Which Is Better at Predicting Quantum-Tunneling Rates: Quantum

Oct 27, 2014 - Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge CB2 1EW, United Kingdom. ‡. Ruđer BoÅ¡ković Institute ...
0 downloads 0 Views 569KB Size
Subscriber access provided by ISTANBUL UNIVERSITESI

Letter

Which is better at predicting quantum-tunneling rates: quantum transition-state theory or free-energy instanton theory? Yanchuan Zhang, Thomas Stecher, Marko Cvitas, and Stuart C. Althorpe J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/jz501889v • Publication Date (Web): 27 Oct 2014 Downloaded from http://pubs.acs.org on November 2, 2014

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Which is better at predicting quantum-tunneling rates: quantum transition-state theory or free-energy instanton theory? Yanchuan Zhang,† Thomas Stecher,† Marko T. Cvitaš,‡ and Stuart C. Althorpe∗,† Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge CB2 1EW, United Kingdom, and

b

Ruđer Bošković Institute, Department of Physical Chemistry,

Bijenička Cesta 54, 10000 Zagreb, Croatia. E-mail: [email protected]

Abstract Quantum transition-state theory (QTST) and free-energy instanton theory (FEIT) are two closely-related methods for estimating the quantum rate coefficient from the free-energy at the reaction barrier. In calculations on one-dimensional models, FEIT typically gives closer agreement than QTST with the exact quantum results at all temperatures below the crossover to deep tunneling, suggesting that FEIT is a better approximation than QTST in this regime. Here we show that this simple trend does not hold for systems of greater dimensionality. We report tests on several collinear and three-dimensional reactions, in which QTST outperforms FEIT over a range of temperatures below cross-over, which can extend down to half the cross-over temperature ∗

To whom correspondence should be addressed University of Cambridge ‡b Ruđer Bošković Institute †

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(below which FEIT outperforms QTST). This suggests that QTST-based methods such as ring-polymer molecular dynamics (RPMD) may often give closer agreement with the exact quantum results than FEIT.

Keywords: Quantum tunneling, dynamics

2 ACS Paragon Plus Environment

Page 2 of 17

Page 3 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Proton-transfer reactions involving tunneling through a barrier are widespread in chemistry, in fields ranging from gas-phase reactions 1–4 to enzyme catalysis. 5–7 Tunneling can accelerate the rates of such reactions by orders of magnitude and can also give rise to unusual isotope effects. Conventional (i.e. Wigner-Eyring-Polanyi 8 ) TST cannot treat tunneling, and an exact quantum treatment 9 is possible only for the simplest gas-phase reactions. There are therefore a variety of approximate methods, 10–27 which aim to treat quantum tunneling in various temperature regimes, an important class of which are those based on evaluating a path-integral free-energy in the region of the reaction barrier. 19–27 These methods scale much better with system size than exact quantum dynamics, since the simulations can be done using the ‘ring-polymer’ techniques first used to calculate static properties. 28,29 There are two main approaches behind such free-energy methods. The older approach is based on the so-called ‘Im F’ expression, 16–19 which is obtained by analytically continuing the partition function into the complex plane. To our knowledge, there is no clear derivation of this approach from first principles, although it has been used extensively in various branches of physics, 30 and its complete steepest-descent approximation gives the famous ‘instanton’ rate derived earlier by Miller from the quantum flux-side correlation function; 15 its partial steepest-descent implementation gives what we will refer to here as the free-energy instanton theory (FEIT), first used by Mills and coworkers. 27 The other approach is quantum transition-state theory (QTST), which was originally obtained heuristically, first in the special case of a centroid constraint, 20–22 then more generally as the TST limit of ring-polymer molecular dynamics (RPMD) rate-theory. 19,23,24 However QTST was shown recently to have a rigorous derivation as the short-time limit of the quantum flux through a path-integral dividing surface. 31–33 The similarities of the QTST and FEIT rate expressions will be explored later in this letter. Previous work, 19 published before the derivation of QTST, reported comparisons between FEIT and QTST for one-dimensional reaction barriers, which showed that FEIT gave closer agreement than QTST with the exact quantum rate-coefficients, at all temperatures below

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 17

the cross-over temperature Tc , 17

Tc =

h ¯ ω‡ , 2πkB

(1)

where ω ‡ is the barrier frequency and kB is the Boltzmann constant. We will refer to this range of temperatures as the ‘deep-tunneling’ regime.∗ Above Tc , the Boltzmann matrix is dominated by fluctuations around a point on top of the barrier; below Tc , it is dominated by fluctuations around the delocalised ‘instanton’ path, which is a periodic orbit on the inverted potential well. These results suggested that QTST may be an approximation to FEIT. It was also found 19 that QTST consistently under-predicted the exact quantum rate for symmetric barriers, and over-predicted it for asymmetric barriers. Here we extend these comparisons to a set of collinear and three-dimensional reactions, to find out whether FEIT continues to give a better approximation than QTST to the exact quantum rate. As mentioned above, the QTST rate is identical to the TST limit of RPMD rate-theory, which corresponds to applying classical rate-theory to the ring-polymer Hamiltonian. 19,23,24 For a system with f degrees of freedom, whose classical Hamiltonian is

H=

f X p2i + V (x1 , ..., xf ) , 2m i i=1

(2)

the ring-polymer Hamiltonian is 23 f N X X p2i,j HN (x, p) = + UN (βN , x), 2mi i=1 j=1 ∗

Other definitions of this term are sometimes used in the literature.

4 ACS Paragon Plus Environment

(3)

Page 5 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

where

UN (βN , x) =

N X

V (x1,j , ..., xf,j )

(4)

j=1

+

f X i=1

N mi X (xi,j+1 − xi,j )2 . 2(βN h ¯ )2 j=1

is the ring-polymer potential energy surface, and βN = β/N with β = 1/kB T . Defining a dividing surface s(x) to be a function of the ring-polymer coordinates which divides reactants from products, one then takes the t → 0+ limit of the classical flux-side time-correlation function in ring-polymer space to obtain the QTST rate, 31 1 kQTST (β)Qr (β) = lim N →∞ (2π¯ h)N f

Z

Z dp

dx e−βN HN (p,x)

×δ[s(x)]vs (p, x)h[vs (p, x)].

(5)

where

vs (p, x) =

f N X X ∂s(x) pi,j i=1 j=1

∂xi,j mi

,

(6)

is the instantaneous flux through s(x). Quantum transition-state theory gives the exact quantum rate if there is no recrossing of s(x) by the quantum flux, 31,33 but it does not give a rigorous upper bound, since coherent recrossing may increase the rate by removing destructive interference. However, if realtime coherence effects are small (as is likely for single-surface chemical reactions at room temperatures), QTST gives a good approximation to an upper bound, in which case the optimal dividing surface s(x) is the surface that minimizes kQTST . At temperatures above the cross-over temperature Tc , the optimal dividing surface can usually be well-approximated

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 17

by

s(x) = q 0 − s0 ,

where q 0 =

PN i

(7)

qi /N is the ring-polymer centroid along the classical reaction coordinate q,

and s0 is located at the top of the barrier. This important special case of QTST corresponds to the centroid-TST method. 20–22 For symmetric or weakly asymmetric barriers, the centroid dividing surface continues to work below Tc , typically down to Tc /2. For asymmetric barriers, however, the dividing surface mixes in other degrees of freedom at temperatures below Tc ; above Tc /2, it can be approximated by 19

s = q 0 cosφ − rsinφ − s0 ,

where r =

(8)

p 2 2 (q+1 + q−1 )/N , and q±1 are degenerate free ring-polymer normal modes with

frequency 2π/β¯ h in the N → ∞ limit. The FEIT rate 27 is obtained by analytically continuing the partition function into the complex plane, and is given by 2 kFEIT Qr (T ) = β¯ h

r

π Qs=0 (T ), 2β|F 00 (0)|

(9)

where Qs=u (T ) is the constrained partition function 1 Qs=u (T ) = lim N →∞ (2π¯ h)N f

Z

Z dp

dx e−βN HN (p,x)

×δ[s(x) − u].

(10)

and F (u) is the ring-polymer free-energy 1 F (u) = − ln[Qs=u (T )]. β

6 ACS Paragon Plus Environment

(11)

Page 7 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

The function s in the delta function is usually chosen to be the same as the ring-polymer dividing surface s of Eq. (5) (i.e. it is chosen to be close to the surface that maximizes the free-energy). Clearly the expressions for kQTST and kFEIT are very similar. In the special case that the mass m associated with the s coordinate is not configuration-dependent, we obtain the simple relation 2π kFEIT = kQTST β¯ h

r

m . |F 00 (0)|

(12)

At temperatures below the cross-over temperature, the steepest-descent limit of kFEIT yields the ‘instanton’ rate. 15–17,19,34 To our knowledge, this correspondence is the nearest thing to a rigorous justification for the use of kFEIT , namely that it provides an improved description of the thermal fluctuations around the instanton. It therefore makes sense only to apply FEIT at temperatures below Tc . Tests on one-dimensional models have found that FEIT gives a consistent improvement on the QTST rate in this regime. Examples of this trend are shown in Fig. 1, for a symmetric Eckart barrier potential, using parameters taken from ref. 19. Figures 2 and 3 show comparisons of kQTST and kFEIT with the exact quantum rates for some collinear and three-dimensional reactions using potential energy surfaces 35–39 taken from the POTLIB website. 40 The free-energy F (u) was calculated using umbrella integration 41 along the reaction coordinate s, with harmonic bias potentials placed at equally spaced windows to yield the mean force f (s), which was then fitted to a spline 42 to give the second derivative F 00 (0) needed to evaluate Eq. (12). For the symmetric collinear reactions, we used s = q 0 (with classical reaction coordinate q taken to be the unstable normal mode of the collinear transition-state); for the asymmetric Cl + H2 reaction, we used the conical reaction coordinate of Eq. (8), obtaining the values of φ (which varied from φ = 0.299 to φ = 0.422 radians from 300 K to 100 K) from instanton calculations at each temperature. For the fully-dimensional H + H2 and D + H2 reactions, we used the dividing surface of ref. 2,43 and

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 17

44, which interpolates between the asymptotic dividing surface

s0 (q) = R∞ − |R|,

(13)

in which |R| is the vector between reactant atom A and reactant diatom BC center-ofmass (and R∞ is chosen large enough to ensure the A-BC interaction is negligible), and the transition-state dividing surface ‡ ‡ s1 (q) = rBC − min[rAB , rAC ] − (rBC − rAB ),

(14)

where rXY is the distance between atoms X and Y and the daggered quantities are the values at the transition-state. Note that this dividing surface is not optimal, as is reflected in the apparent shift in the cross-over temperature in Fig. 3. The collinear calculations used N = 256 (T > 200K) and 512 (T ≤ 200K) polymer beads; the three-dimensional calculations used N = 192 (T ≤ 300K), 128 (T = 350, 400K) and 64 (T ≥ 500K). The collinear Cl + HCl reaction was also tested, but found to give poor results for both kQTST and kFEIT , most probably because of the large amount of dividing-surface recrossing in this heavy-light-heavy reaction. Unlike the one-dimensional results in Fig. 1, the collinear and three-dimensional results in Figs. 2 and 3 do not show consistently better performance of FEIT over QTST at all temperatures below Tc . Instead, there is a range of temperatures below Tc across which QTST is closer to the exact quantum result than FEIT; in the H/D + H2 calculations, this range extends as far down as half the cross-over temperature Tc . Once the temperature is below this range, however, then FEIT does give better results than QTST, where trends first identified in ref. 19 (that QTST underestimates the rate for symmetric barriers and overestimates it for asymmetric barriers) also appear to hold. †



Note that small changes to

Additional calculations would be needed to confirm whether the upward trend in the QTST rate for Cl + H2 continues below 200 K.

8 ACS Paragon Plus Environment

Page 9 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

theses results would be obtained if the dividing surfaces were optimised perfectly, but this is unlikely to change the trends in Figs. 2 and 3. In summary, when applying QTST and FEIT to two- and three-dimensional collinear reactions, we have not been able to find a direct over-the-barrier reaction which replicates the one-dimensional behavior shown in Fig. 1. Instead, we find that QTST gives better agreement than FEIT with the exact quantum rate coefficient over a range of temperature below cross-over. In addition to our findings, many recent studies have shown that RPMD methods can produce results close to accurate quantum rates and experimental results for many gas-phase reactions. 45–48 Further systems need to be explored in the future, but the trend found here suggests that QTST-based methods give more accurate results than FEIT, except at very low temperatures.

Acknowledgments YZ acknowledges funding from the British Council in New Zealand and the University of Cambridge. SCA acknowledges funding from the UK Engineering and Physical Sciences Research Council.

References (1) Shannon, R. J.; Blitz, M. A.; Goddard, A.; Heard, D. E. Accelerated chemistry in the reaction between the hydroxyl radical and methanol at interstellar temperatures facilitated by tunnelling. Nat. Chem. 2013, 5, 745–9. (2) Collepardo-Guevara, R.; Suleimanov, Y. V.; Manolopoulos, D. E. Bimolecular reaction rates from ring polymer molecular dynamics. J. Chem. Phys. 2009, 130, 174713. (3) Nyman, G.; Yu, H.-G. Quantum approaches to polyatomic reaction dynamics. Int. Rev. Phys. Chem. 2013, 32, 39–95.

9 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(4) Li, Y.; Suleimanov, Y. V.; Green, W. H.; Guo, H. Quantum rate coefficients and kinetic isotope effect for the reaction Cl + CH4 → HCl + CH3 from ring polymer molecular dynamics. J. Phys. Chem. A 2014, 118, 1989. (5) Liang, Z.-X.; Klinman, J. P. Structural bases of hydrogen tunneling in enzymes: progress and puzzles. Curr. Opin. Struct. Biol. 2004, 14, 648–55. (6) Masgrau, L.; Ranaghan, K. E.; Scrutton, N. S.; Mulholland, A. J.; Sutcliffe, M. J. Tunneling and classical paths for proton transfer in an enzyme reaction dominated by tunneling: oxidation of tryptamine by aromatic amine dehydrogenase. J. Phys. Chem. B 2007, 111, 3032–47. (7) Rommel, J. B.; Liu, Y.; Werner, H.-J.; Kästner, J. Role of tunneling in the enzyme glutamate mutase. J. Phys. Chem. B 2012, 116, 13682–9. (8) Pollak, E.; Talkner, P. Reaction rate theory: what it was, where is it today, and where is it going? Chaos 2005, 15, 26116. (9) Miller, W. H.; Schwartz, S. D.; Tromp, J. W. Quantum mechanical rate constants for bimolecular reactions. J. Chem. Phys. 1983, 79, 4889. (10) Truhlar, D. G.; Garrett, B. C.; Klippenstein, S. J. Current status of transition-state theory. J. Phys. Chem. 1996, 100, 12771–12800. (11) Hernandez, R.; Miller, W. H. Semiclassical transition state theory. A new perspective. Chem. Phys. Lett. 1993, 214, 129–136. (12) Wang, Y.; Bowman, J. M. One-dimensional tunneling calculations in the imaginaryfrequency, rectilinear saddle-point normal mode. J. Chem. Phys 2008, 129, 121103. (13) Wang, Y.; Bowman, J. M. Mode-specific tunneling using the Qim path: theory and an application to full-dimensional malonaldehyde. J. Chem. Phys. 2013, 139, 154303.

10 ACS Paragon Plus Environment

Page 10 of 17

Page 11 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(14) Miller, W. H.; Zhao, Y.; Ceotto, M.; Yang, S. Quantum instanton approximation for thermal rate constants of chemical reactions. J. Chem. Phys. 2003, 119, 1329. (15) Miller, W. H. Quantum mechanical transition state theory and a new semiclassical model for reaction rate constants. J. Chem. Phys. 1974, 61, 1823. (16) Callan, C. G.; Coleman, S. Fate of the false vacuum. II. First quantum corrections. Phys. Rev. D 1977, 16, 1762. (17) Benderskii, V. A.; Makarov, D. E.; Wight, C. A. One-dimensional models. Adv. Chem. Phys. 1994, 88, 55. (18) Andersson, S.; Nyman, G.; Arnaldsson, A.; Manthe, U.; Jónsson, H. Comparison of Quantum Dynamics and Quantum Transition State Theory Estimates of the H + CH4 Reaction Rate† . J. Phys. Chem. A 2009, 113, 4468. (19) Richardson, J. O.; Althorpe, S. C. Ring-polymer molecular dynamics rate-theory in the deep-tunneling regime: Connection with semiclassical instanton theory. J. Chem. Phys. 2009, 131, 214106. (20) Gillan, M. J. Quantum Simulation of Hydrogen in Metals. Phys. Rev. Lett. 1987, 58, 563–566. (21) Gillan, M. J. Quantum-classical crossover of the transition rate in the damped double well. J. Phys. C 1987, 20, 3621–3641. (22) Voth, G. A.; Chandler, D.; Miller, W. H. Rigorous formulation of quantum transition state theory and its dynamical corrections. J. Chem. Phys. 1989, 91, 7749. (23) Craig, I. R.; Manolopoulos, D. E. Chemical reaction rates from ring polymer molecular dynamics. J. Chem. Phys. 2005, 122, 84106. (24) Craig, I. R.; Manolopoulos, D. E. A refined ring polymer molecular dynamics theory of chemical reaction rates. J. Chem. Phys. 2005, 123, 34102. 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(25) Menzeleev, A. R.; Ananth, N.; Miller III, T. F. Direct simulation of electron transfer using ring polymer molecular dynamics: comparison with semiclassical instanton theory and exact quantum methods. J. Chem. Phys. 2011, 135, 074106. (26) Habershon, S.; Manolopoulos, D. E.; Markland, T. E.; Miller III, T. F. Ring-polymer molecular dynamics: quantum effects in chemical dynamics from classical trajectories in an extended phase space. Annu. Rev. Phys. Chem. 2013, 64, 387–413. (27) Mills, G.; Schenter, G. K.; Makarov, D. E.; Jónsson, H. Generalized path integral based quantum transition state theory. Chem. Phys. Lett. 1997, 278, 91. (28) Chandler, D.; Wolynes, P. G. Exploiting the isomorphism between quantum theory and classical statistical mechanics of polyatomic fluids Exploiting the isomorphism between quantum theory and classical statistical mechanics of polyatomic fluids a ). J. Chem. Phys. 1981, 74, 4078. (29) Parrinello, M.; Rahman, A. Study of an F center in molten KCl. J. Chem. Phys. 1984, 80, 860. (30) Shifman, M., Ed. Instantons in gauge theories; World Scientific: Singapore, 1994. (31) Hele, T. J. H.; Althorpe, S. C. Derivation of a true (t → 0+ ) quantum transition-state theory. I. Uniqueness and equivalence to ring-polymer molecular dynamics transitionstate-theory. J. Chem. Phys. 2013, 138, 084108. (32) Hele, T. J. H.; Althorpe, S. C. On the uniqueness of t → 0+ quantum transition-state theory. J. Chem. Phys. 2013, 139, 084116. (33) Althorpe, S. C.; Hele, T. J. H. Derivation of a true (t → 0+ ) quantum transition-state theory. II. Recovery of the exact quantum rate in the absence of recrossing. J. Chem. Phys. 2013, 139, 084115.

12 ACS Paragon Plus Environment

Page 12 of 17

Page 13 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(34) Althorpe, S. C. On the equivalence of two commonly used forms of semiclassical instanton theory. J. Chem. Phys. 2011, 134, 114104. (35) Siegbahn, P.; Liu, B. An accurate three-dimensional potential energy surface for H3 . J. Chem. Phys. 1978, 68, 2457. (36) Truhlar, D. G.; Horowitz, C. J. Functional representation of Liu and Siegbahn’s accurate ab initio potential energy calculations for H + H2 . J. Chem. Phys. 1978, 68, 2466. (37) Bondi, D. K.; Connor, J. N. L.; Garrett, B. C.; Truhlar, D. G. Test of variational transition state theory with a large-curvature tunneling approximation against accurate quantal reaction probabilities and rate coefficients for three collinear reactions with large reaction-path curvature: Cl+HCl, Cl+DCl, and Cl+MuCl. J. Chem. Phys. 1983, 78, 5981. (38) Schwenke, D. W.; Tucker, S. C.; Steckler, R.; Brown, F. B.; Lynch, G. C.; Truhlar, D. G.; Garrett, B. C. Global potential-energy surfaces for H2 Cl. J. Chem. Phys. 1989, 90, 3110. (39) Lynch, G. C.; Truhlar, D. G.; Brown, F. B.; Zhao, J.-G. A New Potential Energy Surface for H2 Br and Its Use To Calculate Branching Ratios and Kinetic Isotope Effects for the H + HBr Reaction. J. Phys. Chem. 1995, 99, 207. (40) Duchovic, R. J.; Volobuev, Y. L.; Lynch, G. C.; Jasper, A. W.; Truhlar, D. G.; Allison, T. C.; Wagner, A. F.; Garrett, B. C.; Espinosa-Garca, J.; Corchado, J. C. POTLIB. http://comp.chem.umn.edu/potlib. (41) Kästner, J.; Thiel, W. Bridging the gap between thermodynamic integration and umbrella sampling provides a novel analysis method: “Umbrella integration”. J. Chem. Phys. 2005, 123, 144104.

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(42) Stecher, T.; Bernstein, N.; Csányi, G. Free Energy Surface Reconstruction from Umbrella Samples Using Gaussian Process Regression. J. Chem. Theory Comput. 2014, DOI: 10.1021/ct500438v. (43) Yamamoto, T.; Miller, W. H. On the efficient path integral evaluation of thermal rate constants within the quantum instanton approximation. J. Chem. Phys. 2004, 120, 3086–3099. (44) Stecher, T.; Althorpe, S. C. Improved free-energy interpolation scheme for obtaining gas-phase reaction rates from ring-polymer molecular dynamics. Mol. Phys. 2012, 110, 875–883. (45) Pérez de Tudela, R.; Aoiz, F. J.; Suleimanov, Y. V.; Manolopoulos, D. E. Chemical Reaction Rates from Ring Polymer Molecular Dynamics: Zero Point Energy Conservation in Mu + H2 → MuH + H. J. Phys. Chem. Lett. 2012, 3, 493–497. (46) Li, Y.; Suleimanov, Y. V.; Yang, M.; Green, W. H.; Guo, H. Ring Polymer Molecular Dynamics Calculations of Thermal Rate Constants for the O(3 P) + CH4 → OH + CH3 . J. Phys. Chem. Lett. 2013, 4, 48–52. (47) Li, Y.; Suleimanov, Y. V.; Li, J.; Green, W. H.; Guo, H. Rate coefficients and kinetic isotope effects of the X + CH4 → CH3 + HX(X = H, D, Mu) reactions from ring polymer molecular dynamics. J. Chem. Phys. 2013, 138, 094307. (48) Allen, J. W.; Green, W. H.; Li, Y.; ; Guo, H.; Suleimanov, Y. V. Communication: Full dimensional quantum rate coefficients and kinetic isotope effects from ring polymer molecular dynamics for a seven-atom reaction OH + CH4 → CH3 + H2 O. J. Chem. Phys. 2013, 138, 221103.

14 ACS Paragon Plus Environment

Page 14 of 17

Page 15 of 17

1.2

103

1.0 100 10-3 10-6 2

Ratio

k cms−1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

0.8

Quantum QTST FEIT Classial-TST

4

0.6 6

βħω‡

8

10

12

0.4 2

Quantum QTST FEIT

4

6

βħω‡

8

10

12

Figure 1: Comparison of exact quantum, QTST, FEIT and classical TST rate coefficients k for the symmetric Eckart barrier used in ref. 19. The panel on the right plots the ratios of the QTST and FEIT rate coefficients to the exact quantum rate. The blue dashed line indicates the cross-over temperature Tc .

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

104

2.1

100

1.8

Quantum QTST FEIT Classical-TST

10-12

k cms−1

10-16

Ratio

H +H2

10-8

2

4

6 1000/T

8

10

0.6 1.4

10-7

1.2

10-14

H +ClH

10-21

10-35

Quantum QTST FEIT Classical-TST

2

4

6 1000/T

8

10

0.6

H +BrH Quantum QTST FEIT Classical-TST

2

4

6 1000/T

8

10

10-2 10-4

1.3

Quantum QTST FEIT

H +ClH

2

4

6 1000/T

8

10 Quantum QTST FEIT

H +BrH

2

4

6 1000/T

2.7

8

Cl +H2 Quantum QTST FEIT Classical-TST

1.9

10 Quantum QTST FEIT

2.3

2

10

1.1

0.7

Ratio

100

8

0.9

104 102

6 1000/T

1.5

10-15 10-20

4

1.7

Ratio

10-10

1.0

2

0.8

100 10-5

H +H2

1.2

100

10-28

k cms−1

1.5

Quantum QTST FEIT

0.9

Ratio

k cms−1

10-4

k cms−1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 17

Cl +H2

1.5 1.1

3

1000/T

4

5

0.7

2

3

1000/T

4

Figure 2: Same comparison as Fig. 1 for several collinear reactions.

16 ACS Paragon Plus Environment

5

Page 17 of 17

10-12

2.1

10-14

1.8

10-18 10-20 10-22

H +H2

Ratio

k cms−1

10-16

Quantum QTST FEIT Classical-TST

2

10-12

3 1000/T

4

10-18 10-20 10-22

H +H2

1.2

0.6

5

2

3 1000/T

4

2.1

D +H2 Quantum QTST FEIT Classical-TST

2

1.5

5 Quantum QTST FEIT

1.8 Ratio

10-16

1.5

Quantum QTST FEIT

0.9

10-14

k cms−1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

D +H2

1.2

0.9 3 1000/T

4

5

0.6

2

3 1000/T

4

5

Figure 3: Same comparison as Fig. 1 for fully-dimensional H + H2 and D + H2 .

Graphical TOC Entry

17 ACS Paragon Plus Environment