You are on page 1of 12

Article

pubs.acs.org/JPCA

Benchmarking Compound Methods (CBS-QB3, CBS-APNO, G3, G4,


W1BD) against the Active Thermochemical Tables: A Litmus Test for
Cost-Effective Molecular Formation Enthalpies
John M. Simmie* and Kieran P. Somers
Combustion Chemistry Centre & School of Chemistry, National University of Ireland, Galway, Ireland
*
S Supporting Information

ABSTRACT: The theoretical atomization energies of some 45 CxHyOz molecules


present in the Active Thermochemical Tables compilation and of particular interest to
the combustion chemistry community have been computed using five composite model
chemistries as titled. The species contain between 1−8 “heavy” atoms, and a few are
conformationally diverse with up to nine conformers. The enthalpies of formation at 0
and 298.15 K are then derived via the atomization method and compared against the
recommended values. In general, there is very good agreement between our averaged
computed values and those in the ATcT; those for 1,3-cyclopentadiene exceptionally
differ considerably, and we show from isodesmic reactions that the true value for 1,3-
cyclopentadiene is closer to 134 kJ mol−1 than the reported 101 kJ mol−1. If one is
restricted to using a single method, statistical measures indicate that the best methods are
in the rank order G3 ≈ G4 > W1BD > CBS-APNO > CBS-QB3. The CBS-x methods do
on average predict ΔfH⊖(298.15 K) within ≈5 kJ mol−1 but are prone to occasional
lapses. There are statistical advantages to be gained from using a number of methods in tandem, and all possible combinations
have been tested. We find that the average formation enthalpy coming from using CBS-APNO/G4, CBS-APNO/G3, and G3/G4
show lower mean signed and mean unsigned errors, and lower standard and root-mean-squared deviations, than any of these
methods in isolation. Combining these methods also leads to the added benefit of providing an uncertainty rooted in the
chemical species under investigation. In general, CBS-APNO and W1BD tend to underestimate the formation enthalpies of
target species, whereas CBS-QB3, G3, and G4 have a tendency to overestimate the same. Thus, combining CBS-APNO with a
G3/G4 combination leads to an improvement in all statistical measures of accuracy and precision, predicting the ATcT values to
within 0.14 ± 4.21 kJ mol−1, thus rivalling “chemical accuracy” (±4.184 kJ mol−1) without the excessive cost associated with
higher-level methods such as W1BD.

■ INTRODUCTION
The determination of accurate formation enthalpies (ΔfH⊖) of
The accuracy of any ΔfH⊖ derived from quantum chemistry
depends largely on the model chemistry applied and the size of
stable, radical, and ionic species is central to the science of the molecular system, with the theoretical thermodynamicist in
combustion modeling, serving as essential input information for constant pursuit of “chemical accuracy”  that is formation
enthalpies computed within 1 kcal mol−1 (4.184 kJ mol−1) of
a multitude of numerical simulations. Such is the importance of
the true formation enthalpy. The current “gold-standard” in
this fundamental property that great effort has been invested in
model chemistries is the coupled-cluster method with
their tabulation and refinement, with the JANAF,1 CODATA,2
perturbative quadruples, CCSDT(Q), with subsequent extrap-
and Third Millenium databases,3 and the Active Thermochem-
olation of the computed energies to the basis-set limit; see for
ical Tables (ATcT),4−6 serving the combustion community well
example, work on the C2H5 + O2 reaction.8 This is rarely
for the last number of decades.
achievable, and more practical standards are the CCSD(T)
Although experimental determination (primarily calorimetry)
methods9 with perturbative inclusion of triples. The Wn,10−12
is the ideal method to obtain accurate ΔfH⊖, it is both laborious
high-accuracy extrapolated ab initio thermochemistry
and time-consuming. Add to this the number of species which
(HEAT),13−15 focal-point extrapolation,16 and Feller-Peter-
may comprise a chemical kinetic mechanism7 (100s−1000s),
many of which are radicals which may be experimentally
inaccessible, it is clear that obtaining experimental ΔfH⊖ for all Special Issue: 100 Years of Combustion Kinetics at Argonne: A
species of interest in combustion modeling is effectively Festschrift for Lawrence B. Harding, Joe V. Michael, and Albert F.
impossible. However, it is now accepted that quantum Wagner
chemistry, either in isolation or in tandem with experiment, Received: November 14, 2014
can offer an effective means with which to obtain such Revised: January 5, 2015
information on feasible timescales. Published: January 12, 2015

© 2015 American Chemical Society 7235 DOI: 10.1021/jp511403a


J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

son-Dixon17 methods have also shown accuracy to within ≈2 kJ methodology behind ATcT and discussion of uncertainty
mol−1 when compared with well-known experimental for- quantification in general has been recently given by Ruscic.35
mation enthalpies. However, despite advances in computer The isodesmic reaction method is not employed in this work
software and hardware, these methods remain computationally as part of our benchmarking, owing to the problems discussed
expensive and are somewhat restricted in their application to above, rather we use the atomization method alone to derive
molecules containing some five or so nonhydrogen atoms and, ΔfH⊖ throughout. Thus, the formation enthalpies derived
as a consequence, can rarely treat the full panoply of molecules herein are representative of the “pure” method, and are
which are to be encountered in detailed chemical kinetic independent of experimental formation enthalpies required for
models.7 isodesmic reactions. The results aim to provide an assessment
Compound methods offer a cost-effective alternative to the of the performance of each method against a wide-range of
above, albeit at the expense of accuracy. These methods consist stable and, in the future, radical species.
of a series of less expensive geometry optimization, frequency, Recently, Simmie et al.36 carried out a thermochemical study
and single-point energy calculations with empirical corrections of furan derivatives, and they found that a ΔfH⊖ computed by
used to overcome deficiencies when the methods are taking an average ΔfH⊖ from CBS-QB3, CBS-APNO, and G3
benchmarked against formation enthalpies of well-known atomization energies recreated those computed from well-
atomic and small-molecule standards. Numerous compound framed isodesmic working reactions to within ≈4 kJ mol−1,
methods have entered widespread use within the combustion despite the large uncertainties associated with the atomization
community in recent times; in particular, the CBS-x18,19 (x = calculations. The reasons for this interesting behavior were not
QB3 and APNO) and Gaussian-x20−24 (x = 1, 2, 3, and 4) investigated at the time, but Somers37 later ascribed it to a
methods are regularly employed in thermodynamic and kinetics
tendency of the G3 method to consistently overpredict ΔfH⊖,
computations of relevance to combustion and atmospheric
for the CBS-APNO method to consistently underpredict ΔfH⊖,
chemistry.
and for the CBS-QB3 method to arrive at a similar answer to
Variants of the G3 method, such as G3B325 and G3MP226
the isodesmic working reactions.
are also popular, with the former frequently employed for ΔfH⊖
computations by Burcat et al. as part of the Third Millennium A second goal of this work is therefore to assess whether a
Database, with an uncertainty of ±8 kJ mol−1 typically quoted combination of compound methods can provide a more
therein when they have applied this method. rigorous assessment of ΔfH⊖ than any of these methods in
“Standard” methods have been tweaked on numerous isolation, thus offering a cost-effective alternative to high-level
occasions to remedy perceived difficulties and/or with the methods. A primary reason for using two or more model
aim of improving the computational cost; for example, variants chemistries, quite apart from the statistical gain, is that the final
of G4,27,28 of W129,30 and W3,31,32 etc. It appears that divergent result is then less-dependent upon the actual geometrical
evolution is the order of the day. structure if the optimization and frequency calculations, inbuilt
In order to overcome the large uncertainties in ΔfH⊖ when into each composite method, differ. Of course if the two
compound methods are employed, the isodesmic reaction optimizations result in distinctly different geometries and/or
method33 is frequently employed to reduce errors and electronic states then this approach would be compromised, but
uncertainties in the computation. However, a number of in our experience this has rarely happened for the species under
problems arise when employing this method. Well-known consideration here.


formation enthalpies are required from experiment or high-level
theory to create these hypothetical working reactions; the COMPUTATIONAL METHODOLOGY
computed ΔfH⊖ is very much dependent on the quality of the
isodesmic reaction framed (given that there is no unique Five compound methods have been chosen for benchmarking
answer), and the computational cost is increased as numerous as part of this work: CBS-QB3, CBS-APNO, G3, G4, and
quantum chemical calculations must be carried out to W1BD,38 as embedded within the application Gaussian.39
determine the theoretical enthalpy of all the chaperone For the computation of 0 K enthalpies of formation, ΔfH0,
molecules. However, provided good working reactions can be we commence by calculating the theoretical atomization
framed the method can deliver excellent results; as an example, energy, TAE0, for the reaction:
see previous work on cyclic esters.34
The empirical corrections employed in compound methods Cx HyOz → x 3C + y 2 H + z 3O
also reduce their predictive capability, as the databases against
which they are benchmarked, and any empirical corrections which is given by
thus derived tend to consist of a limited number of small
molecule targets. Molecules of interest to combustion scientists TAE0 = xH0(3C) + yH0(2 H) + zH0(3O) − H0(Cx HyOz )
can be large, conformationally complex, species extending well (1)
outside the benchmarked test-set in terms of size and the true
uncertainty in any ΔfH⊖ computed from these methods is where H0 is the 0 K enthalpy of an atom or molecule and is the
somewhat unknown. sum of the electronic and zero-point energies. Zero-point
Thus, we arrive at the goals of this work: to benchmark a energies are automatically computed, adjusted by a built-in
range of popular compound methods against a database of well- scale factor and added to the 0 K electronic energy by each
known formation enthalpies, namely, the Active Thermochem- compound method as part of its predefined series of
ical Tables of Ruscic.6 This database has been chosen as it is computations. The ΔfH0 of the molecule then follows knowing
based on either direct experiment or experiment in the theoretical atomization energy and the experimentally
combination with high-level theory and is undergoing constant known formation enthalpies of the component atoms in their
development and refinement. A brief description of the gaseous state:
7236 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Δf H0(Cx HyOz ) = [xΔf H0(3C) + yΔf H0(2 H) 298.15 K. We display in Table 3, the results of computed
formation enthalpies for all five model chemistries in
+ zΔf H0(3O)] − TAE0 (2) comparison to ATcT values.
Problem Molecules. Some molecules in the ATcT

For formation enthalpies at 298.15 K, ΔfH , the same compendium are problematic in the sense that the composite
equations are employed, with thermal enthalpy corrections methods under discussion here are unable to attain a
added to H0 in eq 1 to account for translational, rotational, and satisfactory solution to their structure. Thus, for example,
vibrational degrees of freedom of each molecule and atom. oxirene, a three-membered heterocyclic ring containing a CC
Standard state formation enthalpies of each gaseous atom, bond resists optimization (or rather optimizes to a structure
ΔfH⊖, also replace the 0 K counterparts in eq 2. with an imaginary frequency) with the B3LYP functional, which
Note that the influence of vibrational anharmonicities, is of course an essential component of CBS-QB3, G4, and
rotational nonrigidities, and torsional degrees of freedom are W1BD. The basis sets used by these three methods are
not accounted for in the predefined routine of calculations in predefined at 6-311G(2d,d,p), 6-31G(2df,p), and cc-pVTZ+d;
any of the compound methods employed, and therefore, in the note that using a 6-311++G(d,p) basis set leads to a successful
final formation enthalpies reported herein. The 0 and 298.15 K convergence but just not to a C2v structure.
formation enthalpies of atomic species used in the above This molecule has a long and checkered history; in a telling
equations are given in Table 1 and are adopted from the phrase it has been described as “hovering on the edge of
ATcT.6 reality”, or it has been subtitled with a Shakespearean quotation
“to be or not to be”, with some theoretical methods concluding
Table 1. Atomic Formation Enthalpies (kJ mol−1) that it is a saddle point and others a local minimum on the
potential energy surface.42,43 Recent extremely expensive
T (K) C (3P) H (2S1/2) O (3P2)
coupled-cluster interference-corrected explicitly correlated
0 711.38 216.034 246.844 second-order perturbation theory results44 ascribe a total
298.15 716.87 217.998 249.229 atomization energy of 1827.8 kJ mol−1 at 0 K, which implies
an enthalpy of formation of 277.0 kJ mol−1, whereas equally
In those cases where conformational diversity exists the expensive W4 calculations45 give 272.1 and 268.2 kJ mol−1 at 0
Gibbs free energies, ΔG⊖ (298.15 K) were determined for each and 298.15 K, respectively. Viewed from this perspective, the
conformer and the Boltzmann distribution computed, taking CBS-APNO and G3 0 K results of 274.5 and 275.4 kJ mol−1,
due account of degeneracies, σ (eq 3). The contribution, xi, respectively, are surprisingly good.
made by each conformer to the overall enthalpy of formation We should also take a moment to note if there are any
for species X is then calculated from considerable differences between any of our computations and
those values recommended in the ATcT. The most obvious
n
example is 1,3-cyclopentadiene where the ATcT value of
xi = σi exp( −ΔGm⊖(i)/RT )/∑ [σi exp( −ΔGm⊖(i)/RT )] ΔfH⊖(298.15K) = 101.30 ± 2.5 differs considerably from our
i=1
result of 137.1 ± 6.7 kJ mol−1. In the compendium
(3)
Thermochemical Data of Organic Compounds, Pedley et al.
n
quote46 a value of 134.3 ± 1.5 kJ mol−1, whereas the NIST
Δf Hm⊖(X) = ∑ [xiΔf Hm⊖(i)] WebBook includes determinations47,48 of 139 and 133.4 kJ
i=1 (4) mol−1. In a thermochemical and kinetic analysis of addition
Thus, in the case of n-hexane nine conformers (out of a reactions to 1,3-cyclopentadiene Zhong and Bozzelli49 use a
possible 12 as shown schematically by Morini et al.)40 with value of 131 kJ mol−1. More recent photoacoustic spectroscopy
accompanying degeneracies or symmetry numbers were and quantum chemical studies by Agapito et al.50 imply a heat
identified: ttt(1), gtt(4), tgt(2), tgg(4), gtg(2), g+t+g−(2), of formation of 134 kJ mol−1.
ggg(2), g+x−t+(4), and t+g+x−(4). These give rise to populations From a consideration of isodesmic reactions involving (i)
of 17.1, 45.0, 11.8, 13.6, 3.2, 5.9, 1.2, 1.4, and 0.8%, respectively, furan, propane and dimethyl ether, (ii) ethyne, 2 molecules of
from G4 calculations. ethene and methane as chaperones, it can be shown that the
Each individual enthalpy of formation was then weighted computed reaction enthalpies of −90.30 ± 1.58 and 275.53 ±
with respect to the population to calculate a final value for that 4.71 kJ mol−1 lead to values for ΔfH (298.15 K) of 135.21 ±
particular molecule (eq 4). In some cases, individual con- 1.80 and 132.47 ± 4.76 kJ mol−1, respectively, from which a
formers are present in the ATcT such as syn- and anti-formic grand-weighted average of 134.9 ± 1.7 kJ mol−1 results. Thus, it
acid, and in these cases, no distributions are needed. appears that the ATcT value for 1,3-cyclopentadiene is
In general, the literature was consulted for possible incorrect.
conformers of a particular molecule, but in any event, these There are similar, but less severe, disparities for oxalic acid.
were then checked by using the application Spartan and We note that an isodesmic reaction:
utilizing a modest basis set and B3LYP functional to generate (COOH)2 + C2H6 = 2CH3COOH
the distributions.41 Only those which contributed to a
significant extent were then retained for further computation. can be framed whose average reaction enthalpy is −43.86 ±

■ RESULTS AND DISCUSSION


The results at 0 K are shown in Table 2; although interesting,
0.96 kJ mol−1, from which the predicted formation enthalpy of
the dominant aa conformer of oxalic acid is −739.8 ± 1.2 kJ
mol−1 where we have adopted the ATcT numbers for the
the primary focus of our attention will be on the results at chaperones ethane and cis acetic acid. This is in very good
298.15 K, which are much more important for chemical kinetic agreement with the atomization value of −735.5 ± 5.3 kJ mol−1
simulations and modeling purposes. for the same conformer, thus indicating that the ATcT value for
7237 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 2. Final Computed Formation Enthalpies of Compounds at 0 K/kJ mol−1


species QB3 APNO G3 G4 W1BD ATcT
1,3-butadiyne 468.41 462.06 462.03 458.61 458.08 458.39
1,3-cyclopentadiene 161.39 148.19 155.65 153.43 144.16 118.20
1-butene (gauche) 28.49 15.49 22.23 23.05 15.39 20.30
1-butene (cis) 29.32 16.30 23.07 24.11 15.85 20.93
1-butyne 188.37 178.03 181.67 181.33 176.13 178.97
2,2,4-trimethylpentane −147.58 −178.69 −159.19 −156.69 −168.22
gauche −160.09 −191.51 −171.70 −169.26 −181.87 −171.20
2,3-butanedione −315.19 −319.76 −313.36 −311.53 −318.11 −310.25
2-butyne 165.90 155.80 160.26 160.51 154.99 158.72
2-propanol (gauche) −249.47 −257.53 −249.86 −247.65 −257.09
2-propanol (trans) −247.99 −256.49 −248.78 −246.28 −255.95
acetic acid (cis) −422.67 −422.63 −418.61 −416.90 −423.43 −419.59
acetone −200.49 −−207.47 −200.81 −199.61 −206.14 −199.39
allene 202.25 196.48 196.62 197.07 194.26 197.63
benzene 107.21 91.80 103.83 102.26 92.65 100.70
carbon monoxide −115.57 −113.78 −114.71 −117.59 −114.07 −113.81
cis-2-butene 21.79 8.40 16.17 16.91 8.30 13.94
cyclobutene 186.71 175.51 183.12 180.29 172.02 173.90
cyclohexane −74.02 −97.33 −82.55 −80.66 −93.44 −82.52
cyclopropane 76.69 65.64 73.82 71.92 65.95 70.76
cyclopropene 298.87 292.28 295.82 293.58 288.91 292.64
dimethyl ether −171.42 −173.81 −167.66 −166.80 −171.77 −166.51
dioxirane 0.42 5.71 9.70 6.86 8.66 9.00
ethanal −156.89 −158.95 −155.71 −155.49 −158.85 −154.98
ethane −66.45 −74.99 −69.15 −67.35 −73.12 −68.13
ethene 64.11 59.75 60.55 60.84 58.11 61.08
ethylene glycol no. 1 −375.41 −374.64 −370.04 −368.17 −378.03
ethylene glycol no. 2 −373.05 −372.87 −367.84 −365.49 −376.18
ethylene glycol no. 3 −372.82 −376.46 −366.17 −364.65 −375.53
ethyne 234.82 233.54 230.90 229.18 228.61 228.89
ethynol 95.86 97.37 95.93 95.82 92.19 94.80
formaldehyde −110.24 −107.06 −107.15 −108.04 −107.36 −105.33
formic acid (syn) −377.05 −372.59 −371.39 −371.35 −374.83 −371.62
formic acid (anti) −360.12 −356.14 −354.84 −354.93 −358.01 −355.28
glyoxal (trans) −214.26 −209.65 −209.62 −210.61 −210.77 −206.85
glyoxal (cis) −194.85 −190.70 −190.25 −192.15 −192.13 −188.42
isobutane −100.31 −116.91 −105.72 −103.79 −113.44 −106.76
isobutene 11.26 −1.83 5.36 5.60 −1.81 3.46
ketene −45.49 −47.19 −47.14 −−45.51 −47.47 −45.47
methane −66.31 −70.91 −67.78 −66.61 −68.60 −66.56
methanediol −385.18 −381.72 −378.43 −378.13 −385.35 −379.21
methanol −193.15 −193.46 −190.06 −189.87 −194.75 −189.82
methyl formate (cis Z) −354.08 −351.59 −348.58 −347.45 −350.69 −344.54
methyl hydroperoxide −123.01 −117.57 −112.96 −114.09 −117.97 −114.99
n-butane (trans) −92.35 −109.03 −97.89 −95.63 −106.21 −98.54
n-butane (gauche) −89.57 −106.14 −95.06 −92.69 −103.17 −98.54
n-hexane ttt −119.90 −144.30 −128.06 −125.53 −140.24
gtt −117.02 −141.50 −125.39 −122.61 −137.14
tgt −117.10 −141.41 −125.38 −122.65 −137.18
tgg −114.34 −139.78 −123.75 −120.02 −134.17
gtg −114.07 −138.68 −122.71 −119.69 −134.04
g+t+g− −113.52 −138.25 −122.20 −119.07 −133.46
ggg −112.14 −138.09 −122.10 −117.97 −131.84
g+x−t+ −107.68 −132.51 −116.30 −113.51 −127.90
t+g+x− −107.19 −132.15 −116.18 −112.80 −127.20
o-benzyne 477.32 463.97 470.99 467.35 464.34 469.60
oxalic acid (c/c) −716.78 −710.04 −707.19 −704.49 −711.63
oxalic acid (t/t) −734.03 −726.40 −724.48 −719.94 −728.90
oxalic acid (c/t) −723.06 −715.92 −713.53 −709.83 −717.78
oxirane (1A1) −43.62 −44.31 −39.64 −41.35 −44.17 −40.17
oxirene (singlet) 270.38 274.54 275.44 269.63 270.16 276.60

7238 DOI: 10.1021/jp511403a


J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 2. continued
species QB3 APNO G3 G4 W1BD ATcT
propane −78.96 −91.57 −82.99 −80.87 −89.26 −82.13
propene 40.32 31.34 35.52 35.95 30.35 35.36
propyne 198.64 191.56 193.73 192.74 189.92 192.86
succinic acid (aAa) −806.20 −805.87 −799.30 −793.72 −806.69
sGs −810.18 −809.03 −802.93 −798.32 −808.41
sGa −804.18 −802.68 −796.99 −792.09 −802.00
gAs −801.13 −800.04 −793.68 −789.33 −800.58
toluene 80.45 60.99 74.99 73.56 63.80 73.69
trans-2-butene 16.56 3.38 10.73 11.40 3.36 9.39
vinyl alcohol (anti) −108.62 −110.86 −109.00 −108.14 −113.89
vinyl alcohol (syn) −112.65 −114.81 −112.72 −111.93 −118.27 −112.45

oxalic acid of −721.4 ± 2.1 is probably somewhat on the high The conformer distribution outlined in Table 4 are
side. consistent except for the W1BD results which give less
In a similar manner, an isodesmic reaction targeting prominence to the sGs conformer than to the aAa. These
cyclobutene can be constructed with ethane, cyclobutane, and were among the most challenging calculations undertaken
ethene as chaperons. This reaction is, not unexpectedly, very (Table 5), and perhaps a relaxation, in the case of the sGs
nearly thermoneutral at 1.74 ± 1.16 kJ mol−1 and based on a conformer, of the convergence-related options for the Berny
formation enthalpy for cyclobutane,46 which is absent from the algorithm is responsible for this anomaly.
ATcT, of 28.4 ± 0.6 kJ mol−1; this results in ΔfH⊖(298.15 K) = Computational Cost. A crude example of the computa-
163.0 ± 1.3 kJ mol−1, which is somewhat higher than the ATcT tional cost in terms of job times is shown in Table 5 from which
value of 156.9 ± 1.6 but in excellent agreement with our it can be seen that G3 emerges as a relatively cheap yet, to be
computed average of 163.9 ± 5.9 kJ mol−1. described below, effective method.
Statistical analyses in subsequent sections therefore do not Comparison of Methods with ATcT. We benchmark each
include the molecules 1,3-cyclopentadiene and oxalic acid but theoretical method against the formation enthalpies recom-
cyclobutene is included in our later calculations. mended in the ATcT via descriptive statistics, and the results
Conformers. Some of the values in ATcT v1.110 are are shown in Tables 6 and 7. In order to assess how well a
composites, for example 1-butene is quoted at −0.03 ± 0.48 kJ combination of different compound methods may perform,
mol−1, but in reality this molecule is comprised of two unweighted average formation enthalpies for various combina-
conformers, namely cis and gauche, with CCCC dihedral angles tions of the CBS-QB3, CBS-APNO, G3, and G4 methods have
of ≈0° and ≈120°, respectively. The degenerate gauche also been computed. Note that combinations involving the
conformer is dominant at 78.5% ± 1.5%, but the differences W1BD method have generally been excluded, owing to its
in enthalpy are slight, so the overall enthalpy of formation at expense, but it is included in a combination of “all methods” as
298.15 K is +0.78 ± 5.59, in very good agreement with the shown in subsequent figures and tables.
listed ATcT value of −0.03 ± 0.48 kJ mol−1. The results for 1- For each method, or combination of methods, the difference
butene and other systems are tabulated in Table 4. between the ATcT recommendation and our computed
In general, most of the conformer distributions are consistent formation enthalpy for species, i, is denoted as Φi = ΔfH⊖ ATcT
across all five composite methods, but the treatment of those − ΔfH⊖ i , and the mean signed errors (MSE) and mean
compounds whose geometries are very sensitive to slight unsigned errors (MUE) are computed as follows where n is the
changes in dihedral angles, such as the gylcols and dibasic acids, number of species in the benchmark set:
can be much less successful.
For oxalic acid, (COOH)2, three conformers prevail with the MSE = ∑ (Φi)/n (5)
OC−CO always anti and the H−OC−C angles either
anti or syn (Figure 1). MUE = ∑ |Φ|i /n (6)
In the case of the higher acid succinic, (CH2COOH)2, 4
principal conformers were identified in which three dihedral In terms of uncertainties, the maximum absolute deviation
angles differ. These are classified in the order shown, Figure 2, (|Φmax|) from the ATcT is determined from eq 6 and is
as syn s, anti a, or gauche g, with the middle dihedral essentially an indicator of “worst-case scenario” performance for
capitalized. Thus, the all trans configuration is denoted as aAa, each method; it is rarely used in thermochemical work as a true
etc. These lowest energy conformers all exhibit syn H−O−C indicator of statistical accuracy or precision. In line with the
O dihedrals. In this, we follow the notation and conclusions of a recommendations of Ruscic, uncertainties are considered for
recent study by Vogt et al.51 95% confidence intervals via computation of the population
The optimization procedures inbuilt in CBS-QB3 and G4 standard deviation (σMSE) and root-mean-square deviations
converge to a structure with an imaginary frequency if a C2h (RMSD):
symmetry is imposed; this is a problem with the functional and
basis sets used for these two cases. The incorporation of one
(or more) diffuse terms in the 6-311G(2d,d,p) basis set, which
σMSE = 2 × ∑ (Φi − MSE)2 /n (7)
is the default option in the model chemistry CBS-QB3, so that
it now becomes CBS-QB3(+)52, does yield a nonimaginary
solution with C2h symmetry.
RMSD = 2 × ∑ (Φi)2 /n (8)

7239 DOI: 10.1021/jp511403a


J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 3. Final Computed Formation Enthalpies of Compounds at 298.15 K/kJ mol−1


species QB3 APNO G3 G4 W1BD ATcT ±
1,3-butadiyne 471.07 464.24 463.89 461.47 460.74 460.11 0.87
1,3-cyclopentadiene 145.90 132.59 140.30 138.00 128.72 101.30 2.50
1-butene 8.28 −4.76 2.24 2.96 −4.83 −0.03 0.48
1-butyne 175.81 165.16 169.07 168.97 163.64 165.39 0.85
2,2,4-trimethylpentane −210.74 −242.31 −221.92 −219.66 −232.28 −223.70 1.50
2,3-butanedione −329.50 −334.43 −327.74 −325.78 −332.37 −326.81 0.98
2-butyne 155.35 144.95 149.65 150.01 144.48 145.76 0.79
2-propanol −272.13 −280.26 −272.30 −270.09 −279.53 −272.81 0.37
acetic acid −434.23 −434.25 −430.05 −428.31 −434.87 −433.71 0.49
acetone −214.95 −224.33 −215.26 −214.24 −220.67 −216.09 0.37
allene 195.66 189.77 190.04 190.47 187.70 190.15 0.37
benzene 91.65 76.16 88.47 86.81 77.14 83.18 0.26
carbon monoxide −111.44 −109.62 −110.56 −113.43 −109.91 −110.53 0.03
cis-2-butene 2.10 −11.22 −3.24 −2.78 −11.32 −7.33 0.53
cyclobutene 171.05 159.78 167.55 164.67 156.42 156.90 1.60
cyclohexane −111.55 −134.71 −119.53 −117.87 −130.86 −122.08 0.68
cyclopropane 60.53 49.45 57.73 55.82 49.84 53.61 0.53
cyclopropene 291.06 284.32 287.98 285.72 281.14 283.91 0.59
dimethyl ether −188.22 −190.50 −184.27 −183.56 −188.43 −184.02 0.44
dioxirane −5.95 −0.72 3.30 0.55 2.36 1.30 1.20
ethanal −166.36 −168.38 −165.00 −164.90 −168.26 −165.46 0.32
ethane −81.61 −90.15 −84.21 −82.48 −88.23 −83.79 0.17
ethene 56.24 51.84 52.70 53.00 50.26 52.56 0.15
ethylene glycol −393.53 −394.68 −386.99 −384.86 −394.21 −389.42 0.49
ethyne 234.81 233.21 230.69 229.44 228.62 228.33 0.15
ethynol 94.53 95.57 94.34 94.75 90.96 92.70 1.40
formaldehyde −113.24 −110.05 −110.13 −111.01 −110.33 −109.17 0.11
formic acid −383.10 −378.60 −377.30 −377.30 −380.80 −378.94 0.27
glyoxal −218.25 −213.70 −213.57 −214.51 −214.65 −212.48 0.59
isobutane −127.920 −144.48 −133.05 −131.24 −140.92 −135.36 0.40
isobutene −8.92 −22.01 −14.57 −14.56 −21.95 −17.60 0.53
ketene singlet −47.50 −49.37 −49.19 −47.46 −49.38 −48.58 0.15
methane −73.95 −78.54 −75.41 −74.22 −76.21 −74.53 0.06
methanediol (sc,sc) −398.12 −394.59 −391.18 −390.87 −398.03 −392.61 0.96
methanol −203.44 −203.71 −200.26 −200.09 −204.88 −200.71 0.18
methyl formate (cis Z) −365.61 −363.24 −360.10 −358.88 −362.14 −357.80 0.59
methyl hydroperoxide −134.33 −129.09 −124.40 −125.60 −129.37 −127.73 0.91
n-butane −118.72 −135.75 −124.35 −121.86 −132.42 −125.85 0.38
n-hexane −155.63 −180.08 −163.49 −161.05 −175.67 −166.94 0.48
o-benzyne 470.39 457.41 464.76 460.69 457.45 460.70 1.40
oxalic acid −742.52 −734.26 −732.26 −728.21 −737.19 −721.40 2.10
oxirane (1A1) −55.01 −55.77 −51.02 −52.70 −55.51 −52.72 0.44
oxirene (singlet) 267.82 272.45 273.50 267.18 267.60 275.90 3.10
propane −100.52 −113.11 −104.35 −102.36 −110.74 −104.41 0.29
propene 26.05 17.07 21.40 21.71 16.12 20.35 0.33
propyne 192.48 185.07 187.42 186.75 183.79 185.80 0.38
succinic acid −829.62 −827.47 −820.98 −816.80 −826.02 −817.75 0.61
toluene 60.59 41.08 55.48 53.85 44.02 50.41 0.37
trans-2-butene −3.18 −16.34 −8.74 −8.31 −16.33 −11.18 0.51
vinyl alcohol −121.93 −123.86 −121.56 −121.16 −127.54 −123.76 0.91

Note first the 2-fold multiplier in the above formulas which We will first consider the performance of individual
provides us with 95% confidence, and second that there is a compound methods against the ATcT benchmark. As a general
subtle difference in our two measures of dispersion, σMSE and trend, the G3, G4, and CBS-QB3 methods tend to overpredict
the RMSD. The former assumes the average deviation from the the formation enthalpies of the compounds studied, thus their
ATcT (MSE) is the central value for the distribution, whereas negative MSE. Conversely, the W1BD and CBS-APNO
the latter assumes the central line is zero deviation from the methods underpredict ΔfH⊖ leading to a positive MSE. The
ATcT. Both measures ultimately highlight the same trends in computed MSEs and MUEs imply the following trend with
the performance of each method. In the text, uncertainties respect to accuracy: G3 ≈ G4 > W1BD > CBS-APNO > CBS-
typically refer to the MSE ± σMSE. QB3 (Figure 3). These results are somewhat surprising, with
7240 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 4. Composition-Averaged Enthalpies of Formation at 298.15 K/kJ mol−1


composition QB3 APNO G3 G4 W1BD ATcT
1-Butene
gauche 0.788 0.781 0.786 0.805 0.763
cis 0.212 0.219 0.214 0.195 0.237
final 8.28 −4.76 2.24 2.96 −4.83 −0.03 ± 0.48
2,2,4-Trimethylpentane
gauche 0.9981 0.9981 0.9978 0.9980 0.9980
transa 0.0019 0.0019 0.0022 0.0020 0.0022
final −210.74 −242.31 −221.90 −219.66 −232.28 −223.7 ± 1.5
2-Propanol
gauche 0.787 0.755 0.755 0.776 0.755
trans 0.213 0.245 0.246 0.224 0.245
final −272.13 −280.26 −272.30 −270.09 −279.53 −272.81 ± 0.37
Ethylene Glycol
no. 1 0.609 0.316 0.547 0.492 0.298
no. 2 0.206 0.144 0.211 0.144 0.127
no. 3 0.185 0.540 0.242 0.364 0.575
final −393.53 −394.68 −386.99 −384.86 −394.21 −389.42 ± 0.49
n-Butane
trans 0.593 0.756 0.750 0.614 0.625
gauche 0.407 0.244 0.250 0.386 0.375
final −118.72 −135.75 −124.35 −121.86 −132.42 −125.85 ± 0.38
n-Hexane
ttt 0.174 0.161 0.146 0.171 0.191
gtt 0.460 0.423 0.423 0.450 0.462
tgt 0.122 0.102 0.105 0.118 0.120
tgg 0.124 0.179 0.185 0.136 0.111
gtg 0.031 0.033 0.035 0.032 0.030
g+t+g- 0.059 0.066 0.068 0.059 0.055
ggg 0.010 0.018 0.019 0.012 0.009
g+x-t+ 0.012 0.010 0.010 0.014 0.014
t+g+x- 0.008 0.009 0.009 0.008 0.008
final −155.63 −180.08 −163.49 −161.05 −175.67 −166.94 ± 0.48
Oxalic Acid
syn/syn 0.003 0.007 0.001 0.007 0.005
anti/anti 0.966 0.919 0.936 0.950 0.966
syn/anti 0.031 0.074 0.063 0.044 0.030
final −742.52 −734.26 −732.26 −728.21 −737.19 −721.4 ± 2.1
Succinic Acid
aAa 0.166 0.128 0.108 0.136 0.480
sGs 0.733 0.765 0.778 0.782 0.433
sGa 0.059 0.056 0.065 0.042 0.033
gAs 0.042 0.052 0.049 0.040 0.053
final −829.62 −827.47 −820.98 −816.80 −826.02 −817.8 ± 0.61
Vinyl Alcohol
anti 0.195 0.229 0.255 0.209 0.169
syn 0.805 0.772 0.745 0.791 0.832
final −121.93 −123.86 −121.56 −121.16 −127.54 −123.76 ± 0.91
a
One imaginary frequency of 26 cm−1.

Figure 2. Key dihedrals of succinic acid.

Figure 1. Conformers of oxalic acid.


both Gx methods outperforming the considerably more
expensive W1BD.
7241 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 5. Crude Timings (Minutes) of Calculations Steps for species, this result manifesting itself as a positive MSE. As a
sGs Succinic Acid; 8 Heavy Atoms, with Total result, combining CBS-APNO with either G3/G4, CBS-QB3/
Computational Time G3, or CBS-QB3/G4 leads to improvements in the computed
MSEs when the results are compared with individual methods.
no. step QB3 APNO G3 G4 W1BD
In fact, adding the CBS-APNO to the G3/G4 combination
1 Opt 19 7 3 81 142 leads to improvements in all statistical indicators of accuracy
2 Freq 17 4 1 45 86 and precision when compared to a G3/G4 combination. As the
3 Opt − 1579 9 − − CBS-QB3, G3, and G4 methods all tend to overpredict the
4 SPE 127 1907 73 79 435 formation enthalpies of the ATcT (negative MSE), combining
5 SPE 6 158 53 53 14760 these methods offers no improvement on a G3/G4
6 SPE 24 143 362 359 56340 combination.
7 SPE − 317 193 277 23280 Also shown in Table 7 are results from a combination of all
8 SPE − − − 355 39240 methods, both including and excluding W1BD computations.
9 SPE − − − 3480 − Adding the CBS-QB3 method to a CBS-APNO/G3/G4
Total Simulation Times combination offers nothing in terms of reducing absolute
hours 3.2 68.6 11.6 78.8 2238 deviations and uncertainties. In turn, incorporating our W1BD
relative 1.0 21.3 3.6 24.5 695.8 results into the average reduces the MUE, RMSD, and standard
deviation, but only marginally, and perhaps not significantly
Rayne and Forrest have shown in a study of a number of enough to warrant the added expense.
compounds, not just of type CxHyOz but also containing Surprisingly, the W1BD method shows poorer performance
nitrogen, sulfur, fluorine, and chlorine, with up to 4 “heavy” than many so-called “cheaper” methods against this test-set of
atoms that W1BD ΔfH⊖(298.15 K) are systematically lower formation enthalpies. On the basis of the present results, the
than comparable G4 ones.53 Some of these results have been following methods, or combination of methods, could be
questioned,54,55 but for all the 18 species in common between argued to offer more accurate and precise results than W1BD at
our work and Rayne’s and Forrest’s, there are only very small significant cost savings: G3, G4, G3/G4, CBS-APNO/G3,
differences in the reported formation enthalpies. CBS-APNO/G4, CBS-APNO/G3/G4, CBS-QB3/CBS-
The CBS-APNO and CBS-QB3 methods tend to perform APNO/G4, CBS-QB3/CBS-APNO/G4, CBS-QB3/CBS-
worst of all, and while on average they tend to predict the ΔfH⊖ APNO/G3/G4. The added advantages of using multiple
to within ≈5 kJ mol−1, in some instances these methods deviate methods which were alluded to earlier would also apply.
absolutely from the ATcT by >10 kJ mol−1, with respective The question also arises as to whether the accuracy of any
|Φmax| of 18.61 and 14.10 kJ mol−1 computed. |Φmax| in the given method deteriorates with the size of the system under
cases of G3, G4, and W1BD are computed as 10.60, 8.70, and study. Plots of MSE versus molecular weight (MW) (Figure 5)
8.82 kJ mol−1, respectively. These trends are visible in Figure 4, show that over the admittedly narrow range explored here,
where statistical measures of variation (2σ about the MSE, and increasing MW leads to increasing MSEs for both CBS-APNO
twice the RMSD) show the following trends in terms of and W1BD. Conversely, a slight, perhaps negligible, decreasing
precision: G3 ≈ G4 > W1BD > CBS-APNO > CBS-QB3. MSE is observed for CBS-QB3 computations as f(MW), and
The CBS-x methods are therefore more prone to producing almost no change for G3 and G4. Figure 5 also illustrates the
an extremely erroneous result, a certain warning flag should one general trend of increasing variation in ΔfH⊖ across methods as
MW increases.


attempt to use either of these methods in isolation for
thermochemical work. Thus, Figures 3 and 4 also delineate the
performance of combinations of two or more different CONCLUSION AND RECOMMENDATIONS
compound methods in terms of accuracy and precision. In identifying the combination CBS-APNO/G3, CBS-APNO/
For combinations of two compound methods, six combina- G4, and/or G3/G4 as the best combinations of methods for
tion are possible using the CBS-QB3, CBS-APNO, G3, and G4 computing atomization energies, we are conscious of the fact
methods. An average formation enthalpy from any combination that the different approaches used by these two methods are
of the CBS-APNO, G3, and G4 methods shows the lowest poles apart. Thus, for example, G3 utilizes HF/6-31G(d) for
MSE and MUE, and RMSD and 2σ deviations, of all individual geometry optimization and frequency determination, whereas
and combined methods. Thus, combining any two of these G4 employs B3LYP/6-31G(2df,p) for that purpose. Subse-
methods may offer an improvement over their individual quent single-point energy calculations are also considerably
implementation. Likewise, the CBS-QB3 and CBS-APNO different as is the fact that G3 reoptimizes the structure at
methods are shown to perform quite poorly in isolation, but MP2(Full)/6-31G(d) before embarking on the single-point
a combination of these methods outperforms either method in corrections. Thus, the results from this combination are to
isolation in terms of accuracy and precision. some extent insulated from errors arising from specific
In terms of computational time, Table 5, a combination of contributions, and very good data can be obtained. The fact
CBS-QB3 and the CBS-APNO methods is of a similar expense that the model chemistry CBS-APNO can only be applied to
to a combination of the G3 and G4 methods, so a combination first row atoms does militate against its more widespread
of the latter two methods offers increased accuracy and application, although a CBS-APNO/G3/G4 combination offers
precision for roughly the same cost. Combining CBS-APNO formation enthalpies which rival chemical accuracy (0.14 ±
with G4 does lead to improvements from the CBS-APNO 4.21 kJ mol−1). These uncertainties are expected to be lower
method in isolation, but combining CBS-APNO with G3, or should one employ this trio of methods in combination with
G3 with G4, would provide a more cost-effective approach. well-framed isodesmic reactions.
Excluding W1BD, CBS-APNO was the only method which For a discussion of the statistical advantages and the relevant
tended to under-predict the formation enthalpy of the target equations, applied in that case to a series of isodesmic reactions
7242 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 6. Calculated Differences (Φi, kJ mol−1) for the Five Compound Methods Tested with ATcT Recommendations and
Uncertaintiesa
molecule QB3 APNO G3 G4 W1BD ATcT ±
1,3-butadiyne −10.99 −4.09 −3.79 −1.39 −0.59 460.11 0.87
1,3-cyclopentadienea −44.60 −31.30 −39.00 −36.70 −27.40 101.30 2.50
1-butene −8.33 4.77 −2.23 −3.03 4.77 −0.03 0.48
1-butyne −10.41 0.19 −3.71 −3.61 1.79 165.39 0.85
2,2,4-trimethylpentane −12.96 18.61 −1.80 −4.04 8.58 −223.70 1.50
2,3-butanedione 2.69 7.59 0.89 −1.01 5.59 −326.81 0.98
2-butyne −9.54 0.76 −3.84 −4.24 1.26 145.76 0.79
2-propanol −0.71 7.49 −0.51 −2.71 6.69 −272.81 0.37
acetic acid 0.49 0.49 −3.61 −5.41 1.19 −433.71 0.49
acetone −1.09 8.21 −0.79 −1.89 4.61 −216.09 0.37
allene −5.55 0.35 0.15 −0.35 2.45 190.15 0.37
benzene −8.42 6.98 −5.32 −3.62 6.08 83.18 0.26
carbon monoxide 0.88 −0.93 0.07 2.88 −0.63 −110.53 0.03
cis-2-butene −9.43 3.87 −4.13 −4.53 3.97 −7.33 0.53
cyclobutene −14.10 −2.90 −10.60 −7.80 0.50 156.90 1.60
cyclohexane −10.48 12.62 −2.58 −4.18 8.82 −122.08 0.68
cyclopropane −6.89 4.11 −4.09 −2.19 3.81 53.61 0.53
cyclopropene −7.19 −0.39 −4.09 −1.79 2.81 283.91 0.59
dimethyl ether 4.18 6.48 0.28 −0.42 4.38 −184.02 0.44
dioxirane 7.20 2.00 −2.00 0.80 −1.10 1.30 1.20
ethanal 0.94 2.94 −0.46 −0.56 2.84 −165.46 0.32
ethane −2.19 6.31 0.41 −1.29 4.41 −83.79 0.17
ethene −3.64 0.76 −0.14 −0.44 2.26 52.56 0.15
ethylene glycol 4.08 5.28 −2.42 −4.52 4.78 −389.42 0.49
ethyne −6.47 −4.87 −2.37 −1.07 −0.27 228.33 0.15
ethynol −1.80 −2.90 −1.60 −2.10 1.70 92.70 1.40
formaldehyde 4.03 0.93 0.93 1.83 1.13 −109.17 0.11
formic acid 4.16 −0.34 −1.64 −1.64 1.86 −378.94 0.27
glyoxal 5.72 1.22 1.12 2.02 2.12 −212.48 0.59
isobutane −7.46 9.14 −2.36 −4.16 5.54 −135.36 0.40
isobutene −8.70 4.40 −3.00 −3.00 4.40 −17.60 0.53
ketene singlet −1.08 0.82 0.62 −1.08 0.82 −48.58 0.15
methane −0.53 3.97 0.87 −0.33 1.67 −74.53 0.06
methanediol (sc,sc) 5.49 1.99 −1.41 −1.71 5.39 −392.61 0.96
methanol 2.69 2.99 −0.41 −0.61 4.19 −200.71 0.18
methyl formate (cis Z) 7.80 5.40 2.30 1.10 4.30 −357.80 0.59
methyl hydroperoxide 6.57 1.37 −3.33 −2.13 1.67 −127.73 0.91
n-butane −7.15 9.85 −1.55 −3.95 6.55 −125.85 0.38
n-hexane −11.34 13.16 −3.44 −5.84 8.76 −166.94 0.48
o-benzyne −9.70 3.30 −4.10 0.00 3.30 460.70 1.40
oxalic acida 21.10 12.90 10.90 6.80 15.80 −721.40 2.10
oxirane (1A1) 2.28 3.08 −1.72 −0.02 2.78 −52.72 0.44
oxirene (singlet) 8.10 3.40 2.40 8.70 8.30 275.90 3.10
propane −3.91 8.69 −0.01 −2.01 6.29 −104.41 0.29
propene −5.65 3.25 −1.05 −1.35 4.25 20.35 0.33
propyne −6.70 0.70 −1.60 −0.90 2.00 185.80 0.38
succinic acid 11.87 9.75 3.25 −0.95 8.25 −817.75 0.61
toluene −10.19 9.31 −5.09 −3.39 6.41 50.41 0.37
trans-2-butene −7.98 5.12 −2.48 −2.88 5.12 −11.18 0.51
vinyl alcohol −1.86 0.14 −2.16 −2.56 3.74 −123.76 0.91
a
Species are omitted from statistical analyses.

but capable of much wider interpretation, see earlier work on A model chemistry which combines the best functional and
alkyl hydroperoxides by Simmie et al.56 basis set for optimization and frequency calculations and not
Usually transferable uncertainties are used by theoreticians too expensive higher-order corrections is badly needed for
developing new methodologies; that is, they benchmark their usage by inexpert computational chemists. It should not have
method against an existing database and then transfer that two geometry optimization routines, as in G3 or CBS-APNO
uncertainty to new species. But there is no evidence that such for example, because occasional failures can occur in the second
an approach has any grounding in reality. optimization which is perforce starting from the previous
7243 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

Table 7. Summary Statistics (kJ mol−1) Based on Results


shown in Table 6
model chemistry MUE MSE |Φmax| ±2σ 2RMSD
G3 2.27 −1.71 10.60 4.76 5.86
G4 2.46 −1.74 8.70 5.15 6.21
W1BD 3.85 3.74 8.82 5.13 8.82
APNO 4.55 3.86 18.61 9.21 12.02
QB3 6.08 −2.78 14.10 13.02 14.16
2 Methods
APNO/G4 2.05 1.06 7.29 4.71 5.17
APNO/G3 2.24 1.08 8.41 5.51 5.91
G3/G4 2.27 −1.72 9.20 4.56 5.72
QB3/APNO 3.06 0.54 10.81 7.50 7.58
QB3/G3 3.85 −2.24 12.35 8.33 9.46
QB3/G4 3.98 −2.26 10.95 8.33 9.47
3 Methods
APNO/G3/G4 1.58 0.14 7.10 4.21 4.22
QB3/APNO/G4 2.37 −0.22 8.27 5.87 5.89
QB3/APNO/G3 2.45 −0.21 9.20 6.21 6.22
QB3/G3/G4 3.23 −2.07 10.83 6.83 7.99
4+ Methods
Figure 4. Measures of dispersion (±2σ from MSE, 2 times root-mean-
QB3/APNO/G3/G4 2.17 −0.59 8.85 5.40 5.53
squared deviation) for individual compound methods, and combina-
all methods 1.80 0.28 6.98 4.75 4.79 tions thereof, upon comparison with the ATcT.

Figure 5. Signed error in standard state formation enthalpy versus


Figure 3. Mean signed errors (MSE), mean unsigned errors (MUE),
molecular mass for compounds studied. Lines are fitted via least-
and maximum absolute deviations (|Φmax|) for individual compound
squares regression. Consult online version of manuscript for
methods, and combinations thereof, upon comparison with the ATcT.
interpretation of color in figure.

geometry. This is particularly the case in the location of


transition states where energy minimization failures are rather commonly used to determine vibrational frequencies, but this
more common. Some progress in this direction has been made necessitates scale factors to correct the resultant zero-point
by da Silva who has combined the M06-2X functional within a energies and vibrational partition functions. Unfortunately,
G3-type framework to create G3X-K theory.57 separate factors seem to be required to scale the frequencies
Three out of the five model chemistries surveyed here (“low” frequencies being treated differently than “high”
employ the B3LYP functional, which although not as popular frequencies)61 to compute zero-point energies, entropies and
now as it was at the time of a 2007 review by Sousa et al.,58 is enthalpies, and worse still, these latter are functions of
still very common. A range of better functionals are available, temperature.62 An effort is needed here to simplify without
see for example recent comprehensive reviews,59,60 but have losing essential details perhaps by way of universal scale
not yet come into widespread use. factors.63 The incorporation of anharmonic effects has not been
The treatment of vibrational and rotational partition seriously attempted for moderately sized molecules for reasons
functions is crucial because from an end-user perspective discussed by Irikura.64
thermochemistry at 0 K is of a much lesser value than the same In order to reach the wider computational chemistry
data at 298.15 K. The harmonic oscillator approximation is community, the composite method should be embedded or
7244 DOI: 10.1021/jp511403a
J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

easily interfaced with all popular computational codes including (9) Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M.
those which carry a licensing fee. A 5th-Order Perturbation Comparison of Electron Correlation
Although the ATcT probably represents the current state of Theories. Chem. Phys. Lett. 1989, 157 (6), 479−483.
the art for thermochemical tabulations, and the results (10) Martin, J. M. L.; de Oliveira, G. Towards Standard Methods for
presented herein reinforce the large majority of recommenda- Benchmark Quality Ab Initio Thermochemistry: W1 and W2 Theory.
J. Chem. Phys. 1999, 111 (5), 1843−1856.
tions in this database, there remains a certain lack of
(11) Boese, A. D.; Oren, M.; Atasoylu, O.; Martin, J. M. L.; Kallay,
transparency in how some formation enthalpies reported M.; Gauss, J. W3 Theory: Robust Computational Thermochemistry in
therein have been derived, 1,3-cyclopentadiene, oxalic and the kJ/mol Accuracy Range. J. Chem. Phys. 2004, 120 (9), 4129−4141.
succinic acids, and cyclobutene being exemplary. A minor but (12) Karton, A.; Rabinovich, E.; Martin, J. M. L.; Ruscic, B. W4
important improvement to benefit auditability of the Theory for Computational Thermochemistry: In Pursuit of Confident
compendium would be direct referencing of any relevant Sub-kJ/mol Predictions. J. Chem. Phys. 2006, 125 (14), 144108.
experiment or theory which were primarily used to derive a (13) Tajti, A.; Szalay, P. G.; Császár, A. G.; Kallay, M.; Gauss, J.;
given value, as done in the Third Millennium Database.3 Valeev, E. F.; Flowers, B. A.; Vazquez, J.; Stanton, J. F. HEAT: High


*
ASSOCIATED CONTENT
S Supporting Information
Accuracy Extrapolated Ab Initio Thermochemistry. J. Chem. Phys.
2004, 121 (23), 11599−11613.
(14) Bomble, Y. J.; Vazquez, J.; Kallay, M.; Michauk, C.; Szalay, P. G.;
0 K Zero-point energy-corrected electronic energies of all Császár, A. G.; Gauss, J.; Stanton, J. F. High-Accuracy Extrapolated Ab
Initio Thermochemistry. II. Minor Improvements to the Protocol and
compounds studied and 298.15 K zero-point and thermal
a Vital Simplification. J. Chem. Phys. 2006, 125 (6), 064108.
energy-corrected electronic energies of all compounds studied. (15) Harding, M. E.; Vazquez, J.; Ruscic, B.; Wilson, A. K.; Gauss, J.;
This material is available free of charge via the Internet at Stanton, J. F. High-Accuracy Extrapolated Ab Initio Thermochemistry.
http://pubs.acs.org.


III. Additional Improvements and Overview. J. Chem. Phys. 2008, 128
(11), 114111.
AUTHOR INFORMATION (16) Császár, A. G.; Allen, W. D.; Schaefer, H. F. In Pursuit of the Ab
Corresponding Author Initio Limit for Conformational Energy Prototypes. J. Chem. Phys.
*E-mail: john.simmie@nuigalway.ie. 1998, 108 (23), 9751−9764.
Notes (17) Feller, D.; Peterson, K. A.; Dixon, D. A. Further Benchmarks of
a Composite, Convergent, Statistically Calibrated Coupled-Cluster-
The authors declare no competing financial interest.


Based Approach for Thermochemical and Spectroscopic Studies. Mol.
Phys. 2012, 110 (19−20), 2381−2399.
ACKNOWLEDGMENTS
(18) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G.
Computational resources were provided by the Irish Centre for A. A Complete Basis Set Model Chemistry. VII. Use of the Minimum
High-End Computing, ICHEC. We are grateful to Alin Elena Population Localization Method. J. Chem. Phys. 2000, 112 (15),
(ICHEC), C. Franklin Goldsmith (Brown), Amir Karton 6532−6542.
(Western Australia), Wim Klopper (Karlsruhe), Errol Lewars (19) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A. A
(Trent), Jan Martin (Weizmann Institute), and Andrew Yeung Complete Basis Set Model Chemistry. V. Extensions to Six or More
(Texas A & M) for their considerable assistance. We thank the Heavy Atoms. J. Chem. Phys. 1996, 104 (7), 2598−2619.
reviewers for their useful comments. (20) Pople, J. A.; Head-Gordon, M.; Fox, D. J.; Raghavachari, K.;


Curtiss, L. A. Gaussian-1 Theory: A General Procedure for Prediction
REFERENCES of Molecular Energies. J. Chem. Phys. 1989, 90, 5622.
(21) Curtiss, L. A.; Jones, C.; Trucks, G. W.; Raghavachari, K.; Pople,
(1) NIST-JANAF Thermochemical Tables, 4th ed.; Chase, M. W. Jr., J. A. Gaussian-1 Theory of Molecular-Energies for 2nd-Row
Ed.; J. Phys. Chem. Ref. Data Monogr. 9; American Institute of
Compounds. J. Chem. Phys. 1990, 93 (4), 2537−2545.
Physics and American Chemical Society: New York, 1998.
(22) Curtiss, L. A.; Raghavachari, K.; Trucks, G. W.; Pople, J. A.
(2) Cox, J. D.; Wagman, D. D.; Medvedev, V. A. CODATA Key
Gaussian-2 Theory for Molecular-Energies of 1st-Row and 2nd-Row
Values for Thermochemistry; Hemisphere, New York, 1989.
Compounds. J. Chem. Phys. 1991, 94 (11), 7221−7230.
(3) Goos, E.; Burcat, A.; Ruscic, B. Ideal Gas Thermochemical
(23) Curtiss, L. A.; Raghavachari, K.; Redfern, P. C.; Rassolov, V.;
Database with Updates from Active Thermochemical Tables. http://
garfield.chem.elte.hu/Burcat/burcat.html (accessed January 20, 2015). Pople, J. A. Gaussian-3 (G3) Theory for Molecules Containing First
(4) Ruscic, B.; Pinzon, R. E.; Morton, M. L.; von Laszevski, G.; and Second-Row Atoms. J. Chem. Phys. 1998, 109 (18), 7764−7776.
Bittner, S. J.; Nijsure, S. G.; Amin, K. A.; Minkoff, M.; Wagner, A. F. (24) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. Gaussian-4
Introduction to Active Thermochemical Tables: Several “Key” Theory. J. Chem. Phys. 2007, 126, 084108.
Enthalpies of Formation Revisited. J. Phys. Chem. A 2004, 108 (45), (25) Baboul, A. G.; Curtiss, L. A.; Redfern, P. C.; Raghavachari, K.
9979−9997. Gaussian-3 Theory Using Density Functional Geometries and Zero-
(5) Ruscic, B.; Pinzon, R. E.; von Laszewski, G.; Kodeboyina, D.; Point Energies. J. Chem. Phys. 1999, 110 (7), 7650−7657.
Burcat, A.; Leahy, D.; Montoya, D.; Wagner, A. F. Active (26) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K.; Rassolov, V.;
Thermochemical Tables: Thermochemistry for the 21st Century. J. Pople, J. A. Gaussian-3 Theory Using Reduced Möller-Plesset Order. J.
Phys. Conf. Ser. 2005, 16, 561−570. Chem. Phys. 1999, 110 (10), 4703−4709.
(6) Ruscic, B. Active Thermochemical Tables. ATcT.anl.gov (27) Chan, B.; Coote, M. L.; Radom, L. G4-SP, G4(MP2)-SP, G4-
(updated Active Thermochemical Tables (ATcT) values based on v SC, and G4(MP2)-SC: Modifications to G4 and G4(MP2) for the
1.110 of the Thermochemical Network (2012); last update Nov 30, Treatment of Medium-Sized Radicals. J. Chem. Theory Comput. 2010,
2013.) 6 (9), 2647−2653.
(7) Simmie, J. M. Detailed Chemical Kinetic Models for the (28) Chan, B.; Deng, J.; Radom, L. G4(MP2)-6x: A Cost-Effective
Combustion of Hydrocarbon Fuels. Prog. Energy Combust. Sci. 2003, Improvement to G4(MP2). J. Chem. Theory Comput. 2011, 7 (1),
29 (6), 599−634. 112−120.
(8) Wilke, J. J.; Allen, W. D.; Schaefer, H. F. Establishment of the (29) Chan, B.; Radom, L. W1x-1 and W1x-2: W1-Quality Accuracy
C2H5+O2 Reaction Mechanism: A Combustion Archetype. J. Chem. with an Order of Magnitude Reduction in Computational Cost. J.
Phys. 2008, 128 (7), 074308. Chem. Theory Comput. 2012, 8 (11), 4259−4269.

7245 DOI: 10.1021/jp511403a


J. Phys. Chem. A 2015, 119, 7235−7246
The Journal of Physical Chemistry A Article

(30) Karton, A.; Martin, J. M. L. Explicitly Correlated Wn Theory: (51) Vogt, N.; Abaev, M. A.; Rykov, A. N.; Shishkov, I. F.
W1-F12 and W2-F12. J. Chem. Phys. 2012, 136 (12), 124114. Determination of Molecular Structure of Succinic Acid in a Very
(31) Chan, B.; Radom, L. W3x: A Cost-Effective Post-CCSD(T) Complex Conformational Landscape: Gas-Phase Electron Diffraction
Composite Procedure. J. Chem. Theory Comput. 2013, 9 (11), 4769− (GED) and Ab Initio Studies. J. Mol. Struct. 2011, 996 (1−3), 120−
4778. 127.
(32) Karton, A.; Gruzman, D.; Martin, J. M. L. Benchmark (52) Parthiban, S.; de Oliveira, G.; Martin, J. M. L. Benchmark Ab
Thermochemistry of the CnH2n+2 Alkane Isomers (N = 2−8) and Initio Energy Profiles for the Gas-Phase SN2 Reactions Y− + CH3X →
Performance of DFT and Composite Ab Initio Methods for CH3Y + X− (X,Y = F,Cl,Br). Validation of Hybrid DFT Methods. J.
Dispersion-Driven Isomeric Equilibria. J. Phys. Chem. A 2009, 113 Phys. Chem. A 2001, 105 (5), 895−904.
(29), 8434−8447. (53) Rayne, S.; Forest, K. Estimated Gas-Phase Standard State
(33) Wheeler, S. E.; Houk, K. N.; Schleyer, P. V. R.; Allen, W. D. A Enthalpies of Formation for Organic Compounds Using the Gaussian-
Hierarchy of Homodesmotic Reactions for Thermochemistry. J. Am. 4 (G4) and W1BD Theoretical Methods. J. Chem. Eng. Data 2010, 55
Chem. Soc. 2009, 131 (7), 2547−2560. (11), 5359−5364.
(34) Würmel, J.; Simmie, J. M. Thermochemistry and Kinetics of (54) Dorofeeva, O. V. Comments on “Estimated Gas-Phase Standard
Angelica and Cognate Lactones. J. Phys. Chem. A 2014, 118, 4172− State Enthalpies of Formation for Organic Compounds Using the
4183. Gaussian-4 (G4) and W1BD Theoretical Methods” (Rayne, S.; Forest,
(35) Ruscic, B. Uncertainty Quantification in Thermochemistry, K. J. Chem. Eng. Data 2010, 55, 5359−5364). J. Chem. Eng. Data 2011,
Benchmarking Electronic Structure Computations, and Active 56 (3), 682−683.
Thermochemical Tables. Int. J. Quantum Chem. 2014, 114 (17), (55) Rayne, S.; Forest, K. Reply to Comments by O. V. Dorofeeva on
1097−1101. J. Chem. Eng. Data 2010, 55, 5359−5364. J. Chem. Eng. Data 2011, 56
(36) Simmie, J. M.; Somers, K. P.; Metcalfe, W. K.; Curran, H. J. (3), 684−685.
Substituent Effects in the Thermochemistry of Furans: A Theoretical (56) Simmie, J. M.; Black, G.; Curran, H. J.; Hinde, J. P. Enthalpies of
(CBS-QB3, CBS-APNO and G3) Study. J. Chem. Thermodyn. 2013, Formation and Bond Dissociation Energies of Lower Alkyl Hydro-
58, 117−128. peroxides and Related Hydroperoxy and Alkoxy Radicals. J. Phys.
(37) Somers, K. P. On the Pyrolysis and Combustion of Furans: Chem. A 2008, 112 (22), 5010−5016.
Quantum Chemical, Statistical Rate Theory and Chemical Kinetic (57) da Silva, G. G3X-K Theory: A Composite Theoretical Method
Modelling Studies. Ph.D. Thesis, National University of Ireland, for Thermochemical Kinetics. Chem. Phys. Lett. 2013, 558, 109−113.
Galway, 2014. (58) Sousa, S. F.; Fernandes, P. A.; Ramos, M. J. General
(38) Barnes, E. C.; Petersson, G. A.; Montgomery, J. A.; Frisch, M. J.; Performance of Density Functionals. J. Phys. Chem. A 2007, 111
Martin, J. M. L. Unrestricted Coupled Cluster and Brueckner Doubles (42), 10439−10452.
Variations of W1 Theory. J. Chem. Theory Comput. 2009, 5 (10), (59) Peverati, R.; Truhlar, D. G. Quest for a Universal Density
2687−2693. Functional: The Accuracy of Density Functionals across a Broad
(39) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Spectrum of Databases in Chemistry and Physics. Philos. Trans. R. Soc.,
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, A 2014, 372, 20120476.
B.; Petersson, G. A.; et al. Gaussian 09, revision D.01; Gaussian, Inc.: (60) Zhang, I. Y.; Xu, X. Doubly Hybrid Density Functional for
Wallingford, CT, 2009. Accurate Description of Thermochemistry, Thermochemical Kinetics
(40) Morini, F.; Knippenberg, S.; Deleuze, M. S.; Hajgato, B. and Nonbonded Interactions. Int. Rev. Phys. Chem. 2011, 30 (1), 115−
Quantum Chemical Study of Conformational Fingerprints in the 160.
Photoelectron Spectra and (e, 2e) Electron Momentum Distributions (61) Laury, M. L.; Boesch, S. E.; Haken, I.; Sinha, P.; Wheeler, R. A.;
of n-Hexane. J. Phys. Chem. A 2010, 114 (12), 4400−4417. Wilson, A. K. Harmonic Vibrational Frequencies: Scale Factors for
(41) Spartan ‘08, v 1.0.0; Wavefunction Inc., Irvine, CA, 2008. Pure, Hybrid, Hybrid Meta, and Double-Hybrid Functionals in
(42) Vacek, G.; Galbraith, J. M.; Yamaguchi, Y.; Schaefer, H. F.; Conjunction with Correlation Consistent Basis Sets. J. Comput.
Nobes, R. H.; Scott, A. P.; Radom, L. Oxirene: To Be or Not to Be. J. Chem. 2011, 32 (11), 2339−2347.
Phys. Chem. 1994, 98 (35), 8660−8665. (62) Merrick, J. P.; Moran, D.; Radom, L. An Evaluation of
(43) Lewars, E. Modeling Marvels, Chapter 3; Springer: Amsterdam, Harmonic Vibrational Frequency Scale Factors. J. Phys. Chem. A 2007,
2008. 111 (45), 11683−11700.
(44) Vogiatzis, K. D.; Haunschild, R.; Klopper, W. Accurate (63) Alecu, I. M.; Zheng, J. J.; Zhao, Y.; Truhlar, D. G.
Atomization Energies from Combining Coupled-Cluster Computa- Computational Thermochemistry: Scale Factor Databases and Scale
tions with Interference-Corrected Explicitly Correlated Second-Order Factors for Vibrational Frequencies Obtained from Electronic Model
Perturbation Theory. Theor. Chem. Acc. 2014, 133 (3), 1−12. Chemistries. J. Chem. Theory Comput. 2010, 6 (9), 2872−2887.
(45) Karton, A.; Daon, S.; Martin, J. M. L. W4−11: A High- (64) Irikura, K. K. Anharmonic Partition Functions for Polyatomic
Thermochemistry. J. Chem. Thermodyn. 2014, 73, 183−189.
Confidence Benchmark Dataset for Computational Thermochemistry
Derived from First-Principles W4 Data. Chem. Phys. Lett. 2011, 510
(4−6), 165−178.
(46) Pedley, J. B.; Naylor, R. D.; Kirby, S. P. Thermochemical Data of
Organic Compounds, 2nd ed.; Chapman & Hall: London, 1986.
(47) Roth, W. R.; Adamczak, O.; Breuckmann, R.; Lennartz, H. W.;
Boese, R. Resonance Energy Calculation: The MM2ERW Force-Field.
Chem. Ber. 1991, 124 (11), 2499−2521.
(48) Furuyama, S.; Golden, D. M.; Benson, S. W. Thermochemistry
of Cyclopentene and Cyclopentadiene from Studies of Gas-Phase
Equilibria. J. Chem. Thermodyn. 1970, 2 (2), 161−169.
(49) Zhong, X.; Bozzelli, J. W. Thermochemical and Kinetic Analysis
of the H, OH, HO2, O, and O2 Association Reaction with
Cyclopentadienyl Radical. J. Phys. Chem. A 1998, 102, 3537−3555.
(50) Agapito, F.; Nunes, P. A.; Cabral, B. J. C.; dos Santos, R. M. B.;
Simoes, J. A. M. Energetic Differences between the Five- and Six-
Membered Ring Hydrocarbons: Strain Energies in the Parent and
Radical Molecules. J. Org. Chem. 2008, 73 (16), 6213−6223.

7246 DOI: 10.1021/jp511403a


J. Phys. Chem. A 2015, 119, 7235−7246

You might also like