Skip to main content
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian
  • Log in
  • My Cart

Main menu

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Home
Home

Advanced Search

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses

New Research In

Physical Sciences

Featured Portals

  • Physics
  • Chemistry
  • Sustainability Science

Articles by Topic

  • Applied Mathematics
  • Applied Physical Sciences
  • Astronomy
  • Computer Sciences
  • Earth, Atmospheric, and Planetary Sciences
  • Engineering
  • Environmental Sciences
  • Mathematics
  • Statistics

Social Sciences

Featured Portals

  • Anthropology
  • Sustainability Science

Articles by Topic

  • Economic Sciences
  • Environmental Sciences
  • Political Sciences
  • Psychological and Cognitive Sciences
  • Social Sciences

Biological Sciences

Featured Portals

  • Sustainability Science

Articles by Topic

  • Agricultural Sciences
  • Anthropology
  • Applied Biological Sciences
  • Biochemistry
  • Biophysics and Computational Biology
  • Cell Biology
  • Developmental Biology
  • Ecology
  • Environmental Sciences
  • Evolution
  • Genetics
  • Immunology and Inflammation
  • Medical Sciences
  • Microbiology
  • Neuroscience
  • Pharmacology
  • Physiology
  • Plant Biology
  • Population Biology
  • Psychological and Cognitive Sciences
  • Sustainability Science
  • Systems Biology
Research Article

A well-scaling natural orbital theory

Ralph Gebauer, Morrel H. Cohen, and Roberto Car
PNAS November 15, 2016 113 (46) 12913-12918; first published November 1, 2016; https://doi.org/10.1073/pnas.1615729113
Ralph Gebauer
aThe Abdus Salam International Centre for Theoretical Physics, 34151 Trieste, Italy;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Morrel H. Cohen
bDepartment of Physics and Astronomy, Rutgers University, Piscataway, NJ 08854;
cDepartment of Chemistry, Princeton University, Princeton, NJ 08544;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: mhcohen@prodigy.net rcar@princeton.edu
Roberto Car
cDepartment of Chemistry, Princeton University, Princeton, NJ 08544;
dDepartment of Physics, Princeton University, Princeton, NJ 08544
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: mhcohen@prodigy.net rcar@princeton.edu
  1. Contributed by Morrel H. Cohen, September 27, 2016 (sent for review December 31, 2014; reviewed by Paul W. Ayers and Adam Wasserman)

  • Article
  • Figures & SI
  • Info & Metrics
  • PDF
Loading

Significance

Computations of locations of nuclei and movement of electrons within molecules and materials are widely used in science and technology. Direct computation of a system’s wave function for that purpose becomes impractical as system size grows. Current alternative methods can have difficulty with strongly correlated electron motion or spurious electron self-interaction. By using “natural spin-orbitals” to describe the motion of individual electrons, solving for them together with their joint and individual probabilities of occurrence within the system, we are able to account better for electron correlation when strong while avoiding self-interaction and maintaining the growth of computation cost with system size at the level of Hartree–Fock theory. Our numerical results for some small test molecules are very good.

Abstract

We introduce an energy functional for ground-state electronic structure calculations. Its variables are the natural spin-orbitals of singlet many-body wave functions and their joint occupation probabilities deriving from controlled approximations to the two-particle density matrix that yield algebraic scaling in general, and Hartree–Fock scaling in its seniority-zero version. Results from the latter version for small molecular systems are compared with those of highly accurate quantum-chemical computations. The energies lie above full configuration interaction calculations, close to doubly occupied configuration interaction calculations. Their accuracy is considerably greater than that obtained from current density-functional theory approximations and from current functionals of the one-particle density matrix.

  • electronic structure
  • correlation
  • density matrix

Computing the ground-state energy of N interacting electrons is central to quantum chemistry, condensed-matter physics, and related sciences. Reducing its complexity significantly below that of the many-body wave function is a major goal. Density-functional theory (DFT)† (1, 2) achieved maximal reduction using electron density as the basic variable. DFT transformed many sciences and technologies, but finding accurate, parameter-free approximations to its exchange–correlation energy functional that avoid self-interaction and capture strong electron correlation is difficult. One-particle density-matrix (1-DM) functional theories (3) have one more degree of complexity. In them, the 1-DM is often represented by its eigenvalues, the occupation numbers, and the corresponding eigenvectors, the natural spin-orbitals (NSOs), e.g., refs. 4 and 5. While avoiding the mean-field form of the 1-DM of DFT (2), the approximations to the exchange–correlation functional of the 1-DM have difficulties like those of the DFT approximations. Two-particle density-matrix (2-DM) functional theories, e.g., refs. 6⇓–8, are less reduced. The ground-state energy is a known, explicit functional of the 2-DM in Coulombic systems. However, whereas useful complete conditions are known for the N representability of the 1-DM, the form of the known complete conditions (7) for the 2-DM renders them unsuitable for practical application. Nevertheless, major progress has been made toward necessary conditions for N representability that can be systematically refined (9, 10). Although not variational, the resulting calculations are almost as accurate as full configuration interaction (FCI) calculations (9, 11, 12). The computational cost of 2-DM calculations scales as at least the sixth power of the basis-set size, significantly worse than the asymptotic third-power scaling of Hartree–Fock theory, DFT, and 1-DM theories. In our natural-orbital-functional theory, OP-NOFT, the basic variables are the NSOs, their occupation numbers, and their joint occupation probabilities (OPs). That allows us to represent the 2-DM accurately, transcending the limitations of the 1-DM theories. OP-NOFTs general form contains single-NSO through 4-NSO joint-occupation probabilities (n-OPs) and scales as the fifth power of the basis-set size. OP-NOFTs simplest formulation, for seniority 0, OP-NOFT-0, approximates doubly occupied configuration interaction (DOCI) (13). OP-NOFT-0 contains only 1- and 2-natural-orbital (NO) OPs and retains the third-power scaling of Hartree–Fock energy-functional minimization with a higher prefactor. OP-NOFT-0 describes the dissociation of simple diatomic molecules and multiatom chains with accuracy that can be comparable to that of DOCI, which uses a compact basis of Slater determinants (SD) but retains combinatorial scaling. In 4-electron systems, illustrated with the Paldus H4 test (14), OP-NOFT-0 yields results identical to those of DOCI. OP-NOFT-0 is powerful at high correlation, i.e., for static correlation at intermediate and large interatomic separations where Hartree–Fock fails due to the multireference character of the ground-state wavefunction. There, OP-NOFT-0 outperforms Hartree–Fock, DFT with standard approximations, and quantum-chemistry methods such as (single-reference) coupled cluster with single, double, and perturbative triple electron–hole excitations [CCSD(T)], a standard of accuracy near equilibrium separations. This introduction of higher-order OPs as variational parameters, with closure of the theory at their level, is an essential part of our work and is responsible for its favorable scaling.

OP-NOFT, General Formulation

The NSO Basis.

We consider time-reversal invariant saturated systems with nondegenerate, singlet ground states. The inverse approach (6⇓–8) starts from the N-representability conditions on the 2-DM. Instead, we take a forward approach: We introduce a specific form for the trial wavefunction and derive the 2-DM explicitly. Our starting point is that of conventional FCI, except that our one-particle basis is the complete set of NSOs of the trial function Ψ, ψk(x)=ϕk(r)χk(σ), with r space and σ-spin coordinates. The NOs ϕk(r) can be real and are independent of the spin function. The complete set of N-electron orthonormal SDs Φn(x1,x2,⋯,xN), n=k1,k2,⋯,kN, formed from its NSOs supports representation of any trial wavefunction Ψ(x1,x2,⋯,xN) as the expansionΨ(x1,⋯,xN)=∑n CnΦn(x1,⋯,xN).[1]As the ground-state wave function can be chosen to be real, so can the trial functions and the normalized Cn (∑nCn2=1). The NSOs vary with the trial function or the coefficients in the search for the ground state. The combinatorial complexity of determining the ground-state energy by variation of the Cn is composed of the separate combinatorial complexities of the signs and magnitudes of the coefficients Cn. We use distinct reductive approximations for their signs and magnitudes. The signs depend on the sign convention chosen for the SDs. We use the Leibniz form for the SDs,Φn(x1,⋯,xN)=1N!∑p sgn{Pp}Ppψk1(x1)⋯ψkN(xN).[2]The sum is over the elements of the symmetric group of order N, the permutations Pp. The sign of Φn is fixed by the ordering k1<k2<⋯<kN in the product of the NSOs ψki in [2]. The SDs and their coefficients can then be specified by listing the NSOs occupied in the SDs, i.e., by the index n.

The 1-DM, the Orthogonality Constraint, the Pair-Difference Constraint.

The 1-DM of Ψ,ρ(x′,x)=N∫dx2⋯dxNΨ(x′,x2,⋯,xN)Ψ(x,x2,⋯,xN),[3]becomesρ(x′,x)=∑i≠j[∑m∌i,jCi,mCj,m]ψi(x′)ψj(x)[4]after [1] and [2] are inserted into [3]. In [4] the subindex m specifies the N−1 NSOs present in Φi,m and Φj,m, excluding ψi and ψj. As the ψ are the NSOs of Ψ, the eigenfunctions of ρ(x′,x), the bracketed quantity in [4] must vanish for i≠j. Regard the coefficients Ci,m and Cj,m as the components of vectors Ci and Cj and the bracket as their scalar product Ci⋅Cj, which must vanish. There are two realizations of this orthogonality constraint (OC). In the first, and most general form, the presence of Φi,m in Ψ does not exclude the presence of Φj,m. The individual terms in the scalar product need not vanish. The second, the pair-difference constraint (PDC), is a special case of the OC, in which the presence of Φi,m excludes Φj,m so that either Ci,m or Cj,m is zero for each m, and the sum vanishes term by term. Under the PDC, those Φ present in the expansion of Ψ must differ from one another by at least two NSOs. The OC is necessary and sufficient for N representability, whereas the PDC is only sufficient. We impose the PDC on the Φ as a simplifying variational approximation. The PDC was proved to be well-satisfied in the FCI result for H8 using the minimal basis set STO-6G (Slater-type basis functions expressed with six contracted Gaussian functions).

Under the OC or PDC, ρ(x′,x) takes the diagonal formρ(x′,x)=∑k p1(k) ψk(x′)ψk(x).[5]Here, the p1(k)=∑nCn2 νk,n, where νk,n=1 if k∈n and 0 otherwise, are the eigenvalues of ρ(x′,x), the occupation numbers or occupation probabilities (1-OP) of its eigenfunctions ψk. They satisfy the necessary and sufficient conditions 0≤p1(k)≤1 and ∑kp1(k)=N. In general, only M>N occupation numbers p1(k) are nonnegligible, and only the corresponding active NSOs need be included in the representation of any trial function, providing a natural cutoff. The 1-DM is thus of algebraic complexity in the 1-OPs and the NSOs.

The 2-DM, the Sign Conjecture, the ξ-Approximation.

The 2-DM of Ψ,π(x1′x2′;x1x2)=N(N−1)∫dx3⋯dxN×Ψ(x1′,x2′,x3,⋯,xN)Ψ(x1,x2,x3,⋯,xN),becomesπ(x1′x2′;x1x2)=∑i<i′,j<j′,mi,i′,j,j′∉mCii′mCjj′m(ψi(x1′)ψi′(x2′)−ψi′(x1′)ψi(x2′))(ψj(x1)ψj′(x2)−ψj′(x1)ψj(x2)).[6]π separates into a part πd diagonal in the indices, i.e., with ii′=jj′, and an off-diagonal part, πod, with ii′≠jj′:πd(x1′x2′;x1x2)=12∑i≠j p11(ij)(ψi(x1′)ψj(x2′)−ψj(x1′)ψi(x2′))(ψi(x1)ψj(x2)−ψj(x1)ψi(x2)),[7]πod(x1′x2′;x1x2)=∑i<i′≠j<j′,mi,i′,j,j′∉mCii′mCjj′m(ψi(x1′)ψi′(x2′)−ψi′(x1′)ψi(x2′))(ψj(x1)ψj′(x2)−ψj′(x1)ψj(x2)).[8]Electron correlation is expressed through πod. The analogous off-diagonal part of ρ(x′,x) is suppressed by the OC, an advantage of the NSO basis. Note that the PDC has eliminated 3-index terms from πod in [8].

The p11(ij)= ∑nCn2 νi,nνj,n in πd are joint two-state occupation probabilities (2-OPs). Mazziotti has reported (10, 15) necessary and sufficient conditions on the 2-OPs that arise from the positivity conditions on the q-OPs, i.e., the p11⋯1(i1,i2,⋯,iq)=∑nCn2 νi1,nνi2,n⋯νiq,n, at any order 2≤q≤N. These conditions derive from the positivity conditions (15) on the diagonal elements of the 2-DM (16). Limiting ourselves to the (2,2) and (2,3) conditions, the following conditions for the 2-OPs hold:sup(p1(i)+p1(j)−1,0)≤p11(ij) ≤ p1(<),[9]sup(p1(i)+p1(j)+p1(k)−1,0)≤p11(ij)+p11(ik)+p11(jk).[10]p1(<) is the lesser of p1(i) and p1(j). In addition, the sum rule∑j(≠i) p11(ij)=(N−1)p1(i),[11]must be satisfied. Conditions 9–11 were first established by Weinhold and Bright Wilson (17). They are necessary but not sufficient conditions for N representability (7, 10, 16, 18). Establishing a complete set of conditions is quantum Merlin–Arthur (QMA)-hard in N because the number of (2,q) positivity conditions increases combinatorially with increasing q≤N. Fortunately, numerical calculations on atoms and molecules indicate that sufficiently accurate lower-bound ground-state energies often result by imposing (2,q)-positivity conditions with q≤3 (9, 19). This suggests that even in the most difficult situations, fermionic problems in atoms and molecules should require only a finite and small set of positivity conditions. Here we shall limit ourselves to conditions 9–11, as we found in our numerical calculations that they are sufficient to produce accurate lower-bounds. If higher-order conditions were found to be necessary, it would not be hard for us to add a few more.

The πd of [7] contains only 2-OPs and products of two distinct NSOs; it has at most algebraic complexity ∼M3 deriving from condition 10. Thus, when only conditions 9–11 are imposed, the combinatorial complexity of the ground-state problem resides entirely in the πod of [8]. We extract the sign s(ii′m) of the coefficient Cii′m in [8] and, relating its magnitude to the joint N-OP p11⋯1(ii′m)≡Cii′m2, we rewrite[8] asπod(x1′x2′;x1x2)=∑i<i′≠j<j′,mi,i′,j,j′∉ms(ii′m)s(jj′m)p11⋯11/2(ii′m)p11⋯11/2(jj′m)(ψi(x1′)ψi′(x2′)−ψi′(x1′)ψi(x2′))(ψj(x1)ψj′(x2)−ψj′(x1)ψj(x2)).[12]We suppose that a variational approximation exists in whichs(ii′m)s(jj′m)=s(ii′)s(jj′),∀m.[13]This sign conjecture reduces the sign complexity to algebraic, scaling as M2. πod simplifies toπod(x1′x2′;x1x2)=∑i<i′≠j<j′s(ii′)s(jj′)[∑mi,i′,j,j′∉mp11⋯11/2(ii′m)p11⋯11/2(jj′m)](ψi(x1′)ψi′(x2′)−ψi′(x1′)ψi(x2′))(ψj(x1)ψj′(x2)−ψj′(x1)ψj(x2)).[14]The quantities p11⋯11/2(ii′m) and p11⋯11/2(jj′m) are mth components of vectors p11⋯11/2(ii′) and p11⋯11/2(jj′). The bracketed quantity in [14] is their scalar product. Express it as∑mi,i′,j,j′∉mp11⋯11/2(ii′m)p11⋯11/2(jj′m)=p11001/2(ii′jj′)p00111/2(ii′jj′)ξ(ii′jj′),[15]where p1100(ii′jj′) is the square magnitude of the vector p11⋯11/2(ii′) and p0011(ii′jj′) that of p11⋯11/2(jj′). p1100(ii′jj′) is the probability that ψi and ψi′ are occupied while ψj and ψj′ are not:p1100(ii′jj′)=∑mi,i′,j,j′∉mp11⋯1(ii′m)=∑n Cn2 νi,nνi′,n(1−νj,n)(1−νj′,n),and the reverse is true for p0011(ii′jj′).

The Schwarz inequality 0≤ξ(ii′jj′)≤1 imposes bounds on ξ(ii′jj′), the cosine of the hyperangle between the vectors. The upper bound ξ=1 is exact for N=2. Substituting [15] into [14] yieldsπod(x1′x2′;x1x2)=∑i<i′≠j<j′s(ii′)s(jj′)[p1100(ii′jj′) p0011(ii′jj′)]1/2ξ(ii′jj′)(ψi(x1′)ψi′(x2′)−ψi′(x1′)ψi(x2′))(ψj(x1)ψj′(x2)−ψj′(x1)ψj(x2)),[16]in which only ξ(ii′jj′) retains combinatorial complexity:ξ(ii′jj′)=∑′p11⋯11/2(ii′m)p11⋯11/2(jj′m)(∑′p11⋯1(ii′m) ∑′p11⋯1(jj′m))1/2,where the primed sums are over all m with i,i′,j,j′∉m.

Inserting 4-OPs likep1111(ii′kl)=∑nCn2 νii′,n νkl,n;k<l≠i,i′,j,j′in place of the N-OPs in ξ reduces the complexity of πod to algebraic. The resulting approximation,ξ(ii′jj′)≈∑k<l″p11111/2(ii′kl)p11111/2(jj′kl)(∑k<l″p1111(ii′kl) ∑k<l″p1111(jj′kl))1/2,[17]is not variational, but obeys the 0,1 bounds of the Schwarz inequality. It is exact for N=4, and scales as M4. In [17] the doubly primed sums are over the indices k<l, which must differ from i,i′,j,j′. Bounds on the p1111 that are the generalizations of [9–11] for 3-OPs and 4-OPs can be formulated.

The OP-NOFT Energy Functional.

The trial energy E[Ψ]=〈Ψ|H^|Ψ〉, the expectation value of the Hamiltonian H^, is an explicit functional of the 1- and 2-DM:E[Ψ]=E[ρ,π]=tr{ρh^}+tr{πw^}.Here h^ is the single-particle kinetic-energy operator plus the external potential, and w^ is the 2-electron Coulomb interaction. E[ρ,π] splits into two parts, Ed diagonal and Eod off-diagonal in the SD:E=Ed+EodEd=tr{ρh^}+tr{πdw^}Eod=tr{πodw^}.[18]The Hartree–Fock wave function minimizes Ed; πod introduces electron correlation into Eod. The explicit forms of Ed and Eod follow from [5], [7], and [16]:Ed=∑i p1(i)hii+∑i<j p11(ij)[Jij−Kij],[19]where hii=〈ψi|h^|ψi〉, and Jij=〈ψiψi|w^|ψjψj〉 and Kij=〈ψiψj|w^|ψjψi〉 are the Coulomb and exchange integrals, respectively. The second term on the right-hand side of [19] originates from πd, the form of which is represented exactly in our theory. The second term on the right-hand side of [19] contains only positive contributions and is essential; the integral relation connecting π and ρ depends only on πd and guarantees that the E is self-interaction-free.Eod=∑i<i′≠j<j′s(ii′)s(jj′)p11001/2(ii′jj′)p00111/2(ii′jj′)ξ(ii′jj′)[Kii′jj′−Kii′j′j],[20]where Kii′,jj′=〈ψiψi′|w^|ψjψj′〉. Eqs. 18–20 define the OP-NOFT energy functional within the PDC. Including the complexity of efficient evaluation of the matrix elements, it scales as M5 if the N-representability conditions for the 3- and 4-OPs can be limited to those deriving from the (3,q) and (4,q) positivity conditions with q≤4.

Proof of the Sign Conjecture.

A variational sign approximation must be a statement about the sign s(ii′m) or s(jj′m) of each coefficient appearing in [8]. To prove [13], we must find at least one statement in which the m dependences of s(ii′m) and s(jj′m) cancel. We have found two and present one here and one in SI Appendix, section S1. The former is valid for the general case of matrix elements [Kii′,jj′−Kii′,j′j] of arbitrary sign, the latter only for positive ones.

Assigning each index l in Cn a sign s(l) and taking s(n) as their product to form s(n)=∏l∈ns(l) is a variational approximation. Consequently s(ii′m)=s(i)s(i′)s(m), ands(ii′m) s(jj′m)=s(i)s(i′)s(j)s(j′),[21]so that [13] is proved, with s(ii′)=s(i)s(i′).

This approximation treats the form and phase, 0 or π, of each NSO as independent variables. The choice of signs for each index is not specified in [21]. Most energy minimization schemes start with random initial NSOs; similarly the choice of signs in [21] should be random, half positive and half negative. The number of the initial NSOs should be greater than the anticipated value of M to allow for unequal numbers of positive and negative signs of the active NSOs.

This complete factorization of s(n) and thence of s(ii′) is a restrictive approximation. A variational approximation yielding unfactorized s(ii′) could be more accurate. In SI Appendix, section S1 we have introduced a different variational approximation and rule for the signs which leads to [13] without factorization for positive matrix elements. Random assignment of signs in [13] and the sign rule of SI Appendix, section S1 yield identical results where tested, the significance of which is discussed there.

OP-NOFT-0

Theory.

The SDs in [1] can be classified by their seniority, the number A of singly occupied one-particle states they contain. For N even and for a global spin singlet (S=0) state, the N-particle Hilbert space divides into sectors of increasing even seniority starting with A=0, where all SDs contain only doubly occupied states. For molecular systems CI expansions converge rapidly with seniority, and DOCI A=0 calculations describe static correlation rather well, as demonstrated in ref. 20.

The PDC is equivalent to a restriction on seniorities in that seniorities differing only by 4 are allowed. Recent CI calculations for systems with even numbers of electrons showed that the seniority 2 sector largely decouples from the seniority 0 sector, supporting the accuracy of the PDC (20). These considerations also apply to systems with an odd number of electrons, in which seniorities would be odd but still differ only by 4.

We now formulate OP-NOFT explicitly in the A=0 sector to illustrate further how an OP-NOFT functional is constructed and to prepare for numerical implementation; it becomes OP-NOFT-0, in which the PDC is automatically satisfied. Tracing out the spins, [5] becomesρ(r′,r)=2∑k p1(k)ϕk(r′)ϕk(r).[22]k now labels M(>N/2) active doubly occupied NO states, and the following conditions hold:0≤p1(k)≤1 and 2∑k p1(k)=N.[23]In [22] and [23] p1(k) is the occupation number of either of the paired NSOs having the NO ϕk.

In the 2-DM, double occupancy results in a major simplification of the structure of πod. We make the orbital and spin components of the NSO indices explicit. They take the form is, with i now the orbital index and s=± the spin index. The only index pairs that can enter πod in [16] are i+,i− and j+,j−. The only sets of two index pairs that can enter the right-hand side of [17] are i+,i−, k+,k− and j+,j−, k+,k−. The occupation numbers νi+ and νi− are equal, with values 0 or 1, so that all 4-NSO OPs in [17] and [16] are identical to the corresponding spin-independent 2-NO OPs, e.g., p1111(i+,i−,k+,k−)=p11(ik). The signs in [16] depend on a single orbital index, s(i+,i−)=s(i), and the ξ depend on two-orbital indices, ξ(i+,i−,j+,j−)=ξ(ij). With these simplifications, the 2-DM of [7] and [16] becomesπ(r1′r2′;r1r2)=πd(r1′r2′;r1r2)+πod(r1′r2′;r1r2),[24]after tracing out the spins, withπd(r1′r2′;r1r2)=2∑ij p11(ij)(2ϕi(r1′)ϕj(r2′)ϕi(r1)ϕj(r2)−ϕi(r1′)ϕj(r2′)ϕj(r1)ϕi(r2)),[25]πod(r1′r2′;r1r2)=2∑i≠j s(i)s(j)[p10(ij)p01(ij)]1/2ξ(ij)ϕi(r1′)ϕi(r2′)ϕj(r1)ϕj(r2).[26]The sum in [25] includes the term i=j, for which p11(ii)=p1(i), and ξ(ij) in [26] is nowξ(ij)≈∑k(≠i,j) p111/2(ik)p111/2(jk)[∑k(≠i,j) p11(ik) ∑k(≠i,j) p11(jk)]1/2.[27]The one- and two-orbital OPs of OP-NOFT-0 lie within the same bounds as in the general case, [9] and [10], and their sum rules become, respectively, [23] and2∑j(≠i) p11(ij)=(N−2)p1(i).[28]The π of [24] satisfies two important sum rules:∫dr2 π(rr2;r′r2)=(N−1)ρ(r,r′)∫dr1 dr2 π(r1r2;r1r2)w(r12)≥0.The OP-NOFT-0 form for π, [24–27], is exact when N=2 with ξ=1. It is equivalent to DOCI for N=4 if the signs are correct. We show numerically that this is the case for the sign choice of SI Appendix, section S1 for the Paldus H4 test, as reported in SI Appendix, section S5. For all of the H4 configurations studied, the OP-NOFT-0 correlation energy coincides with that of DOCI to numerical precision. That the signs are correct for H2 and H4 confirms the validity of the sign rule in those cases and suggests a broader utility. When N>4, the ξ-approximation of [27] and the limitation to the (2,2) and (2,3) positivity conditions break the equivalence to DOCI.

The expectation value E=〈Ψ|H^|Ψ〉 becomesE=2∑i p1(i)〈ϕi|h^|ϕi〉+∑ij p11(ij)(2Jij−Kij)+∑i≠j s(i)s(j)p101/2(ij) p011/2(ij) ξ(ij) Kij,[29]where Jij and Kij are, respectively, positive Hartree and exchange integrals defined in terms of the NOs, Jij=〈ϕiϕi|w^|ϕjϕj〉 and Kij=〈ϕiϕj|w^|ϕjϕi〉. With [27] for ξ(ij), E in [29] is a functional of the NOs and the 1- and 2-state OPs. Kollmar introduced a similar J-K functional but simplified the 2-DM (21). The signs are chosen by a sign rule and are not variables. p10 is related to p11 and p1 by p1(i)=p11(ij)+p10(ij) and is eliminated from the functional. Each sum in the denominator of ξ(ij) in [27] is simplified by the sum rule of [28] to, e.g.,∑k(≠i,j)p11(ik)=12(N−2)p1(i)−p11(ij).As stated above, we assume that the (2,2) and (2,3) positivity conditions are sufficient in practice. Under this circumstance, the infimum of E with respect to the NOs and the OPs, subject to the constraints 23 and 9, 10, and 28, yields a variational approximation to the ground-state energy, apart from the ξ-approximation 27, for N>4.

Eq. 29 is a generalization of the NOFT formulations of 1-DM functional theories, which require only 1-state OPs. The extra complexity from 2-state OPs and implicit 4-state OPs is more than compensated by the substantial gain in accuracy it makes possible. The computational cost of calculating E from [29] scales like Hartree–Fock energy-functional minimization with a greater prefactor [M3 vs. (N/2)3] due to fractional occupation of NOs.

Numerical Results for Simple Molecular Systems.

To test OP-NOFT-0, we studied several diatomic molecules, the Paldus H4 test, and linear chains of H atoms with open boundary conditions. We included all electrons (core and valence) and expanded the NOs in the Gaussian 6–31G** (Pople's notation for a double-zeta split-valence Gaussian basis set), STO-6G, and cc-pVTZ (correlation consistent valence triple-zeta Gaussian basis set with polarization functions) bases. The constrained minimization was performed by damped Car–Parrinello dynamics (22), as detailed in SI Appendix, section S2.

We started the minimization from NOs and OPs obeying the constraints but otherwise random. The signs were taken from the sign rule of SI Appendix, section S1, Table S2. They were kept fixed during optimization.

At convergence, the active subset of NOs had p1≥10−3. The remaining NOs contributed negligibly to the energy. The same active NOs and signs were found for several test cases starting instead from a sufficiently large set of random NOs, half with positive and half with negative signs.‡ It is significant for the rule of [21] for arbitrary matrix-element signs that for the systems tested, the Brillouin–Wigner perturbation-theory-based rule of SI Appendix, Table S1, and the alternative of random initial assignment of signs to pairs yield the same results for positive matrix elements. The procedure of [21] also yields half positive and half negative signs for the pairs when signs are assigned randomly to the individual NOs with no reference to the matrix-element signs.

We report the dissociation energy curves of the dimers H2, LiH, and HF in SI Appendix, section S3, Figs. S1–S3. We performed restricted Hartree–Fock, DFT [PBE (24) and/or PBE0 (25)], complete active space self-consistent field (CASSCF), and coupled cluster with single, double, and perturbative triple electron-hole excitations [CCSD(T)] calculations with the same basis. For H2, our functional depends only on 1-state OPs and the signs s(i); it reduces to the exact expression of Löwdin and Shull (26). When the s(i) are chosen from SI Appendix, Table S2, the OP-NOFT-0 dissociation energy curve coincides with that of CASSCF at all interatomic separations, implying that this sign rule is exact for H2. Even in a system as simple as H2, spin-restricted Hartree–Fock and DFT fail badly at dissociation because these single-reference theories cannot recover the Heitler–London form of the wavefunction.

OP-NOFT-0 becomes identical to DOCI for N=4 when the sign choice is correct. The conditions 9, 10, and 28 simplify in this case as discussed in SI Appendix, section S4. Expression 27 for ξ is exact, but the A=0 restriction is not. We performed the H4 Paldus test using a minimal 1s basis set for OP-NOFT-0, DOCI, and FCI. The DOCI/OP-NOFT-0 equivalence and the accuracy of the OP-NOFT-0 signs are confirmed by the results reported in SI Appendix, section S5. Whereas DOCI only captures about 25–90% of the configuration-dependent correlation energy in the Paldus test, the OP-NOFT-0 dissociation curve of LiH almost coincides with CASSCF in SI Appendix, Fig. S2, which indicates that its 1s electrons are nearly inert so that higher seniorities contribute negligibly to its ground-state energy.

HF, a 10-electron system, provides a complete test of the relation of OP-NOFT-0 to DOCI. The energies obtained with the basis set 6–31G** are shown in SI Appendix, Fig. S3. OP-NOFT-0 and DOCI are above both CASSCF and CCSD(T), although OP-NOFT-0 lies below DOCI because, although size-consistent, the ξ-approximation is nonvariational here. The OP-NOFT-0 signs are correct. The results are discussed further in SI Appendix, section S3.

Linear H chains are relatively simple systems whose energy surfaces present a serious challenge for single-reference methods. Fig. 1 shows symmetric dissociation energy curves of H8. OP-NOFT-0 provides a consistent description of the energy close to and everywhere above the CASSCF reference. The breakdown of CCSD(T) at large separations is caused by its single-reference character. The deviation of OP-NOFT-0 from CASSCF should be attributed mainly to the restriction to the A=0 sector, a conclusion supported by the seniority-restricted CI calculations of ref. 20. Close comparison with those calculations is not entirely straightforward, as ref. 20 used the slightly smaller 6–31G basis and a fixed, symmetric, or broken symmetry, molecular orbital basis, whereas we used self-consistent NOs.

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Symmetric dissociation curve of a linear H8 chain. The squares indicate one-half of the energy of a H16 chain (black square: Hartree–Fock energy; green squares: OP-NOFT-0 energy). All Hartree–Fock results are from spin-restricted calculations.

The OP-NOFT-0 1-DM displays the entanglement due to correlation through variation of the occupation numbers and the von Neumann entanglement entropy with interatomic separation shown in SI Appendix, section S6, Fig. S5. The increase of entanglement entropy with separation signals a dramatic increase of correlation corresponding to multireference character. The OP-NOFT-0 2-DM gives access to electron pair correlations.

To test the dependence of the accuracy of OP-NOFT-0 on electron number, we studied the symmetric dissociation of H16.§ Results for the energy of H16 divided by 2 are shown as squares in Fig. 1. OP-NOFT-0 works equally well for this longer chain. The total energy at dissociation is twice that of H8, and the slightly increased binding energy per atom at equilibrium arises from an increase in the correlation energy, as expected from more effective screening in the larger system.

The N2 molecule is a severe test because of its triple bond. OP-NOFT-0, DOCI, and FCI results obtained with the minimal basis set STO-6G are compared in Fig. 2. They are similar to those for HF, with OP-NOFT-0 and DOCI both above FCI but OP-NOFT-0 below DOCI. The accuracy of our DOCI results was improved by use of the optimized OP-NOFT-0 NOs for the DOCI basis.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Dissociation curve of the N2 molecule.

Note that in all of the systems studied, the positivity condition (2,2) was found to be sufficient at near-equilibrium up to intermediate separations dominated by dynamic correlation because the (2,3) condition was automatically satisfied there. Moreover, only in the case of H8, H16, and N2 at large separations did inclusion of the (2,3) positivity condition turn out to be essential to prevent runaway from the ground-state solution.

Discussion

We have introduced a method for correlated electronic-structure calculations, OP-NOFT, that scales algebraically. Its DOCI-like simplification, OP-NOFT-0, scales favorably with system size, with Hartree–Fock energy-minimization scaling. The close correspondence of the energies calculated via OP-NOFT-0 with DOCI calculations supports the accuracy of limiting the positivity conditions to [9–10] and also of the ξ-approximation 27. OP-NOFT-0 is restricted to the A=0 sector of the Hilbert space. It provides an accurate description of single-bond breaking and is a considerable improvement over single-reference methods in all cases studied.

Adding the computation of interatomic forces to the OP-NOFT-0 energy-minimization methodology would make possible the use of the theory for structural optimization and ab initio molecular dynamics (22).

From the practical point of view, minimization of the functional [29] is significantly more laborious than minimization of the Hartree–Fock or the DFT functional. We attribute this difficulty to the need to include in [29] small occupation numbers. In damped dynamics minimization the forces acting on the corresponding NOs are thus very weak compared with the forces acting on the NOs with occupation numbers close to 1, slowing down the entire procedure considerably. This difficulty is common to all NO-based methods including those based on the 1-DM. Solving it is essential to making OP-NOFT methods applicable in practice. Care must be taken to avoid spurious minima, as in other nonlinear optimization problems.

Acknowledgments

The authors acknowledge illuminating discussions with Paul W. Ayers and Kieron Burke. Refs. 13 and 17 were brought to the authors’ attention by Paul W. Ayers. The authors further thank J. E. Moussa for important comments. M.H.C. and R.C. acknowledge support from the Department of Energy under Grant DE-FG02-05ER46201.

Footnotes

  • ↵1To whom correspondence may be addressed. Email: mhcohen{at}prodigy.net or rcar{at}princeton.edu.
  • Author contributions: R.G., M.H.C., and R.C. designed research, performed research, contributed new reagents/analytic tools, analyzed data, and wrote the paper.

  • Reviewers: P.W.A., McMaster University; and A.W., Purdue University.

  • The authors declare no conflict of interest.

  • ↵†SI Appendix, section S8 is a list of acronyms.

  • ↵‡For H2 the sign rule (s(i≤N)=+1 and s(i>N)=−1) holds near equilibrium, but a more complex pattern emerges at large separation where additional positive signs are needed for the van der Waals tail of the interaction potential (23). In principle, these positive signs could be obtained with our minimization procedure, but their effect is beyond the accuracy of the present calculations.

  • ↵§We do not give the CASSCF energies in this case, as the dimension of the active subspace would make these calculations very expensive.

  • This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1615729113/-/DCSupplemental.

Freely available online through the PNAS open access option.

View Abstract

References

  1. ↵
    1. Hohenberg P,
    2. Kohn W
    (1964) Inhomogeneous electron gas. Phys Rev 136:B864–B871.
    .
    OpenUrlCrossRef
  2. ↵
    1. Kohn W,
    2. Sham LJ
    (1965) Self-consistent equations including exchange and correlation effects. Phys Rev 140:A1133–A1138.
    .
    OpenUrlCrossRef
  3. ↵
    1. Gilbert TL
    (1975) Hohenberg-Kohn theorem for nonlocal external potentials. Phys Rev B 12:2111–2120.
    .
    OpenUrlCrossRef
  4. ↵
    1. Gritsenko O,
    2. Pernal K,
    3. Baerends EJ
    (2005) An improved density matrix functional by physically motivated repulsive corrections. J Chem Phys 122(20):204102.
    .
    OpenUrlCrossRefPubMed
  5. ↵
    1. Lathiotakis NN,
    2. Marques MAL
    (2008) Benchmark calculations for reduced density-matrix functional theory. J Chem Phys 128(18):184103.
    .
    OpenUrlCrossRefPubMed
  6. ↵
    1. Coleman AJ
    (1963) Structure of fermion density matrices. Rev Mod Phys 35:668–686.
    .
    OpenUrlCrossRef
  7. ↵
    1. Garrod C,
    2. Percus JK
    (1964) Reduction of the N-particle variational problem. J Math Phys 5:1756.
    .
    OpenUrlCrossRef
  8. ↵
    1. Percus J
    (2013) On the trail of the 2-body reduced density matrix. Comput Theor Chem 1003:2–7.
    .
    OpenUrlCrossRef
  9. ↵
    1. Zhao Z,
    2. Braams BJ,
    3. Fukuda M,
    4. Overton ML,
    5. Percus JK
    (2004) The reduced density matrix method for electronic structure calculations and the role of three-index representability conditions. J Chem Phys 120(5):2095–2104.
    .
    OpenUrlCrossRefPubMed
  10. ↵
    1. Mazziotti DA
    (2012) Structure of fermionic density matrices: Complete N-representability conditions. Phys Rev Lett 108(26):263002.
    .
    OpenUrlCrossRefPubMed
  11. ↵
    1. Nakata M, et al.
    (2001) Variational calculations of fermion second-order reduced density matrices by semidefinite programming algorithm. J Chem Phys 114:8282.
    .
    OpenUrlCrossRef
  12. ↵
    1. Gidofalvi G,
    2. Mazziotti DA
    (2008) Active-space two-electron reduced-density-matrix method: Complete active-space calculations without diagonalization of the N-electron Hamiltonian. J Chem Phys 129(13):134108.
    .
    OpenUrlCrossRefPubMed
  13. ↵
    1. Weinhold F,
    2. Bright Wilson E
    (1967) Reduced density matrices of atoms and molecules. I. The 2 matrix of double-occupancy, configuration-interaction wavefunctions for singlet states. J Chem Phys 46:2752.
    .
    OpenUrlCrossRef
  14. ↵
    1. Jankowski K,
    2. Paldus J
    (1980) Applicability of coupled-pair theories to quasidegenerate electronic states: A model study. Int J Quantum Chem 18:1243–1269.
    .
    OpenUrlCrossRef
  15. ↵
    1. Mazziotti DA
    (2012) Significant conditions for the two-electron reduced density matrix from the constructive solution of N representability. Phys Rev A 85:062507.
    .
    OpenUrlCrossRef
  16. ↵
    1. Ayers PW,
    2. Davidson ER
    (2007) Linear inequalities for diagonal elements of density matrices. Adv Chem Phys 134:443–483.
    .
    OpenUrl
  17. ↵
    1. Weinhold F,
    2. Bright Wilson E
    (1967) Reduced density matrices of atoms and molecules. II. On the N-representability problem. J Chem Phys 47:2298.
    .
    OpenUrlCrossRef
  18. ↵
    1. Davidson ER
    (1969) Linear inequalities for density matrices. J Math Phys 10:725.
    .
    OpenUrlCrossRef
  19. ↵
    1. Mazziotti DA
    (2012) Two-electron reduced density matrix as the basic variable in many-electron quantum chemistry and physics. Chem Rev 112(1):244–262.
    .
    OpenUrlCrossRefPubMed
  20. ↵
    1. Bytautas L,
    2. Henderson TM,
    3. Jiménez-Hoyos CA,
    4. Ellis JK,
    5. Scuseria GE
    (2011) Seniority and orbital symmetry as tools for establishing a full configuration interaction hierarchy. J Chem Phys 135(4):044119.
    .
    OpenUrlCrossRefPubMed
  21. ↵
    1. Kollmar C
    (2004) The “JK-only” approximation in density matrix functional and wave function theory. J Chem Phys 121(23):11581–11586.
    .
    OpenUrlCrossRefPubMed
  22. ↵
    1. Car R,
    2. Parrinello M
    (1985) Unified approach for molecular dynamics and density-functional theory. Phys Rev Lett 55(22):2471–2474.
    .
    OpenUrlCrossRefPubMed
  23. ↵
    1. Sheng XW,
    2. Mentel ŁM,
    3. Gritsenko OV,
    4. Baerends EJ
    (2013) A natural orbital analysis of the long range behavior of chemical bonding and van der Waals interaction in singlet H2: The issue of zero natural orbital occupation numbers. J Chem Phys 138(16):164105.
    .
    OpenUrlCrossRefPubMed
  24. ↵
    1. Perdew JP,
    2. Burke K,
    3. Ernzerhof M
    (1996) Generalized gradient approximation made simple. Phys Rev Lett 77(18):3865–3868.
    .
    OpenUrlCrossRefPubMed
  25. ↵
    1. Adamo C,
    2. Barone V
    (1999) Toward reliable density functional methods without adjustable parameters: The PBE0 model. J Chem Phys 110:6158.
    .
    OpenUrlCrossRef
  26. ↵
    1. Löwdin PO,
    2. Shull H
    (1956) Natural orbitals in the quantum theory of two-electron systems. Phys Rev 101:1730–1739.
    .
    OpenUrlCrossRef
PreviousNext
Back to top
Article Alerts
Email Article

Thank you for your interest in spreading the word on PNAS.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
A well-scaling natural orbital theory
(Your Name) has sent you a message from PNAS
(Your Name) thought you would like to see the PNAS web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Well-scaling natural orbital theory
Ralph Gebauer, Morrel H. Cohen, Roberto Car
Proceedings of the National Academy of Sciences Nov 2016, 113 (46) 12913-12918; DOI: 10.1073/pnas.1615729113

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Request Permissions
Share
Well-scaling natural orbital theory
Ralph Gebauer, Morrel H. Cohen, Roberto Car
Proceedings of the National Academy of Sciences Nov 2016, 113 (46) 12913-12918; DOI: 10.1073/pnas.1615729113
Digg logo Reddit logo Twitter logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Mendeley logo Mendeley
Proceedings of the National Academy of Sciences: 113 (46)
Table of Contents

Submit

Sign up for Article Alerts

Article Classifications

  • Physical Sciences
  • Applied Physical Sciences

Jump to section

  • Article
    • Abstract
    • OP-NOFT, General Formulation
    • OP-NOFT-0
    • Discussion
    • Acknowledgments
    • Footnotes
    • References
  • Figures & SI
  • Info & Metrics
  • PDF

You May Also be Interested in

Abstract depiction of a guitar and musical note
Science & Culture: At the nexus of music and medicine, some see disease treatments
Although the evidence is still limited, a growing body of research suggests music may have beneficial effects for diseases such as Parkinson’s.
Image credit: Shutterstock/agsandrew.
Scientist looking at an electronic tablet
Opinion: Standardizing gene product nomenclature—a call to action
Biomedical communities and journals need to standardize nomenclature of gene products to enhance accuracy in scientific and public communication.
Image credit: Shutterstock/greenbutterfly.
One red and one yellow modeled protein structures
Journal Club: Study reveals evolutionary origins of fold-switching protein
Shapeshifting designs could have wide-ranging pharmaceutical and biomedical applications in coming years.
Image credit: Acacia Dishman/Medical College of Wisconsin.
White and blue bird
Hazards of ozone pollution to birds
Amanda Rodewald, Ivan Rudik, and Catherine Kling talk about the hazards of ozone pollution to birds.
Listen
Past PodcastsSubscribe
Goats standing in a pin
Transplantation of sperm-producing stem cells
CRISPR-Cas9 gene editing can improve the effectiveness of spermatogonial stem cell transplantation in mice and livestock, a study finds.
Image credit: Jon M. Oatley.

Similar Articles

Site Logo
Powered by HighWire
  • Submit Manuscript
  • Twitter
  • Facebook
  • RSS Feeds
  • Email Alerts

Articles

  • Current Issue
  • Latest Articles
  • Archive

PNAS Portals

  • Anthropology
  • Chemistry
  • Classics
  • Front Matter
  • Physics
  • Sustainability Science
  • Teaching Resources

Information

  • Authors
  • Editorial Board
  • Reviewers
  • Librarians
  • Press
  • Site Map
  • PNAS Updates

Feedback    Privacy/Legal

Copyright © 2021 National Academy of Sciences. Online ISSN 1091-6490