P O Moiety as an Ambidextrous Hydrogen Bond Acceptor - The

Dec 28, 2017 - For example, the electron density function at the P═O moiety of diphenylphosphonic acid in the crystal state suggests that this moiet...
3 downloads 11 Views 3MB Size
Article pubs.acs.org/JPCC

Cite This: J. Phys. Chem. C 2018, 122, 1711−1720

PO Moiety as an Ambidextrous Hydrogen Bond Acceptor Elena Yu. Tupikina,† Michael Bodensteiner,‡ Peter M. Tolstoy,§ Gleb S. Denisov,† and Ilya G. Shenderovich*,‡ †

Institute of Physics, Saint-Petersburg State University, Ulianovskaya st. 1, 198504 St. Petersburg, Russian Federation Faculty of Chemistry and Pharmacy, University of Regensburg, Universitätsstr. 31, 93040 Regensburg, Germany § Center for Magnetic Resonance, Saint-Petersburg State University, Universitetsky pr. 26, 198504 St. Petersburg, Russian Federation ‡

S Supporting Information *

ABSTRACT: Hydrogen bond patterns of crystals of phosphinic, phosphonic, and phosphoric acids and their cocrystals with phosphine oxides were studied using 31P NMR and single-crystal X-ray diffraction. Two main factors govern these patterns and favor or prevent the formation of cocrystals. The first one is a high proton-accepting ability of the PO moiety in these acids. As a result, this moiety effectively competes with other proton acceptors for hydrogen bonding. For example, this moiety is a stronger proton acceptor than the CO moiety of carboxylic acids. The second factor is the inclination of the PO moiety of both the acids and the oxides to form two hydrogen bonds at once. The peculiarity of these bonds is that they weaken each other to a little degree only. In order to highlight this point, we are using the term “ambidextrous”. These two features should govern the interactions of PO moiety with water and other proton donors and acceptors in molecular clusters, the active sites of enzymes, soft matter, and at surfaces.



INTRODUCTION Higher energy and directionality are two factors that differentiate conventional B···H−A hydrogen bonds (H-bonds) from van der Waals interactions, where A and B stand for oxygen, nitrogen, or a halogen. For these reasons, it is rarely a problem to predict the configuration of a simple H-bonded complex in the gas or liquid state and the effects of solvent polarity and H/ D isotope substitution on geometry of such complexes. The situation is different in the solid state. H-bonding remains critical for the structural features of crystalline and amorphous materials.1,2 At the same time, the effect of weaker but abundant van der Waals interactions cannot be ignored anymore. Although these weak interactions cannot compete with conventional H-bonds, they impact on the stability of different H-bonded patterns and the crystal structure as a whole. Among the most amazing examples are nuclear quantum effects on the H/D isotopic polymorphism in the complex of pentachlorophenol with 4-methylpyridine3−5 and on the ligand−receptor recognition.6−8 The latter is responsible for specific activity of deuterated drugs that just entered clinical practice.9 A much less studied phenomenon is the potential ability of some proton-accepting groups to form several Hbonds at once. For example, the electron density function at the PO moiety of diphenylphosphonic acid in the crystal state suggests that this moiety has a potential to form up to three hydrogen bonds at once.10,11 There is a number of publications that specifically highlight the ability of the PO moiety to form two hydrogen bonds at once.12,13 At the same time, it remains unknown how these conjugated hydrogen bonds affect each other. Here we purposefully avoid the term “bifurcated hydrogen bond” because its meaning has to be specified in each specific case.14 © 2017 American Chemical Society

The situation is especially complex for amorphous materials for which the morphology cannot be evaluated by X-ray diffraction (XRD). However, in some of these materials the morphology can be successfully studied using solid-state nuclear magnetic resonance (NMR) spectroscopy. The only requirement of this method is the presence in the given material of NMR active nuclei whose NMR parameters depend on intermolecular interactions. Depending on the chemical composition of the system under study, these nuclei usually are 1 H, 2H, 13C, 15N, 19F, and 31P. The 1H and 2H nuclei are rarely the best choice seeing a small range of chemical shifts. The use of 13C and 15N NMR often requires isotope labeling. In contrast, 19F and 31P are blameless. The objective of this work was to evaluate the general structural patterns in the cocrystals of phosphinic (R2POOH), phosphonic (RPO(OH)2), and phosphoric (PO(OH)3) acids with phosphine oxides (R3PO). We aimed also to learn about the factors that favor or prevent the formation of a certain pattern. To facilitate our task, the composition and morphology were initially studied using 31P NMR and then, when it was possible, were confirmed by XRD analysis. Phosphinic, phosphonic, and phosphoric acids represent three different types of H-bond pattern. These H-bonds are perturbed by interaction with phosphine oxides that results in the change of H-bond pattern. This choice provides two independent 31P spin-labels that are sensitive to morphology. In Figure 1 are shown substances that have been used as model systems in this work. As the model acids, we used Received: November 15, 2017 Revised: December 28, 2017 Published: December 28, 2017 1711

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C

of cyclic dimers R22(8). It is an open question whether the anticooperative effect between these two bonds is large or small. However, it is reasonable to expect that incorporation of a proton acceptor can be energetically favorable.25 We are not aware about the crystal structure of A5. What is known is that A5 forms with B1 a crystal whose structure contained an H-bonded acid dimer terminated by four molecules of the base.26



EXPERIMENTAL SECTION

All chemicals were commercial products and used without additional purification. The solid-state 31P NMR measurements were performed on an Infinityplus spectrometer system (Agilent) operated at 7 T, equipped with a variable-temperature Chemagnetics-Varian 6 mm pencil CPMAS probe. The 31 1 P{ H} static and MAS CP NMR spectra were recorded using a cross-polarization contact time of 1 ms; the typical 90° pulse lengths were about 4.0 μs. The spectra were indirectly referenced to H3PO4 (85% in H2O). The numerical NMR parameters have been extracted from the experimentally obtained spectra using the WSolids1 program package.27 The Gaussian 09.C.01 program package was used for the model calculations at the B3LYP/cc-pVDZ level of approximation.28 The geometry optimization was done at the standard convergence criteria. The B3LYP functional produces reasonable geometries, harmonic frequencies, and lattice energies for molecular crystals with H-bonds of different strength29−32 and does not require the use of different empirical corrections.33 X-ray diffraction data for the single crystals of A3/B1, A5/B1, and A5/B3 were collected at Rigaku Oxford Diffraction SuperNova diffractometers. Using Olex2,34 the structures were solved with the ShelXT structure solution program,35 using intrinsic phasing methods. The model was refined with version 2016/6 of ShelXL using least-squares minimization.36 The atomic coordinates, the bond lengths, the angles, and the thermal parameters have been deposited at the Cambridge Crystallographic Data Centre (CCDC) with numbers 1577504, 1577505, and 1577506.

Figure 1. Substances used as model systems in this work. Triphenylphosphine oxide (B1), trioctyl-λ5-phosphanone (B2), hexamethylphosphoric triamide (B3), dimethylphosphinic acid (A1), diphenylphosphinic acid (A2), methylphosphonic acid (A3), phenylphosphonic acid (A4), and phosphoric acid (A5).

dimethylphosphinic acid (A1), diphenylphosphinic acid (A2), methylphosphonic acid (A3), phenylphosphonic acid (A4), and phosphoric acid (A5). As the model phosphine oxides, we used triphenylphosphine oxide (B1), trioctyl-λ5-phosphanone (trioctylphosphine oxide, B2), and hexamethylphosphoric triamide (B3). The gas-phase proton affinities of these bases are about 910, 950, and 960 kJ/mol, respectively.15 They cover the range from weak to strong bases.16 B1 (the melting point (mp) ≈ 430 K) is an effective crystallization aid; it forms hemihydrates.17−19 B2 (mp ≈ 325 K) is neither hygroscopic nor water-soluble. In contrast, B3 (mp ≈280 K) is very hygroscopic. These differences ensure that the general structural patterns governed by the interactions of PO moiety can be identified. In Figure 2 are shown H-bond crystal patterns of phosphinic and phosphonic acids whose substituents are not bulky and/or



RESULTS Structural changes are monitored in this study using parameters of the 31P NMR chemical shift tensors (CST) of the partners.37,38 These parameters are the anisotropy (Δσ), the principal components (δ11, δ22, and δ33), and the isotropic value (δiso) of CST. Δσ = δzz − (δxx + δyy)/2, where δii are the principal components of CST ordered according to their separation from δiso = (δ11 + δ22 + δ33)/3 in the following way: |δzz − δiso| ≥ |δxx − δiso| ≥ |δyy − δiso|. 31P CST parameters of polycrystalline B1, B2, A1, A2, A3, A4, and frozen B3 are collected in Table 1. These parameters are hard to measure for A5 because of its hygroscopy. Phosphinic Acids. Samples A1/B1 and A2/B1 were obtained by slow evaporation of solutions of A1/B1 and A2/ B1 equimolar mixtures in ethanol/dichloromethane. The 31 1 P{ H} CPMAS NMR spectrum of sample A1/B1 is shown in Figure 3a. This spectrum can be perfectly simulated using a superposition of the 31P{1H} CPMAS spectra of bulk A1 (Figure 3b) and B1 (Figure 3c). Similarly, the 31P{1H} CPMAS NMR spectrum of sample A2/B1 can be simulated using a superposition of the spectra of bulk A2 and B1. Sample A2/B2 was obtained by heating an equimolar mixture of A2 and B2 to 370 K for 2 h. Thus, powdered A2

Figure 2. H-bond crystal patterns of phosphinic (a) and phosphonic acids (b).

cannot participate in strong specific interactions. These patterns exhibit one general feature, namely, H-bonded chains.20−22 However, the structures of these chains are different for these two types of acids. In the case of phosphinic acids there are two H-bonds per molecule, and each POH···OP pair is bound by a linear H-bond (Figure 2a). Using the graph set approach this can be identified as a chain C11(4).23,24 Incorporation of a proton acceptor in such structure can be energetically unfavorable because of a competition with the PO moiety for the same mobile proton. In contrast, there are four H-bonds per molecule of phosphonic acid (Figure 2b). Here, each PO moiety is involved in two H-bonds, representing a chain C22(8) 1712

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C Table 1. Experimental 31P NMR CST of the Model Bases and Acids substance d

B1 B1e B2h B3h A1h A2g A3g A4h

T (K)

Δσa

δisob

δ11c

δ22c

δ33c

300 300 300 210 300 300 300 300

−195 −190 −179 −135 −84 −133 −56 −78

28.8 27.4 47.9 25.4 58.0 29.0 38.1 21.1

96 95 108 90f 105 107 66 71

91 86 108 50f 69 40 48 23

−101 −99 −71 −65f 3 −59 1 −31

a

The anisotropy of CST, ppm. bThe average value of the principal components, ppm. cThe principal components of CST, ppm. dData are taken for the orthorhombic polymorph from ref 37. eData are taken for the monoclinic polymorph from ref 37. fThe margin of error is about 5 ppm. gData are taken from ref 37. hThis work.

Figure 4. 31P{1H} CP NMR static and MAS spectra at 300 K of A3/ B1 cocrystal (a) and its characteristic structural pattern with four Hbonds (b). The numerical data stand for O···O distances.

complex of the 1:1 stoichiometry. The NMR parameters of A3 and B1 in this complex are collected in Table 2. The crystal structure of this complex is discussed in a separate section below. Table 2. Experimental 31P NMR CST of Solid Complexes Formed between the Model Bases and Acids

Figure 3. 31P{1H} CPMAS spectra at 300 K of A1/B1 (a) and A2/B2 (d) samples. Spectra simulated using NMR parameters of bulk A1 (b), B1 (c), A2 (e), and B2 (f).

was dissolved in molten B2. The mixture hardened at 300 K. The 31P{1H} CPMAS NMR spectrum of this sample is shown in Figure 3d. The spectrum can be simulated using a superposition of the 31P{1H} CPMAS spectra of bulk A2 (Figure 3e) and B2 (Figure 3f). Both the direct addition of B3 to solid A1 and A2 and evaporation of solutions of A1/B3 and A2/B3 mixtures in ethanol/dichloromethane gave rise to oily products. In the case when A2 and B3 were in the 3:1 molar ratio, A2 and B3 gave in 31 1 P{ H} CP spectra narrow peaks at 24.8 and 21.1 ppm, respectively. The mixture froze at about 200 K. The resulted spectrum could be simulated using the NMR parameters of bulk A2 and B3. Phosphonic Acids. Slow evaporation of a solution of A3/ B1 equimolar mixture in solution in ethanol gave rise to a crystalline precipitate. The 31P{1H} CP and CPMAS NMR spectra of this precipitate are shown in Figure 4a. The deconvolution of these spectra shows that A3 and B1 form a

complexa

T (K)

Δσb

δisoc

δ11d

δ22d

δ33d

A3/B1 A3/B1 A4/B1 A4/B1 A3/B2 A3/B2 A4/B2 A4/B2 A5/2B1

300 300 300 300 270 270 295 295 300

A5/2B1 A5/B2 A5/B2 A5/B3 A5/B3

300 215 215 300 300

−158 −78 −156 −63 −146 −84 −142 −90 −188 −162 −40 −114 −34 −79 −65

33.4 32.2 31.2 19.3 57.8 26.5 55.5 14.4 18.9 33.7 1.9 62.0 1.8 28.7 1.1

86 66 83 57 108 62 108 69 82 88 15 106 22 55 23

86 51 83 23 103 48 98 21 82 88 15 94 3 55 23

−72 −20 −73 −23 −38 −30 −39 −45 −107 −74 −25 −14 −21 −23 −42

a Parameters correspond to the substance labeled in bold. bThe anisotropy of CST, ppm. cThe average value of the principal components, ppm. dThe principal components of CST, ppm.

Similarly, B1 forms cocrystals of the 1:1 stoichiometry with A4. The corresponding 31P{1H} NMR parameters are collected in Table 2. We did not attempt to make an XRD measurement for this sample. Slow evaporation of a solution of A3/B2 equimolar mixture in ethanol gave rise to an oily product that hardens at 270 K. The 31P{1H} CP and CPMAS NMR spectra of this precipitate are shown in Figure 5a,b. The deconvolution of these spectra shows that A3 and B2 form a complex of the 1:1 stoichiometry (Figure 5c,d). The NMR parameters are collected in Table 2. 1713

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C

to justify a deconvolution of these static spectra into two or three 31P CSTs. We did not study the 31P NMR parameters of samples A4/B3 and 2A4/B3. Phosphoric Acid. Slow evaporation of a solution of 1:2 A5/B1 mixture in dichloromethane gave rise to a crystalline precipitate. The 31P{1H} CP NMR spectra of this precipitate are shown in Figure 6a. The deconvolution of these spectra

Figure 5. 31P{1H} CP static (a) and MAS (b) NMR spectra at 270 K of A3/B2 sample and the deconvolution of the MAS spectrum on the A3 (c) and B2 (d) components. 31P{1H} CP static (e) and MAS (f) NMR spectra at 295 K of A4/B2 sample and the deconvolution of the MAS spectrum on the A4 (g) and B2 (h) components. 31P{1H} CP MAS NMR spectra of 2A4/B2 sample at 295 K (i) and 240 K (j).

Figure 6. 31P{1H} CP static and MAS NMR spectra of A5/2B1 cocrystal (a) and its characteristic structural pattern with six H-bonds (b). The numerical data stand for O···O distances. The unit cell contains two crystallographically independent acid dimers.

shows that A5 and B1 form a complex of the 1:2 stoichiometry. The NMR parameters of A5 and two nonequivalent B1 in this complex are collected in Table 2. The crystal structure of this complex is discussed in a separate section below. Slow evaporation of solutions of 1:1 and 1:2 A5/B2 mixtures in dichloromethane gave rise to oily products. The 31P{1H} CP NMR spectra of these samples at low temperatures are shown in Figure 7. The static spectrum of A5/B2 sample (Figure 7a) can be simulated using two 31P CSTs (Figure 7b,c). The numerical parameters of these CSTs are collected in Table 2. The MAS spectrum of this sample exhibits two signals of the line width of 1 kHz (Figure 7d). The MAS spectrum of A5/2B2 sample exhibits the same two signals and a signal at ca. 48 ppm that can be attributed to bulk B2 (Figure 7e). Slow evaporation of a solution of A5/B3 equimolar mixture in dichloromethane gave rise to a crystalline precipitate. The 31 1 P{ H} CP and CPMAS NMR spectra of this precipitate are shown in Figure 8a. The deconvolution of these spectra shows that A5 and B3 form a complex of the 1:1 stoichiometry. The NMR parameters of A5 and B3 in this complex are collected in Table 2. The crystal structure of this complex is discussed in a separate section below. Model Calculations. In order to understand structural patterns of phosphine oxides H-bonding, we made a number of model calculations. The level of approximation that was used in these calculations is not sufficient to obtain realistic numeric values. Our aim was to observe a trend how these parameters

The quality of the crystals was insufficient for the XRD measurement. Sample 2A4/B2 was obtained by heating a mixture of A4 and B2 in the 2:1 molar ratio to 370 K for 2 h. After cooling to room temperature, the mixture becomes heterogeneous; there are an oily phase and a solid phase. The 31P{1H} CPMAS NMR spectrum of this sample at 295 K is shown in Figure 5i. The mixture hardened upon cooling. Its 31P{1H} CPMAS NMR spectrum at 240 K is shown in Figure 5j. These spectra cannot be simulated using the parameters of A4/B2 sample and of bulk A4 and B2. The positions of signals in these spectra depend on temperature. Mixtures of A3 with B3 and A4 with B3 are homogeneous oily products at room temperature for both the 1:1 and 2:1 molar ratios. At 300 K the static 31P{1H} NMR spectrum of sample A3/B3 exhibits two peaks of equal integrated intensity centered at 25.9 and 25.2 ppm. At 200 K the static 31P{1H} CP NMR spectrum of this sample exhibits a signal that can be roughly simulated using the following parameters: δ11 = δ22 = 55 ppm and δ33 = −34 ppm. There is only one peak at 25.3 ppm in the 31P{1H} CPMAS NMR spectrum. The line width of the peak is about 1 kHz. The NMR parameters of sample 2A3/ B3 are very similar to the former ones. At 300 K there are two peaks centered at 26.6 and 26.3 ppm whose integrated intensity ratio is about 1:2. At 200 K the NMR parameters are δ11 = δ22 = 57 ppm, δ33 = −34 ppm, and δiso = 26.7 ppm. It is not possible 1714

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C

Figure 7. 31P{1H} CP NMR static (a) and MAS (d) spectra of A5/B2 sample at 215 K. The deconvolution of the static spectrum into the B2 (b) and A5 (c) components. 31P{1H} CPMAS NMR spectrum of A5/ 2B2 sample at 230 K (e). Asterisks indicate spinning sidebands.

Figure 9. Optimized structures of model acid−base complexes: H3PO/HF (a), H3PO/2HF (b), B3/HF (c), B3/2HF (d), Coll/HF (e), and Coll/2HF (f). The PH and CH protons are omitted for clarity.

Table 3. Energetic and Structural Parameters of H-Bonds in the Model Acid−Base Complexes

a

Figure 8. 31P{1H} static and MAS CP NMR spectra of A5/B3 cocrystal (a) and its characteristic structural pattern with six H-bonds (b). The numerical data stand for O···O distances.

complex

EHBa (kJ/mol)

O···Hb (Å)

H3PO/HF H3PO/2HF B3/HF B3/2HF Coll/HF

49 87 61 113 64

1.63 1.67 1.54 1.60

H-bond energy; see text for details. bThe O···H distance.

EBase is the Hartree−Fock energy of a solitary molecule of the corresponding base, EHF is the Hartree−Fock energy of a solitary molecule of HF, and n is the number of HF molecules in the complex. H3PO and B3 behave similarly; namely, their PO groups are ready to form two H−bonds at once (Figure 9b,d). Although the formation of the second H-bond results in a decrease of the average per-H-bond interaction energy, the effect is only about 10% for both bases. It goes without saying that this decrease is associated with an increase of the length of these H-bonds and that the interaction energies are larger for B3. In contrast, Coll can form only one H-bond; the addition of the second molecule of HF changes the structure to [CollH]+···[FHF]−.40−44 XRD Structures. The crystal samples turned out to be hygroscopic and especially in the case of A5/2B1 of poor quality. A5/2B1 revealed additional non-merohedral twinning. Nevertheless, we were able to determine the structures for the three combinations: A3/B1, A5/2B1, and A5/B3. Geometrical and displacement parameter restraints were applied to

change with respect to the H−bond structure. In Figure 9 are shown optimized structures of the acid−base complexes under consideration. We have used hydrogen fluoride (HF) as the model acid because of its small size and high proton-donating ability, which exclude any other possible noncovalent interactions that can compete with the PO···H interaction under question.39 Phosphine oxides of a low and high protonaccepting ability were modeled by H3PO (Figure 9a,b) and B3 (Figure 9c,d), respectively. 2,4,6-Trimethylpyridine (collidine; Coll) was taken for comparison because its complexes with HF had been studied both experimentally and theoretically in detail (Figure 9e,f).40−44 Qualitative energetic and structural parameters of the model acid−base complexes are collected in Table 3. The parameter EHB that we associate with the total energy of H-bonding was calculated in the following way: EHB = Ecompl − EBase − nEHF, where Ecompl is the Hartree−Fock energy of the model complex, 1715

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C

are H-bonded to the COH moieties of the molecules of the other chain. Thus, there are two H-bonding chains of C(7)C(5) type. Each PO group is involved in two Hbonds, whereas CO groups remain excluded from the Hbond network. The proton-accepting ability of the CO group is low while the orientation of the lone pairs of the PO group favors to the double H-bonding.10 This pattern is present for different species of such type.57−59 In some cases the inclination of PO groups to multiple H-bonding is compensated by water.60,61 The differences in the proton-accepting properties of the CO and PO moieties should be important for specific biochemical activity of the latter.62−64 We conclude that whereas even a weak base can compete with the CO moiety of a carboxylic group for a proton, a much stronger base is needed in order to outperform the proton accepting ability of the PO moiety. The proton accepting abilities of B1 and B2 are not sufficient to perturb the H-bond network of crystalline A1 and A2. B3 is strong enough for that but it is not a good partner for cocrystallization. We assume that substituted pyridines can be a reasonable alternative, especially taking into account the possibility to use the isotropic 15N NMR chemical shift as a measure of H-bond geometry.65−69 Our qualitative model calculations support these tentative conclusions. The formation of the second H-bond by the PO group under discussion results only in a small decrease of the per-H-bond interaction energy. This decrease is smaller even than that for fluorine anion, one of the strongest H-bond acceptors.70 Thus, the anticooperativity effect for H-bonding to the PO group is very small. In order to highlight this effect, we prefer to introduce a term “ambidextrous”. The PO moiety is an ambidextrous H-bond acceptor; that is, the PO moiety is able to form two H-bonds at once without a marked reduction in the strength of each of them. Whether the onebond or the two-bond structural patterns are energetically favorable in the crystal structure under discussion depends on two factors: the per-H-bond interaction energy of the specific acid−base pair and the effect of these two structural patterns on the total energy of van der Waals interactions. The larger is the proton-accepting ability of a PO group, the larger is the perH-bond interaction energy and the larger is the probability that this group will be involved in multiple H-bonding in the crystalline state. The model calculations do not exhibit any specificity for the per-H-bond interaction energy in phosphine oxides as compared to pyridines. The gas-phase proton affinities of H3PO, B3, and Coll are about 900, 960, and 990 kJ/mol and correlate with the per-H-bond interaction energies in their complexes with HF that are about 49, 61, and 64 kJ/ mol, respectively. In addition to the aforementioned, the ambidextrous nature of PO moieties should affect the structure of H-bonded complexes in the gas and liquid states. For example, when a silica surface is functionalized with propionic acid moieties, the effective acidity of the surface can be manipulated by adding or removing of a small amount of water.71 In contrast, residual water cannot be removed from a silica surface functionalized with phosphonic acid moieties even at 420 K in high vacuum.72 It would be of great interest to compare the structures of small acid−base clusters of pyridines and phosphine oxides. While the former are studied in many details, the latter are rather terra incognita.48,73−76 Of special interest in this context are compounds that contain PO and POH moieties; such compounds can be effective NMR sensors of noncovalent interactions in complex systems.77−80

disordered parts of the structures. Wherever possible bridging hydrogen positions were located from the difference Fourier map and refined without geometrical restraints. The structural pattern of A3/B1 crystal exhibits a cyclic dimer of A3 terminated at each end by a B1 molecule, representing a motif D11(2) R22(8) D33(11) (Figure 4b). In A5/ 2B1 crystal cyclic dimers of A5 are terminated at the both ends by four B1 molecules, a motif D11(2) R22(8) D33(11) D22(7) (Figure 6b). The simplified structural pattern of A5/B3 cocrystal exhibits an alternating chain of cyclic dimers of A5 and two molecules of B3, a motif D11(2) R22(8) D33(11) D22(7) D22(6) D12(3) R24(12) (Figure 8b). The PO group of each B3 is involved in two H-bonds with two cyclic dimers of A3.



DISCUSSION The solid samples A1/B1, A2/B1, and A2/B2 represent the physical mixtures of the components. Oily mixtures A1/B3 and A2/B3 are converting to the physical mixtures of the components upon freezing. The latter happens about 80 K below the melting point of B3. Thus, phosphinic acids A1 and A2 interact with phosphine oxides in the liquid state but do not form cocrystals with them. In this sense, phosphinic acids deviate from benzoic acids. The latter form cocrystals of 1:1 and 2:1 acid:base composition with a large variety of bases including phosphine oxides.45−47 The structural units of these cocrystals resemble H-bonded complexes in solution.48−50 Let us compare H-bond features of carboxylic and phosphinic acids. The O···O distance in a crystal of 2-nitrobenzoic acid (pKa ≈ 2.2) is about 2.66 Å.51 For phosphonic acids the O···O distances are 2.50 and 2.45−2.48 Å for A3 (pKa ≈ 2.9) and A4 (pKa ≈ 2.3), respectively.16,20,21,52 The pKa’s of these acids are similar. However, in the case of phosphinic acids the H-bonds are shorter and presumably stronger. Bulk benzoic acids, including 2-nitrobenzoic acid, tend to aggregate in cyclic dimers while phosphinic acids prefer chains. However, this cannot be the reason for the difference. There are phosphinic acids that aggregate in cyclic dimers due to steric reasons. The O···O distances in these dimers are about 2.50 Å or shorter as well.53−55,43,44 Some ideas can be extracted from the XRD structure of 3-(hydroxy(phenyl)phosphoryl)propanoic acid.56 In Figure 10 is shown a scheme of the H-bond network in this structure. The POH and PO moieties of neighboring molecules form an H-bond chain. Besides, the PO moieties

Figure 10. Structure of a polymorph of 3-(hydroxy(phenyl)phosphoryl)propanoic acid.56 The numerical data stand for the O··· O distances of COH···OP and POH···OP H-bonds. 1716

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C

(i) The structural patterns of H-bond networks formed by phosphinic acids are governed by a high proton-accepting ability of their own PO moiety and its inclination to double H-bonding. Thereby, these H-bond networks differ from ones formed by carboxylic acids. None of the phosphine oxides studied in this work form cocrystals with phosphinic acids at room temperature. However, the formation of cocrystals is not excluded for phosphine oxides exhibiting the proton-accepting ability larger than that of B2, for example, similar to the protonaccepting ability of B3. Alternatively, phosphinic acids could form cocrystals with highly basic substituted pyridines. (ii) The structural pattern of H-bond networks in the cocrystals of phosphonic acids with phosphine oxides exhibits a cyclic dimer of the acid terminated at the both ends by the molecules of the oxide, a motif D(3) R22(8) D(3). However, if the acid is in excess, it does not form the second crystalline phase of bulk acid. Instead, the components form an oily product that becomes glassy upon cooling and represents a solid solution of the base in the acid. (iii) The structural pattern of H-bond networks in the cocrystals of phosphoric acid with phosphine oxides depends on the basicity of the oxide. For B1 the cyclic dimer of the acid is terminated from the both ends by four B1 molecules. This cluster represents an isolated structural unit, a motif D(7) R22(8) D(7). There are only van der Waals interactions between these units. In contrast, PO groups of B2 or B3 form multiple H-bonds that connect the cyclic dimer of the acid into a chain. This H-bond network goes through the whole cocrystal, a chain of R24(12) and R22(8).

Phosphonic acids A3 and A4 are packed in the crystal in a chain arrangement with a double H-bonding at the PO moieties (Figure 2b). This is a weak point that can be attacked by a base. Indeed, the structural pattern in cocrystal A3/B1 is the cyclic dimer of A3 terminated at the both ends by a B1 molecule (Figure 4b). The NMR parameters of cocrystal A4/ B1 leave no doubt that its crystal structure is very similar. Most probably, the same is true for cocrystals A3/B2 and A4/B2. However, the NMR parameters of sample 2A4/B2 suggest that all these cocrystals can be unstable and change to a glassy products if the acid is in excess. This trend is confirmed by the NMR parameters of samples A3/B3 and 2A3/B3. These oily products become glassy at low temperature. Here, a crystal is not formed even at the equimolar ratio, A3/B3. Phosphoric acid forms cocrystals with all studied phosphine oxides. However, the structural patterns are different. In the cocrystal with B1 the cyclic dimer of A5 is terminated at the both ends by four B1 molecules (Figure 6b). This cluster does not form H-bonds with the neighboring ones. In contrast, in the cocrystal with B3 the cyclic dimer of A5 is terminated at the both ends by four B3 molecules that form H-bonds with the next cyclic dimer (Figure 8b). Thus, there are 1D H-bond bands of alternating cyclic dimer and B3 pairs. As a result, the composition of the cocrystal changes to 1:1. The same composition was observed for cocrystal A5/B2. Thus, this crystal structure should be similar to the one of A5/B3. We come explicitly to the question about a competition between the total energies of H-bonds and van der Waals interactions. The best approach to answering this question is the topological analysis of charge density distribution function (ρ(r)) based on Bader’s “atoms in molecules” (AIM) theory.81 According to AIM, the bond critical point (bcp) and the bond path connecting the corresponding atoms are indicative of a bonding interaction, and the topological parameters of ρ(r) reflect the mechanism of this interaction and its strength. The experimental distribution of ρ(r) can be derived from highresolution X-ray diffraction data.82 The energy of the interactions can be estimated using the so-called Espinosa’s correlation scheme.83,84 Although the numerical accuracy of such estimates presumably depends on the bond energy,85,86 the qualitative validity of this approach has been demonstrated for hydrogen bonds and weak noncovalent interactions.87−90 However, such analysis requires crystals of an extremely high quality. We did not have such crystals. Alternatively, periodic quantum chemical calculations can be of aid here.91,92 The progress here has been reviewed recently.93 Nevertheless, the reported structures support the conclusion that the increase in energy of H-bonding interaction results in higher probability of double H-bonding of PO moieties in the crystalline state.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.7b11299. Table S1 (PDF) X-ray diffraction structure for single crystal A3/B1 (CIF) X-ray diffraction structure for single crystal A5/2B1 (CIF) X-ray diffraction structure for single crystal A5/B3 (CIF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]; phone +49-9419434027 (I.G.S.). ORCID

Elena Yu. Tupikina: 0000-0002-0998-8348 Ilya G. Shenderovich: 0000-0001-6713-9080



Notes

The authors declare no competing financial interest.

CONCLUSIONS In this work, we have studied interaction of phosphinic, phosphonic, and phosphoric acids with phosphine oxides in order to evaluate the general structural patterns of their cocrystals and to learn about the factors that favor or prevent the formation of a certain pattern. Our main conclusion is that these patterns are affected by the inclination of PO moiety to form two H-bonds at once without a marked reduction in the strength of each of them. We say that PO moiety is an ambidextrous hydrogen bond acceptor. The larger is the proton-accepting ability of this moiety in the molecule under discussion, the higher is the probability of double H-bonding. We also reach the following specific conclusions:



ACKNOWLEDGMENTS We thank Prof. Dr. Sc. Konstantin A. Lyssenko for helpful discussions. This work was supported by the Russian Foundation of Basic Research (Project 17-03-00590) and the German−Russian Interdisciplinary Science Center (G-RISC) funded by the German Federal Foreign Office via the German Academic Exchange Service (DAAD). The authors gratefully acknowledge the Gauss Centre for Supercomputing e.V. for funding this project by providing computing time on the GCS Supercomputer SuperMUC at Leibniz Supercomputing Centre (LRZ). 1717

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C



(21) Siasios, G.; Tiekink, E. R. T. Crystal Structure of Diphenylphosphinic Acid (Redetermination at 173 K), C12H11O2P. Z. Kristallogr. - Cryst. Mater. 1994, 209, 547. (22) Mehring, M.; Schurmann, M.; Ludwig, R. tert-Butylphosphonic Acid: From the Bulk to the Gas Phase. Chem. - Eur. J. 2003, 9, 837− 849. (23) Bernstein, J.; Davis, R. E.; Shimoni, L.; Chang, N.-L. Patterns in Hydrogen Bonding: Functionality and Graph Set Analysis in Crystals. Angew. Chem., Int. Ed. Engl. 1995, 34, 1555−1573. (24) Etter, M. C. Encoding and Decoding Hydrogen-Bond Patterns of Organic Compounds. Acc. Chem. Res. 1990, 23, 120−126. (25) Borissova, A. O.; Lyssenko, K. A.; Gurinov, A. A.; Shenderovich, I. G. Energy Analysis of Competing Non-Covalent Interaction in 1:1 and 1:2 Adducts of Collidine with Benzoic Acids by Means of X-Ray Diffraction. Z. Phys. Chem. 2013, 227, 775−790. (26) Chekhlov, A. N. Synthesis and Crystal Structure of the 2:1 Adduct of Triphenylphosphine Oxide and Phosphoric Acid. Russ. J. Inorg. Chem. 2005, 50, 1009−1013. (27) Eichele, K. WSolids1 ver. 1.20.20, Universität Tübingen, Tübingen, 2013. (28) 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 C.1; Gaussian, Inc.: Wallingford, CT, 2009. (29) King, M. D.; Buchanan, W. D.; Korter, T. M. Investigating the Anharmonicity of Lattice Vibrations in Water-Containing Molecular Crystals through the Terahertz Spectroscopy of l-Serine Monohydrate. J. Phys. Chem. A 2010, 114, 9570−9578. (30) Manin, A. N.; Voronin, A. P.; Manin, N. G.; Vener, M. V.; Shishkina, A. V.; Lermontov, A. S.; Perlovich, G. L. Salicylamide Cocrystals: Screening, Crystal Structure, Sublimation Thermodynamics, Dissolution, and Solid-State DFT Calculations. J. Phys. Chem. B 2014, 118, 6803−6814. (31) Zhurov, V. V.; Pinkerton, A. A. Inter- and Intramolecular Interactions in Crystalline 2-Nitrobenzoic AcidAn Experimental and Theoretical QTAIM. J. Phys. Chem. A 2015, 119, 13092−13100. (32) Katsyuba, S. A.; Vener, M. V.; Zvereva, E. E.; Brandenburg, J. G. The Role of London Dispersion Interactions in Strong and Moderate Intermolecular Hydrogen Bonds in the Crystal and in the Gas Phase. Chem. Phys. Lett. 2017, 672, 124−127. (33) Cutini, M.; Civalleri, B.; Corno, M.; Orlando, R.; Brandenburg, J. G.; Maschio, L.; Ugliengo, P. Assessment of Different Quantum Mechanical Methods for the Prediction of Structure and Cohesive Energy of Molecular Crystals. J. Chem. Theory Comput. 2016, 12, 3340−3352. (34) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. A Complete Structure Solution, Refinement and Analysis Program (2009). J. Appl. Crystallogr. 2009, 42, 339−341. (35) Sheldrick, G. M. SHELXT−Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, 71, 3−8. (36) Sheldrick, G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (37) Shenderovich, I. G. Effect of Noncovalent Interactions on the 31 P Chemical Shift Tensor of Phosphine Oxides, Phosphinic, Phosphonic, and Phosphoric Acids, and Their Complexes with Lead(II). J. Phys. Chem. C 2013, 117, 26689−26702. (38) Begimova, G. U.; Tupikina, E. Yu.; Yu, V. K.; Denisov, G. S.; Bodensteiner, M.; Shenderovich, I. G. Effect of Hydrogen Bonding to Water on the 31 P Chemical Shift Tensor of Phenyl- and Trialkylphosphine Oxides and α-Amino Phosphonates. J. Phys. Chem. C 2016, 120, 8717−8729. (39) Pollice, R.; Bot, M.; Kobylianskii, I. J.; Shenderovich, I.; Chen, P. Attenuation of London Dispersion in Dichloromethane Solutions. J. Am. Chem. Soc. 2017, 139, 13126−13140. (40) Golubev, N. S.; Shenderovich, I. G.; Smirnov, S. N.; Denisov, G. S.; Limbach, H.-H. Nuclear Scalar Spin-Spin Coupling Reveals Novel Properties of Low-Barrier Hydrogen Bonds in a Polar Environment. Chem. - Eur. J. 1999, 5, 492−497.

REFERENCES

(1) Jeffrey, G. A. An Introduction to Hydrogen Bonding; Oxford University Press: Oxford, 1997. (2) Manriquez, R.; Lopez-Dellamary, F. A.; Frydel, J.; Emmler, T.; Breitzke, H.; Buntkowsky, G.; Limbach, H.-H.; Shenderovich, I. G. Solid-State NMR Studies of Aminocarboxylic Salt Bridges in l-Lysine Modified Cellulose. J. Phys. Chem. B 2009, 113, 934−940. (3) Steiner, T.; Majerz, I.; Wilson, C. C. First O−H−N Hydrogen Bond with a Centered Proton Obtained by Thermally Induced Proton Migration. Angew. Chem., Int. Ed. 2001, 40, 2651−2654. (4) Zhou, J.; Kye, Y. S.; Harbison, G. S. Isotopomeric Polymorphism. J. Am. Chem. Soc. 2004, 126, 8392−8393. (5) Ip, B. C. K.; Shenderovich, I. G.; Tolstoy, P. M.; Frydel, J.; Denisov, G. S.; Buntkowsky, G.; Limbach, H.-H. NMR Studies of Solid Pentachlorophenol-4-Methylpyridine Complexes Exhibiting Strong OHN Hydrogen Bonds: Geometric H/D Isotope Effects and Hydrogen Bond Coupling Cause Isotopic Polymorphism. J. Phys. Chem. A 2012, 116, 11370−11387. (6) Kržan, M.; Vianello, R.; Maršavelski, A.; Repič, M.; Zakšek, M.; Kotnik, K.; Fijan, E.; Mavri, J. The Quantum Nature of Drug-Receptor Interactions: Deuteration Changes Binding Affinities for Histamine Receptor Ligands. PLoS One 2016, 11, e0154002. (7) Block, E.; Jang, S.; Matsunami, H.; Sekharan, S.; Dethier, B.; Ertem, M. Z.; Gundala, S.; Pan, Y.; Li, S.; Li, Z.; et al. Implausibility of the Vibrational Theory of Olfaction. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, E2766−E2774. (8) Franco, M. I.; Turin, L.; Mershin, A.; Skoulakis, E. M. C. Molecular Vibration-Sensing Component in Drosophila Melanogaster Olfaction. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 3797−3802. (9) Gant, T. G. Using Deuterium in Drug Discovery: Leaving the Label in the Drug. J. Med. Chem. 2014, 57, 3595−3611. (10) Lyssenko, K. A.; Grintselev-Knyazev, G. V.; Antipin, M. Yu. Nature of the P−O Bond in Diphenylphosphonic Acid: Experimental Charge Density and Electron Localization Function Analysis. Mendeleev Commun. 2002, 12, 128−130. (11) Chesnut, D. B.; Savin, A. The Electron Localization Function (ELF) Description of the PO Bond in Phosphine Oxide. J. Am. Chem. Soc. 1999, 121, 2335−2336. (12) Ahmed, R.; Altieri, A.; D’Souza, D. M.; Leigh, D. A.; Mullen, K. M.; Papmeyer, M.; Slawin, A. M. Z.; Wong, J. K. Y.; Woollins, J. D. Phosphorus-Based Functional Groups as Hydrogen Bonding Templates for Rotaxane Formation. J. Am. Chem. Soc. 2011, 133, 12304− 12310. (13) Gonschorowsky, M.; Merz, K.; Driess, M. Cyclohexylbis(hydroxymethyl)phosphane: A Hydrophilic Phosphane Capable of Forming Novel Hydrogen-Bonding Networks. Eur. J. Inorg. Chem. 2006, 2006, 455−463. (14) Rozas, I.; Alkorta, I.; Elguero, J. Bifurcated Hydrogen Bonds: Three-Centered Interactions. J. Phys. Chem. A 1998, 102, 9925−9932. (15) Hunter, E. P. L.; Lias, S. G. Evaluated Gas Phase Basicities and Proton Affinities of Molecules: An Update. J. Phys. Chem. Ref. Data 1998, 27, 413−656. (16) Gurinov, A. A.; Lesnichin, S. B.; Limbach, H.-H.; Shenderovich, I. G. How Short is the Strongest Hydrogen Bond in the Proton-Bound Homodimers of Pyridine Derivatives? J. Phys. Chem. A 2014, 118, 10804−10812. (17) Etter, M. C.; Baures, P. W. Triphenylphosphine Oxide as a Crystallization Aid. J. Am. Chem. Soc. 1988, 110, 639−640. (18) Baures, P. W. Monoclinic Triphenylphosphine Oxide Hemihydrate. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1991, 47, 2715−2716. (19) Ng, S. W. A. Second Monoclinic Modification of Triphenylphosphine Oxide Hemihydrate. Acta Crystallogr., Sect. E: Struct. Rep. Online 2009, 65, o1431. (20) Fenske, D.; Mattes, R.; Lons, J.; Tebbe, K.-F. The Crystal Structure of Diphenylphosphinic Acid. Chem. Ber. 1973, 106, 1139− 1144. 1718

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C (41) Shenderovich, I. G.; Tolstoy, P. M.; Golubev, N. S.; Smirnov, S. N.; Denisov, G. S.; Limbach, H.-H. Low-Temperature NMR Studies of the Structure and Dynamics of a Novel Series of Acid−Base Complexes of HF with Collidine Exhibiting Scalar Couplings Across Hydrogen Bonds. J. Am. Chem. Soc. 2003, 125, 11710−11720. (42) Del Bene, J. E.; Perera, S. A.; Bartlett, R. J.; Yanez, M.; Mo, O.; Elguero, J.; Alkorta, I. Two-Bond 19F−15N Spin−Spin Coupling Constants (2hJF‑N) across F−H···N Hydrogen Bonds. J. Phys. Chem. A 2003, 107, 3121−3125. (43) Del Bene, J. E.; Elguero, J. Predicted Signs of One-Bond Spin− Spin Coupling Constants (1hJH‑Y) across X−H−Y Hydrogen Bonds for Complexes with Y = 15N, 17O, and 19F. J. Phys. Chem. A 2004, 108, 11762−11767. (44) Andreeva, D. V.; Ip, B.; Gurinov, A. A.; Tolstoy, P. M.; Denisov, G. S.; Shenderovich, I. G.; Limbach, H.-H. Geometrical Features of Hydrogen Bonded Complexes Involving Sterically Hindered Pyridines. J. Phys. Chem. A 2006, 110, 10872−10879. (45) Al-Farhan, K. A. Triphenylphosphine Oxide-1-Naphthoic Acid (1/1). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2004, 60, o531−o532. (46) Al-Farhan, K. A. Triphenylphosphine Oxide-3-Chloro-Benzoic Acid (1/1). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2003, 59, o179−o180. (47) Lorente, P.; Shenderovich, I. G.; Golubev, N. S.; Denisov, G. S.; Buntkowsky, G.; Limbach, H.-H. 1H/15N NMR Chemical Shielding, Dipolar 15N, 2H Coupling and Hydrogen Bond Geometry Correlations in a Novel Series of Hydrogen-Bonded Acid−Base Complexes of Collidine with Carboxylic Acids. Magn. Reson. Chem. 2001, 39, S18− S29. (48) Limbach, H.-H.; Pietrzak, M.; Sharif, S.; Tolstoy, P. M.; Shenderovich, I. G.; Smirnov, S. N.; Golubev, N. S.; Denisov, G. S. NMR Parameters and Geometries of OHN and ODN Hydrogen Bonds of Pyridine−Acid Complexes. Chem. - Eur. J. 2004, 10, 5195− 5204. (49) Tolstoy, P. M.; Smirnov, S. N.; Shenderovich, I. G.; Golubev, N. S.; Denisov, G. S.; Limbach, H.-H. NMR Studies of Solid State − Solvent and H/D Isotope Effects on Hydrogen Bond Geometries of 1:1 Complexes of Collidine with Carboxylic Acids. J. Mol. Struct. 2004, 700, 19−27. (50) Golubev, N. S.; Smirnov, S. N.; Schah-Mohammedi, P.; Shenderovich, I. G.; Denisov, G. S.; Gindin, V. A.; Limbach, H.-H. Study of Acid-Base Interaction by Means of Low-Temperature NMR Spectra. Structure Salicylic Acid Complexes. Russ. J. Gen. Chem. 1997, 67, 1082−1087. (51) Portalone, G. A. Redetermination of 2-Nitrobenzoic Acid. Acta Crystallogr., Sect. E: Struct. Rep. Online 2009, 65, o954. (52) Ioannou, P. V. Dimethylphosphinato and Dimethylarsinato Complexes of Sb(III) and Bi(III) and Their Chemistry. Monatsh. Chem. 2012, 143, 1349−1356. (53) Aslanov, L. A.; Sotman, S. S.; Rybakov, V. B.; Elepina, L. G.; Nifant’ev, E. E. The Crystal and Molecular Structure of Dicyclohexylphosphinic Acid. J. Struct. Chem. 1980, 20, 646−647. (54) Dubrovina, N. V.; Jiao, H.; Tararov, V. I.; Spannenberg, A.; Kadyrov, R.; Monsees, A.; Christiansen, A.; Borner, A. A New Access to Chiral Phospholanes. Eur. J. Org. Chem. 2006, 2006, 3412−3420. (55) Gushwa, A. F.; Belabassi, Y.; Montchamp, J. L.; Richards, A. F. A Facile Synthesis and Crystallographic Analysis of Seven Trityl Phosphorus Compounds and Two Nickel(II) Phosphine SideProducts. J. Chem. Crystallogr. 2009, 39, 337−347. (56) Hu, Q.-S.; Zhang, X.-Z.; Luo, S.-F.; Sun, Y.-H.; Du, Z.-Y. Two Polymorphs of (2-Carboxyethyl)(Phenyl)Phosphinic Acid. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2011, 67, o195−o197. (57) Feng, L.; Luck, R. L. Synthesis and Structural Characterization of a Dimolybdenum Complex Bridged by the Hydroxymethylphenylphosphinate Ligand. J. Chem. Crystallogr. 2011, 41, 1317−1322. (58) Bartczak, T. J.; Yagbasan, R. 1,2,7a-Trihydroxy-2-Methylperhydro-1-Phosphaindene 1-Oxide. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1991, 47, 1750−1752.

(59) Marsh, R. E. The Centrosymmetric-Noncentrosymmetric Ambiguity: Some More Examples. Acta Crystallogr., Sect. A: Found. Crystallogr. 1994, 50, 450−455. (60) Kaboudin, B.; Saadati, F.; Yokomatsu, T. Reaction of 1-Amino Bisphosphinic Acids with Acid Chlorides: Synthesis of Novel Cyclic 1Hydroxy-1′-amino-1, 1-bisphosphinic Acids. Synlett 2010, 2010, 1837−1840. (61) Swor, C. D.; Nell, B. P.; Zakharov, L. N.; Tyler, D. R. P,P’DiphenylEthylEnediphosphinic Acid Dehydrate. Acta Crystallogr., Sect. E: Struct. Rep. Online 2012, 68, o2456. (62) Pabis, A.; Williams, N. H.; Kamerlin, S. C. L. Simulating the Reactions of Substituted Pyridinio-N-Phosphonates with Pyridine as a Model for Biological Phosphoryl Transfer. Org. Biomol. Chem. 2017, 15, 7308−7316. (63) Barrozo, A.; Blaha-Nelson, D.; Williams, N. H.; Kamerlin, S. C. L. The Effect of Magnesium Ions on Triphosphate Hydrolysis. Pure Appl. Chem. 2017, 89, 715−727. (64) Bora, R. P.; Mills, M. J. L.; Frushicheva, M. P.; Warshel, A. On the Challenge of Exploring the Evolutionary Trajectory from Phosphotriesterase to Arylesterase Using Computer Simulations. J. Phys. Chem. B 2015, 119, 3434−3445. (65) Gurinov, A. A.; Denisov, G. S.; Borissova, A. O.; Goloveshkin, A. S.; Greindl, J.; Limbach, H.-H.; Shenderovich, I. G. NMR Study of Solvation Effect on the Geometry of Proton-Bound Homodimers of Increasing Size. J. Phys. Chem. A 2017, 121, 8697−8705. (66) Kong, S.; Borissova, A. O.; Lesnichin, S. B.; Hartl, M.; Daemen, L. L.; Eckert, J.; Antipin, M. Yu.; Shenderovich, I. G. Geometry and Spectral Properties of the Protonated Homodimer of Pyridine in the Liquid and Solid States. A Combined NMR, X-ray Diffraction and Inelastic Neutron Scattering Study. J. Phys. Chem. A 2011, 115, 8041− 8048. (67) Limbach, H.-H.; Chan-Huot, M.; Sharif, S.; Tolstoy, P. M.; Shenderovich, I. G.; Denisov, G. S.; Toney, M. D. Critical Hydrogen Bonds and Protonation States of Pyridoxal 5′-phosphate Revealed by NMR. Biochim. Biophys. Acta, Proteins Proteomics 2011, 1814, 1426− 1437. (68) Lesnichin, S. B.; Tolstoy, P. M.; Limbach, H.-H.; Shenderovich, I. G. Counteranion-Dependent Mechanisms of Intramolecular Proton Transfer in Aprotic Solution. Phys. Chem. Chem. Phys. 2010, 12, 10373−10379. (69) Shenderovich, I. G.; Buntkowsky, G.; Schreiber, A.; Gedat, E.; Sharif, S.; Albrecht, J.; Golubev, N. S.; Findenegg, G. H.; Limbach, H.H. Pyridine-15N − A Mobile NMR Sensor for Surface Acidity and Surface Defects of Mesoporous Silica. J. Phys. Chem. B 2003, 107, 11924−11939. (70) Kucherov, S. Yu.; Bureiko, S. F.; Denisov, G. S. Anticooperativity of FHF Hydrogen Bonds in Clusters of the Type F−×(HF)n, RF×(HF)n and XF×(HF)n, R = alkyl and X = H, Br, Cl. J. Mol. Struct. 2016, 1105, 246−255. (71) Gurinov, A. A.; Mauder, D.; Akcakayiran, D.; Findenegg, G. H.; Shenderovich, I. G. Does Water Affect the Acidity of Surfaces? The Proton-Donating Ability of Silanol and Carboxylic Acid Groups at Mesoporous Silica. ChemPhysChem 2012, 13, 2282−2285. (72) Mauder, D.; Akcakayiran, D.; Lesnichin, S. B.; Findenegg, G. H.; Shenderovich, I. G. Acidity of Sulfonic and Phosphonic AcidFunctionalized SBA-15 under Almost Water-Free Conditions. J. Phys. Chem. C 2009, 113, 19185−19192. (73) Hilliard, C. R.; Kharel, S.; Cluff, K. J.; Bhuvanesh, N.; Gladysz, J. A.; Blümel, J. Structures and Unexpected Dynamic Properties of Phosphine Oxides Adsorbed on Silica Surfaces. Chem. - Eur. J. 2014, 20, 17292−17295. (74) Trujillo, C.; Sanchez-Sanz, G.; Alkorta, I.; Elguero, J. Thermodynamic and Hydrogen-Bond Basicity of Phosphine Oxides: Effect of the Ring Strain. Comput. Theor. Chem. 2012, 994, 81−90. (75) Cuypers, R.; Sudhoelter, J. R.; Zuilhof, H. Hydrogen Bonding in Phosphine Oxide/Phosphate−Phenol Complexes. ChemPhysChem 2010, 11, 2230−2240. (76) Shenderovich, I. G.; Burtsev, A. P.; Denisov, G. S.; Golubev, N. S.; Limbach, H.-H. Influence of the Temperature-Dependent 1719

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720

Article

The Journal of Physical Chemistry C Dielectric Constant on the H/D Isotope Effects on the NMR Chemical Shifts and the Hydrogen Bond Geometry of the Collidine− HF Complex in CDF3/CDClF2 Solution. Magn. Reson. Chem. 2001, 39, S91−S99. (77) Bibent, N.; Charpentier, T.; Devautour-Vinot, S.; Mehdi, A.; Gaveau, P.; Henn, F.; Silly, G. Solid-State NMR Spectroscopic Studies of Propylphosphonic Acid Functionalized SBA-15 Mesoporous Silica: Characterization of Hydrogen-Bonding Interactions. Eur. J. Inorg. Chem. 2013, 2013, 2350−2361. (78) Karra, M. D.; Sutovich, K. J.; Mueller, K. T. NMR Characterization of Bronsted Acid Sites in Faujasitic Zeolites with Use of Perdeuterated Trimethylphosphine Oxide. J. Am. Chem. Soc. 2002, 124, 902−903. (79) Lee, Y. J.; Bingöl, B.; Murakhtina, T.; Sebastiani, D.; Meyer, W. H.; Wegner, G.; Spiess, H. W. High-Resolution Solid-State NMR Studies of Poly(vinyl phosphonic acid) Proton-Conducting Polymer: Molecular Structure and Proton Dynamics. J. Phys. Chem. B 2007, 111, 9711−9721. (80) Rakiewicz, E. F.; Peters, A. W.; Wormsbecher, F.; Sutovich, K. J.; Mueller, K. T. Characterization of Acid Sites in Zeolitic and Other Inorganic Systems Using Solid-State 31P NMR of the Probe Molecule Trimethylphosphine Oxide. J. Phys. Chem. B 1998, 102, 2890−2896. (81) Bader, R. F. W. Atoms In Molecules. A Quantum Theory; Clarendron Press: Oxford, 1990. (82) Hansen, N. K.; Coppens, P. Testing Aspherical Atom Refinements on Small-Molecule Data Sets. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1978, 34, 909−921. (83) Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen Bond Strengths Revealed by Topological Analyses of Experimentally Observed Electron Densities. Chem. Phys. Lett. 1998, 285, 170−173. (84) Espinosa, E.; Alkorta, I.; Rozas, I.; Elguero, J.; Molins, E. About the Evaluation of the Local Kinetic, Potential and Total Energy Densities in Closed-Shell Interactions. Chem. Phys. Lett. 2001, 336, 457−461. (85) Vener, M. V.; Manaev, A. V.; Egorova, A. N.; Tsirelson, V. G. QTAIM Study of Strong H-Bonds with the O−H···A Fragment (A = O, N) in Three-Dimensional Periodical Crystals. J. Phys. Chem. A 2007, 111, 1155−1162. (86) Vener, M. V.; Egorova, A. N.; Churakov, A. V.; Tsirelson, V. G. Intermolecular Hydrogen Bond Energies in Crystals Evaluated Using Electron Density Properties: DFT Computations with Periodic Boundary Conditions. J. Comput. Chem. 2012, 33, 2303−2309. (87) Nelyubina, Yu. V.; Antipin, M. Yu.; Lyssenko, K. A. Hydrogen Bonds between Zwitterions: Intermediate between Classical and Charge-Assisted Ones. A Case Study. J. Phys. Chem. A 2009, 113, 3615−3620. (88) Sobczyk, L.; Grabowski, S. J.; Krygowski, T. M. Interrelation between H-Bond and Pi-Electron Delocalization. Chem. Rev. 2005, 105, 3513−3560. (89) Nelyubina, Yu. V.; Glukhov, I. V.; Antipin, M. Yu.; Lyssenko, K. A. “Higher Density Does Not Mean Higher Stability” Mystery of Paracetamol Finally Unraveled. Chem. Commun. 2010, 46, 3469−3471. (90) Nelyubina, Y. V.; Lyssenko, K. A. From “Loose” to “Dense” Crystalline Phases of Calcium Carbonate through “Repulsive” Interactions: An Experimental Charge-Density Study. Chem. - Eur. J. 2012, 18, 12633−12636. (91) Dovesi, R.; Orlando, R.; Civalleri, B.; Roetti, C.; Saunders, V. R.; Zicovich-Wilson, C. M. CRYSTAL: a Computational Tool for the Ab Initio Study of the Electronic Properties of Crystals. Z. Kristallogr. Cryst. Mater. 2005, 220, 571−573. (92) Kresse, G.; Hafner, J. Ab Initio Molecular−Dynamics Simulation of the Liquid−Metal  Amorphous−Semiconductor Transition in Germanium. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 14251−14269. (93) Deringer, V. L.; George, J.; Dronskowski, R.; Englert, U. PlaneWave Density Functional Theory Meets Molecular Crystals: Thermal Ellipsoids and Intermolecular Interactions. Acc. Chem. Res. 2017, 50, 1231−1239.

1720

DOI: 10.1021/acs.jpcc.7b11299 J. Phys. Chem. C 2018, 122, 1711−1720