Anionic Halide···Alcohol Clusters in the Solid State - The Journal of

Sep 5, 2014 - Crystallographic Information File. jp506256a_si_001.cif (101.64 kB). Citing Articles; Related Content. Citation data is made available b...
0 downloads 18 Views 4MB Size
Article pubs.acs.org/JPCA

Anionic Halide···Alcohol Clusters in the Solid State Pavel V. Gushchin,† Maxim L. Kuznetsov,‡ Matti Haukka,§ and Vadim Yu. Kukushkin*,†,∥ †

Institute of Chemistry, Saint Petersburg State University, 198504 Stary Petergof, Russian Federation Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisbon, Portugal § Department of Chemistry, University of Jyväskylä, P.O. Box 35, FI-40014, Jyväskylä, Finland ∥ Institute of Macromolecular Compounds of Russian Academy of Sciences, V. O. Bolshoii Pr. 31, 199004, Saint Petersburg, Russian Federation ‡

S Supporting Information *

ABSTRACT: The cationic (1,3,5-triazapentadiene)PtII complexes [1](Cl)2, [2](Cl)2, [3](Br)2, and [4](Cl)2, were crystallized from ROH-containing systems (R = Me, Et) providing alcohol solvates studied by X-ray diffraction. In the crystal structures of [1−4][(Hal)2(ROH)2] (R = Me, Et), the Hal− ion interacts with two or three cations [1−4]2+ by means of two or three or four contacts thus uniting stacked arrays of complexes into the layers. The solvated MeOH or EtOH molecules occupy vacant space, giving contacts with [1−4]2+, and connects to the Hal− ion through a hydrogen bridge via the H(1O)O(1S) H atom forming, by means of the Hal−···HOR (Hal = Cl, Br) contact, the halide− alcohol cluster. Properties of the Cl−···HO(Me) H-bond in [1][(Cl)2(MeOH)2] were analyzed using theoretical DFT methods.



INTRODUCTION The disposition of solvent molecules around ions is an issue of substantial chemical importance, and over the past three decades it has been the subject of great amount of experimental and theoretical investigations. In particular, the studies of anionic clusters contributed greatly to diverse fields of science such as solution chemistry, crystal growth, surface science, catalysis, biochemistry, and many others.1 The thermochemical stabilities and reactivities of these anionic clusters are particularly helpful for understanding the solvent effects that are among the key factors that determine structure−activity relationships. It is worthwhile noticing that substantial fraction of the obtained data on the anionic clusters deals with relatively small systems,2 but the fundamental knowledge can be (and in many instances it was!3−8) easily amplified to substantially larger and more practical systems. For anions, halides weakly bound to neutral solvent molecules stand out among the simplest and, at the same time, among the most interesting models of solvation insofar as they correspond to the most abundant species in the condensed phase. Halide anions form strong H-bonding, as the negatively charged Hal− is a good proton acceptor, and therefore halide clusters are commonly accepted as useful models of ion− solvent interactions that exercise a significant influence on the structural and energetic properties of electrolytic solution. Hence, it is not unusual that these clusters attract significant interest and association of halides with various donors (e.g., H2O, alcohols, or NH3) has repeatedly been investigated in the gas phase and characterized by various experimental methods and also analyzed theoretically.9−14 © 2014 American Chemical Society

Although the polar and monoprotic ROH molecules form less extensive hydrogen-bonded networks than water, alcohols represent a class of solvents for which corresponding halide clusters have received a fair amount of interest over the years, both experimentally and theoretically.15−20 However, besides the spectral data for anionic halide···alcohol clusters in the gas phase, structural information on the clusters Hal−···(HOR) in the solid state was not inspected, albeit many species featuring simultaneously halide anions and alcohols of solvation were studied by X-ray crystallography. In the framework of our ongoing project on reactions of metal-activated nitrile substrates, we developed systems that lead to generation of 1,3,5-triazapentadiene metal complexes.21−30 These cationic species possess abilities of forming crystal structures that easily incorporate solvent molecules (e.g., chloroform,31 nitromethane,32 or water29) in a way that these small molecules and their associations (e.g., halogen bonding31 or hydrogen bonding29,32) could be studied by the conventional X-ray diffraction combined with quantum-chemical methods. In particular, we obtained (1,3,5-triazapentadiene)PtII complexes bearing Hal− as counterions (Figure 1, Table 1).22 Upon crystallization of these cationic complexes from alcohols (or ROH mixtures with some other solvents) we found that the halides serve as highly efficient materials assisting cocrystallization of the metal complexes with the solvent employed and giving a variety of solid ROH solvates Received: June 24, 2014 Revised: September 5, 2014 Published: September 5, 2014 9529

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

[1][(Cl)2(MeOH)2], the Cl− ion interacts with two cations [1]2+ (Figure 2) by means of one NH···Cl− hydrogen bond [N(2)H(2M)···Cl(1)], via the hydrogen bridge with the H(9)C(9) H atom of the phenyl (Table 2, Figure 3), and one weak π(imine)···Cl(1)− [3.406(3) Å] contact with the Pt N(H)C fragment of [1]2+ (Figures 2 and 3), thus uniting stacked arrays of complexes into the layers (Figure 2). The solvated MeOH molecules occupy vacant space, giving contacts with [1]2+, and connect to the Cl− ion through a hydrogen bridge (see section Theoretical Consideration later) via the H(1O)O(1S) H atom forming, by means of the Cl−···HOMe contact, the chloride−methanol cluster (Table 2, Figure 3). Ethanol [1][(Cl)2(EtOH)2] Solvate. In the crystal structure of [1][(Cl)2(EtOH)2], the Cl− ion interacts with two cations [1]2+ by means of one NH···Cl− hydrogen bond (N(2) H(2M)···Cl(1)) (Table 3, Figure 5), and one weak π(imine)··· Cl(1)− [3.618(3) Å] contact with the PtN(H)C fragment of [1]2+ (Figures 4 and 5), thus uniting stacked arrays of complexes into the layers (Figure 4). The solvated EtOH molecules occupy vacant space, giving contacts with [1]2+, and connect to the Cl− ion through a hydrogen bridge (see section Theoretical Consideration later) via the H(1O)O(1S) H atom forming, by means of the Cl−···HOEt contact, the chloride− ethanol cluster (Table 3, Figure 5). Methanol [2][(Cl)2(MeOH)2] Solvate. In the crystal structure of [2][(Cl)2(MeOH)2], the Cl− ion interacts with three cations [1]2+ (Figure 6) by means of two N−H···Cl− hydrogen bonds, and via two hydrogen bridges with the H(11)C(11) and H(15B)C(15) H atoms of the phenyl and the NEt2 groups, respectively (Table 4, Figure 7), thus uniting stacked arrays of complexes into the layers (Figure 6). The solvated MeOH molecules occupy vacant space, giving contact with [1]2+, and connect to the Cl− ion through a hydrogen bridge (see section Theoretical Consideration later) via the H(1O)O(1S) H atom forming, by means of the Cl−···HOMe contact, the chloride− methanol cluster (Table 4, Figure 7). Methanol [4][(Cl)2(MeOH)2] and Ethanol [4][(Cl)2(EtOH)2] Solvates. In the crystal structure of [4][(Cl)2(MeOH)2] and [4][(Cl)2(EtOH)2], the Cl− ion interacts with two [4]2+ by means of a strong hydrogen bond with N(2) atom, via one hydrogen bridge with H(9)C(9) H atom of the phenyl (Tables 5 and 6; Figures 8−11), one weak π(imine)···Cl(1)− (3.444(9) Å for [4][(Cl) 2 (MeOH) 2 ] and 3.538(3) Å for [4][(Cl)2(EtOH)2]) contact with the PtN(H)C fragment of [4]2+, thus uniting stacked arrays of complexes into the layers (Figures 8 and 10). The solvated MeOH or EtOH molecules occupy vacant space, giving contact with [4]2+, and connect to the Cl− ion through a hydrogen bridge (see section Theoretical Consideration later) via the H(1O)O(1S) H atom forming, by means of the Cl−···HOR contact, the chloride−alcohol cluster (Tables 5 and 6; Figures 9 and 11). Layered Arrays Featuring Bromide···Methanol Clusters. Methanol [3][(Br)2(MeOH)2] Solvate. In the crystal structure of [3][(Br)2(MeOH)2], the Br− ion interacts with two cations [3]2+ (Figure 12) by means of one N−H···Br− hydrogen bond (N(2)H(2M)···Br(1)), via two hydrogen bridges with the H(9)C(9) and H(15A)C(15) H atoms of the phenyl and the NC5H10 groups, respectively (Table 7, Figure 13), and one weak π(imine)···Br(1)− [3.661(2) Å] contact with the PtN(H)C fragment of [3]2+ (Figures 12 and 13), thus uniting stacked arrays of complexes into the layers (Figure 12). The solvated MeOH molecules occupy vacant interlayer space, giving contacts with [3]2+, and connect to the Br− ion through

Figure 1. 1,3,5-Triazapentadiene systems.

Table 1. Alcohol Solvates of (1,3,5-Triazapentadiene)PtII Complexes

(Table 1), which then could be studied by X-ray crystallography. It is known that hydrogen bonding in solids is conventionally studied by IR33−39 or solid-state 1H NMR spectroscopies.40−53 Herein we employ another method and we report on the nature of the clusters Hal−···(HOR) thus detected in the solid state and studied by combined crystallographic and theoretical methods.



RESULTS AND DISCUSSION The observed assemblies may be considered as stacked arrays of complexes jointed by hydrogen bondings into layers. Layered Arrays Featuring Chloride···Alcohol Clusters. Methanol [1][(Cl)2(MeOH)2] Solvate. In the crystal structure of 9530

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

Figure 2. Thermal ellipsoid view of [1][(Cl)2(MeOH)2] with the atomic numbering scheme. Thermal ellipsoids are drawn with the 50% probability (left). 3D packing diagram of [1][(Cl)2(MeOH)2] along the a projection (right).

Theoretical Consideration. It is well-known that an X-ray diffraction experiment cannot indicate precise location of H atoms. Therefore, the O−H hydrogens are practically always placed at idealized positions even if a suitable electron maximum could be found from a difference Fourier map. It implies that the results of X-ray crystallography alone do not give an unambiguous answer to the question: is the halide ion bound to alcohol in the clusters [Hal−(ROH)] via the real hydrogen bonds or is this a consequence of packing effects. In the first case, the structures [1−4][(Hal)2(ROH)2] are expected to be preserved in the isolated form whereas in the second case the structures should collapse on going from the solid state to the isolated cluster. With the aim to clarify this situation, we carried out theoretical DFT (M06-2X) calculations for the isolated complex [1][(Cl)2(MeOH)2]. This allowed us to exclude the crystal packing effects from consideration and to investigate the hydrogen bondings only within this cluster. The calculations of [1][(Cl)2(MeOH)2] in the isolated state did not result in significant structural changes compared to the experimental X-ray data (Figure 14). The maximum deviation between theoretical and experimental bond lengths was found for the Pt−N bonds (0.04−0.05 Å), whereas this difference for other bonds of the metallacycles does not exceed 0.011 Å, usually falling with the 3σ interval of the experimental data. The hydrogen bond between the Cl− ion and the second sphere MeOH molecule is preserved as a result of the calculations with the parameters Cl−···H, H−O, and Cl−···O of 2.193, 0.987, and 3.137 Å, correspondingly, and the ClHO angle of 159.7°. The theoretical Cl−···O distances appear to be longer than the experimental ones by 0.088 Å and the Cl−···H−O angles are by 8° lower. The topological analysis of the electron density distribution (AIM) reveals the existence of a bond critical point (BCP) for the Cl−···HOMe interaction. The calculated ρ, ∇2ρ, and Hb values for this BCP are 0.178 e/Å3, 1.764 e/Å5, and −0.009 hartree/Å3, respectively, which are typical for the normal Hbond.54 The positive ∇2ρ and negative Hb values indicate also that the Cl−···HOMe hydrogen bond may be assigned to the medium strength type and it has some degree of covalent character.55 The calculated NBO atomic charge of the Cl− anions in the complex is −0.85 e, and the occupancies of LP[Cl−] decrease by 6−71 me upon the formation of the H-bond (Table 8). At

Table 2. Hydrogen Bonds for [1][(Cl)2(MeOH)2] [Å and deg]a D−H···A

d(H···A)

d(D···A)

∠(DHA)

N(2)−H(2M)···Cl(1)#3 O(1S)−H(1O)···Cl(1) N(2)−H(2N)···O(1S)#2 N(5)−H(5N)···O(1S) C(9)−H(9)···Cl(1)

2.42 2.22 2.15 2.26 2.78

3.163(3) 3.049(3) 2.920(4) 2.997(4) 3.662(4)

143 168 146 141 154

Symmetry transformations used to generate equivalent atoms: #1 −x, −y + 1, −z + 1; #2 −x + 1/2, y + 1/2, z; #3 −x + 1, −y + 1, −z + 1. a

Figure 3. Cl− counterion in the environment (left) and the structure of the [Cl−(MeOH)] cluster (right).

Table 3. Hydrogen Bonds for [1][(Cl)2(EtOH)2] [Å and deg]a D−H···A

d(H···A)

d(D···A)

∠(DHA)

N(2)−H(2M)···Cl(1)#2 O(1S)−H(1O)···Cl(1)#1 N(2)−H(2N)···O(1S)#3 N(5)−H(5N)···O(1S)

2.40 2.22 2.16 2.19

3.169(3) 3.042(3) 2.887(4) 2.986(4)

145 166 166 154

Symmetry transformations used to generate equivalent atoms: #1 −x, −y, −z + 1; #2 −x + 1, −y, −z + 1; #3 x + 1/2, −y − 1/2, −z + 1. a

a hydrogen bridge (see section Theoretical Consideration later) via the H(1O)O(1S) H atom forming, by means of the Br−··· HOMe contact, the bromide−methanol cluster (Table 7, Figure 13). 9531

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

Figure 4. Thermal ellipsoid view of [1][(Cl)2(EtOH)2] with the atomic numbering scheme. Thermal ellipsoids are drawn with the 50% probability (left). 3D packing diagram of [1][(Cl)2(EtOH)2] along the a projection (right).

Figure 5. Cl− counterion in the environment (left) and the structure of the [Cl−(EtOH)] cluster (right).

Figure 6. Thermal ellipsoid view of [2][(Cl)2(MeOH)2] with the atomic numbering scheme. Thermal ellipsoids are drawn with the 50% probability (left). 3D packing diagram of [2][(Cl)2(MeOH)2] along the a projection (right).

Table 4. Hydrogen Bonds for [2][(Cl)2(MeOH)2] [Å and deg]a D−H···A

d(H···A)

d(D···A)

∠(DHA)

N(2)−H(2M)···Cl(1) N(2)−H(2N)···Cl(1)#2 O(1S)−H(1O)···Cl(1) N(5)−H(5N)···O(1S)#3 C(11)−H(11)···Cl(1) C(15)−H(15B)···Cl(1)

2.37 2.48 2.25 2.12 2.84 2.73

3.2333(17) 3.2045(16) 3.0793(17) 2.908(2) 3.609(2) 3.6437(19)

168 140 172 155 138 154

Figure 7. Cl− counterion in the environment (left) and the structure of the [Cl−(MeOH)] cluster (right).

Symmetry transformations used to generate equivalent atoms: #1 −x + 2, −y + 2, −z + 1; #2 −x + 1, −y + 1, −z + 1; #3 −x + 1, −y + 2, −z + 1; #4 x + 1, y, z. a

[1]2+, and MeOH plays, conceivably, the role of a charge transmitter. Indeed, on the one hand, each Cl− ion has three other contacts with neighboring hydrogen atoms of the CH or NH bonds of the ligand. However, these contacts are only

the same time, the overall charge of the methanol molecules in [1][(Cl)2(MeOH)2] is virtually zero. These results suggest that the charge transfer occurs from the Cl− anion to the complex 9532

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

Table 5. Hydrogen Bonds for [4][(Cl)2(MeOH)2] [Å and deg]a D−H···A

d(H···A)

d(D···A)

∠(DHA)

O(1S)−H(1O)···Cl(1) N(5)−H(5)···O(1S)#1 N(2)−H(2A)···O(1S)#2 N(2)−H(2B)···Cl1#3

2.36 2.17 2.18 2.33

3.131(2) 2.956(2) 3.021(2) 3.1856(2)

156.0 152.2 155.6 169.6

Symmetry transformations used to generate equivalent atoms: #1 −x + 1, −y + 1, −z + 1; #2 −x + 2, −y + 1, −z + 1; #3 x, y, z − 1. a

Table 6. Hydrogen Bonds for [4][(Cl)2(EtOH)2] [Å and deg]a D−H···A

d(H···A)

d(D···A)

∠(DHA)

N(2)−H(2M)···Cl(1) O(1S)−H(1O)···Cl(1)#4 N(2)−H(2N)···O(1S)#2 N(5)−H(5N)···O(1S)#3 C(9)−H(9)···Cl(1)#2

2.30 2.24 2.13 2.13 2.74

3.216(3) 3.078(2) 2.931(3) 2.986(3) 3.638(3)

176.7 173.1 168.6 164.2 157.5

Figure 9. Cl− counterion in the trigonal bipyramid environment (left) and the structure of the [Cl−(MeOH)] cluster (right).

toward the O atom and the s-character of the oxygen hybrid orbital forming the H−O bond are enhanced as a result of the formation of this H-bond (Table 8). Both these effects of polarization and rehybridization should shorten the H−O bond56 but, apparently, they are not sufficient to overcome the effect of the increase of the antibonding σ*(H−O) orbital occupancy.

a Symmetry transformations used to generate equivalent atoms: #1 −x, −y, −z; #2 −x + 1, −y, −z; #3 x − 1/2, −y − 1/2, −z; #4 −x + 11/2, y − 1/2, z.



slightly shorter than the sum of the van der Waals radii of Cl and H (2.648−2.869 Å for the calculated equilibrium geometry vs 2.95 Å). On the other hand, the MeOH molecule is linked with the NH group of the ligand by a rather strong N−H···O hydrogen bond with the parameters N−H, H···O, and N···O of 1.026, 1.897, and 2.795 Å, the NHO angle of 144.2° and the ρ, ∇2ρ, and Hb values for the H···O BCP of 0.221 e/Å3, 2.431 e/ Å5, and −0.015 hartree/Å3, respectively. The second-order perturbation theory analysis revealed the overall E(2) energies for the LP(Cl−) → σ*(H−O)MeOH and LP(O)MeOH → σ*(H− N)ligand charge transfers of 20.6 and 15.8 kcal/mol, respectively (Table 8). All these data suggest that the Cl− → 1 charge transfer occurs through the Cl−···H−O(Me)···HN moiety (Figure 15). The H−O bond in the methanol molecule is elongated by 0.021 Å on going from free MeOH to [1][(Cl)2(MeOH)2]. Such an elongation may be explained by an increase of the occupancy of the antibonding σ*(H−O) orbital. The NBO analysis also shows that both polarization of the H−O NBO

FINAL REMARKS It is surprising that despite rapt attention to the Hal−···HOR hydrogen bonding, researchers specializing in this area focused all their efforts on either gas-phase or theoretical structural studies2,15−20,57−84 of halide···alcohol clusters. By contrast, no attention has been drawn to the data for more than 500 X-ray solid-state structures of these species, which, although unprocessed until this work, could be retrieved from the available databases. In particular, apart from our identification of the halide···alcohol clusters, data on numerous X-ray structures for Hal−···HOR species could be obtained from the Cambridge Structural Database. Moreover, in the X-ray structural works providing experimental data on the halide− alcohol solvates, none of crystallographers inspected the relevance of their studies to those of the Hal−···HOR hydrogen bonding and, most likely, this article is the first report of this kind.

Figure 8. Thermal ellipsoid view of [4][(Cl)2(MeOH)2] with atomic numbering scheme. Thermal ellipsoids are drawn with the 50% probability (left). 3D packing diagram of [4][(Cl)2(MeOH)2] along the a projection (right). 9533

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

Figure 10. Thermal ellipsoid view of [4][(Cl)2(EtOH)2] with atomic numbering scheme. Thermal ellipsoids are drawn with the 50% probability (left). 3D packing diagram of [4][(Cl)2(EtOH)2] along the a projection (right).

Table 7. Hydrogen Bonds for [3][(Br)2(MeOH)2] [Å and deg]a

Figure 11. Cl− counterion in the tetrahedron environment (left) and the structure of the [Cl−(EtOH)] cluster (right).

D−H···A

d(H···A)

d(D···A)

∠(DHA)

N(2)−H(2M)···Br(1)#3 O(1S)−H(1O)···Br(1) N(2)−H(2N)···O(1S)#2 N(5)−H(5N)···O(1S) C(9)−H(9)···Br(1) C(15)−H(15A)···Br(1)

2.57 2.40 2.16 2.20 2.77 2.88

3.333(2) 3.171(3) 2.962(4) 3.005(4) 3.701(3) 3.754(4)

146 153 151 152 166 147

a Symmetry transformations used to generate equivalent atoms: #1 −x, −y, −z + 1; #2 x − 1, y, z; #3 −x, −y + 1, −z + 1.

The information presented and discussed here shows a variety of halide···alcohol cluster types. Theoretical DFT calculations demonstrated that the interaction between outersphere chloride ion and methanol molecule in [1][(Cl)2(MeOH)2] may be considered as a typical hydrogen bond of the medium strength type. The charge transfer occurs from Cl− to the ligand in [1]2+, whereas MeOH plays a role as a charge transmitter. The authors hope that this article would stimulate further interest in the field of the Hal−···HOR bonding as it is anticipated that halide···alcohol clusters may play an important role for strategic crystal engineering and material design of halide-containing systems.

plexes [1][(Cl)2], [2][(Cl)2], and [3][(Br)2] crystallize as the [1−3][(Hal)2(ROH)2] (Hal = Cl, Br) alcohol solvates by slow evaporation of a ROH/CHCl3 (1:1, v/v) solutions at room temperature (RT) in air. Complex [4][(Cl)2] crystallizes as the [4][(Cl) 2 (MeOH) 2 ] methanol solvate or as the [4][(Cl)2(EtOH)2] ethanol solvate by slow evaporation of a MeOH/EtNO2 (1:2, v/v) solution at RT in air or by slow evaporation of a EtOH/ClCH2CH2Cl (1:1, v/v) solution at RT in air, respectively. X-ray Crystal Structure Determinations. [1][(Cl)2(MeOH)2], [1][(Cl)2(EtOH)2], [2][(Cl)2(MeOH)2], [4][(Cl) 2 (MeOH) 2 ], [4][(Cl) 2 (EtOH) 2 ], and [3][(Br)2(MeOH)2] were immersed in cryo-oil, mounted in a Nylon loop, and measured at a temperature of 150−100 K. The X-ray diffraction data were collected on a Bruker Kappa Apex II or Agilent Supernova diffractometers using Mo Kα radiation (λ



EXPERIMENTAL SECTION Starting Complexes and Crystal Growth. Complexes [1][(Cl)2], [2][(Cl)2], [3][(Br)2], and [4][(Cl)2] were obtained via the previously described procedures.22,32 Com-

Figure 12. Thermal ellipsoid view of [3][(Br)2(MeOH)2] with the atomic numbering scheme. Thermal ellipsoids are drawn with the 50% probability (left). 3D packing diagram of [3][(Br)2(MeOH)2] along the a projection (right). 9534

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

Figure 13. Br− counterion in the environment (left) and the structure of the [Br−(MeOH)] cluster (right).

Figure 14. Experimental X-ray (A) and equilibrium M06-2X (B) structures of [1][(Cl)2(MeOH)2].

Table 8. Characteristics of the H-Bonds in [1][(Cl)2(MeOH)2] Calculated at the M06-2X Level of Theory X···H−Y

Cl−···H−O(Me)

O···H−N

l(H−Y) l(X···H) l(X···Y) ∠XHY ρ ∇2ρ Hb Δocc. σ*[H−Y] Δocc. LP[X] E(2) LP[X] → σ*[H−Y] %, s-char. of σ[H−Y] at the Y atom in MeOH or [1]2+ in [1][(Cl)2(MeOH)2] pol. of σ[H−Y], %Y in MeOH or [1]2+ in [1][(Cl)2(MeOH)2]

0.987 Å 2.193 Å 3.137 Å 159.7° 0.178 e/Å3 1.764 e/Å5 −0.009 hartree/Å3 49 me −6 to −71 me 20.6 kcal/mol

1.026 Å 1.897 Å 2.795 Å 144.2° 0.221 e/Å3 2.431 e/Å5 −0.015 hartree/Å3 29 me −8 to −34 me 15.8 kcal/mol

20.54% 25.48%

28.92% 30.69%

73.94% 77.95%

72.14% 74.20%

Figure 15. Equilibrium structure of [1][(Cl)2(MeOH)2] with the selected Cl−···H−O(Me)···HN fragment.

= 0.710 73 Å). The Apex2,85 Denzo-Scalepack,86 EvalCCD,87 or CrysAlisPro88 program packages were used for cell refinements and data reductions. The structures were solved by direct methods using the SHELXS-9789 or Superf lip90 programs. A numerical semiempirical absorption correction (SADABS91 or CrysAlisPro88) was applied to all data. Structural refinements were carried out using the SHELXL-9789 and Olex292 graphical user interface. Molecular graphics and analyses were performed with the UCSF Chimera package.93 The NH, NH2, and OH hydrogen atoms were either located from the difference Fourier map and refined isotropically or constrained to ride on their parent atom. Other hydrogen atoms were positioned geometrically and constrained to ride on their parent atoms, with C−H

= 0.95−0.99 Å, N−H = 0.86−0.88 Å, O−H = 0.84 Å, and Uiso = 1.2−1.5 Ueq(parent atom). The crystallographic details are summarized in Table 9. Supplementary crystallographic data for these compounds have been deposited at the Cambridge Crystallographic Data Centre (CCDC 1009519−1009524) and can be obtained free of charge via www.ccdc.cam.ac.uk/data_ request/cif.



COMPUTATIONAL DETAILS The full geometry optimization of [1][(Cl)2(MeOH)2], [1]2+, and MeOH has been carried out at the DFT level of theory using the M06-2X functional94 with the help of the Gaussian9535

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

Table 9. Crystal Data empirical formula fw temp (K) λ (Å) cryst syst space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z ρcalc (Mg/m3) μ(Mo Kα) (mm−1) no. of reflns no. of unique reflns GOOF (F2) Rint R1a (I ≥ 2σ) wR2b (I ≥ 2σ) a

[1][(Cl)2(MeOH)2]

[1][(Cl)2(EtOH)2]

[2][(Cl)2(MeOH)2]

[4][(Cl)2(MeOH)2]

[4][(Cl)2(EtOH)2]

[3][(Br)2(MeOH)2]

C34H46Cl2N10O2Pt 892.80 150(2) 0.71073 orthorhombic Pbca 10.5253(5) 17.2681(8) 20.8581(10) 90 90 90 3791.0(3) 4 1.564 3.887 18458 4327 1.004 0.0651 0.0334 0.0570

C36H50Cl2N10O2Pt 920.85 123(2) 0.71073 orthorhombic Pbca 10.4214(2) 17.7041(3) 21.4018(5) 90 90 90 3948.66(14) 4 1.549 3.734 33594 1856 1.146 0.0333 0.0310 0.0554

C38H54Cl2N10O2Pt 948.90 123(2) 0.71073 triclinic P1̅ 10.0300(3) 10.1874(3) 10.6219(4) 91.285(2) 106.122(3) 102.222(2) 1015.22(5) 1 1.552 3.633 7754 4992 1.069 0.0139 0.0157 0.0380

C38H50Cl2N10O4Pt 976.87 100(2) 0.71073 monoclinic P21/c 8.8123(5) 21.8035(13) 10.7171(6) 90 105.5720(10) 90 1983.6(2) 2 1.636 3.726 16376 5243 1.050 0.0169 0.0184 0.0425

C40H54Cl2N10O4Pt 1004.92 100(2) 0.71073 orthorhombic Pbca 10.7643(2) 17.7840(2) 22.2899(4) 90 90 90 4267.01(12) 4 1.564 3.466 72658 4886 1.091 0.0674 0.0301 0.0388

C40H54Br2N10O2Pt 1061.84 120(2) 0.71073 triclinic P1̅ 8.9470(9) 10.8626(11) 11.7344(12) 94.084(5) 109.329(5) 104.174(5) 1028.84(18) 1 1.714 5.400 18390 5918 1.100 0.0599 0.0280 0.0654

R1 = Σ∥F0| − |Fc∥/Σ|F0|. bwR2 = [Σ[w(F02 − Fc2)2]/Σ[w(F02)2]]1/2.

0995 program package. The calculations were carried out using a quasi-relativistic Stuttgart pseudopotential that described 60 core electrons and the appropriate contracted basis set96 for the platinum atoms and the 6-31G(d) basis set for other atoms. The M06-2X functional is much less time-consuming than the MP2 method used by us previously for the investigation of the halogen and hydrogen bondings31,32 but, at the same time, it describes reasonably the weak dispersion forces.97,98 No symmetry operations have been applied. The Hessian matrix was calculated analytically to prove the location of correct minima (no imaginary frequencies). The topological analysis of the electron density distribution with help of the AIM method of Bader99 was performed using the program AIMAll.100 The atomic charges and bond orbital nature were analyzed by using the natural bond orbital (NBO) partitioning scheme.101



and the Visby Program (Sweden) is gratefully acknowledged. All measurements, apart from the X-ray crystal structure determinations that were conducted in Jyväskylä, were performed at Center for Magnetic Resonance and Center for Chemical Analysis and Materials Research (Saint Petersburg State University).



(1) Wild, D. A.; Bieske, E. J. Infrared Investigations of Negatively Charged Complexes and Clusters. Int. Rev. Phys. Chem. 2003, 22 (1), 129−151. (2) Kovács, A.; Varga, Z. Halogen Acceptors in Hydrogen Bonding. Coord. Chem. Rev. 2006, 250 (5−6), 710−727. (3) Takashima, K.; Riveros, J. M. Gas-Phase Solvated Negative Ions. Mass Spectrom. Rev. 1998, 17 (6), 409−430 and references cited therein.. (4) Viggiano, A. A.; Arnold, S. T.; Morris, R. A. Reactions of MassSelected Cluster Ions in a Thermal Bath Gas. Int. Rev. Phys. Chem. 1998, 17 (2), 147−184 and references therein. (5) Savage, P. B.; Holmgren, S. K.; Gellman, S. H. Anion and IonPair Complexation by a Macrocyclic Phosphine Oxide Disulfoxide. J. Am. Chem. Soc. 1994, 116 (9), 4069−4070. (6) Worm, K.; Schmidtchen, F. P. Molecular Recognition of Anions by Zwitterionic Host Molecules in Water. Angew. Chem., Int. Ed. Engl. 1995, 34 (1), 65−66. (7) Tamao, K.; Hayashi, T.; Ito, Y. Anion Complexation by Bidentate Lewis Acidic Hosts, Ortho-bis(fluorosilyl) Benzenes. J. Organomet. Chem. 1996, 506 (1−2), 85−91. (8) Schmidtchen, F. P.; Berger, M. Artificial Organic Host Molecules for Anions. Chem. Rev. 1997, 97 (5), 1609−1646. (9) Mak, C. C.; Timerghazin, Q. K.; Peslherbe, G. H. Photoexcitation and Charge-Transfer-to-Solvent Relaxation Dynamics of the I− (CH3CN) Complex. J. Phys. Chem. A 2013, 117 (32), 7595−7605. (10) Wild, D. A.; Lenzer, T. Ab Initio Study of the Fluoride− Ammonia Clusters: F−−(NH3)n, n = 1−3. Phys. Chem. Chem. Phys. 2004, 6 (22), 5122−5132. (11) Wild, D. A.; Lenzer, T. Structures and Infrared Spectra of Fluoride−Hydrogen Sulfide Clusters from Ab Initio Calculations: F−− (H2S)n, n = 1−5. Phys. Chem. Chem. Phys. 2005, 7 (22), 3793−3804.

ASSOCIATED CONTENT

S Supporting Information *

Cif files of the reported structures. This material is available free of charge via the Internet at http://pubs.acs.org

■ ■

REFERENCES

AUTHOR INFORMATION

Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS P.V.G. expresses gratitude to the Scientific Council of the President of the Russian Federation (Grant MK-2286.2013.3) and the Government of Saint Petersburg for support of his studies. M.L.K. thanks Fundaçaõ para a Ciência e a Tecnologia (FCT), Portugal, for the financial support (projects PEst-OE/ QUI/UI0100/2013 and PTDC/QUI-QUI/119561/2010) and FCT and IST for the contract within the “FCT Investigator” program. V.Y.K. is much obliged to RFBR for grants 14-0393959 and 13-03-12411. Financial support from RAS Presidium Program (8P, coordinated by acad. N.T. Kuznetsov; Russia) 9536

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

(12) Wild, D. A.; Kuwata, K. T.; Wong, C.-K.; Lobo, J. D.; Deev, A.; Schindler, T. S.; Okumura, M.; Bieske, E. J. Infrared Spectra of MassSelected Br−(NH3)n and I−NH3 Clusters. J. Phys. Chem. A 2010, 114 (14), 4762−4769. (13) Robertson, W. H.; Diken, E. G.; Price, E. A.; Shin, J.-W.; Johnson, M. A. Spectroscopic Determination of the OH− Solvation Shell in the OH−·(H2O)n Clusters. Science 2003, 299 (5611), 1367− 1372. (14) Robertson, W. H.; Johnson, M. A. Molecular Aspects of Halide Ion Hydration: the Cluster Approach. Annu. Rev. Phys. Chem. 2003, 54, 173−213. (15) Bogdanov, B.; McMahon, T. B. An Ab Initio and Density Functional Theory Investigation of the Structures and Energetics of Halide Ion-Alcohol Complexes in the Gas Phase. J. Phys. Chem. A 2000, 104 (33), 7871−7880. (16) Robertson, W. H.; Karapetian, K.; Ayotte, P.; Jordan, K. D.; Johnson, M. A. Infrared Predissociation Spectroscopy of I-B· (CH3OH)n, n = 1,2: Cooperativity in Asymmetric Solvation. J. Chem. Phys. 2002, 116 (12), 4853−4857. (17) Young, R. M.; Yandell, M. A.; Neumark, D. M. Dynamics of Electron Solvation in I−(CH3OH)n clusters (4 ≤ n ≤ 11). J. Chem. Phys. 2011, 134 (12), 124311−10. (18) Morgon, N. H.; Riveros, J. M. Potential Energy Surface for the Photoelectron Spectrum of [CH3OHBr]−. J. Mol. Struct. (THEOCHEM) 2001, 539 (1−3), 135−143. (19) Corbett, C. A.; Martínez, T. J.; Lisy, J. M. Solvation of the Fluoride Anion by Methanol. J. Phys. Chem. A 2002, 106 (42), 10015− 10021. (20) Beck, J. P.; Lisy, J. M. Cooperatively Enhanced Ionic Hydrogen Bonds in Cl−(CH3OH)1−3Ar Clusters. J. Phys. Chem. A 2010, 114 (37), 10011−10015. (21) Gushchin, P. V.; Tyan, M. R.; Bokach, N. A.; Revenco, M. D.; Haukka, M.; Wang, M. J.; Lai, C. H.; Chou, P. T.; Kukushkin, V. Y. Novel Tailoring Reaction for Two Adjacent Coordinated Nitriles Giving Platinum 1,3,5-Triazapentadiene Complexes. Inorg. Chem. 2008, 47 (24), 11487−11500. (22) Gushchin, P. V.; Kuznetsov, M. L.; Haukka, M.; Wang, M.-J.; Gribanov, A. V.; Kukushkin, V. Y.; Novel, A. Reactivity Mode for Metal-Activated Dialkylcyanamide Species: Addition of N,N′-Diphenylguanidine to a cis-(R2NCN)2PtII Center Giving an EightMembered Chelated Platinaguanidine. Inorg. Chem. 2009, 48 (6), 2583−2592. (23) Bokach, N. A.; Kuznetsova, T. V.; Simanova, S. A.; Haukka, M.; Pombeiro, A. J. L.; Kukushkin, V. Y. Nitrile−Amidine Coupling at Pt(IV) and Pt(II) Centers. An Easy Entry to Imidoylamidine Complexes. Inorg. Chem. 2005, 44 (14), 5152−5160. (24) Kopylovich, M. N.; Haukka, M.; Kirillov, A. M.; Kukushkin, V. Y.; Pombeiro, A. J. L. Unsymmetrical NiII−Imidoylamidine Complexes Derived from a Novel Oxime-Mediated Single-Pot Reaction of Nitriles. Chem.Eur. J. 2007, 13 (3), 786−791. (25) Kopylovich, M. N.; Pombeiro, A. J. L.; Fischer, A.; Kloo, L.; Kukushkin, V. Y. Facile Ni(II)/Ketoxime-Mediated Conversion of Organonitriles into Imidoylamidine Ligands. Synthesis of Imidoylamidines and Acetyl Amides. Inorg. Chem. 2003, 42 (22), 7239−7248. (26) Sarova, G. H.; Bokach, N. A.; Fedorov, A. A.; Berberan-Santos, M. N.; Kukushkin, V. Y.; Huakka, M.; Fraústo da Silva, J. J. R.; Pombeiro, A. J. L. A New Family of Luminescent Compounds: Platinum(II) Imidoylamidinates Exhibiting pH-Dependent Room Temperature Luminescence. Dalton Trans. 2006, 31, 3798−3805. (27) Gushchin, P. V.; Luzyanin, K. V.; Kopylovich, M. N.; Haukka, M.; Pombeiro, A. J. L.; Kukushkin, V. Y. NiII-Mediated Coupling between Iminoisoindolinones and Nitriles Leading to Unsymmetrical 1,3,5-Triazapentadienato Complexes. Inorg. Chem. 2008, 47 (8), 3088−3094. (28) Tskhovrebov, A. G.; Bokach, N. A.; Haukka, M.; Kukushkin, V. Y. Different Routes for Amination of Platinum(II)-Bound Cyanoguanidine. Inorg. Chem. 2009, 48 (18), 8678−8688. (29) Kopylovich, M. N.; Tronova, E. A.; Haukka, M.; Kirillov, A. M.; Kukushkin, V. Y.; Fraústo da Silva, J. J. R.; Pombeiro, A. J. L.

Identification of Hexameric Water and Hybrid Water−Chloride Clusters Intercalated in the Crystal Hosts of (Imidoylamidine)nickel(II) Complexes. Eur. J. Inorg. Chem. 2007, 29, 4621−4627. (30) Kopylovich, M. N.; Luzyanin, K. V.; Haukka, M.; Pombeiro, A. J. L.; Kukushkin, V. Y. Copper-Mediated Imine−Nitrile Coupling Leading to Unsymmetric 1,3,5-Triazapentadienato Complexes Containing the Incorporated Iminoisoindolin-1-one Moiety. Dalton Trans. 2008, 38, 5220−5224. (31) Gushchin, P. V.; Starova, G. L.; Haukka, M.; Kuznetsov, M. L.; Eremenko, I. L.; Kukushkin, V. Y. Chloride-Chloroform Clusters Exhibiting Weak Hydrogen and Halogen Bondings Are Fully Characterized in the Solid State by X-ray Diffraction. Cryst. Growth Des. 2010, 10 (11), 4839−4846. (32) Gushchin, P. V.; Kuznetsov, M. L.; Wang, Q.; Karasik, A. A.; Haukka, M.; Starova, G. L.; Kukushkin, V. Y. Lasagna-Type Arrays with Halide-Nitromethane Cluster Filling. The First Recognition of the Hal−···HCH2NO2 (Hal = Cl, Br, I) Hydrogen Bonding. Dalton Trans. 2012, 41 (23), 6922−6931. (33) Sonoda, Y.; Goto, M.; Ikeda, T.; Shimoi, Y.; Hayashi, S.; Yamawaki, H.; Kanesato, M. Intermolecular CH center dot center dot center dot O Hydrogen Bonds in Formyl-Substituted Diphenylhexatriene, a [2 + 2] Photoreactive Organic Solid: Crystal structure and IR, NMR Spectroscopic Evidence. J. Mol. Struct. 2011, 1006 (1−3), 366−374. (34) Kolev, T. Influence of Intermolecular Hydrogen Bonding on IRSpectroscopic Properties of (R)-(−)-1-Phenylglycinium Hydrogen Squarate Monohydrate in Solid-State. IR-LD, Raman Spectroscopy and Theoretical Study. J. Mol. Struct. 2007, 846 (1−3), 139−146. (35) Belkova, N. V.; Epstein, L. M.; Filippov, O. A.; Shubina, E. S. IR Spectroscopy of Hydrides and Its Application to Hydrogen Bonding and Proton Transfer Studies. Spectrosc. Prop. Inorg. Organomet. Compd. 2012, 43, 1−28. (36) Belkova, N. V.; Epstein, L. M.; Shubina, E. S. Dihydrogen Bonding, Proton Transfer and Beyond: What We Can Learn from Kinetics and Thermodynamics. Eur. J. Inorg. Chem. 2010, 23, 3555− 3565. (37) Olschewski, M.; Knop, S.; Lindner, J.; Voehringer, P. From Single Hydrogen Bonds to Extended Hydrogen-Bond Wires: LowDimensional Model Systems for Vibrational Spectroscopy of Associated Liquids. Angew. Chem., Int. Ed. 2013, 52 (37), 9634−9654. (38) Kim, Y. S.; Hochstrasser, R. M. Applications of 2D IR Spectroscopy to Peptides, Proteins, and Hydrogen-Bond Dynamics. J. Phys. Chem. B 2009, 113 (24), 8231−8251. (39) DelBene, J. E. In Anharmonicities, Isotopes, and IR and NMR Properties of Hydrogen-Bonded Complexes in Isotope Effects in Chemistry and Biology; Kohen, A., Limbach, H.-H., Eds.; Taylor & Francis: Boston, 2005; pp 153−174. (40) Uldry, A. C.; Griffin, J. M.; Yates, J. R.; Perez-Torralba, M.; Maria, M. D. S.; Webber, A. L.; Beaumont, M. L. L.; Samoson, A.; Claramunt, R. M.; Pickard, C. J.; Brown, S. P. Quantifying Weak Hydrogen Bonding in Uracil and 4-Cyano-4′-Ethynylbiphenyl: A Combined Computational and Experimental Investigation of NMR Chemical Shifts in the Solid State. J. Am. Chem. Soc. 2008, 130 (3), 945−954. (41) Sharif, S.; Schagen, D.; Toney, M. D.; Limbach, H. H. Coupling of Functional Hydrogen Bonds in Pyridoxal-5′-Phosphate-Enzyme Model Systems Observed by Solid-State NMR Spectroscopy. J. Am. Chem. Soc. 2007, 129 (14), 4440−4455. (42) Chierotti, M. R.; Gobetto, R.; Nervi, C.; Bacchi, A.; Pelagatti, P.; Colombo, V.; Sironi, A. Probing Hydrogen Bond Networks in HalfSandwich Ru(II) Building Blocks by a Combined H-1 DQ CRAMPS Solid-State NMR, XRPD, and DFT Approach. Inorg. Chem. 2014, 53 (1), 139−146. (43) Rees, G. J.; Day, S. P.; Lari, A.; Howes, A. P.; Iuga, D.; Pitak, M. B.; Coles, S. J.; Threlfall, T. L.; Light, M. E.; Smith, M. E.; Quigley, D.; Wallis, J. D.; Hanna, J. V. A Multinuclear Solid State NMR, Density Functional Theory and X-Ray Diffraction Study of Hydrogen Bonding in Group I Hydrogen Dibenzoates. CrystEngComm 2013, 15 (43), 8823−8839. 9537

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

the Vapor Phase. Comparison with Chloride (Cl−) and Variation with Ligand Structure. J. Phys. Chem. 1985, 89 (25), 5552−5556. (65) Hiraoka, K.; Mizuse, S. Gas-Phase Solvation of Cl− with H2O, CH3OH, C2H5OH, I-C3H7OH, N-C3H7OH, and t-C4H9OH. Chem. Phys. 1987, 118 (3), 457−466. (66) Evans, D. H.; Keesee, R. G.; Castleman, A. W. Thermodynamics of Gas-Phase Mixed-Solvent Cluster Ions - Water and Methanol on K+ and Cl− and Comparison to Liquid Solutions. J. Phys. Chem. 1991, 95 (9), 3558−3564. (67) Rosenfeld, R. N.; Jasinski, J. M.; Brauman, J. I. Saturation Effects in the Fluence Dependence of the Infrared Photodissociation of CH3OHF−. Chem. Phys. Lett. 1980, 71 (3), 400−404. (68) Larson, J. W.; McMahon, T. B. Strong Hydrogen-Bonding in Gas-Phase Anions - An Ion-Cyclotron Resonance Determination of Fluoride Binding Energetics to Bronsted Acids from Gas-Phase Fluoride Exchange Equilibria Measurements. J. Am. Chem. Soc. 1983, 105 (10), 2944−2950. (69) Larson, J. W.; McMahon, T. B. Hydrogen-Bonding in Gas-Phase Anions - An Experimental Investigation of the Interaction between Chloride-Ion and Bronsted Acids from Ion-Cyclotron Resonance Chloride Exchange Equilibria. J. Am. Chem. Soc. 1984, 106 (3), 517− 521. (70) Moylan, C. R.; Dod, J. A.; Brauman, J. I. Electron Photodetachment Spectroscopy of Solvated Anions. A Probe of Structure and Energetics. Chem. Phys. Lett. 1985, 118 (1), 38−39. (71) Moylan, C. R.; Dodd, J. A.; Han, C.; Brauman, J. I. Spectroscopy of Reactive Intermediates: Electron Photodetachment from Solvated Anions. J. Chem. Phys. 1987, 86 (10), 5350−5357. (72) Hiraoka, K.; Yamabe, S. Solvation of Halide-Ions with CH3OH in the Gas-Phase. Int. J. Mass Spectrom. Ion Processes 1991, 109, 133− 150. (73) Wilkinson, F. E.; Szulejko, J. E.; Allison, C. E.; McMahon, T. B. Fourier Transform Ion Cyclotron Resonance Investigation of the Deuterium Isotope Effect on Gas Phase Ion/Molecule Hydrogen Bonding Interactions in AlcoholFluoride Adduct Ions. Int. J. Mass Spectrom. Ion Processes 1992, 117, 487−505. (74) Peiris, D. M.; Riveros, J. M.; Eyler, J. R. Infrared Multiple Photon Dissociation Spectra of Methanol-Attached Anions and Proton-Bound Dimer Cations. Int. J. Mass Spectrom. Ion Processes 1996, 159 (1−3), 169−183. (75) Mihalick, J. E.; Gatev, G. G.; Brauman, J. I. Electron Photodetachment Spectroscopy of Solvated Anions: RO·HF− or ROH·F−? J. Am. Chem. Soc. 1996, 118 (49), 12424−12431. (76) Tanabe, F. K. J.; Morgon, N. H.; Riveros, J. M. Relative GasPhase Bromide and Iodide Affinity of Simple Solvent Molecules Determined by FT-ICR. J. Phys. Chem. 1996, 100 (8), 2862−2866. (77) Yang, Y.; Linnert, H. V.; Riveros, J. M.; Williams, K. R.; Eyler, J. R. Photodetachment of Singly Solvated Halide Ions. J. Phys. Chem. A 1997, 101 (13), 2371−2378. (78) Wilkinson, F. E.; Peschke, M.; Szulejko, J. E.; McMahon, T. B. Deuterium Isotope Effects on Gas Phase Ion−Molecule HydrogenBonding Interactions: Alcohol-Alkoxide and Alcohol-Chloride Adduct Ions. Int. J. Mass. Spectrom. Ion Processes 1998, 175 (1−2), 225−240. (79) Cumming, J. B.; French, M. A.; Kebarle, P. Effects of Charge Delocalization on Hydrogen Bonding to Negative Ions and Solvation of Negative Ions. Substituted Phenols and Phenoxide Anions. J. Am. Chem. Soc. 1977, 99 (21), 6999−7003. (80) Berthier, G.; Savinelli, R.; Pullman, A. Theoretical Study of the Binding of the Chloride Anion to Water and Alcohols. Int. J. Quantum Chem. 1997, 63 (2), 567−574. (81) Bradforth, S. E.; Arnold, D. W.; Metz, R. B.; Weaver, A.; Neumark, D. M. Spectroscopy of the Transition-State - Hydrogen Abstraction Reactions of Fluorine. J. Phys. Chem. 1991, 95 (21), 8066−8078. (82) Wladkowski, B. D.; East, A. L. L.; Mihalick, J. E.; Allen, W. D.; Brauman, J. I. The Proton-Transfer Surface of CH3OHF−. J. Chem. Phys. 1994, 100 (3), 2058−2088.

(44) Bradley, J. P.; Pickard, C. J.; Burley, J. C.; Martin, D. R.; Hughes, L. P.; Cosgrove, S. D.; Brown, S. P. Probing Intermolecular Hydrogen Bonding in Sibenadet Hydrochloride Polymorphs by High-Resolution H-1 Double-Quantum Solid-State NMR Spectroscopy. J. Pharm. Sci. 2012, 101 (5), 1821−1830. (45) (a) Jeffrey, G. A. An Introduction to Hydrogen Bonding; Oxford University Press: Oxford, U.K., 1997. (b) Scheiner, St. Hydrogen Bonding. Theoretical Perspectives; Oxford University Press: Oxford, New York, 1999. (46) Belkova, N. V.; Shubina, E. S.; Epstein, L. M. Diverse World of Unconventional Hydrogen Bonds. Acc. Chem. Res. 2005, 38 (8), 624− 631. (47) Epstein, L. M.; Shubina, E. S. New Types of Hydrogen Bonding in Organometallic Chemistry. Coord. Chem. Rev. 2002, 231 (1−2), 165−181. (48) Belkova, N. V.; Epstein, L. M.; Shubina, E. S. Peculiarities of Hydrogen Bonding and Proton Transfer Equilibria of Organic versus Organometallic Bases. ARKIVOC 2008, IV, 120−138. (49) Aliev, A. E.; Harris, K. D. M. In Probing Hydrogen Bonding in Solids Using Solid State NMR Spectroscopy in Supramolecular Assembly via Hydrogen Bonds I; Mingos, D. M. P., Ed.; Springer: Berlin, 2004. (50) Denisov, G. S.; Mavri, J.; Sobczyk, L. In Potential Energy Shape for the Proton Motion in Hydrogen Bonds Reflected in Infrared and NMR Spectra in Hydrogen BondingNew Insights; Grabowski, S. J., Ed.; Springer: Berlin, 2006. (51) Limbach, H. H.; Denisov, G. S.; Golubev, N. S. In Hydrogen Bond Isotope Effects Studied by NMR in Isotope Effects in Chemistry and Biology; Kohen, A., Limbach, H.-H., Eds.; Taylor & Francis: Boston, 2005; pp 193−230. (52) Kolehamainen, E.; Osmialowski, B. 15N NMR Studies of Tautomerism. Int. Rev. Phys. Chem. 2012, 31 (4), 567−629. (53) Brown, S. P. Applications of High-Resolution H-1 Solid-State NMR. Solid State Nucl. Magn. Reson. 2012, 41, 1−27. (54) Grabowski, S. J. What Is the Covalency of Hydrogen Bonding? Chem. Rev. 2011, 111 (4), 2597−2625. (55) Rozas, I.; Alkorta, I.; Elguero, J. Behavior of Ylides Containing N, O, and C Atoms as Hydrogen Bond Acceptors. J. Am. Chem. Soc. 2000, 122 (45), 11154−11161. (56) Alabugin, I. V.; Manoharan, M.; Peabody, S.; Weinhold, F. Electronic Basis of Improper Hydrogen Bonding: A Subtle Balance of Hyperconjugation and Rehybridization. J. Am. Chem. Soc. 2003, 125 (19), 5973−5987. (57) Cabarcos, O. M.; Weinheimer, C. J.; Martinez, T. J.; Lisy, J. M. The Solvation of Chloride by MethanolSurface Versus Interior Cluster Ion States. J. Chem. Phys. 1999, 110 (19), 9516−9526. (58) Nielsen, S. B.; Ayotte, P.; Kelley, J. A.; Johnson, M. A. Infrared Spectroscopic Observation of the Argon Isomer Distribution in Evaporative Ensembles of I− ROH Arm (R = methyl, ethyl, isopropyl) Clusters. J. Chem. Phys. 1999, 111 (21), 9593−9599. (59) Bogdanov, B.; Peschke, M.; Tonner, D. S.; Szulejko, J. E.; McMahon, T. B. Stepwise Solvation of Halides by Alcohol Molecules in the Gas Phase. Int. J. Mass Spectrom. 1999, 185−187, 707−725 and references therein.. (60) DeTuri, V. F.; Su, M. A.; Ervin, K. M. Dynamics of Endoergic Bimolecular Proton Transfer Reactions: F− + ROH → HF + RO− (R = H, CH3, CH3CH2, (CH3)2CH, and (CH3)3C. J. Phys. Chem. A 1999, 103 (11), 1468−1479. (61) DeTuri, V. F.; Ervin, K. M. Competitive Threshold CollisionInduced Dissociation: Gas-Phase Acidities and Bond Dissociation Energies for a Series of Alcohols. J. Phys. Chem. A 1999, 103 (35), 6911−6920. (62) Yamdagni, R.; Payzant, J. D.; Kebarle, P. Solvation of Cl− and O2− with H2O, CH3OH, and CH3CN in Gas-Phase. Can. J. Chem. 1973, 51 (15), 2507−2511. (63) Caldwell, G.; Kebarle, P. Binding-Energies and Structural Effects in Halide Anion ROH and RCOOH Complexes from Gas-Phase Equilibria Measurements. J. Am. Chem. Soc. 1984, 106 (4), 967−969. (64) Sieck, L. W. Thermochemistry of Solvation of Nitrite Anion (NO2−) and Nitrobenzene Anion (C6H5NO2−) by Polar Molecules in 9538

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539

The Journal of Physical Chemistry A

Article

(83) Bergner, A.; Dolg, M.; Küchle, W.; Stoll, H.; Preuss, H. Ab-Initio Energy-Adjusted Pseudopotentials for Elements of Groups 13−17. Mol. Phys. 1993, 80 (6), 1431−1441. (84) Andzelm, J.; Huzinaga, S.; Klobukowski, M.; Radzio, E.; Sakai, Y.; Tatekawi, H. Gaussian Basis Sets for Molecular Calculations; Elsevier: Amsterdam, 1984. (85) Bruker, A. X. S. APEX2 - Software Suite for Crystallographic Programs; Bruker AXS, Inc.: Madison, WI, USA, 2009. (86) Otwinowski, Z.; Minor, W. Processing of X-ray Diffraction Data Collected in Oscillation Mode. In Methods in Enzymology; Macromolecular Crystallography, Part A; Carter, C. W., Sweet, J., Eds.; Academic Press: New York, 1997; Vol. 276, pp 307−326. (87) Duisenberg, A. J. M.; Kroon-Batenburg, L. M. J.; Schreus, A. M. M. An Intensity Evaluation Method: EVAL-14. J. Appl. Crystallogr. 2003, 36 (2), 220−229. (88) Agilent, CrysAlisPro; Agilent Technologies Inc.: Yarnton, Oxfordshire, England, 2013. (89) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr. A 2008, 64 (1), 112−122. (90) Palatinus, L.; Chapuis, G. SUPERFLIP− a Computer Program for the Solution of Crystal Structures by Charge Flipping in Arbitrary Dimensions. J. Appl. Crystallogr. 2007, 40 (4), 786−790. (91) Sheldrick, G. M. SADABS - Bruker AXS Scaling and Absorption Correction; Bruker AXS, Inc.: Madison, WI, USA, 2008. (92) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42 (2), 339−341. (93) Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. UCSF ChimeraA Visualization System for Exploratory Research and Analysis. J. Comput. Chem. 2004, 25 (13), 1605−1612. (94) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120 (1−3), 215−241. (95) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian 09, Revision A.01; Gaussian, Inc.: Wallingford, CT, 2009. (96) Andrae, D.; Haeussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-Adjusted Abinitio Pseudopotentials for the 2nd and 3rd Row Transition-Elements. Theor. Chim. Acta 1990, 77 (2), 123−141. (97) Burns, L. A.; Várquez-Mayagoitia, Á .; Sumpter, B. G.; Sherill, C. D. Density-Functional Approaches to Noncovalent Interactions: A Comparison of Dispersion Corrections (DFT-D), Exchange-Hole Dipole Moment (XDM) Theory, and Specialized Functionals. J. Chem. Phys. 2011, 134 (8), 084107−25. (98) Sumimoto, M.; Kawashima, Y.; Yokogawa, D.; Hori, K.; Fujimoto, H. Influences of Dispersion and Long-Range Corrections on Molecular Structures of Three Types of Lithium Phthalocyanine Dimer. Int. J. Quantum Chem. 2013, 113 (3), 272−276. (99) Bader, R. F. W. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, U.K., 1990. (100) Keith, T. A.; Gristmill, T. K. AIMAll (Version 13.01.27); TK Gristmill Softward: Overland Park, KS, USA, 2013 (aim.tkgristmill. com). (101) Reed, A. E.; Curtiss, L. A.; Weinhold, F. Intermolecular Interactions from a Natural Bond Orbital, Donor-Acceptor Viewpoint. Chem. Rev. 1988, 88 (6), 899−926.

9539

dx.doi.org/10.1021/jp506256a | J. Phys. Chem. A 2014, 118, 9529−9539