-
PDF
- Split View
-
Views
-
Cite
Cite
Tao Cui, Wenhao Lu, Naiyan Pan, Weiying Zheng, A two-grid finite element method for the Schrödinger–Poisson system, IMA Journal of Numerical Analysis, 2025;, draf014, https://doi.org/10.1093/imanum/draf014
- Share Icon Share
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 $| is denoted by
and its inner product is denoted by |$(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)$$ \begin{align}& f(t)= f_{0}\,e^{-\mu (t-\varepsilon_{F})}, \qquad f(t)= f_{0}/(1+e^{\mu (t-\varepsilon_{F})}), \qquad f_{0}>0.\end{align} $$
|$V_{0}\in L^{2}(\varOmega )$| 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
, 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)$$ \begin{align}& -\varDelta z = \xi \quad \hbox{in}\;\;\varOmega,\qquad z=0 \quad \hbox{on}\;\;\partial\varOmega.\end{align} $$ 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})$|.
2. The SP system
The SP model is the coupling of the Poisson equation for the electric potential and the eigenvalue problem of the Hamitonian. After proper nondimensionalization, the Poisson equation has the form
where |$V$| is the electrostatic potential, |$n_{D}\in L^{2}(\varOmega )$| is the doping profile and is the electron-density operator defined by
Here |$(\varepsilon _{l},\psi _{l})$|, |$l\ge 1$|, denote the eigenpairs of the Hamiltonian |${\cal H}[u] = -\varDelta +u+V_{0}$|, namely

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 . Moreover, the electron-density operator
is Fréchet-differentiable and monotone, namely







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



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

Since |$f$| deceases exponentially, we can assume that, with a constant |$C$| independent of |$H$|,
The two-grid finite element method for solving (2.1) is to seek |$V_{h}\in X_{h}$| such that
where is the discrete electron-density operator defined as follows:

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

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 |$\varepsilon _{l}[u] \le \varepsilon _{l,H}[u]$|.
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 |$V_{h}$|.
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
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}]$| and satisfiesMoreover, |$\varGamma $| is oriented such that the interval |$[\varepsilon _{1,H}, \varepsilon _{N_{H},H}]$| lies on its left-hand side.(3.9)$$ \begin{align}& \mathrm{dist}(\varepsilon_{l,H},\varGamma)\geq\delta,\qquad 1\le l\le N_{H}.\end{align} $$
- (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)$$ \begin{align}& \int_{\varGamma}\left|{z}\right|{}^{k}\big|f_{H}^{(-j)}(z)\big|\,\mathrm{d} z \le C\quad \hbox{for}\;\; j=1,2\;\; \hbox{and}\;\; k=0, 1,2,3.\end{align} $$
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
, 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}}$|.














for all
.
Since |$f_{H}$| is non-increasing, the conclusion comes directly from Lemma 3.4.
3.3 Well-posedness of the discrete problem
We end this section by proving the well-posedness of problem (3.3).






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 ,

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

where |$C_{2},C_{3}$| are two positive constants depending only on |$\varOmega $|.
For any |$g\in H^{-1}(\varOmega )$|, the elliptic problem

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

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



Let denote the Banach algebra of linear and bounded operators on
, 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$|-summable operators on is denoted by |$(\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 ]$| as follows:
It is clear that |${\cal M}[\eta ]$| maps |$L^{\frac{qr}{q-r}}(\varOmega )$| to |$L^{r}(\varOmega )$| for |$1\le r\le q$| and maps 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 |$X_{H}$|.







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
It is left to estimate . 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
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 satisfiesMoreover, |$\hat \varGamma $| is oriented such that the interval |$\big [0,(\varepsilon _{1}+\rho _{u})^{-1}\big ]$| lies on its left-hand side.(4.11)$$ \begin{align}& \mathrm{dist}(t,\hat\varGamma)\geq\delta\quad\;\; \forall\, t\in [0,(\varepsilon_{1}+\rho_{u})^{-1}].\end{align} $$
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.







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







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]$|.

5.2 The Newton–Raphson method
Next we define the discrete residual operator |${\cal P}_{h}: X_{h}\to X_{h}^{\prime}$| as follows:
Then (3.3) is equivalent to the nonlinear equation
From Lemma 3.4, |${\cal P}_{h}$| is Frechét-differentiable at |$w_{h}\in X_{h}$| and the Frechét derivative satisfies
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
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.
While |$(e_{k} \ge \epsilon \;\; \& \;\; k< N_{\mathrm{itr}})$|, do
Compute all eigenpairs |$(\varepsilon _{l,H},\psi _{l,H})$|, |$1\le l\le N_{H}$|, by solving (3.5) with |$V_{h}$| replaced by |$V_{h}^{(k)}$|.
Compute the electron density |$n_{H} = \sum \limits _{l=1}^{L_{H}}f_{H}(\varepsilon _{l,H})\psi _{l,H}^{2}$|.
- 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$$ \begin{align*}& (\nabla r_{k},\nabla v_{h})-\big(n^{\prime}_{H}(r_{k}),v_{h}\big) =\big(\nabla V_{h}^{(k)},\nabla v_{h}\big) + (n_{D}-n_{H}, v_{h}) \quad \forall\, v_{h}\in X_{h}, \end{align*} $$$$ \begin{align*} n^{\prime}_{H}(r_{k}) =\sum^{N_{H}}_{i,j=1}\frac{f_{H}(\varepsilon_{i,H})-f_{H}(\varepsilon_{j,H})} {\varepsilon_{i,H}-\varepsilon_{j,H}} \bigg(\int_{\varOmega}r_{k}\psi_{i,H}\psi_{j,H}\bigg)\psi_{i,H}\psi_{j,H}. \end{align*} $$
- Update the approximate solutionEnd while.


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
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
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 |$\tilde{V}_{h}^{(k)}$|.
Compute the electron density |$\tilde n_{H} = \sum \limits _{l=1}^{L_{H}}f_{H}(\varepsilon _{l,H})\psi _{l,H}^{2}$|.
- Solve (5.6) for the correction function |$r_{k}\in X_{h}$|, that is,where the approximate Fréchet derivative at |$\tilde{V}_{h}^{(k)}$| is defined by(5.8)$$ \begin{align}& (\nabla r_{k},\nabla v_{h})-\big(\tilde{n}^{\prime}_{H}(r_{k}),v_{h}\big) =(\nabla \tilde{V}_{h}^{(k)},\nabla v_{h}) + (n_{D}-\tilde{n}_{H}, v_{h}) \quad \forall\, v_{h}\in X_{h},\end{align} $$$$ \begin{align*} \tilde{n}^{\prime}_{H}(r_{k}) =\sum^{L_{H}}_{i,j=1}\frac{f_{H}(\varepsilon_{i,H})-f_{H}(\varepsilon_{j,H})} {\varepsilon_{i,H}-\varepsilon_{j,H}} \bigg(\int_{\varOmega}r_{k}\psi_{i,H}\psi_{j,H}\bigg)\psi_{i,H}\psi_{j,H}. \end{align*} $$
- Update the approximate solutionEnd 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 |$l-1$|.

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, .
|$N_{h}$| . | |$h$| . | |$N_{H}$| . | |$H$| . | ![]() | order . |
---|---|---|---|---|---|
|$125$| | |$1/4$| | |$125$| | |$1/4$| | 4.64e−01 | — |
|$729$| | |$1/8$| | |$125$| | |$1/4$| | 1.89e−01 | 1.29 |
|$4913$| | |$1/16$| | |$729$| | |$1/8$| | 8.25e−02 | 1.15 |
|$35937$| | |$1/32$| | |$729$| | |$1/8$| | 3.99e−02 | 1.05 |
|$274625$| | |$1/64$| | |$4913$| | |$1/16$| | 1.95e−02 | 1.03 |
|$N_{h}$| . | |$h$| . | |$N_{H}$| . | |$H$| . | ![]() | order . |
---|---|---|---|---|---|
|$125$| | |$1/4$| | |$125$| | |$1/4$| | 4.64e−01 | — |
|$729$| | |$1/8$| | |$125$| | |$1/4$| | 1.89e−01 | 1.29 |
|$4913$| | |$1/16$| | |$729$| | |$1/8$| | 8.25e−02 | 1.15 |
|$35937$| | |$1/32$| | |$729$| | |$1/8$| | 3.99e−02 | 1.05 |
|$274625$| | |$1/64$| | |$4913$| | |$1/16$| | 1.95e−02 | 1.03 |
|$N_{h}$| . | |$h$| . | |$N_{H}$| . | |$H$| . | ![]() | order . |
---|---|---|---|---|---|
|$125$| | |$1/4$| | |$125$| | |$1/4$| | 4.64e−01 | — |
|$729$| | |$1/8$| | |$125$| | |$1/4$| | 1.89e−01 | 1.29 |
|$4913$| | |$1/16$| | |$729$| | |$1/8$| | 8.25e−02 | 1.15 |
|$35937$| | |$1/32$| | |$729$| | |$1/8$| | 3.99e−02 | 1.05 |
|$274625$| | |$1/64$| | |$4913$| | |$1/16$| | 1.95e−02 | 1.03 |
|$N_{h}$| . | |$h$| . | |$N_{H}$| . | |$H$| . | ![]() | order . |
---|---|---|---|---|---|
|$125$| | |$1/4$| | |$125$| | |$1/4$| | 4.64e−01 | — |
|$729$| | |$1/8$| | |$125$| | |$1/4$| | 1.89e−01 | 1.29 |
|$4913$| | |$1/16$| | |$729$| | |$1/8$| | 8.25e−02 | 1.15 |
|$35937$| | |$1/32$| | |$729$| | |$1/8$| | 3.99e−02 | 1.05 |
|$274625$| | |$1/64$| | |$4913$| | |$1/16$| | 1.95e−02 | 1.03 |
Number of iterations: fixed-point ative method versus inexact Newton method. (Example 6.2)
. | . | one grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact 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 grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact Newton . |
|$729$| | |$125$| | |$29$| | |$5$| | |$19$| | |$6$| |
|$4913$| | |$729$| | |$36$| | |$5$| | |$19$| | |$6$| |
|$35937$| | |$729$| | |$39$| | |$5$| | |$33$| | |$6$| |
|$274625$| | |$4913$| | |$42$| | |$5$| | |$33$| | |$6$| |
Number of iterations: fixed-point ative method versus inexact Newton method. (Example 6.2)
. | . | one grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact 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 grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact 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 grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact 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 grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact 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 grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact 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 grid . | two grids . | ||
---|---|---|---|---|---|
|$N_{h}$| . | |$N_{H}$| . | fixed-point . | inexact Newton . | fixed-point . | inexact Newton . |
|$729$| | |$125$| | |$8.77$| | |$5.84$| | |$3.72$| | |$2.08$| |
|$4913$| | |$729$| | |$83.78$| | |$36.47$| | |$28.61$| | |$15.21$| |
|$35937$| | |$729$| | |$792.07$| | |$307.06$| | |$425.27$| | |$126.51$| |
|$274625$| | |$4913$| | |$8287.22$| | |$2696.86$| | |$3573.88$| | |$1026.52$| |
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.

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.