Next Article in Journal
Studies on the Alkaloids of the Bark of Magnolia officinalis: Isolation and On-line Analysis by HPLC-ESI-MSn
Next Article in Special Issue
Understanding the Nucleophilic Character and Stability of the Carbanions and Alkoxides of 1-(9-Anthryl)ethanol and Derivatives
Previous Article in Journal
Analytical Study for the Charge-Transfer Complexes of Rosuvastatin Calcium with π-Acceptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Benchmark Study on the Smallest Bimolecular Nucleophilic Substitution Reaction: H−+CH4 →CH4+H

1
Institució Catalana de Recerca i Estudis Avançats (ICREA), Pg. Lluís Companys 23, 08010 Barcelona, Spain
2
Institut de Química Computacional i Catàlisi (IQCC) and Departament de Química, Universitat de Girona, Campus Montilivi, 17071 Girona, Spain
3
Department of Theoretical Chemistry & Amsterdam Center for Multiscale Modeling, VU University, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands
4
Institute of Molecules and Materials, Radboud University Nijmegen, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands
*
Author to whom correspondence should be addressed.
Molecules 2013, 18(7), 7726-7738; https://doi.org/10.3390/molecules18077726
Submission received: 20 May 2013 / Revised: 3 June 2013 / Accepted: 28 June 2013 / Published: 3 July 2013
(This article belongs to the Special Issue Electrophilic & Nucleophilic Substitution)

Abstract

:
We report here a benchmark study on the bimolecular nucleophilic substitution (SN2) reaction between hydride and methane, for which we have obtained reference energies at the coupled cluster toward full configuration-interaction limit (CC-cf/CBS). Several wavefunction (HF, MP2, coupled cluster) and density functional methods are compared for their reliability regarding these reference data.

1. Introduction

Bimolecular substitution (SN2) reactions play an important role in organic chemistry and in biochemistry (DNA replication mechanism). Interestingly, there is a profound solvent effect present which has a major effect on the reaction barriers and intermediates. For example, the prototypical SN2 reaction of chloride with methyl chloride shows in the gas phase a double-well potential (see Figure 1) with deep wells and a reduced barrier. On the other hand, in solution the energy profile turns basically into a unimodal reaction [1,2,3,4,5,6,7,8,9,10,11,12] (see Figure 1), accompanied by a significant increase of the reaction barrier. In previous studies [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] it was shown that coupled cluster methods in general give accurate results for the energy profile of SN2 reactions, while density functional methods give qualitatively correct results but often underestimate barriers [15]. This has led to the design of new and improved density functionals (SSB-D [43], S12g [44] and S12h [44]), where in particular the latter hybrid functional (S12h) was shown to provide accurate results for the complete energy profile of SN2 reactions.
Figure 1. Energy profile for SN2 reaction in gas phase and in solution.
Figure 1. Energy profile for SN2 reaction in gas phase and in solution.
Molecules 18 07726 g001
Here we have studied the smallest SN2 reaction possible, between hydride and methane:
Molecules 18 07726 i001
For this reaction, we have been able to perform coupled cluster calculations [45] up to the level of CCSDT/aug-cc-pVTZ, and through extrapolation techniques we have obtained reference data at CC-cf/CBS (CC-cf=continued fraction [46] extrapolation toward full-CI limit; CBS = Complete Basis Set). Moreover, we have explored the energy profile for this reaction with 28 density functionals, (LDA, [47,48,49] PBE, [50] PBE-D3, [50,51] PBE0, [52,53,54] PBE0-D3, [51,52,53,54] PW91, [55,56] BP86, [57,58] revPBE, [59] OPBE, [50,60,61] OLYP, [61,62] B3LYP, [63,64] B3LYP-D3, [51,63,64] BLYP, [57,62] B2PLYP, [65] M06, [66] M06-2X, [66] M06-L, [67] B97, [68] B97-3, [69] B97-D2, [51] TPSS, [70] TPSS-D3, [51,70] TPSSh, [71] SSB-D, [43] S12g, [44] S12h, [44] CAM-S12h, [44] CAM-B3LYP [72]), among which the most popular ones from the DFT2012 popularity poll [73] and the newly developed S12g/S12h functional [44].

2. Results and Discussion

The complete energy profile for the SN2 reaction of H+CH4 was studied using both wavefunction and density functional methods. The reaction proceeds from the reactants (R, see Figure 1) towards a reactant complex (RC) and then crosses the central barrier (TS) to reach a product complex (PC) and finally products (P). The RC is reached early on, e.g., with a C-H(nucleophile) distance of some 3.84 Å; this is ca. 0.7 Å longer than the case of Cl + CH3Cl (3.15 Å) even though the size of the nucleophile is probably much smaller here [note however that Pauling [74] reported a larger ionic radius for H (2.08 Å) than for Cl (1.81 Å), while Frecer [75,76] through Monte Carlo obtained values of 2.28 Å (H) and 2.30 Å (Cl) respectively]. This is consistent with a very weak ion-molecule interaction in the reactant complexes. Interestingly enough, at the highest level for which we could obtain the energies directly, CCSDT/atz (see Table 1), this leads to an energy profile of this gas-phase reaction that is more reminiscent of an SN2 profile in solution [1,2,77] (see Figure 1). For example, the RC well is almost non-existent with a depth of 0.9 kcal·mol−1 and the barrier is quite steep with a value of 49.4 kcal·mol−1.

2.1. Coupled Cluster Results

We extrapolated the coupled cluster energies to come as close as possible to the full-CI result, for which we use the continued-fraction approximant [46]. There are two possibilities for this extrapolation (Equations 1a,b), using either the CCSD(T) energy (as we do here in this first part), or the CCSDT energy (as discussed later on), for the δ3 ingredient. The resulting ECC-cf energies are also given in Table 1:
Molecules 18 07726 i002
Molecules 18 07726 i003
The results of Table 1 make it clear that there is a clear basis set effect, where it is not really important to increase the basis set size but it is more important to include diffuse functions [42,78]. Given that we deal here with anionic species, for which diffuse functions are important [21,42,79], this comes as no surprise. There is also a significant electron-correlation effect, where energies obtained at the CCSDT level do not seem to have converged to the full-CI limit. For instance, the well depth increases with the atz basis from −0.27 kcal·mol−1 (RHF) to −0.69 kcal·mol−1 at CCSD, −0.90 kcal·mol−1 at CCSD(T), −0.93 kcal·mol−1 at CCSDT and −1.22 kcal·mol−1 at CC-cf. Likewise, the barrier continues to drop from 62.6 kcal·mol−1 at RHF to 49.4 kcal·mol−1 at CCSDT, and reaches 49.0 kcal·mol−1 at CC-cf. It should be noted that the perturbative triples approach in CCSD(T) gives a good approximation for the CCSDT energies (difference 0.05−0.30 kcal·mol−1). Based on the extrapolation towards full-CI and complete basis set with CC-cf at increasingly larger basis sets (CC-cf/CBS), final reference energies of -1.20 kcal·mol−1 (RC) and +48.90 kcal·mol−1 (TS) are obtained.
Table 1. Relative energies (kcal·mol−1) obtained with wavefunction and density functional methods a.
Table 1. Relative energies (kcal·mol−1) obtained with wavefunction and density functional methods a.
dztzqzadzatzaqzdztzqzadzatzaqz
RCRCRCRCRCRCTSTSTSTSTSTS
RHF−2.55−2.02−1.70−0.27−0.27−0.2752.9056.8658.4761.9862.5962.61
MP2−3.41−3.15−2.93−0.87−0.91−0.8743.8946.4447.2849.3650.1950.21
CCSD−3.53−3.26−2.99−0.70−0.69−0.6242.2345.9547.6151.8752.9753.03
CCSD(T)−3.66−3.46−3.23−0.88−0.90−0.8540.2043.4144.7348.8849.6749.63
CCSDT−3.68−3.49n/ab−0.92−0.93n/a b39.9243.1044.4148.6149.42n/ab
CC−cf c−3.80−3.71−3.53−1.13−1.22−1.1939.8342.8544.0748.3048.9548.87
LDA−10.07−7.78−6.86−3.08−2.77−2.7422.8628.0129.8333.2634.1234.16
PBE−7.72−5.69−5.14−1.99−1.85−1.8327.6832.7634.4938.2639.0039.09
PBE−D3−8.01−5.98−5.40−2.21−2.05−2.0327.4332.5134.2538.0538.8038.88
PBE0−5.56−4.18−3.69−1.18−1.08−1.0735.0439.6141.2144.3845.0345.08
PBE0−D3−5.84−4.43−3.92−1.36−1.25−1.2434.7739.3540.9644.1644.8144.86
PW91−7.975.95−5.49−2.46−2.37−2.3727.3232.4634.1237.7638.5238.59
BP86−6.50−4.39−3.74−0.78−0.65−0.6228.3433.5835.4939.3140.2440.32
revPBE−6.48−4.64−4.23−1.54−1.50−1.5130.1635.2036.8740.6241.2941.38
OPBE−4.72−3.53−3.34n/a dn/a dn/a d34.4938.8740.1843.0143.2243.28
OLYP−6.28−4.77−4.57n/a dn/a dn/a d33.4638.2939.7143.2843.8243.92
B3LYP−5.73−4.02−3.50−0.78−0.73−0.7234.4039.4941.2044.7845.8445.91
B3LYP−D3−6.12−4.35−3.80−0.98−0.91−0.8933.9139.0340.7544.3945.4645.53
BLYP−7.34−5.02−4.44−1.22−1.18−1.1828.6334.3236.1740.3541.5641.66
B2PLYP−4.21−3.00−2.51+0.02+0.03+0.0438.4942.5243.9747.4048.4648.53
M06−5.54−3.99−3.92−1.80−1.84−2.0136.8741.9442.5046.2946.5646.13
M06−2X−5.61−4.43−2.42−1.26−1.10−1.0636.0841.3543.5845.1647.1146.95
M06−L−5.46−4.27−3.76−1.27−1.19−1.1839.8044.0045.6748.7949.3549.45
B97−5.72−4.30−3.84−1.24−1.15−1.1134.3138.9940.7044.1444.9645.11
B97−3−4.78−3.55−3.00−0.63−0.51−0.4437.6841.9943.8247.3248.2648.46
B97−D2−6.57−4.62−4.14−1.33−1.21−1.2028.6933.9335.7440.0541.0341.08
TPSS−5.58−4.10−3.69−1.38−1.33−1.3230.7535.8237.3540.3341.2741.30
TPSS−D3−5.98−4.44−4.00−1.60−1.51−1.5030.3735.4737.0040.0140.9740.99
TPSSh−5.05−3.73−3.33−1.12−1.07−1.0633.3438.2539.7642.6143.5143.52
SSB−D−7.89−6.12−5.54−2.15−2.04−2.0130.6535.0436.8240.5541.0841.20
S12g−8.20−6.36−5.77−2.33−2.18−2.1530.3634.9536.7440.5141.1041.21
S12h−6.18−4.74−4.18−1.47−1.37−1.3536.5140.9942.6645.9746.6046.67
CAM−S12h−5.71−4.38−3.83−1.31−1.22−1.2138.3542.8344.5047.7248.3548.41
CAM−B3LYP−5.03−3.62−3.09−0.64−0.61−0.6038.8543.8845.6149.1150.0350.08
a energies relative to reactants, for each method at their own optimized geometry; b not available due to insufficient computational resources; c obtained with equation 1a at CCSD(T) optimized geometry; d not available due to dissociation towards reactants, i.e., no RC−complex found.
The results obtained with the different levels of coupled cluster method, i.e., CCSD, CCSD(T), CCSDT and CC-cf, together with RHF and MP2 (see Table 1) indicate that CCSD may be sufficient for getting good results. CCSD underestimates the RC well depth (e.g., by 0.3−0.6 kcal·mol−1), and overestimates the barrier by ca. 3−4 kcal·mol−1. MP2 works in this respect even better with deviations from CC-cf that are twice as small, even though the computational effort is more or less the same as CCSD. As already noted often before, RHF cannot be trusted for these energy profiles as it gives barriers which are too large.

2.2. Density Functional Energies

All density functionals show the correct energy profile with a shallow well for the RC (−0.4 to −2.7 kcal·mol−1 with the largest basis set aqz), and a substantial barrier (34.2 to 50.1 kcal·mol−1 with the aqz basis). Nevertheless, the results for the 28 density functionals show quite a diversity in the accuracy for the energy profile, even though some general trends are obvious: they tend to overestimate the RC well depth, and (generally) underestimate the reaction barrier. The least reliable functional is not surprisingly LDA, which for the RC predicts a well depth of −2.74 kcal·mol−1, and places the TS at +34.16 kcal·mol−1 (e.g., a deviation of ca. 15 kcal·mol−1). Early GGA functionals (PBE, BP86 [80], BLYP) improve the barrier by ca. 5–7 kcal·mol−1, and a further 3 kcal·mol−1 is obtained by the use of OPTX in OPBE/OLYP. Surprisingly, SSB−D predicts a barrier that is ca. 2.0 kcal·mol−1 lower than OPBE, even though prior studies showed them to behave similarly. This cannot be due to the inclusion of dispersion in SSB−D, since the effect of including Grimme’s dispersion energy is limited (< 0.5 kcal·mol−1, see Table 1). Even better results are obtained with hybrid functionals and reasonable results are obtained: the difference with the CC-cf results is now only 3−4 kcal·mol−1 (at a fraction of the computational cost) for the most often used hybrid functionals (B3LYP, PBE0, M06). The recently developed S12h and M06−2X bring the deviation from CC−cf down to ca. 2 kcal·mol−1, while excellent results (deviation <0.5 kcal·mol−1) are obtained with four functionals: B2PLYP (48.53 kcal·mol−1), M06−L (49.45 kcal·mol−1), B97−3 (48.46 kcal·mol−1) and CAM−S12h (48.41 kcal·mol−1). Of these four, two give an excellent description of the RC well depth: M06−L (−1.18 kcal·mol−1) and CAM−S12h (−1.21 kcal·mol−1), while the other two underestimate it slightly (B97−3, −0.44 kcal·mol−1) or show a non-existent RC (B2PLYP, +0.04 kcal·mol−1). The non−existence of a RC happens also for OPBE and OLYP with the augmented basis set (adz, atz, aqz), where the optimization proceeds towards reactants.

2.3. Structural Parameters

All methods used here confirm the early character of the RC, with a distance between the nucleophile (Nu) and the central carbon observed within the range of 2.97 Å (B2PLYP) to 4.73 Å (RHF) (see Table 2). These two values are, together with LDA (3.06 Å), rather different from the other wavefunction and density functional methods that give values roughly between 3.4 and 4.0 Å. Moreover, this C−Nu distance is the one that distinguishes the several methods for the deviations with respect to the CCSDT/atz results. The variation for the other structural parameters is much smaller (see Table 2).
Table 2. Structural parameters (Å, °) for stationary points, obtained with atz basis set.
Table 2. Structural parameters (Å, °) for stationary points, obtained with atz basis set.
r(C−H) ar(C−LG) br(C−Nu) cr(C−H) d℘(H−C−LG) er(C−LG) fr(C−H) gMAD h
RHF1.0821.0854.7341.081110.051.6901.0590.898
MP21.0841.0883.7901.084110.641.5781.0670.073
CCSD1.0861.0904.0391.086110.411.6291.0670.201
CCSD(T)1.0881.0923.8571.087110.551.6291.0690.020
CCSDT1.0881.0923.8381.087110.561.6331.0690
LDA1.0971.1043.0551.098111.601.5701.0810.786
PBE1.0961.1013.5251.096110.791.6091.0780.314
PBE−D31.0961.1023.3941.096110.911.6121.0790.444
PBE01.0891.0933.5861.089110.741.6031.0710.253
PBE0−D31.0891.0933.4731.089110.831.6061.0720.366
PW911.0941.0993.5231.094110.761.6141.0760.315
BP861.0961.1003.8621.096110.541.6201.0790.033
revPBE1.0971.1014.0851.097110.361.6241.0790.248
OPBE1.094n/ain/ain/ain/ai1.5711.078n/ai
OLYP1.093n/ain/ain/ain/ai1.6041.075n/ai
B3LYP1.0881.0923.7921.088110.521.6371.0690.046
B3LYP−D31.0891.0933.5651.088110.671.6421.0690.272
BLYP1.0941.0994.1001.094110.321.6521.0750.263
B2PLYP1.0931.1002.9671.093111.501.5591.0790.874
M061.0871.0913.6431.087110.571.6211.0710.195
M06−2X1.0871.0913.3091.087110.931.6121.0690.529
M06−L1.0851.0884.0881.085110.331.6461.0690.251
B971.0911.0953.6841.090110.601.6221.0720.154
B97−31.0871.0913.7891.087110.501.6121.0690.053
B97−D21.0951.1003.8921.095110.461.6721.0770.069
TPSS1.0921.0973.7091.092110.491.6601.0750.238
TPSS−D31.0921.0983.4901.092110.641.6451.0730.348
TPSSh1.0901.0943.7131.089110.511.6361.0710.125
SSB−D1.0871.0933.4591.087110.791.5671.0720.384
S12g1.0931.0983.3841.093110.921.5801.0760.457
S12h1.0871.0923.4001.087110.891.6001.0700.439
CAM−S12h1.0871.0913.4171.086110.871.6061.0690.421
CAM−B3LYP1.0871.0903.7241.086110.611.6351.0670.114
a C−H distance in methane−reactant; b C−LG distance in RC, LG=leaving group; c C−Nu distance in RC, Nu=nucleophile; d C−H distance in RC; e angle H−C−LG in RC; f C−LG (C−Nu) distance at TS; g C−H distance at TS; h mean absolute deviation of distances compared to CCSDT/atz values; i not available due to dissociation towards reactants, i.e., no RC−complex found.

2.4. Single−Point Calculations at CCSDT/atz Geometry

The comparison of the energy profiles for the different methods is of course influenced to some extent by the different geometries used with the different methods. Therefore, and in order to make an honest comparison between the different methods we also performed single-point energy calculations using the CCSDT/atz geometries. Given in Table 3 are the results for all wavefunction and density functional methods.
Table 3. Energy profile (aqz basis, kcal·mol−1) using single−point calculations at CCSDT/atz geometry.
Table 3. Energy profile (aqz basis, kcal·mol−1) using single−point calculations at CCSDT/atz geometry.
MethodRCTS MethodRCTS
RHF0.0463.01 B3LYP−D3−0.8945.54
MP2−0.8650.52 BLYP−1.0941.66
CCSD−0.6053.03 B2PLYP−0.6847.82
CCSD(T)−0.8549.63 M06−2.0146.15
CC-cf−1.1948.87 M06−2X−1.0247.01
M06−L−1.0148.70
LDA−2.3534.62 B97−1.1145.14
PBE−1.8139.17 B97−3−0.4448.51
PBE-D3−1.9738.96 B97−D2−1.1641.18
PBE0−1.0745.18 TPSS−1.3141.29
PBE0-D3−1.2244.95 TPSS−D3−1.4940.99
PW91−2.3538.64 TPSSh−1.0543.52
BP86−0.5840.35 SSB−D−1.9941.65
revPBE−1.3941.39 S12g−2.1041.54
OPBE−1.0443.73 S12h−1.3246.79
OLYP−1.7944.03 CAM−S12h−1.1748.50
B3LYP−0.6945.92 CAM−B3LYP−0.5950.09
By comparing the results from Table 3 with those from Table 1, it can be seen that the influence of the geometry is limited. The largest difference for the different methods observed is of the order of 0.4 kcal·mol−1 (e.g., for LDA, −2.74 kcal·mol−1 for the RC at the LDA geometry, and −2.35 kcal·mol−1 at the CCSDT/atz geometry). However, the typical deviation is less than 0.1 kcal·mol−1 (i.e., chemical accuracy); for instance, PBE−D3 shows differences of 0.06 and 0.08 kcal·mol−1 for the RC and TS energies with the two different geometries. All of this indicates that the energy surface is quite flat, as was already obvious from Figure 1.

2.5. Competition with Proton Transfer Pathway

An alternative reaction is possible in which the hydride abstracts a proton from methane. This leads to the formation of a methyl anion and H2:
Molecules 18 07726 i004
This alternative process is however associated with an endothermic reaction energy of +21.35 kcal·mol−1 at S12h/aqz. This pathway is beyond the scope of the present investigation which focuses on the thermoneutral identity SN2 reaction.

3. Experimental

All wavefunction based calculations (Hartree-Fock, Second-order Møller-Plesset Perturbation Theory, Coupled Cluster Theory) have been performed with the Coupled-Cluster techniques for Computational Chemistry (CFOUR) [81,82] program version 1.2, using a variety of Dunning’s correlation-consistent basis sets [83,84]: cc−pVXZ (X=D, T, Q, abbreviated as dz, tz and qz respectively) and aug−cc−pVXZ (X=D, T, Q, abbreviated as adz, atz and aqz respectively). The continued-fraction approximant [46] for obtaining coupled−cluster energies toward the full configuration-interaction limit (CC−cf) has been used with equations 1a,b to reach completeness of electron-correlation energies. The NWChem program [85] version 6.1 was used for all density functional calculations.

4. Conclusions

We have performed a benchmark study on the smallest bimolecular nucleophilic substitution (SN2) reaction possible: H+CH4→ CH4+H, for which we obtained reference data at the near-full-CI coupled cluster limit using the continued−fraction approximant (CC-cf). Unlike typical SN2 reactions in the gas phase, which usually show a double-well potential, the current reaction shows an energy profile that resembles more the unimodal profile of the SN2 reaction in solution, with a relatively shallow reactant-complex well of only −1.19 kcal·mol−1 and a high barrier amounting to 48.87 kcal·mol−1. All other computational methods also clearly show the steep reaction barrier that needs to be overcome (34−50 kcal·mol−1), and the very shallow wells of the (ion-dipole) reactant complex. All density functionals have the tendency to underestimate the reaction barrier, while for the RC the deviations compared to CC-cf are much smaller (< 0.5 kcal·mol−1). Excellent results have been obtained with B97-3, M06-L and the newly developed CAM−S12h functional.

Acknowledgments

The following organizations are thanked for financial support: the Ministerio de Ciencia e Innovación (MICINN, project number CTQ2011-25086/BQU), the DIUE of the Generalitat de Catalunya (project number 2009SGR528, and Xarxa de Referència en Química Teòrica i Computacional), the National Research School Combination—Catalysis (NRSC-C), and the Netherlands Organization for Scientific Research (NWO−CW). Financial support from MICINN (Ministry of Science and Innovation, Spain) and the FEDER fund (European Fund for Regional Development) was provided by grant UNGI08-4E-003. Excellent service by the Centre de Serveis Científics i Acadèmics de Catalunya (CESCA) is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. van Bochove, M.A.; Bickelhaupt, F.M. Nucleophilic substitution at C, Si and P: How solvation affects the shape of reaction profiles. Eur. J. Org. Chem. 2008, 2008, 649–654. [Google Scholar] [CrossRef]
  2. Bickelhaupt, F.M. Understanding reactivity with Kohn−Sham molecular orbital theory: E2−SN2 mechanistic spectrum and other concepts. J. Comput. Chem. 1999, 20, 114–128. [Google Scholar] [CrossRef]
  3. Olmstead, W.N.; Brauman, J.I. Gas-phase nucleophilic displacement reactions. J. Am. Chem. Soc. 1977, 99, 4219–4228. [Google Scholar] [CrossRef]
  4. Chabinyc, M.L.; Craig, S.L.; Regan, C.K.; Brauman, J.I. Gas−phase ionic reactions: Dynamics and mechanism of nucleophilic displacements. Science 1998, 279, 1882–1886. [Google Scholar] [CrossRef]
  5. Chandrasekhar, J.; Smith, S.F.; Jorgensen, W.L. Theoretical examination of the SN2 reaction involving chloride ion and methyl chloride in the gas phase and aqueous solution. J. Am. Chem. Soc. 1985, 107, 154–163. [Google Scholar] [CrossRef]
  6. Vayner, G.; Houk, K.N.; Jorgensen, W.L.; Brauman, J.I. Steric retardation of SN2 reactions in the gas phase and solution. J. Am. Chem. Soc. 2004, 126, 9054–9058. [Google Scholar] [CrossRef]
  7. Chandrasekhar, J.; Jorgensen, W.L. Energy profile for a nonconcerted SN2 reaction in solution. J. Am. Chem. Soc. 1985, 107, 2974–2975. [Google Scholar] [CrossRef]
  8. Chandrasekhar, J.; Smith, S.F.; Jorgensen, W.L. SN2 reaction profiles in the gas phase and aqueous solution. J. Am. Chem. Soc. 1984, 106, 3049–3050. [Google Scholar] [CrossRef]
  9. Jorgensen, W.L.; Buckner, J.K. Effect of hydration on the structure of an SN2 transition state. J. Phys. Chem. 1986, 90, 4651–4654. [Google Scholar] [CrossRef]
  10. Chen, X.; Regan, C.K.; Craig, S.L.; Krenske, E.H.; Houk, K.N.; Jorgensen, W.L.; Brauman, J.I. Steric and solvation effects in ionic SN2 reactions. J. Am. Chem. Soc. 2009, 131, 16162–16170. [Google Scholar]
  11. Kim, Y.; Cramer, C.J.; Truhlar, D.G. Steric effects and solvent effects on SN2 reactions. J. Phys. Chem. A 2009, 113, 9109–9114. [Google Scholar] [CrossRef]
  12. Craig, S.L.; Brauman, J.I. Intramolecular microsolvation of SN2 transition states. J. Am. Chem. Soc. 1999, 121, 6690–6699. [Google Scholar] [CrossRef]
  13. Manikandan, P.; Zhang, J.; Hase, W.L. Chemical dynamics simulations of X + CH3Y → XCH3 + Y gas−phase SN2 nucleophilic substitution reactions. Nonstatistical dynamics and nontraditional reaction mechanisms. J. Phys. Chem. A 2012, 116, 3061–3080. [Google Scholar] [CrossRef]
  14. Kowalski, K.; Valiev, M. Extensive regularization of the coupled cluster methods based on the generating functional formalism: Application to gas-phase benchmarks and to the SN2 reaction of CHCl3 and OH in water. J. Chem. Phys. 2009, 131, 234107. [Google Scholar] [CrossRef]
  15. Swart, M.; Solà, M.; Bickelhaupt, F.M. Energy landscapes of nucleophilic substitution reactions: A comparison of density functional theory and coupled cluster methods. J. Comput. Chem. 2007, 28, 1551–1560. [Google Scholar] [CrossRef]
  16. Galabov, B.; Nikolova, V.; Wilke, J.J.; Schaefer, H.F., III; Allen, W.D. Origin of the SN2 benzylic effect. J. Am. Chem. Soc. 2008, 130, 9887–9896. [Google Scholar] [CrossRef]
  17. Botschwina, P. The saddle point of the nucleophilic substitution reaction Cl + CH3Cl: Results of large-scale coupled cluster calculations. Theor. Chem. Acc. 1998, 99, 426–428. [Google Scholar]
  18. Botschwina, P.; Horn, M.; Seeger, S.; Oswald, R. Stationary points of the potential surface for the reaction F+CH3Cl → FCH3 + Cl: Results of large-scale coupled cluster calculations. Ber. Bunsenges. Phys. Chem. 1997, 101, 387–390. [Google Scholar] [CrossRef]
  19. Gonzales, J.M.; Cox III, R.S.; Brown, S.T.; Allen, W.D.; Schaefer III, H.F. Assessment of density functional theory for model SN2 reactions: CH3X + F (X = F, Cl, CN, OH, SH, NH, PH). J. Phys. Chem. A 2001, 105, 11327–11346. [Google Scholar] [CrossRef]
  20. Gonzales, J.M.; Pak, C.; Cox, R.S.; Allen, W.D.; Schaefer III, H.F.; Császár, A.G.; Tarczay, G. Definitive ab initio studies of model SN2 reactions CH3X + F (X=F, Cl, CN, OH, SH, NH2, PH2). Chem.-Eur. J. 2003, 9, 2173–2192. [Google Scholar] [CrossRef]
  21. Bento, A.P.; Solà, M.; Bickelhaupt, F.M. Ab initio and DFT benchmark study for nucleophilic substitution at carbon (SN2@C) and silicon (SN2@Si). J. Comput. Chem. 2005, 26, 1497–1504. [Google Scholar] [CrossRef]
  22. Bento, A.P.; Solà, M.; Bickelhaupt, F.M. E2 and SN2 reactions of X + CH3CH2X (X = F, Cl); an ab initio and dft benchmark study. J. Chem. Theory Comp. 2008, 4, 929–940. [Google Scholar] [CrossRef]
  23. Bachrach, S.M.; Mulhearn, D.C. Nucleophilic substitution at sulfur: SN2 or addition−elimination? J. Phys. Chem. 1996, 100, 3535–3540. [Google Scholar] [CrossRef]
  24. Angel, L.A.; Ervin, K.M. Dynamics of the gas-phase reactions of fluoride ions with chloromethane. J. Phys. Chem. A 2001, 105, 4042–4051. [Google Scholar] [CrossRef]
  25. Angel, L.A.; Ervin, K.M. Gas-phase reactions of the iodide ion with chloromethane and bromomethane: Competition between nucleophilic displacement and halogen abstraction. J. Phys. Chem. A 2004, 108, 9827–9833. [Google Scholar] [CrossRef]
  26. Angel, L.A.; Garcia, S.P.; Ervin, K.M. Dynamics of the gas−phase reactions of chloride ion with fluoromethane: High excess translational activation energy for an endothermic SN2 reaction. J. Am. Chem. Soc. 2002, 124, 336–345. [Google Scholar] [CrossRef]
  27. Borisov, Y.A.; Arcia, E.E.; Mielke, S.L.; Garrett, B.C.; Dunning, T.H., Jr. A systematic study of the reactions of OH with chlorinated methanes. 1. Benchmark studies of the gas-phase reactions. J. Phys. Chem. A 2001, 105, 7724–7736. [Google Scholar] [CrossRef]
  28. Glukhovtsev, M.N.; Pross, A.; Radom, L. Gas-phase identity SN2 reactions of halide anions with methyl halides: A high-level computational study. J. Am. Chem. Soc. 1995, 117, 2024–2032. [Google Scholar] [CrossRef]
  29. Glukhovtsev, M.N.; Pross, A.; Radom, L. Gas-phase identity SN2 reactions of halide ions at neutral nitrogen: A high-level computational study. J. Am. Chem. Soc. 1995, 117, 9012–9018. [Google Scholar] [CrossRef]
  30. Uggerud, E. Nature of the transition state in gas phase SN2 identity reactions: Correlation between nucleophilicity and proton affinity. J. Chem. Soc. Perkin Trans. 1999, 2, 1459–1463. [Google Scholar]
  31. Wang, H.; Hase, W.L. Kinetics of F + CH3Cl SN2 nucleophilic substitution. J. Am. Chem. Soc. 1997, 119, 3093–3102. [Google Scholar] [CrossRef]
  32. Wladkowski, B.D.; Allen, W.D.; Brauman, J.I. The SN2 identity exchange reaction F + CH3F → FCH3 + F: Definitive ab initio predictions. J. Phys. Chem. 1994, 98, 13532–13540. [Google Scholar] [CrossRef]
  33. Xiong, Y.; Zhu, H.-J.; Ren, Y. A theoretical study of the gas-phase ion pair SN2 reactions of lithium halide and methyl halide with inversion and retention mechanisms. J. Mol. Struct. (THEOCHEM) 2003, 664𢀓665, 279–289. [Google Scholar]
  34. Yu, Z.-X.; Wu, Y.-D. An SN2−like transition state for alkene episulfidation by dinitrogen sulfide. J. Org. Chem. 2003, 68, 6049–6052. [Google Scholar] [CrossRef]
  35. Sun, L.; Song, K.; Hase, W.L.; Sena, M.; Riveros, J.M. Stationary points for the OH + CH3F → CH3OH + F potential energy surface. Int. J. Mass Spectrom. 2003, 227, 315–325. [Google Scholar] [CrossRef]
  36. Schmatz, S.; Botschwina, P.; Stoll, H. Coupled cluster calculations for the SN2 reaction Cl + CH3Br → ClCH3 + Br. Int. J. Mass Spectrom. 2000, 201, 277–282. [Google Scholar] [CrossRef]
  37. Schmatz, S.; Botschwina, P.; Hauschildt, J.; Schinke, R. Symmetry specificity in the unimolecular decay of the Cl···CH3Cl complex: Two-mode quantum calculations on a coupled-cluster [CCSD(T)] potential energy surface. J. Chem. Phys. 2001, 114, 5233–5245. [Google Scholar] [CrossRef]
  38. Parthiban, S.; de Oliveira, G.; Martin, J.M.L. Benchmark ab initio energy profiles for the gas−phase SN2 reactions Y + CH3X → CH3Y + X(X,Y = F,Cl,Br). Validation of hybrid dft methods. J. Phys. Chem. A 2001, 105, 895–904. [Google Scholar] [CrossRef]
  39. Matsubara, H.; Horvat, S.M.; Schiesser, C.H. Methyl radical also reacts by the frontside mechanism: An ab initio study of some homolytic substitution reactions of methyl radical at silicon, germanium and tin. Org. Biomol. Chem. 2003, 1, 1199–1203. [Google Scholar] [CrossRef]
  40. Lee, I.; Kim, C.K.; Sohn, C.K.; Li, H.G.; Lee, H.W. A high-level theoretical study on the gas-phase identity methyl transfer reactions. J. Phys. Chem. A 2002, 106, 1081–1087. [Google Scholar] [CrossRef]
  41. Zhao, Y.; González−García, N.; Truhlar, D.G. Benchmark database of barrier heights for heavy atom transfer, Nucleophilic substitution, Association, And unimolecular reactions and its use to test theoretical methods. J. Phys. Chem. A 2005, 109, 2012–2018. [Google Scholar] [CrossRef]
  42. Zhao, Y.; Truhlar, D.G. Density functional calculations of E2 and SN2 reactions: Implementation effects on calculated interaction energies. J. Chem. Theory Comp. 2010, 6, 1104–1108. [Google Scholar] [CrossRef]
  43. Swart, M.; Solà, M.; Bickelhaupt, F.M. A new all-round DFT functional based on spin states and SN2 barriers. J. Chem. Phys. 2009, 131, 094103. [Google Scholar] [CrossRef]
  44. Swart, M. A new family of hybrid functionals. Chem. Phys. Lett. 2013. [Google Scholar] [CrossRef]
  45. Helgaker, T.; Ruden, T.A.; Jorgensen, P.; Olsen, J.; Klopper, W. A priori calculation of molecular properties to chemical accuracy. J. Phys. Org. Chem. 2004, 17, 913–933. [Google Scholar] [CrossRef]
  46. Goodson, D.Z. Extrapolating the coupled−cluster sequence toward the full configuration−interaction limit. J. Chem. Phys. 2002, 116, 6948–6956. [Google Scholar] [CrossRef]
  47. Dirac, P.A.M. Quantum mechanics of many−electron systems. Proc. R. Soc. London, Ser. A 1929, 123, 714–733. [Google Scholar] [CrossRef]
  48. Slater, J.C. A simplification of the Hartree-Fock method. Phys. Rev. 1951, 81, 385–390. [Google Scholar] [CrossRef]
  49. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin−dependent electron liquid correlation energies for local spin−density calculations —A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  50. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximations made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  51. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  52. Perdew, J.P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  53. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6169. [Google Scholar] [CrossRef]
  54. Ernzerhof, M.; Scuseria, G.E. Assessment of the Perdew-Burke-Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. [Google Scholar] [CrossRef]
  55. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces—Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar]
  56. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation-energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  57. Becke, A.D. Density−functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  58. Perdew, J.P. Density−functional approximation for the correlation-energy of the inhomogeneous electron-gas. Phys. Rev. B 1986, 33, 8822–8824. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Yang, W. Comment on “Generalized gradient approximation made simple”. Phys. Rev. Lett. 1998, 80, 890. [Google Scholar] [CrossRef]
  60. Swart, M.; Ehlers, A.W.; Lammertsma, K. Performance of the OPBE exchange-correlation functional. Molec. Phys. 2004, 102, 2467–2474. [Google Scholar]
  61. Handy, N.C.; Cohen, A.J. Left−right correlation energy. Molec. Phys. 2001, 99, 403–412. [Google Scholar] [CrossRef]
  62. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle−Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  63. Becke, A.D. Density-functional Thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  64. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields. J. Phys. Chem. 1994, 45, 11623–11627. [Google Scholar]
  65. Grimme, S. Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys. 2006, 124, 034108. [Google Scholar] [CrossRef]
  66. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  67. Zhao, Y.; Truhlar, D.G. A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101. [Google Scholar] [CrossRef]
  68. Becke, A.D. Density-functional thermochemistry. 5. Systematic optimization of exchange-correlation functionals. J. Chem. Phys. 1997, 107, 8554–8560. [Google Scholar] [CrossRef]
  69. Keal, T.W.; Tozer, D.J. Semiempirical hybrid functional with improved performance in an extensive chemical assessment. J. Chem. Phys. 2005, 123, 121103. [Google Scholar] [CrossRef]
  70. Tao, J.M.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the density functional ladder: Nonempirical meta− generalized gradient approximation designed for molecules and solids. Phys. Rev. Lett. 2003, 91, 146401. [Google Scholar] [CrossRef]
  71. Staroverov, V.N.; Scuseria, G.E.; Tao, J.; Perdew, J.P. Comparative assessment of a new nonempirical density functional: Molecules and hydrogen−bonded complexes. J. Chem. Phys. 2003, 119, 12129–12137. [Google Scholar] [CrossRef]
  72. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM−B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef]
  73. Swart, M.; Bickelhaupt, F.M.; Duran, M. Popularity poll density functionals 2012 (DFT2012). Available online: http://www.marcelswart.eu/dft−poll/news2012.pdf (accessed on 4 June 2013).
  74. Pauling, L. The Nature of the Chemical Bond, 3rd ed.; Cornell University Press: Ithaca, NY, USA, 1960; pp. 505–563. [Google Scholar]
  75. Frecer, V.; Miertus, S. Polarizable continuum model of solvation for biopolymers. Int. J. Quantum Chem. 1992, 42, 1449–1468. [Google Scholar] [CrossRef]
  76. Swart, M.; van Duijnen, P.T. DRF90: a polarizable force field. Molec. Simul. 2006, 32, 471–484. [Google Scholar] [CrossRef]
  77. Bickelhaupt, F.M.; Baerends, E.J.; Nibbering, N.M.M. The effect of microsolvation on E2 and SN2 Reactions: Theoretical study of the model system F + C2H5F + nHF. Chem. Eur. J. 1996, 2, 196–207. [Google Scholar] [CrossRef]
  78. Wheeler, S.E.; Houk, K.N. Integration grid errors for meta-gga-predicted reaction energies: Origin of grid errors for the M06 suite of functionals. J. Chem. Theory Comp. 2010, 6, 395–404. [Google Scholar] [CrossRef]
  79. Swart, M.; Solà, M.; Bickelhaupt, F.M. Density functional calculations of E2 and SN2 reactions: Effects of the choice of method, Algorithm, And numerical accuracy. J. Chem. Theory Comp. 2010, 6, 3145–3152. [Google Scholar] [CrossRef]
  80. Pierrefixe, S.C.A.H.; Bickelhaupt, F.M. Hypervalence and the delocalizing versus localizing propensities of H3, Li3, CH5 and SiH5. Struct. Chem. 2007, 18, 813–819. [Google Scholar] [CrossRef]
  81. Harding, M.E.; Metzroth, T.; Gauss, J. Parallel calculation of CCSD and CCSD(T) analytic first and second derivatives. J. Chem. Theory Comp. 2008, 4, 64–74. [Google Scholar] [CrossRef]
  82. Stanton, J.F.; Gauss, J.; Harding, M.E.; Szalay, P.G.; with contributions from A.A. Auer, R.J.B., U. Benedikt, C. Berger, D.E. Bernholdt, Y.J. Bomble, L. Cheng, O. Christiansen, M. Heckert, O. Heun, C. Huber, T.-C. Jagau, D. Jonsson, J. Jusélius, K. Klein, W.J. Lauderdale, D.A. Matthews, T. Metzroth, D.P. O’Neill, D.R. Price, E. Prochnow, K. Ruud, F. Schiffmann, W. Schwalbach, S. Stopkowicz, A. Tajti, J. Vázquez, F. Wang, J.D. Watts; including the integral packages MOLECULE (J. Almlöf and P.R. Taylor), PROPS (P.R. Taylor), ABACUS (T. Helgaker, H.J. Aa. Jensen, P. Jørgensen, and J. Olsen), and ECP routines by A.V. Mitin and C. van Wüllen CFOUR 1.2, see http://www.cfour.de, CFOUR 1.2; Austin, TX, USA; Mainz, Germany, 2010.
  83. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  84. Dunning, T.H., Jr.; Peterson, K.A.; Wilson, A.K. Gaussian basis sets for use in correlated molecular calculations. X. The atoms aluminum through argon revisited. J. Chem. Phys. 2001, 114, 9244–9253. [Google Scholar] [CrossRef]
  85. Valiev, M.; Bylaska, E.J.; Govind, N.; Kowalski, K.; Straatsma, T.P.; Van Dam, H.J.J.; Wang, D.; Nieplocha, J.; Apra, E.; Windus, T.L.; et al. NWChem: A comprehensive and scalable open-source solution for large scale molecular simulations. Comp. Phys. Comm. 2010, 181, 1477–1489. [Google Scholar] [CrossRef]
  • Sample Availability: Cartesian coordinates of all stationary points are available from the authors.

Share and Cite

MDPI and ACS Style

Swart, M.; Bickelhaupt, F.M. Benchmark Study on the Smallest Bimolecular Nucleophilic Substitution Reaction: H−+CH4 →CH4+H. Molecules 2013, 18, 7726-7738. https://doi.org/10.3390/molecules18077726

AMA Style

Swart M, Bickelhaupt FM. Benchmark Study on the Smallest Bimolecular Nucleophilic Substitution Reaction: H−+CH4 →CH4+H. Molecules. 2013; 18(7):7726-7738. https://doi.org/10.3390/molecules18077726

Chicago/Turabian Style

Swart, Marcel, and F. Matthias Bickelhaupt. 2013. "Benchmark Study on the Smallest Bimolecular Nucleophilic Substitution Reaction: H−+CH4 →CH4+H" Molecules 18, no. 7: 7726-7738. https://doi.org/10.3390/molecules18077726

Article Metrics

Back to TopTop