Accelerating Eigenvalue Computation for
Nuclear Structure Calculations via Perturbative Corrections
Abstract
We present a new method for computing the lowest few eigenvalues and the corresponding eigenvectors of a nuclear many-body Hamiltonian represented in a truncated configuration interaction subspace, i.e., the no-core shell model (NCSM). The method uses the hierarchical structure of the NCSM Hamiltonian to partition the Hamiltonian as the sum of two matrices. The first matrix corresponds to the Hamiltonian represented in a small configuration space, whereas the second is viewed as the perturbation to the first matrix. Eigenvalues and eigenvectors of the first matrix can be computed efficiently. Perturbative corrections to the eigenvectors of the first matrix can be obtained from the solutions of a sequence of linear systems of equations defined in the small configuration space. These correction vectors can be combined with the approximate eigenvectors of the first matrix to construct a subspace from which more accurate approximations of the desired eigenpairs can be obtained. We call this method a Subspace Projection with Perturbative Corrections (SPPC) method. We show by numerical examples that the SPPC method can be more efficient than conventional iterative methods for solving large-scale eigenvalue problems such as the Lanczos, block Lanczos and the locally optimal block preconditioned conjugate gradient (LOBPCG) method. The method can also be combined with other methods to avoid convergence stagnation.
I Introduction
Nuclear structure calculations require solving -body Schrödinger equations where is the number of nucleons consisting of protons and neutrons. The Configuration Interaction (CI) method or no-core shell method (NCSM) [1], which represents the solution to the Schrödinger’s equation by a linear combination of -body basis functions, reduces the problem to an algebraic eigenvalue problem:
(1) |
where is the matrix representation of the -body nuclear Hamiltonian operator in a configuration space (spanned by a set of -body basis functions), is the th eigenvalue of representing an approximation to a discrete energy level, and is the corresponding eigenvector that contains the coefficients of the -body basis function in the expansion of the approximate eigenfunction in the -body basis.
The dimension () of the matrix depends on the number of nucleons and the size of the CI model space (determined by a truncation parameter ). Although can be quite large, is very sparse, and often only a few of its eigenpairs at the low end of the spectrum are of interest, making iterative methods suitable for solving (1).
The construction of the matrix by the CI method is typically done in a hierarchical fashion where the leading submatrix of corresponds to a matrix constructed from a smaller configuration space. Because the -body basis functions that form the lower dimension CI space associated with a small are typically more important than basis functions outside of such a configuration space, the eigenvectors of tend to be localized; i.e., the leading components of the tend to be the larger in magnitude, while the tailing components are relatively small in magnitude. We illustrate these properties of a nuclear Hamiltonian and its wavefunction using nucleus 12C as an example. The Hamiltonian is constructed with a nucleon-nucleon interaction Daejeon16 [2] where hbar-omega value of describes the harmonic oscillator basis functions. Figure 1(a) shows that the leading submatrix of that is times smaller corresponds to a matrix constructed from a smaller configuration space. Figure 1(b) shows that the eigenvector corresponding to the lowest eigenvalue is localized in its first few components. Therefore, the vector formed by padding the eigenvector of the leading submatrix with zeros can serve as a good initial guess for many iterative methods used to solve large-scale eigenvalue problems (1). Previous works [3, 4] select such initial guesses for algorithms like the Lanczos algorithm [5], the block Lanczos algorithm [6], the Locally Optimal Block Preconditioned Conjugate Gradient (LOBPCG) algorithm [7], and the Residual Minimization Method with Direct Inversion of Iterative Subspace (RMM-DIIS) correction [8, 9, 10]. Another study [11] uses greedy algorithms to incrementally enlarge the submatrix and use the eigenvector of the enlarged submatrix as an improved starting guess.


In this paper, we propose solving the eigenvalue problem (1) with an iterative subspace projection method constructed from basis vectors obtained from eigenvalue and eigenvector perturbation theory. We call this method a Subspace Projection with Perturbative Corrections (SPPC) method. In this approach, the Hamiltonian matrix to be partially diagonalized is viewed as the sum of two matrices, i.e., , where the eigenpairs of are relatively easy and inexpensive to compute, and is viewed as a perturbation to . Perturbative corrections to eigenpairs of in successively higher order can be computed by solving a sequence of linear systems of equations. Together with the initial approximation to the desired eigenvector obtained from , these correction vectors form a low dimensional subspace from which approximate eigenpairs of are extracted through the Rayleigh-Ritz procedure. When is chosen to be a block diagonal matrix diag(,0), where is the matrix representation of the -body nuclear Hamiltonian in a smaller configuration space associated with a smaller value, these linear systems can be solved efficiently in the small configuration space. The overall computational cost of the SPPC method grows linearly with respect to the highest order of perturbation included in the correction subspace. Adding each perturbative correction vector and performing the Rayleigh-Ritz calculation requires multiplying the sparse matrix with a single vector. We show numerically that the low dimensional subspace constructed in a low order SPPC method provides more accurate approximation to a few lowest eigenvalues of than a Krylov subspace of the same dimension constructed from the same starting guess. The method also appears to be more efficient than the locally optimal block preconditioned conjugate gradient (LOBPCG) method in early iterations even when a good preconditioner is available for the LOBPCG method. Although the SPPC method can stagnate as higher perturbative corrections are included, convergence stagnation can be mitigated by combining SPPC with other iterative algorithms for solving large scale eigenvalue problems using the SPPC’s eigenvector approximation as the starting guess for secondary algorithms.
The rest of this paper is organized as follows. In Section II we describe the basic Subspace Projection with Perturbative Corrections (SPPC) method for computing one eigenpair of , and show how successively higher order perturbative corrections can be obtained from solutions of a set of linear systems of equations. We draw the connection between SPPC and previously developed eigenvector continuation methods. In Section III, we present a version of the SPPC algorithm that can be used to compute a few eigenpairs. A few practical implementation details of the SPPC method are discussed in Section IV. Numerical examples that demonstrate the efficiency of the SPPC method relative to other conventional large-scale eigenvalue computation methods are presented in Section V. We also show the effectiveness of combining SPPC with a conventional eigensolver for computing a few lowest eigenpairs of several nuclei.
II Subspace Projection with Perturbative Corrections (SPPC)
We split the matrix into the sum of two matrices and , i.e.,
(2) |
where
(3) |
with the matrix (where ) being the leading submatrix of that corresponds to the representation of the Hamiltonian within a smaller configuration space.
The eigenvectors of , which can be obtained from the eigenvectors of , are computed with a much lower computational cost compared to those of the full matrix . They can be used as good initial guesses for conventional algorithms, such as the Lanczos and the LOBPCG algorithms, for computing the eigenpairs of .
The SPPC method uses the eigenvectors of to initiate a subspace construction procedure to produce a subspace from which improved approximations of the desired eigenpairs can be obtained.
Instead of using the Lanczos algorithm or the LOBPCG method to construct the subspace, we use perturbative corrections to the initial eigenvector approximation to construct the subspace. In this approach, we view as a matrix obtained from perturbing by with , i.e., we can write as
(4) |
It follows from the Rayleigh-Schrödinger perturbation theory [12] that the eigenvalues and eigenvectors of , which we denote by and respectively, can be written in terms of perturbative corrections to the eigenvalues and eigenvectors of , which we denote by and , i.e.,
(5) |
Here, represent the th order perturbative corrections to the eigenpair , and are independent of the parameter .
Substituting (5) into the equation
(6) |
and matching coefficients of the same degree order yields the following set of equations
(7) |
that allow us to compute and in a recursive fashion.
The asymptotic expansion used in (5) assumes that is a small parameter. As a result, the expansion serves as a good approximation to the desired eigenpair only when is sufficiently small, i.e., when falls within the radius of convergence for (5), which is generally much smaller than 1. As a result, (5) cannot be used directly in general to approximate the th eigenpair of [13, 14]. However, the perturbative vectors can be used to construct a subspace
(8) |
from which approximation to and can be obtained.
The idea of using perturbative corrections to construct an approximating subspace was proposed in [13, 15] in the context of an eigenvector continuation (EC) method [16, 17]. In an EC method, the eigenvectors of for some choices of ’s are used to construct a subspace from which approximations to the eigenvectors of for are obtained from the projection of into such a subspace.
It was found in [13, 15] that instead of using eigenvectors of for several choices of to construct a subspace from which approximate eigenvectors of are extracted through the standard Rayleigh-Ritz procedure, more accurate approximations to the desired eigenpairs of can be obtained from the subspace constructed from the eigenvector of as well as perturbative eigenvector corrections as discussed above.
In a Rayleigh-Ritz procedure, we compute an orthonormal basis matrix of and form the projected matrix
(9) |
If is an eigenpair of , then where , yields an approximate eigenpair of . We consider an approximate eigenpair to have converged if its relative residual norm
(10) |
is smaller than a preset tolerance value.
To obtain an approximate th eigenpair of , we choose as the lowest eigenpair of in the above mentioned Rayleigh-Ritz procedure. We present the SPPC method in Algorithm 1.
We should point out that the main distinction between the SPPC method proposed here and the EC method lies in the subspaces constructed in these methods and the cost of construction. Because EC projects onto a subspace constructed from the eigenvectors of for several (nonzero) ’s, the cost of subspace construction may be just as expensive as the cost of solving the target eigenvalue problem with a particular choice of . On the other hand, because the subspace constructed in the SPPC method uses perturbative corrections that can be obtained by solving much smaller linear systems, the cost of subspace construction is significantly lower.
Although the basic idea of SPPC was presented in [13, 15], the computational cost of this method was not carefully analyzed and compared with those of the state-of-the-art large-scale eigensolvers. In [13], the eigenvectors of can be computed analytically for the one-dimensional quartic anharmonic oscillator. However, in general, identifying a that can be diagonalized analytically is not possible. In [15], SPPC is used to perform a -body nuclear structure calculation. However, a different splitting scheme is used, and the eigenvectors of are not easier to compute than those of .
III Targeting First Few Eigenpairs
Although we can use Algorithm 1 to compute each of the first eigenpairs one by one (or in parallel), approximations to larger eigenvalues and the corresponding eigenvectors appear to converge slowly, and sometimes to wrong values, as we will show in section VI. A more effective way to obtain approximations to the first eigenpairs is to combine the SPPC subspace (8) constructed for each eigenpair to create a larger subspace
(11) |
from which approximate eigenpairs can be extracted simultaneously through the Rayleigh-Ritz procedure, by targeting the lowest eigenpairs of the projected matrix. Algorithm 2 shows how this approach works.
IV Practical considerations
In this section, we describe a few practical implementation details of the SPPC algorithm.
IV.1 Computing the Correction Vectors
We refer to the th eigenvector of denoted by as the zero-th order correction to the th eigenvector of .
When is of the form given in (3), can be obtained by computing the th eigenvector of , denoted by , and appending it with zeros to yield
(12) |
It follows from (7) and the block structures of , , and that the first order corrections to the th eigenvalue and eigenvector of are
(13) |
It is easy to verify that
i.e., the first order correction to the eigenvector is simply the residual associated with the zero-th order approximation.
Because , we can simplify the linear system in (7) to
(14) |
The matrix is in a block diagonal form
(15) |
consisting of two blocks where the first block is relatively small and the second block is a scalar multiple of the identity matrix. As a result, the solution of the linear system in (14) essentially reduces to the solution of a much smaller linear system with being the coefficient matrix. If we partition conformally with the blocks in (15) as and the right-hand side of (14) as such that and , can be easily computed as
(16) |
and can be obtained by solving
(17) |
Note that equation (14) is singular because is an eigenvalue of . However, since already includes and we are only interested in contributions in the orthogonal complement of from the solution of (14), we can project out from the right-hand side of (14) before solving this equation. This is equivalent to projecting out from the right-hand side of (17), i.e. we solve
(18) |
IV.2 Rayleigh-Ritz Calculation
After obtaining the correction vectors, we generate an orthonormal basis matrix of the subspace spanned by these vectors and then perform the Rayleigh-Ritz procedure. We lay out these steps for approximating a single eigenpair and the first few eigenpairs.
Targeting the th eigenpair.
We use the Gram-Schmidt process to obtain an orthonormal basis of the subspace . The orthonormal basis forms the columns of the matrix . The th column, denoted by , is generated as follows.
(19) |
We append to such that
(20) |
Note that the projected matrix can be constructed recursively. Assuming has been computed in the previous step, we just need to compute and append an additional row and column to as shown below.
(21) |
Therefore, the major cost for constructing the projected matrix in each step of the SPPC method is in performing a single SpMV in .
Targeting the first few eigenpairs.
To obtain an orthonormal basis for the combined subspace , we replace the Gram-Schmidt process (19) with the block Gram-Schmidt process. If , the block Gram-Schmidt procedure yields
(22) |
We then perform a QR factorization of , i.e.,
(23) |
to generate an orthonormal basis for . We append to such that
(24) |
is an orthonormal basis for .
Again, the projected matrix can be constructed recursively, and the computational cost is dominated by the cost for computing SpMVs in .
IV.3 Computational Cost
We now discuss the overall computational cost of the SPPC method. We can see from Algorithm 1 and Algorithm 2 that the two major components of the SPPC algorithm are: (1) Solving the linear system (14); (2) forming the projected matrix . We have already shown that the projected matrix can be computed recursively using SpMVs in each step of the SPPC algorithm. The reduced linear system (17) of the correction equation (14) can be solved iteratively using, for example, the MINRES algorithm. Because it has a much smaller dimension, the SpMVs performed in each MINRES iteration are relatively cheap. However, forming the right-hand side of the equation (14) requires multiplying with for which has nearly the same complexity as multiplying with . Therefore, it may appear that each SPPC step requires performing SpMVs. We will show below that this is not the case. Both the right-hand side of (14) and the projected matrix can be obtained from the same product. As a result, each step of the SPPC algorithm only requires performing SpMVs.
Using the matrix splitting , we can rewrite as
(25) |
where . Therefore, can be obtained by subtracting , a much lower computational cost, from .
We now show that can be easily obtained from . It follows from (22) and (23) that
(26) |
As a result, we can obtain from by using following identity
(27) |
Note that the analysis of the computational cost assumes , which contains as its columns for , has been stored in memory.
As we will show in section VI, the highest order perturbation is often limited to 15, beyond which no significant improvement in the approximate eigenpair can be observed. Therefore, the dense linear algebra operations such as computing and diagonalizing the projected matrix can be performed with a relatively low cost compared to the cost of multiplying with .
V Combining SPPC with other algorithms
As we will show in the next section, the perturbative correction is typically effective when the order of perturbation is relatively low. The convergence of SPPC can stagnate when increases, i.e., adding higher order perturbative correction may not help because they may be linearly dependent with respect to the basis vectors included in already. In this case, it is useful to combine SPPC with another algorithm that can take the eigenvector approximation produced by SPPC as the starting guess.
Algorithms that can be combined with the SPPC method in a hybrid algorithm include but are not limited to the following:
-
•
The Lanczos algorithm, which is a classical algorithm that generates an orthonormal basis of a Krylov subspace using the Gram-Schmidt procedure. It is initialized with an approximate eigenvector produced from SPPC or a linear combination of approximate eigenvectors. It uses one SpMV per iteration to compute the next basis vector. The eigenpairs are approximated using Ritz pairs obtained from the Rayleigh-Ritz procedure with the basis vectors of the Krylov subspace.
-
•
The block Lanczos algorithm, which is a variation of the Lanczos algorithm that operates in blocks. It is initialized with a block of vectors approximating several eigenvectors of , and builds a Krylov subspace in blocks, with each iteration performing SpMVs. The main advantage of the block Lanczos algorithm over the Lanczos algorithm is that it can make use of approximations to several eigenvectors more effectively and most dense linear algebra operations can take advantage of level 3 BLAS.
-
•
The LOBPCG algorithm, which is an iterative method that solves the equivalent trace minimization formulation of the eigenvalue problem. Similar to the block Lanczos algorithm, it can be initialized with approximations to several eigenvectors and each iteration performs SpMVs. One advantage of the LOBPCG algorithm is that it can utilize a preconditioner if it is available. The use of a preconditioner can accelerate convergence. A common choice for the preconditioner is a block diagonal part of the Hamiltonian matrix or a shifted matrix of the Hamiltonian for some appropriately chosen shift [7, 4]. For our numerical experiments, we use a shifted preconditioner that involves a specific block diagonal part of the Hamiltonian matrix and a constant shift that approximates the lowest eigenvalue.
-
•
The residual minimization method (RMM) combined with the Direct Inversion of Iterative Subspace (DIIS) refinement algorithm (RMM-DIIS), which is a quasi-Newton algorithm for improving a specific eigenpair without computing other eigenpairs, provided that the initial approximation to the desired eigenpairs is sufficiently accurate. Each RMM-DIIS iteration performs one SpMV. The RMM-DIIS can also incorporate a preconditioner to accelerate convergence. For our numerical experiments, we choose a preconditioner that is a specific diagonal part of the Hamiltonian matrix.
VI Numerical Results
In this section, we demonstrate the effectiveness of SPPC algorithm and compare it with other existing algorithms for computing the ground and a few low excited states of several light nuclei. We also show that SPPC can be effectively combined with RMM-DIIS to yield an efficient and accurate hybrid eigensolver for nuclear configuration interaction calculations. We call this hybrid eigensolver the SPPC+RMM-DIIS method.
For all algorithms tested in this section, we use the relative residual norm (10) as the stopping criteria and set the convergence tolerance to be . To ensure a fair comparison with the SPPC, we use the eigenvectors of the zero-order part as the initial guesses for the algorithms. All experiments were conducted using MATLAB.
VI.1 Test Matrices
We use the -body Hamiltonian matrices corresponding to the nuclei 6Li, 7Li, 11B, and 12C in the following numerical experiments. The superscripts indicate the number of nucleons in the nuclei; for example, 7Li indicates Lithium with 3 protons plus 4 neutrons. The Hamiltonian matrices are constructed in a truncated CI space defined by a truncation parameter , using the nucleon-nucleon interaction Daejeon16. For the same nucleus, a larger results in a larger matrix , but the size of the matrix is independent of the interaction. Note that with three-nucleon interactions, the number of nonzero matrix elements is an order of magnitude larger than the one for the nucleon-nucleon interactions, and the number of iterations and the actual eigenvalues of any eigensolver will be different. As we indicated earlier, the construction of can be done in a hierarchical fashion so that a leading submatrix of corresponds to the same nuclear -body Hamiltonian represented in a lower dimensional configuration space associated with a smaller . In this section, if is the matrix representation of a nuclear -body Hamiltonian represented in a configuration space associated with , the submatrix corresponds to the representation of the same Hamiltonian in a configuration space associated with . We list the dimension of (denoted by dim()), the dimension of (denoted by dim()), the number of non-zero elements in (denoted by nnz()), (denoted by nnz()), and the value associated with in Table 1.
Nucleus | dim() | dim() | nnz() | nnz( | |
---|---|---|---|---|---|
6Li | 6 | 197,822 | 17,040 | 106,738,802 | 4,122,448 |
7Li | 6 | 663,527 | 48,917 | 421,938,629 | 14,664,723 |
11B | 4 | 814,092 | 16,097 | 389,033,682 | 2,977,735 |
12C | 4 | 1,118,926 | 17,725 | 555,151,572 | 3,365,009 |
VI.2 Targeting the Lowest Eigenpair
We report the performance of the SPPC, the Lanczos algorithm, the LOBPCG algorithm, and the SPPC+RMM-DIIS for targeting the lowest eigenpair of the matrix . As mentioned earlier, we use a preconditioner for the LOBPCG and the RMM-DIIS algorithms. The primary cost of all these algorithms is the number of SpMVs they perform before reaching convergence. Because each algorithm performs one SpMV per iteration, we can directly compare them by the number of iterations required to reach convergence.
The left plot of Figure 2 shows the convergence history of the algorithms chosen for comparison for 12C with respect to the iteration number. We observe that the SPPC algorithm converges in iterations, which is the least among all methods, while the Lanczos algorithm and the LOBPCG algorithm converge in iterations.


The result shown in the right plot of Figure 2 is for the Hamiltonian associated with 6Li. Several features of the SPPC algorithm are observed. The first observation is that the SPPC method converges more rapidly in the early iterations (up to the 15th iteration), and can be up to at most two orders of magnitude more accurate than other algorithms in these early iterations. However, the SPPC approximation appears to stagnate in subsequent iterations. This suggests that higher order correction vectors produced in later iterations (after iteration 15) do not contribute to improving the subspace constructed by the correction vectors produced in the early iterations. To verify this conjecture, we plot the angle between the current correction vector and the subspace spanned by the previous correction vectors, denoted by , with respect to the iteration number in Figure 3. We observe that is relatively large in the first few SPPC iterations, and gradually decreases to the level of . This is the point at which the new correction vector contributes minimally to the expansion of the subspace.

To overcome this stagnation, we consider a hybrid approach, the SPPC+RMM-DIIS, where we use the SPPC until the point of stagnation, and then switch to the RMM-DIIS. This hybrid approach takes advantage of the fast convergence of the SPPC for the first few iterations and the fast convergence of the RMM-DIIS when initialized with a good initial guess to the desired eigenvector. Specifically, we choose the initial guess in the RMM-DIIS as the Ritz vector produced from the SPPC method at the point of the switch. For 6Li, we use the RMM-DIIS after the th iteration of the SPPC. We observe that the SPPC+RMM-DIIS breaks the stagnation of the SPPC and converges in iterations, while the Lanczos algorithm and the LOBPCG algorithm converge in and iterations, respectively.
Table 2 gives a comparison of the SpMV counts used by several algorithms tested in this section for all four Hamiltonian matrices. With the exception of the Hamiltonian of the nucleus 12C, the SPPC stagnates around iteration number 15. For these cases, we also consider the SPPC+RMM-DIIS. We observe that this hybrid approach converges the fastest with the fewest SpMVs performed.
Nucleus | SPPC | Lanczos | LOBPCG | SPPC+RMM-DIIS |
---|---|---|---|---|
6Li | >30 | 31 | 25 | 23 |
7Li | >30 | 31 | 25 | 24 |
11B | >30 | 30 | 27 | 18 |
12C | 16 | 22 | 22 |
VI.3 Targeting the Five Lowest Eigenpairs
We can use either Algorithm 1 or Algorithm 2 to compute the lowest eigenvalues. In the left plots of Figure 4 and Figure 5, we show the relative residual norms of the approximation to the lowest eigenpairs of the 6Li Hamiltonian at each iteration of Algorithm 1 and Algorithm 2, respectively. The relative residual norms associated with the first three eigenpairs drop below by the th iteration. However, for the fourth and the fifth eigenpairs, the relative residuals obtained by Algorithm 1 jump at a certain point and never become small in subsequent iterations. The convergence failure for these eigenpairs can also be seen from the change of Ritz values at each iteration and how they compare with the true eigenvalues of the Hamiltonian as shown in the right plot of Figure 4. It appears that the 4th and 5th Ritz values move towards a different eigenvalue. This is due to a significant round off error in the orthogonalization process where the new basis contains contribution from the other eigenvector. In contrast, as shown in the right plot of Figure 5, by constructing a larger subspace consisting of perturbative corrections to several eigenvectors, Algorithm 2 computes approximate eigenvalues that do not deviate from the true eigenvalues, and as a result, all of its computed eigenpairs converge within a reasonable accuracy. Similar behaviors are observed for the other three Hamiltonians: 7Li, 11B, and 12C.




We now show that the stagnation in the convergence of the first Ritz pairs in the SPPC method can be eliminated by using a hybrid SPPC+RMM-DIIS. We switch to the RMM-DIIS method from the SPPC method when the relative residual norm of any of the eigenpairs starts to stagnate. In this hybrid approach, a separate RMM-DIIS run is used to refine each approximate eigenpair. It is initialized with the corresponding Ritz vector returned from the SPPC method at the point of the switch.
Figure 6 illustrates the convergence of the SPPC+RMM-DIIS for the 12C Hamiltonian. We choose to switch to the RMM-DIIS from the SPPC at iteration , which is a point of stagnation for most eigenpairs. We observe that the SPPC+RMM-DIIS breaks the stagnation of the SPPC and that all eigenpairs converge rapidly.

In Table 3, we present the SpMV counts of the block Lanczos algorithm, the LOBPCG algorithm, and the SPPC+RMM-DIIS for computing the lowest eigenpairs of the four Hamiltonians. The switch to the RMM-DIIS algorithm occurs for the SPPC+RMM-DIIS after the 15th iteration for all four Hamiltonian matrices. The block Lanczos algorithm and the LOBPCG algorithm are block methods, thus requiring continued iterations until every eigenpair converges and consequently a full SpMVs at each iteration. In contrast, the SPPC+RMM-DIIS targets each eigenpair individually after it switches to RMM-DIIS, so it does not incur more SpMVs for the converged eigenpair as it targets the non-converged eigenpairs. Due to this advantage, we observe that the SPPC+RMM-DIIS converges the fastest.
Nucleus | Block Lanczos | LOBPCG | SPPC+RMM-DIIS |
---|---|---|---|
6Li | 155 | 140 | 127 |
7Li | 175 | 150 | 129 |
11B | 170 | 125 | 94 |
12C | 155 | 165 | 90 |
VII Conclusion
In conclusion, the Subspace Projection with Perturbative Corrections (SPPC) method, combined with the Residual Minimization Method with Direct Inversion of Iterative Subspace (RMM-DIIS), presents a significant advancement in the efficient computation of eigenpairs for large Hamiltonian matrices in nuclear structure calculations. The SPPC method leverages perturbative correction vectors to enhance the accuracy of eigenpair approximations in the initial iterations, substantially reducing the number of sparse matrix-vector multiplications (SpMVs) required for convergence. Although the SPPC may experience stagnation in subsequent iterations, this challenge is effectively mitigated by integrating it with the RMM-DIIS algorithm, which provides robust refinement of eigenvector approximations. Our numerical experiments across several nuclear Hamiltonians demonstrate that the SPPC+RMM-DIIS hybrid approach outperforms traditional methods in terms of SpMVs. This hybrid method offers a promising solution for large-scale nuclear structure calculations, providing a reliable and efficient approach to solving the -body Schrödinger equation.
While the preliminary results of the SPPC are promising, we have not provided a theoretical background explaining why it works. We plan to address this in our future work, discussing the convergence behavior in detail. From a practical standpoint, w e aim to develop a method to automatically detect stagnation, eliminating the need for manual decisions on when to switch to the RMM-DIIS. Additionally, we are interested in implementing the SPPC in a hybrid MPI/OpenMPI code, such as the software MFDn (Many-Fermion Dynamics for nuclear structure) [18, 19, 20], to be run at high-performance computing centers. We also want to explore further optimizations and applications of the SPPC to other large-scale eigenvalue problems in nuclear physics and beyond.
Acknowledgements.
This material is based upon work supported by the Scientific Discovery through Advanced Computing (SciDAC) Program at the U.S. Department of Energy (DOE), Office of Science under Grants DE-SC0023495 and DE-SC0023175 from the Office of Nuclear Physics, and also under funding for the FASTMath Institute from the Office of Advanced Scientific Computing Research through Contract No. DE-AC02-05CH11231. This work used resources at the National Energy Research Scientific Computing Center (NERSC), which is funded by DOE under Contract No. DE-AC02-05CH11231. In addition, D.L. acknowledges partial support from DOE grants DE-SC0024586, DE-SC0023658, DE-SC0013365, and NSF grant PHY-2310620. J.P.V. acknowledges partial support from DOE grants DE-SC0023692 and DE-0023707.References
- Barrett et al. [2013] B. R. Barrett, P. Navrátil, and J. P. Vary, Ab initio no core shell model, Progress in Particle and Nuclear Physics 69, 131 (2013).
- Shirokov et al. [2016] A. Shirokov, I. Shin, Y. Kim, M. Sosonkina, P. Maris, and J. Vary, N3lo nn interaction adjusted to light nuclei in ab exitu approach, Physics Letters B 761, 87 (2016).
- Shao et al. [2018] M. Shao, H. M. Aktulga, C. Yang, E. G. Ng, P. Maris, and J. P. Vary, Accelerating nuclear configuration interaction calculations through a preconditioned block iterative eigensolver, Computer Physics Communications 222, 1 (2018).
- Alperen et al. [2023] A. Alperen, H. M. Aktulga, P. Maris, and C. Yang, Hybrid eigensolvers for nuclear configuration interaction calculations, Computer Physics Communications 292, 108888 (2023).
- Lanczos [1950] C. Lanczos, An iteration method for the solution of the eigenvalue problem of linear differential and integral operators, (1950).
- Golub and Underwood [1977] G. H. Golub and R. Underwood, The block lanczos method for computing eigenvalues, in Mathematical software (Elsevier, 1977) pp. 361–377.
- Knyazev [2001] A. V. Knyazev, Toward the optimal preconditioned eigensolver: Locally optimal block preconditioned conjugate gradient method, SIAM journal on scientific computing 23, 517 (2001).
- Wood and Zunger [1985] D. Wood and A. Zunger, A new method for diagonalising large matrices, Journal of Physics A: Mathematical and General 18, 1343 (1985).
- Jia and Stewart [2001] Z. Jia and G. Stewart, An analysis of the rayleigh–ritz method for approximating eigenspaces, Mathematics of computation 70, 637 (2001).
- Pulay [1980] P. Pulay, Convergence acceleration of iterative sequences. the case of scf iteration, Chemical Physics Letters 73, 393 (1980).
- Hernandez et al. [2021] T. M. Hernandez, R. Van Beeumen, M. A. Caprio, and C. Yang, A greedy algorithm for computing eigenvalues of a symmetric matrix with localized eigenvectors, Numerical Linear Algebra with Applications 28, e2341 (2021).
- Shavitt and Bartlett [2009] I. Shavitt and R. J. Bartlett, Many-body methods in chemistry and physics: MBPT and coupled-cluster theory (Cambridge university press, 2009).
- Franzke et al. [2022] M. C. Franzke, A. Tichai, K. Hebeler, and A. Schwenk, Excited states from eigenvector continuation: The anharmonic oscillator, Physics Letters B 830, 137101 (2022).
- Duguet et al. [2023] T. Duguet, A. Ekström, R. J. Furnstahl, S. König, and D. Lee, Eigenvector continuation and projection-based emulators, arXiv preprint arXiv:2310.19419 (2023).
- Demol et al. [2020] P. Demol, T. Duguet, A. Ekström, M. Frosini, K. Hebeler, S. König, D. Lee, A. Schwenk, V. Somà, and A. Tichai, Improved many-body expansions from eigenvector continuation, Physical Review C 101, 041302 (2020).
- Frame et al. [2018] D. Frame, R. He, I. Ipsen, D. Lee, D. Lee, and E. Rrapaj, Eigenvector continuation with subspace learning, Physical review letters 121, 032501 (2018).
- Sarkar and Lee [2021] A. Sarkar and D. Lee, Convergence of eigenvector continuation, Physical review letters 126, 032501 (2021).
- Sternberg et al. [2008] P. Sternberg, E. G. Ng, C. Yang, P. Maris, J. P. Vary, M. Sosonkina, and H. V. Le, Accelerating configuration interaction calculations for nuclear structure, in SC’08: Proceedings of the 2008 ACM/IEEE Conference on Supercomputing (IEEE, 2008) pp. 1–12.
- Maris et al. [2010] P. Maris, M. Sosonkina, J. P. Vary, E. Ng, and C. Yang, Scaling of ab-initio nuclear physics calculations on multicore computer architectures, Procedia Computer Science 1, 97 (2010).
- Aktulga et al. [2014] H. M. Aktulga, C. Yang, E. G. Ng, P. Maris, and J. P. Vary, Improving the scalability of a symmetric iterative eigensolver for multi-core platforms, Concurrency and Computation: Practice and Experience 26, 2631 (2014).