi _{l,H}$| and using Poincaré’s inequality, we can find a constant |$C$| depending only on |$\varOmega $| such that the estimates hold
There exists a generic constant |$C$| depending only on |$\varOmega $| such that
Since |$\lambda _{l,H}\ge \lambda _{l}$| and |$f$| decays exponentially, we deduce from (2.9) that
Finally, taking |$v_{h}=V_{h}$| in (3.5), we find that
Using (3.19) and Poincaré’s inequality, we obtain
The proof is finished.

4. Finite element error estimates

The purpose of this section is to establish the error estimate between the exact potential |$V$| and the discrete potential |$V_{h}$|⁠. First we estimate the error between |$n[u]$| and |$n_{H}[u]$|⁠. From Lemma 3.2, we have |$L_{H}=L_{H}[u]\le L=L[u]\ll N_{H}$|⁠.

4.1 Solution operators

By the Gagliardo–Nirenberg inequality, Poincaré’s inequality and Young’s inequality, there exists a constant |$C_\varOmega $| depending only on |$\varOmega $| such that, for formula,

Define |$\rho _{u}:=C_\varOmega \varTheta _{u}^{4}$|⁠. Clearly |$a(\rho _{u}+u;\boldsymbol{\cdot },\boldsymbol{\cdot })$| is continuous and coercive on formula, namely

(4.1)

where |$C_{2},C_{3}$| are two positive constants depending only on |$\varOmega $|⁠.

For any |$g\in H^{-1}(\varOmega )$|⁠, the elliptic problem

(4.2)

has a unique solution formula which satisfies formula. This defines a solution operator formula, satisfying |${\cal S}[\rho _{u}+u](g)=w$|⁠. In fact, |${\cal S}[\rho _{u}+u]$| is the inverse of formula and satisfies, upon a constant |$C$| depending only on |$\varOmega $|⁠,

(4.3)

Similarly, we define a discrete solution operator |${\cal S}_{H}[\rho _{u}+u]: H^{-1}(\varOmega ) \to X_{H}$| by the Galerkin approximation of problem (4.2), namely

(4.4)

Without specifications, we write |${\cal S}={\cal S}[u+\rho _{u}]$| and |${\cal S}_{H}={\cal S}_{H}[u+\rho _{u}]$| throughout this section.

 

Lemma 4.1.
Suppose formula and let |$s\in (1/2,1]$| be the regularity constant in (1.2). Then the solution of (4.2) admits formula. Moreover, there exists a constant |$C>0$| depending only on |$\varOmega $| such that
(4.5)

 

Proof.
Set |$p:=6/(3-2s)>3$| and |$q:= p/(p-1)<3/2$|⁠. For |${\boldsymbol{z}}\in{\boldsymbol{L}}^{q}(\varOmega )$|⁠, the elliptic problem
has a unique solution that satisfies
(4.6)
The solution of (4.2) satisfies
By (4.6) and the injection |$W^{1,q}(\varOmega )\hookrightarrow L^{3}(\varOmega )$|⁠, we have
In view of (4.3) and Poincaré’s inequality, it follows that
(4.7)
This yields formula. The error estimate (4.5) comes directly from classical finite element error estimates.

Let formula denote the Banach algebra of linear and bounded operators on formula, endowed with the operator norm. Suppose |${\cal Q}\in \mathscr{B}$| and only has point spectrum |$\{\nu _{l}\}_{l=1}^\infty $|⁠. For any |$p\in [1,+\infty )$|⁠, we define the Schatten |$p$|-norm of |${\cal Q}$| by

An operator |${\cal Q}$| is called |$p$|-summable if |$\left \|{{\cal Q}}\right \|_{p}<\infty $|⁠. The closed ideal of |$p$|-summable operators on formula is denoted by |$(\mathscr{B}_{p},\left \|{\boldsymbol{\cdot }}\right \|_{p})$|⁠.

Similar to (3.13), for |$\eta \in L^{q}(\varOmega )$| with |$q\ge 2$|⁠, we define a multiplying operator |${\cal M}[\eta ]$| as follows:

(4.8)

It is clear that |${\cal M}[\eta ]$| maps |$L^{\frac{qr}{q-r}}(\varOmega )$| to |$L^{r}(\varOmega )$| for |$1\le r\le q$| and maps formula to |$H^{-1}(\varOmega )$| if |$q\ge 3$|⁠. In view of (3.13), we find that |${\cal M}_{H}[\eta ](v)$| is the |$L^{2}$|-projection of |${\cal M}[\eta ](v)$| onto the finite element space |$X_{H}$|⁠.

 

Lemma 4.2.
Suppose formula and |$\eta \in L^{6}(\varOmega )$|⁠. There exists a constant |$C$| depending only on |$\varOmega $| and formula such that

 

Proof.
In view of (4.8), the multiplying operator |${\cal M}[\eta ]$| actually maps formula to |$L^{3/2}(\varOmega )$| and is bounded by |$\left \|{\eta }\right \|_{L^{6}(\varOmega )}$|⁠. Then (4.3) and the injection of |$L^{3/2}(\varOmega )\hookrightarrow H^{-1}(\varOmega )$| show that
Moreover, inequality (4.7) shows that formula. Then we have
The results for |${\cal S}_{H}$| can be proved similarly.

4.2 Estimate of |$n[u]-n_{H}[u]$|

First we split |$n[u]-n_{H}[u]$| into two terms

The first term is easily estimated with (3.1) and the Cauchy–Schwarz inequality. In fact,

From Lemma 2.1, equation (1.1) and Remark 2.2, there exists a constant |$\alpha>0$| such that

(4.9)

It is left to estimate formula. The key ingredient is to use trace representations of |$\hat{n}[u]$| and |$n_{H}[u]$|⁠. Note that |$\mu _{l}= (\varepsilon _{l}+\rho _{u})^{-1}$| and |$\mu _{l,H}= (\varepsilon _{l,H}+\rho _{u})^{-1}$| are eigenvalues of |${\cal S}$| and |${\cal S}_{H}$|⁠, respectively, and |$\psi _{l},\psi _{l,H}$| are the eigenfunctions belonging to |$\mu _{l}$| and |$\mu _{l,H}$|⁠, respectively. Then

(4.10)

Clearly |$\mu _{1},\cdots ,\mu _{L}$| and |$\mu _{1,H},\cdots ,\mu _{L_{H},H}$| are included in the interval |$[0,(\varepsilon _{1}+\rho _{u})^{-1}]$|⁠. By (3.1), we can extend |$\hat{f}_{H}$| by zero to |$(-\infty ,0]$| and the extension |$\hat{f}_{H}\in C^\infty ({\mathbb{R}})$|⁠. For |$k\ge 0$|⁠, the function |$\hat{f}_{k}(t):=t^{-k}\hat{f}_{H}(t)$| is smooth with the extended value |$\hat{f}_{k}(0):=0$| and vanishes on |$(-\infty ,\gamma _{H}]$|⁠, where |$\gamma _{H}:=(2\mu ^{-1}|\!\ln H|+\varepsilon _{F}+1+\rho _{u})^{-1}$|⁠.

Similar to section 3.2, we make two assumptions.

  • (A4) For any integer |$k\ge 0$|⁠, |$\hat{f}_{k}$| is the trace of some holomorphic function |$\hat{f}_{k}(z)$| in |${\mathbb{G}}$|⁠.

  • (A5) There is a bounded contour |$\hat \varGamma \subset{\mathbb{G}}$| which incloses |$[0,(\varepsilon _{1}+\rho _{u})^{-1}]$| and satisfies
    (4.11)
    Moreover, |$\hat \varGamma $| is oriented such that the interval |$\big [0,(\varepsilon _{1}+\rho _{u})^{-1}\big ]$| lies on its left-hand side.

Since |$\hat{f}_{k}\in C^\infty ({\mathbb{R}})$|⁠, assumptions (A4) and (A5) are rather mild by similar analyses given in Remark 3.3. We do not elaborate on the discussions again here.

 

Lemma 4.3.
For formula and integer |$k\ge 1$|⁠, there hold
(4.12)
 
(4.13)

 

Proof.
The proof is easy by using the Residue theorem. Since |$\hat{f}_{k}\equiv 0$| in |$(-\infty ,\gamma _{H})$|⁠, we have
The proof of (4.13) is similar. Since |$X_{H}$| is a subspace of formula, we have the orthogonal decomposition formula with respect to the inner product of formula. Note that |$\{\psi _{l,H}\}_{l=1}^{N_{H}}$| provides an orthonormal basis of |$X_{H}$|⁠. We choose an orthonormal basis of |$X^\perp _{H}$| by |$\big \{\psi ^\perp _{l}\big \}^{\infty }_{l=N_{H}+1}$|⁠. In view of (4.4), the operator |${\cal S}_{H} $| maps formula to |$X_{H}$|⁠, and maps |$X^\perp _{H}$| to |$\{0\}$|⁠. Using the Residue theorem and the fact that |$\hat{f}_{k}\equiv 0$| in |$(-\infty ,\gamma _{H})$|⁠, we have
The proof is finished.

 

Lemma 4.4.
There exits a constant |$C$| depending only on |$\varOmega $| and |$\delta $| such that

 

Proof.
For any |$\eta \in L^{6}(\varOmega )$|⁠, using (4.12) and (4.13) with |$k_{1}=k_{2}=2$|⁠, we have
(4.14)
Direct calculations show that
(4.15)
We shall estimate the three terms on the right-hand side.
For |$p>3/2$|⁠, by (2.9) and (4.1), there exists a constant |$C>0$| depending only on |$p$| and |$\varOmega $| such that |$\mu _{l}^{p} = (\varepsilon _{l}+\rho _{u})^{-p} \le C l^{-2p/3}$|⁠. Then
(4.16)
Note that |$\|({z}-{\cal S})^{-1}\|\le \delta ^{-1}$| for any |$z\in \hat \varGamma $|⁠. By (4.5), (4.16) and Lemma 4.2, the first term on the right-hand side of (4.15) can be estimated as follows:
where the constant |$C>0$| depends only on |$\varOmega $| and |$\delta $|⁠. The second term on the right-hand side of (4.15) can be estimated similarly and is also bounded by formula. For the third term, we use the formula |$\mathrm{tr}({\cal A}{\cal B})=\mathrm{tr}({\cal B}{\cal A})$| and obtain
Substituting these estimates first into (4.15) and then into (4.14), we end up with
The proof is finished.

4.3 Estimate of |$V-V_{h}$|

Now we present the main result of this section, namely the a priori error estimate of |$V_{h}$|⁠.

 

Theorem 4.5.
Let |$\alpha $| be the constant in Theorem 2.3 and define
There exist two positive constants |$C$| and |$\beta $| independent of |$H$| and |$h$| such that

 

Proof.
Let |$\pi _{h}$| be the nodal interpolation operator onto |$X_{h}$|⁠. By Theorem 2.3 and standard interpolation error estimates, we obtain
Take |$k_{1}=1,k_{2}=2$| and |$u=V_{h}$| in (4.13). By Theorem 3.8, Lemma 4.2 and arguments similar to the proof of Lemma 4.4, we obtain
where the constant |$C$| depends only on |$\varOmega $| and |$\delta $|⁠. This shows
(4.17)
Write |$e=V-V_{h}$|⁠, |$\xi _{h}=\pi _{h}V-V_{h}$| and |$\eta =V-\pi _{h}V$| for convenience. Using (3.3) and the equality |$(\nabla V, \nabla v_{h}) = (n[V]-n_{D}, v_{h})$| for all |$v_{h}\in X_{h}$|⁠, we have
(4.18)
Since |$(n_{H}[V]-n_{H}[V_{h}], V-V_{h})\le 0$|⁠, using (2.10) and (4.17), we deduce that
(4.19)
Moreover, (4.9) and Lemma 4.4 indicate that
(4.20)
Inserting (4.19)–(4.20) into (4.18) yields
The proof is finished by using Poincaré’s inequality.

5. Newton-type methods

The purpose of this section is to study Newton-type methods for solving the nonlinear discrete problem (3.3). First we derive the Newton–Raphson method and prove that it has second-order convergence rate. Then we propose an inexact Newton method by truncating the sum on the right-hand side of (3.11). The inexact Newton method has a convergence speed comparable to the Newton–Raphson method, but is vastly superior in terms of computational complexity.

5.1 Stability of |$n_{H}^{\prime}[u]$|

To study the convergence rate, we need the stability of |$n^{\prime}_{H}[u]$| with respect to |$H$| for any |$u\in L^{2}(\varOmega )$|⁠. Recall the bounded contour |$\varGamma $| and the constant |$\delta>0$| in Assumption (A2) of section 3.2. For |$z\in \varGamma $|⁠, |${\cal R}_{z}[u](z)=(z-{\cal H}_{H}[u])^{-1}$| denotes the resolvent of |${\cal H}_{H}[u]$|⁠.

 

Lemma 5.1.
Suppose |$\xi ,u,v\in L^{2}(\varOmega )$| and |$\rho _{u}=C_\varOmega \varTheta _{u}^{4}$|⁠, |$\rho _{v}=C_\varOmega \varTheta _{v}^{4}$| are given as in (4.1). There is a constant |$C>0$| depending only on |$\delta $| and |$\varOmega $| such that

 

Proof.
Let |${\cal H}_{H}[u]=\int ^{\infty }_{-\infty }\lambda \, \mathrm{d} E(\lambda )$| be the spectral decomposition of |${\cal H}_{H}[u]$|⁠. Using (3.9), we have
Together with Lemma 4.2, this shows
The proof is finished.

 

Theorem 5.2.
Suppose |$w, \xi ,\eta \in L^{2}(\varOmega )$|⁠. There exists a constant |$C>0$| depending only on |$\delta $| and |$\varOmega $| such that
(5.1)
 
(5.2)

 

Proof.
From Remark 3.6, it is easy to see that
By Lemma 5.1 and assumption (3.10), we have
Similarly, from Lemma 3.7, we have
Again we get (5.2) by using Lemma 5.1. The proof is finished.

5.2 The Newton–Raphson method

Next we define the discrete residual operator |${\cal P}_{h}: X_{h}\to X_{h}^{\prime}$| as follows:

(5.3)

Then (3.3) is equivalent to the nonlinear equation

(5.4)

From Lemma 3.4, |${\cal P}_{h}$| is Frechét-differentiable at |$w_{h}\in X_{h}$| and the Frechét derivative satisfies

(5.5)

Given an approximate solution |$V^{(k)}_{h}$| with |$k\ge 0$|⁠, the Newton–Raphson method for solving (5.4) is to update |$V^{(k+1)}_{h}=V^{(k)}_{h}-r_{k}$|⁠, where the correction |$r_{k}\in X_{h}$| solves the linearized problem

(5.6)

By Corollary 3.5, it is clear that problem (5.8) or (5.6) has a unique solution.

Now we present the Newton–Raphson method for solving (5.4), or (3.3) equivalently.

 

Algorithm 5.3. (Newton–Raphson method).
Suppose |$\epsilon \in (0,1)$| is the relative tolerance and |$N_{\mathrm{itr}}$| is the number of iterations. Set |$k=0$|⁠, |$e_{0} = 1$| and compute the initial guess |${V}_{h}^{(0)}\in X_{h}$| by solving the linear problem
(5.7)
 
  • While |$(e_{k} \ge \epsilon \;\; \& \;\; k< N_{\mathrm{itr}})$|⁠, do

    1. Compute all eigenpairs |$(\varepsilon _{l,H},\psi _{l,H})$|⁠, |$1\le l\le N_{H}$|⁠, by solving (3.5) with |$V_{h}$| replaced by |$V_{h}^{(k)}$|⁠.

    2. Compute the electron density |$n_{H} = \sum \limits _{l=1}^{L_{H}}f_{H}(\varepsilon _{l,H})\psi _{l,H}^{2}$|⁠.

    3. Solve (5.6) for the correction function |$r_{k}\in X_{h}$|⁠, that is,
      where the Fréchet derivative at |$V_{h}^{(k)}$| is given by (3.11) and has the form
    4. Update the approximate solution
      End while.

 

Theorem 5.3.
Let |$\delta>0$| be the constant in (3.9). There is a constant |$C_{\mathrm{NR}}\ge 1$| depending only on |$\varOmega $| and |$\delta $| such that for any |$V^{(1)}_{h}\in X_{h}$| satisfying formula, the iterative sequence generated by the Newton–Raphson method satisfies

 

Proof.
Define |$e_{k}=V^{(k)}_{h} -V_{h}$|⁠. From (5.5), we know that
Using Corollary 3.5 and inequality (5.2), we have
By Poincaré’s inequality, there exists a constant |$C_{\mathrm{NR}}>0$| depending only on |$\delta $| and |$\varOmega $| such that
Since formula, we conclude the result by induction.

Now we analyze the computational complexity of the stiffness matrix associated with the term |$(n^{\prime}_{H}(r_k), v_h)$| in Algorithm 5.3. Define

Here |$b_{1},\cdots ,b_{N_{h}}$| are nodal basis functions of |$X_{h}$|⁠. By (3.11), we find that |${\mathbb{N}} ={\mathbb{K}}^\top{\mathbb{F}}{\mathbb{K}}$|⁠, where |${\mathbb{K}}$| is a |$N_{H}^{2}\times N_{h}$| matrix and |${\mathbb{F}}$| is a |$N_{H}^{2}\times N_{H}^{2}$| diagonal matrix. Their entries are given by

where |$1\le i,j\le N_{H}$| and |$1\le l\le N_{h}$|⁠. Since |$f_{H}(\varepsilon _{i,H})=0$| for all |$i>L_{H}=O(|\!\ln H|^{3/2})$|⁠, |${\mathbb{F}}$| has only |$O(N_{H}|\!\ln H|^{3/2})$| nonzero diagonal entries. Then only |$O(N_{H}|\!\ln H|^{3/2}\times N_{h})$| entries of |${\mathbb{K}}$| are used in building |${\mathbb{N}}$|⁠. For any |$\mathbf{v}\in{\mathbb{R}}^{N_{h}}$|⁠, the computation of |${\mathbb{N}}{\mathbf{v}}$| consists of three successive steps |${\mathbf{v}}_{1}={\mathbb{K}}{\mathbf{v}}$|⁠, |${\mathbf{v}}_{2}={\mathbb{F}}{\mathbf{v}}_{1}$| and |${\mathbb{N}}{\mathbf{v}}={\mathbb{K}}^\top{\mathbf{v}}_{2}$|⁠. The total number of operations is |$O(N_{h}N_{H}|\!\ln H|^{3/2})$| too. Therefore, the computational complexity of dealing with |${\mathbb{N}}$| needs |$O\big (N_{h}^{3/2}|\!\ln h|^{3/2}\big )$| operations, provided that |$h=O(H^{2})$|⁠, and is very time-consuming.

5.3 Inexact Newton method

To further reduce the computational complexity, we truncate the sum |$n_{H}^{\prime}$| in (3.11) and define an approximate operator |$\tilde n_{H}^{\prime}\approx n_{H}^{\prime}$| as follows

 

Algorithm 5.4. (Inexact Newton method).

Suppose |$\epsilon \in (0,1)$| is the relative tolerance and |$N_{\mathrm{itr}}$| is the number of iterations. Set |$k=0$|⁠, |$e_{0} = 1$| and the initial guess |$\tilde{V}_{h}^{(0)}={V}_{h}^{(0)}$| by solving (5.7).

  • While |$(e_{k} \ge \epsilon \;\; \& \;\; k< N_{\mathrm{itr}})$|⁠, do

    1. Compute the eigenpairs |$(\varepsilon _{l,H},\psi _{l,H})$|⁠, |$1\le l\le L_{H}:=L_{H}[\tilde{V}_{h}^{(k)}]$|⁠, by solving the discrete eigenvalue problem (3.5) with |$V_{h}$| replaced by |$\tilde{V}_{h}^{(k)}$|⁠.

    2. Compute the electron density |$\tilde n_{H} = \sum \limits _{l=1}^{L_{H}}f_{H}(\varepsilon _{l,H})\psi _{l,H}^{2}$|⁠.

    3. Solve (5.6) for the correction function |$r_{k}\in X_{h}$|⁠, that is,
      (5.8)
      where the approximate Fréchet derivative at |$\tilde{V}_{h}^{(k)}$| is defined by
    4. Update the approximate solution
      End while.

From (3.6), we know that |$L_{H}:=L_{H}[\tilde V_{h}^{(k)}]\le C|\!\ln H|^{3/2}$|⁠. Usually, an efficient algorithm for computing the eigen-pairs |$(\varepsilon _{i,H},\psi _{i,H})$|⁠, |$1\le i\le L_{H}$|⁠, needs |$O(N_{H}L_{H}^{2})=O(N_{H}|\!\ln H|^{3})$| operations. Similar to |${\mathbb{N}}$|⁠, the stiffness matrix associated with |$\big (\tilde n^{\prime}_{H}[\tilde V_{h}^{(k)}](\tilde r_{k}), v_{h}\big )$| can be written as |$\tilde{\mathbb{N}} = \tilde{\mathbb{K}}^\top \tilde{\mathbb{F}}\tilde{\mathbb{K}}$|⁠, where |$\tilde{\mathbb{F}}$| is a diagonal matrix and |$\tilde{\mathbb{K}}$| is a low-rank matrix. Their nonzero entries are given by

where |$1\le i,j\le L_{H}$| and |$1\le l\le N_{H}$|⁠. The build of |$\tilde{\mathbb{K}}$| and |$\tilde{\mathbb{F}}$| needs |$O(N_{h}|\!\ln H|^{3})$| operations totally, and the computation of |$\tilde{\mathbb{N}}{\mathbf{v}}$| also needs |$O(N_{h}|\!\ln H|^{3})$| operations for any |$\mathbf{v}\in{\mathbb{R}}^{N_{h}}$|⁠. Therefore, the computational complexity of dealing with |$\tilde{\mathbb{N}}$| only needs |$O(N_{h}|\!\ln H|^{3})$| operations. Comparing |$\tilde{\mathbb{N}}$| with matrix |${\mathbb{N}}$| in the Newton–Raphson method, the computational complexity is reduced from |$O(N_{h}^{3/2}|\!\ln h|^{3/2})$| to |$O(N_{h}|\!\ln H|^{3})$|⁠. Tables 2 and 3 in the next section show the efficiencies of both the Newton–Raphson method and the inexact Newton method.

6. Numerical experiments

We report two numerical experiments to demonstrate the competitive behavior of the two-grid finite element method. The first one verifies the convergence rate of the two-grid method. The second one demonstrates the efficiency of the inexact Newton method.

Our code is based on the finite element toolbox PHG developed in the State Key Laboratory of Scientific and Engineering Computing (LSEC), Chinese Academy of Sciences. The numerical experiments are carried out on the super computer LSSC-IV of LSEC. The computational domain is set by |$\varOmega =(0,1)^{3}$| and the distribution function is chosen as |$f(t)=f_{0}\,e^{-t/10}$|⁠.

Throughout this section, we fix the number of computed eigenvalues by |$L_{H}=10$|⁠. Computing more eigenvalues does not improve the results due to the fact that |$f_{H}(\varepsilon _{l,H}[u])=0$| for |$l>L_{H}$|⁠. In practice, we can also compute the eigenvalues |$\varepsilon _{l,H}$| one by one from (3.5) until |$f_{H}(\varepsilon _{l,H}[u])=0$|⁠. Then the number |$L_{H}$| is set by |$l-1$|⁠.

 

Example 6.1.
The operator formula has a sequence of eigenvalues |$\lambda _{n_{x},n_{y},n_{z}}$| and corresponding functions |$\phi _{n_{x},n_{y},n_{z}}$|⁠, which are given by
We arrange the eigenvalues in ascending order and replace triple indices |$(n_{x},n_{y},n_{z})$| by single indices such that the eigenvalues are written as |$\lambda _{1}\leq \lambda _{2}\leq \lambda _{3}\leq \cdots $|⁠, and the corresponding eigenfunctions are denoted by |$\phi _{1}, \phi _{2}, \phi _{3},\cdots $|⁠.

Define |$V_{0}= \sin (\pi x)\sin (\pi y)\sin (\pi z)$|⁠, |$f_{0}=1$|⁠, |$\varepsilon _{F}=0$| and |$n_{D}=\sum \limits _{l=1}^\infty f(\lambda _{l}-\varepsilon _{F})\phi _{l}^{2} -\varDelta V_{0}$|⁠. The exact solution of the problem is given by |$V=V_{0}$|⁠.

To illustrate the convergence rates of the numerical solutions, we choose a tetrahedral mesh for the calculations. The fine mesh is obtained by uniform refinements of the coarse mesh. Table 1 shows that optimal convergence rates are obtained for the numerical potential when |$h\leq 2H^{2}$|⁠, that is, formula.

Table 1

Convergence orders of |$V_{h}$|⁠. (Example 6.1)

|$N_{h}$||$h$||$N_{H}$||$H$|formulaorder
|$125$||$1/4$||$125$||$1/4$|4.64e−01
|$729$||$1/8$||$125$||$1/4$|1.89e−011.29
|$4913$||$1/16$||$729$||$1/8$|8.25e−021.15
|$35937$||$1/32$||$729$||$1/8$|3.99e−021.05
|$274625$||$1/64$||$4913$||$1/16$|1.95e−021.03
|$N_{h}$||$h$||$N_{H}$||$H$|formulaorder
|$125$||$1/4$||$125$||$1/4$|4.64e−01
|$729$||$1/8$||$125$||$1/4$|1.89e−011.29
|$4913$||$1/16$||$729$||$1/8$|8.25e−021.15
|$35937$||$1/32$||$729$||$1/8$|3.99e−021.05
|$274625$||$1/64$||$4913$||$1/16$|1.95e−021.03
Table 1

Convergence orders of |$V_{h}$|⁠. (Example 6.1)

|$N_{h}$||$h$||$N_{H}$||$H$|formulaorder
|$125$||$1/4$||$125$||$1/4$|4.64e−01
|$729$||$1/8$||$125$||$1/4$|1.89e−011.29
|$4913$||$1/16$||$729$||$1/8$|8.25e−021.15
|$35937$||$1/32$||$729$||$1/8$|3.99e−021.05
|$274625$||$1/64$||$4913$||$1/16$|1.95e−021.03
|$N_{h}$||$h$||$N_{H}$||$H$|formulaorder
|$125$||$1/4$||$125$||$1/4$|4.64e−01
|$729$||$1/8$||$125$||$1/4$|1.89e−011.29
|$4913$||$1/16$||$729$||$1/8$|8.25e−021.15
|$35937$||$1/32$||$729$||$1/8$|3.99e−021.05
|$274625$||$1/64$||$4913$||$1/16$|1.95e−021.03
Table 2

Number of iterations: fixed-point ative method versus inexact Newton method. (Example 6.2)

  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$29$||$5$||$19$||$6$|
|$4913$||$729$||$36$||$5$||$19$||$6$|
|$35937$||$729$||$39$||$5$||$33$||$6$|
|$274625$||$4913$||$42$||$5$||$33$||$6$|
  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$29$||$5$||$19$||$6$|
|$4913$||$729$||$36$||$5$||$19$||$6$|
|$35937$||$729$||$39$||$5$||$33$||$6$|
|$274625$||$4913$||$42$||$5$||$33$||$6$|
Table 2

Number of iterations: fixed-point ative method versus inexact Newton method. (Example 6.2)

  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$29$||$5$||$19$||$6$|
|$4913$||$729$||$36$||$5$||$19$||$6$|
|$35937$||$729$||$39$||$5$||$33$||$6$|
|$274625$||$4913$||$42$||$5$||$33$||$6$|
  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$29$||$5$||$19$||$6$|
|$4913$||$729$||$36$||$5$||$19$||$6$|
|$35937$||$729$||$39$||$5$||$33$||$6$|
|$274625$||$4913$||$42$||$5$||$33$||$6$|
Table 3

Computing time (⁠|$seconds$|⁠). (Example 6.2)

  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$8.77$||$5.84$||$3.72$||$2.08$|
|$4913$||$729$||$83.78$||$36.47$||$28.61$||$15.21$|
|$35937$||$729$||$792.07$||$307.06$||$425.27$||$126.51$|
|$274625$||$4913$||$8287.22$||$2696.86$||$3573.88$||$1026.52$|
  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$8.77$||$5.84$||$3.72$||$2.08$|
|$4913$||$729$||$83.78$||$36.47$||$28.61$||$15.21$|
|$35937$||$729$||$792.07$||$307.06$||$425.27$||$126.51$|
|$274625$||$4913$||$8287.22$||$2696.86$||$3573.88$||$1026.52$|
Table 3

Computing time (⁠|$seconds$|⁠). (Example 6.2)

  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$8.77$||$5.84$||$3.72$||$2.08$|
|$4913$||$729$||$83.78$||$36.47$||$28.61$||$15.21$|
|$35937$||$729$||$792.07$||$307.06$||$425.27$||$126.51$|
|$274625$||$4913$||$8287.22$||$2696.86$||$3573.88$||$1026.52$|
  one gridtwo grids
|$N_{h}$||$N_{H}$|fixed-pointinexact Newtonfixed-pointinexact Newton
|$729$||$125$||$8.77$||$5.84$||$3.72$||$2.08$|
|$4913$||$729$||$83.78$||$36.47$||$28.61$||$15.21$|
|$35937$||$729$||$792.07$||$307.06$||$425.27$||$126.51$|
|$274625$||$4913$||$8287.22$||$2696.86$||$3573.88$||$1026.52$|

 

Example 6.2.
The purpose of this example is to illustrate the efficiency of the inexact Newton method. We test the inexact Newton method by solving the problem with both one-grid finite element method, meaning that both the Poisson equation and the Schrödinger eigenvalue problem are solved on the fine mesh |${\cal T}_{h}$|⁠, and two-grid finite element method, meaning that the Poisson equation is solved on |${\cal T}_{h}$| and the Schrödinger eigenvalue problem is solved on the coarse mesh |${\cal T}_{H}$|⁠. Define
The exact solution of the problem is given by |$V=V_{0}$|⁠.

The information about both fine meshes and coarse meshes are listed in Table 1. As a comparison, we also present the numbers of fixed-point iterations for solving the nonlinear discrete Poisson problem. The tolerance for relative residuals is set by |$10^{-8}$|⁠.

Table 2 lists the numbers of iterations required for solving the coupled SP system with both the fixed-point iterative method and the inexact Newton method. Clearly the inexact Newton method needs much less iterations than the fixed-point iterative method and is robust as the number of DOFs increases.

In Fig. 1, we show the numbers of iterations for both the inexact Newton method and the fixed-point iterative method on a fine mesh with |$N_{h}=729$| and a coarse mesh with |$N_{H}=125$|⁠. We observe that the inexact Newton method has second-order convergence, while the fixed-point iterative method on a single mesh has first-order convergence.

Fixed-point iterative method versus inexact Newton method.
Fig. 1.

Fixed-point iterative method versus inexact Newton method.

Table 3 lists the computing time for solving the problem by different methods. We find that the two-grid inexact Newton method takes less time than other methods. Moreover, the two Newton-type methods are much superior to the fixed-point method in terms of computing time.

Funding

Tao Cui was supported by National Key R & D Program of China 2019YFA0709600 and 2019YFA0709602. Weiying Zheng was supported in part by National Key R & D Program of China 2019YFA0709600 and 2019YFA0709602, and the National Science Fund for Distinguished Young Scholars 11725106.

References

Amrouche
,
C.
,
Bernardi
,
C.
,
Dauge
,
M.
&
Girault
,
V.
(
1998
)
Vector potential in three-dimensional nonsmooth domains
.
Math. Methods Appl. Sci.
,
21
,
823
864
.

Ancona
,
M. G.
&
Iafrate
,
G. J.
(
1989
)
Quantum correction of the equation of state of an electron gas in a semiconductor
.
Phys. Rev. B
,
39
,
9536
9540
.

Ancona
,
M. G.
&
Tiersten
,
H. F.
(
1987
)
Macroscopic physics of the silicon inversion layer
.
Phys. Rev.
,
35
,
7959
7965
.

Ando
,
T.
,
Fowler
,
A. B.
&
Stern
,
F.
(
1982
)
Electronic properties of two-dimensional systems
.
Rev. Mod. Phys.
,
54
,
437
672
.

Beaumont
,
S. P.
&
Sotomayor-Torres
,
C. M.
(
1990
)
Science and Engineering of One and Zero-Dimensional Semiconductors
.
Boston, MA
:
Springer US
.

Ben Abdallah
,
N.
,
Cáceres
,
M. J.
,
Carrillo
,
J. A.
&
Vecil
,
F.
(
2009
)
A deterministic solver for a hybrid quantum-classical transport model in nanoMOSFETs
.
J. Comput. Phys.
,
228
,
6553
6571
.

Ben Abdallah
,
N.
&
Unterreiter
,
A.
(
1998
)
On the stationary quantum drift diffusion model
.
Z. Angew. Math. Phys.
,
49
,
251
275
.

de Falco
,
C.
,
Gatti
,
E.
,
Lacaita
,
A. L.
&
Sacco
,
R.
(
2005
)
Quantum-corrected drift-diffusion models for transport in semiconductor devices
.
J. Comput. Phys.
,
204
,
533
561
.

Kaiser
,
H. C.
&
Rehberg
,
J.
(
1997
)
On stationary Schrödinger–Poisson equations modelling an electron gas with reduced dimension
.
Math. Methods Appl. Sci.
,
20
,
1283
1312
.

Kaiser
,
H. C.
&
Rehberg
,
J.
(
2000
)
About a stationary Schrödinger–Poisson system with Kohn-Sham potential in a bounded two or three-dimensional domain
.
Z. Angew. Math. Phys.
,
41
,
33
72
.

Landauer
,
R.
(
1988
)
Spatial variation of currents and fields due to localized scatterers in metallic conduction
.
IBM J. Res. Dev.
,
32
,
306
316
.

Laux
,
S. E.
&
Frank
,
S.
(
1986
)
Electron states in narrow gate-induced channels in Si
.
Appl. Phys. Lett.
,
49
,
91
93
.

Monk
,
P.
(
2003
)
Finite Element Methods for Maxwells Equations
. Oxford:
Oxford University Press
.

Mu
,
P.
&
Zheng
,
W.
(
2023
)
A positivity-preserving finite element method for quantum drift-diffusion model
.
J. Comput. Math.
,
41
, 909–932.

Nier
,
F.
(
1993
)
A variational formulation of Schrödinger–Poisson systems in dimension d |$\leq $| 3
.
Comm. Partial Differ. Equ.
,
18
,
1125
1147
.

Ohkura
,
Y.
(
1990
)
Quantum effects in Si n-MOS inversion layer at high substrate concentration
.
Solid State Electron.
,
33
,
1581
1585
.

Pinnau
,
R.
(
2002
)
A review on the quantum drift diffusion model
.
Transp. Theory Statist. Phys.
,
31
,
367
395
.

Reed
,
M.
&
Simon
,
B.
(
1975
)
Methods of Modern Mathematical Physics
.
New York, San Francisco, and London
:
Academic Press
.

Spinelli
,
A. S.
,
Benvenuti
,
A.
&
Pacelli
,
A.
(
1998
)
Self-consistent 2-D model for quantum effects in n-MOS transistors
.
IEEE
,
45
,
1342
1349
.

Stern
,
F.
(
1972
)
Self-consistent results for n-type Si inversion layers
.
Phys. Rev. B
,
5
,
4891
4899
.

Vinter
,
B.
(
1984
)
Subbands and charge control in a two-dimensional electron gas field-effect transistor
.
Appl. Phys. Lett.
,
44
,
307
309
.

Wu
,
C.-Y.
,
Banerjee
,
S.
,
Sadra
,
K.
,
Streetman
,
B. G.
&
Sivan
,
R.
(
1996
)
Quantization effects in inversion layers of PMOSFETs on Si (100) substrates
.
IEEE
,
17
,
276
278
.

Zeider
,
E.
(
1990
)
Nonlinear Functional Analysis and its Applications, II/B: Nonlinear Monotone Operators
.
Springer-Verlag
:
New York Inc.

Zhang
,
L.
,
Cao
,
L.
&
Luo
,
J.
(
2014
)
Multiscale analysis and computation for a stationary Schrödinger–Poisson system in heterogeneous nanostructures
.
Multiscale Model. Simul.
,
12
,
1561
1591
.

This is an Open Access article distributed under the terms of the Creative Commons Attribution NonCommercial-NoDerivs licence (https://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial reproduction and distribution of the work, in any medium, provided the original work is not altered or transformed in any way, and that the work properly cited. For commercial re-use, please contact [email protected]