We investigate the exact full configuration interaction quantum Monte Carlo algorithm (without the initiator approximation) applied to weak sign-problem fermionic systems, namely, systems in which the energy gap to the corresponding sign-free or “stoquastized” state is small. We show that the minimum number of walkers required to exactly overcome the sign problem can be significantly reduced via an importance-sampling similarity transformation even though the similarity-transformed Hamiltonian has the same stoquastic gap as the untransformed one. Furthermore, we show that in the off-half-filling Hubbard model at U/t = 8, the real-space (site) representation has a much weaker sign problem compared to the momentum space representation. By applying importance sampling using a Gutzwiller-like guiding wavefunction, we are able to substantially reduce the minimum number of walkers in the case of 2 × ℓ Hubbard ladders, enabling us to get exact energies for sizable ladders. With these results, we calculate the fundamental charge gap ΔEfund = E(N + 1) + E(N − 1) − 2E(N) for the ladder systems compared to strictly one-dimensional Hubbard chains and show that the ladder systems have a reduced fundamental gap compared to the 1D chains.
I. INTRODUCTION
The fermionic sign problem is one of the most fundamental limitations of quantum Monte Carlo (QMC) methods.1,2 In projector QMC methods, where the aim is to project out a specific state—usually the ground state—through a stochastic evolution of signed walkers, the problem manifests itself through the exponential growth of an undesired distribution of walkers of “bosonized” or “stoquastized” character, leaving the desired fermionic distribution lost in the noise of the simulation. This instability leads to diverging variances in expectation values of measured observables.
There are three known strategies to counter this problem. The first is to place constraints on the evolving distribution of walkers to prevent the growth of the undesired distribution. The fixed-node approximation of Diffusion Monte Carlo (DMC),3–5 the constrained path and phaseless approximations of auxiliary-field quantum Monte Carlo (AFQMC),6–12 and the initiator approximation13 of the full configuration interaction quantum Monte Carlo (FCIQMC) method14 all fall into this category. The difficulty is that these approximations generally lead to biased results (unless an exact form of the constraint, e.g., via a symmetry, can be applied), and furthermore, these approximations are either uncontrolled or can only be controlled (reduced) through time-consuming protocols, such as multi-determinant trial wavefunctions in the fixed-node and phaseless approximations, or by increasing the walker number in the initiator approximation. These methods are ultimately scaling exponentially in the computational resources needed to achieve a desired accuracy. The second strategy is to perform unconstrained projections, attempting to measure a transient fermionic signal before the undesired solution becomes dominant. Such simulations are practical only for very small systems and otherwise cost exponential time. The third strategy, which is advocated in the FCIQMC method (without the initiator approximation), is to introduce walker annihilation processes into the algorithm, which destabilize the undesired solution15 and allow the desired fermionic solution to emerge without bias. The difficulty here is that a minimum number of walkers is required to achieve this, and this minimum number is generally found to scale with the size of the Hilbert space. This makes the algorithm exponentially costly in memory, though usually with a small prefactor. When exact FCIQMC is applicable, the quality of the sampled wavefunction is comparable to that obtained by exact diagonalization, but with a lower memory requirement.
When FCIQMC is used with a small number of walkers, the unphysical stoquastic solution becomes dominant and the fermionic solution is lost in noise. The stability of the fermionic solution can be quantified by the energy gap between the exact fermionic ground-state and that of the undesired solution.15 However, as we will show in this paper, the stoquastized gap alone does not fully describe the computational difficulty to overcome the sign problem. In particular, we will show that two Hamiltonians with exactly the same gap can, nevertheless, require very different minimal numbers of walkers to converge. We show that guiding wavefunctions can be used to further mitigate the sign problem by reducing the minimum of walkers required to fully resolve the sign problem, thereby arriving at an exact scheme with lower computational cost. (Previously, it has been found that by employing a trial wavefunction, a “fixed-node Hamiltonian” can be defined by which this stoquastized gap can be reduced at the expense of introducing an approximation to the Hamiltonian.)16–19
If a Hamiltonian in a particular basis has no sign problem, then its stoquastized gap is identically zero; conversely, if this gap is found to be zero, then that Hamiltonian has no sign problem. A weak sign problem is one in which the gap is non-zero, but small. There are some systems, such as periodic 1D Hubbard models, in which the gap remains small even as the system size grows (i.e., it is not size-extensive). In these cases, the FCIQMC algorithm can be applied to very large systems, essentially accessing the thermodynamic limit. In other systems, the gap grows with system size, ultimately limiting the system sizes, which can be treated. Even here, one can still obtain exact-diagonalization quality results for systems far exceeding conventional exact diagonalization.
A characteristic of weak sign-problem fermionic systems is that their wavefunctions tend to be highly spread out throughout the Hilbert space, somewhat similar to the related stoquastized systems (in which all imaginary-time Feynman path contributions are strictly positive, leading to diffuse, non-compact wavefunctions). For such systems, methods based on truncations of the Hilbert space will lead to large errors. Therefore, techniques such as the initiator approximation of FCIQMC, or similar selected CI methods of quantum chemistry, struggle to describe such systems. Paradoxically, therefore, weak sign-problem systems are highly challenging systems, and it is necessary to devise alternative techniques—not based on truncations—to address them. The FCIQMC method without the initiator approximation is one such method, but even there, it needs to be adapted to efficiently yet faithfully sample the exact fermionic ground-state. Devising such a method is the main purpose of the present study.
It is an interesting fact that not every fermionic quantum system has a sign problem in FCIQMC, even in non-trivial cases. In some systems, it turns out that it is possible to subdivide the Hilbert space of the problem in such a way that, during the propagation step of the FCIQMC algorithm, all walkers that arrive at a given point in the Hilbert space always have the same sign, even though there are both types of walkers distributed throughout the space. Such a case occurs in one-dimensional Hubbard models with open boundary conditions, irrespective of the filling, as long as one uses a real-space basis.20
The Hamiltonian for a Hubbard chain of length ℓ is given by
and are the fermionic creation and annihilation operators at lattice site i with electron spin σ, respectively. are the corresponding number operators. t is the hopping amplitude, and U is the on-site interaction. The Hubbard model is a seemingly simple lattice model but yet already shows the rich physical properties of more realistic system setups21–23 and has been studied with a variety of different computational methods.24
For open boundary conditions—sites ℓ and 1 are not regarded as neighbors—the system is sign-problem-free for all fillings. For periodic (anti-periodic) boundary conditions—sites ℓ and 1 are connected with hopping amplitude −t (t)—there is no sign problem for an odd (even) number of up-electrons and an odd (even) number of down-electrons.
In these systems, calculations of systems with up to 150 sites in the case of the one-dimensional Hubbard chains have been shown to be calculable with FCIQMC.20 In such large spaces, another systematic bias called population control bias, which is independent of the sign problem, becomes notable. It has been shown that the application of importance sampling in combination with an a posteriori reweighing procedure can practically remove the bias.
When moving away from the filling rules in the periodic 1D Hubbard model mentioned previously or moving to 2D Hubbard systems, there is an inevitable sign problem. However, we will show that these systems still show only a weak sign problem when using a real-space basis, meaning that the number of FCIQMC walkers to fully resolve the sign problem is very small compared to the total Hilbert space size.
To illustrate that in a real-space basis the sign problem of the Hubbard model is comparatively weak, we calculated ΔEstoq of two half-filled Hubbard 4 × 4 systems. For U/t = 8, in a real-space basis, we obtain ΔEstoq = 3.792. In a reciprocal-space basis, we get 97.201, which means an increase of almost two orders of magnitude. These numbers are also shown in Fig. 2. Even for U/t = 4—where the k-space representation typically is much more suitable25—ΔEstoq in the real-space basis is still much lower than in the reciprocal-space basis (8.401 compared to 40.720).
The sign problem is especially weak if the lattice geometry still resembles the geometry of a sign-problem-free 1D system, such as a two-legged ladder. Still, the minimum number of walkers may be too high to be affordable.
In this paper, we analyze the FCIQMC algorithm applied to weak sign-problem fermionic systems, epitomized by Hubbard models in a real-space basis in periodic 1D and ladder geometry (see Fig. 1). In such problems, the initiator approximation is severe, while the exact algorithm can be applied affordably using importance sampling. We observe a significant reduction of the minimum number of walkers in weak-sign-problem systems. We use the method to report unbiased results for the interaction regime of U/t = 8 at and close to half-filling.
Ladder-like lattice geometry. Dashed lines indicate periodic boundary conditions.
Ladder-like lattice geometry. Dashed lines indicate periodic boundary conditions.
II. THE SIGN PROBLEM IN FCIQMC
A. Overview
FCIQMC is an imaginary-time projection method to (in its simplest formulation) sample the ground state wavefunction |Ψ0⟩ of a Hamiltonian by solving the stationary Schrödinger equation.14 The master equation is given by
with S being a scalar parameter, the shift. Starting from an initial wavefunction |Ψ(0)⟩ that has non-zero overlap with the true ground state, |Ψ(τ)⟩ will converge to the ground state in the large-τ limit.
FCIQMC applies this master equation in a finite basis of Slater determinants (although recently FCIQMC has been extended to a spin-adapted version working in a basis of configuration state functions26,27). The evolving wavefunction coefficients are represented by a number of walkers Ni that reside on a specific Slater determinant |Di⟩. is, then, stochastically applied to the instantaneous walker distribution in discretized time steps Δτ as
The stochastic application of consists of three steps: spawning, death/cloning, and annihilation. In the spawning step, each walker attempts to spawn from the determinant |Di⟩ it is sitting on to a determinant |Dj⟩ that it is connected to by a non-zero Hamiltonian matrix element Hji. The spawning attempt succeeds with a probability , where is the probability that |Dj⟩ is selected as a target determinant starting from |Di⟩. In the death/cloning step, the walker population Ni of every occupied determinant is multiplied by Δτ(Hii − S). If after this step the walker population is lower than a certain minimum occupation threshold (mostly chosen to be 1), the population is stochastically rounded either up to the occupation threshold or down to 0. If it is rounded to 0, the determinant is removed from memory until a new spawn happens onto this particular determinant. This makes the instantaneous representation of the FCIQMC wavefunction very memory-efficient. The annihilation step is an important step for reducing the sign problem. As the last step of an iteration, all spawns to the same target determinant are communicated to the same processor. There, positive and negative spawns cancel each other. This is easily possible since FCIQMC works in a discrete basis of Slater determinants unlike, e.g., diffusion Monte Carlo (DMC) where such immediate cancellations are not straightforwardly possible and, therefore, of limited effectiveness.5
The ground-state energy can be measured by averaging either the projected energy or the shift. The projected energy is defined as
where |D0⟩ is the reference determinant (which in real-space Hubbard systems close to half-filling is always a Néel determinant). The shift is not only used to control the population but can also be used as an energy estimator. It is updated according to
every A iterations. The third term is an extension to the usual shift update formula and was presented in Ref. 28. The first logarithmic term just leads to a constant population. The second term has the desired population Nw in the denominator so that the average total population will converge to that value precisely. The prefactor γ2/4 ensures critical damping in a simplified model and still proves useful in actual simulations.
Typically, in ab initio systems, FCIQMC is applied in conjunction with the initiator approximation13 and has been successfully applied in various ab initio systems,29–32 as well as the Hubbard model in a reciprocal-space basis for U/t ≤ 4.25,33 In the initiator method, determinants are dynamically marked as initiators during the simulation when their walker population exceeds the preset initiator threshold tinit. Only initiators are allowed to spawn to every other determinant since they are deemed to be sign-coherent because of their well-established population of a single sign. Non-initiators can only passively acquire population and spawn to occupied determinants only, with the rare exception that two simultaneous non-initiator spawns on an empty determinant are also allowed. This leads to an effective, dynamically updated truncation of the Hamiltonian. This introduces a bias into the sampled wavefunction and energy estimators. The initiator approximation is not suitable for treating the real-space Hubbard model for U/t ≲ 16 where the wavefunction is strongly spread-out across the Hilbert space. Even for tinit close or equal to the occupation threshold, the initiator truncation becomes so severe that the energy estimates are very slowly converging with respect to Nw, even when using improved formulation of the initiator method, such as the adaptive-shift method.34,35 Therefore, in the following, we will discuss calculations and algorithmic improvement within non-initiator FCIQMC only.
B. Quantifying the sign problem
In the absence of any annihilations, the FCIQMC wavefunction would collapse to the ground state of the so-called “stoquastized” version of the original Hamiltonian.15 The matrix elements of the stoquastized Hamiltonian are defined as
In other words, the matrix elements of have the same magnitude as , but the matrix is a sign-free version of it. It is easy to see that the lowest eigenvalue of is always lower than or equal to the lowest eigenvalue E0 of .36 The larger the stoquastized gap , the larger is the tendency of the shift to collapse to the ground state with eigenvalue instead of the sought-for true fermionic ground state with eigenvalue E0. On the same note, the overlap with any reasonable trial wavefunction vanishes because the fermionic solution is exponentially suppressed. This also makes the determination of the ground-state energy using projected energy estimators impossible.
ΔEstoq is an important concept to quantify the sign problem. It can be regarded as a global quantity that quantifies how far the stoquastized and fermionic differ. However, there is no monotonic relation between the stoquastized gap and the actual computational effort (i.e., the minimum number of walkers required) to solve a system with FCIQMC. In a practical simulation, the actual sampling of the wavefunction is also important. This fact will be demonstrated and exploited later in this paper.
To render the correct fermionic solution of a given sign-problematic Hamiltonian with FCIQMC, the Hilbert space has to be occupied with a minimum number of walkers simultaneously. This is such that the necessary annihilations can actually take place and the fermionic solution can emerge. Still, the minimum number of walkers Nmin to get the numerically exact ground-state solution and overcome the sign problem is typically smaller than the number of determinants in the Hilbert space. The minimum number of walkers is also called the annihilation plateau since in most cases, it can be identified by a plateau in the walker number in the walker growth phase, caused by walker annihilations. However, especially in the real-space Hubbard systems studied in this paper, the plateau can be hard to identify. Therefore, we use a different method to determine Nmin, which we call the fixed-N0 method. In this method, the population on the reference determinant is held fixed, rather than the total population. In order to achieve this, the shift is updated according to
N0(τ) is the population of the reference determinant at imaginary time τ and j runs over all connected determinants, including j = 0. The shift is now equivalent to Eproj, and there is only a single energy estimator. Plugging Eq. (7) into Eq. (3) leads to
so the population on the reference determinant is held exactly constant at a specified value N0 and especially at a given sign. This ensures sign coherence automatically at the expense of a total walker population Nw that is not predefined. For N0 = 1, the averaged total population would converge to Nmin exactly. Choosing N0, this low, typically, leads to large fluctuations of Nw, however, making it difficult to average. Therefore, in all calculations, in this paper, we will use N0 = 50, which will lead to slightly higher than Nmin.37
III. HUBBARD SYSTEMS WITH WEAK SIGN PROBLEMS
As mentioned in the Introduction, the 1D Hubbard model is sign-problem-free for open boundary conditions for all fillings and for periodic boundary conditions for an odd number of up-electrons and odd number of down-electrons. A proof for this is given in Appendix A.
Moving away from these system setups to different fillings or going to 2D will inevitably introduce a sign problem. The strength of the sign problem, however, strongly depends on the basis representation, the geometry of the lattice, and the filling.
As can be seen in Fig. 2, the sign problem in a real-space basis is typically much weaker than comparable systems in a reciprocal space basis, i.e., ΔEstoq is an order of magnitude smaller.
Gaps between fermionic and stoquastized ground states for various Hubbard models at U/t = 8 in real-space and momentum space representations. ΔEstoq is about an order of magnitude larger for reciprocal space systems compared to real-space representations, making the real-space Hubbard model relatively weakly sign-problematic. Furthermore, it can be seen that the ladder systems (i.e., 2 × ℓ geometries) have a smaller stoquastized gap compared to the square geometries of the same size. In addition, half-filled (hf) systems have smaller stoquastized gaps than systems with one hole, while the gaps of the latter increase more slowly with the increasing system size.
Gaps between fermionic and stoquastized ground states for various Hubbard models at U/t = 8 in real-space and momentum space representations. ΔEstoq is about an order of magnitude larger for reciprocal space systems compared to real-space representations, making the real-space Hubbard model relatively weakly sign-problematic. Furthermore, it can be seen that the ladder systems (i.e., 2 × ℓ geometries) have a smaller stoquastized gap compared to the square geometries of the same size. In addition, half-filled (hf) systems have smaller stoquastized gaps than systems with one hole, while the gaps of the latter increase more slowly with the increasing system size.
A. Non-extensive and extensive sign problems
We next analyze the sign problem of 1D systems with even n↑ or n↓ with periodic boundary conditions. According to Eqs. (A3) and (A4), there are non-negative matrix elements in the Hamiltonian; however, the sign problem is non-size-extensive. This means that the sign problem is actually weakening with increasing system size.
To understands this, let us see how the sign problem emerges. Let be the amplitude to propagate a walker on |Di⟩ to |Dj⟩ in one time-step of Δτ. Then, the amplitude to propagate a walker from |Di⟩ to |Dj⟩ after K such steps is given by the sum over all possible Feynman paths of length K, which start at i and end at j,
For ground-state problems, we are interested in this quantity in the K → ∞ limit, and we will refer to this transition amplitude as
Given the stoquastized problem, the amplitude to propagate a walker from |Di⟩ to |Dj⟩ is given by
with the corresponding ground-state quantity being
The stoquastized quantities and are by definition non-negative. A sign problem, i.e., a gap between the fermionic and stoquastized ground states, is created when there are at least two paths in with different signs, leading to
The maximal sign problem is created when there is full destructive interference in the fermionic contribution, i.e., Cji = 0. Let us now consider what happens in different geometries.
1. 1D case
In the 1D case, the number of pathways between two determinants is severely limited due to the lattice geometry. In open boundary conditions, there is only one essential pathway—thus no sign problem—in periodic boundary conditions, there are only two essential pathways that could destructively interfere. The sign problem measure is reduced if the chain length ℓ is increased or if the on-site interaction U/t is increased. An illustrative explanation for these relations is given in Appendix B. This leads to a non-extensive sign problem for 1D Hubbard systems, which is shown in Fig. 3(a) in terms of the stoquastic gap as a function of the chain length. For systems with over 100 sites, the sign problem becomes negligible even off-half-filling and with periodic boundary conditions as , then, becomes equal to E0 within statistical errors. Thus, systems of this size can be solved with less than 108 walkers when accounting for the population control bias.
Gaps between the fermionic and stoquastized solution for a 1D chain and a 2D ladder. For the chains, the gap is 0 for all systems with odd n↑ and odd n↓. It is not size-extensive for all other fillings, and ΔEstoq decreases with system size. The sign problem in the 2 × ℓ ladder case is size-extensive, i.e., ΔEstoq increases when the length of the ladder system is increased. However, at U/t = 16, this increase is rather slow, whereas at U/t = 8, it is much faster. This implies that the large-U limit of the ladder systems remains weakly sign-problematic even for large systems. On the other hand, the low U limit becomes increasingly sign problematic in the real-space representation.
Gaps between the fermionic and stoquastized solution for a 1D chain and a 2D ladder. For the chains, the gap is 0 for all systems with odd n↑ and odd n↓. It is not size-extensive for all other fillings, and ΔEstoq decreases with system size. The sign problem in the 2 × ℓ ladder case is size-extensive, i.e., ΔEstoq increases when the length of the ladder system is increased. However, at U/t = 16, this increase is rather slow, whereas at U/t = 8, it is much faster. This implies that the large-U limit of the ladder systems remains weakly sign-problematic even for large systems. On the other hand, the low U limit becomes increasingly sign problematic in the real-space representation.
On the same note, this leads to the counterintuitive situation that for sign-problematic 1D systems of intermediate size (e.g., around 50 sites), where is still smaller than E0 outside of statistical errors, is much harder or even impossible to solve. Even if the non-size-extensive gap is already very small at this system size, due to the very large Hilbert space size, a significant or even unreachable number of walkers would be needed to establish a sign-coherent solution.
2. 2D case
In the 2D case for a square lattice, the length of the elementary sign-problematic cycles does not change with the increasing system size. It always consists of a four-site plaquette. Therefore, the special circumstances that cause the non-extensivity of the sign problem in the 1D case no longer exist here. In addition to the elementary cycles, adding more four-site plaquettes even increases the number of sign-problematic cycles. This leads to an extensive sign problem, which means that the stoquastic gaps increase with system size. This is shown in Fig. 3(b). This severely limits the treatable system size. Depending on the exact shape of the lattice, however, the sign problem can—even though it is size-extensive—still be weak and systems well beyond the capability of exact diagonalization can be treated numerically exactly when FCIQMC is enhanced by an importance sampling scheme using a simple guiding wavefunction. An example for a Hubbard system class with a very weak sign problem is the ladder systems, shown in Fig. 1, which are 2 × ℓ Hubbard lattices with a 2 × 2 primitive cell. Ladder systems have been studied mostly using the density-matrix renormalization group (DMRG) as they still show a relatively low entanglement entropy, making it possible to calculate larger clusters.38
We end this section by noting that our finding that the real-space representation results in a weaker sign problem compared to the momentum space representation is in line with a similar conclusion arrived at by Blunt, in a study of fixed-node and partial-node FCIQMC,19 for ab initio systems. There, he found that fully localized basis sets tend to lead to weaker sign problems (as measured by the population plateau), compared to split-localized and canonical Hartree–Fock or natural orbitals. Taken together with our results, we believe this finding may be quite general, namely, real-space or localized basis sets—while leading to highly spread-out, non-compact, wavefunctions—also lead to the weakest sign problems. Put another way, we hypothesize that orbital representations that lead to compact wavefunctions tend also to lead to severe sign problems, implying that the observed compactness may be due to large amounts of annihilation.
We support this by the following argument: Suppose that we perform a deterministic simulation using the FCIQMC master equation [Eq. (3)]. At equilibrium, the net contribution onto |Di⟩ is ΔNi(τ) = 0. When all other Nj equal their respective true Cj coefficient, Ni is then given by
Ni will be small if (i) Hii is large, (ii) all connected Nj or their respective matrix elements Hji are small in magnitude, or (iii) there are a lot of canceling contributions from connected determinants |Dj⟩ because of opposite signs. Case (iii) leads to a large contribution to the sign problem by |Di⟩ even though it might be unimportant in the FCI expansion. This indicates that compact wavefunctions due to the orbital representation can on the same note cause a more severe sign problem.
This highlights the general challenge we face, namely, to successfully overcome the sign problem, we may need to abandon the notion of seeking orbital representations that lead to maximally compact wavefunctions.
IV. REDUCING THE MINIMUM NUMBER OF WALKERS BY IMPORTANCE SAMPLING
ΔEstoq can be modified by changing the basis of the many-particle Hamiltonian. However, finding and applying these many-particle basis rotations that could potentially reduce the gap, in general, are equally hard or even harder than overcoming the sign problem itself.39,40 Single-particle basis rotations can reduce the sign problem, but this can be used only in very limited cases.41
For the FCIQMC method, however, we find empirically that ΔEstoq is not the only defining factor on how difficult it is to overcome the sign problem. Since the number of walkers ultimately determines the necessary computational resources for an FCIQMC simulation, the minimum number of walkers Nmin to obtain the sign-coherent solution is the important algorithmic quantity. We also define the ratio as the FCIQMC-related strength of the sign problem. Examples of the relationship between s and Nmin with respect to ΔEstoq are shown in Fig. 4.
The minimum number of walkers, both in absolute numbers Nmin and relative to the Hilbert space size , as a function of the stoquastized gap for various Hubbard systems in the real-space representation at half-filling (hf) and with one hole at U/t = 8. s is a measure of the FCIQMC-related strength of the sign problem. It can be seen that although Nmin increases with system size for a given type of system, s decreases for the same systems.
The minimum number of walkers, both in absolute numbers Nmin and relative to the Hilbert space size , as a function of the stoquastized gap for various Hubbard systems in the real-space representation at half-filling (hf) and with one hole at U/t = 8. s is a measure of the FCIQMC-related strength of the sign problem. It can be seen that although Nmin increases with system size for a given type of system, s decreases for the same systems.
We find that s is a monotonic function of ΔEstoq when moving from half-filling to one hole for a given system size. Increasing the length of the ladder will, however, increase ΔEstoq but decrease s. On the other hand, when looking purely at Nmin, we find that it behaves exponentially with respect to ΔEstoq for a given filling. The one-hole systems, however, systematically have a higher Estoq even though a larger but half-filled system might have a larger Nmin.
Therefore, we conclude that the exact shape of the wavefunction also influences how computationally expensive it is to overcome the FCIQMC sign problem. Thus, we investigate how guiding wavefunctions that leave ΔEstoq unchanged influence the strength of the FCIQMC sign problem.
A. Gutzwiller-like guiding wavefunction
Previously, we have seen that the FCIQMC-related strength of the sign problem s does not only depend on ΔEstoq but is also system-dependent (see Fig. 3). Hence, instead of modifying ΔEstoq, we can also make the wavefunction more compact by applying a simple diagonal guiding wavefunction |Ψg⟩. Here, we will use a Gutzwiller-like wavefunction ansatz.
It is common practice in various QMC methods to use a guiding wavefunction to improve convergence and reduce stochastic fluctuations.42–45 We define the Gutzwiller wavefunction as
where |ΨHF⟩ is the Hartree–Fock determinant.46 |ΨHF⟩ spans all determinants in the real-space basis, making the evaluation of ⟨Di|ΨGutzwiller⟩ require additional computational overhead. To this end, we simplify |ΨGutzwiller⟩ to
where |Di⟩ are the real-space Slater determinants. The overlaps with another real-space determinant |Dj⟩ are, therefore, simply given by
In FCIQMC, it is applied by scaling the spawning attempts from determinant |Di⟩ to |Dj⟩ by
In contrast to using full |ΨGutzwiller⟩, these weights are cheap to calculate as the diagonal elements are required anyway for the death/cloning step.
The matrix elements of the matrix that is sampled when weighing the spawns like that are
This is a similarity-transformed version of the original Hamiltonian; thus, the spectrum is conserved. Since the transformation is diagonal and the matrix elements are merely a scaled version of Hji, it is true that
so is also a similarity-transformed version of . Therefore, the spectrum of is conserved, too. Thus, ΔEstoq is left unchanged in the importance-sampled Hamiltonian.
On the other hand, the distribution of the weights in the ground-state wavefunction expansion that is sampled in an FCIQMC run is changed. The ground-state wavefunction of is given by
where Ci are the real-space wavefunction coefficients of the untransformed wavefunction. For positive values of g, the transformation increases the compactness of the wavefunction, i.e., the ℓ1 norm of the wavefunction is concentrated on fewer determinants since determinants with larger diagonal elements (i.e., a larger number of double occupancies) are suppressed. The effect on a wavefunction is shown for a small 2 × 3 Hubbard ladder in Fig. 5 where exact diagonalization is trivial.
Exact ground-state wavefunction of a 3 × 2 Hubbard ladder system at U/t = 8 for the untransformed (g = 0.0) and transformed Hamiltonian (g = 0.2). The determinants are ordered by weight in the respective wavefunction from large to small. The coefficients are normalized with respect to the ℓ1 norm.
Exact ground-state wavefunction of a 3 × 2 Hubbard ladder system at U/t = 8 for the untransformed (g = 0.0) and transformed Hamiltonian (g = 0.2). The determinants are ordered by weight in the respective wavefunction from large to small. The coefficients are normalized with respect to the ℓ1 norm.
B. Improved convergence with importance sampling
When running an FCIQMC simulation with a walker number Nw smaller than Nmin, the shift S is a biased energy estimator. Since for Nw < Nmin the sign problem is unresolved, the sampled FCIQMC wavefunction |Ψ(τ)⟩ is some superposition of the equally valid solutions |Ψ⟩ and −|Ψ⟩. Therefore, the population on the reference determinant |D0⟩ also does not have a constant sign. Thus, Eproj on average does not converge to the correct ground-state energy and cannot be used as an energy estimator.
The shift S is not only used to control the population but can also be used as an energy estimator. For a large enough population Nw ≥ Nmin, the average shift converges to the exact ground state energy E0. For Nw < Nmin, the Hilbert space is not occupied by enough walkers simultaneously such that enough annihilations can take place because opposite sign walkers do not meet in the same iteration on the respective determinant. A sign-incoherent wavefunction is sampled in this case. Because of only partially occurring annihilations, the Hilbert space is overpopulated with respect to the fermionic ground-state wavefunction |Ψ0⟩ but underpopulated with respect to the stoquastized version as some but not all necessary annihilations take place on average. Therefore, the average shift, which is sensitive to the number of annihilated walkers as it is the population control parameter, is in between the two respective ground-state energies in these cases, .
Applying a guiding wavefunction such as |Ψg⟩ that compactifies the ground-state solution and resembles the true solution |Ψ0⟩ can greatly improve the convergence of with respect to Nw and, thus, lower Nmin. It increases the number of annihilations and in doing so enables the algorithm to resolve the sign problem for a lower number of walkers. Figure 6 shows the convergence curve for a 2 × 8 ladder system at half-filling and with one hole for different g. It also displays the mean number of annihilations for the half-filled system as a function of Nw. Also shown is Nmin as determined using the fixed-N0 scheme with N0 = 50. As expected, it is situated at a walker number slightly above where convergence is reached.
Convergence of the average shift and mean number of annihilations with respect to the total walker number Nw for a 2 × 8 real-space Hubbard lattice at U/t = 8 at half-filling and with one hole. Application of |Ψg⟩ as a guiding wavefunction improves convergence considerably and lowers Nmin as it makes annihilations more efficient. The effect is increased when the parameter g is increased. Also shown as vertical lines in the top plot is the value for Nmin determined in a simulation with the shift update scheme according to Eq. (7) with N0 = 50.
Convergence of the average shift and mean number of annihilations with respect to the total walker number Nw for a 2 × 8 real-space Hubbard lattice at U/t = 8 at half-filling and with one hole. Application of |Ψg⟩ as a guiding wavefunction improves convergence considerably and lowers Nmin as it makes annihilations more efficient. The effect is increased when the parameter g is increased. Also shown as vertical lines in the top plot is the value for Nmin determined in a simulation with the shift update scheme according to Eq. (7) with N0 = 50.
C. Statistical considerations in importance-sampled calculations
It is a natural question to ask what the effect of an approximate guiding wavefunction is compared to that using the exact ground-state wavefunction |Ψ0⟩ if known. Figure 7 shows the dynamics of the walker population and the averaged values for Nmin in a fixed-N0 run for a small 2 × 3 Hubbard system. We compare runs where |Ψg⟩ with increasing g and |Ψ0⟩ are applied as guiding wavefunctions. Increasing g only ever decreases the plateau Nmin down to a minimum that is reached when |Ψ0⟩ is applied. Using |Ψg⟩ that is only an approximation to |Ψ0⟩, however, introduces large fluctuations, which increases the time needed to reduce statistical errors in the energy estimators to the desired level and eventually makes the algorithm unstable, especially when aiming at low N0 in the fixed-N0 mode.
Dynamics and averaged minimum walker numbers Nmin for a small 2 × 3 real-space Hubbard system at U/t = 8. Nmin has been determined using the fixed-N0 method with N0 = 10 000. Increasing the parameter g in the simple Gutzwiller-like guiding wavefunction monotonically lowers Nmin but increases the fluctuations both in Nw and in the energy estimators considerably (also see Fig. 8). The value of Nmin that is achieved by applying the exact ground-state wavefunction |Ψ0⟩ as a guiding wavefunction presents a lower bound for the Nmin achieved with increasing g but with considerably lower fluctuations.
Dynamics and averaged minimum walker numbers Nmin for a small 2 × 3 real-space Hubbard system at U/t = 8. Nmin has been determined using the fixed-N0 method with N0 = 10 000. Increasing the parameter g in the simple Gutzwiller-like guiding wavefunction monotonically lowers Nmin but increases the fluctuations both in Nw and in the energy estimators considerably (also see Fig. 8). The value of Nmin that is achieved by applying the exact ground-state wavefunction |Ψ0⟩ as a guiding wavefunction presents a lower bound for the Nmin achieved with increasing g but with considerably lower fluctuations.
We investigate the effect of the Gutzwiller-like guiding wavefunction on the stochastic noise further, this time using the larger 2 × 8 ladder system. As shown in the top panel of Fig. 8, increasing g only ever decreases Nmin. Furthermore, we calculate the standard deviation of the shift energy estimator for FCIQMC calculations conducted at their respective Nmin for g values ranging from 0.0 (no importance sampling) to 0.2 (see Fig. 8, bottom panel).
Reduction of Nmin by applying |Ψg⟩ as a guiding wavefunction calculated using the fixed-N0 method with N0 = 50 for the 2 × 8 ladder system at U/t = 8 (top). Further increasing the Gutzwiller parameter g monotonically reduces Nmin, but the additional effect weakens. The standard deviation of the shift energy estimator of calculations at the respective walker number Nmin for equal central processing unit (CPU) times, however, shows an optimum at around g ≈ 0.1 and sharply increases for larger g (bottom).
Reduction of Nmin by applying |Ψg⟩ as a guiding wavefunction calculated using the fixed-N0 method with N0 = 50 for the 2 × 8 ladder system at U/t = 8 (top). Further increasing the Gutzwiller parameter g monotonically reduces Nmin, but the additional effect weakens. The standard deviation of the shift energy estimator of calculations at the respective walker number Nmin for equal central processing unit (CPU) times, however, shows an optimum at around g ≈ 0.1 and sharply increases for larger g (bottom).
For equal CPU time—which corresponds to proportionately more iterations in imaginary time for lower Nw—the lowest standard deviation σ is reached with importance sampling at around gopt ≈ 0.1. For larger g, σ sharply increases, making exact calculations unfeasible at some point. This effectively puts a boundary on the applicability of |Ψg⟩ when deviating from guiding wavefunctions that resemble the exact solution. In some cases, it can be useful, however, to exceed gopt when one can accept larger statistical deviations in order to further lower Nmin to make a calculation affordable.
V. FUNDAMENTAL GAPS IN 1D AND 2D-LADDER SYSTEMS
We can now use the observation that the sign problem in 1D systems is not size-extensive and the possibility to reduce Nmin in problems with a size-extensive but weak sign problem, such as the 2D ladder systems (both at half-filling and with one hole). We can use this to calculate the fundamental many-particle gaps in these systems. The fundamental gaps of the Hubbard model are given by
where E(hf) is the energy of the half-filled system and E(1e/h) is the energy with one excess electron or one hole, respectively. Since for the Hubbard model, E(1e) = E(1h) + U, it is only necessary to calculate the half-filled and the one-hole system.
In Fig. 9, we show ΔEfund for chains up to 102 sites in both open and periodic boundary conditions using the NECI code.47 The calculations involved in open boundary conditions are all sign-problem-free. The calculations for the periodic boundary case involve sign problematic calculations; however, these have a non-size-extensive sign problem. Thus, calculations of more than 100 sites are doable. Figure 10 shows the ladder systems with 2 × 6, 2 × 8, 2 × 10, and 2 × 12 sites for both U/t = 8 and 16 calculated with FCIQMC. In the ladder systems, the sign problem is size-extensive. Therefore, the number of walkers to overcome the sign problem scales exponentially with system size. Importance sampling, however, allows for the convergence of the 2 × 12 with an affordable number of walkers. The results and the minimum walker numbers for U/t = 8 are given in Table I. In the most challenging system, the 2 × 12 ladder system with one hole at U/t = 8, with g = 0.25, we were able to reduce Nmin from at least several billion walkers—which well exceeds the applicability of our available hardware—to roughly 300 × 106 for which we can calculate the ground-state energy on 32 nodes with 640 cores in 96 h with an error on the order of 10−4.
Fundamental many-particle gaps for Hubbard chains as a function of the inverse chain length 1/ℓ at U/t = 8 and 16 calculated using FCIQMC with the open boundary condition (obc) and periodic boundary condition (pbc). Even though one-hole systems with pbc do have a sign problem, in principle, due to the non-size-extensivity of the sign-problem, calculations of over 100 sites close to the thermodynamic limit (tdl) can be converged. The analytic Bethe ansatz result in the tdl is given as a reference.
Fundamental many-particle gaps for Hubbard chains as a function of the inverse chain length 1/ℓ at U/t = 8 and 16 calculated using FCIQMC with the open boundary condition (obc) and periodic boundary condition (pbc). Even though one-hole systems with pbc do have a sign problem, in principle, due to the non-size-extensivity of the sign-problem, calculations of over 100 sites close to the thermodynamic limit (tdl) can be converged. The analytic Bethe ansatz result in the tdl is given as a reference.
Fundamental many-particle gaps for Hubbard ladders as a function of the inverse chain length 1/ℓ at U/t = 8 and 16 calculated using importance-sampled FCIQMC with periodic boundary conditions (pbcs). By using importance sampling with a simple Gutzwiller-like guiding wavefunction, systems of up to 2 × 12 can be calculated without any bias related to the sign problem with an affordable number of walkers (see also Table I).
Fundamental many-particle gaps for Hubbard ladders as a function of the inverse chain length 1/ℓ at U/t = 8 and 16 calculated using importance-sampled FCIQMC with periodic boundary conditions (pbcs). By using importance sampling with a simple Gutzwiller-like guiding wavefunction, systems of up to 2 × 12 can be calculated without any bias related to the sign problem with an affordable number of walkers (see also Table I).
Numerical results for the ground-state energies E0 for the 2 × ℓ systems at half-filling and with one hole at U/t = 8, respectively. Additionally, we give the values for Nmin measured using the fixed-N0 method for N0 = 50 with the respective g.
Lattice . | E0 . | Nmin [×103] w/o imp. samp. . | Nmin [×103] w/imp. samp. . | g . |
---|---|---|---|---|
(a) Half-filling | ||||
2 × 6 | −5.1600 | 48 | 11 | 0.15 |
2 × 8 | −6.8469 | 718 | 97 | 0.15 |
2 × 10 | −8.5382 | 12 638(2) | 1085(1) | 0.15 |
2 × 12 | −10.2353 | n.a. | 21 275(10) | 0.15 |
(b) One hole | ||||
2 × 6 | −6.6890 | 149 | 55 | 0.15 |
2 × 8 | −8.4518 | 3168(1) | 627 | 0.15 |
2 × 10 | −10.1687 | 80 494(10) | 13 410(5) | 0.15 |
2 × 12 | −11.8794 | n.a. | 297 850(9774) | 0.25 |
Lattice . | E0 . | Nmin [×103] w/o imp. samp. . | Nmin [×103] w/imp. samp. . | g . |
---|---|---|---|---|
(a) Half-filling | ||||
2 × 6 | −5.1600 | 48 | 11 | 0.15 |
2 × 8 | −6.8469 | 718 | 97 | 0.15 |
2 × 10 | −8.5382 | 12 638(2) | 1085(1) | 0.15 |
2 × 12 | −10.2353 | n.a. | 21 275(10) | 0.15 |
(b) One hole | ||||
2 × 6 | −6.6890 | 149 | 55 | 0.15 |
2 × 8 | −8.4518 | 3168(1) | 627 | 0.15 |
2 × 10 | −10.1687 | 80 494(10) | 13 410(5) | 0.15 |
2 × 12 | −11.8794 | n.a. | 297 850(9774) | 0.25 |
In Fig. 11, we compare the FCIQMC result–which is numerically unbiased within stochastic error bars when accounting for the (very small) population control bias—for the 2 × 12 system at U/t = 8 with one hole with results from the density-matrix renormalization group (DMRG).48 The DMRG numbers are shown as a function of 1/M, where M is the bond dimension of the matrix product state (MPS). Fiedler-type ordering of the sites is used. The MPS representation is known to be optimal in the 1D case because of the low entanglement of the ground state. In systems that resemble 1D systems, such as the ladder systems, we have been dealing with in this paper, this advantage still carries over. We see good agreement between the FCIQMC result for 3 × 108 walkers and g = 0.25 and DMRG for M ≥ 4000.
Convergence of the DMRG-optimized variational energy, for the 2 × 12 Hubbard ladder with one hole, with the bond dimension M in Fiedler ordering. A comparison with the FCIQMC-optimized ground-state energy is shown (range gives stochastic errors; g = 0.25; 3 × 108 walkers).
Convergence of the DMRG-optimized variational energy, for the 2 × 12 Hubbard ladder with one hole, with the bond dimension M in Fiedler ordering. A comparison with the FCIQMC-optimized ground-state energy is shown (range gives stochastic errors; g = 0.25; 3 × 108 walkers).
VI. CONCLUSION AND OUTLOOK
In this paper, we have described a certain class of Hubbard systems that show very weak sign problems. Thus, they can be treated in FCIQMC numerically exactly without the initiator approximation. In the special case of 1D systems, the sign problem is non-size-extensive, i.e., systems at various fillings can be treated for system sizes of up to 102 sites when accounting for the population control bias. 2 × ℓ ladder systems show a size-extensive yet very weak sign problem. Using importance sampling with an easy-to-evaluate Gutzwiller-like guiding wavefunction, however, can reduce the required number of walkers such that systems well beyond the scope of exact diagonalization can be treated numerically exactly within stochastic error bars.
In an upcoming publication, we will discuss controlled approximations in systems beyond what is possible in a numerically exact manner. Importance-sampled FCIQMC will be used as a basis to treat very-weak sign problem subspaces of larger systems. Furthermore, using importance sampling with more sophisticated guiding wavefunctions, if evaluated efficiently, could potentially reduce the minimum number of walkers further, expanding the scope of treatable systems.
ACKNOWLEDGMENTS
The authors would like to thank Michael Willatt for fruitful discussions. The authors acknowledge the Max Planck Society for support.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Niklas Liebermann: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Software (lead); Validation (lead); Visualization (lead); writing – original draft (lead); writing – review & editing (lead). Khaldoon Ghanem: Conceptualization (supporting); Investigation (supporting); Methodology (supporting); Software (supporting). Ali Alavi: Conceptualization (supporting); Formal analysis (supporting); Funding acquisition (lead); Investigation (supporting); Methodology (equal); Project administration (lead); Resources (lead); Supervision (lead); Validation (supporting); writing – original draft (supporting); writing – review & editing (supporting).
DATA AVAILABILITY
All FCIQMC calculations were conducted using NECI (https://github.com/ghb24/NECI_STABLE). DMRG calculations were conducted using BLOCK (https://github.com/sanshar/Block). The data that support the findings of this study are available from the corresponding author upon reasonable request.
APPENDIX A: SIGN-PROBLEM-FREE 1D SYSTEMS: PROOF
The 1D Hubbard model is sign-problem-free in the following cases:
for open boundary conditions and
for periodic boundary conditions for odd n↑ and n↓.
This can be easily seen in the usual fermionic second quantization formalism. To this end and without loss of generality,49 we will write a Slater determinant |Di⟩ with the fermionic creation operators at site p acting onto the vacuum state |⟩ in the convention of spin-first ordering as
with a positive sign. n↑ (n↓) are the number of up-spin (down-spin) electrons. αp and βq number the spatial sites in ascending order. The Hubbard Hamiltonian only allows for hoppings of the same spins to nearest-neighboring sites, so without loss of generality, an excitation of an ↑-electron from site j to a site j ± 1 is given by
(where j ± 1 is not occupied with an ↑-electron already). Since both and were commuted with the same number of operators, trivially, there is no sign change. For an excitation of a ↓-electron, the same applies. This proves that all 1D Hubbard systems with open boundary conditions are sign-problem-free. With an excitation that exploits periodic boundary conditions, however, we get
had to be commuted with n↑ − 1 creation operators, whereas did not have to be commuted at all. Thus, there is no sign change only for odd n↑. For a ↓-electron, we get
In addition, here, n↓ has to be odd for no sign change to occur.
APPENDIX B: NON-SIZE-EXTENSIVE SIGN PROBLEM IN 1D SYSTEMS: HEURISTIC EXPLANATION
Without loss of generality, let us look at two determinants |Di⟩ and |Dj⟩ in a system with an even number of ↑-electrons and an odd number of ↓-electrons. Between |Di⟩ and |Dj⟩, a single ↑-electron is moved from site i to site j,
The two main contributing pathways consist of one moving ↑-electrons not exploiting periodic boundary conditions and one moving them exploiting the boundary conditions. In the FCIQMC language, according to Eq. (3), a walker sitting on |Di⟩ is creating
walkers on |Dj⟩, where ndocc is the number of doubly occupied sites shared by |Di⟩ and |Dj⟩, |j − i| is the number of sites between site i and site j, and is the number of ↓-electrons between i and j. In contrast, exploiting periodic boundary conditions leads to
On the one hand, the contribution of these pathways to the fermionic solution is given by Pnon-periodic + Pperiodic. On the other hand, the contribution to the stoquastic solution is . Therefore, we can conclude that the stoquastized gap due to these contributions is largest when Pnon-periodic = −Pperiodic because of maximal cancellation. One deviates from the equality if (i) nsites is large, (ii) is large, and (iii) U is large, implying a reduction of the sign-problematic contribution.
REFERENCES
It is noted that does not necessarily have to have only negative off-diagonal entries, which is the definition of a stoquastic matrix, in order for it to be sign-problem-free. We transform a stoquastic matrix Mstoq to M according to M = DMstoqD, with D being a diagonal matrix with diagonal entries of either +1 or −1. This transformation introduces positive off-diagonal elements to M but does not change the spectrum, i.e., the stoquastic gap remains zero, and it can be simulated in a sign-problem-free manner.
The way of updating the shift according to Eq. (7) typically leads to larger fluctuations in the energy estimators, which would require longer running times. Therefore, we only use this method to establish and then continue with updating the shift according to Eq. (5) with a target population set to .
Note that the proof regarding the existence of a sign problem holds for any ordering. When not choosing the spin-first ordering, the corresponding Hamiltonian will not be stoquastic since a change of ordering will multiply whole columns and their corresponding rows of the Hamiltonian with −1. This does not affect the sign problem, however, which makes the proof valid in any case.