Next Article in Journal
Analysis of Flame Suppression Capabilities Using Low-Frequency Acoustic Waves and Frequency Sweeping Techniques
Next Article in Special Issue
Novel Coordination Mode in the Potassium Mefenamate Trihydrate Polymeric Structure
Previous Article in Journal
Identifying the Attack Sources of Botnets for a Renewable Energy Management System by Using a Revised Locust Swarm Optimisation Scheme
Previous Article in Special Issue
Solvent Effect on the Stability and Reverse Substituent Effect in Nitropurine Tautomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NMR Properties of the Cyanide Anion, a Quasisymmetric Two-Faced Hydrogen Bonding Acceptor

by
Ilya G. Shenderovich
1,2,* and
Gleb S. Denisov
2
1
Institute of Organic Chemistry, University of Regensburg, Universitaetstrasse 31, 93053 Regensburg, Germany
2
Department of Physics, St. Petersburg State University, 198504 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2021, 13(7), 1298; https://doi.org/10.3390/sym13071298
Submission received: 18 June 2021 / Revised: 15 July 2021 / Accepted: 17 July 2021 / Published: 19 July 2021

Abstract

:
The isotopically enriched cyanide anion, (13C≡15N), has a great potential as the NMR probe of non-covalent interactions. However, hydrogen cyanide is highly toxic and can decompose explosively. It is therefore desirable to be able to theoretically estimate any valuable results of certain experiments in advance in order to carry out experimental studies only for the most suitable molecular systems. We report the effect of hydrogen bonding on NMR properties of 15N≡13CH···X and 13C≡15NH···X hydrogen bonding complexes in solution, where X = 19F, 15N, and O=31P, calculated at the ωB97XD/def2tzvp and the polarizable continuum model (PCM) approximations. In many cases, the isotropic 13C and 15N chemical shieldings of the cyanide anion are not the most informative NMR properties of such complexes. Instead, the anisotropy of these chemical shieldings and the values of scalar coupling constants, including those across hydrogen bonds, can be used to characterize the geometry of such complexes in solids and solutions. 1J(15N13C) strongly correlates with the length of the N≡C bond.

1. Introduction

Hydrogen cyanide and cyanide salts are ubiquitous compounds. They can be both the starting ingredients of life [1,2] and the reason for its end [3,4,5]. The most distinctive feature of the cyanide anion is that its two ends are capable of almost equal non-covalent interactions [6]. As a result, its orientation in crystalline cyanide salts is often disordered [7,8,9,10,11]. The same feature is responsible for the polymerization of hydrogen cyanide, resulting in materials with different characteristics and properties [12,13,14,15]. Dimers [16,17,18] and linear clusters [19,20] of hydrogen cyanide, its crystal structure [21], and its aggregates with the cyanide anion [22] and other proton acceptors [23,24,25,26] were studied both experimentally and theoretically. The knowledge gained in these studies was used in practical applications [27,28,29,30,31,32].
A deeper understanding of the properties of non-covalent interactions involving the cyanide anion in condensed matter systems can be obtained using Nuclear Magnetic Resonance (NMR) spectroscopy. Although it can be done using naturally abundant (12C≡14N) [33,34], (13C≡14N) [8,35], and (12C≡15N) [36], more can be learned with isotopically enriched (13C≡15N) [37,38,39,40,41]. It appears that the isotropic 13C and 15N NMR chemical shifts of the cyanide anion change characteristically when the corresponding atom participates in covalent or non-covalent interactions and can therefore be used to study these interactions [42,43]. In this sense, the properties of these isotropic chemical shifts are similar to the 15N isotropic chemical shift of pyridines [44,45]. However, even for pyridines, non-covalent interactions of different nature can lead to similar changes in the 15N isotropic chemical shift [46]. Direct covalent or non-covalent interactions are not the only factors that can have a measurable effect on the isotropic chemical shift. The vibrational wave function of the molecule [47,48,49], solvent effects [50,51], molecular exchange [52], and the crystal field [53] are important as well. All of these factors are particularly important for the cyanide anion, which has two similar active centers for non-covalent interactions. However, it is this feature that can be used to solve the problem.
Isotropic chemical shifts are not the only NMR parameters that can be useful. In solution, the structure of hydrogen bonded complexes can be elucidated using scalar spin-spin couplings including those across hydrogen bonds [54,55,56,57,58,59]. For 12C≡15N-H···19F dissolved in a CDF3/CDF2Cl mixture, three scalar couplings were measured experimentally at 130 K: |1J(15N1H)| = 92 Hz, |2hJ(15N19F)| = 61 Hz, and |1hJ(1H19F)| = 63 Hz [60]. Even when some couplings are not available or averaged to zero by rapid molecular exchange, the 1J(13C15N) coupling can still provide important information about the qualitative structure of such complexes. In solids, the anisotropy of the chemical shielding is much more sensitive to external influences [61,62] and molecular dynamics [63,64] than the isotropic value. Consequently, (13C≡15N) is an ideal hydrogen bond acceptor for NMR studies both in solutions and in solids, because several independent parameters can be measured for the same sample under the same experimental conditions. However, hydrogen cyanide is highly toxic and can decompose explosively. It is therefore desirable to theoretically estimate any valuable results of certain experiments in advance and to carry out experimental studies only for the most suitable molecular systems.
In this work we report on a computational study of a variety of 15N≡13C-H···X and 13C≡15N-H···X hydrogen bonded complexes, where X = 19F-R, 15N-R, O=31P-R and cover a wide range of basicity. (Li15N13C)4···(1H415N)+ and (Li13C15N)4···(1H415N)+ hydrogen bonded aggregates have been used to study the effect of competing non-covalent interactions. The objective of this study was to explore the dependence of NMR parameters on the properties of hydrogen bonding. For each of these parameters, we assessed the amplitude of possible changes and the ability to use these changes to differentiate between different interactions.

2. Materials and Methods

Gaussian 09.D.01 program package was used [65]. Geometry optimizations were done in the ωB97XD/def2tzvp approximation [66,67]. The identity of minima was confirmed by the absence of imaginary vibrational frequencies. NMR parameters were calculated using the Gauge-Independent Atomic Orbital GIAO method [68] in the ωB97XD/pcJ-2 approximation [69,70]. All calculations were done using the Polarizable Continuum Model (PCM) with water as a solvent [71,72,73], unless otherwise noted. The default SCRF=PCM method was used to construct the solute cavity.
The ωB97XD functional and the pcJ-2 basis set are capable of correctly reproducing the experimental values of the chemical shielding and scalar spin-spin coupling [50,53,74]. On the contrary, both the PCM and the solvation model based on density (SMD) approximation [75] underestimate solvent effects on the geometry of hydrogen bonded complexes [51,76]. The resulting deviations can only be neglected in specific cases [77,78,79]. For polar solvents, a change in the values of the dielectric constant has a negligible effect. There are a number of approaches that can be used to properly simulate solvent effects [80,81,82,83,84,85]. None of these approaches are universal and effortlessly applicable to an arbitrary molecular system. Consequently, the geometries of the molecular complexes used below do not correspond to either the averaged or the most expected experimental geometries in a polar medium. These geometries are used only for a qualitative assessment of the NMR parameters that can be expected in solution for such complexes.
Chemical shieldings are tensor quantities. For the cyanide anion, the principal components of the 13C and 15N shielding tensors, σ11 ≤ σ22 ≤ σ33, can be easily represented in the molecular coordinate system. One of these components is parallel to the C≡N bond and the other two are normal to this bond and are equal. In this paper the isotropic value of these shielding tensors, σiso = (σ11 + σ22 + σ33)/3, and their span, Ω = σ33 − σ11, will be discussed. More detailed description can be found elsewhere [86,87,88].

3. Results and Discussion

Figure 1 shows the structures of NCH and CNH complexes with various proton acceptors and their relative energies. For most of these complexes, the NCH···X form is energetically more preferable than the CNH···X form. The opposite situation is observed for complexes with pyridine-4-amine and 1-azabicyclo[2.2.2]octane, Figure 1e,f. This result is associated with the fact that the energy of hydrogen cyanide is 65 kJ/mol lower than the energy of hydrogen isocyanide, Figure 1a. This difference is much smaller in their complexes with proton acceptors. Consequently, one can say that the binding energy of the CNH···X form is much higher [89], and this form may be stable if the carbon atom participates in another covalent or non-covalent interaction [43,90,91,92]. The analysis of the energetically most favorable structures with the hydrogen bonded cyanide anion at different conditions is beyond the scope of this study. The NMR parameters of both forms are reported below.

3.1. The Effect of Hydrogen Bonding on the NMR Parameters of NCH and CNH

The length of the N≡C bond decreases in the series (N≡C), C≡NH, N≡CH, Table 1. However, the NMR parameters of these species are mainly determined by the location of the proton, and not by this contraction. Any of these parameters can be used to distinguish between CNH and NCH. On the contrary, the difference with (NC), is large only for the atom carrying the proton. The only exception is 1J(15N13C). The absolute value of this coupling decreases almost linearly with increasing N≡C distance. The difference of σiso(15N) in (NC) and CNH is similar to that in pyridine and pyridinium [44]. 1J(15N1H) in CNH is the same as in pyridinium [93]. 2J(15N1H) and 2J(13C1H) in NCH and CNH can be measured in the absence of proton exchange.
NCH is a weak proton donor. The proton transfer only takes place in its complex with F anion, (NC)···H-F, Table 2. The N≡C distance in this complex is shorter than that in free (NC), but greater than that in CNH. All other complexes have the NC-H···X structure with the N≡C distance greater than that in free NCH, but shorter than that in free CNH. For the weakest proton acceptor studied here, the O=P group, the N≡C distances in (NC-H)2···O=P, and free NCH are very similar. Note, that in this work we use the (NC-H)2···O=P complex, Figure 1g. The O=P group tends to form two hydrogen bonds at the same time [94,95,96]. Moreover, the presence of the second bond does not affect the strength of the first one. CNH is a strong proton donor. The structure CN-H···X is present only for pyridine and the P=O group. In these CN-H···X complexes the N≡C distance is very similar to that in (NC)···H-F, Table 2. It is noteworthy that, for the same proton acceptor, the H···X distance in CN-H···X is shorter than in NC-H···X.
σiso(15N) and Ω(15N) can be used to distinguish between different CN-H···X complexes, Table 3. On the contrary, these parameters are very similar in all (CN)···H-X and NC-H···X complexes. σiso(13C) and Ω(13C) are very similar for all NC-H···X complexes. However, these parameters can be used to distinguish between the NC-H···X, (NC)···H-X, and CNHX complexes.
The geometry of NC-H···X and (NC)···H-X complexes can be quantitatively characterized using 1J(13C1H), Table 4. The long range hJ(13CX) couplings across the hydrogen bond are also large enough to be observed experimentally. The location of the mobile proton in these complexes can be proved using any scalar coupling. This is trivial for 1J(13C1H) and hJ(1HX). However, because the N≡C distances in NC-H···X are shorter than in (NC)···H-X, 1J(15N13C) ≥ 20 Hz in NC-H···X and about 10 Hz in (NC)···H-X. 2J(15N1H) can be measured in NC-H···X but not in (NC)···H-X. Surprisingly, 3hJ(15N19F) in (NC)···H-F is large.
A combination of 1J(15N1H), and 1hJ(1HX) can be used to quantitatively characterize the geometry of CN-H···X and (CN)···H-X complexes, Table 5. The long range hJ(15NX) across the hydrogen bond are large, except 3hJ(15N31P). 2J(13C1H) can be measured in CN-H···X only. 1J(15N13C) < 10 Hz is characteristic for (CN)···H-X. In CN-H···X the value of this coupling is similar to that in (NC)···H-X. 3hJ(13C19F) in (CN)···H-F is large enough to be observed experimentally.
It is worth taking a closer look at the complex of the cyanide anion with hydrogen fluoride. It is obvious that the fluoride anion in solution interacts not only with hydrogen cyanide but also with a cation. The later interaction is simulated here using lithium fluoride. For hydrogen cyanide the structure of this complex is NCH···FLi both at the gas phase and PCM approximations, Table 6. For hydrogen isocyanide the structure of this complex is CNH···FLi at the gas phase approximation and (CN)···(H-FLi)+ at the PCM approximation. At the gas phase approximation, the H···F distance in CN-H···FLi is much shorter than in NC-H···FLi. The geometry of NC-H···FLi shows that this distance changes much more than other distances when the influence of the solvent is considered.
The geometric changes cause changes of the chemical shieldings, Table 7. However, these relationships are anything but self-evident. For NC-H···FLi, the use of the PCM approximation leads to large changes in σiso(15N) and Ω(15N), although the N≡C distance changes only slightly. CN-H···FLi and (CN)···(H-FLi)+ have similar σiso(13C) and very different Ω(13C). The ambiguity of these changes suggests that correctly interpreting the geometry of hydrogen bonded complexes based on NMR data may require comparing several NMR parameters.
In contrast, the values of scalar couplings seem to be only marginally dependent on the use of the PCM approximation, Table 8. For NC-H···FLi the main changes, as expected, are observed for 1J(13C1H), 1hJ(1H19F), and 2hJ(13C19F). The change of CN-H···FLi to (CN)···(H-FLi)+ leads to strong changes in 1J(15N1H), 1hJ(1H19F), and 2hJ(15N19F). The contraction for the N…F distance in (CN)···(H-FLi)+ causes an increase of 3hJ(13C19F).
The complex of hydrogen cyanide with tetrabutylammonium fluoride is the only complex that has so far been characterized in detail experimentally by means of NMR. The experimentally measured |1J(15N1H)| = 92 Hz [60]. Consequently, this complex has the CN-H···F structure. Other coupling constants are |2hJ(15N19F)| = 61 Hz, and |1hJ(1H19F)| = 63 Hz [60]. The values of these three coupling constants are close to those calculated for CN-H···FLi in the gas phase approximation, Table 8. To a first approximation, this result again indicates that the values of these constants depend mainly on the distances and much less on considering the influence of the solvent field. However, the closeness of the experimental and calculated values should not be overestimated. Compare the expected changes in these coupling constants when going from CN-H···FLi to (CN)···(H-FLi)+, Table 8. 1J(15N1H) changes from −110 Hz to −57 Hz. The experimental value |1J(15N1H)| = 92 Hz is very reasonable. 1hJ(1H19F) changes from −72 Hz to 47 Hz. The experimental value |1hJ(1H19F)| = 63 Hz is consistent with the trend. 2hJ(15N19F) changes from −79 Hz to −153 Hz. The experimental value |2hJ(15N19F)| = 61 Hz contradicts to the trend. Similar problems were observed for the complex of pyridine with hydrogen fluoride [50].

3.2. Proton Bound Dimers of Cyanide

Three proton bound dimers of cyanide are possible: [NCHCN], [CNHCN], and [CNHNC]. Note that the most energetically favorable structure of HN2C2 is N≡C-C≡NH+. However, cyanogen does not form spontaneously in hydrogen cyanide solutions [97] and is not discussed here. Proton bound hydrogen cyanide homodimers and more complex clusters [98,99,100] are also not discussed here.
[NC···H···CN] is the energetically most favorable proton bound dimer of cyanide, Table 9. In this centrosymmetric complex, the proton is equally strongly bound to both carbon atoms. The N≡C and C···H distances are very short.
The hetero- and homodimers [CN]···H-CN and [CN]···H-NC have similar energies and are asymmetric. The reason why some proton bound homodimers are symmetric and others not was explained in the past [101,102] and recently reviewed [103]. In [CN]···H-CN, the hydrogen bond is stronger, and the position of the proton is more symmetric. In this complex the N···H distance is shorter than the H-C distance. However, the C≡N distance in the [CN] unit is similar to that in free [CN], while in the H-CN unit this distance is closer to that in free H-CN, Table 1 and Table 9. For this reason, we call the former unit in this complex a base, and the latter an acid, regardless of the distance to the proton. The geometry of [CN]···H-NC is similar to that of proton bound homodimers of pyridine [104] and its derivatives [105]. Some of the NMR parameters of this complex and its derivatives can be found elsewhere [106,107,108].
In all these complexes σiso(13C), Ω(13C), σiso(15N) and Ω(15N) are very similar and can hardly be used to characterize both the complexes and the specific units, except for H-NC in [CN]···H-NC, Table 9. Scalar couplings are more characteristic. Note how large 2hJ(13C13C) and 2hJ(13C15N) can be across these hydrogen bonds.

3.3. Solvation by the Cyanide Anions

There are examples of hydrogen bonded aggregates in which several equal units are solvating one center. For example, aggregates (FH)n…F with n ≤ 6 were extensively investigated both experimentally and theoretically [49,59,109,110]. An increase in n leads to a weakening of each of the hydrogen bonds due to the anticooperativity of their interaction. Here, this effect will be used to determine the amplitude of changes in NMR parameters that are possible for the same type of interaction.
The experimentally observed CN-H···F complex was discussed above. The energy of this complex is similar to that of NC-H···F, Table 10. For larger (CN-H)n···F and (NC-H)n···F aggregates with n = 2−4 (NC-H)n···F are energetically more favorable because hydrogen cyanide has lower energy than isocyanide. Here we are interested whether it is possible to recognize aggregates with different n by NMR. The geometry of these aggregates gradually changes with increasing n. These changes lead to changes in the NMR parameters. How large are the changes in the NMR parameters and are the trends of these changes monotonic?
In (NC-H)n···F with n = 2−4 all shielding parameters, σiso(13C), Ω(13C), σiso(15N) and Ω(15N), are similar for all n and can hardly be used to distinguish between these complexes, Table 11. In contrast, in (CN-H)n···F only for Ω(13C) the changes are small. The dependence of σiso(15N) and Ω(15N) on the N…H distance is well known. Surprisingly, in this complex, σiso(13C) also strongly depends on this distance. Note that these changes do not correlate with changes in the N≡C distance, Table 10.
The composition of (NC-H)n···F can be clearly obtained using 1J(13C1H), 1hJ(1H19F), and 2hJ(13C19F), Table 12. For (CN-H)n···F the same can be done using 1J(15N1H) and 2hJ(15N19F). The behavior of 1hJ(1H19F) is not monotonic, Table 12. The reason for this is as follows [59]. J(1H19F) is very large and positive at short H…F distances. As this distance increases, J(1H19F) decreases, goes through zero, and becomes negative. The absolute value of this coupling increases again, passes through a maximum, and then decreases to zero. In (NC-H)n···F the H…F distances are larger than in (CN-H)n···F for the same n and J(1H19F) is close to its negative maximum already at n = 2, Table 10 and Table 12. However, J(1H19F) does not depend exclusively on the H…F distance and differs in (NC-H)n···F and (CN-H)n···F for the same H…F distance. Note that the same behavior is observed for 2hJ(15N19F) in (CN-H)n···F as n changes from 1 to 3. The long range 2hJ(15N1H), 2hJ(13C1H), and 3hJ(15N19F) are similar for all n = 2−4. In contrast, 3hJ(13C19F) varies greatly with n. 1J(15N13C) cannot be used to distinguish between different n = 2−4, but this coupling is characteristic for (NC-H)n···F and (CN-H)n···F.
The cyanide anion is a quasisymmetric two-faced acceptor of interactions. How strong is the anticooperativity of such interactions? Here we compare ([NC])4···[H4N]+, (LiNC)4···[H4N]+, ([CN])4···[H4N]+, and (LiCN)4···[H4N]+ aggregates. The hydrogen bonds in ([NC])4···[H4N]+ and ([CN])4···[H4N]+ are weak. The C…H and N…H distances are large and the N≡C distances are close to that in free [C≡N], Table 13 and Table 1. The energies of these aggregates are very similar. Although the addition of terminal Li cations weakens these bonds, the effect is small. At the same time, the presence of a new interaction can change the preferred orientation of the cyanide anions, Table 13.
Surprisingly, the presence of Li has a great impact on σiso(15N) and Ω(15N), regardless of whether it interacts with nitrogen or carbon atoms, Table 14. Its impact on σiso(13C) and Ω(13C) can be measurable but is smaller.
Scalar couplings in these aggregates are small and are not easy to measure, Table 15. However, the value of 1J(15N13C) in (Li13C≡15N)4···(H415N)+ is very remarkable. It shows that this coupling increases above 15 Hz when the carbon atom participates in a strong interaction and remains below 15 Hz when such interaction is either weak or directed towards the nitrogen atom, Table 15 and Table 1.

4. Conclusions

This work reports on the geometry and NMR parameters of hydrogen cyanide and isocyanide hydrogen bonded to proton acceptors, whose ability to accept protons was distributed over a wide range. The main objective of this study was to find NMR parameters that can be used to unambiguously determine the structure of such complexes. The most important conclusions are as follows.
For weak proton acceptors, the most energetically favorable structure is NC-H···X. The NMR parameters of such structures are similar to those of free hydrogen cyanide. The most characteristic parameters of these structures are 1J(13C1H) > 200 Hz and |1J(15N13C)| > 20 Hz.
For strong proton acceptors, the most energetically favorable structure may be C≡N···H···X. The most characteristic parameters of these structures are the 15N isotropic chemical shift, the span of the 15N chemical shift tensor, and 1J(15N1H).
When the cyanide anion interacts with weak acids, both [NC]···H-X and [CN]···H-X structures are equally possible. The NMR parameters of such structures are similar to those of free cyanide anion. However, if the terminal nitrogen atom in [NC]···H-X participates in an additional interaction, its chemical shift can vary greatly.
The most energetically favorable structure of a complex of hydrogen cyanide and the cyanide anion is symmetric, [NC···H···CN], with 1hJ(13C1H) about 100 Hz.
The spin-spin coupling across the N≡C bond, 1J(15N13C) strongly correlates with the bond length and can be used to identify the structure of hydrogen bonded complexes in cases where other spin-spin couplings are averaged to zero due to proton and molecular exchange, Figure 2. Deviations from this correlation are possible only if both atoms participate in external interactions.

Author Contributions

Conceptualization, I.G.S.; methodology, I.G.S. and G.S.D.; data curation, G.S.D.; writing—original draft preparation, I.G.S.; writing—review and editing, G.S.D.; visualization, I.G.S.; supervision, I.G.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Foundation of Basic Research (Project 20-03-00231).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Acknowledgments

The authors gratefully acknowledge the Gauss Centre for Supercomputing e.V. (www.gauss-centre.eu accessed on 17 June 2021) for funding this project by providing computing time on the GCS Supercomputer SuperMUC at Leibniz Supercomputing Centre (LRZ, www.lrz.de (accessed on 19 July 2021)).

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Patel, B.H.; Percivalle, C.; Ritson, D.J.; Duffy, C.D.; Sutherland, J.D. Common origins of RNA, protein and lipid precursors in a cyanosulfidic protometabolism. Nat. Chem. 2015, 7, 301–307. [Google Scholar] [CrossRef] [Green Version]
  2. Perrin, Z.; Carrasco, N.; Chatain, A.; Jovanovic, L.; Vettier, L.; Ruscassier, N.; Cernogora, G. An Atmospheric Origin for HCN-Derived Polymers on Titan. Processes 2021, 9, 965. [Google Scholar] [CrossRef]
  3. Akimov, M.G.; Fomina-Ageeva, E.V.; Dudina, P.V.; Andreeva, L.A.; Myasoyedov, N.F.; Bezuglov, V.V. ACTH(6–9)PGP Peptide Protects SH-SY5Y Cells from H2O2, tert-Butyl Hydroperoxide, and Cyanide Cytotoxicity via Stimulation of Proliferation and Induction of Prosurvival-Related Genes. Molecules 2021, 26, 1878. [Google Scholar] [CrossRef] [PubMed]
  4. Leavesley, H.B.; Li, L.; Prabhakaran, K.; Borowitz, J.L.; Isom, G.E. Interaction of Cyanide and Nitric Oxide with Cytochrome c Oxidase: Implications for Acute Cyanide Toxicity. Tox. Sci. 2008, 101, 101–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Donato, D.B.; Nichols, O.; Possingham, H.; Moore, M.; Ricci, P.F.; Noller, B.N. A critical review of the effects of gold cyanide-bearing tailings solutions on wildlife. Environ. Int. 2007, 33, 974–984. [Google Scholar] [CrossRef]
  6. Meot-Ner, M. The ionic hydrogen bond. Chem. Rev. 2005, 105, 213–284. [Google Scholar] [CrossRef] [Green Version]
  7. Chen, J.-Y.; Yoo, C.-S. Physical and chemical transformations of sodium cyanide at high pressures. J. Chem. Phys. 2009, 131, 144507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Kuhn, N.; Eichele, K.; Steimann, M.; Al-Sheikh, A.; Doser, B.; Ochsenfeld, C. Hydrogen bonds with cyanide ions? The structures of 1,3-diisopropyl-4,5-dimethylimidazolium cyanide and 1-isopropyl-3,4,5-trimethylimidazolium cyanide. Z. Anorg. Allg. Chem. 2006, 632, 2268–2275. [Google Scholar] [CrossRef]
  9. Černák, J.; Orendáč, M.; Potočňák, I.; Chomič, J.; Orendáčová, A.; Skoršepa, J.; Feher, A. Cyanocomplexes with one-dimensional structures: Preparations, crystal structures and magnetic properties. Coord. Chem. Rev. 2002, 224, 51–66. [Google Scholar] [CrossRef]
  10. Markley, T.J.; Toby, B.H.; Pearlstein, R.M.; Ramprasad, D. New Synthesis Routes to Lithium and Cesium Cyanide Salts. Inorg. Chem. 1997, 36, 3376–3378. [Google Scholar] [CrossRef]
  11. Lely, J.A.; Bijvoet, J.M. The crystal structure of lithium cyanide. Recl. Trav. Chim. Pays Bas 1942, 61, 244–252. [Google Scholar] [CrossRef]
  12. Ruiz-Bermejo, M.; de la Fuente, J.L.; Pérez-Fernández, C.; Mateo-Martí, E. A Comprehensive Review of HCN-Derived Polymers. Processes 2021, 9, 597. [Google Scholar] [CrossRef]
  13. Fernández, A.; Ruiz-Bermejo, M.; de la Fuente, J.L. Modelling the kinetics and structural property evolution of a versatile reaction: Aqueous HCN polymerization. Phys. Chem. Chem. Phys. 2018, 20, 17353–17366. [Google Scholar] [CrossRef]
  14. Mamajanov, I.; Herzfeld, J. HCN polymers characterized by SSNMR: Solid state reaction of crystalline tetramer (diaminomaleonitrile). J. Chem. Phys. 2009, 130, 134504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Mamajanov, I.; Herzfeld, J. HCN polymers characterized by solid state NMR: Chains and sheets formed in the neat liquid. J. Chem. Phys. 2009, 130, 134503. [Google Scholar] [CrossRef] [Green Version]
  16. Bläsing, K.; Bresien, J.; Labbow, R.; Schulz, A.; Villinger, A. A Dimer of Hydrogen Cyanide Stabilized by a Lewis Acid. Angew. Chem. Int. Ed. 2018, 57, 9170–9175. [Google Scholar] [CrossRef] [PubMed]
  17. Alfredsson, M.; Ojamäe, L.; Hermansson, K.G. A comparison of Hartree—Fock, MP2, and DFT results for the HCN dimer and crystal. Int. J. Quantum Chem. 1996, 60, 767–778. [Google Scholar] [CrossRef]
  18. Buxton, L.W.; Campbell, E.J.; Flygare, W.H. The vibrational ground state rotational spectroscopic constants and structure of the HCN dimer. Chem. Phys. 1981, 56, 399–406. [Google Scholar] [CrossRef]
  19. de Oliveira, P.M.C.; Silva, J.A.; Longo, R.L. Benchmark, DFT assessments, cooperativity, and energy decomposition analysis of the hydrogen bonds in HCN/HNC oligomeric complexes. J. Mol. Model. 2017, 23, 1–10. [Google Scholar] [CrossRef]
  20. King, B.F.; Weinhold, F. Structure and spectroscopy of (HCN)n clusters: Cooperative and electronic delocalization effects in C–H⋅⋅⋅ N hydrogen bonding. J. Chem. Phys. 1995, 103, 333–347. [Google Scholar] [CrossRef]
  21. Dulmage, W.J.; Lipscomb, W.N. The crystal structures of hydrogen cyanide, HCN. Acta Cryst. 1951, 4, 330–334. [Google Scholar] [CrossRef] [Green Version]
  22. Bläsing, K.; Harloff, J.; Schulz, A.; Stoffers, A.; Stoer, P.; Villinger, A. Salts of HCN-Cyanide Aggregates: [CN(HCN)2] and [CN(HCN)3]. Angew. Chem. Int. Ed. 2020, 59, 10508–10513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Grabowski, S.J. Hydrogen Bonds with BF4 Anion as a Proton Acceptor. Crystals 2020, 10, 460. [Google Scholar] [CrossRef]
  24. Szatyłowicz, H.; Krygowski, T.M.; Fonseca Guerra, C.; Bickelhaupt, F.M. Complexes of 4-substituted phenolates with HF and HCN: Energy decomposition and electronic structure analyses of hydrogen bonding. J. Comput. Chem. 2013, 34, 696–705. [Google Scholar] [CrossRef]
  25. McDowell, S.A.; Buckingham, A.D. Cooperative and diminutive hydrogen bonding in Y⋯ HCN⋯ HCN and NCH⋯ Y⋯ HCN trimers (Y= BF, CO, N2). J. Chem. Phys. 2010, 132, 064303. [Google Scholar] [CrossRef] [PubMed]
  26. Clutter, D.R.; Thompson, W.E. Infrared spectroscopic study of polycrystalline NH4CN. J. Chem. Phys. 1969, 51, 153–159. [Google Scholar] [CrossRef]
  27. Şahin, Ö.; Özdemir, Ü.Ö.; Seferoğlu, N.; Aydıner, B.; Sarı, M.; Tunç, T.; Seferoğlu, Z. A highly selective and sensitive chemosensor derived coumarin–thiazole for colorimetric and fluorimetric detection of CN ion in DMSO and aqueous solution: Synthesis, sensing ability, Pd(II)/Pt(II) complexes and theoretical studies. Tetrahedron 2016, 72, 5843–5852. [Google Scholar] [CrossRef]
  28. Li, J.J.; Wei, W.; Qi, X.L.; Xu, X.; Liu, Y.C.; Lin, Q.H.; Dong, W. Rational design, synthesis of reaction-based dual-channel cyanide sensor in aqueous solution. Spectrochim. Acta A 2016, 152, 288–293. [Google Scholar] [CrossRef]
  29. Rybarczyk-Pirek, A.J.; Małecka, M.; Palusiak, M. Use of Quantum Theory of Atoms in Molecules in the Search for Appropriate Hydrogen Atom Locations in X-ray Diffraction Based Studies. Cryst. Growth Des. 2016, 16, 6841–6848. [Google Scholar] [CrossRef]
  30. Li, J.; Qi, X.; Wei, W.; Liu, Y.; Xu, X.; Lin, Q.; Dong, W. A “donor-two-acceptor” sensor for cyanide detection in aqueous solution. Sens. Actuators B Chem. 2015, 220, 986–991. [Google Scholar] [CrossRef]
  31. Le Questel, J.Y.; Berthelot, M.; Laurence, C. Hydrogen-bond acceptor properties of nitriles: A combined crystallographic and ab initio theoretical investigation. J. Phys. Org. Chem. 2000, 13, 347–358. [Google Scholar] [CrossRef]
  32. Herail, M.; Berthelot, M.; Proutiere, A. Prediction of hydrogen bond basicity in nitriles from dipole moments, mesomeric effects and electrostatic potentials. J. Phys. Org. Chem. 1995, 8, 421–428. [Google Scholar] [CrossRef]
  33. Bechtel, H.A.; Steeves, A.H.; Wong, B.M.; Field, R.W. Evolution of chemical bonding during HCN⇄HNC isomerization as revealed through nuclear quadrupole hyperfine structure. Angew. Chem. Int. Ed. 2008, 47, 2969–2972. [Google Scholar] [CrossRef] [PubMed]
  34. Wong, B.M. Nuclear quadrupole hyperfine structure in HC14N/H14NC and DC15N/D15NC isomerization: A diagnostic tool for characterizing vibrational localization. Phys. Chem. Chem. Phys. 2008, 10, 5599–5606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Stokes, H.T.; Ailion, D.C.; Case, T.A. Molecular reorientations in the ordered phases of KCN and NaCN studied by NMR. Phys. Rev. B 1984, 30, 4925. [Google Scholar] [CrossRef] [Green Version]
  36. Geisheimer, A.R.; Wren, J.E.; Michaelis, V.K.; Kobayashi, M.; Sakai, K.; Kroeker, S.; Leznoff, D.B. Aggregation of [Au(CN)4] Anions: Examination by Crystallography and 15N CP-MAS NMR and the Structural Factors Influencing Intermolecular Au··· N Interactions. Inorg. Chem. 2011, 50, 1265–1274. [Google Scholar] [CrossRef]
  37. Seliman, A.A.; Altaf, M.; Kawde, A.N.; Wazeer, M.I.; Isab, A.A. NMR and kinetic studies of the interactions of [Au(cis-DACH)Cl2]Cl and [Au(cis-DACH)2]Cl3 with potassium cyanide in aqueous solution. J. Coord. Chem. 2014, 67, 3431–3443. [Google Scholar] [CrossRef]
  38. He, C.; Smith, M.A. A comprehensive NMR structural study of Titan aerosol analogs: Implications for Titan’s atmospheric chemistry. Icarus 2014, 243, 31–38. [Google Scholar] [CrossRef]
  39. Werner-Zwanziger, U.; Chapman, K.W.; Zwanziger, J. Multinuclear NMR study of zinc dicyanide. Z. Phys. Chem. 2012, 226, 1205–1218. [Google Scholar] [CrossRef]
  40. Shaikh, M.N.; Al-Maythalony, B.A.; Monim-Ul-Mehboob, M.; Fettouhi, M.; Wazeer, M.I.; Isab, A.A.; Ahmad, S. Mercury (II) cyanide complexes with alkyldiamines: Solid-state/solution NMR, computational, and antimicrobial studies. J. Coord. Chem. 2012, 65, 2074–2086. [Google Scholar] [CrossRef]
  41. He, C.; Lin, G.; Upton, K.T.; Imanaka, H.; Smith, M.A. Structural investigation of HCN polymer isotopomers by solution-state multidimensional NMR. J. Phys. Chem. A 2012, 116, 4751–4759. [Google Scholar] [CrossRef]
  42. Cabral, B.J. Born-Oppenheimer molecular dynamics, hydrogen bond interactions and magnetic properties of liquid hydrogen cyanide. J. Mol. Liq. 2018, 272, 778–786. [Google Scholar] [CrossRef]
  43. Alkorta, I.; Rozas, I.; Elguero, J. Isocyanides as hydrogen bond acceptors. Theor. Chem. Acc. 1998, 99, 116–123. [Google Scholar] [CrossRef]
  44. 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 Serious of Hydrogen-Bonded Acid-Base Complexes of Collidine with Carboxylic Acids. Magn. Reson. Chem. 2001, 39, S18–S29. [Google Scholar] [CrossRef]
  45. Solum, M.S.; Altmann, K.L.; Strohmeier, M.; Berges, D.A.; Zhang, Y.; Facelli, J.C.; Pugmire, R.J.; Grant, D.M. 15N Chemical Shift Principal Values in Nitrogen Heterocycles. J. Am. Chem. Soc. 1997, 119, 9804–9809. [Google Scholar] [CrossRef]
  46. Gurinov, A.A.; Rozhkova, Y.A.; Zukal, A.; Čejka, J.; Shenderovich, I.G. Mutable Lewis and Brønsted Acidity of Aluminated SBA-15 as Revealed by NMR of Adsorbed Pyridine-15N. Langmuir 2011, 27, 12115–12123. [Google Scholar] [CrossRef]
  47. Gräfenstein, J. Efficient calculation of NMR isotopic shifts: Difference-dedicated vibrational perturbation theory. J. Chem. Phys. 2019, 151, 244120. [Google Scholar] [CrossRef]
  48. Golubev, N.S.; Melikova, S.M.; Shchepkin, D.N.; Shenderovich, I.G.; Tolstoy, P.M.; Denisov, G.S. Interpretation of H/D Isotope Effects on NMR Chemical Shifts of [FHF] Ion Based on Calculations of Nuclear Magnetic Shielding Tensor Surface. Z. Phys. Chem. 2003, 217, 1549–1563. [Google Scholar] [CrossRef]
  49. Shenderovich, I.G.; Limbach, H.-H.; Smirnov, S.N.; Tolstoy, P.M.; Denisov, G.S.; Golubev, N.S. H/D Isotope Effects on the Low-Temperature NMR Parameters and Hydrogen Bond Geometries of (FH)2F and (FH)3F Dissolved in CDF3/CDF2Cl. Phys. Chem. Chem. Phys. 2002, 4, 5488–5497. [Google Scholar] [CrossRef]
  50. Shenderovich, I.G.; Denisov, G.S. Solvent effects on acid-base complexes. What is more important: A macroscopic reaction field or solute-solvent interactions? J. Chem. Phys. 2019, 150, 204505. [Google Scholar] [CrossRef]
  51. 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. [Google Scholar] [CrossRef] [Green Version]
  52. 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. [Google Scholar] [CrossRef]
  53. Shenderovich, I.G. Electric field effect on 31P NMR magnetic shielding. J. Chem. Phys. 2020, 153, 184501. [Google Scholar] [CrossRef]
  54. Alkorta, I.; Elguero, J.; Denisov, G.S. A review with comprehensive data on experimental indirect scalar NMR spin–spin coupling constants across hydrogen bonds. Magn. Res. Chem. 2008, 46, 599–624. [Google Scholar] [CrossRef] [PubMed]
  55. Del Bene, J.E.; Elguero, J. Probing the proton-transfer coordinate of complexes with F−H…P hydrogen bonds using one-and two-bond spin–spin coupling constants. Magn. Res. Chem. 2007, 45, 714–719. [Google Scholar] [CrossRef] [PubMed]
  56. Alkorta, I.; Elguero, J. Review on DFT and ab initio Calculations of Scalar Coupling Constants. Int. J. Mol. Sci. 2003, 4, 64–92. [Google Scholar] [CrossRef]
  57. 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. [Google Scholar] [CrossRef]
  58. Dingley, A.J.; Grzesiek, S. Direct observation of hydrogen bonds in nucleic acid base pairs by internucleotide 2JNN couplings. J. Am. Chem. Soc. 1998, 120, 8293–8297. [Google Scholar] [CrossRef]
  59. Shenderovich, I.G.; Smirnov, S.N.; Denisov, G.S.; Gindin, V.A.; Golubev, N.S.; Dunger, A.; Reibke, R.; Kirpekar, S.; Malkina, O.L.; Limbach, H.-H. Nuclear Magnetic Resonance of Hydrogen Bonded Clusters between F and (HF)n: Experiment and Theory. Ber. Bunsenges. Phys. Chem. Chem. Phys. 1998, 102, 422–428. [Google Scholar] [CrossRef]
  60. Golubev, N.S.; Detering, C.; Smirnov, S.N.; Shenderovich, I.G.; Denisov, G.S.; Limbach, H.-H.; Tolstoy, P.M. H/D Isotope Effects on NMR Chemical Shifts of Nuclei Involved in a Hydrogen Bridge of Hydrogen Isocyanide Complexes with Fluoride Anion. Phys. Chem. Chem. Phys. 2009, 11, 5154–5159. [Google Scholar] [CrossRef]
  61. Begimova, G.U.; Tupikina, E.Y.; Yu, V.K.; Denisov, G.S.; Bodensteiner, M.; Shenderovich, I.G. Effect of Hydrogen Bonding to Water on the 31P Chemical Shift Tensor of Phenyl- and Trialkylphosphine Oxides and a-Amino Phosphonates. J. Phys. Chem. C 2016, 120, 8717–8729. [Google Scholar] [CrossRef]
  62. Shenderovich, I.G. Effect of Noncovalent Interactions on the 31P 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. [Google Scholar] [CrossRef]
  63. Shenderovich, I.G. For Whom a Puddle Is the Sea? Adsorption of Organic Guests on Hydrated MCM-41 Silica. Langmuir 2020, 36, 11383–11392. [Google Scholar] [CrossRef] [PubMed]
  64. 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. [Google Scholar] [CrossRef]
  65. 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 D.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  66. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  68. Cheeseman, J.R.; Trucks, G.W.; Keith, T.A.; Frisch, M.J. A Comparison of Models for Calculating Nuclear Magnetic Resonance Shielding Tensors. J. Chem. Phys. 1996, 104, 5497–5509. [Google Scholar] [CrossRef]
  69. Deng, W.; Cheeseman, J.R.; Frisch, M.J. Calculation of Nuclear Spin-Spin Coupling Constants of Molecules with First and Second Row Atoms in Study of Basis Set Dependence. J. Chem. Theory Comput. 2006, 2, 1028–1037. [Google Scholar] [CrossRef]
  70. Jensen, F. Segmented contracted basis sets optimized for nuclear magnetic shielding. J. Chem. Theory Comput. 2015, 11, 132–138. [Google Scholar] [CrossRef] [PubMed]
  71. Cossi, M.; Barone, V.; Cammi, R.; Tomasi, J. Ab initio study of solvated molecules: A new implementation of the polarizable continuum model. Chem. Phys. Lett. 1996, 255, 327–335. [Google Scholar] [CrossRef]
  72. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  73. Scalmani, G.; Frisch, M.J. Continuous surface charge polarizable continuum models of solvation. I. General formalism. J. Chem. Phys. 2010, 132, 114110. [Google Scholar] [CrossRef]
  74. Shenderovich, I.G. Experimentally Established Benchmark Calculations of 31P NMR Quantities. Chem. Methods 2021, 1, 61–70. [Google Scholar] [CrossRef]
  75. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  76. Shenderovich, I.G. Simplified Calculation Approaches Designed to Reproduce the Geometry of Hydrogen Bonds in Molecular Complexes in Aprotic Solvents. J. Chem. Phys. 2018, 148, 124313. [Google Scholar] [CrossRef] [PubMed]
  77. Filarowski, A.; Koll, A. Integrated intensity of OH absorption bands in bent hydrogen bonds in ortho-dialkylaminomethyl phenols. Vibr. Spectrosc. 1996, 12, 15–24. [Google Scholar] [CrossRef]
  78. Pápai, I. Theoretical study of the Cu(H2O) and Cu(NH3) complexes and their photolysis products. J. Chem. Phys. 1995, 103, 1860–1870. [Google Scholar] [CrossRef]
  79. Borisenko, V.E.; Filarovski, A.I. The electrooptical parameters of aniline and its halogen derivatives in hydrogen bonded complexes. J. Mol. Struct. 1989, 196, 353–370. [Google Scholar] [CrossRef]
  80. Hok, L.; Mavri, J.; Vianello, R. The Effect of Deuteration on the H2 Receptor Histamine Binding Profile: A Computational Insight into Modified Hydrogen Bonding Interactions. Molecules 2020, 25, 6017. [Google Scholar] [CrossRef] [PubMed]
  81. Shenderovich, I.G.; Denisov, G.S. Adduct under Field—A Qualitative Approach to Account for Solvent Effect on Hydrogen Bonding. Molecules 2020, 25, 436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Dominikowska, J.; Palusiak, M. Tuning Aromaticity of para-Substituted Benzene Derivatives with an External Electric Field. ChemPhysChem 2018, 19, 590–595. [Google Scholar] [CrossRef] [PubMed]
  83. Pylaeva, S.A.; Elgabarty, H.; Sebastiani, D.; Tolstoy, P.M. Symmetry and Dynamics of FHF Anion in Vacuum, in CD2Cl2 and in CCl4. Ab Initio MD Study of Fluctuating Solvent–Solute Hydrogen and Halogen Bonds. Phys. Chem. Chem. Phys. 2017, 19, 26107–26120. [Google Scholar] [CrossRef] [Green Version]
  84. Balevicius, V.; Maršalka, A.; Klimavičius, V.; Dagys, L.; Gdaniec, M.; Svoboda, I.; Fuess, H. NMR and XRD study of hydrogen bonding in picolinic acid N-oxide in crystalline state and solutions: Media and temperature effects on potential energy surface. J. Phys. Chem. A 2018, 122, 6894–6902. [Google Scholar] [CrossRef] [PubMed]
  85. Bora, R.P.; Mills, M.J.; 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. [Google Scholar] [CrossRef] [PubMed]
  86. Shenderovich, I.G.; Limbach, H.-H. Solid State NMR for Nonexperts: An Overview of Simple but General Practical Methods. Solids 2021, 2, 139–154. [Google Scholar] [CrossRef]
  87. Bryce, D.L.; Bernard, G.M.; Gee, M.; Lumsden, M.D.; Eichele, K.; Wasylishen, R.E. Practical Aspects of Modern Routine Solid-State Multinuclear Magnetic Resonance Spectroscopy: One-Dimensional Experiments. Can. J. Anal. Sci. Spectrosc. 2001, 46, 46–82. [Google Scholar] [CrossRef]
  88. Duer, M.J. (Ed.) Solid-State NMR Spectroscopy. Principles and Applications; Blackwell Science Ltd.: Oxford, UK, 2002. [Google Scholar]
  89. Shahi, A.; Arunan, E. Hydrogen bonding, halogen bonding and lithium bonding: An atoms in molecules and natural bond orbital perspective towards conservation of total bond order, inter-and intra-molecular bonding. Phys. Chem. Chem. Phys. 2014, 16, 22935–22952. [Google Scholar] [CrossRef]
  90. Ramabhadran, R.O.; Hua, Y.; Flood, A.H.; Raghavachari, K. C vs N: Which end of the cyanide anion is a better hydrogen bond acceptor? J. Phys. Chem. A 2014, 118, 7418–7423. [Google Scholar] [CrossRef]
  91. Ramabhadran, R.O.; Hua, Y.; Li, Y.J.; Flood, A.H.; Raghavachari, K. From Atomic to Molecular Anions: A Neutral Receptor Captures Cyanide Using Strong C−H Hydrogen Bonds. Chem. Eur. J. 2011, 17, 9123–9129. [Google Scholar] [CrossRef]
  92. Hinchliffe, A. Hydrogen bonding in HCN…HF, HCN…HCl, CH3CN…HF and CH3CN…HCl; an ab initio SCF-MO study. Adv. Mol. Relax. Int. Pr. 1981, 19, 227–237. [Google Scholar] [CrossRef]
  93. 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. [Google Scholar] [CrossRef] [PubMed]
  94. Tupikina, E.Y.; Bodensteiner, M.; Tolstoy, P.M.; Denisov, G.S.; Shenderovich, I.G. P=O Moiety as an Ambidextrous Hydrogen Bond Acceptor. J. Phys. Chem. C 2018, 122, 1711–1720. [Google Scholar] [CrossRef]
  95. Arp, F.F.; Bhuvanesh, N.; Blümel, J. Hydrogen peroxide adducts of triarylphosphine oxides. Dalton Trans. 2019, 48, 14312–14325. [Google Scholar] [CrossRef] [PubMed]
  96. Ahn, S.H.; Lindhardt, D.; Bhuvanesh, N.; Blümel, J. Di(hydroperoxy)cycloalkanes Stabilized via Hydrogen Bonding by Phosphine Oxides: Safe and Efficient Baeyer−Villiger Oxidants. ACS Sustainable Chem. Eng. 2018, 6, 6829–6840. [Google Scholar] [CrossRef]
  97. Brotherton, T.K.; Lynn, J.W. The synthesis and chemistry of cyanogen. Chem. Rev. 1959, 59, 841–883. [Google Scholar] [CrossRef]
  98. Provasi, P.F.; Aucar, G.A.; Sanchez, M.; Alkorta, I.; Elguero, J.; Sauer, S.P. Interaction Energies and NMR Indirect Nuclear Spin− Spin Coupling Constants in Linear HCN and HNC Complexes. J. Phys. Chem. A 2005, 109, 6555–6564. [Google Scholar] [CrossRef]
  99. Del Bene, J.E.; Alkorta, I.; Elguero, J. Ab initio study of ternary complexes X:(HCNH)+: Z with X, Z= NCH, CNH, FH, ClH, and FCl: Diminutive cooperative effects on structures, binding energies, and spin–spin coupling constants across hydrogen bonds. J. Phys. Chem. A 2011, 115, 12677–12687. [Google Scholar] [CrossRef]
  100. Sanchez, M.; Provasi, P.F.; Aucar, G.A.; Alkorta, I.; Elguero, J. Theoretical study of HCN and HNC neutral and charged clusters. J. Phys. Chem. B 2005, 109, 18189–18194. [Google Scholar] [CrossRef] [PubMed]
  101. Chan, B.; Del Bene, J.E.; Radom, L. What factors determine whether a proton-bound homodimer has a symmetric or an asymmetric hydrogen bond? Mol. Phys. 2009, 107, 1095–1105. [Google Scholar] [CrossRef]
  102. Chan, B.; Del Bene, J.E.; Radom, L. Proton-Bound Homodimers: How Are the Binding Energies Related to Proton Affinities? J. Am. Chem. Soc. 2007, 129, 12197–12199. [Google Scholar] [CrossRef]
  103. Shenderovich, I.G. Actual Symmetry of Symmetric Molecular Adducts in the Gas Phase, Solution and in the Solid State. Symmetry 2021, 13, 756. [Google Scholar] [CrossRef]
  104. Kong, S.; Borissova, A.O.; Lesnichin, S.B.; Hartl, M.; Daemen, L.L.; Eckert, J.; Antipin, M.Y.; 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. [Google Scholar] [CrossRef] [PubMed]
  105. 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. [Google Scholar] [CrossRef] [PubMed]
  106. Alkorta, I.; Blanco, F.; Elguero, J. A theoretical structural analysis of the factors that affect 1JNH, 1hJNH and 2hJNN in N–H··· N hydrogen-bonded complexes. Mag. Res. Chem. 2009, 47, 249–256. [Google Scholar] [CrossRef]
  107. Krivdin, L.B. Calculation of 15N NMR chemical shifts: Recent advances and perspectives. Prog. Nucl. Magn. Reson. Spectrosc. 2017, 102, 98–119. [Google Scholar] [CrossRef]
  108. Benedict, H.; Limbach, H.-H.; Wehlan, M.; Fehlhammer, W.P.; Golubev, N.S.; Janoschek, R. Solid State 15N NMR and Theoretical Studies of Primary and Secondary Geometric H/D Isotope Effects on Low-Barrier NHN−Hydrogen Bonds. J. Am. Chem. Soc. 1998, 120, 2939–2950. [Google Scholar] [CrossRef]
  109. Bulychev, V.P.; Buturlimova, M.V.; Tokhadze, K.G. Calculation of interaction of the stretching and bending vibrations of HF in the hydrogen-bonded complex [F(HF)2]. Opt. Spectrosc. 2012, 113, 588–592. [Google Scholar] [CrossRef]
  110. Grabowski, S.J. Cooperativity of hydrogen and halogen bond interactions. In Highlights in Theoretical Chemistry, Proceedings of the 8th Congress on Electronic Structure: Principles and Applications (ESPA 2012), Barcelona, Spain, 26–29 June 2012; Novoa, J., Ruiz López, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; Volume 5. [Google Scholar] [CrossRef]
Figure 1. The structure of NCH and CNH complexes with proton acceptors and their relative energies ΔE = E(CNH-form) − E(NCH-form) obtained at the ωB97XD/def2tzvp and PCM = water approximations. (a) No acceptor, (b) fluoride anion (F), (c) ammonia (NH3), (d) pyridine, (e) pyridine-4-amine (4AP), (f) 1-azabicyclo[2.2.2]octane (ABCO), (g) phosphoric acid (P=O).
Figure 1. The structure of NCH and CNH complexes with proton acceptors and their relative energies ΔE = E(CNH-form) − E(NCH-form) obtained at the ωB97XD/def2tzvp and PCM = water approximations. (a) No acceptor, (b) fluoride anion (F), (c) ammonia (NH3), (d) pyridine, (e) pyridine-4-amine (4AP), (f) 1-azabicyclo[2.2.2]octane (ABCO), (g) phosphoric acid (P=O).
Symmetry 13 01298 g001
Figure 2. The 1J(15N13C) spin-spin coupling as a function of the N≡C distance in different hydrogen bonded complexes of the cyanide anion. NC-H···X (red circles), [NC]···H-X (green hexagons), CN-H···X (blue squares), [CN]···H-X (violet diamonds), LiNC···H-X (green down triangle), and LiCN···H-X (violet up triangle).
Figure 2. The 1J(15N13C) spin-spin coupling as a function of the N≡C distance in different hydrogen bonded complexes of the cyanide anion. NC-H···X (red circles), [NC]···H-X (green hexagons), CN-H···X (blue squares), [CN]···H-X (violet diamonds), LiNC···H-X (green down triangle), and LiCN···H-X (violet up triangle).
Symmetry 13 01298 g002
Table 1. The N≡C distances, the isotropic chemical shielding, σiso, the span, Ω, of the shielding tensors of the 15N and 13C nuclei, and scalar coupling constants of NC, NCH, and CNH obtained at the PCM = water approximation.
Table 1. The N≡C distances, the isotropic chemical shielding, σiso, the span, Ω, of the shielding tensors of the 15N and 13C nuclei, and scalar coupling constants of NC, NCH, and CNH obtained at the PCM = water approximation.
CompoundN≡C, Å15N, ppm13C, ppm1J(15N13C), HzJ(15N1H), HzJ(13C1H), Hz
σisoΩσisoΩ
15N≡13C1.1664−4658510403−7--
15N≡13C1H1.1436−2454865302−25−11267
13C≡15N1H1.15857840113392−14−12123
Table 2. Geometry of NCH···X and CNH···X complexes obtained at PCM = water.
Table 2. Geometry of NCH···X and CNH···X complexes obtained at PCM = water.
Base (X)NCH···XCNH···X
N≡C, ÅC···H, ÅH···X, 1 ÅC≡N, ÅN···H, ÅH···X, 1 Å
F1.15991.57541.00931.16361.52020.9875
H3N1.14611.10721.90931.16461.63101.0790
pyridine1.14581.10341.90511.15941.08861.5705
4AP1.14631.11091.85431.16501.70321.0515
ABCO1.14711.12821.77531.16501.68681.0578
P=O1.14481.0871.901.15831.0291.70
1 The distance between the NCH or CNH proton and the proton accepting atom of the base.
Table 3. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of NCH···X and CNH···X complexes obtained at PCM = water.
Table 3. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of NCH···X and CNH···X complexes obtained at PCM = water.
Base (X)NCH···XCNH···X
15N, ppm13C, ppm15N, ppm13C, ppm
σisoΩσisoΩσisoΩσisoΩ
F−3356625383−1754614397
H3N−1753857335−2756013398
pyridine−20560573343946117388
4AP−1954156336−2856111401
ABCO−2154655341−2455911403
P=O−20544603315943417389
Table 4. Scalar coupling constants in NCH ···X obtained at PCM = water.
Table 4. Scalar coupling constants in NCH ···X obtained at PCM = water.
Base (X)1J(15N13C), Hz2J(15N1H), Hz1J(13C1H), Hz1hJ(1HX), Hz2hJ(13CX), Hz3hJ(15NX), Hz
F−10−231291 1314 1−22 1
H3N−22−122584 220 21 2
pyridine−22−122594 2−21 21 2
4AP−22−122554 2−24 21 2
ABCO−20−122432 2−23 21 2
P=O−24−12266−2 37 30 3
1 X = 19F. 2 X = 15N. 3 X = 31P.
Table 5. Scalar coupling constants in CNH ···X obtained at PCM = water.
Table 5. Scalar coupling constants in CNH ···X obtained at PCM = water.
Base (X)1J(15N13C), Hz1J(15N1H), Hz2J(13C1H), Hz1hJ(1HX), Hz3hJ(13CX), Hz2hJ(15NX), Hz
F−8−53371 127 1−93 1
H3N−711−66 2−2 212 2
pyridine−13−9422−5 2−3 216 2
4AP−831−97 2−2 211 2
ABCO−841−73 2−2 29 2
P=O−15−11425−2 30 3−3 3
1 X = 19F. 2 X = 15N. 3 X = 31P.
Table 6. Geometry of NCH···FLi and CNH···FLi hydrogen-bonded complexes obtained at the gas phase and PCM = water approximations.
Table 6. Geometry of NCH···FLi and CNH···FLi hydrogen-bonded complexes obtained at the gas phase and PCM = water approximations.
PCMNCH···FLiCNH···FLi
N≡C, ÅC···H, ÅH···F, ÅC≡N, ÅN···H, ÅH···F, Å
gas1.14601.09801.69351.16021.04381.5239
water1.14841.14611.46661.16051.20791.1738
Table 7. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of NCH···FLi and CNH···FLi complexes obtained at the gas phase and PCM = water approximations.
Table 7. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of NCH···FLi and CNH···FLi complexes obtained at the gas phase and PCM = water approximations.
PCMNCH···FLiCNH···FLi
15N, ppm13C, ppm15N, ppm13C, ppm
σisoΩσisoΩσisoΩσisoΩ
gas−32561613305743612395
water−16538523441649718361
Table 8. Scalar coupling constants in NCH ···FLi and CNH ···FLi obtained at the gas phase and PCM = water approximations.
Table 8. Scalar coupling constants in NCH ···FLi and CNH ···FLi obtained at the gas phase and PCM = water approximations.
Complex1J(15N13C), HzJ(15N1H), HzJ(13C1H), Hz1hJ(1H19F), HzhJ(13C19F), HzhJ(15N19F), Hz
gas
NCH···FLi−22−12256−66146−7
CNH···FLi−13−11024−7215−79
water
NCH···FLi−20−12232−109252−17
CNH···FLi−11−57164742−153
Table 9. The structure of NC···HCN, NC···HNC, and CN···HNC complexes, their relative energies ΔE, the isotropic chemical shielding, σiso, the span, Ω, of the shielding tensors, and scalar coupling constants obtained at PCM = water.
Table 9. The structure of NC···HCN, NC···HNC, and CN···HNC complexes, their relative energies ΔE, the isotropic chemical shielding, σiso, the span, Ω, of the shielding tensors, and scalar coupling constants obtained at PCM = water.
Parameter[NC···H···CN][CN]···H-CN[CN]···H-NC
ΔE, kJ/mol05164
Base: N≡C, Å1.15151.16551.1611
Base: X···H, Å1.2220 11.1756 21.4362 2
Acid: H-Y, Å1.2220 31.2917 31.0847 4
Acid: N≡C, Å1.15151.15281.1532
Base: σiso (15N), ppm−6−17−22
Base: Ω(15N), ppm525548555
Acid: σiso (15N), ppm−6−1928
Acid: Ω(15N), ppm525544481
Base: σiso (13C), ppm322116
Base: Ω(13C), ppm377385394
Acid: σiso (13C), ppm323223
Acid: Ω(13C), ppm377376382
Base: 1J(15N13C), Hz−13−8−9
Acid: 1J(15N13C), Hz−13−13−14
Base: J(13C1H), Hz95145
Acid: J(13C1H), Hz959524
Base: J(15N1H), Hz−8−44−9
Acid: J(15N1H), Hz−8−8−87
2hJ(XHY), Hz113 5−57 620 7
1 X = C. 2 X = N. 3 Y = C. 4 Y = N. 5 2hJ(13C13C). 6 2hJ(13C15N). 7 2hJ(15N15N).
Table 10. Geometry of (NCH)n···F and (CNH)n··· F aggregates obtained at PCM = water.
Table 10. Geometry of (NCH)n···F and (CNH)n··· F aggregates obtained at PCM = water.
nΔE 1
kJ/mol
(NCH)n··· F(CNH)n···F
N≡C, ÅC···H, ÅH···F, ÅC≡N, ÅN···H, ÅH···F, Å
141.15991.57541.00931.16361.52020.9875
2481.14851.14881.46741.15901.10861.3338
3931.14691.11971.58191.15841.06121.4651
41431.14611.10621.66211.15831.04121.5588
1 Relative energies ΔE = E(CNH-form) − E(NCH-form)
Table 11. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of (NCH)n···F and (CNH)n··· F aggregates obtained at PCM = water.
Table 11. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of (NCH)n···F and (CNH)n··· F aggregates obtained at PCM = water.
n(NCH)n···F(CNH)n···F
15N, ppm13C, ppm15N, ppm13C, ppm
σisoΩσisoΩσisoΩσisoΩ
1−3356625383−1754614397
2−15537513443447020386
3−16539553394745129387
4−18540573365544018389
Table 12. Scalar coupling constants in (NCH)n···F and (CNH)n··· F aggregates obtained at PCM = water.
Table 12. Scalar coupling constants in (NCH)n···F and (CNH)n··· F aggregates obtained at PCM = water.
Complex1J(15N13C), HzJ(15N1H), HzJ(13C1H), Hz1hJ(1H19F), HzhJ(13C19F), HzhJ(15N19F), Hz
(NCH)···19F−10−231291314−22
(NCH)2···19F−20−12230−108228−16
(NCH)3···19F−22−12250−87171−11
(NCH)4···19F−22−12257−68132−8
(CNH)···19F−8−5337127−93
(CNH)2···19F−13−8722−6729−109
(CNH)3···19F−14−10325−7319−80
(CNH)4···19F−15−11025−6113−58
Table 13. Geometry of (NC)43−···(H4N)+, (LiNC)43−···(H4N)+, (CN)43−···(H4N)+, and (LiCN)43−···(H4N)+ hydrogen-bonded aggregates obtained at PCM = water.
Table 13. Geometry of (NC)43−···(H4N)+, (LiNC)43−···(H4N)+, (CN)43−···(H4N)+, and (LiCN)43−···(H4N)+ hydrogen-bonded aggregates obtained at PCM = water.
BΔE 1
kJ/mol
([BNC])4···(NH4)+([BCN])4···(NH4)+
N≡C, ÅC···H, ÅN-H, ÅC≡N, ÅN···H, ÅN-H, Å
No−61.16442.02331.04001.16551.87511.0374
Li+101.16242.02951.03821.16151.88341.0359
1 Relative energies ΔE = E(CNH-form) − E(NCH-form).
Table 14. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of (15N≡13C)43−···(H4N)+, (Li15N≡13C)4···(H4N)+, (13C≡15N)43−···(H4N)+, and (Li13C≡15N)4···(H4N)+ aggregates obtained at PCM = water.
Table 14. The isotropic chemical shielding, σiso, and the span, Ω, of the shielding tensors of the 15N and 13C nuclei of (15N≡13C)43−···(H4N)+, (Li15N≡13C)4···(H4N)+, (13C≡15N)43−···(H4N)+, and (Li13C≡15N)4···(H4N)+ aggregates obtained at PCM = water.
B([B15N≡13C])4···(NH4)+([B13C≡15N])4···(NH4)+
15N, ppm13C, ppm15N, ppm13C, ppm
σisoΩσisoΩσisoΩσisoΩ
No−3757215398−3957913399
Li+−10969127384452035378
Table 15. Scalar coupling constants in (15N≡13C)43−···(1H415N)+, (Li15N≡13C)43−···(1H415N)+, (13C≡15N)43−···(1H415N)+, and (Li13C≡15N)43−···(1H415N)+ aggregates obtained at PCM = water.
Table 15. Scalar coupling constants in (15N≡13C)43−···(1H415N)+, (Li15N≡13C)43−···(1H415N)+, (13C≡15N)43−···(1H415N)+, and (Li13C≡15N)43−···(1H415N)+ aggregates obtained at PCM = water.
Complex1J(15N13C), HzhJ(15N1H), HzhJ(13C1H), HzhJ(13C15N), HzhJ(15N15N), Hz
([15N≡13C])4···(H415N)+−90−7−151
(Li15N≡13C)4···(H415N)+−5−1−6−150
([13C≡15N])4···(H415N)+−750−16
(Li13C≡15N)4···(H415N)+−1740−15
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shenderovich, I.G.; Denisov, G.S. NMR Properties of the Cyanide Anion, a Quasisymmetric Two-Faced Hydrogen Bonding Acceptor. Symmetry 2021, 13, 1298. https://doi.org/10.3390/sym13071298

AMA Style

Shenderovich IG, Denisov GS. NMR Properties of the Cyanide Anion, a Quasisymmetric Two-Faced Hydrogen Bonding Acceptor. Symmetry. 2021; 13(7):1298. https://doi.org/10.3390/sym13071298

Chicago/Turabian Style

Shenderovich, Ilya G., and Gleb S. Denisov. 2021. "NMR Properties of the Cyanide Anion, a Quasisymmetric Two-Faced Hydrogen Bonding Acceptor" Symmetry 13, no. 7: 1298. https://doi.org/10.3390/sym13071298

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop