Home The Partition of Unity Finite Element Method for the Schrödinger Equation
Article Open Access

The Partition of Unity Finite Element Method for the Schrödinger Equation

  • Daniele Boffi , Ondrej Certik , Francesca Gardini EMAIL logo and Gianmarco Manzini
Published/Copyright: May 16, 2024

Abstract

A Schrödinger equation for the system’s wavefunctions in a parallelepiped unit cell subject to Bloch-periodic boundary conditions must be solved repeatedly in quantum mechanical computations to derive the materials’ properties. Recent studies have demonstrated how enriched finite element type Galerkin methods can substantially lower the number of degrees of freedom necessary to produce accurate solutions with respect to the standard plane-waves method. In particular, the flat-top partition of unity finite element method enriched with the radial eigenfunctions of the one-dimensional Schrödinger equation offers a very effective way of solving the three-dimensional Schrödinger eigenvalue problem. We investigate the theoretical properties of this approximation method, its well-posedness and stability, we prove its convergence and derive suitable bound for the ℎ- and 𝑝-refinement in the L 2 and energy norm for both the eigenvalues and the eigenfunctions. Finally, we confirm these theoretical results by applying this method to the eigenvalue problem of the one-electron Schrödinger equation with the harmonic potential, for which the exact solution is known.

MSC 2020: 65L60; 34L16

1 Introduction

According to the density functional theory (DFT) [25, 21, 17], we can derive material properties from ab initio quantum mechanics calculations by solving the system of the Khon–Sham equations [20].

This approach requires repeatedly solving the steady-state Schrödinger and Poisson equations on a parallelepiped unit cell with Bloch-periodic boundary conditions [3]. The Schrödinger equation is an eigenvalue problem for the Hamiltonian differential operator describing the energy of a given multi-electron atomic or molecular system. The eigenvalues are the discrete energy levels, while the eigenfunctions are the wavefunctions of the corresponding quantum states [28, 24].

DFT represents a quantum mechanical computational methodology that facilitates the examination of the electronic structures and properties of molecular and solid-state systems. The foundational principle of DFT resides in the conceptualization of electronic systems via electron density profiles, thereby circumventing the direct resolution of the complex many-body Schrödinger equation. Within the framework of DFT, the more difficult many-body problem is effectively transformed into a tractable single-electron effective potential equation, analogous to a three-dimensional Schrödinger equation. The method delineated herein seeks to harness DFT within this scope.

The many-body Schrödinger equation delineates the interactions among the electrons in a system, presenting a formidable computational challenge, particularly for substantial systems. DFT streamlines this challenge by postulating the electron density as a fundamental descriptor. Electron density conveys the positional probability density of electrons within a spatial domain. The seminal Hohenberg–Kohn theorems assert that the ground-state electron density coherently determines the ground-state energy of a system. Consequently, it is feasible to ascertain the ground-state electron density and subsequently employ it to deduce diverse system properties, obviating the need to explicitly solve the many-body Schrödinger equation.

In the DFT paradigm, the derivation of the effective potential equation incorporates the electron-electron interactions as an effective field influencing each electron in isolation. This effective field amalgamates the external potential attributable to the atomic nuclei, and the exchange-correlation potential, which encapsulates the complexities of electron-electron repulsions. The objective is to pinpoint the electron density that minimizes the total energy of the system, an endeavor accomplished through the resolution of the effective potential equation.

DFT combined with the effective potential equation makes elucidating a system’s electronic structure and properties possible. By resolving the effective potential equation, one can determine the electron density and, thereby, calculate various properties, including the total energy, electronic charge distribution and the configuration of molecular orbitals. These properties yield profound insights into molecular and material behaviors and reactivities.

In essence, the proposed method leverages DFT by simplifying the many-body Schrödinger equation to a manageable single-electron, three-dimensional equation. The solution of this equation grants access to the electron density and ancillary properties of the system, thereby enabling a comprehensive exploration of the electronic structure and properties of molecular entities.

It is by far recognized that solving the Schrödinger equation is the most time-consuming step in such calculations. The plane wave (PW) pseudopotential method is currently the most advanced technology implemented in application software to accomplish this task. The PW method employs a Fourier basis set and calls for the resolution of a discrete standard eigenproblem.

However, Galerkin methods with enrichment can be highly competitive with the PW approach to obtain a numerical solution with a required accuracy using the same number of basis functions (or even less), i.e., by significantly reducing the number of degrees of freedom. An incomplete list of research articles exploring this possibility includes the references [2, 32, 22, 36, 26, 33, 34, 11, 27, 8, 18, 12]. The two most important issues with the enriched Galerkin method are the need to solve a generalized rather than a regular eigenvalue problem and the poor conditioning of the associated system matrices. These two crucial points can negatively impact computational costs. The partition of unity finite element method (PUFEM) [14] offers a very effective way to address these points and numerically solve the Schrödinger eigenvalue problem in a parallelepiped unit cell subject to Bloch-periodic boundary conditions. The partition of unity (PUM) [23, 5] is a generalization of the finite element method (FEM) for the spatial discretization of partial differential equations (PDEs); see also [4, 30]. Its application to the Schrödinger eigenvalue problem with Bloch-periodic boundary conditions was recently investigated in [32, 26]. The basic idea is that we cover the three-dimensional domain by overlapping patches so that each patch compactly supports a nonnegative weight function. These weight functions form a partition of unity and have the flat-top property, i.e., it is identically equal to one over some finite subset of its support. We employ 𝑝-th degree orthogonal polynomials on each patch built as the tensor product of univariate rescaled and translated Legendre polynomials. This formulation guarantees the 𝑝-th order completeness. We enrich the PUFEM by incorporating the radial Schrödinger’s eigenfunctions. The discrete variational form results in a generalized eigenvalue problem. Using the variational lumping of [31], we obtain a regular eigenvalue problem for which effective eigensolvers are available. This approach for solving the Schrödinger eigenvalue problem with the harmonic potential is accurate, stable and effective. In fact, for a given required accuracy in the energy eigenvalues, the suggested strategy provides optimal rates of convergence in the energy and the application of orbital enrichment dramatically lower the number of degrees of freedom.

In this work, we provide the proper mathematical and theoretical setting of the Schrödinger eigenvalue problem in a unit cube with Bloch-periodic condition for the partition of unity method. Our approach is based on the spectral approximation theory of compact operators, cf. [6] , and is, in practice, independent of the type of partition of unity functions that we can use in the PUFEM construction. To derive the convergence rates for the eigenvalue and eigenfunction approximation, we need the error bounds for the source problem, which is equivalent to a diffusion-reaction equation (the reaction term being the potential of the Hamilton operator). We obtain these estimates from the basic assumptions on the PUM of the original works of references [23, 5].

The paper is organized as follows. In Section 2, we present the Schrödinger eigenvalue problem on the unit cube with Bloch-periodic boundary conditions. In Section 3, we review the partition of unity finite element method. In Section 4, we prove the convergence of the method and deduce error estimates for the eigenvalue and eigenfunction approximation. In Section 5, we briefly detail the implementation of the method. In Section 6, we assess the numerical behavior of the method and show that it agrees with the theoretical expectations. In Section 7, we offer our final remarks and mention a few possible future research topics along the lines of this work.

2 The Schrödinger Eigenvalue Problem

In the context of atomic and molecular quantum theory, the Schrödinger eigenvalue problem on the domain Ω reads as

(2.1) H ψ ( x ) = ε ψ ( x ) for all x Ω ,

where H : = 1 2 Δ + V eff ( x ) is the Hamilton operator. The eigenvalue 𝜀 and the eigenfunction 𝜓 are the energy level and the corresponding wavefunction of the physical system described by ℋ. From density functional theory [7, 35, 16], the effective potential V eff ( x ) in ℋ collects all local and nonlocal interaction terms.

To describe a periodic condensed matter system, we consider the parallelepiped unit-cell Ω R d , with primitive lattice vectors a d , d = 1 , 2 , 3 , and the lattice translation vectors R = n 1 a 1 + n 2 a 2 + n 3 a 3 , for all n d Z , d = 1 , 2 , 3 . Then we assume that the effective potential is periodic in this lattice structure; hence, we have V eff ( x + R ) = V eff ( x ) . The wavefunction 𝜓 solving Schrödinger’s equation (2.1) must satisfy Bloch’s theorem; see [3, Chapter 8]. Therefore, for any eigenfunction 𝜓, there is a wavevector 𝐤 such that ψ ( x + R ) = ψ ( x ) exp ( i k R ) . For k = 0 (Γ-point), the wavefunction is periodic. For k 0 , there is a phase shift exp ( i k R ) associated with the translation 𝐑. In this setting, we write problem (2.1) as

1 2 Δ ψ ( x ) + V eff ( x ) ψ ( x ) = ε ψ ( x ) in Ω ,
(2.2) ψ ( x + a d ) = exp ( i k a d ) ψ ( x ) on Γ d ,
ψ ( x + a d ) n ̂ ( x ) = exp ( i k a d ) ψ ( x ) n ̂ ( x ) on Γ d ,
where n ̂ ( x ) is the outward unit normal at 𝐱 and Γ d is the set of the bounding faces of the domain Ω. The self-adjointness of the Hamiltonian ℋ implies that the eigenvalues 𝜀 are real numbers even though the boundary conditions and thus the wavefunction are complex-valued.

We consider the functional space of complex-valued functions

V : = { v H 1 ( Ω ; C ) : v ( x + a d ) = v ( x ) exp ( i k a d ) for all x Γ d , d = 1 , 2 , 3 } ,

endowed with the usual H 1 norm denoted by V ; the L 2 -scalar product will be denoted by

(2.3) ψ , v : = Ω v ( x ) ̄ ψ ( x ) d x .

We introduce the sesquilinear form A : V × V C ,

(2.4) A ( ψ , v ) : = Ω ( 1 2 v ( x ) ̄ ψ ( x ) + v ( x ) ̄ V eff ( x ) ψ ( x ) ) d x ,

and assume that V eff ( x ) is uniformly bounded from below and above, i.e., there exist two strictly positive constants V and V such that V V eff ( x ) V for almost every x Ω . Consequently, the sesquilinear form A ( , ) is coercive and continuous, that is, there exist positive constants 𝛼 and 𝐶 such that

A ( v , v ) α v V 2 for all v V , | A ( ψ , v ) | C ψ V v V for all ψ , v V .

The variational formulation of (2.1) reads as: find ( ε , ψ ) R × V , with ψ 0 , Ω = 1 , such that

(2.5) A ( ψ , v ) = ε ψ , v for all v V .

From the classic eigenvalue theory [9, 6], equation (2.5) admits an infinite discrete set of strictly positive eigenvalues, which diverge to infinity. Moreover, such eigenvalues can have a multiplicity greater than one with a finite-dimensional eigenspace. The eigenfunctions form a basis of 𝑉 that is orthogonal with respect to the L 2 -inner product (2.3) and the inner product defined by the sesquilinear form (2.4).

The discretization of (2.1) is carried out by introducing a sequence of conforming, finite-dimensional subspaces of 𝑉, denoted by V PU and numbered with the integer label . This label denotes the refinement level in the case of the ℎ-refinement and the degree of the polynomials in the case of the 𝑝-refinement. The construction of the approximation spaces V PU will be detailed in the next section. The discrete eigenvalue problem reads as: find ( ε , ψ ) R × V PU , with ψ 0 , Ω = 1 , such that

(2.6) A ( ψ , v ) = ε ψ , v for all v V PU .

We also consider the continuous source problem associated with the eigenvalue problem (2.5): find ψ s V such that

(2.7) A ( ψ s , v ) = g , v for all v V ,

where we assume that g L 2 ( Ω ) . It is standard to prove the well-posedness of problem (2.7), i.e., existence and uniqueness of its solution, by using the Lax–Milgram lemma [10]. In fact, due to the boundedness assumption on the potential field V eff , the bilinear form 𝒜 in (2.4) is coercive and the right-hand side of (2.7) is a continuous functional. Moreover, due to regularity result [1, 15], there exists r > 0 , depending only on Ω, such that ψ s H 1 + r ( Ω ; C ) . The following stability estimate holds:

(2.8) | ψ s | 1 + r C g 0 .

Eventually, it is known that r = 1 if the domain Ω is convex, as in our case, so we can deduce that ψ s H 2 ( Ω ; C ) .

Similarly, the discrete source problem reads as: find ψ s V PU such that

(2.9) A ( ψ s , v ) = g , v for all v V PU .

The well-posedness of problem (2.9) follows from the fact that all the approximation spaces V PU are conforming subspaces of 𝑉, and the continuous problem on 𝑉 is well-posed.

3 Partition of Unity Finite Element Method

In this section, we discuss the construction of the approximation space V PU used in the partition of unity finite element method (PUFEM). Since the indication of the refinement level is not important, we remove the sub-index ℓ to simplify the notation. The main ingredients of the PUFEM are

  1. a set of overlapping patches { Ω i } i , i = 1 , 2 , , n , forming a finite, open covering of the computational domain Ω, i.e., Ω i = 1 n Ω i ;

  2. a set of Lipschitz-continuous functions { ϕ i } i = 1 , 2 , , n such that supp ( ϕ i ) Ω ̄ i and forming a partition of unity, i.e., i = 1 n ϕ i ( x ) = 1 for all x Ω ;

  3. a set of local approximation spaces V i H 1 ( Ω i ) .

As it is commonly done, we consider the functions in V i extended to the entire domain Ω in order to give sense to the next formula. It turns out that only their value in Ω i is relevant. Then we define the global approximation space as

(3.1) V PU : = { v V : v ( x ) = i = 1 n ϕ i ( x ) v i ( x ) for some  v i V i } = ( i = 1 n ϕ i V i ) V .

The global regularity of the approximation space V PU stems from the global regularity of the partition of unity functions ϕ i and the overlapping of the patches Ω i . Since, by construction, the global space V PU is a subspace of 𝑉, its functions automatically satisfy the periodicity condition (2.2).

Let diam ( Ω i ) = sup x , x ′′ Ω i | x x ′′ | denote the diameter of the 𝑖-th patch. According to [23, 5], we consider the following assumption on the partition of unity.

  1. A positive integer 𝑀, and two positive real constants C and C G exist such that

    1. card { Ω i : x Ω i } M for every x Ω ;

    2. ϕ i , Ω i C ;

    3. ϕ i , Ω i C G diam ( Ω i ) .

The constant 𝑀 controls the overlap of the patches so that no more than 𝑀 patches can overlap at any point x Ω . The overlap of the patches ensures that the functions ϕ i form a sufficiently regular partition of unity. The bounds in (ii) and (iii) imply the H 1 -regularity of the functions of space V PU . A partition of unity satisfying A is called an ( M , C , C G ) partition of unity subordinate to the cover { Ω i } . Furthermore, the partition of unity is said to be of degree m ( m N { 0 } ) if ϕ i C m ( Ω ) for every 𝑖.

The property A of the partition of unity and the local approximation properties of the spaces V i determine the approximation result of References [23, 5], which we report below for the sake of completeness and future reference in the paper.

Theorem 1

Let { ϕ i } be a partition of unity subordinate to the cover { Ω i } satisfying assumption A. Assume that each function u H 1 ( Ω ) can be locally approximated by a function u i V i , i = 1 , 2 , , n , such that

u u i 0 , Ω i Ω ϵ 0 ( i ) , ( u u i ) 0 , Ω i Ω ϵ 1 ( i ) .

Then the function

u PU ( x ) = i = 1 n ϕ i ( x ) u i ( x ) for all x Ω

approximates 𝑢 in V PU and is such that

(3.2) u u PU 0 , Ω M C ( i = 1 n ϵ 0 2 ( i ) ) 1 2 ,
(3.3) ( u u PU ) 0 , Ω 2 M ( i = 1 n ( ( C G diam ( Ω i ) ) 2 ϵ 0 2 ( i ) + C 2 ϵ 1 2 ( i ) ) ) 1 2 .

In view of the previous theorem, the space V PU inherits the approximation properties of the local spaces V i . In particular, if the patches are star-shaped with respect to a ball of points and the local spaces contain the polynomials of degree at most 𝑝, we can bound the approximation errors ϵ 0 ( i ) and ϵ 1 ( i ) in the ℎ-refinement setting using the standard estimates of the polynomial interpolation in Sobolev spaces, cf. [13, 10]. Therefore, for every patch Ω i , it holds that

ϵ 0 ( i ) + diam ( Ω i ) ϵ 1 ( i ) C h min ( p , r ) + 1 | u | r + 1 , Ω i Ω .

Hence, using (3.2)–(3.3) from Theorem 1, we find the global approximation estimates for u PU ,

u u PU 0 , Ω + h | u u PU | 1 , Ω C h min ( p , r ) + 1 | u | r + 1 , Ω .

The constant 𝐶 is independent of ℎ, but depends on C , C G and 𝑀.

According to the standard convergence theory of the conforming finite element methods, cf. Cea’s lemma and the duality argument, we obtain the following bounds on the error for the solution of the source problem (2.9).

Lemma 1

Let g L 2 ( Ω ) be the load function on the right-hand side of problem (2.7). Let ψ s H 1 + r ( Ω ) , r 1 , be the solution of (2.7) and ψ s the solution of its partition of unity approximation (2.9). There exists a positive constant 𝐶 such that

ψ s ψ s 0 , Ω C h min ( p , r ) + 1 | ψ s | r + 1 , Ω , | ψ s ψ s | 1 , Ω C h min ( p , r ) | ψ s | r + 1 , Ω ,

where 𝐶 is independent of ℎ, but depends on C , C G and 𝑀.

If we consider the 𝑝-refinement setting, we can bound the approximation errors ϵ 0 ( i ) and ϵ 1 ( i ) as follows:

ϵ 0 ( i ) C p ( r + 1 ) u r + 1 , Ω i Ω and ϵ 1 ( i ) C p r u r + 1 , Ω i Ω .

Hence, using (3.2)–(3.3) from Theorem 1, we find the global approximation estimates for u PU ,

u u PU 0 , Ω C p ( r + 1 ) u r + 1 , Ω and | u u PU | 1 , Ω C p r u r + 1 , Ω .

Again, according to the standard convergence theory of the conforming finite element methods, we obtain the following bounds with respect to 𝑝 on the error for the solution of the source problem (2.9).

Lemma 2

Let g L 2 ( Ω ) be the load function on the right-hand side of problem (2.7). Let ψ s H 1 + r ( Ω ) , r 1 , be the solution of (2.7) and ψ s the solution of its partition of unity approximation (2.7). There exists a positive constant 𝐶 such that

ψ s ψ s 0 , Ω C p ( r + 1 ) | ψ s | r + 1 , Ω , | ψ s ψ s | 1 , Ω C p r | ψ s | r + 1 , Ω ,

where 𝐶 is independent of 𝑝, but depends on C , C G and 𝑀. Moreover, if the solution ψ s is analytic, we have that

| ψ s ψ s | 1 , Ω C exp ( b p )

for some constants 𝐶 and 𝑏 independent of 𝑝.

4 Convergence Analysis and Error Estimates

4.1 Spectral Approximation for Compact Operators

The eigenvalue problem we are interested in can be reformulated in the context of the approximation of eigenpairs of compacts operators. In this section, we review the main results of the theory of spectral approximation for compact operators [6, 9, 19]. To this end, we first define the continuous and discrete solution operators 𝑇 and T associated with problems (2.7) and (2.9), respectively.

Let T : L 2 ( Ω ) L 2 ( Ω ) be the operator mapping the loading term 𝑔 into ψ s = : T g , namely,

(4.1) T g V such that A ( T g , v ) = g , v for all v V .

Thanks to the well-posedness of the continuous source problem (2.7), and the self-adjointness and coercivity of A ( , ) , operator 𝑇 is well-defined, self-adjoint and positive-definite with respect to both A ( , ) and , . Moreover, the compact embedding of 𝑉 in L 2 ( Ω ) implies that operator 𝑇 is compact.

We define the discrete solution operator T : L 2 ( Ω ) V PU L 2 ( Ω ) in a similar way. Hence, such operator remaps the loading term 𝑔 into ψ s = : T g , namely,

T g V PU such that A ( T g , v ) = g , v for all v V PU .

Thanks to the well-posedness of the continuous source problem (2.7), and the self-adjointness and coercivity of A ( , ) , operator T is well-defined, self-adjoint and positive-definite with respect to both A ( , ) and , . Moreover, operator T is compact since its range is finite dimensional.

The eigenmodes of the operators 𝑇 and T are linked to the eigensolutions of the continuous and discrete eigenproblems by the following relation: ( ϵ , ψ ) is an eigenpair of the continuous problem if and only if ( 1 / ϵ , ψ ) is an eigenpair for the operator 𝑇, i.e., T ψ = ( 1 / ϵ ) ψ . A similar property must also hold in the discrete setting so that T ψ = ( 1 / ϵ ) ψ , where ( ϵ , ψ ) is an eigenpair of the discrete problem. Thanks to this correspondence, we will carry out the convergence analysis in the framework of the spectral approximation of compact operators.

To state the correct spectral approximation of operator 𝑇, we resort to the uniform convergence of the family of discrete operators { T } to 𝑇 (see [9, Proposition 7.4], cf. also [6]). Precisely, we prove that

(4.2) T T L ( L 2 ( Ω ) ) 0 , as  ,

or, equivalently,

(4.3) ( T T ) g 0 C ρ ( ) g 0 for all g L 2 ( Ω ) ,

with ρ ( ) tending to zero as ℓ goes to infinity.

Typically, the uniform convergence (4.3) follows from a priori estimates of the source problem without assuming any additional regularity on 𝑔. In addition to the convergence of the eigenmodes, condition (4.2), or its equivalent (4.3), implies that no spurious eigenvalues pollute the spectrum. Indeed, each discrete eigenvalue approximates a continuous eigenvalue. On the other hand, each continuous eigenvalue with a multiplicity m 1 is approximated by exactly 𝑚 discrete eigenvalues counted with their multiplicity.

Major results concerning the order of convergence of the eigenpairs are provided by the two following theorems. Their proofs are found in [6, Theorems 7.1–7.4], and [9, Theorems 9.3–9.7]. First, we define the gap between two subspaces 𝑈, W L 2 ( Ω ) as

δ ( U , W ) = max ( δ ̂ ( U , W ) , δ ̂ ( W , U ) ) ,

where

δ ̂ ( U , W ) = sup u U , u 0 = 1 inf w W u w 0 .

Theorem 2

Let the uniform convergence (4.2) hold true. Let 𝜆 and E λ denote an eigenvalue of 𝑇 with multiplicity 𝑚 and its corresponding eigenspace. Then there exist 𝑚 discrete eigenvalues λ 1 , , λ m (repeated according to their respective multiplicities) converging to 𝜆. Moreover, if E λ , denotes the direct sum of the eigenspaces of the discrete eigenvalues λ 1 , , λ m , then it holds

δ ( E λ , E λ , ) C ( T T ) | E λ L ( L 2 ( Ω ) ) .

We state the order of convergence of the eigenvalues in the following theorem.

Theorem 3

Let the uniform convergence (4.2) hold true. Let η 1 , , η m be a basis of the eigenspace E λ corresponding to the eigenvalue 𝜆. Then, for i = 1 , , m ,

(4.4) | λ λ i | C ( j , k = 1 m | ( T T ) η k , η j | + ( T T ) | E λ L ( L 2 ( Ω ) ) 2 ) ,

where λ 1 , , λ m are the 𝑚 discrete eigenvalues converging to 𝜆 repeated accordingly to their multiplicities.

4.2 Convergence Analysis for the ℎ- and 𝑝-Refinement

In this section, we provide the convergence analysis for the ℎ- and 𝑝-refinement of the partition of unity approximation of the eigensolutions of the Schrödinger equation.

Theorem 4

The family of operators T converges uniformly to the operator 𝑇, that is,

T T L ( L 2 ( Ω ) ) 0 for .

Proof

Consider the solutions ψ s and ψ s to the continuous and discrete source problems (2.7) and (2.9), respectively. In the case of the ℎ-refinement, we let t = min ( p , r ) , where 𝑝 is the polynomial order of the method and 𝑟 the regularity index of the solution, so that ψ s H 1 + r ( Ω ) . We apply the L 2 estimate of Theorem 1 with g L 2 ( Ω ) and inequality (2.8), and we find that

ψ s ψ s 0 , Ω C h t + 1 g 0 , Ω ,

where h is the mesh size parameter of the ℓ-th mesh refinement level, so that h 0 for . Then it follows that

T T L ( L 2 ( Ω ) ) = sup g L 2 ( Ω ) T g T g 0 g 0 = sup g L 2 ( Ω ) ψ s ψ s 0 g 0 C h t + 1 .

In the case of the 𝑝-refinement (recall that p = ), we apply the L 2 estimate of Theorem 2 with g L 2 ( Ω ) and inequality (2.8), and we find that

ψ s ψ s 0 , Ω C p ( r + 1 ) g 0 , Ω .

Then it follows that

T T L ( L 2 ( Ω ) ) = sup g L 2 ( Ω ) T g T g 0 g 0 = sup g L 2 ( Ω ) ψ s ψ s 0 g 0 C p ( r + 1 ) .

The order of convergence in the approximation of the eigenspaces is established in the following theorem.

Theorem 5

Let 𝜆 be an eigenvalue of 𝑇, with multiplicity 𝑚, and denote the corresponding eigenspace by E λ . Then exactly 𝑚 discrete eigenvalues λ 1 , , λ m (repeated according to their respective multiplicities) converge to 𝜆. Moreover, let E λ , be the direct sum of the eigenspaces corresponding to the discrete eigenvalues λ 1 , , λ m converging to 𝜆. Then

δ ( E λ , E λ , ) { C h t + 1 ( h -refinement ) , C p ( r + 1 ) ( p -refinement ) .

Proof

The proof follows the same argument of the previous theorem after restricting ( T T ) to E λ L 2 ( Ω ) . Thanks to the L 2 a priori error estimate for the discretization of the source problem, it holds

( T T ) | E λ L ( L 2 ( Ω ) ) = sup g E λ T g T g 0 g 0 = sup g E λ ψ s ψ s 0 g 0 { C h t + 1 ( h -refinement ) , C p ( r + 1 ) ( p -refinement ) .

The theorem’s assertion follows by using Theorem 2 and the above observation. ∎

Now, using the estimate for the gap δ ( E λ , E λ , ) , we obtain an upper bound for the approximation error of the eigenfunctions, cf. [6, 9]. This result is a straightforward consequence of Theorem 5.

Theorem 6

Let 𝜓 be a unit eigenfunction associated with the eigenvalue 𝜀 of multiplicity 𝑚 and let w ( 1 ) , , w ( m ) denote linearly independent eigenfunctions associated with the 𝑚 discrete eigenvalues of problem (2.6) converging to 𝜀. Then there exists ψ span { w ( 1 ) , , w ( m ) } such that

ψ ψ 0 { C h t + 1 ( h -refinement ) , C p ( r + 1 ) ( p -refinement ) ,

where 𝑟 is the regularity index of 𝜓; for the ℎ-refinement, t = min { p , r } , 𝑝 is the order of the method; for the 𝑝-refinement, 𝑝 is the polynomial order (recall that p = ).

Furthermore, in the following theorem, we prove the double order convergence of the eigenvalue approximation. Since the eigenvalues of the solution operators 𝑇 and T are the reciprocal of the eigenvalues 𝜀 and their approximations ε , we state the result as follows.

Theorem 7

Let 𝜀 be an eigenvalue of problem (2.5) with multiplicity 𝑚, and ε 1 , , ε m the 𝑚 discrete eigenvalues of problem (2.6) converging to 𝜀. Then the following optimal double-order convergence holds for all i = 1 , , m :

(4.5) | ε ε i | { C h 2 t ( h -refinement ) , C p 2 r ( p -refinement ) ,

where 𝑟 is the regularity index of 𝜓; for the ℎ-refinement, t = min { p , r } , 𝑝 is the order of the method; for the 𝑝-refinement, 𝑝 is the polynomial order (recall that p = ).

Proof

To prove the bound (4.5), we apply Theorem 3. Let again η 1 , , η m be a basis of the eigenspace E ε corresponding to the eigenvalue 𝜀. Using the definition of operator 𝑇 from (4.1) and the Galerkin orthogonality of the source problem, e.g., A ( T η k , ( T T ) η j ) = 0 , we note that

( T T ) η j , η k = A ( T η k , ( T T ) η j ) = A ( ( T T ) η k , ( T T ) η j ) .

We bound the right-most term above by using the continuity of the bilinear form A ( , ) and the a priori error estimate in the energy norm, cf. Lemmas 1 (ℎ-refinement) and 2 (𝑝-refinement), so that

| ( T T ) η j , η k | { C h 2 t ( h -refinement ) , C p 2 r ( p -refinement ) .

Theorem 5 implies that the second term of (4.4) is decreasing proportionally to h 2 ( t + 1 ) (ℎ-refinement) and p 2 ( r + 1 ) (𝑝-refinement), and the assertion of the theorem follows immediately. ∎

Finally, we provide an error estimate in the energy norm for the eigenfunction approximation.

Theorem 8

With the same notation as in Theorem 6, we have

| ψ ψ | 1 { C h t ( h -refinement ) , C p r ( p -refinement ) ,

where 𝑟 is the regularity index of 𝜓; for the ℎ-refinement, t = min { p , r } , 𝑝 is the order of the method; for the 𝑝-refinement, 𝑝 is the polynomial order (recall that p = ).

Proof

From the definition of 𝑇 and T , it follows that

ψ ψ = ε T ψ ε T ψ = ( ε ε ) T ψ + ε ( T T ) ψ + ε T ( ψ ψ ) .

Then

| ψ ψ | 1 | ε ε | | T ψ | 1 + ε | ( T T ) ψ | 1 + ε | T ( ψ ψ ) | 1 .

The first term at the right-hand side of the previous equation is of order h 2 t (ℎ-refinement) or p 2 r (𝑝-refinement), cf. Theorem 7. The second one is of order h t (ℎ-refinement), cf. Lemma 1, or p r (𝑝-refinement), cf. Lemma 2. To bound the third term, we use coercivity (with coercivity constant α ) of the bilinear form A ( , ) , the definition of the discrete solution operator T , the Cauchy–Schwarz inequality, the continuity of operator T with respect to the L 2 -inner product. We find the inequality chain

α | T ( ψ ψ ) | 1 2 A ( T ( ψ ψ ) , T ( ψ ψ ) ) = ψ ψ , T ( ψ ψ ) ψ ψ 0 T ( ψ ψ ) 0 T L ( L 2 ( Ω ) ) ψ ψ 0 2 { C h 2 t + 2 ( h -refinement ) , C p 2 ( r + 1 ) ( p -refinement ) .

The assertion of the theorem is a consequence of these upper bounds. ∎

4.3 Filtering PUFEM and Enriched PUFEM

While in the classical setting of PUFEM functions in V i are polynomials in Ω i , we are interested in the enriched version of PUFEM, where the elements of V i may be polynomials in Ω i or other functions, defined in an appropriate way in order to better approximate the exact solution. In order to formalize this concept, we might write the 𝑖-th local approximation space V i as

V i = V i P + V i E ,

where V i P = span { π k ( i ) } and V i E = span { ψ k ( i ) } . The notation refers to the fact that usually V i P will be the polynomial part of the space and V i E the enrichment.

Figure 1 
                  1D structure of the overlapping patches (the top line), and integration cells (bottom line) corresponding to the various patches.
The first and last integration cells are due to the overlap with two patches that are not shown in the figure.
Figure 1

1D structure of the overlapping patches (the top line), and integration cells (bottom line) corresponding to the various patches. The first and last integration cells are due to the overlap with two patches that are not shown in the figure.

We implement PUFEM and enriched PUFEM by assembling the global stiffness and mass matrices from local contributions. We compute these local contributions on a set of nonoverlapping integration cells that partition the computational domain by taking into account possible patch overlappings. For example, in the 1D case, see Figure 1, we distinguish two different types of integration cells:

  1. cells containing a single patch;

  2. cells shared by two overlapping patches.

In the former case, the local stiffness and mass matrices only contain the contributions from the basis functions { π k ( i ) } of the associated patch. In the latter case, such matrices depend on union of the basis functions of the overlapping patches. This procedure leads to a “redundant” construction and, eventually, to singular matrices.

The same consideration holds for the enrichment part of the local approximation spaces. Moreover, the enrichment procedure may suffer of numerical instability because the refinement may lead to an ill-conditioned basis for V i since some components in V i E might be better and better approximated by V i P . Similar issues exist also in the 2D and 3D cases; technical details are reported in the final appendix.

We now describe a filtering procedure that we are going to use in order to address these issues. At the end of the process, we are going to have a reduced space V ̃ i that will be presented as

V ̃ i = V ̃ i P V ̃ i D ,

where V ̃ i D is that part of V ̃ i E that is orthogonal to V ̃ i P .

The reduction procedure is based on a threshold tolerance τ > 0 that will be responsible to filtering the elements of V i that cause the ill-conditioning. The eigenvalues responsible for the ill-conditioning are very close to the machine precision, typically of the order of 10 14 , and are well-separated from the “good” eigenvalues, which are generally clustered around 1 after a matrix rescaling. Therefore, we can set up the threshold tolerance to a small value just above the machine precision, e.g., in the range [ τ = 10 13 , 10 12 ] . Hence, as long as 𝜏 is sufficiently small not to affect the “good” eigenvalues and sufficiently big to remove all the “bad” ones, the accuracy of the final approximation is independent of its choice.

The filtering process consists of three main steps: an initial rescaling of the functions forming a basis of V i P and V i E ; a reduction step based on a first internal orthogonalization leading to two reduced approximation spaces V ̃ i P and V ̃ i E ; a final orthogonalization step between the basis of V ̃ i P and V ̃ i E .

Consider the so called overlap (mass) matrix

S i = [ S i , P P S i , P E S i , E P S i , E E ] ,

where the entries are the block sub-matrices ( S i , P P ) | a b = ( π a i , π b i ) , ( S i , P E ) | a b = ( π a i , ψ b i ) , ( S i , E P ) | a b = ( ψ a i , π b i ) , and ( S i , E E ) | a b = ( ψ a i , ψ b i ) for every admissible value of the index pair ( a , b ) .

First Step

We rescale the basis functions π a ( i ) and ψ a ( i ) , and introduce two new sets of basis functions { π ̂ a ( i ) } and { ψ ̂ a ( i ) } that are defined as follows:

π ̂ a ( i ) = | Ω i | 1 / 2 π a ( i ) 0 , Ω i π a ( i ) and ψ ̂ a ( i ) = | Ω i | 1 / 2 π a ( i ) 0 , Ω i ψ a ( i ) .

According to [29], we adopt the weighted inner product on the functional space L 2 ( Ω i , C , { ϕ i } ) and, in practice, we build the enriched functional spaces through a double orthogonalization and reduction process that is aimed at improving the stability of the discrete problem without significantly affecting its convergence properties.

Second Step

By using these scaled basis functions, we recompute a new overlap matrix S ̂ i and note that its diagonal entries are equal to | Ω i | . Since the overlap matrix S ̂ i is real and symmetric, and so are the diagonal matrix blocks S ̂ i , P P and S ̂ i , E E , we perform the eigendecomposition

D i , P = O i , P T S ̂ i , P P O i , P and D i , E = O i , E T S ̂ i , E E O i , E ,

where D i , and O i , with { P , E } , are the eigenvalue and eigenvector matrices of, respectively, S ̂ i , P P and S ̂ i , E E . Note that the eigenvector matrices O i , are orthogonal, i.e., O i , T O i , = I i , . Using these matrices, we define a basis transformation providing two new sets of basis functions, namely, { π ̃ a ( i ) } and { ψ ̃ a ( i ) } given by the linear combinations of { π ̂ a ( i ) } and { ψ ̂ a ( i ) } , and using the components of the matrices O i , P and O i , E as the coefficients. We now make use of the tolerance 𝜏 and write

π ̃ a ( i ) = b ( O i , P ) a , b π ̂ b ( i ) = b : ( D i , P ) b b > τ ( O i , P ) a , b π ̂ b ( i ) + b : ( D i , P ) b b τ ( O i , P ) a , b π ̂ b ( i ) = π ̃ a ( i , 1 ) + π ̃ a ( i , 2 ) , ψ ̃ a ( i ) = b ( O i , P ) a , b ψ ̂ b ( i ) = b : ( D i , P ) b b > τ ( O i , E ) a , b ψ ̂ b ( i ) + b : ( D i , P ) b b τ ( O i , E ) a , b ψ ̂ b ( i ) = ψ ̃ a ( i , 1 ) + ψ ̃ a ( i , 2 ) .

In these relations, we split the summation over the index 𝑏 to separate the contributions corresponding to the eigenvalues bigger or smaller than the given tolerance factor 𝜏. This eigendecomposition has two crucial properties:

  1. { π ̃ a ( i ) } is an orthogonal basis set in V i P and { ψ ̃ a ( i ) } is an orthogonal basis set in V i E ;

  2. π ̃ a ( i , 2 ) 0 , Ω i 2 τ and ψ ̃ a ( i , 2 ) 0 , Ω i 2 τ .

Property (i) follows immediately from the definitions introduced so far. For example, for the basis set { π ̃ a ( i ) } , a straightforward calculation proves that

( π ̃ a ( i ) , π ̃ a ( i ) ) = b b ( O i , P ) a , b ( O i , P ) a , b ( π ̂ b ( i ) , π ̂ b ( i ) ) = [ ( O i , P ) T S ̂ O i , P ] a a = ( D i , P ) a a = ( D i , P ) a a δ a a .

Property (ii) readily follows by taking a = a in the calculation above and considering only the summation terms providing π ̃ a ( i , 2 ) . We obtain π ̃ a ( i , 2 ) 0 , Ω i 2 = ( D i , P ) a a τ . The same arguments prove the corresponding relations for ψ ̃ a ( i , 2 ) . At this point, we perform the spectral reduction by taking π ̃ a ( i ) π ̃ a ( i , 1 ) and ψ ̃ a ( i ) ψ ̃ a ( i , 1 ) . We denote the spaces generated by the linear combinations of these reduced set of basis functions by V ̃ i P and V ̃ i E , respectively. Note that these functions still form two sets of orthogonal functions.

Third Step

In the third and final step, we mutually orthogonalize the functions π ̃ a ( i ) and ψ ̃ a ( i ) so that the decomposition

V ̃ i = V ̃ i P V ̃ i D ,

holds, where now V ̃ i D is the space of functions of V ̃ i E that are orthogonal to V ̃ i P .

If we reformulate problem (2.6) by considering the space V \PU defined in (3.1) with V ̃ i instead of V i , we can repeat the convergence analysis of the previous sections by using the same arguments as long as the right approximation property is satisfied. In the next section, we will discuss how the spectral reduction strategy can be effectively implemented.

5 Implementation for the Schrödinger Equation

In our implementation, we use the flat-top partition of unit that was initially proposed in [2]. The design of the partition of unity method that we consider in this paper first requires a set of overlapping patches { Ω i } , i = 1 , 2 , , n . Let h = 1 / n be the mesh size parameter and consider the univariate partition of the unit interval [ 0 , 1 ] in every space direction given by the points x i s = 2 h i , s = 1 , , d , where d = 1 , 2 , 3 . Then we define the 𝑖-th patch in the 𝑠-th direction as

Ω i s = [ x i s α h , x i s + α h ] ,

where the real factor 𝛼 is between 1 and 2. The 𝑑-dimensional patch cover is then defined by taking the tensor product of the one-dimensional coverings. We label the multidimensional patches by using the multi-index i ¯ = ( i 1 , , i d ) , i s = 1 , , n . For example, if d = 3 , the patch associated with i ¯ = ( i 1 , i 2 , i 3 ) is

Ω i ¯ = [ x i 1 1 α h , x i 1 1 + α h ] × [ x i 2 2 α h , x i 2 2 + α h ] × [ x i 3 3 α h , x i 3 3 + α h ] .

The patches of this construction satisfy the overlap condition, i.e., assumption (A1), because a point of the domain Ω can belong at most to 2 d patches.

To define the set of partition of unity functions { ϕ i ¯ } , we can choose any nonnegative, multidimensional B-spline W : = [ 1 , 1 ] d R and compose it with a homothetic transformation T i ¯ that translates and rescales it onto the patch Ω i ¯ . We first obtain the auxiliary functions

W i ¯ ( x ) = { W T i ¯ ( x ) , x Ω i ¯ , 0 otherwise .

Then we normalize such functions and we obtain the partition of unity functions

ϕ i ¯ ( x ) = W i ¯ ( x ) j ¯ W j ¯ ( x ) for all i ¯ .

Each function ϕ i ¯ is nonnegative, compactly supported by its corresponding patch Ω i ¯ , and the set { ϕ i ¯ } is clearly a partition of unity due to the normalization condition.

The approximation space V i in each patch Ω i is given by the tensor product of the polynomials of degree 𝑝 in every direction.

5.1 Local Reduction of the Space Basis Functions

In the flat-top partition of unity method (FT-PUM), we consider overlapping patches with support { Ω i } i = 1 , , n , which provide a covering for the domain Ω. On the 𝑖-th patch with support Ω i , the local approximation consists of polynomial functions π k ( i ) and nonpolynomial (special) enrichment functions ψ k ( i ) . The global approximation is formed by gluing together these local approximations “weighted” by continuous functions ϕ i , which have the partition of unity property and the flat-top property. The flat-top property implies that the 𝑖-th function ϕ i is such that ϕ i = 1 on the (closed) sub-domain Ω i FT Ω i (the superscript FT stands for flat top). We reformulate the global approximation space (3.1) as

V PU = i = 1 n ϕ i ( x ) V i , where V i = span { { π k ( i ) } k , { ψ k ( i ) } k } .

Due to the flat-top property, we can also write

V PU = ϕ 1 V 1 ϕ 2 V 2 ϕ n V n .

Following [29], we devise a transformation over the support Ω i of each patch that yields a set of linearly independent local basis functions. This strategy makes the global approximation above more stable, cf. [32]. The final map 𝐓 consists of a sequence of matrix transformations from the linearly independent (unstable) local basis to a stable local basis. The resulting modifications of the Gram (overlap) matrix are presented in [2, pages 8–9].

Let T R M × N , M < N ( N = dim ( V PU ) ) be the rectangular (composite) matrix of the basis transformation, u R n N the vector collecting the (initial) degrees of freedom, and u ̃ R M the vector collecting the reduced degrees of freedom. Let T denote the Moore–Penrose inverse matrix of the rectangular matrix 𝐓 so that we can write

u = T T u ̃ , u ̃ = ( T ) T u .

Consider the discrete generalized eigenvalue problem that originated from the Bloch-periodic Schrödinger eigenproblem. Let H R N × N denote the matrix corresponding to the discretized Hamiltonian operator, and let S R N × N denote the matrix corresponding to the discretization of the right-hand side. Note that 𝐇 is the finite element stiffness matrix augmented with a term from the discretization of the potential energy, while 𝐒 is a mass matrix. Hereafter, we will refer to 𝐇 and 𝐒 as the Hamiltonian matrix and the overlap matrix, respectively, following a terminology widely accepted in the XFEM literature. The discrete generalized eigenvalue problem reads as

Hu = ϵ Su .

We multiply this equation by 𝐓 (from the left) and use the relation u = T T u ̃ (on the right), and we find that

THT T u ̃ = ϵ TST T u ̃ .

Introducing the notation H ̃ = THT T R M × M and S ̃ = TST T R M × M , we rewrite the transformed discrete eigenproblem as

H ̃ u ̃ = ϵ S ̃ u ̃ .

Since M < N (or even M N ), the transformed discrete eigenproblem is smaller in size than the original discrete eigenproblem. The global transformation matrix 𝐓 is constructed by computing patch-based stable transformation matrices T i . According to Section 4.3, we split the 𝑖-th local approximation space V i as follows:

V i = V i P V i E = V i P V i D ,

where V i P = span { π k ( i ) } , V i E = span { ψ k ( i ) } , and V i D is that part of V i E that is orthogonal to V i P . By splitting the polynomial and nonpolynomial parts of the space (here labeled by 𝒫 and ℰ, respectively), we write the overlap matrix S i as the block matrix

S i = [ S i , P P S i , P E S i , E P S i , E E ] .

All enrichment functions are scaled appropriately, and so the weighted L 2 ( Ω i , C , { ϕ i } ) -norm of all basis functions is set to the volume of the patch denoted by | Ω i | . Denote D i = diag ( S i ) . Then define the scaling transformation T i = | Ω i | i 1 / 2 D i 1 / 2 so that S i is transformed as

S ̂ i = T i S i ( T i ) T = [ S ̂ i , P P S ̂ i , P E S ̂ i , E P S ̂ i , E E ] .

Note that the diagonal entries of S ̂ i are equal to | Ω i | .

We introduce the matrices D i , and O i , with { P , E } , which, according to previous section, are the eigenvalue and eigenvector matrices of the eigendecomposition

D i , P = O i , P T S ̂ i , P P O i , P and D i , E = O i , E T S ̂ i , E E O i , E .

We use thresholding to discard eigenvectors that correspond to eigenvalues that are below a given tolerance. We denote these modified eigenvalue and eigenvector matrices as D ̃ i , P 1 / 2 , D ̃ i , E 1 / 2 and O ̃ i , P T , O ̃ i , E T . Then we define the block matrices

T i 1 = [ T ̃ i 1 0 0 I E ] , T i 2 = [ I P P 0 0 T ̃ i 2 ] ,

where

T ̃ i 1 = D ̃ i , P 1 / 2 O ̃ i , P T and T ̃ i 2 = D ̃ i , E 1 / 2 O ̃ i , E T .

These matrices represent projection operators that remove the near-null space of S ̂ i , P P and S ̂ i , E E . First, we apply T i 1 (and ( T i 1 ) T ) to S ̂ i and we obtain

S i 1 = T i 1 S ̂ i ( T i 1 ) T = [ I i , P P T ̃ i 1 S ̂ i , P E S ̂ i , P E ( T ̃ i 1 ) T S ̂ i , E E ]

since

D ̃ i , P 1 / 2 O ̃ i , P T S ̂ i , P P O i , P D ̃ i , P 1 / 2 = D ̃ i , P 1 / 2 D ̃ i , P D ̃ i , P 1 / 2 = I i , P P .

Then we apply T i 2 (and ( T i 2 ) T ) to S ̂ i 1 to get

S i 2 = T i 2 S i 1 ( T i 2 ) T = [ I i , P P T ̃ i 2 T ̃ i 1 S ̂ i , P E S ̂ i , P E ( T ̃ i 1 ) T ( T ̃ i 2 ) T I i , E E ]

since again

D ̃ i , E 1 / 2 O ̃ i , E T S ̂ i , E E O i , E D ̃ i , E 1 / 2 = D ̃ i , E 1 / 2 D ̃ i , E D ̃ i , E 1 / 2 = I i , E E .

We simplify the notation of the off-diagonal block matrix terms by introducing

S i , P E = T ̃ i 2 T ̃ i 1 S ̂ i , P E and S i , E P = S ̂ i , P E ( T ̃ i 1 ) T ( T ̃ i 2 ) T

so that we can write

S i 2 = T i 2 S i 1 ( T i 2 ) T = [ I i , P P S i , P E S i , E P I i , E E ] .

To remove all the polynomial components in the enrichment space (off-diagonal terms in S i 2 ), we consider the block matrix

T i 3 = [ I i , P P 0 S i , E P I i , E E ]

so that

S i 3 = T i 3 S i 2 ( T i 3 ) T = [ I i , P P 0 0 I E E S i , E P S i , P E ] = [ I i , P P 0 0 S i , D D ] ,

where we introduced the Schur complement

S i , D D : = I E E S i , E P S i , P E . = I E E S i , E P ( I P P ) 1 S i , P E .

Lastly, we perform an eigendecomposition of S i , D D to obtain an orthonormal basis of V i with respect to the L 2 -weighted inner product, and consider the transformation

T i 4 = [ I i , P P 0 0 T ̃ i 4 ] , where T ̃ i 4 = D ̃ i , D 1 / 2 O ̃ i , D T .

The corresponding transformation applied to the overlap matrix reads as

S i 4 = T 4 S i 3 ( T i 4 ) T .

Summarizing, the stable transformation matrix T i on the 𝑖-th patch can be decomposed as

T i = T i 4 T i 3 T i 2 T i 1

and is such that

T i S i ( T i ) T = [ T i 4 T i 3 T i 2 T i 1 ] S i [ ( T i 1 ) T ( T i 2 ) T ( T i 3 ) T ( T i 4 ) T ] = [ I i , P P 0 0 I i , D D ] .

Finally, we also apply the transformation matrix to the stiffness matrix deriving from the Hamiltonian ℋ.

Figure 2 
                  Relative error curves versus ℎ for the first four eigenvalues of the one-dimensional problem using the PUFEM with 
                        
                           
                              
                                 p
                                 =
                                 
                                    1
                                    ,
                                    2
                                    ,
                                    3
                                 
                              
                           
                           
                           p=1,2,3
                        
                     .
As indicated by the legend, the results for the enriched calculation are marked by a full symbol.
Figure 2

Relative error curves versus ℎ for the first four eigenvalues of the one-dimensional problem using the PUFEM with p = 1 , 2 , 3 . As indicated by the legend, the results for the enriched calculation are marked by a full symbol.

Figure 3 
                  Relative error curves versus ℎ for the first four eigenvalues of the two-dimensional problem using the PUFEM with 
                        
                           
                              
                                 p
                                 =
                                 
                                    1
                                    ,
                                    2
                                    ,
                                    3
                                 
                              
                           
                           
                           p=1,2,3
                        
                     .
As indicated by the legend, the results for the enriched calculation are marked by a full symbol.
Figure 3

Relative error curves versus ℎ for the first four eigenvalues of the two-dimensional problem using the PUFEM with p = 1 , 2 , 3 . As indicated by the legend, the results for the enriched calculation are marked by a full symbol.

6 Numerical Experiments

We assess the performance of the PUFEM by solving the one-dimensional and two-dimensional Schrödinger eigenvalue problem for the harmonic potential. This problem admits an exact solution that can be used to assess the accuracy of our eigensolver. In the one-dimensional (1D) case, the eigenvalues are the integer numbers, i.e., 1 , 2 , 3 , rescaled by the factor 1 / 2 and with multiplicity equal to one. In the two-dimensional (2D) case, the eigenvalues are still the integer numbers, but now their multiplicity is equal to the eigenvalue, i.e., the sequence of eigenvalues is 1 , 2 , 2 , 3 , 3 , 3 , 4 , 4 , 4 , 4 , 5 , 5 , 5 , 5 , 5 , . The corresponding eigenfunctions are the Hermite functions, which are the Hermite polynomials multiplied by the Gauss bell-shaped function exp ( r 2 ) , with r = x in 1D and r = x 2 + y 2 in 2D.

We consider two distinct applications of the PUFEM for the polynomial degree p = 1 , 2 , 3 in every patch without any enrichment, and assuming that all patch polynomial spaces are enriched by the first eigenfunction. Figures 2 and 3 report the relative errors for the first four eigenvalues. When we add the first eigenfunction to the approximation spaces, the corresponding eigenvalue must be computed exactly (up to the rounding errors due to the finite precision arithmetic and the ill-conditioning of the system). The top-left plots of both figures clearly reflect this behavior. In all other eigenvalues, the enrichment is clearly improving the accuracy of the numerical approximation by reducing the error level, without, however, changing the convergence rate. It must be noted that, in order to better appreciate the asymptotic behavior of our results, the computational tests should be performed on finer meshes.

7 Conclusions

In this work, we investigated the well-posedness and stability of the PUFEM based on the flat-top partition of unity, demonstrated its convergence, and derived appropriate error estimates for the eigenvalues and for the eigenfunctions in the L 2 and energy norm for both the ℎ- and 𝑝-refinement. We also validated these theoretical findings by applying this procedure to the known exact solution for the eigenvalue problem of the one-electron Schrödinger equation with harmonic potential.

Award Identifier / Grant number: 89233218CNA000001

Funding statement: The work of the second and last author was partially supported by the Laboratory Directed Research and Development (LDRD) program of Los Alamos National Laboratory. Los Alamos National Laboratory is operated by Triad National Security, LLC, for the National Nuclear Security Administration of the U.S. Department of Energy (Contract No. 89233218CNA000001).

A Practical Implementation Details

In the following pseudocode description of the PUFEM/FT-PUM, we separate the algorithm into distinct stages, such as building integration cells and patches, assembling Hamiltonian and overlap matrices, and solving the generalized eigenvalue problem for clarity in understanding each computational step. We discuss the use of real and integer topological data structures to define the integration cells, patches and mesh connectivity in the next subsections.

An error in the conversion from LaTeX to XML has occurred here. [!t]

The PUFEM/FT-PUM software encapsulated within the software implementation offers a structured and modular framework for handling the complexities of the problem. Numerical integration to compute the stiffness and mass matrix entries is carried out inside the integration cells. The integration cells are built by considering the intersection of the overlapping patches. These latter ones are obtained from an underlying cartesian mesh into a set of logical cells. Note the logical cells are never used in the algorithm implementation. The primary input variables are set in the calling main file and shared with all invoked routines (or reset in these last ones as parameters):

  • dims, number of dimensions (set in the main routine);

  • Np(1:dims), number of patches in every directions

  • xmin(1:dims), xmax(1:dims), computational domain as tensor product of all the directions

  • ppu, type of flat-top PU shape functions

  • penr, maximum degree of the polynomial basis;

  • npenr, the number of enriching functions;

  • alpha, patch extension outside the logical cell;

  • Nq, number of quadrature nodes

Such variables include the number of patches ( N p ), boundary definitions ( x min , x max ), and polynomial and enrichment function details (penr, npenr).

The parameters are flexibly set to accommodate various problem instances. For example, from penr, we derive the degree of the polynomial approximation, and therefore, the total number of polynomial basis functions is equal to penr + 1 in 1D, ( penr + 1 ) ( penr + 2 ) / 2 in 2D, and ( penr + 1 ) ( penr + 2 ) ( penr + 3 ) / 6 in 3D. Similarly, the primary variable npenr provides the total number of enriching functions, e.g., the Hermite functions for the Schrödinger equation with the harmonic potential.

The initialization and allocation of arrays for cells, patches, basis functions, and quadrature nodes follow a thorough consideration of computational efficiency and memory management. Then the routine build_integration_cells_patches builds the data structures of the problem, which we distinguish between real and topological.

The following subsections outline the procedure mainly for the one-dimensional case. However, it provides a foundation for extension to multidimensional scenarios. The careful consideration of boundary conditions, both periodic and nonperiodic, makes it possible to have a versatile application of the software to various physical problems.

Here is an explanation of how the FT-PUM algorithm handles the polynomial enrichment details, based on the previous sections.

  • Construction of basis functions: The algorithm constructs the local basis functions supported by the patches. These basis functions are enriched with additional enriching functions to better capture the solution’s behavior within each patch, especially in regions where the solution might have a higher gradient or other complexities.

  • Array allocation for enrichment: We describe the allocation of arrays to store information about the enriched basis functions. For instance, ib(penr+1+npenr,N_p) is an array where ib(i,e) stores the global index of the i-th basis function (polynomial and nonpolynomial) of the e-th patch.

  • Integration and assembly: During the integration and assembly phase entries of the stiffness and mass matrices are computed. The enrichment allows for a more accurate integration over the cells, particularly where patches overlap, as the enriched basis functions provide a better representation of the solution.

  • Solving the generalized eigenvalue problem: Finally, with the enriched system matrices, the algorithm solves a generalized eigenvalue problem to find the eigenvalues (eigs) and eigenvectors (c), which correspond to the energy levels and states of the quantum system that are described by the Schrödinger equation.

By incorporating polynomial enrichment into the finite element method, the FT-PUM algorithm can achieve high accuracy and convergence rates for quantum mechanical problems that may exhibit rapid changes in the solution domain, such as those encountered in electronic structure calculations.

A.1 Real Data Structures

In the scheme, integration cells and patches are used in a different way. Patches support the local basis functions, while integration cells are the portion of the domain where the entries of the stiffness and mass matrices are practically computed (or integrated). The patches can overlap and normally do that in the integration cells. Instead, the integration cells cannot have a substantial intersection.

The real data structures define the integration cells and the patches as 1D segments or tensor product of 1D segments in the multi-dimensional case. For example, in the 1D case, these data structures are set as follows:

  • cells(1:2,i)=[verts(i),verts(i+1)] contains the one-dimensional coordinates of the nodes defining the i-th integration cell, which is delimited by cells(1,i)=verts(i) and cell(2,i)=verts(i+1);

  • patches(1:2,i), contains the one-dimensional coordinates of the nodes defining the i-th patch, which is delimited by patches(1,i) and patches(2,i).

The auxiliary data structure verts is created to store the position vector (coordinates) of the nodes defining the integration cells. In the multi-dimensional case, these data structures are built direction by direction.

A.2 Integer Topological Data Structures

The integer topological data structures are used to describe the mesh connectivity. In the one-dimensional case, we consider the following data structures.

  • ib contains the global indexing of the patch’s basis functions (including polynomial and nonpolynomial basis functions). For example, ib(i,e) stores the global index of the i-th basis function (including polynomial and nonpolynomial basis functions) of the e-th patch, which is determined by looping on i=1,penr+1+npenr and e=1,Np.

  • cells2patches contains the indices of the patches that may overlap inside the integration cells. For example, cells2patches(i,e) contains the indices of all the patches seen by the e-th integration cell; note that the first dimension (index i) is 2, so

    • if the integration cell coincides with the flat-top region, then only i = 1 is set, and this is the global index of the patch that contains the integration cells;

    • if the integration cell is the overlap of two patches, then cells2patches(1,e) and cells2patches(2,e) are the two indices of the patches that contain cell e.

    Moreover, we distinguish between the periodic and nonperiodic boundary conditions, which corresponds to the allocation
    • cells2patches(dims,nconnect,2*Np-1+2) in the periodic case;

    • cells2patches(dims,nconnect,2*Np-1) in the nonperiodic case,

    where nconnect is equal to 2 and is the maximum number of patches that can overlap inside a given integration cell.

  • patch_idx(i,e) corresponds to cells2patches(i,e); this returns the kind of patch. In particular, this information selects the subset of basis functions in the single patch or the two overlapping patches seen by the integration cell that we consider to compute the local stiffness and mass matrix.

A graphical representation of the correspondence between patches and integration cells is provided by Figures 4 and 5. Figures 6 and 7 show the cell/patch partitions of a one-dimensional domain with nonperiodic and periodic boundary conditions using four logical cells.

Figure 4 
                     1D structure of the overlapping patches (midline), logical cells (the top line) and integration cells (bottom line).
Top labels 
                           
                              
                                 
                                    p
                                    
                                       i
                                       −
                                       2
                                    
                                 
                              
                              
                              p_{i-2}
                           
                        , 
                           
                              
                                 
                                    p
                                    
                                       i
                                       −
                                       1
                                    
                                 
                              
                              
                              p_{i-1}
                           
                        , 
                           
                              
                                 
                                    p
                                    i
                                 
                              
                              
                              p_{i}
                           
                        , 
                           
                              
                                 
                                    p
                                    
                                       i
                                       +
                                       1
                                    
                                 
                              
                              
                              p_{i+1}
                           
                         enumerate patches.
Each 
                           
                              
                                 
                                    p
                                    
                                       1
                                       :
                                       
                                          2
                                          ,
                                          i
                                       
                                    
                                 
                              
                              
                              p_{1:2,i}
                           
                         corresponds to the first and second vertex of
patch 𝑖, whose coordinates are stored in the array
vert(i).
Figure 4

1D structure of the overlapping patches (midline), logical cells (the top line) and integration cells (bottom line). Top labels p i 2 , p i 1 , p i , p i + 1 enumerate patches. Each p 1 : 2 , i corresponds to the first and second vertex of patch 𝑖, whose coordinates are stored in the array vert(i).

Figure 5 
                     1D structure of the integration cells (bottom line), logical cells (the top line) and overlapping patches (midline).
The upper indices 
                           
                              
                                 
                                    i
                                    −
                                    2
                                 
                              
                              
                              i-2
                           
                        , 
                           
                              
                                 
                                    i
                                    −
                                    1
                                 
                              
                              
                              i-1
                           
                        , 𝑖, 
                           
                              
                                 
                                    i
                                    +
                                    1
                                 
                              
                              
                              i+1
                           
                        , 
                           
                              
                                 
                                    i
                                    +
                                    2
                                 
                              
                              
                              i+2
                           
                         over the bottom line enumerate the integration cells.
The dots labeled by 
                           
                              
                                 
                                    v
                                    i
                                 
                              
                              
                              v_{i}
                           
                         (below the bottom line) are the vertices defining the integration cells, whose coordinates are stored in
vert(i).
Figure 5

1D structure of the integration cells (bottom line), logical cells (the top line) and overlapping patches (midline). The upper indices i 2 , i 1 , 𝑖, i + 1 , i + 2 over the bottom line enumerate the integration cells. The dots labeled by v i (below the bottom line) are the vertices defining the integration cells, whose coordinates are stored in vert(i).

Figure 6 
                     1D structure of a four-cell partitioning with nonperiodic boundaries: logical cells on the top line, overlapping patches on the midline, integration cells on the bottom line.
Figure 6

1D structure of a four-cell partitioning with nonperiodic boundaries: logical cells on the top line, overlapping patches on the midline, integration cells on the bottom line.

Figure 7 
                     1D structure of a four-cell partitioning with periodic boundaries: logical cells on the top line, overlapping patches on the midline, integration cells on the bottom line.
Figure 7

1D structure of a four-cell partitioning with periodic boundaries: logical cells on the top line, overlapping patches on the midline, integration cells on the bottom line.

In the two- and three-dimensional cases, we suitably modify the definitions of ib and cells2patches given above for the one-dimensional case as follows.

  • In 2D, we use ib(i,j,e1,e2), cells2patches(i,j,e1,e2) instead of ib(i,e), cells2patches(i,e). Herein, the index pair (e1,e2) addresses the two-dimensional patch or cell corresponding to the one-dimensional partitions labeled by e1=1:Np(1) and e2=1:Np(2). According to this definition, ib(i,j,e1,e2) is the number of the (i,j)-th basis functions for i , j = 1 , , penr + 1 + npenr of the (e1,e2)-th patch.

  • in 3D, we use ib(i,j,k,e1,e2,e3), cells2patches(i,j,k,e1,e2,e3), where the index triple (e1,e2,e3) addresses the two-dimensional patch or cell corresponding to the one-dimensional partitions addressed by e1=1:Np(1), e2=1:Np(2) and e3=1:Np(3). According to this definition, ib(i,j,k,e1,e2,e3) is the number of the (i,j,k)-th basis functions for i , j , k = 1 , , penr + 1 + npenr of the (e1,e2,e3)-th patch.

The data structure patch_idx is defined accordingly.

The content of these data structures differs depending on the boundary conditions, in particular depending on the periodic and the nonperiodic case. For example, in the two-dimensional case, we note that

  • the allocation ib of the global indexing of patch’s basis functions (including polynomial and nonpolynomial basis functions) is the same for the periodic and nonperiodic cases, and it is allocated as ib(penr+1+npenr,penr+1+npenr,Np(1),Np(2)).

  • cells2patches contains the indices of the patches that may overlap inside the integration cells. The allocation is

    • cells2patches(dims,nconnect,2*Np(1)-1+2,2*Np(2)-1+2) in the periodic case;

    • cells2patches(dims,nconnect,2*Np(1)-1,2*Np(2)-1) in the nonperiodic case,

    where dims is the number of dimensions and nconnect is the maximum number of patches that can share an integration cell. Note that dims = 2 , nconnect = 2 dims = 2 2 = 4 , the last two indices range through the number of integration cells in the two directions, and the integration cell is indexed by (e1,e2) ranging through
    • e1 = 1 , , 2 N p ( 1 ) 1 + 2 , e2 = 1 , , 2 N p ( 2 ) 1 + 2 in the periodic case;

    • e1 = 1 , , 2 N p ( 1 ) 1 , e2 = 1 , , 2 N p ( 2 ) 1 in the nonperiodic case.

    In summary,
    • if the integration cell labeled by (e1,e2) is within the flat-top region of a given patch, then only the global index of that patch is stored;

    • if the integration cell is shared by two or four overlapping patches, then their indices are stored.

  • patch_idx corresponds to cells2patches; this returns the kind of patch. In particular, we distinguish the case the integration cell coincides with the flat-top region of a patch from the case the integration cell is given by the intersection of two or four distinct patches.

An example of the cells-to-patches enumeration for a 4 × 4 partitioning of the domain Ω = [ 0 , 1 ] × [ 0 , 1 ] is given in Figures 8 (nonperiodic case) and 9 (periodic case). The generalization to the three-dimensional case is straightforward.

Figure 8 
                     Nonperiodic 2D case: the left panel shows the logical cells (dashed lines) and the integration cells.
The numbering is on the patch.
One patch is outlined at the right-most part of the panel, showing its integration cells and the numbers of the various patches overlapping inside them.
Figure 8

Nonperiodic 2D case: the left panel shows the logical cells (dashed lines) and the integration cells. The numbering is on the patch. One patch is outlined at the right-most part of the panel, showing its integration cells and the numbers of the various patches overlapping inside them.

Figure 9 
                     Periodic 2D case: the left panel shows the logical cells (dashed lines) and the integration cells.
The numbering is on the patch.
One patch is outlined at the right-most part of the panel, showing its integration cells and the numbers of the various patches overlapping inside them.
The black lines on the top and left side represent the 1D integration cells, identified here by the subsegments delimited by the dots.
Note that the integration cells have a different numbering but they are still numbered by the same indices.
Figure 9

Periodic 2D case: the left panel shows the logical cells (dashed lines) and the integration cells. The numbering is on the patch. One patch is outlined at the right-most part of the panel, showing its integration cells and the numbers of the various patches overlapping inside them. The black lines on the top and left side represent the 1D integration cells, identified here by the subsegments delimited by the dots. Note that the integration cells have a different numbering but they are still numbered by the same indices.

Acknowledgements

The first, third, and last authors are members of the Gruppo Nazionale Calcolo Scientifico-Istituto Nazionale di Alta Matematica (GNCS-INdAM). The third author acknowledges the contribution of the National Recovery and Resilience Plan, Mission 4 Component 2 – Investment 1.4 – National center for Hpc, big data and quantum computing, spoke 6.

References

[1] S. Agmon, Lectures on Elliptic Boundary Value Problems, Van Nostrand Math. Stud. 2, D. Van Nostrand Co., Princeton, 1965. Search in Google Scholar

[2] C. Albrecht, C. Klaar, J. E. Pask, M. A. Schweitzer, N. Sukumar and A. Ziegenhagel, Orbital-enriched flat-top partition of unity method for the Schrödinger eigenproblem, Comput. Methods Appl. Mech. Engrg. 342 (2018), 224–239. 10.1016/j.cma.2018.07.042Search in Google Scholar

[3] N. W. Ashcroft and N. D. Mermin, Solid State Physics, Thomson Learning, Toronto, 1976. Search in Google Scholar

[4] I. Babuška, G. Caloz and J. E. Osborn, Special finite element methods for a class of second order elliptic problems with rough coefficients, SIAM J. Numer. Anal. 31 (1994), no. 4, 945–981. 10.1137/0731051Search in Google Scholar

[5] I. Babuška and J. M. Melenk, The partition of unity method, Internat. J. Numer. Methods Engrg. 40 (1997), no. 4, 727–758. 10.1002/(SICI)1097-0207(19970228)40:4<727::AID-NME86>3.3.CO;2-ESearch in Google Scholar

[6] I. Babuška and J. Osborn, Eigenvalue problems, Handbook of Numerical Analysis, Vol. II, North-Holland, Amsterdam (1991), 641–787. 10.1016/S1570-8659(05)80042-0Search in Google Scholar

[7] R. F. W. Bader, T. T. Nguyen-Dang and Y. Tal, A topological theory of molecular structure, Rep. Progr. Phys. 44 (1981), no. 8, 893–948. 10.1088/0034-4885/44/8/002Search in Google Scholar

[8] A. S. Banerjee, L. Lin, W. Hu, C. Yang and J. E. Pask, Chebyshev polynomial filtered subspace iteration in the discontinuous Galerkin method for large-scale electronic structure calculations, Chem. Phys. 145 (2016), no. 15, Article ID 154101. 10.1063/1.4964861Search in Google Scholar

[9] D. Boffi, Finite element approximation of eigenvalue problems, Acta Numer. 19 (2010), 1–120. 10.1017/S0962492910000012Search in Google Scholar

[10] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods, 3rd ed., Texts Appl. Math. 15, Springer, New York, 2008. 10.1007/978-0-387-75934-0Search in Google Scholar

[11] J. S. Chen, W. Hu and M. Puso., Orbital HP-Clouds for solving Schrödinger equation in quantum mechanics, Comput. Methods Appl. Mech. Engrg. 196 (2007), 3693–3705. 10.1016/j.cma.2006.10.030Search in Google Scholar

[12] D. Davydov, T. Gerasimov, J.-P. Pelteret and P. Steinmann, Convergence study of the h-adaptive PUM and the hp-adaptive FEM applied to eigenvalue problems in quantum mechanics, Adv. Model. Simul. Eng. 4 (2017), Paper No. 7. 10.1186/s40323-017-0093-0Search in Google Scholar

[13] T. Dupont and R. Scott, Polynomial approximation of functions in Sobolev spaces, Math. Comp. 34 (1980), no. 150, 441–463. 10.1090/S0025-5718-1980-0559195-7Search in Google Scholar

[14] T.-P. Fries and T. Belytschko, The extended/generalized finite element method: an overview of the method and its applications, Internat. J. Numer. Methods Engrg. 84 (2010), no. 3, 253–304. 10.1002/nme.2914Search in Google Scholar

[15] P. Grisvard, Singularities in boundary value problems and exact controllability of hyperbolic systems, Optimization, Optimal Control and Partial Differential Equations (Iaşi 1992), Internat. Ser. Numer. Math. 107, Birkhäuser, Basel (1992), 77–84. 10.1007/978-3-0348-8625-3_8Search in Google Scholar

[16] E. K. U. Gross and R. M. Dreizler, Density Functional Theory, NATO Sci. Ser. 337, Springer, New York, 2013. Search in Google Scholar

[17] P. Hohenberg and W. Kohn, Inhomogeneous electron gas, Phys. Rev. (2) 136 (1964), 864–871. 10.1103/PhysRev.136.B864Search in Google Scholar

[18] B. Kanungo and V. Gavini, Large-scale all-electron density functional theory calculations using an enriched finite-element basis, Phys. Rev. B 95 (2017), Article ID 035112. 10.1103/PhysRevB.95.035112Search in Google Scholar

[19] T. Kato, Perturbation Theory for Linear Operators, 2nd ed., Grundlehren Math. Wiss. 132, Springer, Berlin, 1976. Search in Google Scholar

[20] W. Kohn and L. J. Sham, Self-consistent equations including exchange and correlation effects, Phys. Rev. (2) 140 (1965), 1133–1138. 10.1103/PhysRev.140.A1133Search in Google Scholar

[21] M. Levy, Universal variational functionals of electron densities, first-order density matrices, and natural spin-orbitals and solution of the 𝑣-representability problem, Proc. Natl. Acad. Sci. USA 76 (1979), no. 12, 6062–6065. 10.1073/pnas.76.12.6062Search in Google Scholar PubMed PubMed Central

[22] L. Lin, J. Lu, L. Ying and W. E., Adaptive local basis set for Kohn–Sham density functional theory in a discontinuous Galerkin framework I: Total energy calculation, J. Comput. Phys. 231 (2012), 2140–2154. 10.1016/j.jcp.2011.11.032Search in Google Scholar

[23] J. M. Melenk and I. Babuška, The partition of unity finite element method: basic theory and applications, Comput. Methods Appl. Mech. Engrg. 139 (1996), no. 1–4, 289–314. 10.1016/S0045-7825(96)01087-0Search in Google Scholar

[24] A. Messiah, Quantum Mechanics, Dover, New York, 2014. Search in Google Scholar

[25] R. G. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules, Int. Ser. Monogr. Chem. 16, Oxford University, Oxford, 1994. Search in Google Scholar

[26] J. E. Pask and N. Sukumar, Partition of unity finite element method for quantum mechanical materials calculations, Extreme Mech. Lett. 11 (2017), 8–17. 10.1016/j.eml.2016.11.003Search in Google Scholar

[27] J. E. Pask, N. Sukumar and S. E. Mousavi, Linear scaling solution of the all-electron Coulomb problem in solids, Int. J. Multiscale Comput. Eng. 10 (2012), 83–99. 10.1615/IntJMultCompEng.2011002201Search in Google Scholar

[28] E. Schrödinger, Undulatory theory of the mechanics of atoms and molecules, Phys. Revi. 28 (1926), 1049–1070. 10.1103/PhysRev.28.1049Search in Google Scholar

[29] M. A. Schweitzer, Stable enrichment and local preconditioning in the particle-partition of unity method, Numer. Math. 118 (2011), no. 1, 137–170. 10.1007/s00211-010-0323-6Search in Google Scholar

[30] M. A. Schweitzer, Generalizations of the finite element method, Cent. Eur. J. Math. 10 (2012), no. 1, 3–24. 10.2478/s11533-011-0112-1Search in Google Scholar

[31] M. A. Schweitzer, Variational mass lumping in the partition of unity method, SIAM J. Sci. Comput. 35 (2013), no. 2, A1073–A1097. 10.1137/120895561Search in Google Scholar

[32] N. Sukumar and J. E. Pask, Classical and enriched finite element formulations for Bloch-periodic boundary conditions, Internat. J. Numer. Methods Engrg. 77 (2009), no. 8, 1121–1138. 10.1002/nme.2457Search in Google Scholar

[33] S. Yamakawa and S.-A. Hyodo, Electronic state calculation of hydrogen in metal clusters based on Gaussian-FEM mixed basis function, J. Alloys Compd. 356–357 (2003), 231–235. 10.1016/S0925-8388(03)00353-0Search in Google Scholar

[34] S. Yamakawa and S.-A. Hyodo, Gaussian finite-element mixed-basis method for electronic structure calculations, Phys. Rev. B 71 (2005), Article ID 035113. 10.1103/PhysRevB.71.035113Search in Google Scholar

[35] W. Yang and P. W. Ayers, Density-functional theory, Computational Medicinal Chemistry for Drug Discovery, CRC Press, Boca Raton (2003), 103–132. 10.1201/9780203913390.ch4Search in Google Scholar

[36] G. Zhang, L. Lin, W. Hu, C. Yang and J. E. Pask, Adaptive local basis set for Kohn–Sham density functional theory in a discontinuous Galerkin framework II: Force, vibration, and molecular dynamics calculations, J. Comput. Phys. 335 (2017), 426–443. 10.1016/j.jcp.2016.12.052Search in Google Scholar

Received: 2023-04-22
Revised: 2024-03-05
Accepted: 2024-04-16
Published Online: 2024-05-16
Published in Print: 2025-01-01

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/cmam-2023-0115/html
Scroll to top button