30.11.2012 Views

First-principles computational studies of the torsional potential ...

First-principles computational studies of the torsional potential ...

First-principles computational studies of the torsional potential ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

<strong>First</strong>-<strong>principles</strong> <strong>computational</strong> <strong>studies</strong> <strong>of</strong> <strong>the</strong><br />

<strong>torsional</strong> <strong>potential</strong> energy surface <strong>of</strong> <strong>the</strong> sec-butyl<br />

radical<br />

Ya Kun Chen and Yan Alexander Wang<br />

Abstract: <strong>First</strong>-<strong>principles</strong> calculations were carried out to investigate <strong>the</strong> <strong>torsional</strong> <strong>potential</strong> energy surface (PES) <strong>of</strong> <strong>the</strong> secbutyl<br />

radical. All <strong>the</strong> wave function methods employed predict a cis-like stable conformation with a dihedral angle <strong>of</strong> about<br />

47° in addition to <strong>the</strong> trans-like global minimum conformation and a gauche conformation. However, most <strong>of</strong> <strong>the</strong> popular<br />

density functional approaches predict only <strong>the</strong> latter two local minima and lack <strong>the</strong> cis conformation that was experimentally<br />

observed. On <strong>the</strong> o<strong>the</strong>r hand, some density functional methods that incorporate <strong>the</strong> exact exchange and asymptotically corrected<br />

correlation functionals can locate <strong>the</strong> cis conformation successfully. The basis-set effect was also measured using popular<br />

B3LYP and MP2 Hamiltonians: only moderate shape changes were found for PES pr<strong>of</strong>iles upon basis-set variations.<br />

The stationary structures and <strong>the</strong>ir Hessians were obtained at both MP2 and B3LYP levels, with or without incorporating<br />

<strong>the</strong> zero-point energies. Opposite to <strong>the</strong> relative stability within <strong>the</strong> Born–Oppenheimer approximation, <strong>the</strong> cis conformation<br />

is more stable than <strong>the</strong> gauche conformation upon <strong>the</strong> zero-point correction, consistent with <strong>the</strong> experiment observations.<br />

Key words: sec-butyl, <strong>torsional</strong> <strong>potential</strong> energy surface, density functional <strong>the</strong>ory, wave function <strong>the</strong>ory.<br />

Résumé : On a effectué des calculs théoriques à partir des principes premiers pour étudier la surface d’énergie potentielle<br />

(SEP) de torsion du radical sec-butyle. Toutes les méthodes de fonction d’onde utilisées prédisent qu’il existe sous une<br />

conformation de type cis stable avec un angle dièdre de 47° à côté d’une conformation de type trans avec minimum global<br />

et d’une conformation gauche. Toutefois, la plupart des approches populaires par la fonctionnelle de la densité ne prédisent<br />

l’existence que des deux derniers minima locaux sans prédire celle de la conformation cis observée expérimentalement. Par<br />

ailleurs, certaines méthodes de la fonctionnelle de la densité qui incorporent l’échange exact et des fonctionnelles de corrélation<br />

corrigées d’une façon asymptotique permettent de localiser avec succès la conformation cis. On a aussi mesuré l’effet<br />

de l’ensemble de base en faisant appel aux hamiltoniens populaires B3LYP et MP2 : toutefois, avec ces changements d’ensembles<br />

de base, on n’a pu observer que des changements modérés de la forme des pr<strong>of</strong>ils de la surface d’énergie potentielle<br />

(SEP). On a obtenu les structures stationnaires et leurs matrices hessiennes aux niveaux MP2 ainsi que B3LYP, avec<br />

et sans incorporation des énergies du point zéro. Contrairement à la stabilité relative observée dans l’approximation de<br />

Born–Oppenheimer, la conformation cis est plus stable que la conformation gauche lorsqu’on introduit une correction pour<br />

l’énergie au point zéro et ceci est en accord avec les observations expérimentales.<br />

Mots‐clés : sec-butyle, surface d’énergie potentielle de torsion, théorie de la fonctionnelle de la densité, théorie de la fonction<br />

d’onde.<br />

[Traduit par la Rédaction]<br />

Introduction<br />

Torsional <strong>potential</strong> energy surfaces (PESs) <strong>of</strong> many chemical<br />

species have been studied numerous times because <strong>the</strong>y<br />

provide <strong>the</strong> key information for molecular <strong>the</strong>rmodynamic<br />

state functions. At <strong>the</strong> same time, it is also <strong>of</strong> great interest<br />

to compare <strong>the</strong> <strong>the</strong>oretically predicted PES and available experimental<br />

data to develop and improve <strong>computational</strong> methods<br />

and to provide deep insight into <strong>the</strong> nature <strong>of</strong> <strong>torsional</strong><br />

barriers.<br />

Even molecules as simple as ethane have attracted much<br />

<strong>the</strong>oretical attention for <strong>the</strong> investigation <strong>of</strong> <strong>the</strong> physical<br />

nature <strong>of</strong> its <strong>torsional</strong> barrier. The rotational barriers <strong>of</strong><br />

ethane and its congeners were mainly attributed to hyperconjugation<br />

interactions by some researchers, challenging <strong>the</strong><br />

popular belief in many textbooks <strong>of</strong> <strong>the</strong> importance <strong>of</strong> steric<br />

repulsive interactions. 1 However, using generalized valence<br />

bond <strong>the</strong>ory, a recent <strong>the</strong>oretical study, in which <strong>the</strong> <strong>torsional</strong><br />

barrier <strong>of</strong> ethane was decomposed adiabatically into steric repulsion<br />

and hyperconjugation as well as several o<strong>the</strong>r insignificant<br />

contributions, manifested that <strong>the</strong> steric repulsion is<br />

<strong>the</strong> main factor to <strong>the</strong> stabilization <strong>of</strong> staggered conformation<br />

and hyperconjugation is only <strong>of</strong> a secondary effect. 2<br />

Received 10 February 2011. Accepted 23 June 2011. Published at www.nrcresearchpress.com/cjc on 9 November 2011.<br />

Y.K. Chen and Y.A. Wang. Department <strong>of</strong> Chemistry, University <strong>of</strong> British Columbia, 2036 Main Mall, Vancouver, BC V6T 1Z1,<br />

Canada.<br />

Corresponding author: Yan Alexander Wang (e-mail: yawang@chem.ubc.ca).<br />

This article is part <strong>of</strong> a Special Issue to commemorate <strong>the</strong> International Year <strong>of</strong> Chemistry.<br />

Can. J. Chem. 89: 1469–1476 (2011) doi:10.1139/V11-109 Published by NRC Research Press<br />

1469


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

1470 Can. J. Chem. Vol. 89, 2011<br />

Without <strong>the</strong> ambiguity in <strong>the</strong> physical origin <strong>of</strong> <strong>torsional</strong><br />

barriers, it is widely accepted that conjugation effect dominates<br />

<strong>the</strong> <strong>torsional</strong> barrier <strong>of</strong> <strong>the</strong> conjugated systems. However,<br />

this well-adopted understanding does not promise good<br />

agreement between experimental results and <strong>the</strong>oretical predictions<br />

in many conformational <strong>studies</strong>. The <strong>torsional</strong> energy<br />

pr<strong>of</strong>ile <strong>of</strong> 2,2′-bifuran was studied using various forms <strong>of</strong><br />

Hamiltonians, in which density functional <strong>the</strong>ory (DFT)<br />

methods were found inappropriate for <strong>the</strong> conformation study<br />

<strong>of</strong> this system. 3 Not only does <strong>the</strong> form <strong>of</strong> <strong>the</strong> Hamiltonian<br />

matter, but <strong>the</strong> basis set employed also plays a significant<br />

role. When biphenyl PES was studied using several wave<br />

function <strong>the</strong>ory (WFT) methods, <strong>the</strong> quality <strong>of</strong> basis sets<br />

was found to be more important than that <strong>of</strong> <strong>the</strong> Hamiltonians.<br />

4,5 Besides <strong>the</strong> individual Hamiltonian or <strong>the</strong> basis-set<br />

effect, delicate selection <strong>of</strong> certain combinations <strong>of</strong> Hamiltonians<br />

and basis sets can be critical for an accurate calculation<br />

<strong>of</strong> <strong>the</strong>se (semi)rigid molecules. Theoretical study <strong>of</strong> <strong>the</strong> <strong>torsional</strong><br />

barriers <strong>of</strong> biisothianaph<strong>the</strong>ne showed that an accurate<br />

and balanced description <strong>of</strong> both conjugation and steric repulsion<br />

between nonvicinal hydrogens could be only<br />

achieved using high-level Hamiltonians and large basis sets. 6<br />

Therefore, one must be aware <strong>of</strong> <strong>the</strong> limitations <strong>of</strong> popular<br />

<strong>the</strong>oretical tools; extensive <strong>studies</strong> are required to avoid false<br />

predictions even for small molecules based on calculations<br />

using a single Hamiltonian with a single basis set.<br />

O<strong>the</strong>r than <strong>the</strong>se well-documented molecules, it might become<br />

even more challenging to study <strong>the</strong> conformations <strong>of</strong><br />

some open-shell and complex species, including radicals<br />

whose PESs are usually so flat that an accurate description<br />

<strong>of</strong> <strong>the</strong> electronic energies involved goes beyond <strong>the</strong> reach <strong>of</strong><br />

popular <strong>computational</strong> methods. For example, in <strong>the</strong> case <strong>of</strong><br />

haloethyl radicals, a <strong>the</strong>oretical study showed that <strong>the</strong> DFT<br />

methods with <strong>the</strong> generalized gradient approximation (GGA)<br />

outperformed <strong>the</strong> local Møller-Plesset (MP2) method in predicting<br />

<strong>the</strong> dissociation energy <strong>of</strong> haloethyl radicals. 7 On <strong>the</strong><br />

o<strong>the</strong>r hand, large spin contamination was found from DFT<br />

calculations <strong>of</strong> OOBr and OOCl, contradicting <strong>the</strong> popular<br />

view that DFT methods suffer less spin contamination than<br />

WFT methods. Sometimes, poor results can come from<br />

many different methods in a single case. In a conformational<br />

study <strong>of</strong> annulene molecules, both MP2 and DFT<br />

methods were found to fail one way or ano<strong>the</strong>r. 8 All <strong>the</strong> uncertainties<br />

and controversies mentioned above suggest that<br />

some popular <strong>the</strong>oretical methods are still unable to work<br />

equally well for conformation <strong>studies</strong> <strong>of</strong> various molecular<br />

systems. Therefore, a convincing conclusion for <strong>the</strong> conformation<br />

<strong>studies</strong> <strong>of</strong>ten requires systematical benchmarking<br />

comparisons.<br />

In this study, <strong>the</strong> <strong>torsional</strong> PES <strong>of</strong> sec-butyl was <strong>the</strong>oretically<br />

investigated. The available experimental data were<br />

mainly from hyperfine coupling constant (HFCC) study <strong>of</strong><br />

its muoniated sec-butyl formed from muonium addition to 2butene<br />

isomers. 9,10 As a short-lived, light isotope <strong>of</strong> hydrogen,<br />

<strong>the</strong> muonium, which is composed <strong>of</strong> a positive muon<br />

(antimuon) and an electron, is a good probing atom for <strong>the</strong><br />

study <strong>of</strong> various systems. 11<br />

Before any comparison is made, it is important to point out<br />

that all sec-butyl isotopomers share <strong>the</strong> same PES within <strong>the</strong><br />

Born–Oppenheimer (BO) framework, since <strong>the</strong> electronic<br />

Hamiltonian only parametrically depends on nuclear posi-<br />

tions and has nothing to do with nuclear masses. Thus, a<br />

conformational study within <strong>the</strong> BO approximation is equally<br />

applicable for both regular (unmuoniated) and muoniated secbutyl<br />

radicals (from muonium addition to 2-butene).<br />

Based on <strong>the</strong> analysis <strong>of</strong> muon spin resonance (mSR) data,<br />

<strong>the</strong> muoniated sec-butyl radical formed from cis-2-butene<br />

has a large 0 K HFCC value on muon and this value is different<br />

from <strong>the</strong> muon HFCC from muonium addition to<br />

trans-2-butene. Fur<strong>the</strong>r temperature-dependent HFCC study<br />

also supported <strong>the</strong> existence <strong>of</strong> two distinct muoniated secbutyl<br />

radicals, both with large values <strong>of</strong> HFCCs on muon.<br />

However, calculations using <strong>the</strong> popular hybrid B3LYP-<br />

DFT Hamiltonian, normally believed to be well suited to<br />

predict HFCCs <strong>of</strong> small organic molecules, 12 only located<br />

two equilibrium structures on <strong>the</strong> ground-state <strong>torsional</strong><br />

PES: 9,10 one corresponding to <strong>the</strong> radical formed from muonium<br />

addition to trans-2-butene and <strong>the</strong> o<strong>the</strong>r, a gauche conformation<br />

with HFCC on muon only about one-third <strong>of</strong> <strong>the</strong><br />

experimental value. The absence <strong>of</strong> <strong>the</strong> experimentally observed<br />

muoniated sec-butyl formed from cis-2-butene from<br />

B3LYP calculations prompted us to carry out extensive <strong>the</strong>oretical<br />

<strong>studies</strong> <strong>of</strong> <strong>the</strong> <strong>torsional</strong> PES <strong>of</strong> <strong>the</strong> sec-butyl radical.<br />

Besides providing static conformations, <strong>the</strong> PES also <strong>of</strong>fers<br />

valuable information for temperature-dependent HFCC <strong>studies</strong>.<br />

The dynamic (vibrational) and <strong>the</strong>rmodynamic corrections<br />

obtained based on <strong>the</strong> PES can be used to calculate <strong>the</strong><br />

temperature-dependent behaviour <strong>of</strong> HFCCs by solving <strong>torsional</strong><br />

Schrödinger equations. Something worrisome is that<br />

<strong>the</strong> non-BO effect may in turn be significant in a molecule<br />

containing muon, which may weaken <strong>the</strong> validity <strong>of</strong> our<br />

study based on <strong>the</strong> BO approximation. However, a study <strong>of</strong><br />

<strong>the</strong> reaction rate <strong>of</strong> Mu + H 2 indicates that <strong>the</strong> non-BO effect<br />

only changes <strong>the</strong> reaction barrier height by 3.8%. 13 Therefore,<br />

we can expect much less non-BO effect in <strong>the</strong> muoniated<br />

sec-butyl radicals in <strong>the</strong> current study.<br />

To our knowledge, some o<strong>the</strong>r experimental results from<br />

<strong>the</strong> homolysis <strong>of</strong> butyl radicals are also available. Thermodynamic<br />

and kinetic <strong>studies</strong> <strong>of</strong> <strong>the</strong> decomposition <strong>of</strong> <strong>the</strong> sec-butyl<br />

radical into propene and <strong>the</strong> methyl radical were performed, 14<br />

but detailed conformational information <strong>of</strong> <strong>the</strong> sec-butyl<br />

radical could not be inferred from <strong>the</strong> homolysis study, since<br />

<strong>the</strong> experiment carried out at high temperature averaged out<br />

<strong>the</strong> behaviours <strong>of</strong> all distinct conformations. Therefore, <strong>the</strong>oretical<br />

calculations have to be done to illuminate possible<br />

conformations <strong>of</strong> <strong>the</strong> sec-butyl radical. Thus far, <strong>the</strong> first ab<br />

initio study on sec-butyl conformations can be traced back to<br />

<strong>the</strong> Hartree–Fock (HF) calculations in <strong>the</strong> 1980s, 15 in which<br />

<strong>the</strong> dynamic correlation effect was not taken into account.<br />

Therefore, a systematical study <strong>of</strong> <strong>the</strong> sec-butyl radical with<br />

various correlated Hamiltonians and different basis sets is<br />

much needed.<br />

Computational details<br />

All calculations in this study were carried out using <strong>the</strong> Gaussian<br />

03 package within <strong>the</strong> BO framework. 16 Different first<strong>principles</strong><br />

electronic Hamiltonians, including spin-unrestricted<br />

HF, MP2, CCSD, CISD, QCISD, CCSD(T), and a variety<br />

<strong>of</strong> Kohn–Sham DFT methods were utilized. Wave functions<br />

were expanded in terms <strong>of</strong> atomic-centred one-electron<br />

basis functions including Pople-type basis sets (6–31g(d),<br />

Published by NRC Research Press


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

Chen and Wang 1471<br />

Fig. 1. Hamiltonian and basis-set effects on <strong>the</strong> <strong>potential</strong> energy surface <strong>of</strong> <strong>the</strong> sec-butyl radical. (a)–(d) Hamiltonian effects using <strong>the</strong> ccpVDZ<br />

basis set. (e) and (f) Basis-set effects based on B3LYP and MP2 Hamiltonians, respectively.<br />

Energy (kcal/mol)<br />

Energy (kcal/mol)<br />

Energy (kcal/mol)<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

TS0<br />

cis<br />

TS1<br />

gauche<br />

trans<br />

0<br />

0 30 60 90 120 150 180<br />

2.5<br />

2<br />

1.5<br />

TS0<br />

Torsional angle<br />

CCSD<br />

CISD<br />

MP2<br />

B3LYP<br />

HF<br />

TS2<br />

TS180<br />

( a ) WFT and B3LYP<br />

( b)<br />

DFT (no cis)<br />

1<br />

(TS1)<br />

(cis)<br />

TS2<br />

0.5<br />

0<br />

gauche<br />

TS180<br />

trans<br />

0 30 60 90 120 150 180<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Torsional angle<br />

( c)<br />

DFT (flat)<br />

BB95<br />

PBEPBE<br />

PW91B95<br />

PW91PW91<br />

6-311++g(3df,2pd)<br />

6-311++g(d,p)/ultrafine grid<br />

6-311++g(d,p)<br />

6-31g(d,p)<br />

AUG-cc-pVTZ<br />

cc-pVTZ<br />

AUG-cc-pVDZ<br />

cc-pVDZ/ultrafine grid<br />

cc-pVDZ<br />

0<br />

0 30 60 90 120 150 180<br />

Torsional angle<br />

( e)<br />

B3LYP<br />

6–311++g(d,p), and 6–311++g(3df,2pd)), 17,18 and Dunning’s<br />

correlation-consistent basis sets (cc-pVDZ, cc-pVTZ, AUGcc-pVDZ,<br />

and AUG-cc-pVTZ). 19 For <strong>the</strong> DFT calculations,<br />

two-electron integrals were numerically integrated with <strong>the</strong><br />

default grid quality in Gaussian 03 (75 radical shells and<br />

302 angular points per shell) unless o<strong>the</strong>rwise specified.<br />

Geometry optimization was performed using <strong>the</strong> z-matrix<br />

coordinate system. Completely relaxed <strong>torsional</strong> PESs <strong>of</strong> <strong>the</strong><br />

Energy (kcal/mol)<br />

Energy (kcal/mol)<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

TS0<br />

cis<br />

TS1<br />

gauche<br />

TS2<br />

trans<br />

0<br />

0 30 60 90 120 150 180<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Torsional angle<br />

( d)<br />

DFT (cis)<br />

MPwB1K<br />

MPw1K<br />

MPw1B95<br />

BHandH<br />

PBE0<br />

SVWN<br />

6-311++g**<br />

6-311+g**<br />

6-311g**<br />

6-31g**<br />

AUG-cc-pVTZ<br />

cc-pVTZ<br />

AUG-cc-pVDZ<br />

cc-pVDZ<br />

TS180<br />

0<br />

0 30 60 90 120 150 180<br />

Torsional angle<br />

()MP2 f<br />

sec-butyl radical as a function <strong>of</strong> <strong>the</strong> dihedral angle (q) <strong>of</strong> <strong>the</strong><br />

four backbone carbon atoms were plotted using selected combinations<br />

<strong>of</strong> electronic Hamiltonians and basis sets (Fig. 1).<br />

The dihedral angle was scanned from 0° to 180° with a step<br />

size <strong>of</strong> 5°, while all <strong>the</strong> degrees <strong>of</strong> freedom o<strong>the</strong>r than <strong>the</strong><br />

backbone dihedral angle were allowed to be fully optimized.<br />

The PES pr<strong>of</strong>ile <strong>of</strong> q in <strong>the</strong> range from 180° to 360° is essentially<br />

<strong>the</strong> mirror image <strong>of</strong> <strong>the</strong> pr<strong>of</strong>ile from 0° to 180°, if<br />

Published by NRC Research Press


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

1472 Can. J. Chem. Vol. 89, 2011<br />

certain symmetry constraints are applied at proper points.<br />

However, for <strong>the</strong> simplicity and uniformity <strong>of</strong> <strong>the</strong> calculations,<br />

<strong>the</strong> symmetry <strong>of</strong> <strong>the</strong> molecule was subject to no constraint<br />

throughout unless o<strong>the</strong>rwise noted.<br />

In addition, <strong>the</strong> optimized conformations were calculated at<br />

both <strong>the</strong> MP2 and B3LYP levels to obtain <strong>the</strong> diagonalized<br />

Hessian matrices to identify <strong>the</strong> nature <strong>of</strong> <strong>the</strong> stationary structures<br />

and to include <strong>the</strong> zero-point vibrational corrections.<br />

Because <strong>the</strong> harmonic approximation is usually inappropriate<br />

for <strong>the</strong> shallow minimum <strong>of</strong> <strong>the</strong> cis-structure, anharmonic vibrational<br />

calculations were also carried out at <strong>the</strong> MP2 level<br />

<strong>of</strong> <strong>the</strong>ory to obtain <strong>the</strong> anharmonic zero-point corrections.<br />

In this study, single-reference spin-unrestricted Hamiltonians<br />

were used. Spin contamination, believed to be responsible<br />

for <strong>the</strong> failure <strong>of</strong> UMP2 in determining <strong>the</strong> reaction<br />

barrier <strong>of</strong> radical formation reactions, 20,21 was found to be<br />

very small for all <strong>the</strong> spin-unrestricted Hamiltonians in <strong>the</strong><br />

current study. This is not surprising because it is well established<br />

that <strong>the</strong> simple alkyl radical has a localized unpaired<br />

electron at <strong>the</strong> radical centre. As a result, a single-reference<br />

method should be adequate to treat alkyl radicals well. Thus,<br />

spin-unrestricted single-reference methods were utilized in<br />

this study.<br />

Results and discussion<br />

The <strong>torsional</strong> PES pr<strong>of</strong>iles calculated using different Hamiltonians<br />

and basis sets are shown in Fig. 1, with <strong>the</strong>ir lowest<br />

points as <strong>the</strong> common reference zero point to simplify <strong>the</strong><br />

comparison. The cc-pVDZ basis set was utilized in<br />

Figs. 1a–1d because geometry optimization is usually carried<br />

out with a relatively small basis set and cc-pVDZ was found<br />

to be adequate for this purpose. 3<br />

In Fig. 1a, <strong>the</strong> <strong>torsional</strong> energy curve obtained from <strong>the</strong><br />

HF calculation lies above all o<strong>the</strong>r <strong>potential</strong> energy pr<strong>of</strong>iles.<br />

The PES pr<strong>of</strong>iles from <strong>the</strong> CCSD, B3LYP, and MP2 calculations<br />

lie at <strong>the</strong> bottom across each o<strong>the</strong>r, whereas <strong>the</strong> <strong>potential</strong><br />

energy curve obtained from <strong>the</strong> CISD calculation is<br />

situated in <strong>the</strong> middle. The PES calculated at <strong>the</strong> B3LYP<br />

level drops at a slower rate with respect to <strong>the</strong> dihedral angle<br />

increment when compared with all <strong>the</strong> WFT methods. The<br />

B3LYP calculation predicts two stable conformations when<br />

<strong>the</strong> dihedral angle q is less than 180°, whereas all <strong>the</strong> WFT<br />

methods predict three stable conformations. The most stable<br />

conformation is uniformly predicted to be a trans-like structure<br />

at q ≈ 170° (conformation I in Fig. 2f). A stable gauche<br />

conformation (structure II in Fig. 2d) is also universally predicted<br />

at q ≈ 80°. This gauche local minimum is ~0.3–<br />

0.6 kcal/mol (1 cal = 4.184 J) energetically higher than <strong>the</strong><br />

global minimum, depending on <strong>the</strong> Hamiltonian used.<br />

The existence <strong>of</strong> a third cis-like conformation (structure III<br />

in Fig. 2b) is worth some elaboration. All <strong>the</strong> WFT methods<br />

predict <strong>the</strong> existence <strong>of</strong> a stable structure on a relatively flat<br />

segment <strong>of</strong> <strong>the</strong> PES curves around q ≈ 50°, whereas selected<br />

B3LYP methods predict an accelerated energy drop with increased<br />

dihedral angle in <strong>the</strong> same range. This cis-like conformation<br />

was earlier recognized as one <strong>of</strong> <strong>the</strong> three stable<br />

conformations in <strong>the</strong> very first <strong>the</strong>oretical conformation study<br />

<strong>of</strong> <strong>the</strong> sec-butyl radical by Chen et al. 15 The correlation effects,<br />

which are important to describe dispersion interactions<br />

for a correct prediction <strong>of</strong> stable conformations, was not in-<br />

cluded by <strong>the</strong> HF calculations performed by Chen et al. 15<br />

though. Thus, methods taking account <strong>of</strong> correlation effects<br />

must be employed here for a reliable conformation study.<br />

Both DFT (B3LYP) and high-level post-HF WFT methods<br />

can incorporate correlation effects, but predict different <strong>torsional</strong><br />

pr<strong>of</strong>iles for <strong>the</strong> sec-butyl radical in this study. As verified<br />

by a mSR experiment, <strong>the</strong> cis-like conformation does<br />

exist and carries a large value <strong>of</strong> HFCC on <strong>the</strong> b-muon. 9,10<br />

In this regard, WFT methods outperform B3LYP-DFT methods<br />

decisively.<br />

When q = 0°, <strong>the</strong> HF calculation predicts that <strong>the</strong> barrier<br />

TS0 (Fig. 2a), <strong>the</strong> global maximum and <strong>the</strong> transition state<br />

connecting <strong>the</strong> cis-conformation with its mirror image, is<br />

2.3 kcal/mol higher (less stable) than <strong>the</strong> global mininum,<br />

<strong>the</strong> trans-like conformation. This indicates that <strong>the</strong> HF calculation<br />

predicts stronger bulky repulsion (by about 0.4 kcal/mol)<br />

between <strong>the</strong> two terminal methyl groups <strong>of</strong> <strong>the</strong> sec-butyl<br />

radical than any o<strong>the</strong>r correlated methods owing to <strong>the</strong> neglect<br />

<strong>of</strong> <strong>the</strong> correlation effects. The barrier separating conformations<br />

II and III, TS1 (Fig. 2c), is located around q ≈<br />

55° using correlated WFT methods and is at q ≈ 60° using<br />

<strong>the</strong> HF Hamiltonian. This TS1 barrier is very low energetically<br />

with respect to conformation III and can be easily<br />

overcome. The height <strong>of</strong> <strong>the</strong> energy barrier between conformations<br />

I and II, TS2 (Fig. 2e), in comparison to conformation<br />

II is about 0.5 kcal/mol from WFT calculations,<br />

whereas it drops to as low as ~0.1 kcal/mol using <strong>the</strong><br />

B3LYP method. At q ≈ 180°, <strong>the</strong> actual <strong>torsional</strong> barrier<br />

between conformation I and its mirror image was not obtained<br />

directly from <strong>the</strong> PES curve because <strong>the</strong> symmetry<br />

was not constrained during <strong>the</strong> relaxed <strong>potential</strong> energy<br />

scan. Instead, <strong>the</strong> C s symmetry was used to help acquire<br />

<strong>the</strong> <strong>torsional</strong> barrier at q ≈ 180°. Based on separate calculations,<br />

this TS180 <strong>torsional</strong> barrier is 0.13 kcal/mol at <strong>the</strong><br />

B3LYP/cc-pVDZ level and 0.29 kcal/mol at MP2/cc-pVDZ.<br />

For <strong>the</strong> sake <strong>of</strong> simplicity, no fur<strong>the</strong>r effort was devoted to<br />

obtain this barrier height using o<strong>the</strong>r methods.<br />

Besides <strong>the</strong> comparison between WFT methods and <strong>the</strong><br />

representative B3LYP Hamiltonian, DFT methods with various<br />

functionals were also examined using <strong>the</strong> cc-pVDZ basis<br />

set. Based on <strong>the</strong> shape <strong>of</strong> <strong>the</strong> calculated <strong>potential</strong> energy<br />

pr<strong>of</strong>iles between q = 40° and 60°, where <strong>the</strong> presence <strong>of</strong> conformation<br />

III is <strong>of</strong> interest, <strong>the</strong> functionals are classified into<br />

three different categories (Figs. 1b–1d).<br />

Figure 1b shows <strong>the</strong> functionals that predict no minimum<br />

within <strong>the</strong> range <strong>of</strong> interest. This group comprises some popular<br />

hybrid or pure GGA functionals. The PES pr<strong>of</strong>ile obtained<br />

from <strong>the</strong> VSXC calculation (inset <strong>of</strong> Fig. 1b) is<br />

erratically different from o<strong>the</strong>r functionals. The VSXC functional<br />

fails to describe <strong>the</strong> <strong>potential</strong> energy needed to approach<br />

<strong>the</strong> energy maximum at q = 0° with <strong>the</strong> big overall<br />

steric repulsion between <strong>the</strong> terminal methyl groups. The<br />

o<strong>the</strong>r functionals in this group yield very similar <strong>potential</strong> energy<br />

pr<strong>of</strong>iles with only two stable conformations (I and II).<br />

At any particular value <strong>of</strong> <strong>the</strong> dihedral angle, <strong>the</strong> energy difference<br />

between any two functionals in this group (VSXC<br />

excluded) is within 0.2 kcal/mol.<br />

In Fig. 1c, <strong>the</strong> second group <strong>of</strong> functionals predict relative<br />

flat features around q = 60° before <strong>the</strong> accelerated energy<br />

drop to conformation II. Usually, geometry optimization has<br />

a hard time to converge to a stationary point if <strong>the</strong> <strong>potential</strong><br />

Published by NRC Research Press


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

Chen and Wang 1473<br />

Fig. 2. Optimized stationary structures on <strong>the</strong> <strong>torsional</strong> <strong>potential</strong> energy surface <strong>of</strong> <strong>the</strong> sec-butyl radical.<br />

gradient is very small in a case like this. Thus, fur<strong>the</strong>r calculations<br />

to verify <strong>the</strong> existence <strong>of</strong> a minimum might not be<br />

successful. Note that some <strong>of</strong> <strong>the</strong>se functionals contain components<br />

like PW X91 or B C95 that were designed for systems<br />

characterized by weak interactions, 22–24 which may partly<br />

compensate for <strong>the</strong> weakness <strong>of</strong> <strong>the</strong> DFT models in Fig. 1b<br />

so that a flat segment around q = 55° is predicted.<br />

In Fig. 1d, calculations using <strong>the</strong> third group <strong>of</strong> functionals<br />

predict a shallow <strong>potential</strong> well lying on <strong>the</strong> <strong>torsional</strong> PES<br />

around q = 50°, similar to <strong>the</strong> PES pr<strong>of</strong>iles obtained from<br />

MP2 and o<strong>the</strong>r WFT calculations. However, it is well-known<br />

that <strong>the</strong> SVWN functional, as with any o<strong>the</strong>r LDA-type functionals,<br />

should not be trusted unless <strong>the</strong> electron density<br />

undergoes only a small variation in space. The isolated molecular<br />

systems usually exhibit big density gradients especially<br />

around <strong>the</strong> nuclei and <strong>the</strong>ir electronic structure cannot<br />

be well predicted by LDA calculations. Just as in this case,<br />

<strong>the</strong> SVWN calculation predicts a distinct PES curve, quite<br />

different from its peers. This indicates that SVWN fails to<br />

capture <strong>the</strong> nonlocal exchange and correlation effects. Different<br />

from what it is usually referred to, BHandH, comprising<br />

50% exact exchange and 50% LDA-level exchange and LYP<br />

correlation functionals as implemented in Gaussian 03, 16 is<br />

also inappropriate for <strong>the</strong> conformation study <strong>of</strong> <strong>the</strong> sec-butyl<br />

radical because it produces a much deeper <strong>potential</strong> well<br />

about conformation III. 25 The calculation using <strong>the</strong> PBE0 (or<br />

PBE1PBE) exchange-correlation functional, a hybrid functional<br />

with about 25% exact exchange, predicts a <strong>potential</strong><br />

energy curve with a shallow <strong>potential</strong> well around q = 50°,<br />

in contrast to <strong>the</strong> flat surface predicted by <strong>the</strong> pure PBEPBE<br />

calculation. The o<strong>the</strong>r three functionals in this group combine<br />

mPW (modified PW X91) exchange with exact exchange and<br />

PW C91 or meta-B C95 correlation functionals that were designed<br />

for a better long-range behaviour. PBE0 and MPW1K<br />

calculations produce similar <strong>potential</strong> energy curves to those<br />

from correlated WFT calculations. MPW1B95 and<br />

mPWB1K, on <strong>the</strong> o<strong>the</strong>r hand, predict low-lying curves in <strong>the</strong><br />

range <strong>of</strong> q =0°–60°, a clear indication <strong>of</strong> <strong>the</strong> underestimated<br />

repulsion <strong>of</strong> <strong>the</strong> terminal methyl groups compared with o<strong>the</strong>r<br />

Hamiltonians.<br />

Based on <strong>the</strong> PES curves <strong>of</strong> each exchange-correlation<br />

functional combination, it is evident that incorporating <strong>the</strong><br />

exact exchange is crucial for locating <strong>the</strong> cis-like conformation<br />

III. This implies that <strong>the</strong> nonlocal exchange-correlation<br />

effects may play an important role in this conformation study.<br />

At <strong>the</strong> same time, correlation functionals should also be carefully<br />

chosen, since mPW3LYP in group one, using a similar<br />

exchange-correlation functional as for mPW1B95, predicts a<br />

PES pr<strong>of</strong>ile without any sign <strong>of</strong> <strong>the</strong> cis conformation III.<br />

The basis-set effect was examined using both B3LYP and<br />

MP2 Hamiltonians. Based on some early <strong>studies</strong>, 4,5 <strong>the</strong> quality<br />

<strong>of</strong> basis functions can be more important than <strong>the</strong> form <strong>of</strong><br />

<strong>the</strong> Hamiltonians in conformational analysis. The cc-pVDZ<br />

and 6–31g(d) basis sets were believed to be too small to include<br />

correct dispersion interaction, which lead to overstabilization<br />

<strong>of</strong> <strong>the</strong> coplanar conformer <strong>of</strong> biphenyl. 4,5 In a <strong>the</strong>oretical<br />

study <strong>of</strong> radicals, calculations using <strong>the</strong> Pople-type basis<br />

sets with various Hamiltonians approach converged results<br />

only after <strong>the</strong> size <strong>of</strong> <strong>the</strong> basis set went beyond 6–311g(d). 26<br />

To visualize <strong>the</strong> basis-set effect, <strong>the</strong> <strong>torsional</strong> PES pr<strong>of</strong>iles<br />

<strong>of</strong> <strong>the</strong> sec-butyl radical at <strong>the</strong> B3LYP and MP2 levels with<br />

different basis sets are plotted in Figs. 1e and 1f, respectively.<br />

Because <strong>of</strong> its single determinant nature, <strong>the</strong> Kohn–Sham<br />

DFT method is usually less sensitive to <strong>the</strong> variation <strong>of</strong> <strong>the</strong><br />

basis set than multideterminant post-HF methods. In Fig. 1e,<br />

<strong>the</strong> PES pr<strong>of</strong>iles appear to be insensitive to <strong>the</strong> selection <strong>of</strong><br />

<strong>the</strong> basis sets when <strong>the</strong>ir sizes grow bigger than cc-pVDZ or<br />

6–31g(d,p), both <strong>of</strong> which predict <strong>the</strong> appearance <strong>of</strong> a small<br />

cusp at q ≈ 55°. Calculations with larger basis sets predict<br />

smooth and high-lying <strong>potential</strong> energy curves compared with<br />

those with cc-pVDZ or 6–31g(d,p). The energy difference between<br />

calculations using any two basis sets is no more than<br />

0.2 kcal/mol at any given dihedral angle. The convergence <strong>of</strong><br />

basis sets is quickly achieved beyond <strong>the</strong> 6–311++g(d,p)<br />

basis set. With <strong>the</strong> size <strong>of</strong> <strong>the</strong> basis sets increased within<br />

Published by NRC Research Press


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

1474 Can. J. Chem. Vol. 89, 2011<br />

Table 1. Relative energies (in kcal/mol; 1 cal = 4.184 J) and some geometric parameters <strong>of</strong> <strong>the</strong> stationary structures <strong>of</strong> <strong>the</strong> sec-butyl radical calculated at MP2 and B3LYP levels using<br />

<strong>the</strong> cc-pVDZ basis set.<br />

Structure Hamiltonian C1–C2 (Å) C2–C3 (Å) C3–C4 (Å) C2–H4 (Å) C1–C2–C3 (°) q (°) Eele (kcal/mol) EZPC (kcal/mol) EZPC,Mu (kcal/mol)<br />

TS0 MP2 1.499 1.505 1.532 1.094 123.7 0.0 1.788 1.547 (1.772) 1.719 (1.906)<br />

B3LYP 1.493 1.500 1.532 1.094 124.4 0.0 1.740 1.694 1.891<br />

III (cis) MP2 1.500 1.503 1.535 1.095 120.9 47.8 0.683 0.814 (0.711) 0.819 (0.735)<br />

B3LYP NA NA NA NA NA NA NA NA NA<br />

TS1 MP2 1.499 1.501 1.539 1.093 120.6 55.2 0.779 0.476 (0.607) 0.572 (0.641)<br />

B3LYP NA NA NA NA NA NA NA NA NA<br />

II (gauche) MP2 1.499 1.501 1.543 1.097 120.4 74.1 0.312 0.512 (0.486) 0.904 (0.854)<br />

B3LYP 1.492 1.497 1.546 1.096 122.0 84.6 0.553 0.788 1.308<br />

TS2 MP2 1.497 1.500 1.539 1.095 121.6 116.7 0.841 0.509 (0.918) 0.719 (1.121)<br />

B3LYP 1.492 1.497 1.542 1.096 122.6 114.0 0.682 0.632 0.897<br />

I (trans) MP2 1.498 1.500 1.532 1.097 120.4 169.0 0.000 0.000 (0.000) 0.000 (0.000)<br />

B3LYP 1.492 1.495 1.532 1.096 121.5 167.9 0.000 0.000 0.000<br />

TS180 MP2 1.497 1.498 1.531 1.094 120.4 180.0 0.299 –0.293 (–0.176) –0.157 (–0.064)<br />

B3LYP 1.491 1.494 1.531 1.094 121.1 180.0 0.126 –0.283 –0.121<br />

Note: Bond lengths are in Å. Bond angles and dihedral angles (q) are in degrees. The relative energies are measured from <strong>the</strong> global minimal conformation III. The harmonic vibrational zero-pointcorrected<br />

energy values are shown in <strong>the</strong> last two columns for hydrogen and muoniated isotopomers, respectively. For <strong>the</strong> MP2 calculation results, <strong>the</strong> anharmonic vibrational zero-point-corrected energies are<br />

shown in paren<strong>the</strong>ses. NA, not available; Eele, relative energy.<br />

<strong>the</strong> Pople basis-set family, <strong>the</strong> PES curve moves upwards<br />

without exception. However, <strong>the</strong> basis-set effect within <strong>the</strong><br />

Dunning family <strong>of</strong> basis sets is complicated. Adding diffuse<br />

functions to <strong>the</strong> cc-pVDZ or cc-pVTZ basis sets greatly<br />

raises <strong>the</strong> relative energies <strong>of</strong> <strong>the</strong> gauche conformation (II)<br />

and <strong>the</strong> TS2 barrier separating conformations I and II. On<br />

<strong>the</strong> o<strong>the</strong>r hand, incrementing splitting valence functions or<br />

polarization functions to <strong>the</strong> basis set results in a moderate<br />

relative energy rise (to more unstable) <strong>of</strong> conformation II<br />

and TS2, but a dramatic energy increase <strong>of</strong> TS0. This implies<br />

that, when q is between 70° and 130°, long-range interactions<br />

may dominate <strong>the</strong> PES where diffuse basis<br />

functions play an important role. Meanwhile, adding splitting<br />

valence and polarization functions noticeably influences<br />

<strong>the</strong> PES curves at small torsion angles.<br />

O<strong>the</strong>r than <strong>the</strong> basis-set effects, <strong>the</strong> quality <strong>of</strong> numeric integration<br />

also affects <strong>the</strong> accuracy <strong>of</strong> a DFT calculation. A<br />

recent <strong>the</strong>oretical study <strong>of</strong> <strong>the</strong> sec-butyl cation showed that<br />

<strong>the</strong> quality <strong>of</strong> numerical integration <strong>of</strong> a DFT study that uses<br />

grid summation can be important for an accurate prediction<br />

<strong>of</strong> <strong>the</strong> equilibrium geometries. 27 Thus, an ultra-fine grid <strong>of</strong><br />

higher quality consisting <strong>of</strong> 99 radical shells and 590 angular<br />

points per shell was also tested, but no significant difference<br />

to <strong>the</strong> default grid quality was found for <strong>the</strong> system we<br />

studied here.<br />

Unlike <strong>the</strong> Kohn–Sham DFT and HF methods, <strong>the</strong> MP2<br />

method, as with many o<strong>the</strong>r correlated WFT methods using<br />

excited configurations, is normally sensitive to <strong>the</strong> size <strong>of</strong><br />

<strong>the</strong> basis set employed because correlated WFT methods utilize<br />

virtual orbitals that are less optimized during <strong>the</strong> SCF<br />

procedure. Thus, in most cases, MP2 calculations require a<br />

large basis set to achieve a certain level <strong>of</strong> accuracy. A study<br />

<strong>of</strong> biphenyl <strong>torsional</strong> PES indicated that <strong>the</strong> unbalanced treatment<br />

<strong>of</strong> repulsive and attractive dispersion interactions occurred<br />

if only a small Dunning correlation consisted basis<br />

set was used. 4,5 In <strong>the</strong> present study, except for those using<br />

small basis sets like cc-pVDZ and 6–31g(d,p) with which<br />

<strong>the</strong> repulsion between <strong>the</strong> terminal methyl groups is overestimated,<br />

<strong>the</strong> PES pr<strong>of</strong>iles obtained from o<strong>the</strong>r Pople basis sets<br />

lie below <strong>the</strong> <strong>potential</strong> energy pr<strong>of</strong>iles obtained from <strong>the</strong><br />

Dunning correlation-consisted basis sets. The congregations<br />

<strong>of</strong> <strong>the</strong> PES pr<strong>of</strong>iles within each basis-set family suggest a<br />

fast convergence within each basis-set family. On <strong>the</strong> o<strong>the</strong>r<br />

hand, some significant difference persists between <strong>the</strong> two<br />

sets <strong>of</strong> PES curves from <strong>the</strong> two basis-set families, indicating<br />

that ei<strong>the</strong>r or both <strong>of</strong> <strong>the</strong>se basis-set families are still far away<br />

from <strong>the</strong> complete basis-set limit.<br />

To draw a convincing conclusion, calculations using more<br />

accurate Hamiltonians with larger basis sets are always<br />

desired. Limited by our computer resources and <strong>the</strong> complexity<br />

<strong>of</strong> <strong>the</strong> open-shell systems involved, only single-point<br />

CCSD(T)/6–311++g(d,p) calculations using MP2/cc-pVTZoptimized<br />

geometries were carried out for benchmarking purposes.<br />

In <strong>the</strong> end, <strong>the</strong> resulting “true” <strong>torsional</strong> PES pr<strong>of</strong>ile<br />

calculated from CCSD(T) very closely resembles <strong>the</strong> MP2<br />

one, which lends confidence to <strong>the</strong> results from <strong>the</strong> MP2 calculations.<br />

As occurs here, disagreement between <strong>the</strong> DFT and WFT<br />

methods also took place in earlier <strong>studies</strong>. Some good agreements<br />

to <strong>the</strong> experimental data from DFT calculations were<br />

attributed to <strong>the</strong> parameterization with which <strong>the</strong> functionals<br />

Published by NRC Research Press


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

Chen and Wang 1475<br />

were developed. DFT methods were reported to be unsuitable<br />

for <strong>the</strong> conformation study <strong>of</strong> 2,2′-bifuran for missing a shallow<br />

gauche conformation on <strong>the</strong> <strong>torsional</strong> PES curve. 8 It was<br />

also noted that MP2 and DFT gave very similar geometries,<br />

whereas DFT failed to describe <strong>the</strong> energetics <strong>of</strong> <strong>the</strong> <strong>torsional</strong><br />

<strong>potential</strong> <strong>of</strong> biisothianaph<strong>the</strong>ne. 6 In addition, many popular<br />

density functionals are believed inappropriate to describe<br />

noncovalent interactions and are difficult to improve systematically.<br />

In contrast to DFT methods, correlated WFT methods<br />

are in principle controllably adequate and, although with<br />

much higher <strong>computational</strong> cost, can be progressively improved.<br />

High-level correlated post-HF methods can <strong>the</strong>n be<br />

used as a benchmark to justify low-level calculations. Alternatively,<br />

comparison between <strong>the</strong>oretical predictions and<br />

experimental data provides ano<strong>the</strong>r way to validate <strong>the</strong> <strong>the</strong>ories.<br />

Up to today, one <strong>of</strong> <strong>the</strong> most convincing ways to tell<br />

whe<strong>the</strong>r a <strong>the</strong>oretical method can be trusted on a specific system<br />

is to make a comparison with available experimental<br />

data. In our current study, all WFT methods predict <strong>the</strong> existance<br />

<strong>of</strong> <strong>the</strong> cis-like conformation III irrespective <strong>of</strong> <strong>the</strong> selection<br />

<strong>of</strong> <strong>the</strong> basis sets. A comparison with available<br />

experiments has confirmed <strong>the</strong> existence <strong>of</strong> <strong>the</strong> cis-like conformation<br />

III by low-temperature HFCC measurements. 9,10<br />

Thus, <strong>the</strong> WFT methods and several DFT methods in group<br />

III (Fig. 1d) outperform o<strong>the</strong>r <strong>the</strong>oretical methods by being<br />

more in line with experimental results.<br />

The stable geometries <strong>of</strong> <strong>the</strong> sec-butyl radical were optimized<br />

at both B3LYP and MP2 levels with <strong>the</strong> cc-pVDZ<br />

basis set. The optimized geometric parameters <strong>of</strong> those stable<br />

conformations are listed in Table 1. These two employed<br />

methods predict very similar bond lengths and bond angles<br />

for <strong>the</strong> same stationary structure except that <strong>the</strong> dihedral angle<br />

<strong>of</strong> <strong>the</strong> gauche conformation calculated at <strong>the</strong> B3LYP level<br />

differs by about 10° from <strong>the</strong> MP2 calculation. The C2–C3<br />

bond length descends almost monotonically as <strong>the</strong> torsion angle<br />

increases from 0° to 180°. This indicates that <strong>the</strong> C2–C3<br />

bond grows stronger from <strong>the</strong> cis-conformation to <strong>the</strong> transconformation<br />

because <strong>of</strong> decreased repulsion between <strong>the</strong><br />

two terminal methyl groups. The hydrogen–hydrogen repulsion<br />

and hyperconjugation involving C2 and C3 exert only<br />

minor corrections to <strong>the</strong> PES pr<strong>of</strong>iles because <strong>the</strong> energy barrier<br />

from conformation III to conformation II is very small.<br />

Based on <strong>the</strong> optimized structures, <strong>the</strong> C1–C2 bond distance<br />

reaches maximum in TS1. This implies that <strong>the</strong> internal rotation<br />

about <strong>the</strong> C1–C2 bond plays an important role in forming<br />

<strong>the</strong> rotational barrier between conformations II and III.<br />

In an early <strong>the</strong>oretical study by Chen et al., 15 <strong>the</strong> shallow<br />

<strong>potential</strong> well embracing conformation III was merely regarded<br />

as a shoulder extension to <strong>the</strong> electronically more stable<br />

conformation II. If only <strong>the</strong> static electronic <strong>potential</strong><br />

energy is considered, muoniated sec-butyl formed from mounium<br />

addition to cis-2-butene will not exist or can be easily<br />

turned into <strong>the</strong> gauche conformation II by an extremely small<br />

fluctuation (e.g., through vibration) to overcome <strong>the</strong> tiny barrier<br />

TS1, contradicting <strong>the</strong> experimental observation that <strong>the</strong><br />

muon HFCC <strong>of</strong> this structure remained visibly strong over a<br />

wide temperature range. 9,10 There is no doubt that o<strong>the</strong>r effects<br />

stabilizing this molecular species must be taken into<br />

consideration.<br />

In addition to <strong>the</strong> electronic energy, <strong>the</strong> zero-point vibrational<br />

correction also affects <strong>the</strong> relative energy difference<br />

between two structures in reality. More complete <strong>the</strong>rmodynamic<br />

correction must be considered if <strong>the</strong> experiment is<br />

conducted at a higher temperature. To make a sound comparison<br />

to <strong>the</strong> experimental data at low temperatures, both harmonic<br />

and anharmonic vibrational corrections <strong>of</strong> <strong>the</strong> stable<br />

structures and <strong>the</strong> corresponding <strong>the</strong>rmal corrections were<br />

calculated and are summarized in Table 1. It turns out that<br />

<strong>the</strong> zero-point corrected energy <strong>of</strong> <strong>the</strong> muoniated isotopomer<br />

<strong>of</strong> conformation III becomes lower than that <strong>of</strong> <strong>the</strong> muoniated<br />

conformation II, opposite to <strong>the</strong> prediction based on<br />

electronic energies alone. Thus, <strong>the</strong> large value <strong>of</strong> <strong>the</strong> muon<br />

HFCC <strong>of</strong> <strong>the</strong> muoniated radical formed from cis-2-butene<br />

can be definitely assigned after including zero-point correction.<br />

Ano<strong>the</strong>r change in relative stability owing to <strong>the</strong> inclusion<br />

<strong>of</strong> <strong>the</strong> vibrational correction is <strong>the</strong> transition state TS180<br />

connecting conformation I and its mirror image: <strong>the</strong> transitionstate<br />

structure becomes <strong>the</strong> lowest point after <strong>the</strong> vibrational<br />

correction. Similar effects <strong>of</strong> <strong>the</strong> zero-point correction were<br />

also reported in o<strong>the</strong>r <strong>studies</strong> but without detailed discussion.<br />

7<br />

Conclusion<br />

<strong>First</strong>-<strong>principles</strong> calculations using spin-unrestricted HF,<br />

MP2, CCSD, and a variety <strong>of</strong> DFT Hamiltonians with different<br />

basis sets were employed to investigate <strong>the</strong> stable conformations<br />

<strong>of</strong> sec-butyl. Several DFT methods only predict two<br />

equilibrium structures with backbone dihedral angles <strong>of</strong> about<br />

85° and 170°, whereas all WFT methods and some DFT<br />

methods yield an extra cis-like stable conformation with a dihedral<br />

angle <strong>of</strong> about 55°, followed by a low <strong>torsional</strong> barrier<br />

towards <strong>the</strong> gauche conformation, in agreement with existing<br />

experimental results. Inclusion <strong>of</strong> exact exchange and longrange<br />

behaviour-corrected correlation functionals is crucial<br />

to locate this elusive cis-like conformation for DFT methods.<br />

The selection <strong>of</strong> <strong>the</strong> basis sets using <strong>the</strong> B3LYP or <strong>the</strong> MP2<br />

Hamiltonian does not affect <strong>the</strong> number <strong>of</strong> stable conformations<br />

found but only modifies <strong>the</strong> shape <strong>of</strong> <strong>the</strong> PES pr<strong>of</strong>iles.<br />

Noting <strong>the</strong> stable conformation <strong>of</strong> any molecular species in<br />

reality is not simply solely determined by electronic energies,<br />

calculations including <strong>the</strong> zero-point vibrational effect were<br />

fur<strong>the</strong>r carried out. The zero-point corrected energy shows<br />

that <strong>the</strong> cis-like muoniated conformer is more stable than <strong>the</strong><br />

gauche one, in line with <strong>the</strong> available experiment data.<br />

Acknowledgement<br />

Financial support for this project was provided by grants<br />

from <strong>the</strong> Natural Sciences and Engineering Research Council<br />

<strong>of</strong> Canada (NSERC).<br />

References<br />

(1) von Ragué Schleyer, P.; Kaupp, M.; Hampel, F.; Bremer, M.;<br />

Mislow, K. J. Am. Chem. Soc. 1992, 114 (17), 6791. doi:10.<br />

1021/ja00043a026.<br />

(2) Mo, Y.; Gao, J. Acc. Chem. Res. 2007, 40 (2), 113. doi:10.<br />

1021/ar068073w.<br />

(3) Sancho-García, J. C.; Karpfen, A. Theor. Chem. Acc. 2006,<br />

115 (5), 427. doi:10.1007/s00214-006-0123-3.<br />

(4) Sancho-García, J. C.; Cornil, J. J. Chem. Theory Comput.<br />

2005, 1 (4), 581. doi:10.1021/ct0500242.<br />

(5) Tsuzuki, S.; Uchimaru, T.; Matsumura, K.; Mikami, M.;<br />

Published by NRC Research Press


Can. J. Chem. Downloaded from www.nrcresearchpress.com by University <strong>of</strong> British Columbia on 01/03/12<br />

For personal use only.<br />

1476 Can. J. Chem. Vol. 89, 2011<br />

Tanabe, K. J. Chem. Phys. 1999, 110 (6), 2858. doi:10.1063/1.<br />

477928.<br />

(6) Viruela, P. M.; Viruela, R.; Ortí, E.; Brédas, J. L. J. Am. Chem.<br />

Soc. 1997, 119 (6), 1360. doi:10.1021/ja961586l.<br />

(7) Ihee, H.; Zewail, A. H.; Goddard, W. A., III. J. Phys. Chem. A<br />

1999, 103 (33), 6638. doi:10.1021/jp990867n.<br />

(8) King, R. A.; Crawford, T. D.; Stanton, J. F.; Schaefer, H. F., III.<br />

J. Am. Chem. Soc. 1999, 121 (46), 10788. doi:10.1021/<br />

ja991429x.<br />

(9) Fleming, D. G.; Bridges, M. D.; Arseneau, D. J.; Chen, Y. K.;<br />

Wang, Y. A. J. Phys. Chem. A 2011, 115 (13), 2778. doi:10.<br />

1021/jp109676b.<br />

(10) Chen, Y. K.; Fleming, D. G.; Wang, Y. A. J. Phys. Chem. A<br />

2011, 115 (13), 2765. doi:10.1021/jp1096212.<br />

(11) Claxton, T. A. Chem. Soc. Rev. 1995, 24 (6), 437. doi:10.1039/<br />

cs9952400437.<br />

(12) Hermosilla, L.; Calle, P.; Garcia de la Vega, J. M.; Sieiro, C. J.<br />

Phys. Chem. A 2005, 109 (6), 1114. doi:10.1021/jp0466901.<br />

(13) Mielke, S. L.; Schwenke, D. W.; Peterson, K. A. J. Chem.<br />

Phys. 2005, 122 (22), 224313. doi:10.1063/1.1917838.<br />

(14) Knyazev, V. D.; Dubinsky, I. A.; Slagle, I. R.; Gutman, D. J.<br />

Phys. Chem. 1994, 98 (43), 11099. doi:10.1021/j100094a018.<br />

(15) Chen, Y.; Rauk, A.; Tschuikow-Roux, E. J. Phys. Chem. 1990,<br />

94 (16), 6250. doi:10.1021/j100379a019.<br />

(16) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;<br />

Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.;<br />

Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar,<br />

S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.;<br />

Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada,<br />

M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida,<br />

M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.;<br />

Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;<br />

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev,<br />

O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala,<br />

P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg,<br />

J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.;<br />

Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.;<br />

Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.;<br />

Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.;<br />

Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R.<br />

L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.;<br />

Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson,<br />

B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.<br />

Gaussian 03; Gaussian, Inc.: Wallingford, CT, 2004.<br />

(17) Frisch, M. J.; Pople, J. A.; Binkley, J. S. J. Chem. Phys. 1984,<br />

80 (7), 3265. doi:10.1063/1.447079.<br />

(18) Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; von Ragué<br />

Schleyer, P. J. Comput. Chem. 1983, 4 (3), 294. doi:10.1002/<br />

jcc.540040303.<br />

(19) Kendall, R. A.; Dunning, T. H., Jr; Harrison, R. J. J. Chem.<br />

Phys. 1992, 96 (9), 6796. doi:10.1063/1.462569.<br />

(20) Wong, M. W.; Radom, L. J. Phys. Chem. 1995, 99 (21), 8582.<br />

doi:10.1021/j100021a021.<br />

(21) Wong, M. W.; Radom, L. J. Phys. Chem. A 1998, 102 (12),<br />

2237. doi:10.1021/jp973427+.<br />

(22) Perdew, J. P. Unified Theory <strong>of</strong> Exchange and Correlation<br />

Beyond <strong>the</strong> Local Density Approximation. In Electronic<br />

Structure <strong>of</strong> Solids; Ziesche, P., Eschrig, H., Eds.; Akademie<br />

Verlag: Berlin, 1991; pp 11–20.<br />

(23) Burke, K.; Perdew, J. P.; Wang, Y. Derivation <strong>of</strong> a Generalized<br />

Gradient Approximation: The PW91 Density Functional. In<br />

Electronic Density Functional Theory: Recent Progress and<br />

New Directions; Dobson, J. F., Vignale, G., Das, M. P., Eds.;<br />

Plenum: New York, 1997; pp 81–121.<br />

(24) Becke, A. D. J. Chem. Phys. 1996, 104 (3), 1040. doi:10.1063/<br />

1.470829.<br />

(25) Becke, A. D. J. Chem. Phys. 1993, 98 (2), 1372. doi:10.1063/<br />

1.464304.<br />

(26) Mayer, P. M.; Parkinson, C. J.; Smith, D. M.; Radom, L. J.<br />

Chem. Phys. 1998, 108 (2), 604. doi:10.1063/1.476256.<br />

(27) Vrček, V.; Kronja, O.; Saunders, M. J. Chem. Theory Comput.<br />

2007, 3 (3), 1223. doi:10.1021/ct600308b.<br />

Published by NRC Research Press

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!