857
Views
0
CrossRef citations to date
0
Altmetric
Jürgen Troe Special Issue

Calculation of reaction rate constants from a classical Wigner model based on a Feynman path integral open polymer

ORCID Icon, ORCID Icon & ORCID Icon
Article: e1970264 | Received 31 Mar 2021, Accepted 29 Jul 2021, Published online: 21 Sep 2021

Abstract

The Open Polymer Classical Wigner (OPCW) method [J. Chem. Phys. 152, 094111 (2020)] is applied to a flux-Heaviside trace for calculating reaction rate constants. The obtained expression for the reaction rate constant is tested on a symmetric Eckart potential. The OPCW method is shown to converge toward the exact classical Wigner result. The OPCW method shows some promise for rate constant calculations and suggestions for future tests are made.

GRAPHICAL ABSTRACT

1. Introduction

Reaction rates are of central importance in chemistry and methods to calculate reaction rate constants are therefore of interest. In many cases classical mechanics is sufficient in order to calculate rate constants, but quantum mechanics can be essential for instance for light atoms, such as hydrogen, and in cold environments.

Methods to obtain reaction rate constants from a Wigner distribution include the classical Wigner method [Citation1, Citation2], with alternatives such as Feynman–Kleinert Linearised Path Integral [Citation3] (FK-LPI), Local Gaussian Approximation Linearised Semi-Classical Initial Value Representation [Citation4] (LGA-LSC-IVR), and to some extent the transition state theory with quantum partition functions of Wigner [Citation5] and Eyring [Citation6] and the quantum transition state theory of Pollak and Liao [Citation7]. Other notable approximate quantum mechanical methods to calculate reaction rate constants are e.g. the instanton reaction rate theory [Citation8–10], the statistical adiabatic channel model [Citation11] and Ring Polymer Molecular Dynamics (RPMD) [Citation12].

This article presents reaction rate constant calculations using a recent implementation of the classical Wigner method by the same authors [Citation13]. Section 2 is an introduction to a general formulation of chemical reaction rates, then section 3 combines this theory with the Open Polymer Classical Wigner (OPCW) method. Section 4 notes some important corrections to consider for running the calculations. Sections 56, and 7 respectively give the details of the computations that have been run, show the results of these computations, and contain the conclusions that have been drawn.

2. Reaction rate constants

Miller and coworkers published three quantum mechanical traces that can be used to acquire bimolecular reaction rate constants [Citation14, Citation15]. These traces are (1) Css(t)=Treβ2Hˆhxˆeβ2HˆeiHˆthxˆeiHˆt(1) (2) Cfs(t)=Treβ2HˆFˆeβ2HˆeiHˆthxˆeiHˆt(2) (3) Cff(t)=Treβ2HˆFˆeβ2HˆeiHˆtFˆeiHˆt(3) where h(x) is the Heaviside function, xˆ is the position operator, β=(kBT)1, where kB is Boltzmann's constant and T is the absolute temperature, Hˆ is the Hamiltonian operator of the system, i is the imaginary unit, t is time, ℏ is the reduced Planck constant, and Fˆ is the probability density flux operator. These functions can be used to obtain reaction rate constants, kr, through the relations (4) krQR=limtddtCss(t)(4) (5) =limtCfs(t)(5) (6) =0dtCff(t)=12dtCff(t)(6)

where QR is the canonical partition function per unit volume of the reactants.

It can be noted that the traces presented here have a symmetrised Boltzmann operator [Citation16], i.e. eβ2Hˆ on both sides of the first operator in each trace. The relations Equation4 – Equation6 are just as valid for the real part of the trace with an asymmetrically placed Boltzmann operator [Citation14]. Important to notice is that these relations are defined for potentials where the reactants start infinitely far from each other and the products end up infinitely far from each other. This requires the potential to be unbound. This restriction applies to the potential employed in this article. In the computations presented in this article the trace used for the calculation of rate constants is Cfs(t).

3. Reaction rate constant calculation with the open polymer classical wigner method

In this section expressions for calculating reaction rate constants with the OPCW method will be derived. These derivations follow the ones in the original OPCW article [Citation13] quite closely, but use the symmetrised Boltzmann operator and will be specialised to the flux-side trace, Cfs(t) in Equation Equation2.

Initially Cfs(t) is written as the integral over position, x1, of a matrix element. (7) Cfs(t)=dx1x1|eβ2HˆFˆeβ2HˆeiHˆthxˆeiHˆt|x1(7)

The Boltzmann operators are each divided into N2 factors of eβNHˆ, where N must be an even number. N−1 identity operators are inserted between these operators. (8) Cfs(t)=j=1Ndxjx1|eβNHˆ|x2x2|eβNHˆ|x3xN2|eβNHˆFˆ|xN2+1xN1|eβNHˆ|xN×xN|eβNHˆeiHˆth(xˆ)eiHˆt|x1(8) This is an imaginary time, iβ, Feynman path integral [Citation17], which can be expressed with Wigner transforms instead of matrix elements, (9) Cfs(t)=j=1Ndxjdpj2πeij=1Npjx(jmodN)+1xj×eβNHˆWx1+x22,p1×eβNHˆWx2+x32,p2eβNHˆFˆWxN2+xN2+12,pN2eβNHˆWxN1+xN2,pN1×eβNHˆeiHˆth(xˆ)eiHˆtWxN+x12,pN.(9) In the limit N the Wigner transform of the Boltzmann operator simplifies to the classical Boltzmann factor. Also, the Boltzmann operators can be separated from the flux and Heaviside operators in this limit. We obtain (10) Cfs(t)=limNj=1Ndxjdpj2π×eij=1Npjx(jmodN)+1xj×eβNj=1NHyj,pjFˆWyN2,pN2×eiHˆth(xˆ)eiHˆtWyN,pN(10) where yj=xj+x(jmodN)+12 and H(yj,pj) is the classical Hamiltonian. If V(yj) is independent of momentum, all momenta except pN2 and pN can be integrated out analytically as done in the original derivation of the OPCW method [Citation13]. This gives (11) Cfs(t)=limNmN2πβN2β2πmNNj=1Ndxj×dpN2dpNeβNj=1NVyj×emN22βj=1N21xj+1xj2+j=N2+1N1xj+1xj2×eipN2xN2+1xN2eβNpN222mFˆWyN2,pN2×eipNx1xNeβNpN22m×eiHˆth(xˆ)eiHˆtWyN,pN,(11) where m is the mass. For the flux operator Fˆ=12m(δ(xˆ)pˆ+pˆδ(xˆ)), the Wigner transform is (12) FˆWx,p=pmδ(x),(12) where δ(x) is the Dirac delta function. Using the momentum integrals of Appendix A in our previous article [Citation13] a quantity F(yN2,xN2+1xN2) can be defined as (13) FyN2,xN2+1xN2=dpN2FˆWyN2,pN2eβNpN222meipN2xN2+1xN2dpN2eβNpN222meipN2xN2+1xN2=dpN2pN2mδyN2eβNpN222meipN2xN2+1xN22πmNβeβNmN2xN2+1xN2222β2=iNβxN2+1xN2δyN2.(13)

This can be put into Equation Equation11, giving (14) Cfs(t)=limNmN2πβN2β2πmNNj=1Ndxj×dpNeβNj=1NVyj×emN22βj=1N1xj+1xj2×iNβxN2xN2+1δyN2×eipNx1xNeβNpN22m×eiHˆth(xˆ)eiHˆtWyN,pN.(14) We apply the classical Wigner approximation, (15) eiHˆth(xˆ)eiHˆtWyN,pNh(xˆ)WxyN,pN,t,pyN,pN,t=hxyN,pN,t,(15) where x(yN,pN,t) is the position at time t of a classical trajectory starting from yN and pN. Then, by assuming finite N, we obtain (16) Cfs,y(t)mN2πβN2β2πmNN×j=1NdxjdpNeipNx1xN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2×iNβxN2xN2+1δyN2hxyN,pN,t,(16) which is the expression corresponding to the y1-version of the OPCW method for the flux-Heaviside trace. Due to the symmetrisation of the Boltzmann operator, the expression contains yN2 rather than y1 and below it is simply referred to as the y-version of the OPCW method. The trace will be called Cfs,y(t). As shown in Appendix 1, we can also formulate an alternative version of the flux-Heaviside trace where the delta function now takes xN2 as argument. This results in: (17) Cfs,x(t)mN2πβN2β2πmNNj=1Ndxj×dpNeipNx1xN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2×iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1iN2βxN2xN2+2δxN2+1×hxyN,pN,t,(17) which below is called the x-version of the OPCW method for the flux-Heaviside trace.

4. Correction factors for evaluation of traces with monte carlo

In this work the position integrals in Equations Equation16 and Equation17 are evaluated with Metropolis Monte Carlo [Citation18]. In Monte Carlo integration there is a weight function and an integrand. Here, the weight function used is (18) mN2πβN2β2πmNN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2δyy,x,(18) where yy,x is either yN2 or xN2+1. The integrand is thus (19) eipNx1xNiNβxN2xN2+1hxyN,pN,t(19) for the y-version of the OPCW method and (20) eipNx1xN×iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1iN2βxN2xN2+2βzβNγ,k(μkz)μk2×hxyN,pN,t(20) for the x-version of the OPCW method. This means that the quantity actually calculated is Cfs(t)Zδ, where Zδ is (21) Zδ=mN2πβN2β2πmNNj=1NdxjdpN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2×δyy,x,(21) which is similar to a canonical partition function, but the integrand includes a delta function and lacks a factor of eipN(x1xN). By sampling the integrand eipN(x1xN) in the same Monte Carlo run as Cfs(t)Zδ a correction factor ZδZδ can be obtained, where (22) Zδ=mN2πβN2β2πmNNj=1NdxjdpN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2×eipNx1xNδyy,x.(22) Zδ is similar to Zδ, but it has eipN(x1xN) in the integrand, and is thus a partition function with a delta function in it.

The quantities obtained from the Monte Carlo integration are thus Cfs(t)Zδ and ZδZδ. The final expression for Cfs(t) is (23) Cfs(t)=Cfs(t)Zδ×ZδZδ×Zδ,(23) where the additional correction factor Zδ has to be computed separately.

5. Computational details

In this section we first define the Eckart potential. We then give some details on how the traces in Equations Equation16 and Equation17 were evaluated and how the time propagation was done. Finally we consider some numerically exact expressions against which our OPCW results can be compared.

5.1. Potential and system parameters

The symmetric Eckart barrier is given by [Citation19] (24) VEckart(x)=6ωπsech2πmω12x,(24) where ω is the absolute value of the angular frequency at the top of the barrier, i.e. at the transition state.

5.2. Traces

The integrals in equations Equation16 and Equation17 were evaluated through the Metropolis Monte Carlo method [Citation18]. The maximum stepsize for the Monte Carlo was chosen and updated according to the procedure in the original OPCW work [Citation13]. A Maxwell-Boltzmann distribution of the inverse temperature β/N was used to sample the momentum, pN. For the sampling, the ran2 pseudo-random number generator of Press et al. [Citation20] was used.

A molecular dynamics trajectory was run for each 10th Monte Carlo step using the velocity Verlet algorithm [Citation21, Citation22], with a timestep of 0.05 ω1 and a total time of 10 ω1.

The standard deviations of the results where calculated with the block average method, as explained in e.g. [Citation23, Citation24], using a minimum block size of 106 Monte Carlo steps.

To obtain the correction factor Zδ for the Eckart potential the numerical matrix multiplication scheme [Citation25] was used to calculate the matrix elements of the Boltzmann operator. As an alternative, we show how to obtain Zδ by Monte Carlo in Appendix 2.

5.3. Expressions for comparison

The classical limit of the flux-Heaviside trace is easily shown to be (25) kr,CMQR=12πβeβV(0).(25) The corresponding quantum mechanical result can be calculated from an analytic transmission coefficient [Citation19Citation26], (26) ϰEckart(E)=cosh24πE6ω1cosh24πE6ω+cosh576π2,(26) where E is energy. The trace is then obtained by numerically integrating (27) kr,QMQR=12π0dEeβEϰEckart(E).(27) In this work the integral is performed by Romberg integration [Citation27], using an integration interval from 0 to 100 ω and a convergence criterion of 1010.

The reactants are represented as a free particle, which has the partition function per unit of length (28) QR=m2πβ1.(28) This partition function is the same in the classical and quantum mechanical cases.

To obtain exact classical Wigner flux-Heaviside traces for the Eckart potential the numerical matrix multiplication scheme [Citation25] was used to calculate the matrix elements of the Boltzmann operator. An RPMD comparison was calculated using the formulation of Craig and Manolopoulos [Citation28].

6. Results and discussion

In Figures  Cfs(t) for the Eckart barrier at βω=1, 3, and 6 are shown for the x and y versions of OPCW. In Figure  we further show the x version result of OPCW for βω=12. At this low temperature, the convergence of the OPCW result with respect to the number of Monte Carlo steps becomes challenging and we chose a lower value of N to alleviate this. The rate constants and tunnelling factors, Γ=kr,QM/kr,CM, are presented in Tables  and .

Figure 1. Cfs(t) for an Eckart potential with βω=1. The traces are calculated with the y-version of OPCW for different N, exact Classical Mechanics (CM), exact Classical Wigner (CW), and exact Quantum Mechanics (QM). The QM-line only shows the result in the long time limit (not the time dependent function). The numbers in parenthesis are the number of Monte Carlo steps used. The top and bottom lines of each type show the standard deviation of the results (but uncertainties are often so small that the lines are not clearly distinguished).

Figure 1. Cfs(t) for an Eckart potential with βℏω=1. The traces are calculated with the y-version of OPCW for different N, exact Classical Mechanics (CM), exact Classical Wigner (CW), and exact Quantum Mechanics (QM). The QM-line only shows the result in the long time limit (not the time dependent function). The numbers in parenthesis are the number of Monte Carlo steps used. The top and bottom lines of each type show the standard deviation of the results (but uncertainties are often so small that the lines are not clearly distinguished).

Figure 2. As in Figure  but for the x-version of OPCW.

Figure 2. As in Figure 1 but for the x-version of OPCW.

Figure 3. As in Figure  but for βω=3.

Figure 3. As in Figure 1 but for βℏω=3.

Figure 4. As in Figure  but for βω=3 and the x-version of OPCW.

Figure 4. As in Figure 1 but for βℏω=3 and the x-version of OPCW.

Figure 5. As in figure but for βω=6.

Figure 5. As in figure 1 but for βℏω=6.

Figure 6. As in figure but for βω=6 and the x-version of OPCW.

Figure 6. As in figure 1 but for βℏω=6 and the x-version of OPCW.

Figure 7. As in figure but for βω=12 and the x-version of OPCW.

Figure 7. As in figure 1 but for βℏω=12 and the x-version of OPCW.

Table 1. Rate constants (kr) for different inverse temperatures (β) for the Eckart barrier described in Equation Equation24. Comparision of OPCW[x-version, N = 40], Ring Polymer Molecular Dynamics (RPMD), classical mechanics (CM), exact Classical Wigner (CW), and quantum mechanics (QM). Note that for β=12 only N = 20 was used for OPCW.

Table 2. Tunneling factors (Γ) for different inverse temperatures (β) for the Eckart barrier described in Equation Equation24. Comparision of OPCW[x-version, N = 40], Ring Polymer Molecular Dynamics (RPMD), exact Classical Wigner (CW), and exact results. Note that for β=12 only N = 20 was used for OPCW.

At βω=1 the quantum mechanical rate constant is only 6% bigger than the classical one. For the three lower temperatures the differences are significantly bigger, i.e. the quantum mechanical rate constant is 50% bigger for βω=3, a factor 5 bigger for βω=6 and finally a factor of 2000 bigger at the lowest temperature. The classical Wigner method gives rate constants that are intermediate between classical mechanics and quantum mechanics, but for the three lower temperatures they represent a significant improvement over the classical result.

The OPCW results converge to the exact classical Wigner result as the number of beads is increased. From Figures  it can be seen that the y-version of the OPCW method needs more beads than the x-version. The x-version of the OPCW method has excellent convergence with respect to the number of beads. Cfs,y(t) converge from above and Cfs,x(t) converge from below. Both the x and y versions of OPCW give results that are close to the exact classical Wigner result.

The RPMD and classical Wigner methods give similar results for all temperatures. Results from a single, symmetric, one-dimensional, potential is, however, not enough to draw general conclusions.

7. Conclusion and outlook

In this work, the recently developed Open Polymer Classical Wigner (OPCW) method [Citation13] is extended to the flux-Heaviside trace of Miller [Citation15] and applied to a symmetric Eckart potential, for four temperatures. For all but the lowest temperature, the OPCW method can be converged to the exact classical Wigner result by increasing the number of beads in the path integral polymer. For the lowest temperature, a strict convergence of the OPCW result was computationally too demanding, but a result some thirty percent from the exact Classical Wigner model was obtained, see Figure . This should be compared to the classical prediction that is about two thousand times smaller than exact quantum mechanics. Particularly the x-version of the OPCW method shows promise for further tests.

In the future, it would be of interest to apply the OPCW model to reaction rate constant calculations of multidimensional systems, such as the double well potential of Topaler and Makri [Citation29] coupled to the harmonic bath of Caldeira and Leggett [Citation30]. For such a system it could be beneficial to modify the OPCW calculations by making the bath classical, as illustrated in the preceding article [Citation13]. There a ten-dimensional problem was considered. A more elaborate approach would be to keep an intermediate number of beads for the bath degrees of freedom, while still retaining the full number for the reaction coordinate. Such an approach will be the subject of future research. It would also be of interest to try the method on asymmetric potentials. Such potentials are challenging not just for the OPCW method but for all classical Wigner implementations, since the positioning of the dividing surface is not obvious. However, Liu and Miller have shown that this can be done successfully [Citation4].

For systems where the reactants are bound it could be of interest to use the Heaviside-Heaviside trace of Miller et al. [Citation14] instead of the flux-Heaviside trace.

Acknowledgments

Financial support from the Swedish Research Council (Vetenskapsrådet), diary numbers 2016-03275 and 2020-05293, is acknowledged. The computations in this work were performed at Chalmers Centre for Computational Science and Engineering (C3SE), a partner center of the Swedish National Infrastructure for Computing (SNIC).

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

Swedish Research Council 2020-05293; Swedish Research Council 2016-03275

References

Appendices

1

To obtain an expression corresponding to the x1-version of OPCW we start from equation Equation8. There the xˆ-dependent parts of Fˆ have to operate to the right in the matrix element before rewriting as Wigner transforms. To be able to operate to the right, the flux operator has to be rearranged into a form where all xˆ-dependence is to the right of all pˆ-dependence. Using that pˆ=iddx this can be achieved as (A1) Fˆ=12mδ(xˆ)pˆ+pˆδ(xˆ)=12mδ(xˆ),pˆ+pˆδ(xˆ)+pˆδ(xˆ)=12miδ(xˆ)ddxddxδ(xˆ)+2pˆδ(xˆ)=12miδ(xˆ)ddxdδ(xˆ)dxδ(xˆ)ddx+2pˆδ(xˆ)=12midδ(xˆ)dx+2pˆδ(xˆ),(A1) where the square brackets denote a commutator. Using this new form of the flux operator and operating xˆ to the right we obtain (A2) Cfs(t)=j=1Ndxjx1|eβNHˆ|x2x2|eβNHˆ|x3xN2|eβNHˆpˆ|xN2+11mδxN2+1dδxN2+1dxN2+1+xN2|eβNHˆ|xN2+1i2mdδxN2+1dxN2+1xN1|eβNHˆ|xNxN|eβNHˆeiHˆth(xˆ)eiHˆt|x1.(A2) Rewriting as Wigner transforms, taking the limit where the Wigner transform of the Boltzmann operator is the classical Boltzmann factor, and assuming that V(yj) is independent of momentum, a new way to write Equation Equation11 can be obtained: (A3) Cfs(t)=limNmN2πβN2β2πmNNj=1Ndxj×dpN2dpNeβNj=1NVyj×emN22βj=1N21xj+1xj2+j=N2+1N1xj+1xj2×eipN2xN2+1xN2eβNpN222m×pˆWyN2,pN21mδxN2+1+i2mdδxN2+1dxN2+1×eipNx1xNeβNpN22meiHˆth(xˆ)eiHˆtWyN,pN.(A3)

Knowing that pˆWx,p=p, pN2 can be integrated out as (A4) dpN2pˆWyN2,pN2eβNpN222meipN2xN2+1xN2dpN2eβNpN222meipN2xN2+1xN2=dpN2pN2eβNpN222meipN2xN2+1xN22πmNβeβNmN2xN2+1xN2222β2=iNmβxN2+1xN2(A4) and we get (A5) Cfs(t)=limNmN2πβN2β2πmNN×j=1NdxjdpNeβNj=1NVyj×emN22βj=1N1xj+1xj2×βzβNγ,k(μkz)μk2iNβxN2+1xN2δxN2+1dδxN2+1dxN2+1+i2mdδxN2+1dxN2+1eipNx1xN×eβNpN22meiHˆth(xˆ)eiHˆtWyN,pN.(A5) If f(x) is a function of x then dxf(x)dδ(x)dx=dx(df(x)dx)δ(x). This can be used to eliminate the derivative of the delta function through (A6) dxN2+1eβNVyN2+VyN2+1×emN22βxN2+1xN22+xN2+2xN2+12i2mdδxN2+1dxN2+1=dxN2+1eβNVyN2+VyN2+1×emN22βxN2+1xN22+xN2+2xN2+12i2mδxN2+1×(1)βN12dVyN2dyN2+12dVyN2+1dyN2+1dVyN2+1dyN2+1βzβNγ,k(μkz)μk2mN22β2xN2+1xN22xN2+2xN2+1=dxN2+1eβNVyN2+VyN2+1×emN22βxN2+1xN22+xN2+2xN2+12×iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1βzβNγ,k(μkz)μk2iN2β2xN2+1xN2xN2+2δxN2+1,(A6) leading to (A7) Cfs(t)=limNmN2πβN2β2πmNNj=1Ndxj×dpNeβNj=1NVyjemN22βj=1N1xj+1xj2×iNβxN2+1xN2δxN2+1dVyN21dxN2+1+iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1iN2β2xN2+1xN2xN2+2δxN2+1βzβNγ,k(μkz)μk2×eipNx1xNeβNpN22meiHˆth(xˆ)eiHˆtWyN,pN=limNmN2πβN2β2πmNNj=1Ndxj×dpNeβNj=1NVyjemN22βj=1N1xj+1xj2×iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1iN2β2xN2+1xN2xN2+22xN2+1+2xN2βzβNγ,k(μkz)μk2×δxN2+1eipNx1xN×eβNpN22meiHˆth(xˆ)eiHˆtWyN,pN=limNmN2πβN2β2πmNNj=1Ndxj×dpNeβNj=1NVyjemN22βj=1N1xj+1xj2×iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1iN2βxN2xN2+2βzβNγ,k(μkz)μk2δxN2+1×eipNx1xNeβNpN22m×eiHˆth(xˆ)eiHˆtWyN,pN.(A7) Up to this point the new derivation is equivalent to what was done for the y-version up to Equation Equation14. However, once this expression is subjected to the classical Wigner approximation and by assuming finite N, this is not the case anymore, as we obtain (A8) Cfs,x(t)mN2πβN2β2πmNN×j=1NdxjdpNeipNx1xN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2×iβ4NmdVyN2dyN2+dVyN2+1dyN2+1+dVyN2+1dyN2+1iN2βxN2xN2+2βzβNγ,k(μkz)μk2δxN2+1×hxyN,pN,t.(A8) Equation EquationA8 is the expression corresponding to the x1-version of OPCW for the flux-Heaviside trace. Due to the symmetrisation of the trace xN2+1 rather than x1 is used. This trace is called Cfs,x(t) in the main text.

2

Here we show how to calculate the Zδ factor using a Monte-Carlo implementation of the effective frequency theory of Feynman and Kleinert (FK). For an introduction to the FK theory, we refer the reader to Hagen Kleinerts book [Citation31]. The theory develops approximations to path integrals involving the Boltzmann operator. The factor Zδ is given by Equation Equation22: (A9) Zδ=mN2πβN2β2πmNNj=1NdxjdpN×eβNpN22m+j=1NVyj+mN222β2j=1N1xj+1xj2×eipNx1xNδyy,x.(A9) By using δ(yy,x) we have chosen the dividing surface to lie at x=0. The variable pN may be integrated out giving Zδ=mN2πβN2Nj=1Ndxj×eβNj=1NV(yj)+mN222β2j=1N(x(jmodN)+1xj)2δ(yy,x).

This equation can be recognised as simply (A10) Zδ=0|eβHˆ|0.(A10) In multi-dimensional problems, it is necessary to evaluate equation EquationA10 by Monte Carlo methods. In the FK formulation, equation EquationA10 takes the form (see page 477 in [Citation31]) (A11) Zδ=dxc12πa2(xc)12π2πmβeβW1(xc)e12xc2/a2(xc).(A11) Here, W1(xc) is the so-called centroid potential and a2(xc) is the smearing width, both defined for a certain value of the path centroid xc. Feynman and Kleinert showed how to determine W1(xc) and a2(xc) as a function of xc. To implement equation EquationA11 by Monte Carlo, we need a weight-function f(xc). Let us denote the integrand of equation EquationA11 by g(xc). We first note that g(xc) is symmetric around xc=0 for our Eckart potentials. We can get a number xc2 that characterises the spread of g(xc) by performing a Monte Carlo evaluation of (A12) xc2g=dxcxc2g(xc)dxcg(xc),(A12) by using g(xc) as weight function. Afterwards we construct a gaussian function f(xc)=eβkxc2, choosing k so that f(xc) has the same xc2 if we substitute g for f in Eq. EquationA12. Then we rewrite Eq. EquationA11 as (A13) Zδ=dxcg(xc)f(xc)f(xc)=dxcf(xc)×g(xc)f(xc)f=πβk×g(xc)f(xc)f,(A13) where f is defined as in equation EquationA12. This equation is again evaluated by Monte Carlo. In Table  below is shown the results of evaluating equation EquationA10 exactly and by using equation EquationA13. We have also included the low temperature case βω=12. We further show the result obtained by evaluating equation EquationA11 directly through a one-dimensional integration.

Table A1. Zδ for the Eckart potential at various temperatures. Exact results were obtained using Eq. EquationA10 and the numerical matrix multiplication scheme. FK Monte Carlo results were obtained from Eq. EquationA13 using 15000 MC steps. In each case, the variance was estimated using three independent calculations. In parenthesis is shown the force constant of the reference function f(xc).

It is seen that the FK and accurate results are quite close to each other.