With a transcorrelated Hamiltonian, we perform a many body perturbation calculation on the uniform electron gas in the high density regime. By using a correlation factor optimized for a single determinant Jastrow ansatz, the second order correlation energy is calculated as . This already reproduces the exact logarithmic term of the random phase approximation (RPA) result, while the constant term is roughly 7% larger than the RPA one. The close agreement with the RPA method demonstrates that the transcorrelated method offers a viable and potentially efficient method for treating metallic systems.
I. INTRODUCTION
Electronic structure calculations usually suffer from numerical problems originating from the Coulomb potential . On the one hand, the short-range singularity leads to non-smoothness in the many body wave function, characterized by a cusp at the electron coalescence point r12 = 0,1 which causes a slow convergence with respect to the basis-set in configuration expansions of the wave function. On the other hand, the slow decay of the long-range tail of the Coulomb potential also causes a major problem in calculations of metallic systems. Any straightforward perturbation treatment leads to divergent values in the thermodynamic limit, and in order to get a meaningful perturbation result, one has to sum over some sub-sequences of the perturbation expansion up to infinite order, a technique which is usually referred to as the random phase approximation (RPA).2–6
One way to deal with the short-range singularity of the wave function is to make use of the Jastrow ansatz,7
to split the cusp into a pair correlation factor
With a factor u fulfilling the asymptotic condition
the smoothness of the function Φ is one order higher than that of the function Ψ.8 This product form is widely used in quantum Monte Carlo methods,9–11 where the expectation value of the Hamiltonian can be evaluated by random sampling. One can, thus, minimize this expectation value of energy by adjusting the correlation factor τ. Due to the involved high-dimensional integrals, it is difficult to use such kind of variational treatment of the Jastrow wave function in conventional quantum chemistry, where the function Φ needs to be further approximated by configuration expansions. Alternatively, the transcorrelated (TC) method of Boys and Handy12–19 offers a fairly efficient way to deal with the Jastrow ansatz. With this method, the original Schrödinger equation is transformed into a non-Hermitian eigenvalue problem,
via an exact similarity transformation, which removes the involved exponential factor, and results in an exact effective Hamiltonian containing up to three body terms,
This method was initially studied by Boys and Handy for the single determinant ansatz and has only recently been combined with configuration expansion methods of quantum chemistry.20–28 These works reveal a high level of efficiency of the transcorrelated Hamiltonian in handling short range correlations, that the resolving of the cusp leads to significant speed up of the basis convergence, and furthermore, more speed up can be achieved when more generalized factors are used (instead of the simple r12-type or F(r12)-type factors).
Regarding the study of long-range correlation with the transcorrelated method, there exist only a few works dealing with the single determinant ansatz for uniform electron gases (UEGs).29–31 By choosing a proper correlation factor τ, the leading order singularity of the Coulomb potential can be removed from the effective Hamiltonian for both the short-range and the long-range singularities. We, therefore, expect that the divergence of the perturbation energy for metallic systems can be cured to some extent. It should at least be possible to get non-divergent results in the thermodynamic limit at low order perturbation theory without any RPA-type treatment. It would, then, be interesting to know whether this kind of non-RPA treatment leads to meaningful results. In this work, we will test the performance of the many body perturbation (MBPT) treatment of the transcorrelated Hamiltonian on UEG in the high density regime, which has a well-established RPA result,5,32
In Sec. II, we describe the perturbation treatment of the TC Hamiltonian, and in Secs. III and IV, we present the details of the calculations of the first and the second order energies, where a Jastrow factor optimized for the Hartree–Fock reference function is used. In Sec. V, we also present some calculation results with the widely used Coulomb–Yukawa factor, which will be followed by some concluding remarks in Sec. VI.
II. Perturbation treatment of the transcorrelated Hamiltonian
Uniform electron gases are described by the Hamiltonian
The system is characterized by the charge density ρ (or, equivalently, the Wigner–Seitz radius rs or the Fermi wavevector kF). In order to get the explicit dependence of the physical quantities on rs, we re-scale the spatial coordinates
so that the system keeps a fixed density,
but has a re-scaled Hamiltonian,
The formal volume Ω and the particle number N will finally be taken as ∞ in the thermodynamic limit. In this paper, we will only use this re-scaled form and, thus, will ignore the prime on the variables. We only need to remember that the final energies should be divided by according to Eq. (10). Then, we use Jastrow ansatz (1) for the ground state wave function of , and the two body correlation factor u(ri, rj) can be written as u(ri − rj) due to the translational symmetry. The transcorrelated Hamiltonian (5) can be derived straightforwardly in the second quantization formalism31 in terms of plane wave orbitals,
where is the Fourier transformation of u(r). is the Fourier transformation of (▿u(r))2 and can be calculated via the convolution theorem,
Here, we have used the fact that in the thermodynamic limit, the function is actually a 1D function of k due to the rotational symmetry.
It is of crucial importance to take a proper Jastrow factor u. It has to fulfill asymptotic conditions at both the limits,31
where is the Fourier transformation of the Coulomb potential and ρ is the constant density given in Eq. (9). These conditions remove the leading order singular term of the Coulomb potential and lead to finite second order perturbation energy. However, they do not guarantee the quality of the perturbation energy. Based on some recent numerical studies of the TC method on small molecular systems,25,26 the correlation factor optimized for a single determinant reference Φ (e.g., the Hartree–Fock wave function) also performs well when Φ is further treated by configuration expansions, at least for weakly correlated systems.
Optimizations of the correlation factor for a single determinant Jastrow ansatz (eτΦ0) can be realized by solving either a variational equation or a transcorrelated equation for the factor.33 Based on a minimization of the variational energy, one can derive a variational equation,
The explicit presence of the exponential factor eτ in this equation usually leads to major numerical difficulties in solving the equation. For small- or medium-sized systems, it can be directly treated using the variational quantum Monte Carlo method. For large or extended systems, however, the exponential factor can only be treated approximately; for example, the linked cluster expansion34–36 offers one of such kind approximate treatment. This is essentially a RPA-type treatment of the single determinant Jastrow ansatz. Alternatively, it is possible to avoid the treatment of the exponential factor by using the transcorrelated equation for the correlation factor,12,14
The exponential factor is removed from the above equation by the similarity transformation, and thus, Eq. (19) can be solved more easily compared with Eq. (18), especially for large or extended systems. Despite this remarkable simplification, the solutions of Eq. (19) are found to be very close to those of Eq. (18) both for molecular systems33 and for the UEG.31
For UEG in the high density limit, it is interesting to note that both the variational equation [Eq. (18)] and the transcorrelated equation [Eq. (19)] lead to the same final equation,29,31,34–36
where T2(k) is a volume factor defined by
with Θ being the Heaviside step function. In the thermodynamic limit, T2(k) can be calculated as
The correlation factor optimized for the Hartree–Fock reference is, then, solved as
With this optimized factor , we can build the transcorrelated Hamiltonian (12), which is, then, split into two parts for a perturbation treatment,
where for simplicity, we take only the kinetic terms in ,
With this partition of the TC Hamiltonian, the perturbation energies can be written as
where E0 is simply the total kinetic energy and E1 contains the exchange energy and the first order correlation energy . The second order energy contributes only to the correlation energy.
III. THE FIRST ORDER PERTURBATION ENERGY
The first order correlation energy is composed of the expectation values of the two body operator and the three body operator . For calculations of these values, we need to find all possible contractions of the creation and annihilation operators. From the two body operator, we get a direct contribution and an exchange contribution,
Here, only the term of makes a non-vanishing contribution, while the other terms coming from have no contribution to the expectation value. The expectation value of the three body operator also has two possible contractions,
The volume factor O3 is defined as
which can be calculated analytically in the thermodynamic limit.31
In the thermodynamic limit, the summation should be replaced by an integration
This replacement will be used in the following context without any notification.
The first term in Eq. (31) is partly canceled by the first term in Eq. (30), and we finally have three terms in the expression of the first order correlation energy,
For a more transparent treatment, we reformulate expression (23) of as
The expression of the first order energy contains only quadratic terms of , and the prefactor αrs of in Eq. (38) will be canceled by the prefactor in the final energy expression according to Eq. (10). If rs is set to be 0 in the remaining integrands in the first order energy, it is found that will diverge, while the expression of and still gives finite results. The singular term in originates from the integration over small k. In order to get an analytic result of the singular term, we rearrange the expression of as
In the first line of expression (40), T2(k) is simplified with its leading term in the small k region, and this simplified integrand is subtracted out from the original integrand in the second and third lines. The simplified integral in the first line can be calculated analytically as
In the second and third lines, the singularities of the two terms cancel each other, and hence, they contribute at most only to the constant term in the correlation energy. For the calculation of the contribution to the constant, we can simply take out the common factor and set rs = 0 in the remaining integrand, which gives
In a similar way, the integral in the last line can be easily calculated as
and also contribute at most only to the constant term of the correlation energy and can also be simplified with the above method. After this simplification, it is still too difficult to be calculated analytically due to the complicated integrand. There is, however, no problem for a numerical calculation of such 1D integrals, which leads to
By adding all the results together, we get the final result for the first order correlation energy,
The first order correlation energy is essentially the correlation energy produced by the single determinant Jastrow ansatz. It is interesting to find that the leading logarithmic term is the same as that obtained by Talman based on the linked cluster expansion.36 This logarithmic term makes up roughly 90% of the exact RPA logarithmic term , which is also a typical portion of correlated energy one may expect from variational quantum Monte Carlo calculations based on the single determinant Jastrow ansatz.
IV. THE SECOND ORDER PERTURBATION ENERGY
The transcorrelated effective potential depends on rs nonlinearly. Besides the original Coulomb potential w(k), which is proportional to rs, there are some terms that rely on linearly and others quadratically. According to Eq. (38), in the high density regime where rs ∼ 0, scales roughly linearly with rs. To get an estimation of the contribution to the second order energy E2 from each quadratic term in the effective potential , we can first set rs = 0 in the denominator of expression (38) of ,
If this only leads to finite integrals in the calculation of E2, we can ignore this quadratic term for the current calculation since it gives at most a contribution to E2 according to Eq. (29). Based on this estimation, we find there is only one quadratic term left, where approximation (47) leads to divergent integrals in the calculation of E2. This term is given as
which is a contraction of the three body potential in Eq. (14) by taking k = k′. Consequently, for the calculation of the two leading terms in the second order energy E2, we only need to take the following approximation of the effective potential:
By taking this two body potential approximation of , the second order energy consists of two contributions,
obtained by a direct contraction and an exchange contraction, respectively.
The direct contraction term of E2 can be written as
Q(k, t) and A(k, t) are given as integrals,
Detailed calculations dealing with Q(k, t) and A(k, t) are presented in Appendixes A and B. The results are written as
where X(k) is piecewisely defined for k < 2,
and for k ≥ 2,
where we find that the last term equals . Therefore, we have
where we have used a simple relation
which can be obtained from Eqs. (51) and (20). As far as the singularity at k ∼ 0 is concerned, the first integrand is similar to the integrand of ,
We can reformulate the expression of as
where the remaining integrals are not any more singular at k ∼ 0 in the high density limit. Then, we can use the simple expression (47), and the integrals can be numerically evaluated as
and the sum with the first order energy gives
The exchange contraction term of E2 can be written as
where
which can be calculated analytically.31 Therefore, in the thermodynamic limit, we have
which can be calculated numerically as
The total correlation energy is finally estimated as
This result has precisely the same logarithmic term as that of the RPA result, while the constant term is only roughly 7% larger.
The optimized and the Coulomb–Yukawa correlation factor, together with their difference.
The optimized and the Coulomb–Yukawa correlation factor, together with their difference.
V. CALCULATIONS WITH THE COULOMB–YUKAWA FACTOR
The above calculations are based on the optimized Jastrow factor for the given system, which is somewhat complicated for the analytic calculations. Can we use some simpler factors instead, which still fulfill the asymptotic conditions?
In quantum Monte Carlo, a Coulomb–Yukawa-type function,
is broadly used as the Jastrow factor. Here, the expression is reformulated for re-scaled Hamiltonian (11). The Fourier transformation of u(r) has the following simple form:
where the parameters are determined to fulfill the asymptotic conditions. This factor is much simpler for the integral calculations and serves as a good candidate for the test.
To get an idea about the difference between the optimized (opt) and the Coulomb–Yukawa (CY) factors, we have plotted and on a double-log scale in Fig. 1, together with their difference . At short range (k ∼∞), the two factors have a clear power law decay , and their difference also decays sharply when k > 1. (Recall that in our rescaled problem, kF ≡ 1.) At long range (k ∼ 0), the two factors also show a clear power law divergence . Their difference, however, also diverges . Since the leading order [O(k−2)] term of is used to cure the divergence of the original second order energy, the final value of will be determined by the O(k−1) order terms of . We, therefore, expect that the Coulomb–Yukawa factor will lead to a somewhat different result for the second order energy.
In order to get an idea about the performance of this factor in the second order transcorrelated perturbation calculations, we only focus on the leading terms of the perturbation energies. The leading term in the first order energy is given by in Eq. (35),
which does not lead to a logarithmic term, but rather a more singular term.
According to Eq. (60), the term will be completely canceled by the second order energy. The leading term of the total second order perturbation energy can be evaluated by
which reproduces only half of the exact logarithmic term.
This example shows that in order to get a meaningful perturbation energy, it is not enough to construct a correlation factor only based on the asymptotic conditions. The current study suggests that a correlation factor optimized for a single determinant reference function (e.g., the Hartree–Fock reference function) serves as a good candidate. This optimization can be realized by solving either the variational equation [Eq. (18)] or the transcorrelated equation [Eq. (19)] for the factor.
VI. CONCLUSIONS AND REMARKS
By using a proper Jastrow factor, the leading singularity of the Coulomb potential can be removed from the transcorrelated Hamiltonian. This cures the divergence problem of the MBPT method for metallic systems at low orders. Test calculations on uniform electron gases show that already at the second order perturbation level, one can get meaningful results for the correlation energy. We may expect that this method can be developed into an efficient method for real metals, curing problems at both the short-range and the long-range.
The selection of the Jastrow factor is crucial for calculations. The necessary condition that the factor should fulfill the long- and short-range asymptotic conditions does not guarantee performance. For example, the Coulomb–Yukawa-type factor does not lead to a meaningful perturbation result. On the other hand, the transcorrelated equation for the correlation factor [Eq. (19)] may provide a practical means to optimize suitable correlation factors for realistic systems. This will be investigated in future work. We also note that this method only cures the divergence at low orders; starting at fourth order, there are still divergent integrals in the perturbation energies.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Hongjun Luo: Conceptualization (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). Ali Alavi: Conceptualization (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.
APPENDIX A: CALCULATION OF Q(k, t)
The function Q(k, t) can be calculated separately for k ≥ 2 and k < 2.
1. Case k ≥ 2
For k ≥ 2, the integration volume is a simple unit sphere,
In the calculation of integrals, we also need to evaluate the leading orders of Q,
and, thus,
2. Case k < 2
For k < 2, the integration volume is the unit sphere centered at 0 subtracting its intersection with the unit sphere centered at k,
Again, we need the leading terms of the Taylor expansion of Q(k, t) with respect to t, which can be calculated as
Then, we can get
APPENDIX B: CALCULATION OF AND
1. Case k ≥ 2
According to Eq. (A1),
This integral has the form . Since Q(k, t) is finite at t = 0, the leading term of F(x) has the order x6. We can, then, use the simple relation
The calculation with F(5) is lengthy but straightforward, and in the end, we are left with only the following type integrals:
The final result of the integral is
The integral can be simply calculated by partial integration
2. Case k < 2
In the same way as above, the integral can be calculated as
For the calculation of the logarithmic term in , we split out the leading order term of the above integral and write it as
with X(k) ∼ O(k2) being the high order part,
The integral on A can also be calculated simply by partial integration,