• Tidak ada hasil yang ditemukan

Chapter I: Cayley modification for path-integral simulations

1.3 Theory

The theory introduced here adapts and advances previous mathematical results on the numerical approximation of general second order Langevin stochastic partial differential equations with space-time white noise.39

RPMD

We consider a quantum particle in 1D with Hamiltonian operator given by ห†

๐ป = ๐‘ห†2 2๐‘š

+๐‘‰(๐‘žห†) (1.3)

where ห†๐‘ž, ห†๐‘, and ๐‘š represent the particle position, momentum, and mass, respec- tively, and๐‘‰(๐‘žห†) is a potential energy surface. All results presented here are easily generalized to multiple dimensional quantum systems.

The thermal equilibrium properties of the system are described by the quantum mechanical Boltzmann partition function,

๐‘„ =Tr[๐‘’โˆ’๐›ฝ๐ปห†] , (1.4)

where๐›ฝ= (๐‘˜๐ต๐‘‡)โˆ’1is the inverse temperature. Using a path-integral discretization, ๐‘„can be approximated by a classical partition function๐‘„๐‘›of a ring-polymer with ๐‘›beads,5

๐‘„๐‘›= ๐‘š๐‘› (2๐œ‹โ„)๐‘›

โˆซ ๐‘‘๐‘›๐’’

โˆซ

๐‘‘๐‘›๐’—๐‘’โˆ’๐›ฝ ๐ป๐‘›(๐’’,๐’—) , (1.5) where๐’’ = (๐‘ž0, . . . , ๐‘ž๐‘›โˆ’1)is the vector of bead positions, and๐’—is the corresponding vector of velocities. The ring-polymer Hamiltonian is given by

๐ป๐‘›(๐’’,๐’—) =๐ป0

๐‘›(๐’’,๐’—) +๐‘‰ext

๐‘› (๐’’), (1.6)

which includes contributions from the physical potential ๐‘‰ext

๐‘› (๐’’) = 1 ๐‘›

๐‘›โˆ’1

โˆ‘๏ธ

๐‘—=0

๐‘‰(๐‘ž๐‘—) (1.7)

and the free ring-polymer Hamiltonian ๐ป0

๐‘›(๐’’,๐’—) = ๐‘š๐‘› 2

๐‘›โˆ’1

โˆ‘๏ธ

๐‘—=0

h ๐‘ฃ2

๐‘—+๐œ”2

๐‘›(๐‘ž๐‘—+1โˆ’๐‘ž๐‘—)2i

, (1.8)

where ๐‘š๐‘› = ๐‘š/๐‘›, ๐œ”๐‘› = ๐‘›/(โ„๐›ฝ) and ๐‘ž๐‘› = ๐‘ž0. If we let ๐‘› = 1 in Eq. (1.5), the classical partition function of the system (governed by a classical Hamiltonian, Eq. 1.6 with ๐‘› = 1) is recovered, i.e., ๐‘„1 = ๐‘„๐‘๐‘™. In the limit ๐‘› โ†’ โˆž, the path- integral approximation converges to the exact quantum Boltzmann statistics for the system, such that๐‘„โˆž =๐‘„. The thermal ensemble of ring-polymer configurations associated with Eq. (1.5) can be sampled using either molecular dynamics (leading to PIMD methods) or Monte Carlo (leading to PIMC methods).

The classical equations of motion associated with the ring-polymer Hamiltonian in Eq. (1.6),

ยค

๐‘ž๐‘— =๐‘ฃ๐‘—, (1.9)

ยค ๐‘ฃ๐‘— =๐œ”2

๐‘›(๐‘ž๐‘—+1+๐‘ž๐‘—โˆ’1โˆ’2๐‘ž๐‘—) โˆ’ 1 ๐‘š

๐‘‰โ€ฒ(๐‘ž๐‘—),

yield the RPMD model for the real-time dynamics of the system.15,16 RPMD provides a means of approximately calculating Kubo-transformed thermal time- correlation functions, such as the position autocorrelation function

หœ

๐ถ๐‘ž ๐‘ž(๐‘ก) = 1 ๐‘„

Tr[๐‘’โˆ’๐›ฝ๐ปห†๐‘žหœ(0)๐‘žห†(๐‘ก)] (1.10)

where the Kubo-transformed position operator หœ๐‘žis

หœ ๐‘ž = 1

๐›ฝ

โˆซ ๐›ฝ

0

๐‘’๐œ†๐ปห†๐‘ž ๐‘’ห† โˆ’๐œ†๐ปห†๐‘‘๐œ† (1.11) and the time-evolved operator ห†๐‘ž(๐‘ก) is๐‘’๐‘–๐ป ๐‘กห† /โ„๐‘ž ๐‘’ห† โˆ’๐‘–๐ป ๐‘กห† /โ„.

Specifically, the RPMD approximation to Eq. (1.10) is ๐ถหœ๐‘ž ๐‘ž(๐‘ก) = 1

๐‘„๐‘›

โˆซ ๐‘‘๐‘›๐’’

โˆซ

๐‘‘๐‘›๐’—๐‘’โˆ’๐›ฝ ๐ป๐‘›(๐’’,๐’—)๐‘žยฏ(0)๐‘žยฏ(๐‘ก) (1.12) where ยฏ๐‘žis the bead-averaged position

ยฏ

๐‘ž(๐‘ก) = 1 ๐‘›

๐‘›โˆ’1

โˆ‘๏ธ

๐‘—=0

๐‘ž๐‘—(๐‘ก) , (1.13)

and the pair(๐’’(๐‘ก),๐’—(๐‘ก))are evolved by the RPMD equations of motion in Eq. (1.9) with initial conditions drawn from the classical Boltzmann-Gibbs measure.

The RPMD equations of motion can be compactly rewritten as ๐’’ยค

๐’—ยค

= ๐‘จ ๐’’

๐’—

+

0 ๐‘ญ(๐’’)/๐‘š๐‘›

, where ๐‘จ =

0 ๐‘ฐ

โˆ’๐›€2 0

, (1.14)

๐‘ญ(๐’’) = โˆ’โˆ‡๐‘‰ext

๐‘› (๐’’), ๐‘ฐ is an๐‘›ร—๐‘›identity matrix,0is an array of zeros, and๐›€2is the๐‘›ร—๐‘›symmetric positive semi-definite matrix

๐›€2=๐œ”2

๐‘›

๏ฃฎ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฏ

๏ฃฐ

2 โˆ’1 0 ยท ยท ยท 0 โˆ’1

โˆ’1 2 โˆ’1 0 ยท ยท ยท 0 ..

. ..

. ..

. ...

... ... 0 ยท ยท ยท 0 โˆ’1 2 โˆ’1

โˆ’1 0 ยท ยท ยท 0 โˆ’1 2

๏ฃน

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃบ

๏ฃป

. (1.15)

We recognize ๐›€2 as the 1D discrete Laplacian endowed with periodic boundary conditions; its spectral radius that scales as๐‘›2, and since๐›€2 is circulant, it can be diagonalized by the๐‘›ร—๐‘›orthogonal real discrete Fourier transform (DFT) matrix.

In particular, the spectral decomposition of๐›€can be written as

๐›€=๐‘ผ๐›€๐’…๐‘ผT, where๐›€๐’… =diag(0, ๐œ”1,๐‘›, . . . , ๐œ”๐‘›โˆ’1,๐‘›) (1.16) is a diagonal matrix of eigenvalues given by

๐œ”๐‘— ,๐‘› =

๏ฃฑ๏ฃด

๏ฃด

๏ฃฒ

๏ฃด๏ฃด

๏ฃณ

2๐œ”๐‘›sin ๐œ‹ ๐‘—

2๐‘›

if ๐‘— is even, 2๐œ”๐‘›sin

๐œ‹(

๐‘—+1) 2๐‘›

else,

(1.17)

In nontrivial applications, the RPMD equations of motion in Eq. (1.14) cannot be solved analytically. It is then necessary to employ approximate numerical integration of the equations of motion. As we discuss next, designing good numerical integrators for Eq. (1.14) is complicated by the interplay between the time-evolution of the free ring-polymer (obtained by setting๐‘ญ =0 in Eq. 1.14) and the contributions from the physical forces ๐‘ญ.

Cayley removes instabilities in a free ring-polymer mode

RPMD is an example of highly oscillatory Hamiltonian dynamics.40 To understand why numerical integration of such systems is tricky and why the Cayley modification is needed, it helps to consider the equations of motion for a particular normal mode of the free ring polymer with Matsubara frequency๐œ” >0:

๐‘žยค

ยค ๐‘ฃ

= ๐‘จ ๐‘ž

๐‘ฃ

where ๐‘จ=

0 1

โˆ’๐œ”2 0

, (1.18)

which are also the equations of motion for a linear oscillator with natural frequency ๐œ”. If ๐œ” is large, Eq. (1.18) is highly oscillatory. Solving Eq. (1.18) amounts to approximating the matrix exponential exp(ฮ”๐‘ก๐‘จ)whereฮ”๐‘กis a timestep size. A good 2ร—2 matrix approximation ๐‘ดฮ”๐‘กshould satisfy:

(P1) Accuracy โˆฅ๐‘ดฮ”๐‘กโˆ’exp(ฮ”๐‘ก๐‘จ) โˆฅ=๐‘‚(ฮ”๐‘ก3).

(P2) Strong Stability For all ๐œ” > 0, and for all ฮ”๐‘ก smaller than some constant independent of๐œ”, ๐‘ดฮ”๐‘ก is a strongly stable symplectic matrix.

(P3) Time-Reversibility For all๐œ” > 0 andฮ”๐‘ก > 0, ๐‘ดฮ”๐‘ก is reversible with respect to the velocity flip matrix ๐‘น=

1 0 0 โˆ’1

, i.e., ๐‘น ๐‘ดฮ”๐‘ก๐‘น =๐‘ดโˆ’1ฮ”๐‘ก.

We briefly comment on each of these criteria for a good approximation. Prop- erty (P1) is a basic requirement that ensures second-order accuracy on finite-time intervals. Property (P3) is particularly useful for sampling from the stationary distribution, since a reversible map can be readily Metropolized41โ€“43, and since time-reversibility in a volume-preserving numerical integrator leads to a doubling of the accuracy order (see Propositions 5.2 and Theorem 6.2 of Ref.43, respectively).

Property (P2) is the most interesting. A symplectic matrix๐‘บ is stable if all powers of the matrix๐‘บare bounded. A symplectic matrix๐‘บisstrongly stableif๐‘บis stable and all sufficiently close symplectic matrices are also stable. In other words, ๐‘บ is strongly stable if there exists an๐œ– > 0, such that all symplectic matrices ๐‘บ๐œ– that are within a distance๐œ– away from ๐‘บare also stable. A sufficient condition for ๐‘บ to be strongly stable is that the eigenvalues of๐‘บare on the unit circle in the complex plane and are distinct; both the necessary and sufficient conditions for strong stability of symplectic matrices are known.44

-1 0 1 real -1

0 1

imaginary

-1 0 1

real -1

0 1

imaginary

(a)๐‘ก = ๐œ‹3 (b)๐‘ก = ๐œ‹4

Figure 1.1: Eigenvalues of2ร—2symplectic matrices.Eigenvalues of a symplectic matrix ๐‘บ = exp(๐‘ก๐‘จ) (black dots) are plotted in the complex plane along with eigenvalues of a perturbed symplectic matrix ๐‘บ๐œ– = exp( (1/2)๐‘ก๐‘ฉ)exp(๐‘ก๐‘จ)exp( (1/2)๐‘ก๐‘ฉ) (grey dots). The elements of ๐‘จ and ๐‘ฉ are specified in the text. For both values of๐‘ก, ๐‘บ is stable since its eigenvalues lie on the unit circle. When the eigenvalues of๐‘บare not distinct, then as shown in (a),๐‘บ๐œ– has an eigenvalue with modulus greater than one, and hence, ๐‘บ๐œ– loses stability.

However, if the eigenvalues of๐‘บare distinct, then๐‘บis strongly stable, and as shown in (b), ๐‘บ๐œ– is stable since its eigenvalues remain on the unit circle.

Figure 1.1 illustrates the concept of strong stability. In particular, for different values of๐‘ก(as indicated in each panel), the black dots correspond to the eigenvalues of the symplectic matrix๐‘บ=exp(๐‘ก๐‘จ)with๐œ” =3, and the grey dots are the eigenvalues of a perturbation of๐‘บwhich preserves the symplectic nature of the matrix, specifically ๐‘บ๐œ– = exp( (1/2)๐‘ก๐‘ฉ)exp(๐‘ก๐‘จ)exp( (1/2)๐‘ก๐‘ฉ) where ๐‘ฉ =

0 ๐œ– ๐œ– 0

and ๐œ– = 0.15. For any๐‘ก, note that the two eigenvalues of๐‘บare always on the unit circle, and hence, ๐‘บ is always stable, but as the figure shows,๐‘บis not always strongly stable. Indeed, in Figure 1.1 (a), we see that the two eigenvalues of ๐‘บ, represented by a single black dot, are both equal to (โˆ’1,0), which violates the condition for strong stability, and in this case, we see that one of the eigenvalues of๐‘บ๐œ– has modulus greater than unity, which implies that ๐‘บ๐œ– is unstable. In Figure 1.1 (b), the two eigenvalues of ๐‘บ are distinct and equal to (0,ยฑ1), and hence, ๐‘บ is strongly stable. Since ๐‘บ is strongly stable, and ๐œ– is sufficiently small, ๐‘บ๐œ– has eigenvalues that are on the unit circle, and hence, is itself stable. For a more detailed discussion of the concept of strong stability of symplectic matrices, see Section 42 of Ref.45.

A natural candidate for an approximation๐‘ดฮ”๐‘กthat satisfies these criteria is the Verlet integrator, which is ubiquitous in the classical simulation of molecular systems.46โ€“49 For a single Matsubara frequency of the free ring polymer, the Verlet integrator gives

๐‘ดฮ”๐‘ก =

"

1โˆ’ ฮ”๐‘ก2๐œ”2

2 ฮ”๐‘ก

โˆ’1

2ฮ”๐‘ก ๐œ”2(2โˆ’ ฮ”๐‘ก2๐œ”2

2 ) 1โˆ’ ฮ”๐‘ก2๐œ”2

2

#

. (1.19)

However, forฮ”๐‘ก > 2/๐œ”, the eigenvalues of ๐‘ดฮ”๐‘ก are real and distinct, so that one of them has modulus> 1, and therefore the powers of ๐‘ดฮ”๐‘ก grow exponentially. Thus,

-1 0 1 real -1

0 1

imaginary

-1 0 1

real -1

0 1

imaginary

(a) exp(ฮ”๐‘ก๐‘จ) (b) cay(ฮ”๐‘ก๐‘จ)

Figure 1.2: Eigenvalues of the exponential vs. Cayley maps. Eigenvalues of exp(ฮ”๐‘ก๐‘จ) (a) and cay(ฮ”๐‘ก๐‘จ) (b) at 50 different timestep sizes between 0.05 and 5.0 (evenly spaced) and with๐œ”=3, color-coded from blue (smallest) through green and yellow to red (largest).

For exp(ฮ”๐‘ก๐‘จ), the eigenvalues rotate around the unit circle multiple times. However, for cay(ฮ”๐‘ก๐‘จ), the eigenvalues start near(1,0), but never reach (โˆ’1,0). Since the eigenvalues of cay(ฮ”๐‘ก๐‘จ)are always distinct, it provides strong symplectic stability, whereas the matrix exponential loses strong stability every time the eigenvalues hit the horizontal axis. In both panels, the eigenvalue associated with the ring-polymer centroid motion is excluded.

numerical stability requires ฮ”๐‘ก < 2/๐œ”, and Verlet does not satisfy (P2), since this numerical stability requirement is not uniform in๐œ”.

Surprisingly, the exact solution for the normal-mode dynamics also does not satisfy (P2). To see why, note that the eigenvalues of the matrix exponential exp(ฮ”๐‘ก๐‘จ)are ๐‘’ยฑ๐‘–๐œ”ฮ”๐‘กand (P2) requires that๐‘’๐‘–๐œ”ฮ”๐‘ก โ‰  ๐‘’โˆ’๐‘–๐œ”ฮ”๐‘กwhich is violated if and only if

ฮ”๐‘ก = ๐œ‹ ๐‘˜

๐œ” for all๐‘˜ โ‰ฅ 1 . (1.20)

At these timesteps, the exact solution violates strong stability. This is illustrated in Figure 1.2 (a), where the two eigenvalues of exp(ฮ”๐‘ก๐‘จ) are plotted in the complex plane for a range of time-step sizes. Although the two eigenvalues of exp(ฮ”๐‘ก๐‘จ) lie on the unit circle for allฮ”๐‘ก, strong stability fails to hold whenever the eigenvalues are both equal to(ยฑ1,0).

A simple strategy to avoid these artificial resonances is to use a random timestep size ๐›ฟ๐‘ก, e.g., take as timestep size an exponential random variable ๐›ฟ๐‘ก with mean ฮ”๐‘ก. Averaging exp(๐›ฟ๐‘ก๐‘จ) over the exponential probability density function yields ๐‘ดฮ”๐‘ก = E(exp(๐›ฟ๐‘ก๐‘จ)) = (๐‘ฐ โˆ’ ฮ”๐‘ก๐‘จ)โˆ’1, where here ๐‘ฐ is the 2 ร—2 identity matrix.

Unfortunately, as can be easily verified, this matrix satisfies none of our criteria:

it is neither symplectic, nor reversible, nor sufficiently accurate. However, we can easily turn this approximation into one that satisfies (P1), by simply composing 1/2 step of this integrator with 1/2 step of its adjoint ๐‘ดโˆ’1ฮ”๐‘ก. This correction yields the Cayley transform of the matrixฮ”๐‘ก๐‘จ,

cay(ฮ”๐‘ก๐‘จ) โ‰ก (๐‘ฐโˆ’ (1/2)ฮ”๐‘ก๐‘จ)โˆ’1(๐‘ฐ+ (1/2)ฮ”๐‘ก๐‘จ). (1.21)

In fact, the Cayley transform satisfies all three of the specified criteria for a good numerical integrator. It is time-reversible since ๐‘นcay(ฮ”๐‘ก๐‘จ)๐‘น = (๐‘น โˆ’ (1/2)ฮ”๐‘ก๐‘น ๐‘จ)โˆ’1(๐‘น + (1/2)ฮ”๐‘ก๐‘จ๐‘น) = cay(ฮ”๐‘ก๐‘จ)โˆ’1, where we used that ๐‘นโˆ’1 = ๐‘น.

It is a symplectic matrix since

cay(ฮ”๐‘ก๐‘จ)๐‘‡๐‘ฑcay(ฮ”๐‘ก๐‘จ) = ๐‘ฑ where ๐‘ฑ =

0 1

โˆ’1 0

where we used the fact that๐‘จis a Hamiltonian matrix (See Ref.50, Section 2.5). More importantly, it is a strongly stable symplectic matrix for all ฮ”๐‘ก > 0, as illustrated in Figure 1.2 (b); in contrast with the exponential map, for all ๐œ” > 0 and ฮ”๐‘ก > 0 the eigenvalues of the Cayley map are(4โˆ’ฮ”๐‘ก2๐œ”2ยฑ4๐‘–ฮ”๐‘ก ๐œ”)/(4+ฮ”๐‘ก2๐œ”2), which are distinct and of unit modulus. Thus, not only is every matrix power of cay(ฮ”๐‘ก๐‘จ) bounded, but the Cayley map is strongly stable uniformly in๐œ” andฮ”๐‘ก.

Cayley removes instabilities in microcanonical RPMD

For numerical integration of the conservative RPMD equations of motion (Eq. 1.9 or Eq. 1.14), it is standard practice9,16to employ a symmetrically split second-order integrator of the form in Eq. (1.2).

Furthermore, it is standard practice to exactly perform the free ring-polymer time evolution step,16using an exponential map of the form exp(ฮ”๐‘ก ๐ฟ0) =exp(ฮ”๐‘ก๐‘จ)where ๐‘จis the matrix associated with the dynamics of the free ring-polymer Hamiltonian,

๐’’ยค ๐’—ยค

= ๐‘จ ๐’’

๐’—

. (1.22)

In practice, the exact exponential map is executed by successively(i)changing from the Cartesian bead positions and velocities to the normal modes of the free ring polymer,(ii)numerically integrating each of the uncoupled normal mode equations of motion, and(iii)translating the time-evolved normal mode coordinates back into the Cartesian bead positions and velocities. Therefore, the numerical stability of standard RPMD numerical integration may be analyzed in normal mode coordinates, where the free ring-polymer equations of motion in Eq. (1.22) decouple into a system of๐‘›independent oscillators with natural frequencies given by the eigenvalues of the matrix๐›€in Eq. (1.17).

By applying Eq. (1.20) to each normal mode coordinate, we find that strong stability of the exact free ring-polymer time evolution is violated when

ฮ”๐‘ก = ๐œ‹ ๐‘˜

๐œ”๐‘— for all๐‘˜ โ‰ฅ 1 and 1โ‰ค ๐‘— โ‰ค ๐‘›โˆ’1. (1.23) Unstable pairs of ฮ”๐‘ก and ๐‘› are plotted using solid lines in Fig. 1.3(b) for selected values of ๐‘— and ๐‘˜. The horizontal asymptotes in this figure reflect the fact that the eigenvalues of๐›€converge to the eigenvalues of the continuous Laplacian endowed with periodic boundary conditions.

Unlike the exact free ring-polymer step used in standard RPMD numerical integra- tion, the Cayley modification exp(ฮ”๐‘ก ๐ฟ0) โ‰Šcay(ฮ”๐‘ก๐‘จ)is strongly stable for allฮ”๐‘ก >0 uniformly in๐‘›. To see this, note that the Cayley transform can be equivalently com- puted in either bead or normal mode coordinates. More precisely, let๐›€2 =๐‘ผ๐›€2

๐’…๐‘ผT by the spectral decomposition of๐›€ given in Eq. (1.16). Direct computation then shows that

cay(ฮ”๐‘ก๐‘จ) =

๐‘ผ 0 0 ๐‘ผ

cay

ฮ”๐‘ก

0 ๐‘ฐ

โˆ’๐›€2

๐’… 0

๐‘ผT 0 0 ๐‘ผT

.

Using this correspondence, one can invoke the preceding results on the one- dimensional oscillator, to conclude that cay(ฮ”๐‘ก๐‘จ)is second-order accurate, strongly stable symplectic, and time-reversible.

Since the Cayley transform meets our criteria (P1)-(P3), and under suitable condi- tions on the force ๐‘ญ, the Cayley modification to the RPMD numerical integrator is provably stable and second-order accurate on finite-time intervals with a stability requirement that is uniform with respect to the number of ring polymer beads. On the other hand, standard RPMD integrators may display artificial resonance instabilities because the free RP step is not always strongly stable. These insta- bilities often manifest as exponential growth in energy when strong stability is lost, as will be discussed in Section 1.4.

We emphasize that the improved numerical stability of the Cayley modification comes at zero cost in terms of algorithmic complexity or computational expense, and it preserves the same order of accuracy for the overall timestep. Use of this improved integration algorithm simply involves replacing the exact normal mode free ring-polymer step in the standard RPMD integrator with the Cayley modification.

Algorithmic comparison: Standard vs. Cayley

For complete clarity, we now present a side-by-side comparison of the full RPMD timestep (Eq. 1.2) with the free ring-polymer motion exp(ฮ”๐‘ก ๐ฟ0)implemented using either the standard exponential map (i.e., exact normal mode evolution) or via the Cayley modification. In both cases, the full RPMD timestep associated with the splitting in Eq. (1.2) is implemented using the algorithm

Velocity half-step: ๐’— โ†๐’—+ ฮ”๐‘ก

2 ๐‘ญ ๐‘š๐‘›

Free ring-polymer step: (๐’’,๐’—) โ†FRP(๐’’,๐’—;ฮ”๐‘ก) Force evaluation: ๐‘ญ=โˆ’โˆ‡๐‘‰ext

๐‘› (๐’’)

Velocity half-step: ๐’— โ†๐’—+ ฮ”๐‘ก

2 ๐‘ญ ๐‘š๐‘›

(1.24)

In standard RPMD numerical integration, the free ring-polymer step is performed exactly, using:

1. Convert bead Cartesian coordinates to normal modes using the orthogonal transformation:

๐”=๐‘ผ๐’’ and ๐‹=๐‘ผ๐’— (1.25)

where๐‘ผis the real DFT matrix defined in Eq. (1.16).

2. From ๐‘ก to ๐‘ก + ฮ”๐‘ก, exactly evolve the free ring polymer in the normal mode coordinates:

๐œš๐‘—(๐‘ก+ฮ”๐‘ก) ๐œ‘๐‘—(๐‘ก+ฮ”๐‘ก)

=exp(ฮ”๐‘ก๐‘จ๐‘—) ๐œš๐‘—(๐‘ก)

๐œ‘๐‘—(๐‘ก)

(1.26) where

๐‘จ๐‘— =

0 1

โˆ’๐œ”2

๐‘— 0

,

for 0โ‰ค ๐‘— โ‰ค ๐‘›โˆ’1 with๐œ”๐‘— defined in Eq. (1.17).

3. Convert back to bead Cartesian coordinates using the inverse of๐‘ผ, which is just its transpose, since๐‘ผis orthogonal.

In the Cayley modification, the only change is to use the following in place of Eq. (1.26):

๐œš๐‘—(๐‘ก+ฮ”๐‘ก) ๐œ‘๐‘—(๐‘ก+ฮ”๐‘ก)

=cay(ฮ”๐‘ก๐‘จ๐‘—) ๐œš๐‘—(๐‘ก)

๐œ‘๐‘—(๐‘ก)

, (1.27)

where cay is the Cayley transform given in Eq. (1.21).