We explore the application of an extrapolative method that yields very accurate total and relative energies from variational and diffusion quantum Monte Carlo (VMC and DMC) results. For a trial wave function consisting of a small configuration interaction (CI) wave function obtained from full CI quantum Monte Carlo and reoptimized in the presence of a Jastrow factor and an optional backflow transformation, we find that the VMC and DMC energies are smooth functions of the sum of the squared coefficients of the initial CI wave function and that quadratic extrapolations of the non-backflow VMC and backflow DMC energies intersect within uncertainty of the exact total energy. With adequate statistical treatment of quasi-random fluctuations, the extrapolate and intersect with polynomials of order two method is shown to yield results in agreement with benchmark-quality total and relative energies for the C2, N2, CO2, and H2O molecules, as well as for the C2 molecule in its first electronic singlet excited state, using only small CI expansion sizes.

Quantum Monte Carlo (QMC) methods are a broad family of stochastic wave-function-based techniques that accurately approximate the solution of the Schrödinger equation of an electronic system. The variational quantum Monte Carlo (VMC) method1,2 obtains expectation values corresponding to an analytic trial wave function ΨT(R) in real space and provides a framework for optimizing wave function parameters,3,4 such as those in the multideterminant-Jastrow-backflow form,
(1)
where {DI} are M Slater determinants, eJ(R) is a Jastrow correlation factor,5,6 and X(R) are the backflow-transformed electronic coordinates.7 Diffusion quantum Monte Carlo (DMC)2,8 is a real-space projection method that recovers the lowest-energy solution Φ(R) of the Schrödinger equation compatible with the fixed-node condition that Φ(RT(R) be nonnegative everywhere. We refer to the VMC and DMC methods collectively as continuum QMC (cQMC) methods.
In the configuration interaction (CI) ansatz, the solution of the Schrödinger equation is expressed as
(2)
where {|DI⟩} are all possible Slater determinants that can be constructed on a given orbital basis. Equation (2) is exact in the limit of an infinite expansion using a complete basis set, but in practical methods, finite expansions and bases are used. Full CI quantum Monte Carlo (FCIQMC)9,10 is a second-quantized projection technique in which random walkers sample the discrete Hilbert space defined by {|DI⟩} in order to determine approximate values of the CI coefficients {cI}.

The nominal computational cost of cQMC calculations scales polynomially with system size N, typically as N2N4, and the quality of the resulting energies depends on the accuracy of the trial wave function for VMC and on the accurate location of its nodes for DMC. The cQMC methods excel at describing explicit dynamic and long-ranged correlations, but the error incurred by the fixed-node approximation is often significant. By contrast, FCIQMC is formally an exponentially scaling method that trivially captures static correlations but requires a very large number of walkers to provide a good description of dynamic correlations. The complementary nature of the strengths of cQMC and FCIQMC makes combining these methods highly desirable. Several ways of combining cQMC and FCIQMC have been presented in the literature, such as using DMC to assist in the extrapolation to the thermodynamic limit of FCIQMC energies of the electron gas11 or using VMC-optimized Jastrow factors in FCIQMC with the transcorrelated method.12,13 Here, we shall focus on the use of selected CI wave functions generated with FCIQMC to construct multideterminantal trial wave functions for cQMC calculations.

Multideterminant expansions have been used for decades in cQMC calculations of atomic and molecular systems, including ground-state energy calculations,14–19 excitation energies,20–23 and geometry optimizations.22,24 The use of truncated CI expansions in cQMC presents the problem that no reliable criteria exist to truncate wave functions for different systems in a consistent manner, resulting in energy differences of questionable accuracy. One possible approach is to use extremely large multideterminantal wave functions19–27 under the expectation that the fixed-node error in the total energies will become smaller than the target error. While algorithmic developments have vastly reduced the computational cost associated with the use of multideterminantal wave functions in cQMC,25,28–31 this remains an expensive choice. Using trial wave functions without a Jastrow factor reduces the nominal computational burden20,21,27 at the cost of losing the accurate, compact description of dynamic correlation afforded by fully optimized trial wave functions. By including explicit correlations, in the present paper we are able to explore the use of relatively small multideterminantal wave functions to perform an extrapolation of the cQMC total energy to the full-CI, complete orbital-basis limit. We test our method on a variety of molecular systems, obtaining total and relative energies within uncertainty of benchmark-quality results from the literature.

The rest of this paper is structured as follows: In Sec. II, we present the methodological details of our extrapolation method, which we illustrate with calculations of the carbon dimer and the water molecule. We then apply our method to several atomic and molecular systems, and we report the results in Sec. III. Our conclusions and outlook are presented in Sec. IV. Hartree atomic units ( = |e| = me = 4πϵ0 = 1) are used throughout; the uncertainties and error bars we report refer to standard 68.3% (one-sigma) confidence intervals except when explicitly noted otherwise.

Let Mgen be the number of determinants occupied at a given point in an equilibrated FCIQMC calculation, representing the CI wave function,
(3)
where cI is obtained as the sum of the signed weights of the walkers occupying the Ith determinant. The values of the first few coefficients { c I } I M gen converge relatively quickly in FCIQMC calculations and can be expected to be reasonably close to their full CI (FCI) values. This makes FCIQMC an ideal method for quickly generating good-quality selected CI wave functions of moderate sizes—studying the suitability of other CI solvers for this purpose is beyond the scope of this paper.
Let us consider the wave function obtained by truncating Eq. (3) to size MMgen. The sum of the squares of the coefficients of the resulting wave function relative to that at size Mgen is
(4)
which goes to 1 as MMgen. This CI wave function of size M can be combined with a Jastrow factor and, optionally, with a backflow transformation to produce a multideterminant-Jastrow(-backflow) trial wave function for cQMC, as given in Eq. (1). The wave function parameters can be (re-)optimized in the context of VMC, producing a trial wave function with which to compute VMC and DMC energies. Repeating this procedure by truncating the original CI wave function to different sizes yields a set of VMC and DMC energies that can be plotted as a function of w.
Plots of this kind, although using other CI solvers, can be found in the literature; see Fig. 3 of Ref. 4 or Fig. 4 of Ref. 29, for example. The present work is, in fact, inspired by the observation that the VMC and DMC curves in these plots appear to be smooth and would seem to be about to intersect just off the right-hand side of the graph. In Fig. 1, we plot the VMC and DMC energies we obtain for the ground state of the all-electron C2 molecule (see Table I) using Hartree–Fock orbitals expanded in the cc-pCVTZ basis set,32,33 along with quadratic fits to the data of the form,
(5)
where a, b, and c are fit parameters. In this simple example, the fits to the VMC and DMC data intersect at w = 1.031, corresponding to a total energy of −75.9287 Ha, not far off the exact nonrelativistic total energy estimate of −75.9265 Ha given in Ref. 34. We refer to this way of estimating the total energy of a system as the extrapolate and intersect with polynomials of order two (xspot) method.
FIG. 1.

VMC and DMC total ground-state energy of the carbon dimer using multideterminant-Jastrow trial wave functions as a function of w, using Hartree–Fock orbitals expanded in the cc-pCVTZ basis set. Quadratic fits to the data are extended beyond w = 1 to show their intersection, which is in good agreement with the estimated exact nonrelativistic total energy of the system.34 

FIG. 1.

VMC and DMC total ground-state energy of the carbon dimer using multideterminant-Jastrow trial wave functions as a function of w, using Hartree–Fock orbitals expanded in the cc-pCVTZ basis set. Quadratic fits to the data are extended beyond w = 1 to show their intersection, which is in good agreement with the estimated exact nonrelativistic total energy of the system.34 

Close modal
TABLE I.

Atoms and molecules considered in this work, along with their electronic states and geometries.

System Geometry34,35 State
  3P 
  4S 
  3P 
C2  r CC = 1.2425 A ̊   1 Σ g +  
C 2 *   r CC = 1.2425 A ̊   B1Δg 
H2 r OH = 0.9572 A ̊   1A1 
  HOH = 104.5°   
N2  r NN = 1.0977 A ̊   1 Σ g +  
CO2  r CO = 1.1600 A ̊   1 Σ g +  
  OCO = 180°   
System Geometry34,35 State
  3P 
  4S 
  3P 
C2  r CC = 1.2425 A ̊   1 Σ g +  
C 2 *   r CC = 1.2425 A ̊   B1Δg 
H2 r OH = 0.9572 A ̊   1A1 
  HOH = 104.5°   
N2  r NN = 1.0977 A ̊   1 Σ g +  
CO2  r CO = 1.1600 A ̊   1 Σ g +  
  OCO = 180°   

In what follows, we develop the methodology to enable the application of the xspot method in practice using as test systems the C, N, and O atoms, the ground-state C2, N2, H2O, and CO2 molecules, and the C2 molecule in its lowest-lying singlet electronic excited state, which we refer to simply as C 2 * . These atoms and molecules are simulated as all-electron, both in the sense that no effective-core potentials are used and that excitations from “core” orbitals are allowed in the CI wave function. In Table I, we give the states and geometries we have used for these systems.

The extrapolation shown in Fig. 1 might seem simplistic from a quantum chemical perspective, given that all calculations involved have been performed with the same, finite orbital basis, so one would expect an orbital-basis dependent result, which should itself be extrapolated to the complete-basis limit. For instance, the FCIQMC energy tends to a basis-set dependent FCI limit as the number of walkers tends to infinity, and this must in turn be extrapolated to the basis-set limit in order to obtain the exact energy of the system.

In what follows, we will conceptually combine the choice of molecular orbitals (e.g., Hartree–Fock, natural orbitals, …) with the choice of basis set (e.g., cc-pCVDZ, cc-aug-pVTZ, …), so we shall discuss the completeness of the (molecular) orbital basis instead of that of the basis set alone to emphasize this point.

The xspot extrapolation procedure can be easily justified in the hypothetical case of using an infinite, “complete” orbital basis. The FCIQMC energy with this “complete” orbital basis would tend to the exact total energy of the system E0 in the infinite walker-number limit, and the sum of the squared CI coefficients would also tend to that of the exact wave function, w0. The exact wave function has no dynamic correlation left to recover, so the Jastrow factor and backflow displacement in the cQMC trial wave function would optimize to zero, and the VMC and DMC energies would both coincide with E0. At finite expansion sizes, w < w0, the VMC and DMC methods yield variational energies satisfying EVMCEDMCE0, which, assuming these to be smooth functions of w, validates the xspot method with the “complete” orbital basis.

We note that in a truncated CI wave function, the infinite, “complete” orbital basis is effectively finite since a finite number of determinants can only contain a finite number of distinct orbitals. Conversely, a sufficiently small selected CI wave function with a finite orbital basis is indistinguishable from a CI wave function of the same size with the “complete” orbital basis—assuming the finite basis contains the first few orbitals in the “complete” orbital basis.

As the orbitals in a finite basis get used up, the cQMC energies can be expected to plateau as a function of w as they tend to their orbital-basis dependent limit. We refer to this phenomenon as “orbital-basis exhaustion” and to the onset of this plateau as the exhaustion limit wexh.. Note that orbital bases such as natural orbitals can be constructed so as to compactly describe the system with fewer orbitals, which has the side effect of reducing the value of wexh.. We discuss this aspect further in Sec. II C.

As a proxy for the degree of orbital-basis exhaustion, in Table II, we show the fraction of orbitals used in CI wave functions of the same size for Hartree–Fock orbitals expanded in four different basis sets in the cc-pVxZ and cc-pCVxZ families32,33 for the all-electron carbon dimer. Based on these numbers, we use the cc-pCVTZ basis throughout this paper to ensure we have enough leeway to increase the multideterminant wave function size before hitting the exhaustion limit. We provide an a posteriori assessment of this choice in Sec. III.

TABLE II.

Fraction of spatial orbitals in the Hartree-Fock orbital basis that appear in the first 300 configuration state functions of the full CI wave function for the ground state of the carbon dimer using four different basis sets.

Basis set Orbitals used (%)
cc-pVDZ  28/28 = 100 
cc-pCVDZ  32/36 = 89 
cc-pVTZ  38/60 = 82 
cc-pCVTZ  40/86 = 47 
Basis set Orbitals used (%)
cc-pVDZ  28/28 = 100 
cc-pCVDZ  32/36 = 89 
cc-pVTZ  38/60 = 82 
cc-pCVTZ  40/86 = 47 

Finite-orbital-basis FCIQMC and cQMC calculations performed at w < wexh. behave as if one were using the “complete” orbital basis. Therefore, it is legitimate to expect that the extrapolation of quadratic fits to these VMC and DMC data intersect at w = w0 and E = E0, provided that the VMC and DMC energies are smooth functions of w representable by a second-order polynomial for wh.o. < w < wexh., where wh.o. is a threshold below which higher-order polynomials would be needed.

Note that the initial FCIQMC wave function with Mgen determinants is not required to be below the exhaustion limit since it simply serves to construct selected CI wave functions of size MMgen, which are required to be below the exhaustion limit, and to define the arbitrary point at which w = 1 in the plots; w = 1 has no special significance in this method.

In our calculations, we choose CI wave function sizes so that the points are more or less evenly spaced on the w axis, and we make sure that different points correspond to wave functions containing a different number of distinct spatial orbitals so as to capture the effect of simultaneously growing the CI expansion and the orbital basis.

In reality, cQMC energies do not exactly follow smooth curves, but it is reasonable to assume that a smooth underlying trend E(w) exists and that the cQMC energy Ei deviates from it by a quasi-random amount qi. Considering also the statistical uncertainty ΔEi, the ith point in a set of cQMC energies can then be modeled as
(6)
where ζi is a random number drawn from the standard normal distribution. In order to make this generic model for the quasi-random fluctuations useful in practice, we make the approximation that qi = ξiα, where ξi is a random number drawn from the standard normal distribution and α is a constant amplitude independent of w. In Fig. 2, we illustrate our model of cQMC energy data as a function of w.
FIG. 2.

Illustration of the expected behavior of cQMC energies as a function of w. The cQMC energies (circles and squares) deviate from the underlying smooth trend (lines) by a quasi-random amount (of amplitude represented by the width of the shaded area around the lines). The smooth trend can be represented by a quadratic function, E(w) (dashed line), for wh.o. < w < wexh. (shaded middle region), while for w < wh.o. higher-order contributions become important, and for w > wexh. orbital-basis exhaustion sets in. At the value of w corresponding to the exact wave function, the quadratic function gives the exact energy, E0 = E(w0).

FIG. 2.

Illustration of the expected behavior of cQMC energies as a function of w. The cQMC energies (circles and squares) deviate from the underlying smooth trend (lines) by a quasi-random amount (of amplitude represented by the width of the shaded area around the lines). The smooth trend can be represented by a quadratic function, E(w) (dashed line), for wh.o. < w < wexh. (shaded middle region), while for w < wh.o. higher-order contributions become important, and for w > wexh. orbital-basis exhaustion sets in. At the value of w corresponding to the exact wave function, the quadratic function gives the exact energy, E0 = E(w0).

Close modal
We estimate the value of α by performing a preliminary least-squares fit to the bare data, E ̃ ( w ) , and evaluating
(7)
i.e., we obtain α as the root-mean-square deviation of the data from the fit value not accounted for by statistical uncertainty alone. For this procedure to produce a meaningful result, the number of data points in each curve must be significantly greater than the number of parameters in the quadratic fit function; we use at least seven data points for all fits reported in this paper.

In order to account for the statistical uncertainty and quasi-random fluctuations in the xspot method, we use a Monte Carlo resampling technique in which we generate 100 000 instances of each VMC and DMC dataset in which a random amount proportional to α 2 + Δ E i 2 is added to the original energy values. We then perform fits on these shifted data, find the intersection point for each such instance, and obtain the final result by averaging over instances; see Fig. 3 for an illustration of this process. This procedure provides meaningful uncertainties on the intersection energies, which account for both the cQMC statistical uncertainty and quasi-random deviations from the smooth trend.

FIG. 3.

Illustration of the Monte Carlo resampling scheme used to compute statistics on the intersection between two curves. For each curve, having obtained an estimate of α (width of the shaded region) from the original energy data (not shown), we create a synthetic instance of the dataset by shifting the original points by a random amount proportional to α 2 + Δ E i 2 (squares of the same color saturation), perform a quadratic fit (line), and find the intersection between both fits (circled diamond). This process is repeated over the random instances (three shown in the illustration), from which statistics on the intersection are obtained.

FIG. 3.

Illustration of the Monte Carlo resampling scheme used to compute statistics on the intersection between two curves. For each curve, having obtained an estimate of α (width of the shaded region) from the original energy data (not shown), we create a synthetic instance of the dataset by shifting the original points by a random amount proportional to α 2 + Δ E i 2 (squares of the same color saturation), perform a quadratic fit (line), and find the intersection between both fits (circled diamond). This process is repeated over the random instances (three shown in the illustration), from which statistics on the intersection are obtained.

Close modal

We demonstrate the full statistical procedure of the xspot method on multideterminant-Jastrow data for the carbon dimer in Fig. 4. Notice that the distribution of intersection points shown in the inset of Fig. 4 has a tail extending toward low E and large w. These tails become more problematic the more parallel the two intersecting curves are, eventually preventing the evaluation of an intersection point at all. It is, therefore, important to try to apply the xspot method to curves that are as close to perpendicular as possible.

FIG. 4.

VMC and DMC energies of the ground-state C2 molecule as a function of w, as shown in Fig. 1, using the full statistical treatment of the xspot method. The mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown are the estimated exact nonrelativistic energy34 as a dotted line with a shaded area of ±1 kcal/mol around it and the intersection point between the curves. The inset shows the statistical distribution of intersection points as a color map with overlaid contour curves.

FIG. 4.

VMC and DMC energies of the ground-state C2 molecule as a function of w, as shown in Fig. 1, using the full statistical treatment of the xspot method. The mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown are the estimated exact nonrelativistic energy34 as a dotted line with a shaded area of ±1 kcal/mol around it and the intersection point between the curves. The inset shows the statistical distribution of intersection points as a color map with overlaid contour curves.

Close modal

In Fig. 5, we include additional VMC and DMC data using an inhomogeneous backflow transformation (“bVMC” and “bDMC”) for the carbon dimer. We list the intersections between pairs of curves in Table III, all of which are within uncertainty of each other. The VMC and bDMC curves provide the best-resolved results, which is to be expected since these curves intersect at the widest angle among all pairs of curves, as shown in Fig. 5. By contrast, the DMC and bVMC curves intersect at a narrow angle and incur a small but non-zero fraction of “missed” intersections, i.e., random instances of the data whose fits fail to intersect at w > 1, which signals the presence of heavy tails in the intersection distribution, resulting in a large uncertainty on the intersection energy.

FIG. 5.

VMC, DMC, bVMC, and bDMC energies of the ground-state C2 molecule as a function of w. The mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown are the estimated exact nonrelativistic energy34 as a dotted line with a shaded area of ±1 kcal/mol around it and the intersection points between each of the six possible pairs of curves; error bars on these are only shown in the inset for clarity.

FIG. 5.

VMC, DMC, bVMC, and bDMC energies of the ground-state C2 molecule as a function of w. The mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown are the estimated exact nonrelativistic energy34 as a dotted line with a shaded area of ±1 kcal/mol around it and the intersection points between each of the six possible pairs of curves; error bars on these are only shown in the inset for clarity.

Close modal
TABLE III.

Location of all six pairwise intersections of the VMC, DMC, bVMC, and bDMC curves shown in Fig. 5 for the C2 molecule. “Missed intersections” refer to random instances of the curves that do not intersect at w > 1 in the Monte Carlo resampling procedure.

Curves w0 E0 (a.u.) Miss (%)
VMC-DMC  1.031(3)  −75.9288 ± 0.0014  0.00 
VMC-bVMC  1.028(5)  −75.9271 ± 0.0025  0.00 
VMC-bDMC  1.027(2)  −75.9260 ± 0.0008  0.00 
DMC-bVMC  1.05(2)  −75.9333 ± 0.0074  2.75 
DMC-bDMC  1.015(6)  −75.9242 ± 0.0012  0.00 
bVMC-bDMC  1.025(5)  −75.9258 ± 0.0012  0.00 
Curves w0 E0 (a.u.) Miss (%)
VMC-DMC  1.031(3)  −75.9288 ± 0.0014  0.00 
VMC-bVMC  1.028(5)  −75.9271 ± 0.0025  0.00 
VMC-bDMC  1.027(2)  −75.9260 ± 0.0008  0.00 
DMC-bVMC  1.05(2)  −75.9333 ± 0.0074  2.75 
DMC-bDMC  1.015(6)  −75.9242 ± 0.0012  0.00 
bVMC-bDMC  1.025(5)  −75.9258 ± 0.0012  0.00 

For the four curves in Fig. 5, the estimated amplitude of the quasi-random fluctuations ranges from α = 0.17 to 0.41 mHa. Throughout this paper, we converge the cQMC calculations so that the uncertainties satisfy ( Δ E i ) 2 < α 2 , i.e., so that quasi-random fluctuations represent the main contribution to the uncertainty on the fits and intersections; see the supplementary material for the list of values of α obtained. The statistical uncertainties on the cQMC energies can thus be neglected for all practical purposes.

As alluded to in Sec. II A, the choice of molecular orbitals plays a crucial role in the behavior of the cQMC energies as a function of w, modifying the point wexh. at which the effects of exhaustion start to become noticeable. In Fig. 6, we demonstrate this for the H2O molecule by comparing the cQMC energies obtained using Hartree–Fock orbitals expanded in the cc-pCVTZ basis set and natural orbitals expanded in the same basis set constructed so as to diagonalize the one-body density matrix in coupled cluster singles and doubles (CCSD). While CCSD natural orbitals produce lower cQMC energies and correspond to larger values of w at fixed expansion sizes, the cQMC energies we obtain using Hartree–Fock orbitals follow the quadratic trend throughout the whole w range considered, while those obtained with natural orbitals plateau very early on, preventing their meaningful extrapolation.

FIG. 6.

VMC and DMC energies of the H2O molecule as a function of w, both using Hartree–Fock orbitals (left) and CCSD natural orbitals (right) expanded in the cc-pCVTZ basis set. In both cases the same numbers of CSFs are used. Mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown are the estimated exact nonrelativistic energy35 as a dotted line with a shaded area of ±1 kcal/mol around it, and the intersection point between the VMC and DMC curves in the left panel. The cQMC energies obtained with natural orbitals plateau with w, preventing the quadratic extrapolations from reaching the exact energy.

FIG. 6.

VMC and DMC energies of the H2O molecule as a function of w, both using Hartree–Fock orbitals (left) and CCSD natural orbitals (right) expanded in the cc-pCVTZ basis set. In both cases the same numbers of CSFs are used. Mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown are the estimated exact nonrelativistic energy35 as a dotted line with a shaded area of ±1 kcal/mol around it, and the intersection point between the VMC and DMC curves in the left panel. The cQMC energies obtained with natural orbitals plateau with w, preventing the quadratic extrapolations from reaching the exact energy.

Close modal

In our calculations, we use Hartree–Fock orbitals expanded in the cc-pCVTZ Gaussian basis set32,33 obtained using molpro.36 We perform a small-scale FCIQMC calculation using the neci package37 with configuration state functions (CSFs) instead of determinants as walker sites,38 which reduces the number of FCIQMC walkers required to accurately represent the wave function; note that the use of CSFs is not a requirement of the xspot method. The FCIQMC population is grown to 106 walkers and equilibrated, and the coefficients of the Mgen = 10 000 most-occupied CSFs are recorded from a population snapshot. From this information, we build CI expansions with the M CSFs with the largest absolute coefficients, where M ≤ 1500 ≪ Mgen. In our cQMC calculations, the orbitals are corrected to obey the electron–nucleus cusp condition.39 

The CSF coefficients are reoptimized in the presence of a Jastrow factor of the Drummond–Towler–Needs form5,6 and an optional inhomogeneous backflow transformation including electron–electron, electron–nucleus, and electron–electron–nucleus terms.7 We do not optimize any of the parameters in the molecular orbitals, which provide degrees of freedom that overlap significantly with those in the backflow transformation. Note that even though CSFs are used, the presence of the Jastrow factor and the backflow transformation prevents the cQMC trial wave function from formally being an exact spin state.40 We optimize our wave function parameters using linear least-squares energy minimization3,4 with 106 statistically independent VMC-generated electronic configurations, a number large enough that the optimized cQMC energy can be assumed to lie reasonably close to its variational minimum; note that any remaining optimization error can be considered to be absorbed into the quasi-random error.

The resulting trial wave function is then used to run two DMC calculations with time steps 0.001 and 0.004 a.u. and target populations of 2048 and 512 configurations, respectively, except for bDMC runs on CO2 at 500 and 1000 CSFs, for which we use 65 536 and 16 384 configurations. These energies are then linearly extrapolated to the zero time-step, infinite-population limit.2,41

We use the casino package2 to run the cQMC calculations and use multi-determinant compression30 to reduce the computational expense of evaluating the trial wave function. We perform the fits to the data and find their intersections using our custom polyfit tool.42 The cQMC energies obtained for all systems can be found in the supplementary material.

In this section, we test the xspot method on all eight systems under consideration to assess the different aspects discussed in Sec. II and determine the broader applicability of the method. The VMC and bDMC energies and fits we obtained for the eight systems are shown in Fig. 7; additional plots containing the bVMC and DMC energies can be found in the supplementary material.

FIG. 7.

VMC and bDMC energies of the atoms and molecules considered in this work as a function of w. The mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown in each plot are the relevant benchmark energy (see details in the text and Table IV) as a dotted line with a shaded area of ±1 kcal/mol around it and the intersection point between the VMC and bDMC curves. The insets show the statistical distributions of intersection points as color maps with overlaid contour curves.

FIG. 7.

VMC and bDMC energies of the atoms and molecules considered in this work as a function of w. The mean values of the fits to the data are shown as lines, and the translucent areas around them represent 95.5% (two-sigma) confidence intervals. Also shown in each plot are the relevant benchmark energy (see details in the text and Table IV) as a dotted line with a shaded area of ±1 kcal/mol around it and the intersection point between the VMC and bDMC curves. The insets show the statistical distributions of intersection points as color maps with overlaid contour curves.

Close modal

All the curves in Fig. 7 are relatively smooth and provide a well-defined intersection. The apparent non-monotonicity of the bDMC curve for the carbon atom is an artifact of the use of a fit function that formally allows non-monotonic behavior, and should be interpreted accordingly. E(w) can be regarded as approaching the intersection with negligible slope, and the region w > w0 should be ignored since E(w) does not have a physical meaning there. All of the other fits appear to be monotonic in the range shown.

The fraction of orbitals used, a proxy for the degree of orbital-basis exhaustion, is plotted for each of the systems in Fig. 8. We do not use up all of the orbitals in the basis in any of our calculations, and the curves in Fig. 7 do not seem to exhibit symptoms of orbital-basis exhaustion. The bVMC energies do seem to plateau somewhat, which we discuss briefly in the supplementary material; note that we do not use the bVMC data to obtain our final results.

FIG. 8.

Number of distinct spatial orbitals in the FCIQMC trial wave functions relative to that in the orbital basis as a function of the number of CSFs for the various systems considered in this work.

FIG. 8.

Number of distinct spatial orbitals in the FCIQMC trial wave functions relative to that in the orbital basis as a function of the number of CSFs for the various systems considered in this work.

Close modal

In Table IV, we compare the total energies obtained from applying the xspot method to VMC and bDMC energy data with benchmark-quality estimates of the exact nonrelativistic energies of the systems from the literature, along with our best bDMC result and prior cQMC results for reference. The atomization energies of the ground-state molecules are shown in Table V. For excited-state C 2 * , we compare the vertical excitation energy with that calculated with internally contracted multi-reference coupled cluster (ic-MRCC) theory;44–46 we have computed the total energy of C 2 * shown in Table IV by adding the ic-MRCC excitation energy to the estimated ground-state energy of C2 from Ref. 34.

TABLE IV.

Total energies in Ha obtained with the xspot method using the VMC and bDMC data for the various atoms and molecules considered in this work, along with results from prior multi-determinant cQMC studies, our best individual bDMC energy for each system, and benchmark-quality nonrelativistic total energies from the literature.

System Prior cQMC Our best bDMC xspot Benchmark
−37.844 38(5)a  −37.8442(0)  −37.8439 ± 0.0003  −37.8450b 
−54.588 29(7)a  −54.5881(0)  −54.5897 ± 0.0014  −54.5893b 
−75.065 91(8)a  −75.0647(1)  −75.0664 ± 0.0004  −75.0674b 
H2 −76.438 9(1)c  −76.4336(1)  −76.4410 ± 0.0013  −76.4389d 
C2  −75.922 9(6)a  −75.9172(1)  −75.9260 ± 0.0008  −75.9265b 
C 2 *     −75.8374(1)  −75.8455 ± 0.0011  −75.8465b,e 
N2  −109.537 2(3)a  −109.5344(1)  −109.5394 ± 0.0018  −109.5425d 
CO2    −188.5837(2)  −188.5981 ± 0.0042  −188.6015c 
System Prior cQMC Our best bDMC xspot Benchmark
−37.844 38(5)a  −37.8442(0)  −37.8439 ± 0.0003  −37.8450b 
−54.588 29(7)a  −54.5881(0)  −54.5897 ± 0.0014  −54.5893b 
−75.065 91(8)a  −75.0647(1)  −75.0664 ± 0.0004  −75.0674b 
H2 −76.438 9(1)c  −76.4336(1)  −76.4410 ± 0.0013  −76.4389d 
C2  −75.922 9(6)a  −75.9172(1)  −75.9260 ± 0.0008  −75.9265b 
C 2 *     −75.8374(1)  −75.8455 ± 0.0011  −75.8465b,e 
N2  −109.537 2(3)a  −109.5344(1)  −109.5394 ± 0.0018  −109.5425d 
CO2    −188.5837(2)  −188.5981 ± 0.0042  −188.6015c 
a

Reference 18.

b

Reference 34.

c

Reference 43.

d

Reference 35.

e

References 44–46.

TABLE V.

Atomization and excitation energies in mHa of the various molecules considered in this work, corresponding to the total energies in Table IV, obtained from the xspot method, along with benchmark-quality nonrelativistic relative energies from the literature.

xspot Benchmark
H2O → 2H + O  374.6 ± 1.4  371.4a 
C2 → 2C  238.3 ± 0.9  236.5b 
C2 C 2 *   80.6 ± 1.3  80.0c 
N2 → 2N  360.1 ± 2.3  363.9a 
CO2 → C + 2O  621.4 ± 4.3  621.7a 
xspot Benchmark
H2O → 2H + O  374.6 ± 1.4  371.4a 
C2 → 2C  238.3 ± 0.9  236.5b 
C2 C 2 *   80.6 ± 1.3  80.0c 
N2 → 2N  360.1 ± 2.3  363.9a 
CO2 → C + 2O  621.4 ± 4.3  621.7a 
a

Reference 35.

b

Reference 34.

c

References 44–46.

All of the total energies reported in Table IV are within statistical uncertainty of their corresponding benchmark values. An important observation is that our individual cQMC energies are not lower than those from prior cQMC calculations, implying that we incur a lower computational cost, but our xspot results are in general closer to the benchmarks than cQMC results from prior studies. Our xspot energies are on average 1.1 standard errors above the benchmark, with a root-mean-square deviation of 1.9 standard errors. These results are compatible with the xspot method being exact when the method’s assumptions are satisfied. The relative energies are likewise in agreement with the benchmark values.

We find that the magnitude α of the quasi-random fluctuations of the cQMC energies is up to 0.7 mHa. These fluctuations are particularly visible in the VMC data for N, O, and H2O in Fig. 7, for example; in the supplementary material, we give the values of α we have obtained for each of the curves. The magnitude of the quasi-random fluctuations does not seem to increase too rapidly with system size, but their effect on the extrapolated energy becomes more pronounced the further the cQMC data are from the intersection in the plots. This increasing uncertainty on the xspot total energies, reaching 4 mHa for the CO2 molecule, hints at a limitation of the methodology: the cQMC energies and values of w obtained using modest-sized multideterminant expansions with a fixed basis set to move away from the intersection point with increasing system size, which in turn exacerbates the effects of quasi-random noise on the uncertainty of the xspot energy; one would have to use bigger basis sets and larger multideterminantal expansions to get data closer to the intersection in order to reduce this uncertainty, increasing the computational cost of the approach.

We have presented an empirical extrapolation strategy for cQMC energies as a function of the sum of the squared multideterminant coefficients in the initially selected CI wave function from which the trial wave function is constructed. This approach is made possible by the smoothness of the energies as a function of the CI expansion size, and we have presented a simple statistical procedure to handle the quasi-random non-smoothness in the data, which we show to work very well in practice. We find that Hartree–Fock orbitals expanded in standard basis sets provide the type of gradual convergence required for the xspot method to work well. The results from the tests we have conducted are compatible with the xspot method being capable of obtaining exact total energies, with the caveat that trial wave function complexity must increase with system size in order to control the uncertainty on the results.

See the supplementary material for the cQMC data used in this paper, a table of the magnitude of the quasi-random fluctuations encountered, and a discussion of connected extrapolation approaches. The supplementary material additionally cites Refs. 47–51.

P.L.R. and A.A. acknowledge the support from the European Center of Excellence in Exascale Computing, TREX, funded by the Horizon 2020 program of the European Union under Grant No. 952165. The views and opinions expressed are those of the authors only and do not necessarily reflect those of the European Union or the European Research Executive Agency. Neither the European Union nor the granting authority can be held responsible for them.

The authors have no conflicts to disclose.

Seyed Mohammadreza Hosseini: Investigation (equal); Software (equal); Writing – original draft (equal). Ali Alavi: Conceptualization (equal); Funding acquisition (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal). Pablo López Ríos: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
W. L.
McMillan
, “
Ground state of liquid He4
,”
Phys. Rev.
138
,
A442
(
1965
).
2.
R. J.
Needs
,
M. D.
Towler
,
N. D.
Drummond
,
P.
López Ríos
, and
J. R.
Trail
, “
Variational and diffusion quantum Monte Carlo calculations with the CASINO code
,”
J. Chem. Phys.
152
,
154106
(
2020
).
3.
J.
Toulouse
and
C. J.
Umrigar
, “
Optimization of quantum Monte Carlo wave functions by energy minimization
,”
J. Chem. Phys.
126
,
084102
(
2007
).
4.
C. J.
Umrigar
,
J.
Toulouse
,
C.
Filippi
,
S.
Sorella
, and
R. G.
Hennig
, “
Alleviation of the Fermion-sign problem by optimization of many-body wave functions
,”
Phys. Rev. Lett.
98
,
110201
(
2007
).
5.
N. D.
Drummond
,
M. D.
Towler
, and
R. J.
Needs
, “
Jastrow correlation factor for atoms, molecules, and solids
,”
Phys. Rev. B
70
,
235119
(
2004
).
6.
P.
López Ríos
,
P.
Seth
,
N. D.
Drummond
, and
R. J.
Needs
, “
Framework for constructing generic Jastrow correlation factors
,”
Phys. Rev. E
86
,
036703
(
2012
).
7.
P.
López Ríos
,
A.
Ma
,
N. D.
Drummond
,
M. D.
Towler
, and
R. J.
Needs
, “
Inhomogeneous backflow transformations in quantum Monte Carlo calculations
,”
Phys. Rev. E
74
,
066701
(
2006
).
8.
D. M.
Ceperley
and
B. J.
Alder
, “
Ground state of the electron gas by a stochastic method
,”
Phys. Rev. Lett.
45
,
566
(
1980
).
9.
G. H.
Booth
,
A. J. W.
Thom
, and
A.
Alavi
, “
Fermion Monte Carlo without fixed nodes: A game of life, death, and annihilation in Slater determinant space
,”
J. Chem. Phys.
131
,
054106
(
2009
).
10.
D.
Cleland
,
G. H.
Booth
, and
A.
Alavi
, “
Communications: Survival of the fittest: Accelerating convergence in full configuration-interaction quantum Monte Carlo
,”
J. Chem. Phys.
132
,
041103
(
2010
).
11.
M.
Ruggeri
,
P. L.
Ríos
, and
A.
Alavi
, “
Correlation energies of the high-density spin-polarized electron gas to meV accuracy
,”
Phys. Rev. B
98
,
161105(R)
(
2018
).
12.
A. J.
Cohen
,
H.
Luo
,
K.
Guther
,
W.
Dobrautz
,
D. P.
Tew
, and
A.
Alavi
, “
Similarity transformation of the electronic Schrödinger equation via Jastrow factorization
,”
J. Chem. Phys.
151
,
061101
(
2019
).
13.
J. P.
Haupt
,
S. M.
Hosseini
,
P.
López Ríos
,
W.
Dobrautz
,
A.
Cohen
, and
A.
Alavi
, “
Optimizing Jastrow factors for the transcorrelated method
,”
J. Chem. Phys.
158
,
224105
(
2023
).
14.
C.
Filippi
and
C. J.
Umrigar
, “
Multiconfiguration wave functions for quantum Monte Carlo calculations of first-row diatomic molecules
,”
J. Chem. Phys.
105
,
213
(
1996
).
15.
M. D.
Brown
,
J. R.
Trail
,
P.
López Ríos
, and
R. J.
Needs
, “
Energies of the first row atoms from quantum Monte Carlo
,”
J. Chem. Phys.
126
,
224110
(
2007
).
16.
P.
Seth
,
P. L.
Ríos
, and
R. J.
Needs
, “
Quantum Monte Carlo study of the first-row atoms and ions
,”
J. Chem. Phys.
134
,
084105
(
2011
).
17.
F. R.
Petruzielo
,
J.
Toulouse
, and
C. J.
Umrigar
, “
Approaching chemical accuracy with quantum Monte Carlo
,”
J. Chem. Phys.
136
,
124116
(
2012
).
18.
M. A.
Morales
,
J.
McMinis
,
B. K.
Clark
,
J.
Kim
, and
G. E.
Scuseria
, “
Multideterminant wave functions in quantum Monte Carlo
,”
J. Chem. Theory Comput.
8
,
2181
(
2012
).
19.
E.
Giner
,
R.
Assaraf
, and
J.
Toulouse
, “
Quantum Monte Carlo with reoptimised perturbatively selected configuration-interaction wave functions
,”
Mol. Phys.
114
,
910
(
2016
).
20.
A.
Scemama
,
A.
Benali
,
D.
Jacquemin
,
M.
Caffarel
, and
P.-F.
Loos
, “
Excitation energies from diffusion Monte Carlo using selected configuration interaction nodes
,”
J. Chem. Phys.
149
,
034108
(
2018
).
21.
A.
Scemama
,
M.
Caffarel
,
A.
Benali
,
D.
Jacquemin
, and
P.-F.
Loos
, “
Influence of pseudopotentials on excitation energies from selected configuration interaction and diffusion Monte Carlo
,”
Results Chem.
1
,
100002
(
2019
).
22.
M.
Dash
,
J.
Feldt
,
S.
Moroni
,
A.
Scemama
, and
C.
Filippi
, “
Excited states with selected configuration interaction-quantum Monte Carlo: Chemically accurate excitation energies and geometries
,”
J. Chem. Theory Comput.
15
,
4896
(
2019
).
23.
M.
Dash
,
S.
Moroni
,
C.
Filippi
, and
A.
Scemama
, “
Tailoring CIPSI expansions for QMC calculations of electronic excitations: The case study of thiophene
,”
J. Chem. Theory Comput.
17
,
3426
(
2021
).
24.
M.
Dash
,
S.
Moroni
,
A.
Scemama
, and
C.
Filippi
, “
Perturbatively selected configuration-interaction wave functions for efficient geometry optimization in quantum Monte Carlo
,”
J. Chem. Theory Comput.
14
,
4176
(
2018
).
25.
A.
Scemama
,
T.
Applencourt
,
E.
Giner
, and
M.
Caffarel
, “
Quantum Monte Carlo with very large multideterminant wavefunctions
,”
J. Comput. Chem.
37
,
1866
(
2016
).
26.
M. C.
Per
and
D. M.
Cleland
, “
Energy-based truncation of multi-determinant wavefunctions in quantum Monte Carlo
,”
J. Chem. Phys.
146
,
164101
(
2017
).
27.
A.
Scemama
,
Y.
Garniron
,
M.
Caffarel
, and
P.-F.
Loos
, “
Deterministic construction of nodal surfaces within quantum Monte Carlo: The case of FeS
,”
J. Chem. Theory Comput.
14
,
1395
(
2018
).
28.
P. K. V. V.
Nukala
and
P. R. C.
Kent
, “
A fast and efficient algorithm for Slater determinant updates in quantum Monte Carlo simulations
,”
J. Chem. Phys.
130
,
204105
(
2009
).
29.
B. K.
Clark
,
M. A.
Morales
,
J.
McMinis
,
J.
Kim
, and
G. E.
Scuseria
, “
Computing the energy of a water molecule using multideterminants: A simple, efficient algorithm
,”
J. Chem. Phys.
135
,
244105
(
2011
).
30.
G. L.
Weerasinghe
,
P. L.
Ríos
, and
R. J.
Needs
, “
Compression algorithm for multideterminant wave functions
,”
Phys. Rev. E
89
,
023304
(
2014
).
31.
C.
Filippi
,
R.
Assaraf
, and
S.
Moroni
, “
Simple formalism for efficient derivatives and multi-determinant expansions in quantum Monte Carlo
,”
J. Chem. Phys.
144
,
194105
(
2016
).
32.
T. H.
Dunning
, Jr.
, “
Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen
,”
J. Chem. Phys.
90
,
1007
(
1989
).
33.
D. E.
Woon
and
T. H.
Dunning
, Jr.
, “
Gaussian basis sets for use in correlated molecular calculations. V. Core-valence basis sets for boron through neon
,”
J. Chem. Phys.
103
,
4572
(
1995
).
34.
L.
Bytautas
and
K.
Ruedenberg
, “
Correlation energy extrapolation by intrinsic scaling. IV. Accurate binding energies of the homonuclear diatomic molecules carbon, nitrogen, oxygen, and fluorine
,”
J. Chem. Phys.
122
,
154110
(
2005
).
35.
D.
Feller
,
K. A.
Peterson
, and
D. A.
Dixon
, “
A survey of factors contributing to accurate theoretical predictions of atomization energies and molecular structures
,”
J. Chem. Phys.
129
,
204105
(
2008
).
36.
H.-J.
Werner
,
P. J.
Knowles
,
F. R.
Manby
,
J. A.
Black
,
K.
Doll
,
A.
Heßelmann
,
D.
Kats
,
A.
Köhn
,
T.
Korona
,
D. A.
Kreplin
,
Q.
Ma
,
T. F.
Miller
III
,
A.
Mitrushchenkov
,
K. A.
Peterson
,
I.
Polyak
,
G.
Rauhut
, and
M.
Sibaev
, “
The Molpro quantum chemistry package
,”
J. Chem. Phys.
152
,
144107
(
2020
).
37.
K.
Guther
,
R. J.
Anderson
,
N. S.
Blunt
,
N. A.
Bogdanov
,
D.
Cleland
,
N.
Dattani
,
W.
Dobrautz
,
K.
Ghanem
,
P.
Jeszenszki
,
N.
Liebermann
,
G. L.
Manni
,
A. Y.
Lozovoi
,
H.
Luo
,
D.
Ma
,
F.
Merz
,
C.
Overy
,
M.
Rampp
,
P. K.
Samanta
,
L. R.
Schwarz
,
J. J.
Shepherd
,
S. D.
Smart
,
E.
Vitale
,
O.
Weser
,
G. H.
Booth
, and
A.
Alavi
, “
NECI: N-electron configuration interaction with an emphasis on state-of-the-art stochastic methods
,”
J. Chem. Phys.
153
,
034107
(
2020
).
38.
W.
Dobrautz
,
S. D.
Smart
, and
A.
Alavi
, “
Efficient formulation of full configuration interaction quantum Monte Carlo in a spin eigenbasis via the graphical unitary group approach
,”
J. Chem. Phys.
151
,
094104
(
2019
).
39.
A.
Ma
,
M. D.
Towler
,
N. D.
Drummond
, and
R. J.
Needs
, “
Scheme for adding electron-nucleus cusps to Gaussian orbitals
,”
J. Chem. Phys.
122
,
224322
(
2005
).
40.
C.-J.
Huang
,
C.
Filippi
, and
C. J.
Umrigar
, “
Spin contamination in quantum Monte Carlo wave functions
,”
J. Chem. Phys.
108
,
8838
(
1998
).
41.
R. M.
Lee
,
G. J.
Conduit
,
N.
Nemec
,
P.
López Ríos
, and
N. D.
Drummond
, “
Strategies for improving the efficiency of quantum Monte Carlo calculations
,”
Phys. Rev. E
83
,
066706
(
2011
).
42.
See https://github.com/plopezrios/polyfit for our custom regression tool can be found online.
43.
M.
Caffarel
,
T.
Applencourt
,
E.
Giner
, and
A.
Scemama
, “
Communication: Toward an improved control of the fixed-node error in quantum Monte Carlo: The case of the water molecule
,”
J. Chem. Phys.
144
,
151103
(
2016
).
44.
P. K.
Samanta
, private communication (2018).
45.
P. K.
Samanta
,
D.
Mukherjee
,
M.
Hanauer
, and
A.
Köhn
, “
Excited states with internally contracted multireference coupled-cluster linear response theory
,”
J. Chem. Phys.
140
,
134108
(
2014
).
46.
M.
Hanauer
and
A.
Köhn
, “
Pilot applications of internally contracted multireference coupled cluster theory, and how to choose the cluster operator properly
,”
J. Chem. Phys.
134
,
204111
(
2011
).
47.
A. A.
Holmes
,
C. J.
Umrigar
, and
S.
Sharma
, “
Excited states using semistochastic heat-bath configuration interaction
,”
J. Chem. Phys.
147
,
164111
(
2017
).
48.
H. G. A.
Burton
and
P.-F.
Loos
, “
Rationale for the extrapolation procedure in selected configuration interaction
,”
J. Chem. Phys.
160
,
104102
(
2024
).
49.
D. M.
Ceperley
, “
The statistical error of Green’s function Monte Carlo
,”
J. Stat. Phys.
43
,
815
(
1986
).
50.
M.
Taddei
,
M.
Ruggeri
,
S.
Moroni
, and
M.
Holzmann
, “
Iterative backflow renormalization procedure for many-body ground-state wave functions of strongly interacting normal Fermi liquids
,”
Phys. Rev. B
91
,
115106
(
2015
).
51.
W.
Fu
,
W.
Ren
, and
J.
Chen
, “
Variance extrapolation method for neural-network variational Monte Carlo
,”
Mach. Learn.: Sci. Technol.
5
,
015016
(
2024
).