You are on page 1of 7

DOI: 10.1002/chem.

201504016 Full Paper

& Density Functional Calculations

How Many Water Molecules Does it Take to Dissociate HCl?


Alba Vargas-Caamal,[a] Jose Luis Cabellos,[a] Filiberto Ortiz-Chi,[b] Henry S. Rzepa,[c]
Albeiro Restrepo,*[d] and Gabriel Merino*[a]
In memory of Jon Mikel Azpiroz—we dearly miss him

Abstract: The potential energy surfaces of the HCl(H2O)n (n of clusters according to the nondissociated, partially dissoci-
is the number of water molecules) clusters are systematically ated, and fully dissociated character of the H¢Cl bond. Our
explored using density functional theory and high-level ab computations show that if temperature is not controlled
initio computations. On the basis of electronic energies, the during the experiment, the values obtained for the dipole
number of water molecules needed for HCl dissociation is moment (or for any measurable property) are susceptible to
four as reported by some experimental groups. However, change, providing a different picture of the number of water
this number is five owing to the inclusion of entropic fac- molecules needed for HCl dissociation in a nanoscopic drop-
tors. Wiberg bond indices are calculated and analyzed, and let.
the results provide a quadratic correlation and classification

Introduction the beam deflection method, Guggemos et al.[20] observed that


“the addition of a DCl molecule to a water cluster results in
Water plays a central role in our understanding of chemical a strongly enhanced susceptibility. There is evidence for a noticea-
and biological processes, and it has the special ability to disso- ble rise in the dipole moment occurring at n … 5–6. This size is
ciate molecules. In particular, acid dissociation, leading to consistent with predictions for the onset of ionic dissociation”.
proton transfer[1, 2] and solvation of the fragments,[3–13] is one of These results exceed by two the number of water molecules
the most important chemical processes. The dissociation of required to dissociate HCl reported in earlier experimental and
HCl has been the focus of recent studies in ambient bulk theoretical works.[21, 22] For instance, Gutberlet et al. reported
water,[1, 2, 14, 15] as well as at low temperatures in confined geo- the experimental observation of a nanoscopic aqueous droplet
metries such as on amorphous ice surfaces, on ice nanocrys- of acid formed within a superfluid helium cluster at 0.37 K. By
tals, and in microsolvation environments, which are relevant to using high-resolution mass-selective infrared laser spectrosco-
atmospheric or interstellar problems.[16–19] Very recently, after py, these authors found that successive aggregation of HCl
measuring the dipole moments of DCl-doped water clusters by with water molecules, HCl(H2O)n, readily results in the forma-
tion of hydronium at n = 4.[23]
The literature related to the dissociation of acids in aqueous
[a] A. Vargas-Caamal, J. L. Cabellos, G. Merino
Departamento de F†sica Aplicada environments and the microsolvation of the resulting anions
Centro de Investigaciûn y de Estudios Avanzados, Unidad M¦rida and protons is extensive.[15, 20, 24–44] In 2011, Leopold published
Km 6 Antigua Carretera a Progreso. Apdo. Postal 73 a comprehensive review of this topic,[45] in which Table 1 col-
Cordemex, 97310, M¦rida, Yuc. (M¦xico) lects experimental and theoretical evidence about the number
E-mail: gmerino@mda.cinvestav.mx
of water molecules needed to dissociate some simple acids,
[b] F. Ortiz-Chi
Instituto Tecnolûgico Superior de Calkin† suggesting four water molecules for the case of HCl. A note ac-
Av. Ah-Canul s/n, Carr. Fed. Calkin†-Campeche companying Table 1 explicitly calls for the reader to “note excel-
CP 24900, Calkin†, Campeche (M¦xico) lent agreement with the experimental work”, making reference
[c] H. S. Rzepa to papers by the groups of Maillard[24] and Suhm,[26, 46] who
Department of Chemistry, Imperial College London
used several IR spectroscopy techniques to draw their conclu-
South Kensington campus, London, SW7 2AZ (UK)
sions. As a generalization, Leopold summarized the results
[d] A. Restrepo
Instituto de Qu†mica, Universidad de Antioquia UdeA found in the literature in the following way: “the overall picture
Calle 70 No. 52–21, Medell†n (Colombia) is that for acids that ultimately ionize in aqueous solution, three
E-mail: albeiro.restrepo@udea.edu.co to five water molecules typically are needed to induce ionization
Supporting information for this article is available on the WWW under in the lowest-energy isomer of a given cluster. This range is en-
http://dx.doi.org/10.1002/chem.201504016. This includes the variation of
tirely consistent with available experimental evidence”.[45]
the dipole moments as a function of temperature for the HCl(H2O)n clusters,
and the Cartesian coordinates for all the local minimum structures found High-level ab initio and DFT computations on molecular clus-
by Bilatu. ters face a tremendous challenge in producing accurate poten-

Chem. Eur. J. 2016, 22, 2812 – 2818 2812 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

tial energy surfaces (PESs) for intermolecular interactions B2PLYPD3/def2-TZVP results. All computations are performed
owing to the exponential dependency of the number of local using the very-tight integral accuracy and ultrafine grid op-
minima on the number of molecules.[47–49] In most cases, chem- tions available in the Gaussian 09 (revision D.01) package.[70]
ical intuition is the weapon of choice to guess the best candi- Each structure is characterized as a true minimum by harmonic
dates. However, this biased approach quickly becomes imprac- vibrational frequency analysis.
tical because of the overwhelming number of possibilities, To validate our methodology, we optimized and character-
which, even for clusters of relatively small sizes, is bound to ized the structures of the HCl–H2O and HCl(H2O)2 clusters (one
miss relevant regions of the corresponding PESs. This is a very and seven local minima, respectively) at the CCSD(T)/def2-
sensitive issue, especially in the cases where several local TZVP level of theory, obtaining very similar structures and rela-
minima are very close in energy to the global minimum with- tive energies to those found at the B2PLYPD3 level, supporting
out being structurally related. Water clusters are the archetypal the use of this approach.
example of complicated energy landscapes populated by mul-
titudes of structurally unrelated local minima,[47, 50–57] which in
Energetics and structures
several cases are not discernible from the corresponding
global minimum using either experimental or theoretical meth- We uncovered minima on the PESs for the interactions of HCl
ods. The situation is further complicated by low interconver- with up to six water molecules of previously unnoticed com-
sion barriers and by consideration of temperature, entropy, plexity and structural diversity. A total of 431 different local
and internal degrees of freedom,[58–62] which usually shuffle sta- minima were located, distributed as 1, 7, 22, 82, 148, and 171
bility orders in favor of more open structures as the tempera- isomers for n = 1, 2, 3, 4, 5, and 6, respectively (n is the
ture increases, eventually leading to ideal gas behavior at suffi- number of water molecules). Despite the staggering number
ciently high temperatures.[47, 52] Computations on interactions of structures found here, the PESs in question cannot be con-
of water with HCl, dissociated or not, have been performed sidered fully characterized; this is not an attainable goal with
through intuitive construction of initial guesses for cluster currently available experimental or computational methods be-
structures[27, 28, 30, 50] (see Leopold’s review[45] for a more compre- cause the number of possible local minima increases exponen-
hensive list of theoretical calculations on the subject), with the tially with the size of the system. Nonetheless, we are confi-
aim of identifying the structural preferences for all possible dent that our searches capture the most important structural
combinations of interacting species that could be found in features of the HCl(H2O)n clusters.
HCl/water mixtures. We found structures containing nondissociated, partially dis-
In view of the preceding discussion, in this article we have sociated, and fully dissociated HCl. A wide variety of interac-
attempted to understand the microsolvation process of HCl, tions stabilizing the clusters are at play: water–water and HCl–
and specifically, to clarify why the number of water molecules water hydrogen bonds (HBs) with the possibility of HCl acting
needed to achieve dissociation may vary depending on the ex- as donor and as acceptor of one or two HBs, long-range Cl···H
perimental technique employed. To this end, we adopt a sto- interactions, and microsolvation of the H3O + cation in the
chastic, structurally unbiased approach to sample the potential form of Eigen cations (each proton interacting with a lone pair
energy surfaces exhaustively for interactions between hydro- from a vicinal water molecule), and what we refer to as quasi-
chloric acid and one to six water molecules. Our results, based Eigen cations, in which Cl¢ takes over the stabilizing role of
on the electronic energies, indicate that the number of water one of the water molecules (in well-defined Eigen cations,[27, 45]
molecules needed to dissociate HCl is four. However, this con- the H3O + moiety is not in direct contact with the Cl¢ anion).
clusion changes upon consideration of entropic factors. At No Zundel (H2O···H···OH2) + cations were located. However, as
room temperature, the number of water molecules needed to in the case of the Eigen cations, a number of partially dissociat-
dissociate HCl is five. Therefore, the effects of temperature on ed structures, in which the stabilizing role of the second water
the stability and dipole moments are discussed in detail. molecule is assumed by Cl¢ , were identified (Cl¢···H + ···OH2).
Figure 1 A shows the global minimum structures of the
HCl(H2O)n (n = 1–6) clusters optimized using the B2PLYPD3 ap-
Computational Details
proach considering the zero-point energy (ZPE) corrections. In
With the steady growth in computational power at affordable our nomenclature, HCl-nx-y, x indicates the number of water
cost, stochastic explorations of moderately sized atomic and molecules, and y denotes the energetic relative ordering posi-
molecular clusters using quantum Hamiltonians have become tion with respect to the lowest free energy structures. The
a viable choice for obtaining detailed information on the most noteworthy structural variation up to n = 3 involves the
nature of intermolecular interactions and to overcome the key O···H interaction contracting rapidly (from 1.863 to 1.569 æ)
problems associated with biased initial guesses. Our computa- until a hydrogen atom is transferred completely to form a hy-
tional procedure employs a modified kick algorithm called dronium ion for n = 4 (structure HCl-n4-36). So, the overall pic-
Bilatu[48, 63–65] to generate starting structures, establishing a hier- ture provided by these computations is that for HCl, four
archical screening at the PBE0[66]/D95 V level, and the water molecules are needed to induce full dissociation.
B2PLYPD3 dispersion-corrected double-hybrid functional[67, 68] in Interestingly, our computations show that acid dissociation
conjunction with the def2-TZVP[69] basis set. Therefore, the depends on the inclusion of the entropic factors. Figure 1 also
geometrical and energetic discussion is based on the collects the lowest free energy forms at 298.15 K and 1 atm. A

Chem. Eur. J. 2016, 22, 2812 – 2818 www.chemeurj.org 2813 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

Figure 1. A) Lowest energy structures (DEZPE) and B) lowest free energy structures at 298.15 K and 1 atm (DG298.15) for HCl(H2O)n, n = 1–6. Wiberg Bond Indices
for the H···Cl interactions (WBIHCl) are also shown. Isomer populations (%xi) were estimated from standard Boltzmann distributions using the Gibbs free ener-
gies (298 K, 1 atm). ClH···O distances in æ. Data taken from the B2PLYPD3/def2-TVZP optimized geometries.

comparison between the lowest energy and the lowest free Guggemos et al.[20] reported the addition of a DCl (not HCl)
energy structures reveals small structural differences for n = 2 molecule to a water cluster. In principle, this substitution could
and n = 3, in that the hydrogen atoms of the water molecules affect the energetics. Table 1 summarizes the free energy dif-
point in different directions. Strictly speaking, the structures ferences obtained for the n = 2 case upon substitution of the
are different, but the motif is the same. The first notable HCl unit by DCl. It is apparent that the substitution of HCl by
change is perceived for n = 4. At room temperature, the most DCl does not alter the relative energies.
stable structure is an associated pentameric form with an O···H
interaction distance of 1.516 æ (structure HCl-n4-1). The dissoci-
Bonding analysis
ation is complete for the n = 5 case (structure HCl-n5-1), which
is structurally closely related to the global minimum in the A convenient way to analyze proton transfer structurally is to
water pentamer.[54] It is clear that entropy factors are important correlate the distances between all atoms involved in the
for explaining the size evolution of these clusters. Cl···H···O hydrogen bonds (see Figure 2).[71] Cases of the lowest
Thus, temperature is a key variable in determining the equi- free energy form for n = 6 are not included in this correlation.
librium structure of a hydrated molecule. For n = 4, isomer
populations estimated through Boltzmann distributions sug-
gest that at 298.15 K and 1 atm, the lowest energy form con- Table 1. Comparison of free relative energies obtained with HCl and DCl
tributes only approximately 0.14 % (DG = 2.6 kcal mol¢1), where- for n = 2.
as the lowest DG structure (see Figure 1 B) contributes around Isomer DGHCl DGDCl
11 %. There is no ambiguity about the fully dissociated charac- [kcal mol¢1] [kcal mol¢1]
ter of the global minima and for n = 5, 6: the lowest DG struc- 1 0.00 0.00
ture for n = 5 is a quasi-Eigen cation (structure HCl-n5-1), and 2 0.00 0.00
the structure for n = 6 contains an Eigen cation (structure HCl- 3 0.16 0.14
n6-1). The presence of Eigen and quasi-Eigen cations and the 4 0.16 0.14
5 4.35 4.10
absence of Zundel cations are remarkable because it is known 6 5.07 5.04
that Eigen cations are the dominant forms present in bulk 7 5.07 5.04
samples of aqueous solutions of dissociated acids. In summary,
dissociation of HCl occurs with four water molecules without
correction for entropic effects and ZPE, as reported by Mail- For linear hydrogen bonds, q1, defined as (r1¢r2)/2, is the dis-
lard[24] and Suhm,[26, 46] but occurs with five water molecules if tance from the hydrogen atom to the center of the hydrogen
such effects are included. bond, and q2 is the distance between the Cl and O atoms (r1

Chem. Eur. J. 2016, 22, 2812 – 2818 www.chemeurj.org 2814 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

Figure 2. Correlation between q2 and q1. The squares indicate the positions Figure 3. Wiberg bond indices (WBI) for the H¢Cl interaction in clusters re-
of the lowest free energy structures for each molecularity. q2 and q1 are in æ. ported in this work. Partially transferred protons appear in two categories:
Data taken from the B2PLYPD3/def2-TVZP geometries. the right peak shows those in the early stages of H¢Cl dissociation, and the
left peak comprises structures for advanced HCl dissociation. Data taken
from the B2PLYPD3/def2-TVZP optimized geometries.

and r2 are the Cl¢H and O¢H distances, respectively). High


negative (positive) values of q1 correspond to hydrogen atoms tion of complex energy landscapes should be considered com-
that are clearly attached to the chlorine (oxygen) atom, and q1 plete. This is particularly applicable to our case because of the
… 0 indicates partially transferred protons, as the hydrogen large number of structural possibilities arising from the relative
atoms are equidistant from Cl and O. As in other proton trans- positions of hydrogen atoms not involved in the stabilizing hy-
fer cases,[72] a beautiful quadratic correlation is obtained re- drogen-bonding networks, which change the individual dipole
gardless of the number of water molecules present. Figure 2 moments significantly. Additionally, it is clear that the entropic
provides solid evidence for proton transfer beyond subjective effects play a key role in determining the equilibrium structure
visual inspection of the structures. Clearly, the lowest free of the title clusters.
energy structures for n ‹ 4 are in the attached region, and Using our calculated structures, we attempted to reproduce
complete transfer is evident for n = 5. the dipole moment distributions reported by Guggemos
Further evidence about the state of HCl dissociation in gas- et al.[20] Following the postulates of quantum mechanics, we
phase water clusters is provided in Figure 3; here, we analyze computed the expected dipole, < mn > , as the average over
the Wiberg bond indices for the H¢Cl interactions (as a refer- dipole moments of all individual structures in the given PES,
ence, the computed bond index for isolated H¢Cl is 0.94) for min, weighted by xin, the estimated isomer populations on that
the same clusters as selected for Figure 2. Three distinct areas surface obtained from standard Boltzmann distributions of the
can be discerned. The undissociated region covers the [0.8– Gibbs free energies. The results at 298.15 K are plotted in
0.9] interval. There are two intervals of partial proton transfer: Figure 4. Experimental points were taken directly from the
[0.65–0.75], corresponding to early to intermediate stages of work by Guggemos et al.[20]
the transfer process, and [0.25–0.45], corresponding to inter- There is a perfect match between the expected dipole
mediate to late stages of proton transfer. Finally, the [0.1–0.2] moment and the experimental value (Figure 4 A) only for n = 3.
interval encloses structures that all exhibit fully transferred pro- For larger n, there is no clear correlation between the two
tons. Figure 3 provides clear-cut evidence for proton transfer, quantities. However, if < mn > are computed including only the
which is consistent with the results discussed previously in undissociated (or dissociated) structures, the picture changes
Figure 2. drastically. The expected dipole moments for the undissociated
forms increase considerably from n = 4 to n = 6, whereas
< mn > for the dissociated arrangements decrease in the same
Dipole moment distribution
range (see Figure 4 B). In the particular case of n = 4, the
Guggemos and co-workers[20] wrote “we point out that the dy- weight of the undissociated structures is very low. Interestingly,
namically averaged dipole moment within a finite highly fluxio- the expected dipole moments for the associated clusters are
nal system can be qualitatively different from that computed for very close to those obtained experimentally for n = 4 and 5.
a static minimum-energy framework”. We complement this On the contrary, for n = 6, the computed value for the dissoci-
statement by adding that the differences may also be quantita- ated forms is very close to that reported by Guggemos et al.
tive. This issue is further complicated, because chemical and The n = 5 case is interesting. Besides our stochastically locat-
physical processes are not exclusive domains of the lowest ed structures, we took the lowest DG isomer shown in
energy structures in a given PES, and as mentioned above, Figure 1 and optimized the molecular geometries for all possi-
many more structures may still be missing because no explora- ble combinations of the “above” and “under” relative positions

Chem. Eur. J. 2016, 22, 2812 – 2818 www.chemeurj.org 2815 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

a result of a temperature variation, and this could affect the ex-


pected dipole moment. In this sense, it has been shown that
the number of water molecules needed to dissociate NaCl, for
example, is sensitive to temperature.[73] It is apparent from
Figure 5 that changes in temperature induce very strong

Figure 5. Variation in the expected dipole moments as a function of temper-


ature for the HCl(H2O)n clusters. The expected dipoles moments are estimat-
ed from standard Boltzmann distributions of the Gibbs free energies at
1 atm.

changes in the dipole moments. At very low temperatures, the


variation is almost negligible (except for n = 6). However, it is
not possible to predict a lowering or increasing of the dipole
Figure 4. Expected dipole moments for the HCl(H2O)n clusters. A) Expected
dipole moments for all structures in a given PES. B) Expected dipole mo- moment as a consequence of the temperature even when the
ments for undissociated and dissociated structures in a given PES. In both population is separated into dissociated and associated forms
cases the expected dipoles moments are estimated from standard Boltz- (see Supporting Information). Therefore, if the temperature is
mann distributions of the Gibbs free energies at 298.15 K, 1 atm.
not controlled during the experiment, the values obtained for
the dipole moments are susceptible to change, providing a dif-
ferent picture of the number of water molecules needed for
of the four peripheral protons (16 combinations in total). The HCl dissociation in water. The main reason is that at high tem-
resulting structures were energetically very close to the lowest peratures, the number of hydrogen bonds decreases, and less
DG isomer, with DG values no higher than 0.43 kcal mol¢1 rela- compact clusters are obtained. In other words, we recover the
tive to the reference structure. Without these additional struc- gas behavior (no intermolecular interactions, and thus, no clus-
tures, isomers had populations with higher values, with other ter formation) at high temperatures.
geometries having smaller significance. Inclusion of the new
structures complicates the population picture with 11 isomers
Conclusion
now having contributions in the 3–6 % range. Individual dipole
moments are also more susceptible to changes in the orienta- On the basis of electronic energies, the number of water mole-
tion of peripheral protons: amongst the new structures, ex- cules needed for HCl dissociation is four as reported by the
pected dipole moments cover the 2.46–5.09 D range, (com- groups of Maillard,[24] Suhm,[26, 46] and others. However, on the
pared with 3.39 D for the reference structure). However, the ex- basis of Gibbs free energies, this number is five. For n = 4, the
pected averaged dipole moment only changed from 3.77 to lowest DG structure is also partially dissociated. The nondisso-
3.80 D. These results illustrate clearly the preceding discussion ciated, partially dissociated, and fully dissociated character of
regarding contributions from individual structures to the prop- the H¢Cl bond can be characterized unambiguously by Wiberg
erties of complex energy landscapes, and at the same time bond indices, and supported by quadratic correlations be-
provide solid support for our approach to calculating expected tween the distance between Cl and O atoms and the distance
properties as statistical averages weighted by isomer popula- from the proton to the center of the corresponding Cl···H···O
tions derived from Boltzmann distributions. interaction. Very narrow energy windows for the partially and
What about the temperature effect on the dipole moment fully dissociated structures in the corresponding PES preclude
distribution? Clearly, the isomer populations will change as the assignment of macroscopic experimental observables to in-

Chem. Eur. J. 2016, 22, 2812 – 2818 www.chemeurj.org 2816 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

dividual cluster structures. Although our search is comprehen- [25] V. E. Bondybey, M. Beyer, U. Achatz, S. Joos, G. Niedner-Schatteburg, Isr.
sive and exhaustive, it is by no means complete because of J. Chem. 1999, 39, 213 – 219.
[26] M. Farnik, M. Weimann, M. A. Suhm, J. Chem. Phys. 2003, 118, 10120 –
the large number of structural possibilities arising from the rel- 10136.
ative positions of peripheral hydrogen atoms in the clusters. [27] A. A. Hassanali, J. Cuny, M. Ceriotti, C. J. Pickard, M. Parrinello, J. Am.
Paraphrasing Guggemos et al., we point out that the dynamical- Chem. Soc. 2012, 134, 8557 – 8569.
ly averaged dipole moment within a finite highly fluxional [28] C. T. Lee, C. Sosa, M. Planas, J. J. Novoa, J. Chem. Phys. 1996, 104, 7081 –
7085.
system can be qualitatively and quantitatively different from that [29] M. Letzner, S. Gruen, D. Habig, K. Hanke, T. Endres, P. Nieto, G. Schwaab,
computed for a static minimum-energy framework. It is manda- L. Walewski, M. Wollenhaupt, H. Forbert, D. Marx, M. Havenith, J. Chem.
tory to separate the undissociated from the dissociated struc- Phys. 2013, 139, 154304.
tures in order to reproduce the overall experimental trend. Our [30] S. Odde, B. J. Mhin, S. Lee, H. M. Lee, K. S. Kim, J. Chem. Phys. 2004, 120,
9524 – 9535.
computations show that if temperature is not controlled [31] T. Schindler, C. Berg, G. Niedner-Schatteburg, V. E. Bondybey, Chem.
during the experiment, the values obtained for the dipole Phys. Lett. 1994, 229, 57 – 64.
moment (or for any measurable property) are susceptible to [32] Y. K. Yoon, G. B. Carpenter, Acta Crystallogr. 1959, 12, 17 – 20.
change, providing a different picture of the number of water [33] H. S. Rzepa, The Winnower 3:e142410.09115 (2015). DOI:10.15200/
winn.142410.09115.
molecules needed for HCl dissociation in water. [34] M. Masia, H. Forbert, D. Marx, J. Phys. Chem. A 2007, 111, 12181 – 12191.
[35] J. Zischang, D. Skvortsov, M. Yong Choi, R. A. Mata, M. A. Suhm, A. F. Vi-
lesov, J. Phys. Chem. A 2015, 119, 2636 – 2643.
Acknowledgements [36] M. Onč‚k, P. Slav†ček, M. F‚rn†k, U. Buck, J. Phys. Chem. A 2011, 115,
6155 – 6168.
The Moshinsky Foundation, Tecnolûgico Nacional de M¦xico [37] A. M. Morrison, S. D. Flynn, T. Liang, G. E. Douberly, J. Phys. Chem. A
2010, 114, 8090 – 8098.
(069-14-PD) and PRODEP ( ITESCAM-PTC-026) supported the [38] V. Lesch, S. Jeremias, A. Moretti, S. Passerini, A. Heuer, O. Borodin, J.
work in M¦xico. A. R. thanks Universidad de Antioquia for pro- Phys. Chem. B 2014, 118, 7367 – 7375.
viding partial funding for this work via “Estrategia de sostenibi- [39] J. S. Mancini, A. K. Samanta, J. M. Bowman, H. Reisler, J. Phys. Chem. A
lidad”. AVC and JLC thank Conacyt for the PhD and postdoctor- 2014, 118, 8402 – 8410.
[40] Ł. Walewski, H. Forbert, D. Marx, ChemPhysChem. 2013, 14, 817 – 826.
al fellowship, respectively. [41] S. Re, Y. Osamura, Y. Suzuki, H. F. Schaefer, J. Chem. Phys. 1998, 109,
973 – 977.
[42] H. Forbert, M. Masia, A. Kaczmarek-Kedziera, N. N. Nair, D. Marx, J. Am.
Keywords: ab initio calculations · dissociation · global
Chem. Soc. 2011, 133, 4062 – 4072.
optimization · hydrogen bonds · microsolvation [43] E. M. Cabaleiro-Lago, J. M. Hermida-Ramûn, J. Rodr†guez-Otero, J. Chem.
Phys. 2002, 117, 3160 – 3168.
[1] D. Marx, ChemPhysChem 2006, 7, 1848 – 1870. [44] J. S. Mancini, J. M. Bowman, Phys. Chem. Chem. Phys. 2015, 17, 6222 –
[2] D. Marx, ChemPhysChem 2007, 8, 209 – 210. 6226.
[3] W. Domcke, A. L. Sobolewski, Science 2003, 302, 1693 – 1694. [45] K. R. Leopold, Annu. Rev. Phys. Chem. 2011, 62, 327 – 349.
[4] S. M. Hurley, T. E. Dermota, D. P. Hydutsky, A. W. Castleman, Science [46] M. Weimann, M. Farnik, M. A. Suhm, Phys. Chem. Chem. Phys. 2002, 4,
2002, 298, 202 – 204. 3933 – 3937.
[5] M. Miyazaki, A. Fujii, T. Ebata, N. Mikami, Science 2004, 304, 1134 – 1137. [47] N. Acelas, G. Hincapie, D. Guerra, J. David, A. Restrepo, J. Chem. Phys.
[6] W. H. Robertson, E. G. Diken, E. A. Price, J. W. Shin, M. A. Johnson, Sci- 2013, 139, 044310.
ence 2003, 299, 1367 – 1372. [48] M. Saunders, J. Comput. Chem. 2004, 25, 621 – 626.
[7] W. H. Robertson, M. A. Johnson, Science 2002, 298, 69. [49] D. Guerra, J. David, A. Restrepo, J. Comput. Methods Sci. Eng. 2014, 14,
[8] J. W. Shin, N. I. Hammer, E. G. Diken, M. A. Johnson, R. S. Walters, T. D. 93 – 102.
Jaeger, M. A. Duncan, R. A. Christie, K. D. Jordan, Science 2004, 304, [50] J. K. Gregory, D. C. Clary, J. Phys. Chem. A 1997, 101, 6813 – 6819.
1137 – 1140. [51] C. Z. Hadad, A. Restrepo, S. Jenkins, F. Ramirez, J. David, Theor. Chem.
[9] A. Stace, Science 2001, 294, 1292 – 1293. Acc. 2013, 132, 1376 – 1387.
[10] A. F. Voegele, K. R. Liedl, Angew. Chem. Int. Ed. 2003, 42, 2114; Angew. [52] G. HincapieÈ, N. Acelas, M. Castano, J. David, A. Restrepo, J. Phys. Chem.
Chem. 2003, 115, 2162 – 2164. A 2010, 114, 7809 – 7814.
[11] J. M. Headrick, Science 2005, 309, 1326. [53] J. F. P¦rez, C. Z. Hadad, A. Restrepo, Int. J. Quantum Chem. 2008, 108,
[12] J. M. Headrick, E. G. Diken, R. S. Walters, N. I. Hammer, R. A. Christie, J. 1653 – 1659.
Cui, E. M. Myshakin, M. A. Duncan, M. A. Johnson, K. D. Jordan, Science [54] F. Ram†rez, C. Z. Hadad, D. Guerra, J. David, A. Restrepo, Chem. Phys.
2005, 308, 1765 – 1769. Lett. 2011, 507, 229 – 233.
[13] T. S. Zwier, Science 2004, 304, 1119 – 1120. [55] C. P¦rez, M. T. Muckle, D. P. Zaleski, N. A. Seifert, B. Temelso, G. C.
[14] K. Ando, J. T. Hynes, Adv. Chem. Phys. 1999, 110, 381 – 430. Shields, Z. Kisiel, B. H. Pate, Science 2012, 336, 897 – 901.
[15] K. Laasonen, M. L. Klein, J. Am. Chem. Soc. 1994, 116, 11620 – 11621. [56] Y. Wang, V. Babin, J. M. Bowman, F. Paesani, J. Am. Chem. Soc. 2012,
[16] R. Bianco, J. T. Hynes, Acc. Chem. Res. 2006, 39, 159 – 165. 134, 11116 – 11119.
[17] J. P. Devlin, N. Uras, J. Sadlej, V. Buch, Nature 2002, 417, 269 – 271. [57] V. Babin, F. Paesani, Chem. Phys. Lett. 2013, 580, 1 – 8.
[18] T. Huthwelker, M. Ammann, T. Peter, Chem. Rev. 2006, 106, 1375 – 1444. [58] W. Lin, F. Paesani, J. Phys. Chem. A 2013, 117, 7131 – 7141.
[19] V. F. McNeill, T. Loerting, F. M. Geiger, B. L. Trout, M. J. Molina, Proc. Natl. [59] T. Takayanagi, K. Takahashi, A. Kakizaki, M. Shiga, M. Tachikawa, Chem.
Acad. Sci. USA 2006, 103, 9422 – 9427. Phys. 2009, 358, 196 – 202.
[20] N. Guggemos, P. Slavicek, V. V. Kresin, Phys. Rev. Lett. 2015, 114, 043401. [60] L. Walewski, H. Forbert, D. Marx, J. Phys. Chem. Lett. 2011, 2, 3069 –
[21] Z. Kisiel, E. Bialkowska-Jaworska, L. Pszczûlkowski, A. Milet, C. Strunie- 3074.
wicz, R. Moszynski, J. Sadlej, J. Chem. Phys. 2000, 112, 5767 – 5776. [61] S. Sugawara, T. Yoshikawa, T. Takayanagi, M. Tachikawa, Chem. Phys. Lett.
[22] D. Skvortsov, S. J. Lee, M. Y. Choi, A. F. Vilesov, J. Phys. Chem. A 2009, 2011, 501, 238 – 244.
113, 7360 – 7365. [62] A. Milet, C. Struniewicz, R. Moszynski, P. E. S. Wormer, J. Chem. Phys.
[23] A. Gutberlet, G. Schwaab, ©. Birer, M. Masia, A. Kaczmarek, H. Forbert, 2001, 115, 349 – 356.
M. Havenith, D. Marx, Science 2009, 324, 1545 – 1548. [63] J. L. Cabellos, F. Ortiz-Chi, A. Ram†rez, G. Merino, Bilatu 1.0. Cinvestav,
[24] C. Amirand, D. Maillard, J. Mol. Struct. 1988, 176, 181 – 201. M¦rida, 2013.

Chem. Eur. J. 2016, 22, 2812 – 2818 www.chemeurj.org 2817 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Paper

[64] R. Grande-Aztatzi, P. R. Martinez-Alanis, J. L. Cabellos, E. Osorio, A. Marti- prich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K.
nez, G. Merino, J. Comput. Chem. 2014, 35, 2288 – 2296. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford,
[65] A. Ramirez-Manzanares, J. Pena, J. M. Azpiroz, G. Merino, J. Comput. J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi,
Chem. 2015, 36, 1456 – 1466. R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayak-
[66] C. Adamo, V. Barone, J. Chem. Phys. 1999, 110, 6158 – 6170. kara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C.
[67] L. Goerigk, S. Grimme, J. Chem. Theory Comput. 2011, 7, 291 – 309. Gonzalez, J. A. Pople; Gaussian, Inc.: Wallingford CT, 2004.
[68] S. Grimme, J. Chem. Phys. 2006, 124, 034108. [71] H. H. Limbach, P. M. Tolstoy, N. P¦rez-Hern‚ndez, J. Guo, I. G. Shendero-
[69] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297 – 3305. vich, G. S. Denisov, Isr. J. Chem. 2009, 49, 199 – 216.
[70] M. J. Frisch, G. W. T., H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. [72] J. D. Gonzalez, E. Florez, J. Romero, A. Reyes, A. Restrepo, J. Mol. Model.
Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, 2013, 19, 1763 – 1777.
J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. [73] C.-K. Siu, B. S. Fox-Beyer, M. K. Beyer, V. E. Bondybey, Chem. Eur. J. 2006,
Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. 12, 6382 – 6392.
Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O.
Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V.
Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev,
A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Moroku- Received: October 7, 2015
ma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dap- Published online on January 15, 2016

Chem. Eur. J. 2016, 22, 2812 – 2818 www.chemeurj.org 2818 Ó 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like