Molecules 25 02150 v2
Molecules 25 02150 v2
Article
Dimerization of Acetic Acid in the Gas Phase—NMR
Experiments and Quantum-Chemical Calculations
Ondřej Socha 1,2 and Martin Dračínský 1, *
1 Institute of Organic Chemistry and Biochemistry, Czech Academy of Sciences, Flemingovo nám. 2,
166 10 Prague, Czech Republic; [email protected]
2 Faculty of Mathematics and Physics, Charles University, Ke Karlovu 3, 166 10 Prague, Czech Republic
* Correspondence: [email protected]; Tel./Fax: +42-02-2018-3139
Abstract: Due to the nature of the carboxylic group, acetic acid can serve as both a donor and acceptor of
a hydrogen bond. Gaseous acetic acid is known to form cyclic dimers with two strong hydrogen bonds.
However, trimeric and various oligomeric structures have also been hypothesized to exist in both the
gas and liquid phases of acetic acid. In this work, a combination of gas-phase NMR experiments and
advanced computational approaches were employed in order to validate the basic dimerization model
of gaseous acetic acid. The gas-phase experiments performed in a glass tube revealed interactions of
acetic acid with the glass surface. On the other hand, variable-temperature and variable-pressure
NMR parameters obtained for acetic acid in a polymer insert provided thermodynamic parameters
that were in excellent agreement with the MP2 (the second order Møller–Plesset perturbation theory)
and CCSD(T) (coupled cluster with single, double and perturbative triple excitation) calculations
based on the basic dimerization model. A slight disparity between the theoretical dimerization model
and the experimental data was revealed only at low temperatures. This observation might indicate
the presence of other, entropically disfavored, supramolecular structures at low temperatures.
1. Introduction
Noncovalent interactions play a significant role in many chemical and biological processes.
For instance, molecular recognition, protein folding, and crystal packing are phenomena where these
interactions shape the structure, properties, and function of molecules and materials. Hydrogen bonding
is the most important noncovalent interaction. Hydrogen bonds are typically an order of magnitude
weaker than covalent bonds, and yet they determine, for example, the secondary structure of large
biopolymers such as DNA or proteins.
Carboxylic acids possess a functional group that can simultaneously serve as a hydrogen-bond
donor and acceptor. Therefore, carboxylic acids can form a number of supramolecular structures,
and they are often used as model systems to gain an insight into the hydrogen-bonding interaction [1].
Acetic acid, which is an archetypal carboxylic acid, serves as an excellent example to demonstrate this.
Crystalline acetic acid is comprised of linear chains of hydrogen-bonded molecules [2]. In the gas
phase, on the other hand, the presence of centrosymmetric cyclic dimers of acetic acid has been confirmed
by several experimental studies [3–6]. Each of the dimers contains two equivalent strong O–H···O=C
hydrogen bonds. The dimerization enthalpy and entropy have been estimated to be about −15 kcal/mol
and −37 cal/K·mol, respectively, by several experiments, such as IR experiments and vapor-density or
thermal-conductivity measurements [7]. However, NMR spectroscopy experiments of gaseous acetic
acid have estimated the dimerization enthalpy and entropy as about −18 kcal/mol and −42 cal/K·mol,
respectively [8]. There is no clear consensus about the prevailing supramolecular structure in liquid
acetic acid [9]. While neutron diffraction [10] and time-domain Raman spectroscopy [11,12] experiments
have suggested that liquid acetic acid primarily consists of cyclic dimers, other experiments, such as
large-angle X-ray scattering [13] and IR [14] and Raman [15] spectroscopy, have suggested that the
main structural patterns in liquid acetic acid are linear chains similar to those found in the solid.
Furthermore, other studies have indicated that acetic acid monomers and cyclic trimers coexist with
the dimers and chains in the liquid phase [16,17].
Acetic acid and its supramolecular complexes have also been investigated in many theoretical
studies [18–22]. Thanks to its small size, acetic acid can be used for benchmarking different computational
approaches including various density-functional-theory (DFT) functionals or post-Hartree–Fock
methods [23–25]. Even expensive methods, such as coupled clusters (CC) and quadratic configuration
interaction, are feasible [26]. Besides the estimation of binding energies, quantum-chemical calculations
have been utilized for predictions of NMR, IR, and Raman spectroscopic parameters [27,28].
In this work, we combined NMR experiments with advanced computational approaches to better
understand the hydrogen-bonding interactions of acetic acid. The experimental part of this study
aimed to precisely measure the thermodynamic parameters of acetic acid dimerization using gas-phase
NMR spectrometry and to explain the previously observed discrepancies of NMR-derived dimerization
enthalpy and entropy with those obtained by other experiments. The goal of the computational
part was to accurately calculate the dimerization Gibbs energy and proton chemical shifts of the
monomeric and dimeric forms of acetic acid in the gas phase. Several DFT and post-Hartree–Fock
methods were employed for energy and NMR computations. In addition, we utilized path-integral
molecular dynamics (PIMD) simulations to obtain an insight into nuclear quantum effects such as
nuclear delocalization and proton tunneling in the acetic acid dimer. A comparison of theoretical and
experimental thermodynamic and spectroscopic parameters was used to discuss the accuracy of the
computational methods, as well as to provide a validation of the basic dimerization model of gaseous
acetic acid.
2M ↔ D (1)
where M and D stand for the monomer and dimer of acetic acid, respectively. The equilibrium constant
K is then defined in terms of partial pressures:
pD
K= . (2)
p2M
According to the Dalton law of partial pressures, the total pressure ptot can be obtained using
the formula:
ptot = pM + 2pD . (3)
The observed chemical shift δ is calculated as a weighted average of monomeric and dimeric
shifts δM and δD , respectively:
pM
δ= (δM − δD ) + δD (4)
ptot
where the partial pressure of the monomeric form pM is derived using the above equations, which leads
to the expression:
p
1 + 8Kptot − 1
pM = (5)
4K
Molecules 2020, 25, 2150 3 of 16
The equilibrium constant K is directly related to the change of enthalpy ∆H and entropy ∆S of
dimerization through the van’t Hoff equation:
∆H − T∆S
K = exp − (6)
RT
was more than ten times less water in the original sample. Therefore, we concluded that the water
content in the acetic acid used in our experiments was far below 1 mol%.
Measurements in the FEP inserts were performed at constant temperature. The FEP inserts were
not airtight, and the acetic acid vapors continuously leaked from the insert for several hours or days,
depending on the temperature. After the evaporation of the liquid, the partial pressure of the acetic
acid showed an exponential decay during the time t, which was best observed on the integral intensities
I(t) of the methyl-group NMR signal. The pressure at a given time (snapshot) was calculated using
the formula:
I (t)
ptot (t) = p0 (8)
I (t ) t=t
0
I (t)
ptot (t) = p0 − pcorr (9)
I (t ) t=t
0
The chemical shift of the monomer ∆δM was fixed at the value of 3.76 ppm, which was obtained
through extrapolation to zero pressure from the measurement at 150 ◦ C. To improve the convergence
of the optimization procedure, the Savitzky–Golay smoothing filter of the 3rd order with a 21-point
window was applied to the pressure data ptot (t). Further window widening had only a negligible effect
on the optimized parameters. There were between 600 and 1200 pressure points in each dataset.
The uncertainty of the optimized parameters was estimated from the p0 pressure uncertainty.
Two contributions to p0 uncertainty were considered: systematic contribution (10% of p0 ) and
dataset-specific contribution (5% of p0 ). Both originated from the uncertainty of the point where the
liquid evaporated, see below. Though these contributions could be merged into one, they tend to
behave differently in terms of error propagation. The systematic error seems to have had only a minor
influence on the uncertainties of the optimized thermodynamic parameters. The main source of the
error was thus a slight inconsistency in p0 determination between individual data sets. Standard
deviations of the fitted parameters were estimated using the Monte Carlo method. In 1000 iterations,
the predictors p0 were re-sampled from normal distributions with variance parameters reflecting the
above uncertainties. The set of re-optimized parameters was statistically analyzed, yielding standard
deviations of all optimized parameters and ∆H and ∆S covariance. All fitted parameters including
their uncertainties are in the Supplementary Materials.
A
EX CBS
corr = Ecorr + (10)
X3
where X = 2,3,4,5 is the “cardinal number” of the basis set used to calculate the respective energy EX and
A is the system-dependent free parameter. As can be seen in Figure S8, the double-ζ basis set yielded
inconsistent results; therefore, these values were excluded from the extrapolation. The Hartree–Fock
(HF) energy calculated with Aug-cc-pV5Z was then added to ECBS corr in order to obtain the MP2/CBS
binding energy ECBS
MP2
. More detailed values are in Table S3 in the Supplementary Materials. A different
approach to CBS extrapolation was employed in the case of CCSD(T). Due to prohibitively demanding
requirements of the CCSD(T) method, only a single-point energy calculation with the Aug-cc-pVTZ
basis set was possible. However, it has been observed that the difference between the MP2 and CCSD(T)
Molecules 2020, 25, 2150 6 of 16
calculated energy is only slightly dependent on the basis set. Formula (4), which has been proposed to
exploit this fact [44], was employed here:
∆ECBS
CCSD(T )
= ∆ECBS
MP2 + ∆ECCSD ( T ) − ∆EMP2 (11)
smaller basis set
For all MP2 and CCSD(T) calculations, the entropy change ∆S and the thermal contributions to
the enthalpy change ∆Hthermal of dimerization were calculated at the MP2/Aug-cc-pVTZ level. These
values were subsequently added to the MP2- and CCSD(T)-calculated single-point energy values ∆E
according to the equation:
The values of ∆G0 and ∆H were reported for direct comparison with experimental values.
No special treatment of vibronic modes for the correction of entropy or enthalpy was applied.
For the B3LYP and MP2 calculations of 1 H chemical shielding, the gauge-independent atomic
orbital (GIAO) method [45] with the Aug-pcS-2 basis set [46] was used. These computations were
performed on the structures optimized with the Aug-cc-pVQZ basis set using the same method as the
subsequent NMR calculations.
PIMD simulations were performed by the CASTEP program [47], version 17.2, using an NVT
ensemble, the temperature of 300 K, a Langevin thermostat, a 0.5-fs integration time step, ultrasoft
pseudopotentials [48], and a planewave cutoff energy of 300 eV. The integrals were taken over the
Brillouin zone using a Monkhorst–Pack [49] grid of the minimum k-point sampling of 0.1 Å−1 .
Electron-correlation effects were modeled using the generalized-gradient approximation of Perdew,
Burke, and Ernzerhof [50]. The acetic acid monomer and dimer were placed in a cubic periodic box of
15 × 15 × 15 Å−1 , and the atomic positions were optimized by energy minimization prior to the PIMD
runs at the same computational level. No symmetry constraints were applied during the simulation
of the dimer. The production runs of the monomer and dimer simulation of the lengths of 4 and
6 ps, respectively, were preceded by 2-ps equilibration. The path-integral propagation used a Trotter
decomposition of all nuclei into 16 beads.
Time-averaged NMR parameters were computed from 1000 snapshots from the MD and PIMD
simulations. The B3LYP functional with the Aug-pcS-2 basis set was employed for the calculation of
NMR shieldings.
Figure 1. The 1 H–NMR spectrum of acetic acid in a glass insert at the temperature of 60 ◦ C. The spectrum
was not referenced.
An unexpected temperature-dependent line broadening of the COOH signal was observed in the
spectra of gaseous acetic acid in the glass insert. Strangely, the broadening became more prominent
as the temperature increased. This dependence is depicted in Figure 3 and is also noticeable in the
spectra (Figure 2 and Figures S4–S6). Though some broadening can be expected in solution, where it
can indicate a chemical exchange between various supramolecular structures, it is surprising in the
gas phase. In our model, we assumed only one process, which was the dimerization of acetic acid
monomers. We expected this process to have a relatively high exchange rate at the temperatures
attained in our experiments, leading to narrow averaged signals. This broadening pointed to another
unexpected exchange process. One of the possible explanations may be related to water presence in the
sample. However, we did not expect a considerable interaction, such as hydrogen bonding between
acetic acid and water molecules in the gas phase because of the high purity of the sample. Water content
was determined to be far below 1% (molar) using liquid NMR (see the Methods and Supplementary
Materials for more details). Moreover, we expected the interaction with water to be energetically
less favorable than the binding to another acetic acid molecule. The only other explanation of the
unexpected line broadening was the interaction of acetic acid molecules with the walls of the glass
insert, specifically with the oxygen-containing functional groups on its surface. The existence of these
interactions has already been indicated in a previous study [51].
The temperature dependence of the chemical-shift difference ∆δ = δOH − δMe was further analyzed.
The basic dimerization Model (2) was fitted to the experimental data. The optimized parameters
(∆δM , ∆δD , ∆H and ∆S) showed a strong dependence on the pressure ptot , despite the fact that they
were expected to be pressure-independent. This supported the hypothesis of the presence of another
chemical exchange with molecules adsorbed on the glass walls.
To eliminate the influence of the glass surface on the observed data, we decided to utilize a different
material for the NMR tube insert. There are commercially available tube liners made of FEP, which is
a highly inert material. Unfortunately, these inserts are not airtight, which causes complications when
they are used to measure gas samples. However, we exploited the slow gas leakage for the measurement
of the pressure dependence of NMR spectra, where pressure is a function of time. The inserts were
filled with a few drops of liquid sample and subsequently heated up to the desired temperature.
Figure 4 shows an example of the time and thus indirect pressure dependence of the 1 H–NMR spectra
of acetic acid. The signal broadening was no longer present in the spectra, which suggested that this
was, indeed, a glass-specific effect. This fact also provided a possible explanation for the inaccuracy of
the resultant values of ∆H and ∆S in the previous NMR study discussed in the introduction, in which
glass NMR tubes were used [8].
Molecules 2020, 25, 2150 8 of 16
Figure 2. Temperature series of 1 H–NMR spectra of acetic acid (c = 7:9 mM) in the glass insert.
The spectra were not referenced.
Figure 3. Temperature dependence of the half width of the 1 H–NMR signal of the acetic acid COOH
group in the glass insert.
Molecules 2020, 25, 2150 9 of 16
Figure 4. An example of the time dependence of the acetic acid 1 H–NMR spectra in the fluorinated
ethylene propylene copolymer (FEP) insert at the temperature of 80 ◦ C. The gas-phase signals are
highlighted by red color. The spectra were not referenced.
The signal intensities varied during the initial phase of the experiment in the FEP insert due to
the line shape broadening caused by the exchange between the liquid and gas states. It is evident
from Figure 5 that the line broadening was exhibited by both acetic acid signals. Whereas the COOH
linewidth showed a strong dependence on time, the signal of the methyl group tended to have
a constant width, with the only disruption during the l + g → g transition.
During the measurements in the FEP inserts, another set of acetic acid signals was also present in
the spectra. These, however, did not come from the inside of the FEP insert; rather, they came from
the small space between the insert and the enclosing NMR tube. The vapors escaping from the insert
tended to condensate there; therefore, the signals corresponded to liquid acetic acid. The signals of the
gas-state molecules inside the insert were thus not affected by those “outer” molecules.
Figure 5. The 1 H–NMR line broadening in the presence of acetic acid vapors and condensate.
A transition is observed after the liquid evaporation.
The thermodynamic parameters ∆H and ∆S were globally fitted to four datasets at temperatures
25, 60, 80, and 100 ◦ C (Figure 6), yielding the values ∆H = (−15.4 ± 0.5) kcal/ mol, ∆S = (−36.6 ± 1.5)
cal/ mol·K, and ∆G0 = (−4.48 ± 0.12) kcal/ mol. The chemical-shift difference ∆δ of the dimer was
separately determined for each dataset, with the average value being ∆δD = 10.55 ppm. The chemical
shift of the monomer ∆δM = 3.76 ppm was obtained through the extrapolation of the chemical-shift
dependence at 150 ◦ C to zero pressure.
A visual analysis of the fitted data (Figure 6) showed a worse fit at lower temperatures, especially
at 25 ◦ C. This could have partly been caused by a higher noise in the data at lower temperatures,
which is also apparent in the picture. However, the possibility of another chemical exchange between
other supramolecular structures cannot be excluded. Since these supramolecular forms (e.g., trimers
and tetramers) would be entropically less stable than dimers, they would only exist in the acetic acid
Molecules 2020, 25, 2150 10 of 16
vapors at low temperature and/or high pressure. This hypothesis would explain why our model,
which considers only monomers and dimers, fit better to the high-temperature data.
3.2. Computations
Three different computational methods were employed for the calculation of the binding energy
of the acetic acid dimer. B3LYP is a widely used DFT functional due to its speed and good results for
organic molecules. The second-order Møller–Plesset perturbation theory was chosen as a popular
representative of post-Hartree–Fock methods. MP2 scaled sensibly with the system size and allowed for
the geometry optimization of the acetic acid dimer with a quadruple-ζ basis set. The quite demanding
coupled-cluster method with single, double excitations, and triple excitations as a perturbation CCSD(T)
is one of the most precise methods for electronic-energy calculations today.
Figure 6. The pressure dependence of the chemical shift of acetic acid vapors in the FEP insert fitted
using Expression (2). The chemical shift ∆δM was obtained independently through extrapolation from
a 150 ◦ C dataset. Note that the ptot values used in the fit were based on the experimentally obtained
concentration. The experimental temperatures were (a) 25 ◦ C, (b) 60 ◦ C, (c) 80 ◦ C and (d) 100 ◦ C.
Table 1. Interatomic distances (Å) and the O–C–O valence angle (^, ◦ ) in the cyclic dimer of acetic acid
obtained at the second order Møller–Plesset perturbation theory (MP2) level and their differences from
the values obtained at the B3LYP level. GD3: Grimme’s dispersion correction.
We hypothesize that the large difference between the experimental and calculated values of ∆δD
could have originated from the lack of vibrational averaging in the calculations. In the case of the monomer,
the vibrational corrections to both the COOH and methyl protons have been found to be similar in several
cases [55] and thus might largely have cancelled out in the reported ∆δM value. However, the vibrational
correction of the carboxylic hydrogen might have been significantly higher in the dimeric case because of
the more prominent potential anharmonicity than in the case of the monomer.
Table 2. The calculated values of the dimerization enthalpy (kcal/mol), entropy (cal/mol·K), and Gibbs
energy (kcal/mol) of acetic acid at different computational levels. The results for the MP2 and
CCSD(T) (coupled cluster with single, double and perturbative triple excitation) methods rely on
MP2/Aug-cc-pVTZ thermochemical calculation. Aug-cc-pVDZ: augmented correlation-consistent
double-ζ; V5Z: quintuple-ζ; CBS: complete basis set.
In order to estimate the importance of the vibrational-averaging contribution to the chemical shifts,
we performed PIMD simulations of the monomer and dimer of acetic acid. This method was chosen for
its ability to comprehend nuclear quantum effects such as quantum nuclear delocalization and tunneling,
exhibited especially by low-mass atoms like hydrogen. It has previously been demonstrated that PIMD
simulations combined with DFT shielding calculations provide excellent agreement with experimental
solid-state NMR chemical-shift changes induced by temperature change or by deuterium isotope
substitution, which are the consequences of vibrational averaging [56,57]. However, these calculations
also revealed that the PBE functional used in these simulations underestimated the energetic barriers
of proton transfer [58]. Furthermore, the PIMD simulations and subsequent shielding calculations
are computationally very demanding because a large number (hundreds to thousands) of geometry
snapshots must be used to obtain sufficiently converged results.
As can be seen in Table 3, the PIMD-based vibrational correction to the monomer ∆δM value was small
and negative, and when added to the B3LYP or MP2 values, the agreement with experiment was improved.
However, in the case of the dimer, the shielding calculations did not converge, even after averaging
1000 geometry snapshots from the PIMD simulation. This poor convergence led to differences between
the shieldings of the equivalent OH protons in the dimer. Furthermore, the probability distributions of the
calculated ∆δD values were very broad (Figure S9), and insufficient convergence could thus have led to
large errors in the calculated ∆δD values. We also noticed that the use of the PBE functional in the PIMD
simulations led to the oversampling of the dimer geometries, which were characterized by high values of
the chemical-shift difference ∆δD (Figure S10). Therefore, the values of the vibrational correction for the
dimer were probably biased towards positive values and unreliable.
Molecules 2020, 25, 2150 13 of 16
Table 3. The calculated 1 H–NMR chemical-shift differences δOH –δMe (ppm) of the acetic acid monomer
and dimer. The computations were performed on the structures optimized with the Aug-cc-pVQZ
basis set using the same method as the subsequent NMR calculation.
4. Conclusions
We performed gas-phase NMR measurements of acetic acid at variable pressures and temperatures.
These experiments allowed us to extract the thermodynamic parameters of acetic acid dimerization.
These values were in good agreement with previous values based on IR experiments and vapor-density
or thermal-conductivity measurements. We noticed, however, that the gas-phase measurements of
acetic acid had to be performed in inert polymer tubes because interactions of acetic acid with the
surface of glass tubes led to significantly biased results. These experiments also provided chemical
shifts of isolated monomeric and dimeric forms of acetic acid.
The temperature and pressure dependence of the observed chemical shifts of acetic acid was
in very good agreement with the dependences derived for the basic dimerization model. We only
observed minor deviations for the lowest experimental temperature (25 ◦ C). These deviations may
indicate the presence of other supramolecular form(s), such as those of the trimer or tetramer, of acetic
acid at low temperatures in the gas phase.
The experimental thermodynamic parameters and chemical shifts were compared with those
obtained by DFT, MP2, and CCSD(T) calculations. The calculated thermodynamic parameters of the
dimerization and the chemical-shift difference δOH –δMe of the acetic acid monomer were in very good
agreement with the experimental values. However, the calculated chemical-shift difference δOH –δMe
of the acetic acid dimer was ca 1 ppm higher than the experimental value. We hypothesize that this
overestimation was caused by neglecting the vibrational averaging in the calculations. We intended
to estimate the vibrational contribution to the chemical shifts by PIMD simulations and subsequent
shielding calculations for geometry snapshots. Unfortunately, we found out that the computational
method used in the PIMD simulations (the PBE functional) was not suitable for this task because it
oversampled the geometries with very short OH···O hydrogen-bond distances. However, these results
demonstrate the importance of an accurate treatment of vibrational corrections to NMR parameters.
Author Contributions: Conceptualization, O.S. and M.D.; methodology, O.S.; investigation, O.S.; writing—original
draft preparation, O.S.; writing—review and editing, M.D.; supervision, M.D. All authors have read and agreed to
the published version of the manuscript.
Funding: This research was funded by the Czech Science Foundation, grant no. 20-01472S, and by the Charles
University Grant Agency, project no. 1494119.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to
publish the results.
References
1. Davies, J.A.; Hanson-Heine, M.W.D.; Besley, N.A.; Shirley, A.; Trowers, J.; Yang, S.F.; Ellis, A.M. Dimers of
acetic acid in helium nanodroplets. Phys. Chem. Chem. Phys. 2019, 21, 13950–13958. [CrossRef]
2. Nahringbauer, I. Hydrogen bond studies .39. Reinvestigation of crystal structure of acetic acid (at +5 degrees
c and −190 degrees C). Acta Chem. Scand. 1970, 24, 453–462. [CrossRef]
3. Frurip, D.J.; Curtiss, L.A.; Blander, M. Vapor-phase association in acetic and trifluoroacetic
acids—thermal-conductivity measurements and molecular-orbital calculations. J. Am. Chem. Soc. 1980, 102,
2610–2616. [CrossRef]
4. Emmeluth, C.; Suhm, M.A. A chemical approach towards the spectroscopy of carboxylic acid dimer isomerism.
Phys. Chem. Chem. Phys. 2003, 5, 3094–3099. [CrossRef]
5. Derissen, J.L. Reinvestigation of molecular structure of acetic acid monomer and dimer by gas electron
diffraction. J. Mol. Struct. 1971, 7, 67–80. [CrossRef]
6. Bertie, J.E.; Michaelian, K.H. The raman spectrum of gaseous acetic acid at 21 ◦ C. J. Chem. Phys. 1982, 77,
5267–5271. [CrossRef]
7. Togeas, J.B. Acetic acid vapor: 2. A statistical mechanical critique of vapor density experiments.
J. Phys. Chem. A 2005, 109, 5438–5444. [CrossRef] [PubMed]
8. Lumbrosobader, N.; Coupry, C.; Baron, D.; Clague, D.H. Dimerization of carboxylic acids: A vapor-phase
NMR Study. J. Magn. Reson. 1975, 17, 386–392. [CrossRef]
9. Lengvinaite, D.; Aidas, K.; Kimtys, L. Molecular aggregation in liquid acetic acid: Insight from molecular
dynamics/quantum mechanics modelling of structural and NMR properties. Phys. Chem. Chem. Phys. 2019,
21, 14811–14820. [CrossRef] [PubMed]
10. Bertagnolli, H. The structure of liquid acetic-acid—an interpretation of neutron-diffraction results by
geometrical models. Chem. Phys. Lett. 1982, 93, 287–292. [CrossRef]
11. Heisler, I.A.; Mazur, K.; Yamaguchi, S.; Tominaga, K.; Meech, S.R. Measuring acetic acid dimer modes by
ultrafast time-domain Raman spectroscopy. Phys. Chem. Chem. Phys. 2011, 13, 15573–15579. [CrossRef]
[PubMed]
12. Lütgens, M.; Friedriszik, F.; Lochbrunner, S. Direct observation of the cyclic dimer in liquid acetic acid by
probing the C=O vibration with ultrafast coherent Raman spectroscopy. Phys. Chem. Chem. Phys. 2014, 16,
18010–18016. [CrossRef] [PubMed]
13. Takamuku, T.; Kyoshoin, Y.; Noguchi, H.; Kusano, S.; Yamaguchi, T. Liquid structure of acetic acid-water
and trifluoroacetic acid-water mixtures studied by large-angle x-ray scattering and NMR. J. Phys. Chem. B
2007, 111, 9270–9280. [CrossRef] [PubMed]
14. Flakus, H.T.; Hachula, B. The source of similarity of the IR spectra of acetic acid in the liquid and solid-state
phases. Vib. Spectrosc. 2011, 56, 170–176. [CrossRef]
15. Nakabayashi, T.; Kosugi, K.; Nishi, N. Liquid structure of acetic acid studied by Raman spectroscopy and ab
initio molecular orbital calculations. J. Phys. Chem. A 1999, 103, 8595–8603. [CrossRef]
16. Wu, J.P. Gaussian analysis of Raman spectroscopy of acetic acid reveals a significant amount of monomers that
effectively cooperate with hydrogen bonded linear chains. Phys. Chem. Chem. Phys. 2014, 16, 22458–22461.
[CrossRef]
17. Fathi, S.; Bouazizi, S.; Trabelsi, S.; Gonzalez, M.A.; Bahri, M.; Nasr, S.; Bellissent-Funel, M.C. Structural
investigation of liquid acetic acid by neutron scattering, DFT calculations and molecular dynamics simulations.
Complementarity to x-ray scattering results. J. Mol. Liq. 2014, 196, 69–76. [CrossRef]
Molecules 2020, 25, 2150 15 of 16
18. Pašalić, H.; Tunega, D.; Aquino, A.J.A.; Haberhauer, G.; Gerzabek, M.H.; Lischka, H. The stability of the
acetic acid dimer in microhydrated environments and in aqueous solution. Phys. Chem. Chem. Phys. 2012,
14, 4162–4170. [CrossRef]
19. Lim, V.T.; Bayly, C.I.; Fusti-Molnar, L.; Mobley, D.L. Assessing the conformational equilibrium of carboxylic
acid via quantum mechanical and molecular dynamics studies on acetic acid. J. Chem. Inf. Model. 2019, 59,
1957–1964. [CrossRef]
20. Zhang, M.H.; Chen, L.H.; Yang, H.M.; Ma, J. Theoretical study of acetic acid association based on hydrogen
bonding mechanism. J. Phys. Chem. A 2017, 121, 4560–4568. [CrossRef]
21. Chocholoušová, J.; Vacek, J.; Hobza, P. Acetic acid dimer in the gas phase, nonpolar solvent, microhydrated
environment, and dilute and concentrated acetic acid: Ab initio quantum chemical and molecular dynamics
simulations. J. Phys. Chem. A 2003, 107, 3086–3092. [CrossRef]
22. Colominas, C.; Teixido, J.; Cemeli, J.; Luque, F.J.; Orozco, M. Dimerization of carboxylic acids: Reliability of
theoretical calculations and the effect of solvent. J. Phys. Chem. B 1998, 102, 2269–2276. [CrossRef]
23. Řezáč, J.; Riley, K.E.; Hobza, P. S66: A Well-balanced database of benchmark interaction energies relevant to
biomolecular structures. J. Chem. Theory Comput. 2011, 7, 2427–2438. [CrossRef] [PubMed]
24. Řezáč, J.; Hobza, P. Advanced corrections of hydrogen bonding and dispersion for semiempirical quantum
mechanical methods. J. Chem. Theory Comput. 2012, 8, 141–151. [CrossRef]
25. Brauer, B.; Kesharwani, M.K.; Martin, J.M.L. Some observations on counterpoise corrections for explicitly
correlated calculations on noncovalent interactions. J. Chem. Theory Comput. 2014, 10, 3791–3799. [CrossRef]
26. Miller, C.E.; Francisco, J.S. A quadratic configuration interaction study of the proton affinity of acetic acid.
Chem. Phys. Lett. 2002, 364, 427–431. [CrossRef]
27. Saunders, C.M.; Tantillo, D.J. Application of computational chemical shift prediction techniques to the
cereoanhydride structure problem—carboxylate complications. Mar. Drugs 2017, 15, 171. [CrossRef]
28. Rudolph, W.W.; Fischer, D.; Irmer, G. Vibrational spectroscopic studies and DFT calculations on
NaCH3 CO2 (aq) and CH3 COOH(aq). Dalton Trans. 2014, 43, 3174–3185. [CrossRef]
29. de Nevers, N. Air pollution control engineering; Waveland Pr Inc.: Long Grove, IL, USA, 2010.
30. Dean, J.A. Lange’s Handbook of Chemistry, 15th ed.; McGraw-Hill, Inc.: New York, NY, 1999.
31. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.;
Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016.
32. Becke, A.D. Density-functional thermochemistry 3. The role of exact exchange. J. Chem. Phys. 1993, 98,
5648–5652. [CrossRef]
33. Lee, C.T.; Yang, W.T.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional
of the electron-density. Phys. Rev. B 1988, 37, 785–789. [CrossRef]
34. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density
functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104.
[CrossRef]
35. Dunning, T.H. Gaussian-basis sets for use in correlated molecular calculations.1. The atoms boron through
neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [CrossRef]
36. Kendall, R.A.; Dunning, T.H.; Harrison, R.J. Electron-affinities of the 1st-row atoms revisited–systematic
basis-sets and wave-functions. J. Chem. Phys. 1992, 96, 6796–6806. [CrossRef]
37. Boys, S.F.; Bernardi, F. Calculation of small molecular interactions by differences of separate total
energies—some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [CrossRef]
38. Møller, C.; Plesset, M.S. Note on an approximation treatment for many-electron systems. Phys. Rev. 1934, 46,
618–622. [CrossRef]
39. Bartlett, R.J.; Purvis, G.D. Many-body perturbation-theory, coupled-pair many-electron theory, and importance
of quadruple excitations for correlation problem. Int. J. Quantum Chem. 1978, 14, 561–581. [CrossRef]
40. Čížek, J. On the use of the cluster expansion and the technique of diagrams in calculations of correlation effects
in atoms and molecules. In Advances in Chemical Physics; LeFebvre, R., Moser, C., Eds.; John Wiley & Sons, Ltd.:
London, UK, 1969; Volume 14, pp. 35–89.
41. Purvis, G.D.; Bartlett, R.J. A full coupled-cluster singles and doubles model—the inclusion of disconnected
triples. J. Chem. Phys. 1982, 76, 1910–1918. [CrossRef]
42. Scuseria, G.E.; Janssen, C.L.; Schaefer, H.F. An efficient reformulation of the closed-shell coupled cluster
single and double excitation (Ccsd) equations. J. Chem. Phys. 1988, 89, 7382–7387. [CrossRef]
Molecules 2020, 25, 2150 16 of 16
43. Helgaker, T.; Klopper, W.; Koch, H.; Noga, J. Basis-set convergence of correlated calculations on water.
J. Chem. Phys. 1997, 106, 9639–9646. [CrossRef]
44. Hobza, P. Theoretical studies of hydrogen bonding. In Annual Reports Section C (Physical Chemistry);
Webb, G.A., Ed.; Royal Society of Chemistry: Cambridge, UK, 2004; Volume 100, pp. 3–27.
45. Wolinski, K.; Hinton, J.F.; Pulay, P. Efficient implementation of the gauge-independent atomic orbital method
for NMR chemical-shift calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [CrossRef]
46. Jensen, F. Basis set convergence of nuclear magnetic shielding constants calculated by density functional
methods. J. Chem. Theory Comput. 2008, 4, 719–727. [CrossRef] [PubMed]
47. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.J.; Refson, K.; Payne, M.C. First principles
methods using CASTEP. Z. Kristallogr. 2005, 220, 567–570. [CrossRef]
48. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B
1990, 41, 7892–7895. [CrossRef] [PubMed]
49. Monkhorst, H.J.; Pack, J.D. Special points for brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192.
[CrossRef]
50. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett.
1996, 77, 3865–3868. [CrossRef]
51. Stapleton, J.J.; Suchy, D.L.; Banerjee, J.; Mueller, K.T.; Pantano, C.G. adsorption reactions of carboxylic acid
functional groups on sodium aluminoborosilicate glass fiber surfaces. Acs Appl. Mater. Interfaces 2010, 2,
3303–3309. [CrossRef]
52. Alvarez-Idaboy, J.R.; Galano, A. Counterpoise corrected interaction energies are not systematically better
than uncorrected ones: Comparison with CCSD(T) CBS extrapolated values. Theor. Chem. Acc. 2010, 126,
75–85. [CrossRef]
53. Sheng, X.W.; Mentel, L.; Gritsenko, O.V.; Baerends, E.J. Counterpoise correction is not useful for short and
van der waals distances but may be useful at long range. J. Comput. Chem. 2011, 32, 2896–2901. [CrossRef]
54. Mentel, L.M.; Baerends, E.J. Can the counterpoise correction for basis set superposition effect be justified?
J. Chem. Theory Comput. 2014, 10, 252–267. [CrossRef]
55. Ruud, K.; Åstrand, P.O.; Taylor, P.R. Zero-point vibrational effects on proton shieldings: Functional-group
contributions from ab initio calculations. J. Am. Chem. Soc. 2001, 123, 4826–4833. [CrossRef]
56. Dračínský, M.; Hodgkinson, P. Effects of quantum nuclear delocalisation on NMR parameters from path
integral molecular dynamics. Chem. Eur. J. 2014, 20, 2201–2207. [CrossRef] [PubMed]
57. Dračínský, M.; Bouř, P.; Hodgkinson, P. Temperature dependence of NMR parameters calculated from path
integral molecular dynamics simulations. J. Chem. Theory Comput. 2016, 12, 968–973. [CrossRef] [PubMed]
58. Pohl, R.; Socha, O.; Slavíček, P.; Šála, M.; Hodgkinson, P.; Dračínský, M. Proton transfer in guanine-cytosine
base pair analogues studied by NMR spectroscopy and PIMD simulations. Faraday Discuss. 2018, 212,
331–344. [CrossRef] [PubMed]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).