than. ‘42.. .t m“ I. .. 1;. i: t h» .MEle . .3: E i . $3? ‘ 1.1.0.! (-0.): iiiul! 0,1. I). y .. I"! 3| I . .. KN? . .513. 5.... .2. . .51.». 5“...) ‘ . arbbx: ; 1.: {3 3.53.51» fl .\.r)‘l-\5 ‘ . i...1:.vl.c.x‘.£l. ; , .KKAI Iii :0!» ;§.‘IA, 3| L. .$:\t\n. )ill): .t‘h-S911 t ‘ 17.3! . vaL I. ill .. :Et.‘ . 4. 3.31.1? 1.1.55.1.)€\u!\~ .1. .1 . . .,V V ‘ 53.3.31: .. . I ‘ .. .Hu H;W§.Muw.vm m lllllllllllllllHIHHHIHIllllHUI!“llllllllllllllll 3 3 1293 014203 This is to certify that the dissertation entitled Solute-Solvent Interactions in Nonaqueous Solutions presented by Philip W. Schultz has been accepted towards fulfillment of the requirements for Ph . D . d . Chemistry egree m Date 11/16/95 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY , Michigan State I Unlverslty PLACE N RETURN BOX Io romovo this checkout from your rocord. TO AVOID FINES rotum on or baton dot. duo. DATE DUE DATE DUE DATE DUE MSU In An mm Action/Equal Opportunity Imitation mm: SOLUTE-SOLVENT INTERACTIONS IN NONAQUEOUS SOLVENTS By Philip Wayne Schultz A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1995 ABSTRACT SOLUTE-SOLVENT INTERACTIONS IN NONAQUEOUS SOLVEN TS By Philip Wayne Schultz The interactions of the cyanate, thiocyanate and selenocyanide anions in the nonaqueous solvents: methanol, formamide and N-methylformamide were investigated by infrared and multinuclear N MR spectroscopies. In this series of protic solvents, the cyanate anion is more highly solvated than the thiocyanate and selenocyanide anions. In order to understand the complicated nature of the spectral envelopes for the stretching modes, the nature of the experimental vibrational modes were studied by ab initio calculations at the MP2 level, in response to perturbations of these anions by a variety of ionic and hydrogen bonding interactions. These studies provided a firm basis for assigning the perturbations, in frequency and absorptivity, by the formation of hydrogen bonds with this set of anions. These assignments have been augmented with the synthesis and crystal structure determination of model solvates. The three anions were studied also in nitromethane, a relatively ‘inert’ solvent, so that the salt and protic solvent concentration could be varied independently to help identify the component bands of the stretching modes and to provide insight into the solvation phenomena. The concentration dependence of the infrared vibrations and the NMR resonances yielded the stoichiometry of the anion-solvent complex and the thermodynamics of the complex formation. In the amide solvents, the cyanate anion is hydrogen bonded at the nitrogen with one and two solvent molecules. These two interactions are also present in methanol solutions with an additional complex where a methanol molecule is bound to each end of the anion. The thiocyanate and selenocyanide anions are substantially less solvated than the cyanate anion with the interactions occurring mainly at the nitrogen end. These assignments are consistent with the prediction that the negative electrostatic potential is concentrated off the nitrogen atom. These investigations suggest that the extent of self association of the protic solvent is an important factor in the anion-solvent equilibria, resulting in the minimization of the entropy loss upon hydrogen bond formation. Acknowledgments The author would like to express his sincere appreciation to Dr. Alexander I. Popov and Dr. George E. Leroi for their direction and encouragement during the course of this study. Additional thanks need to be expressed to Dr. James F. Harrison who provided the author substantial assistance. He also wishes to thank the research groups of Drs. Leroi and Harrison for their encouragement and the many productive discussions. Great thanks also go to the NMR support staff for their help with the NMR spectrometers and the extensive use of the spectrometers. The author is deeply indebted to the computer facility for providing a tremendous amount of computer resources for this work. Dr. Karen C. Inman, the author’s wife, deserves great thanks for her encouragement and her source of ideas to spur this work to completion and to give this work a quality that would not have been possible without her. Her optimism was a tremendous help. Finally, the author wishes to acknowledge the Creator of this physical world who has given us curious minds and provided us with a fascinating world to discover. Table of Contents List of Tables List of Figures List of Abbreviations I. Introduction A. Introduction B. Solvent-Solvent Interactions C. Ion-Solvent Interactions D. Ion-Ion Interactions E. Model Crystal Structures F. Proposed Work H. Experimental Methods A. Materials B. Spectroscopic Techniques C. Theoretical Calculations D. Data Treatment 111. Theoretical Calculations A. Introduction B. Results OCN', SCN' and SeCN‘ Electrostatic effects on OCN‘, SCN' and SeCN' Ionic and hydrogen bonding Atomic charges in ionic and hydrogen bonded complexes Molecular orbitals in ionic and hydrogen bonded complexes Electrostatic potential in ionic and hydrogen bonded complexes sowéww— Vibrational Frequencies in ionic and hydrogen bonded complexes viii xiv XX 14 31 36 37 38 39 45 46 48 49 55 55 58 70 74 87 92 I 8. Cation Solvation 98 C. Conclusions 98 IV. Solvation of OCN' 103 A. Introduction 104 B. Results 105 1. Infrared Measurements 105 2. Assignments 117 3. NMR Measurements 120 4. Thermodynamic Parameters 121 C. Conclusions 129 V. Solvation of SCN‘ and SeCN' 132 A. Introduction 133 B. Results 135 1. Infrared Measurements 135 2. Assignments 145 3. Thermodynamics 148 C. Conclusions 160 VI. Crystal Structures 162 A. Introduction 163 B. Results 164 1. Crystal Structures for NaNCS°(DMF)2 and NaNCS0(NMF)2 164 a. Amide Structure 172 b. Thiocyanate 172 c. Hydrogen Bonding 172 2. Summary of NaSCN°(DMF)2 and NaSCN-(NMF)2 173 3. Crystal structures for Na(l2C4)2NCS°MeOH and 173 Na(12C4)2NCS a. Infrared Spectroscopy 183 b. Ab initio Calculations 183 4. Summary of Na(12C4)2NCS'MeOH and Na(12C4)2NCS 183 vi VII. Conclusions and Future Work A. Conclusions B. Future Work VIII. References vii 185 186 187 189 Table 1. Table 2. Table 3. Table 4. Table 5. Table 6. Table 7. Table 8. Table 9. List of Tables AH°vap for selected solvents Entropy of vaporization (AS° ) for selected solvents vap Physical properties of selected solvents at 25°C Donor and acceptor numbers of selected solvents Wavenumber and band width of the vCN mode of the thiocyanate anion in various solvents Assignments of the six component vCN bands in tetrabutylammonium thiocyanate in methanol by Gill et al. Assignment of the four component vCN bands of tetrabutylammonium thiocyanate in methanol by Chabanel and coworkers Spectroscopic parameters for the vCN mode of the cyanate anion alkali cyanates in DMSO Thermodynamics of alkali cyanate ion pairs in DMSO at 298K Table 10. Complexes of metal thiocyanates and their vibrational spectra in various aprotic solvents viii 10 13 14 21 22 24 Table 11. Table 12. Table 13. Table 14. Table 15. Table 16. Table 17. Table 18. Table 19. Table 20. MNDO calculations of the vCN stretch in LiNCS complexes compared with the experimental frequency Thermodynamic parameters for the ion pair, dimer and tetramer formation of alkali thiocyanates in various solvents Frequency (cm") of the vCN mode of LiSeCN in various solvents Parameters for NMR spectra Frequencies of the stretching modes of OCN', SCN' and SeCN' species in solution Calculated equilibrium bond lengths, stretching frequencies and atomic charges for OCN', SCN' and SeCN' Potential energy distribution and force constants for the stretching modes of OCN‘, SCN' and SeCN' Effect of a positive charge on the bond lengths, force constants and vibrational frequencies for OCN', SCN' and SeCN’, calculated upon fixing a bare proton 3 A along the internuclear axis for either end Calculated bond lengths, stretching frequencies, integrated intensities, force constants and atomic charges for OCN' in ionic and hydrogen bonded cyanate complexes Calculated bond lengths, stretching frequencies, integrated intensities, force constants and atomic charges for SCN‘ in ionic and hydrogen bonded cyanate complexes ix 29 30 31 40 50 55 56 62 65 67 Table 21. Table 22. Table 23. Table 24. Table 25. Table 26. Table 27. Table 28. Table 29. Calculated bond lengths, stretching frequencies, integrated intensities, force constants and atomic charges for SeCN' in ionic and hydrogen bonded cyanate complexes Occupancies of 2px, 2py and 2pz atomic orbitals in alkali metal, alkaline earth metal and hydrogen bonded cyanates Calculated changes in bond lengths, force constants and frequency shifts for selected ionic cyanates and selenocyanides with respect to the free anions Calculated bond lengths and vibrational frequencies at the HF/ d95v+** level for OCN’, LiNCO, and Li(HOH)3NCO Infrared parameters for the vCN mode of ~O.1M NBu4OCN in several solvents Peak position, band width, and molar absorptivity parameters of component bands of the VCN spectral envelope of ~0.1M OCN‘ in DMF, NM, NMF, FA and MeOH Summary of peak positions and band widths of the curve-fitted component bands of the vCN stretch of ~0.1M NBu4OCN/NM solutions with NMF, FA or MeOH added Assignments for the vCN and VCO & 50m modes for ~0.1M OCN' in neat solutions with MeOH, FA or NMF and titration studies with these solvents 14N and 170 NMR chemical shifts of ~0.1 NBu4OCN in DMF, NM, NMF, FA and MCOH 68 73 96 99 106 106 116 119 121 Table 30. Table 31. Table 32. Table 33. Table 34. Table 35. Table 36. Table 37. Table 38. Association constants, K1, K2, and KD for the formation of OCN""HB, OCN""(HB)2 and BH'"BH, where HB can be NMF, FA or MeOH determined by 14N and 170 NMR spectroscopy Association constants, K1, K2, and KD for the formation of OCN""HB, OCN""(HB)2 and BH"'BH, where HB can be NMF, FA or MeOH determined by infrared spectroscopy Thermodynamic parameters for the formation of 1:1 and 1:2 OCN""solvent complexes in nitromethane solutions Summary of position and band widths for the VCN stretch of ~0.25M NBu4SCN in NM, DMF, NMF, FA and MeOH Summary of the deconvolved parameters for the component bands in ~0.25M NBu4SCN in NM, DMF, NMF, FA, and MeOH Summary of the peak position, band widths (cm“) and molar absorptivities from the curve-fitted component bands of the vCN stretch of 0.15M NBu4SCN with NMF, FA or MeOH added Summary of the peak positions and band widths from the curve- fitted component bands of the vCN stretch of 0.025M NBu4SeCN with MeOH added Association constants, K,+K2 and KD, for the formation of the two 1:1 complexes and self association of the protic solvent, determined by 1“N NMR measurements Association constants (K,+K2 and KB) of SCN' with MeOH, FA and NMF from the stretching modes of SCN' xi 127 128 129 137 137 144 145 155 159 Table 39. Table 40. Table 41. Table 42. Table 43. Table 44. Table 45. Table 46. Table 47. Individual association constants using the extinction coefficients from ab initio calculations Thermodynamic parameters for the association of SCN' with MeOH, FA and NMF from the NMR measurements Positional parameters (A) and equivalent isotropic thermal parameters (A2) with estimated standard deviations in parenthesis for bis(N-methylformamide)—sodium thiocyanate Positional parameters and isotropic temperature with estimated standard deviations in parenthesis factors for bis(N,N-dimethylformamide)-sodium thiocyanate Intermolecular distances (A) with estimated standard deviations in parenthesis for bis(N-methylformamide)-sodium thiocyanate Intermolecular distances with estimated standard deviations in parenthesis for bis(N,N—dimethylformarnide)- sodium thiocyanate Atomic coordinates (104) and equivalent isotropic displacement parameters (A2x103) with estimated standard deviations in parenthesis for the Na(12C4)2NCS complex Atomic coordinates (104) and equivalent isotropic displacement parameters (A2x103) with estimated standard deviations in parenthesis for the N a( 12C4)2NCS°MeOH complex Selected bond lengths with estimated standard deviations in parenthesis for the Na(12C4)2NCS complex 159 160 169 170 171 171 178 179 180 Table 48. Selected bond lengths with estimated standard deviations in 181 parenthesis for the N a( 12C4)2NCS°MeOH Table 49. Ab initio calculations for axial and nonaxial hydrogen bonds 184 with SCN‘ Figure Figure Figure Figure Figure Figure Figure Figure Figure List of Figures . RDF for pure FA, NMF, and DMF . The VCN stretch of NBu4SCN in methanol and the component bands determined by Gill and coworkers for spectral envelopes . Concentration dependence of the 39K chemical shift in various solvents . Concentration dependence of the 23N a chemical shift of NaSCN, NaBr, NaBPh4, NaClO4 in DMF . Experimental infrared band envelope for the VCN stretch of NaSCN in THF with calculated component bands . Pictorial view of the LiNCS ion pair, dimer, and tetramer . Frequency (cm") of the vCN stretch in alkali and alkaline earth thiocyanate ion pairs as a function of the polarizing power of the metal cation . Resonance structures for the thiocyanate anion . VCN/P of various metal thiocyanates correlated to the polarization of the metal cations xiv 12 17 19 20 25 26 26 27 Figure 10. Figure 11. Figure 12. Figure 13. Figure 14. Figure 15. Figure 16. Figure 17. Figure 18. Figure 19. Figure 20. Figure 21. Figure 22. Li(9-Crown-3)SCN Symmetric (LiNCS)2 dimer (left) and asymmetric [Li(TMPA)2NCS]2 Diagrammatic representation of the coordination in (KNCS)2-DB24C8 Contour plots for the valence molecular orbitals of OCN‘ Contour plots for the valence molecular orbitals of SCN' Contour plots for the valence molecular orbitals of SeCN‘ Atomic charges at O, C and N in MNCO species, where M can be a hydrogen bond donor, an alkali metal cation, or an alkaline earth metal cation, versus the Coulombic potential provided by the cation or hydrogen bond donor; r is the correlation coefficient with and without (in parenthesis) the hydrogen bonded species Contour plots for the valence molecular orbitals in LiNCO Contour plots for the valence molecular orbitals in LiOCN Contour plots for the valence molecular orbitals in BeNCO+ Contour plots for the valence molecular orbitals in Li2NCO+ Contour plots for the valence molecular orbitals in LizOCN“ Contour plots for the valence molecular orbitals in LiNCS XV 32 34 35 59 60 61 72 75 76 77 78 79 80 Figure 23. Figure 24. Figure 25. Figure 26. Figure 27. Figure 28. Figure 29. Figure 30. Figure 31. Figure 32. Contour plots for the valence molecular orbitals in BeNCS+ Contour plots for the valence molecular orbitals in LizNCS+ Contour plots for the valence molecular orbitals in LiNCSe Contour plots for the valence molecular orbitals in BeNCSe” Contour plots for the valence molecular orbitals in LiZNCSe* Contour plots for the 11: and 21: molecular orbitals in OCN' Li”NCO‘, Be2*NCO' and Li+OCN', Contour plots of a 1: molecular orbital in HF extending into the ON bond of OCN', thereby increasing the electron density surrounding the C-N bond in the OCN""HF complex Electrostatic potential for OCN', SCN‘ and SeCN‘ Molecular electrostatic potential of OCN', OCN""HOH, OCN""HF and Li‘NCO’ Ab initio frequencies of the vCN stretch in alkali cyanates versus experimental frequencies of the vCN stretch of alkali cyanates in DMSO Figure 33. Ab initio frequencies of the VCN stretch in alkali thiocyanates Figure 34. versus experimental frequencies of the VCN stretch of alkali thiocyanates in aprotic solvents Plot of the FCN and FCO force constants versus their respective bond lengths in ionic and hydrogen bonded OCN‘ complexes xvi 81 82 83 84 85 86 88 9O 91 93 93 97 Figure 35. Figure 36. Figure 37. Figure 38. Figure 39. Figure 40. Figure 41. Figure 42. Figure 43. Figure 44. Figure 45. Pictorial view of the Li(HOH)3NCO ion pair vCN stretch of ~0.1M NBu4OCN in DMF, NM, NMF, FA, and MeOH vCN stretch of ~0.1M NBu4OCN in NM and the Fourier deconvolution of the vCN stretch as a function of MeOH added vCN stretch of ~0. 1M NBu4OCN in NM and the Fourier deconvolution of the vCN stretch as a function of FA added VCN stretch of ~0.1M NBu4OCN in NM and the Fourier deconvolution of the vCN stretch as a function of NMF added Absorptions for the component bands in the 2200-2100 cm'1 region for ~0.1M NBu4OCN/N M solutions as a function of MeOH, FA or NMF concentration \ICN & 250CN stretch for ~ 0.1M NBu4OCN in NM as a function of MeOH added VCN & 250CN stretch for ~ 0.1M NBu4OCN in NM as a function of FA or N MF added '4N and '70 NMR chemical shifts for 0.1M NBu4OCN as a function of MeOH added and temperature in the case for the l"N resonance VCN for ~0.25M NBu4SCN in DMF, NM, NMF, FA and MeOH Fourier deconvolution of vCN for 0.25M NBu4SCN in MeOH xvii 100 107 109 110 111 112 114 115 122 138 139 Figure 46. VCN stretch for ~0.25M NBu4SCN in DMF Figure 47. VCN stretch for ~0. 1M NBu4SCN in NM with MeOH added Figure 48. Fourier deconvolution of the vCN stretch for ~0.1M NBu4SCN in NM as a function of MeOH added Figure 49. Absorptivities for component bands of the vCN mode of ~0.1M NBu4SCN in NM as a function of MeOH added Figure 50. vCS stretch for ~0.1M NBu4SCN in NM as a function of MeOH added Figure 51. 1“N NMR chemical shift for ~0.1M NBu4SCN in NM as a function of MeOH concentration and temperature Figure 52. 1“N NMR line widths for ~0. 1M NBu4SCN in NM as a function of MeOH concentration and temperature Figure 53. Ortep structure for N aNCS°(DMF)2 Figure 54. Packing diagram for NaNCs-(DMF)2 Figure 55. Ortep plot for NaNCs-(NMF)2 Figure 56. Packing diagram for NaNCS°(NMF)2 Figure 57. Ortep plot for Na(12C4)2NCS Figure 58. Packing diagram for Na(l2C4)2NCS Figure 59. Ortep plot for Na(12C4)2NCS°MeOH xviii 140 142 143 146 151 152 154 165 166 167 168 174 175 176 Figure 60. Packing diagram for Na(12C4)2NCS°MeOH 177 xix 1. 2. Salts Solvents NBu4OCN NBu4SCN NBu4SeCN NBu4NO3 NBu4Cl NaSCN KSeCN KCl BaO P205 CaH2 NaClO4 NM DMF NMF FA MeOH List of Abbreviations Tetrabutylammonium cyanate Tetrabutylammonium thiocyanate Tetrabutylammonium selenocyanide Tetrabutylammonium nitrate T etrabutylammonium chloride Sodium thiocyanate Potassium selenocyanide Potassium chloride Barium oxide Phosphorous pentoxide Calcium hydride Sodium perchlorate Nitromethane N,N-dimethylformamide N-methylformamide formamide Methanol XX 3. Ligand HMPA H20 EtOH PY THF AC DMSO 12C4 Hexamethylphosphoric triamide Water Ethanol Pyridine Tetrahydrofuran Acetonitrile Acetone Dimethylsulfoxide 1 2-Crown-4 xxi Chapter I Introduction A. Introduction Although electrolyte solutions play an important role in many biochemical and industrial processes, our knowledge of the structure of such solutions, especially in the case of nonaqueous and mixed solvents, is rather rudimentary.l Many types of interactions may occur in an electrolyte solution. Their strength varies from weak van der Waals forces to strong electrostatic attractions. This wide range of attractive forces and the plethora of interaction sites result in a large number of solution equilibria that are difficult to sort out experimentally. In an electrolyte solution, the types of interactions that occur can be broken down into three broad categories: solvent-solvent, ion-solvent, and ion-ion. The relative importance of the resulting equilibria depends upon the nature of the electrolyte and of the solvent. B. Solvent-Solvent Interactions Interactions between neutral molecules have been of long standing interest. Van der Waals invoked such interactions between gas molecules to explain the deviations from ideal gas law behavior in 1873.2 The dispersion forces that produce the non-ideality are due to transitory electric moments in molecules that arise from the motions of the electrons.3 In nonpolar solvents, dispersion forces alone are responsible for the cohesion of the liquid state. Larger deviations from the ideal gas law are seen in vapors of molecules with a permanent dipole moment. In addition to the dispersion forces, polar molecules may interact through dipole-dipole interactions and hydrogen bonding. The strength of cohesive forces in the liquid state of polar molecules is reflected in the enthalpy of vaporization (AH°V8P), as shown in Table 1. For molecules of similar molecular weights (MW), the enthalpy of vaporization is dramatically larger for polar solvents than for nonpolar ones. Moreover, the stronger attractive forces in polar solvents provide a local structure, or order, Table l. AH°vap for selected solvents“ Molecule MW (amu) Aflow,p (kJ/mole) CH4 16 8.2 C211,, 30 14.7 H20 1 8 40.6 CH3OH 32 35.3 that the nonpolar solvents do not have. The higher order of polar solvents is revealed by the entropy of vaporization (AS°vap), as shown in Table 2. All nonpolar solvents have approximately the same molar entropy of vaporization, known as Trouton's constant, which illustrates that nonpolar solvents have a comparable amount of disorder.5 Since the standard molar entropy of all vapors is roughly constant, this illustrates that nonpolar solvents have a comparable amount of disorder in the liquid state. The larger entropy of vaporization for polar solvents, such as water and methanol, reflects their greater aggregation in the condensed phase. Table 2. Entropy of vaporization (AS°,,,p) for selected solvents4 Solvent 115",“ UK" mole") CCI4 86 C,H.,, 85 C6H6 87 H20 109 CquH 104 Hydrogen bonding between solvent molecules can result in a very strong interaction. This phenomenon is one of the reasons that water has such unusual chemical and physical properties.3 A hydrogen bond in a D-H---A system, where DH is the hydrogen bond donor and A is the hydrogen bond acceptor, has a number of unique characteristics. The H---A bond length is shorter than the sum of the van der Waals radii of the two atoms and longer than the covalent bond length of HA. In the case of an ethanol- pyridine hydrogen bond, the HA bond length is 1.80 A.3 This distance is dramatically less than the van der Waals radius sum of 2.70 A. Moreover, the D-H---A bond angle is almost linear, which may not be the most favorable dipole-dipole alignment.3 The strength of a hydrogen bond, in the gas phase, can be deduced from the change in the stretching frequency of the DH mode, since this bond weakens upon hydrogen bonding.6 In the gas phase, the strength of hydrogen bonds lie between 10 and 35 kJ/mole.3 Table 3. Physical properties of selected solvents at 25°C7 Solvent MW (amu) bp ( 0 C) Dielectric Dipole Moment Constant (D) water 1 8 100 78 l .8 methanol 32 65 33 l .7 formamide 45 21 1 l 11 3.4 N-methyl- 59 l 80 l 82 3 .9 formamide N ,N-dimethyl— 73 153 37 3.9 fomamide nitromethane 61 101 37 3.6 Dispersion forces, dipole-dipole interactions, and hydrogen bonding help determine a solvent’s unique set of physical properties. Examples of these physical properties for a selected series of solvents are gathered in Table 3. The amide solvents qualitatively show how intermolecular forces affect the physical properties of a solvent. Although formamide (FA), N-methylformamide (NMF), and N,N-dimethylformamide (DMF) all have large dipole moments, the strength of the cohesive forces is vastly different as seen from the wide range of boiling points. The effect of hydrogen bonding can be clearly seen by comparing N MF and DMF. These two solvents have the same dipole moment yet greatly different dielectric constants. A comparison of nitromethane (NM) and NMF, two solvents with similar molecular weights, shows the effect of hydrogen bonding on the boiling point. The large difference in boiling points for N MF and FA suggests that the hydrogen bonding is much stronger in FA than in NMF. An understanding of the structure of these solvents would provide a basis for interpreting the differences in their physical properties. The structure of a number of solvents has been studied by X-ray diffraction, neutron diffraction, and electron diffraction.8'l3 The role of hydrogen bonding, dipole- dipole interactions, and dispersion forces in determining the three dimensional structure is demonstrated by the amide solvents. A comparison of the radial distribution functions (RDF) for FA, NMF, and DMF, as reproduced in Figure 1, shows the vastly different structure of these solvents. The RDF shows both the local order, i.e. nearest neighbor interactions, and longer range ordering. Beyond the immediate hydrogen bond interaction at 290 pm (10"2 m), FA has longer range correlations (order) at 350-500 pm and at 700- 800 pm.9 Liquid FA consists of a three dimensional network of hydrogen bonded molecules comprised of ring and chain structures. Liquid NMF also has an extensive hydrogen bonding network, except that it is much more disordered at longer ranges. In fact, N MP is essentially disordered beyond 500 pm.ll On the other hand, DMF, which has no ability to form hydrogen bonds, displays no local order.14 The physical properties of the amide solvents can be interpreted in light of their liquid structures. The dielectric constant of NMF is much greater than that of FA (64%) yet the dipole moment in NMF is only 15% greater than the dielectric constant of FA.7 Moreover, the dielectric constant of NW is almost five times the dielectric constant of DMF, even though these solvents have similar dipole moments. The wide range of dielectric constants in the amide solvents (37- 182) has been attributed to the differences in solution structure. FA and NMF have higher dielectric constants than DMF, which reflects the greater amount of self association in the hydrogen bonding amide solvents. The difference between FA and N MP has been explained through the types of hydrogen bonding. Since FA has several ‘sites available for hydrogen bonding, it can easily form cyclic dimers. The presence of cyclic dimers in FA reduces the ability of this solvent to 'shield ionic charges despite the greater association seen in neat FA versus neat NMF. These observations demonstrate that not only is an understanding of the self association of a solvent helpful in understanding its physical properties, but also the extent of the self association is important in determining the solvent’s physical properties. [Dill-11:10.1 / pm'I L #1 l l 1 I 1 I] 200 400 500 800 IOCO r / pm Figure l. RDF for pure FA, NMF, and DMF” C. Ion-Solvent Interactions The strength of an ion-solvent interaction is determined by the properties of the ion and the solvent. Ionic solvation is manifested primarily through ionic charge-solvent dipole interactions. The donor—acceptor approach to molecular interactions has been helpful in establishing an empirical scale that ranks the strength of ion-solvent interactions. In this exposition, each solvent has a donor number and an acceptor number that are empirical measures of the strength of a cation-solvent interaction and an anion-solvent interaction, respectively.” The donicity of a solvent is defined as the negative molar enthalpy (in kcal/mole) for the reaction of a donor (D) with SbC15 as a reference acceptor, both at 10‘3 M in dichloroethane.‘5 D + SbC15 Z D-SbCls The acceptor number (AN) is related to the NMR chemical shift of 31P in Et3P0 in the solvent under study, where 8m is the 3'P chemical shift of EtjPO at infinite dilution in the solvent relative to the 3'P chemical shift of Et3PO at infinite dilution in hexane and 55t3p0.SbC]5 is the chemical shift of the Et3POOSbC15 complex in 1,2 dichloroethane.l5 Therefore, the acceptor number is based on a scale from 0 to 100 where the acceptor number of hexane is 0 and the EtaPO'SbCI5 complex is 100. The donor and acceptor 5cm, x 100 AN 5 5(Et3Po-8bC15) numbers for many solvents have been determined. Generally, solvents that can hydrogen bond have high acceptor numbers and solvents with amine and amide functions have large donor numbers.”16 Table 4. Donor and acceptor numbers of selected solventsl6 Solvent Donor Number Acceptor Number H20 l 8,30’1 55 MeOH I9 42 FA 24 40 NMF 24 32 DMF 27 16 NM 0 21 ‘The donor number of 30 for water was determined spectroscopically” Many techniques have been employed to investigate ion-solvent interactions. X-ray diffraction has been used to study the hydration of many metal cations by determining the metal-oxygen distance and the hydration number in aqueous solutions.‘4 Vibrational spectroscopy has been employed to follow ion-solvent interactions by analysis of perturbations in the normal vibrational modes of the ion or the solvent.18 The hydration of the Al3+ ion has been studied by 27A1 and ‘70 NMR.19 Generally, ion-solvent interactions observed by NMR spectroscopy are in the fast chemical exchange regime; however, the solvent exchange for the Al3+(H20)n complex is unusually slow so that the hydrated complex can be observed directly from the bulk solution.‘9 Popov and coworkers used 23Na NMR to determine the relative ion-solvent complexation strengths of several nonaqueous solvents with Na+ (from NaClO4).20 The strength of the solvation decreased in the order HMPA > H20 > DMSO > DMF ~ N MF ~ FA ~ MeOH > EtOH > PY >> THF > AC ~ AN >> NM.20 Except for pyridine, there is a good correlation between the relative solvating ability of the solvents and their donor numbers.20 The peculiarity with pyridine can be rationalized in terms of the 'hard-soft‘ interactions between the donors and acceptors. Pyridine is a softer base than the other solvents in this study.” Keil et a]. used proton NMR spectroscopy to determine that the hydration number of Li+ is 4;22 this compares well with the solvation numbers of lithium in l-methyl-Z- pyrrolidone (4.5),23 in acetone (4.3),24 and in DMF (3.8).25 Hertz and coworkers have obtained important structural information about ion- solvent geometries from the study of relaxation times of magnetically active nuclei. The 7Li and 'H relaxation times yielded a Li-O distance of 2.56 i 0.14 A for hydrated Li+ in aqueous solutions.”5 NMR relaxation was also applied to determine the structure of the fluoride ion hydrate and the solvation of I' by formamide.”28 Ion-solvent interactions can be observed directly by vibrational spectroscopy, owing to the faster timescale of this technique.29 Cation hydration has been studied through the metal—oxygen stretching mode {v(M-O)}. The frequency of the v(M-O) vibration in metal hydrates correlates well with the metal-oxygen bond 1ength.3°'3' In addition, the v(M-O) stretch has been used to determine the solvation number of Li+ in various nonaqueous solvents?“32 Luck et al. found that alkali cations shift the v(OH) stretch of H20 in nitromethane up to 60 cm‘l.33 The utility of infrared spectroscopy in the study of anion-solvent interactions is exemplified by the thiocyanate anion. Since thiocyanate is a linear triatomic anion, it has three fundamental vibrational modes: v3 {VCN}, vl {Va}, and the doubly degenerate v2 {850,}. It is important to point out at the onset that the labels VCN and VCS are purely descriptive since the normal modes of these stretching vibrations are the result of displacements in all the atoms in SCN‘.34 Perelygin et al. showed that the vibrational 10 modes of the thiocyanate anion are very sensitive to the nature of the solvent, especially when the solvent can hydrogen bond with the anion.35 As shown in Table 5, hydrogen bonding solvents, such as H20, D20, and MeOH, markedly affect the vCN stretch of the thiocyanate anion. Protic solvents shift the vCN mode to higher frequencies and broaden the spectral envelope. The shift and broadening of the vibrational modes of the SCN' anion were also reported by Koda et al, who observed that the addition of water to a solution of tetrabutylammonium thiocyanate (NBu4SCN) in DMF shifted both the vCN and vcS modes to higher frequencies and broadened both spectral envelopes.36 Bacelon et al. assigned a similar shift in tetrabutylammonium thiocyanate / phenol / CCl4 mixtures to a complex where the phenol molecule is hydrogen bonded to the nitrogen end of the thiocyanate anion.37'39 Table 5. Wavenumber and bandwidth of the VCN mode of the thiocyanate anion in various solvents35 Solvent VCN (cm") AVCN (cm”, F WHH ) (CH3)2NCHO 2058 I 3 (C11,),so 2058 14 11,0 2076 36 13,0 2078 41 CH3OH 2068 5 l Hochstrasser and coworkers studied the effect of hydrogen bonding solvents on the vibrational relaxation of SCN‘ (as well as N3' and OCN‘) in D20 and MeOH, where they found that the relaxation of the VCN mode was consistent with hydrogen bonding to the nitrogen end of the thiocyanate anionm‘ Moreover, neutron diffraction studies of ll NaSCN in D20 and H20 by Kameda and coworkers show that the thiocyanate anion is hydrogen bonded at the nitrogen end of the anion by 1.8 i- 0.2 water molecules with a CN'"H20 angle of 120°.42 Gill et a1. inferred the solvation structure of the thiocyanate anion in a solution of NBu4SCN in methanol from deconvolution of the infrared band envelope of the vCN vibration.43 The spectral envelope was decomposed into six component bands as shown in Figure 2. The analysis was based on the second and fourth derivatives of the vCN absorption, which yielded the position and the number of component bands. These authors assigned the component bands to hydrogen bonding by methanol to the nitrogen end and the ON bond of the thiocyanate anion, based on infrared studies of the association of phenol with various alkenes and alkynes,“4 where the hydrogen bond is pointed to the It electrons of the unsaturated bond. In their assignment (Table 6), a hydrogen bond to the 7: cloud of the CN bond lowers the frequency of the ch stretch, whereas a hydrogen bond to the nitrogen end increases the frequency of the vCN stretch with respect to its value in the free SCN’ anion. These novel assignments are a significant departure from results of previous infrared studies of the thiocyanate anion in protic solvents and in ion pair environments. Moreover, no crystal structures of any metal thiocyanates show such bonding; rather, metal thiocyanate interactions are localized to either the nitrogen or the sulfur end of the thiocyanate anion (vide infra). Chabanel’s group has also studied the vcN stretch of NBu4SCN in MeOH. The vCN mode was found to consist of only four component bands, and assignments of these constituents (Table 7) were based on that group’s previous studies of alkali thiocyanate ion pairs.45 The vCN mode for the free thiocyanate anion is assigned at 2058 cm".45 The lowest frequency mode of the set is assigned to a structure similar to those seen in alkali thiocyanate dimers, where two alkali cations bound to the nitrogen end of the thiocyanate anion reduce the frequency of the vCN mode. A single hydrogen bond to the nitrogen or 12 sulfur end of the thiocyanate anion is assigned as raising the frequency of the vcN stretch, which is analogous to the observations for M+NCS' and M+SCN’ ion pairs. Since both CH3OHm'SCN and CH3OH---’SCN---HOCH3 should increase the frequency of the vCN stretch, the authors were not able to conclusively assign the 2090 cm“ component.” Previous studies of the ionic association of alkali thiocyanates in NMF revealed that the vCN absorption of the thiocyanate anion is broad (~32 cm") and asymmetric.46 Moreover, the peak position and lineshape of the vCN stretch were found to be independent of the cation used. Closer analysis of the spectral envelope suggested that the vCN stretch was a convolution of three bands, at 2066, 2055, and 2045 cm", which are a result of anion-solvent interactions.46 0.20 0.10 - 0.10 - 0.14 0.12 0.10 0.00 - 0.06 - 0.04 0.02 - wavonumbor/cm ' ‘ Figure 2. The vCN stretch of NBu48CN in methanol and the component bands determined by Gill and coworkers for the spectral envelopes“3 13 Table 6. Assignments of the six component vCN bands in tetrabutylammonium thiocyanate in methanol by Gill et al.43 Assignment Peak position (cm") noon, ' _ 2010 S- CE N I I HOCH3 HOCH3 ' _ ‘1 2034 3" CE N S— CE N_ 2054 noon3 I _ 2068 S—CEN --H0CH3 i10CH3 uocn3 I _ 2076 s—CEN --H0CH3 S—CEN—--HOCH3 ano 14 Table 7. Assignment of the four component VCN bands of tetrabutylammonium thiocyanate in methanol by Chabanel and coworkers“5 ‘ Assignment Peak position (cm°’) ,HOCH3 s— CE N" 2042 ‘HOCH3 S—CE N‘ 2058 S-CE N'- - -HOCH3 2070 H3COH- - - “s—cs N 2090 or H3COH- - - 's—CE N - - -H0CH3 D. Ion-Ion Interactions: The vexing problem of electrolyte solutions is the determination of the degree of ionic association and the factor(s) that affect this association. The two most important solvent properties that have been shown to influence ionic association are the dielectric constant and the solvating ability of the solvent.'6 The role of the dielectric constant is obvious from Coulomb's law; solvents with a high dielectric constant are more effective at screening opposing ions. A number of empirical scales have been proposed for a solvent’s 15 solvating ability, such as the donor number and the acceptor number.”47 These scales are a measure of the strength of the ion-solvent interaction. Only solvents with both high solvating ability and high dielectric constants will minimize ion pairing, whereas electrolyte solutions in solvents with either low dielectric constants or low solvating ability will have ionic association. Ionic association in solvents of high dielectric constant (8 > 30) and donicity, has been interpreted in terms of the Eigen mechanism”52 where A+ and B’ refer to the free cation and anion, respectively, and S stands for the solvent molecule. A+SB' is A+ + B' —‘-— A+SB' —"~— A+B' a solvent-separated ion pair and A+B' is a contact ion pair. In solvents of a very low dielectric constant (e < 20) and donicity, the ion pair can aggregate into dimers, (A*B')2, tetrarners, (A*B')4, and higher order multimers, (A*B'),‘.53 In addition, an ion pair A+B' —“— (A+B')2 —‘— (AIB'M _"~— (AIB'L. can associate with either an anion or a cation to form a triple ion, which results in a charged species as illustrated below. A+ + A+B' "—‘~— AIB'A+ Ionic association was first studied by electrical conductance.54 Ostwald and coworkers were able to distinguish free ions from ion pairs, since ion pairs do not conduct current.l Electrical conductance has been recently applied to the study of ion pairing in solutions of tetraalkylammonium salts and cryptate salts.”56 One drawback of this method is that it cannot distinguish contact ion pairs from solvent separated ion pairs. 16 There are a variety of more modern methods to probe solutions. Of these, multinuclear NMR and vibrational spectroscopy have had wide application to the study of ionic association since these two techniques are sensitive to local electrostatic interactions and are applicable to a wide variety of solution problems'w'60 One reason for this broad application of NMR spectroscopy is that almost all elements have a magnetically active isotope.59 Moreover, the resonance range of a given magnetically active nucleus is so large that one rarely encounters overlapped signals from two different nuclei. Ionic association greatly perturbs the electronic environment surrounding a nucleus, which can result in a distinct difference between the chemical shift of the free ion and the ion pair. In addition to this general advantage in specificity and the sensitivity of the chemical shift of a nucleus to a local electric field, there are several special NMR methods that can be used to study solution phenomena, such as nuclear relaxation and nuclear Overhauser techniques.57'58'60 The investigation of electrolyte solutions by NMR was pioneered in the mid- 19608 by Deverall and Richards in their study of the chemical shifts of alkali nuclei in a series of salt solutions?“63 In all cases, the chemical shift depended on the concentration of the salt and the counter ion used. These observations were extended by Bloor and Kidd, who studied the concentration dependence of the 39K chemical shift of fourteen potassium salts in aqueous solution.“'65 In all cases, the concentration dependence of the 39K chemical shift, which results from cation anion interactions, was linear as shown in Figure 3. The concentration dependence of the 39K chemical shift results from cation-anion interactions. The linear dependence of the 39K chemical shift upon electrolyte concentration is not indicative of contact ion pair formation, rather, it has been interpreted as collisional ion pairs.57 In solvents of intermediate and low dielectric constants (8<30), contact ion pairs exist.”70 The existence of contact ion pairs is observed in NMR spectroscopy by a nonlinear dependence of the chemical shift on the solute concentration.” In the case of 1:1 17 0 v ”K chanted mm I”... i -l o . l A A A A A A 1 2 3 4 5 6 7 Potassium ion W’ mm 1120) Figure 3. Concentration dependence of the ”K chemical shift in various solvents"3 electrOlytes undergoing exchange between free ions (M+ and X‘) and an ion pair (M+X‘), the observed chemical shift can be expressed as a function of the mole fraction of the free A + + B' = A+B' ion (Xf) and the ion pair (Xip) and their respective chemical shifts, Sr and Sip. Dabs = Xfo + Xipfiip = 5f + Xip(8ip - 5f) (1) The ion pair formation constant is given by 18 . = (M+X') = C,-Cf ‘P (M*)(X‘> (3.21. (2) where Ct is the total electrolyte concentration, Cf is the concentration of the free ion, and Y: is the mean ion activity coefficient of the salt. These two equations can be combined so that 508. is a function of total electrolyte concentration, C,, with the parameters Kip, 719 5f, —1 i (1 + 4K,I,y,2C,)“2 KipYiZCt 5bs = O (5.p - 5.) +5. (3) and 61p. The ion pair formation constant can be determined by fitting equation 3 to the observed chemical shift and the concentration of the electrolyte. The concentration dependence of the 23Na chemical shift of NaBr, NaSCN, NaBPh4, and NaClO4 in dimethylforrnarnide are illustrative of ion pairing.7| As shown in Figure 4, the chemical shift of 23Na is nonlinear and of the form given by equation 3. The relaxation time of a magnetically active nucleus can also be used to study ion pairing?9 Hertz et al. showed that the ”F relaxation time is extremely sensitive to ion pair formation. By measuring the concentration dependence of the relaxation time of the I9F resonance in a solution of (CH3)4N*F in water or D20,”73 Hertz was able to determine that the fluoride anion is ion paired in these solvents, even at low concentrations.73 Likewise, vibrational spectroscopy and more specifically infrared spectroscopy, provides another perspective for studying ion-ion interactions. A significant advantage of vibrational spectroscopy over magnetic resonance is the timescale of the measurement. Since the lifetime of an ion-ion complex is on the order of 1098, ion pairs are not directly observed by NMR spectroscopy; however IR spectroscopy with its faster timescale can observe ion-ion aggregates directly29 via distinct vibrational bands corresponding to 19 different chemical environments.49 The strong electrostatic interactions within ion pairs that affect the chemical shift of probe nuclei also perturb the vibrational modes of a molecular ion involved in an ionic association. 5 oh __A_. . wk. .\.\ .\o \ .\- 4-1: \- \- \I SCN ut-t Nd! U 3—1 u 10:! Figure 4. Concentration dependence of the 23Na chemical shift of NaSCN, NaBr, NaBPh4, NaClO4 in DMF7| Although infrared spectroscopy can directly observe ions in multiple chemical environments, the bands that result tend to be overlapped as shown in Figure 5.5' In order to obtain quantitative information about each chemical species, the observed spectral envelope must be deconvolved into its component bands. A common decomposition procedure is to fit the experimental spectrum with synthetic Gaussian and/or Lorenztian component bands.74 Petrucci et al. used Voight profiles (a Gaussian-Lorenztian product) in order to fit the band envelope of the vcN stretch of N aSCN in THF. They were able to 20 identify three component bands that were assigned as a NaNCS ion pair (2057 cm"), a dimer (2043 cm'l), and a triple ion (2074 cm'l), NaNCSNa".5| Figure 5. Experimental infrared band envelope for the VCN stretch of NaSCN in THF, with calculated component bands51 Chabanel and coworkers studied the ion pairing of alkali cyanates in dimethylsulfoxide by following the vCN mode of the cyanate anion.” As shown in Table 8, complexation of the cyanate anion with an alkali metal markedly affects the position and intensity of the VCN band. The association was assigned as a M‘NCO’ interaction, with the cation bound to the nitrogen end of the cyanate anion. This assignment is based on earlier work with alkali thiocyanates.” The thermodynamics of these ion pairings were also determined, and are summarized in Table 9. As is generally the case with ion pair formations, the process is entropy driven. 21 Table 8. Spectroscopic parameters for the VCN mode of the cyanate anion and alkali cyanates in DMSO75 Smies Position (cm”) Av,,L(FWHH, cm") Intensity (M" cm") OCN' 2138 10.3 2150 LiNCO 2172 12.4 1920 NaNCO 2159 14.3 1650 KNCO 2148 13.1 1770 RbNCO 2 145 a a CsNCO 2142 a a adata not reported The small enthalpy change accompanying the formation of an ion pair arises from the Coulombic attraction between the cation and the anion, which produces a negative enthalpy contribution, plus the enthalpy required to break the ion-molecule bonds involved in ion solvation. In the alkali cyanate examples, the strong Coulombic attraction in the ion pair is balanced by energy needed to disrupt the structure of the solvated ions. The positive entropy contribution is the result of the loss of entropy due to the ion pair formation being more than balanced by solvent molecules being liberated from the solvated ions. The increased freedom of the solvent molecules is the predominant entropy contribution. The study of the thermodynamics of ion pairing shows how important the solvent is in ion pair formation. Chabanel and coworkers have also studied the aggregation of LiNCO ion pairs to form dimers and triple ions.7‘5'77 Addition of LiClO4 to a solution of LiNCO in DMSO results in a new vibrational band at 2203 cm" and a reduction of the intensity of the ion pair band at 2172 cm" in the vCN region. The 2203 cm‘1 band has been attributed to a triple ion 22 which may be LiNCOLi“, LizNCO“, or Li20CN+ since Chabanel and coworkers were not able to conclusively identify this species.76 The same group has followed the infrared spectrum of OCN' as function of LiNCO concentration in DMSO; new bands are observed at 2189 cm", and 1330 cm’l at high LiNCO concentrations. Chabanel and coworkers have assigned these bands to the lithium cyanate dimer, (LiNCO)2.77 Table 9. Thermodynamics of alkali cyanate ion pairs in DMSO at 298K75 Sfles K (M") AG" (kJmol’) AH" (kJmol’) AS" (JK"mol") LiNCO 110:1:10 -ll.6i0.3 1.2:121 43:1:3 NaNCO 46 i 5 -9.5 :1: 0.5 3.2 i 1.5 43 i 5 KNCO 16 i 4 -6.9 :l: 0.8 0.5 :1: 1.5 25 :t 5 The sensitivity of the vCN stretching mode of the cyanate anion to its local environment can be seen by comparing the solution studies of Chabanel and coworkers with the gas phase studies of KNCO by Devere.78 The VCN stretch of KNCO in the vapor phase consists of two bands at 2207 cm" and 2185 cm". In order to assign these bands, Devere studied the vibrational spectrum of KNCO at several temperatures. Analysis of the vapors by mass spectrometry shows that it is a combination of monomer and dimer KNCO, and that the relative intensity of the (KNCO)2 correlates with the 2185 cm'1 infrared band. This is in contrast to the solution phase studies of the vCN stretch in lithium cyanate, where the absorption due to dimers is lower in frequency than the ion pairs. Moreover, the 59 cm‘1 difference between the VCN stretch of the KN CO ion pair in the gas and solution phase reveals a substantial difference in polarization of the cyanate anion in these media. 23 From these studies, it can be seen that the vCN stretch is quite diagnostic of the local environment surrounding the cyanate anion. Ionic association affects the frequency, the bandwidth, and the intensity the vCN mode. In solution, the frequency of the vCN vibration increases in the order: OCN' < (MNCO)2 < MNCO < triple ion. The ionic association of alkali thiocyanates has been studied in many nonaqueous solvents.”82 Several types of ionic aggregates have been identified from their vibrational spectra. These results, summarized in Table 10, show the assignments for ion pairs, dimers, triple ions and tetramers. It is interesting to note that the vCN and vCS modes are not shifted in the same direction upon complexation. In fact, this observation has been helpful in identifying the structure in unknown metal thiocyanate complexes.83 Ion pairing in alkali metal thiocyanates, which bond to the nitrogen end of the thiocyanate anion, increases the frequency of both the vCN and vCS stretches. Ionic association to the sulfur end of the thiocyanate anion increases the frequency of vCN mode but decreases the frequency of vCS mode. Triple ions, where a metal cation is bound to each end of the thiocyanate anion, have a VCN frequency above that of the corresponding ion pairs. Dimerization and tetramerization of the ion pair reduces the frequency of the vCN mode below that in the free anion . Pictorial views of the LiNCS ion pair, dimer, and tetramer are shown in Figure 6. Through the careful study of many thiocyanate ion pairs, Chabanel et a1. has been able to correlate the frequency of the vCS stretch with the polarizing power of the metal cation. As shown in Figure 7, the polarizing power of the metal cation in a metal thiocyanate ion pair is linearly related to the difference between the vCS stretch in the ion pair and the vCS stretch in the free thiocyanate anion.85 In the extreme example of Al“, the vCS stretch of the thiocyanate anion increases by more than 100 cm". Chabanel and coworkers explain the changes in the vibrational frequencies of the thiocyanate ion in ion pairs through the resonance structures that can be drawn for the thiocyanate anion. When a positive ion interacts with the nitrogen end of the thiocyanate anion, the polarization in the 24 Table 10' complexes 0f metal thiocyanates and their vibrational spectra in various aprotic sol vents79'82'84 Assignment vCN (cm") Vcs (cm'l) SCN"I 2056 735 Ion Pairs LiNCS‘I 2072 765 NaNCS‘ 2066 754 KNCS‘ 2056 742 RbNCSa 2056 742 CsNCSa ; 2056 742 AgSCNb I 2080 721 Triple Ions Li’SCN'Li“ 2087 d Na‘SCN'Na‘c 2081 “ Dimers (LiNCS)2c 2034 788 (NaNCS)2c 2044 769 (KNCS);e 2042 758 Tetramer (LiNCS); 1993 ‘1 “in DMF, bin DMSO, ° nitromethane, dobscured by solvent and cin THF 25 ”-0-. O 8 Ion Pair 0 C O 0 Li Dimer Tetramer Figure 6. Pictoral view of the LiNCS ion pair, dimer, and tetramer85 anion shifts so that the negative charge on the nitrogen increases. As seen in the resonance structures (Figure 8), the greater negative charge on the nitrogen of the thiocyanate anion is accompanied by a reduction in the bond order of the CN bond and an increase in the bond order of CS; the latter will give rise to an increase in the frequency of the of the vCS stretch. Likewise, a positive electrostatic interaction to the sulfur end of the thiocyanate anion will result in an increase in negative charge on the sulfur atom. This change in the polarization in the anion will decrease the bond order of the CS bond and thereby lower the frequency of the vCS stretch. This is observed for AgSCN ion pairs.80 The changes in the frequency of the VCN stretch that arise from ionic association cannot be explained simply through the change in polarization of the anion since the frequency generally increases whether the electrostatic interaction is on the nitrogen end or the sulfur end of the thiocyanate anion. There have been several attempts to explain why the VCN stretch does not reflect the resonance structures.”88 Recently, Chabanel and A13” 850 — in”. ‘68“ OZnZ+ 'e 3 8 800 ~ ca“ ’ oMg2+ 2+ 2* Bo . CO Li+ No+ 750 - ”<2 .RD 1 4 l P Figure 7. Frequency (cm") of the vCS stretch in alkali and alkaline earth thiocyanate ion pairs as a function of the polarizing power of the metal cation85 Figure 8. Resonance structures for the thiocyanate anion 27 coworkers noted that there are two effects that can contribute to the shift of the vcN mode in alkali and alkaline earth thiocyanate ion pairssz'gs First, the stabilization of the 0 molecular orbitals, especially near the cation, in the M"NCS' ion pairs causes a strengthening in the CN bond. Second, polarization from the sulfur to the nitrogen lowers the C-N bond order as shown by the resonance structures. The authors derive this argument from their study of a variety of metal thiocyanates, where they were able to correlate the shift in the vCN mode to the polarizability of the metal ion (Figure 9) and the polarizing power of the metal cation, using the following relation: AVCN = AP + BaP.82 Here, AVCN is the frequency shift of the vCN mode from the value of the free SCN' in 98% THF and 2% DMF.82 A and B are the fitted parameters with P and a the polarizability and polarizing power of the metal .Tl Figure 9. VCN/P of various metal thiocyanates correlated to the polarization of the metal cations82 28 cation, respectively. The resulting vCN stretching frequency of a given metal thiocyanate ion pair reflects a balance of the stabilization in the o orbitals and the reduction of the bond order by the change in the polarization of the anion.”85 This balance is observed best in the alkaline earth thiocyanates. The vcN stretch in the Be(NCS)+ and Ca(NCS)“ ion pairs shift to higher frequency whereas the vCN stretch in a Ba(NCS)+ ion pair is at a lower frequency than that of the free thiocyanate anion.82 The authors explain this difference in the alkaline earth thiocyanates by proposing that the stabilization of the 0' orbitals falls off faster with distance than the resonance effect, so that at long distances (or large cations) the resonance effect predominates and lowers the frequency of the vCN stretch. The VCN stretch of the dimers and tetramers are also lower than the free thiocyanate anion. In these complexes, the coordination of the metal cations to the nitrogen end of the thiocyanate anion is angular rather than linear, which results in an increase in the polarization of the charge to the nitrogen end of the thiocyanate anion. The stabilization of the 0' orbitals will be lessened since there is less overlap between the positive charge and the lone pair of the nitrogen. The net result will be that the resonance effect will predominate and the frequency of the VCN modes in these species will be lower than in the free ion.82 Chabanel and coworkers have modeled the ion pairing, dimerization, and tetramerization in lithium thiocyanates through semi-empirical MN DO calculations.89 Although these results are not quantitatively reliable, the calculated vCN stretch reflects the same trend seen in their previous experimental studies, where higher aggregation of LiN CS reduces the frequency of the vCN stretch (T able 11).89 The energies of all the molecular orbitals in the thiocyanate anion are stabilized by 5-6 eV in the ion pair and by an additional 0.5 eV in the dimer and tetramer, with respect to the thiocyanate anion. Moreover, complexation of the thiocyanate anion by lithium increases the calculated bond length of C-N from 1.177 A in the anion to 1.204 A in the dimer. The calculated atomic charge on the sulfur atom is more positive in the order of tetramer > dimer > ion pair > free anion. 29 The atomic charge on the sulfur and the ON bond length reveal that a positive electrostatic interaction on the nitrogen end of the thiocyanate anion results in change in the polarization of the anion that can be explained by the resonance structures. Although the MNDO calculations qualitatively reflect the electrostatic effects on the vCN stretch of the thiocyanate anion, this level of computation poorly predicts the vCS mode in these complexes. Despite the poor performance of the MN DO calculations on the vCS stretch, they provide support for Chabanel's model of the electrostatic effects on the stretching modes of the thiocyanate anion. Table 11. MNDO calculations of the Von stretch in LiNCS complexes compared with the experimental frequency85'89 Complex Calculated V0,, (cm") Experimental VCN (cm") LiNCS 2313 2072 (LiNCS)2 2244 2034 (LiNCSM 2207 1993 The thermodynamics of the formation of several alkali thiocyanate ion pairs and dimers were also determined; they are summarized in Table 12.70 Again, the ion pair formation is entropically driven with a small positive enthalpy component. This large entropy change results from the freeing up of solvent molecules that surround the unpaired ions, thus providing greater overall disorder in the solution. The same is true for the dimer and tetramer formation where the aggregation is also entropy driven. In fact, the larger entropy term of the dimers and the tetramers is a result of greater release of solvent molecules. The notable exception of the KNCS dimer is due to the poor solvation of the 30 K+ cation which is demonstrated by the small entropy increase as a result of the dimer formation. The ionic association of LiSeCN in several solvents has been studied by infrared spectroscopy. Chabanel and coworkers used the vCN mode of SeCN‘ to identify the free SeCN‘, the LiNCSe ion pair and the dimer, as shown in Table 13.”91 Ionic interactions with SeCN' have a similar effect on the vCN mode as observed for the thiocyanate anion. Ion pairing raises the frequency of the vCN stretch, whereas dimerization and tetramerization lower the frequency of the VCN vibration. Table 12. Thermodynamic parameters for the ion pair, dimer and tetramer formation of alkali thiocyanates in various solvents70 Solvent Salt Kass (M' 1) AH" AS” (It/mole") (JK’Imole'l) Ion Pair DMF LiNCS 2.4 3.5 34 Dimer THF LiNCS 0.24 23 66 NaNCS 45 l 37 KNCS 60 -3 2 Tetramer BuOEt LiNCS ‘ 29 l 30 ‘data not reported E. Model Crystal Structures Several of the structures proposed for metal thiocyanate complexes in solution have been confirmed through the use of single crystal X—ray crystallography. Clearly, the 31 crystal structure will not provide a perfect model of solution structure since it can emulate only a limited number of the broad range of dynamic solute-solvent geometries present in a solution. In spite of this drawback, a variety of solution interactions can be observed in solids, such as ion-ion, ion-solvent, and solvent-solvent interactions. Since these types of interactions in the crystal lattice can be easily varied by the appropriate synthesis, this method can correlate a give crystal structure with complexation in solution by comparison of the vibrational modes in each system. Table 13. Frequency (cm") of the vCN mode of LiSeCN in various solvent9| Solvent SeCN' LiNCSe (LiNCSe); (LiNCSe), DMF’ 2066 2080 f ' MeCNb 2068 208 l f ‘ DEC‘ ‘ 2073 2040 ’ Bu20d ‘ 207 l 2029 ' TEA‘ ' f r 2000 ‘dimethylformamide, bacetonitrile, cdiethylcarbonate, dbutylether ‘triethylamine, rnot observed or minor components To understand the effect of ion pairing and higher order aggregation on the thiocyanate anion in the solid state, the structure of a 'free' thiocyanate anion must be obtained. Since the principle of electroneutrality must be maintained, a 'free' thiocyanate anion can never be observed in the absence of a cation. As in the solution state, the perturbation of the thiocyanate anion by a cation can be minimized through the use of bulky cations. Crown ethers and cryptands have long been used to isolate cations from anions.92 32 LiSCN complexed by 4,7,13-tri0xa-l,10-diazabicyclo[8.5.5]icosane is just such a system.93 According to the analysis of the crystal diffraction, the cryptand surrounds the lithium cation in a three-dimensional structure so that the thiocyanate lies beyond the van der Waals radius of the cation. The positive charge is well screened from the thiocyanate anion, so this crystal provides a good model of the ‘free’ SCN‘ in the solid state. The CS bond length is 1.72 i 0.03 A and the CN bond length is 0.97 i 0.02 A.93 The ion pair interaction and its effect on the geometry of the thiocyanate anion can also be observed from crystal structures. The lithium thiocyanate ion pair is modeled by the crown ether complex of LiSCN with 1,5,9-trioxacyclododecane (9-Crown-3) shown in Figure 10.94 The lithium cation is coordinated to the nitrogen end of the thiocyanate anion. The bond length of the CS bond is reduced to 1.635 :1: 0.006 A, the CN bond is extended 0.0 0 Figure 10. Li(9-Crown-3)SCN93 to 1.160 :1: 0.006 A, and the thiocyanate anion is almost linear at 178.3 i 05°. The lithium cation is 1.958 10.009 A away from the nitrogen end of the thiocyanate anion and forms a 169.4 :1: 06° angle with the CN bond. The sulfur atom of the thiocyanate anion is not within the van der Waal radius of any other atom. Bright and Trutes synthesized a similar ion pair of RbNCS with 2,3,11,12-Dibenzo-1,4,7,10,13,16-hexaoxocyclo-octadeca-2,11- 33 diene (Dibenzo-l8-Crown—6).95 The ion pair structure is similar to the Li(9-Crown-3)SCN structure.94 The rubidium cation is surrounded by the crown ether and the thiocyanate anion is coordinated to the cation through the nitrogen end. In the rubidium complex, the CS bond length is 1.58 i 0.03 and the CN bond length is 1.16 i 0.03; the thiocyanate anion is significantly distorted from linearity, with an angle of 163 i 2°. The Rb-N bond length is 2.94 d: 0.02 A and the Rb-N-C angle is 108.8 i- 0.02.. The sulfur atom of the thiocyanate anion is 3.6-3.8 A from three of the carbon atoms belonging to one of the benzene rings in the crown ether. The crystal structure of the UN CS complex supports that proposed for the LiNCS ion pair in solutions, it is almost linear as would be expected, and the perturbation of the anion's geometry follows the same trend as previous MNDO calculations.89 On the other hand, the rubidium complex provides an example of the care that must be exercised in drawing inferences to solutions from crystal structures, since packing forces and other interactions, such as the benzene-thiocyanate anion interaction, may occur in addition to the ion-ion interaction that one may wish to model. A crystal structure having a distorted dimer-like structure was prepared by Armstrong et al.96’97 The structure of this lithium thiocyanate tetramethylpropylene diarnine (TMPA) is a distorted dimer that resembles two monomers in a loose association. This structure differs from the symmetric dimer proposed by Chabanel85 where each thiocyanate anion tilts from one cation towards another; both are shown below in Figure 11. The CN bond lengths in the asymmetric dimer are 1.146 and 1.153 i0.002 A. The CS bond lengths are 1.601 and 1.606 i 0.002 A. The shorter Li-N distances are 1.978 and 1.992 i 0.003 A with Li-N-C angles of 167.0 i 0.20 and 159.8 i 0.20, respectively. The longer Li-N distances are 2.061 and 2.095 i 0.003 A with Li-N-C angles of 115.3 and 108.8 i 0.2°, respectively. In addition to solving the crystal structure for this complex, the authors studied the vCN mode of the thiocyanate anion, which was measured at 2033 cm'1 in the solid state. Moreover, infrared studies of dilute solutions of this complex in benzene, where the dimer has dissociated into monomers, show that the vCN stretch is unaffected by 34 the monomer-dimer equilibrium and remains at 2033 cm". N o infrared bands in the 2060— 2070 cm'1 range, which would be indicative of LiNCS ion pairs, were observed in these solutions. This careful approach of correlating the structure of solutions with the structure of crystal complexes through the infrared vibrations shows that previously inferred solution structures may be in error. (I) I III 111 ‘P (I) / 01 Li \ TMP 3’4, > Figure 11. Symmetric (LiNCS)2 dimer (left) and asymmetric [Li(TMPA)2NCS]2 dimer (right)”97 A more symmetric dimer structure was synthesized from KNCS and 6,8,9,10,12,13,20,21,23,24,26,27-dodecahydrodibenzo[b,n]-1,4,7,10,13,16,19,22- octaoxacyclotetracosin (Dibenzo-24-crown-8).98 In this interesting complex shown schematically in Figure 12, two potassium cations are coordinated by a single crown ether. The CN bond length is longer than the LiNCS ion pair at 1.18 i 0.01 A and the CS bond length is shorter than the LiNCS ion pair at 1.60 i 0.01 A. The two K-N-C bond angles are 130.3 :1: 03° and 136.3 ‘1: 03°. The triple ion [MNCSM]+, which has been identified in solution studies, is supported by the KSCN acetonitrile oxide heptamer crystalline complex.99 In this complex each [C is surrounded by a nitrile oxide ring. The thiocyanate anion forms a linear bridge to two potassium cations through bonding at the nitrogen and sulfur ends 35 Figure 12. Diagrammatic representation of the coordination in (KNCS)2-DB24C899 Additional support is seen in bis(N-methylformamide) tetrakis(thiocyanato) cobalt(II) mercury(lI),100 where the thiocyanate anion bridges the cobalt and mercury cations. The cobalt cation interacts with the thiocyanate anion through the nitrogen end, the mercury cation through the sulfur end. The CN bond length is 1.10 i 0.03 A and the CS bond length is 1.67 :1: 0.02 A The crystal structures of these thiocyanate complexes show that ionic association perturbs the geometry of the thiocyanate anion in a way that is unique to each complex. The perturbation of the thiocyanate anion by metal cations in solution has long been studied by vibrational spectroscopy in order to identify possible structures of ionic association complexes. The limited examples where the vibrational spectroscopy of known structures 36 with vibrational spectroscopy of unknown systems has shown that one can easily infer the wrong structure from solution studies alone. More work needs to be done to correlate the vibrational spectroscopy of crystal alkali thiocyanate complexes with solution values to . determine if there is a strong relationship. F. Proposed Work The structure of anion-solvent complexes and the thermodynamics of anion solvation are poorly understood, as exemplified by the investigations of the solvation of the thiocyanate anion where there are some conflicting results. The research described in this dissertation utilizes a variety of experimental and computational methods to help determine the structure and thermodynamics of anion-solvent complexes. The thiocyanate anion, along with the closely related OCN' and SeCN' anions, will be used to probe anion-solvent interactions. Methanol and the amide solvents, NMF and FA, will be employed to allow a wide variety of hydrogen bonding interactions. DMF is utilized as a reference. The principal investigations of these interactions will be accomplished by a combination of NMR and infrared spectroscopies. A qualitative picture of the number of solvated species and the structures of the ion-solvates will be obtained by studying each anion in the neat solutions of the hydrogen bonding solvents as well as in less interacting solvents. Quantitative determination of the thermodynamics of the interactions will be made with the help of an ‘inert’ solvent, nitromethane, where the anion and hydrogen bonding solvent concentrations can be independently varied. Ab initio calculations of selected ion- molecule species will be performed to help in the assignment of infrared bands and to provide insight into the nature of the hydrogen bonding interactions. Finally, crystal structures of appropriate solvate structures will be synthesized and their structures solved to add additional support to proposed crystal structures and vibrational assignments. Chapter 11 Experimental Methods A. Materials Methanol (absolute, EM Science, 500ml) was dehydrated by refluxing over CaH2 (5 g) for two hours. The dried solvent was distilled under a dry N 2 atmosphere onto 4A molecular sieves (Linde). The fast 50 ml of the distillate was discarded and the next 300 ml was collected. N,N-dimethylformarnide (99%, EM Science), N-methylforrnamide (99% Aldrich), and formamide (99% EM Science) were dried by adding 5 g of BaO (Baker) to 500 ml of each solvent. Each mixture was allowed to sit for twelve hours, then distilled over fresh BaO under reduced pressure onto 4A Molecular sieves. Nitromethane (96%, Aldrich) was dried by the addition of 5g of CaH2 to 500 ml and allowed to sit for 24 hours. The dried solvent was distilled under reduced pressure onto fresh CaHz. All solvents were stored in dark glass bottles in a dry box under a nitrogen atmosphere. Tetrabutylammonium cyanate (>96%, Fluka) was dissolved into ethyl acetate, and filtered to remove insoluble material. The solvent was removed and the purified cyanate salt was dried under vacuum over P205 at 45°C for 24 hours. Tetrabutylammonium thiocyanate (>99%, Fluka) was dried in a vacuum oven over P205 at 45°C for 24 hours. Tetrabutylammonium selenocyanide was prepared from tetrabutylammonium bromide (99% Aldrich) and potassium selenocyanide (98%, Aldrich). Potassium selenocyanaide was dissolved in acetone and the solution was filtered to remove insoluble material. The potassium selenocyanide was recovered from the acetone solution through precipitation by ethyl ether. Tetrabutylammonium selenocyanide was prepared by mixing equirnolar amounts of tetrabutylammonium chloride and potassium selenocyanide in acetone. The solution was filtered to remove the KC] precipitate and the solvent was removed from the filtrate to obtain NBu4SeCN. The crude tetrabutylammonium selenocyanide was recrystalized in bis(2-ethoxy)ethane (Eastman). The purified tetrabutylammonium selenocyanide was dried in a vacuum oven at 45°C for 24 hours. NaSCN (99.99%, Baker) was dried at 120°C for 24 hours. 38 39 Solutions with known concentrations were prepared by weighing the solute on an analytical balance into a 5 or 25 ml volumetric flask and diluting with the appropriate solvent. All solution preparation was done in a dry box under a nitrogen atmosphere. Ternary solutions of the tetrabutylammonium salt and hydrogen bonding solvent in nitromethane were prepared from stock nitromethane solutions of the salt or the hydrogen bonding solvent. Aliquots (l, 2, 3, and 4 ml) of the salt or hydrogen bonding solvent solution were pipetted into 5ml volumetric flasks so that the concentration of the salt and of the hydrogen bonding solvent could be varied independently. Crystal solvates of NaSCN-NMF and N aSCN-DMF were prepared by slow evaporation of saturated NaSCN (99.4% Baker) salt solutions in each solvent in a dry N 2 atmosphere. Na(12C4)ZSCN-MeOH was prepared by dissolving stoichiometric amounts of N aSCN and 12-crown-4 (12C4, Aldrich, Inc.), in a 1:2 mole ratio, into a minimum of dry MeOH. The solution was diluted to 15 ml with bis(2—ethoxy)ethane (Eastman). The mixtures were allowed to slowly evaporate in a dry N2 atmosphere until the initial formation of crystal solvates was observed. Once crystal solvates were visible, the solutions were capped in order to slow the crystal growth. All crystals were harvested from the mother liquor and placed in a sealed glass capillary. Likewise, Na(12C4)ZSCN was prepared by dissolving stoichiometric amounts of NaSCN and 12-crown-4 (12C4, Aldrich, Inc.), in a 1:2 mole ratio, into a minimum of dichloroethane (Baker) and diluted with bis(2-ethoxy)ethane. The crystal structures were solved by Dr. Don Ward, the Crystallographic Facility Manager. 8. Spectroscopic Techniques Infrared spectra were recorded on a N icolet 520P (N icolet, Inc.) FT IR spectrometer. Three sets of windows were used to cover the 4800-400 cm‘l spectral region: Can (4800-900 cm'l), Irtran-2 (4800—700 cm'l), and Polyethylene (650-400 40 cm‘l). The resolution in all cases was 1 cm‘l- The number of scans was varied from 200 to 10000 depending on the signal intensity of the band studied. Teflon spacers were used to vary the pathlength of the cell, the pathlength value being determined by the fringe method.'°' The cell was flushed with 1-2 ml of the new sample to be studied between each sample, and the spectrometer was purged for five minutes prior to the collection of a new spectrum Multinuclear NMR spectra ('H, 13C, MN, '70, and 77Se) were obtained on VXR- 300 and a VXR-SOO FT-NMR spectrometers (Varian, Inc.). Each sample was placed in a 5 mm NMR tube with a coaxial insert containing an external reference and lock solvent. The parameters used for each nucleus are summarized in the following table. In addition, the 170 spectra were 64K zero filled in order to increase the resolution. The temperature probe in the spectrometer was calibrated by the known temperature dependence of the proton chemical shifts in methanol. Table 14. Parameters for NMR spectra Nucleus Acquisition Time Dely Reference 'H l s l s 1% TMS in CD3C1 13C l s 8 s 1M NBu4NO3 in (CD3)2CO 1“N 1 s 0 8 1M NBu4NO3 in (CD,)2CO ”O 0.025 s 0 8 D20 77Se 1 s 9 8 1M diphenylselenide in (CD,)2CO A crystal of bis(N-methylformamide)-sodium thiocyanate, CSHmN3OZNaS, having approximate dimensions 0.08 x 0.10 x 0.42 mm, was mounted on a Riguaku AFC6S diffractometer with graphite monochromated MoKor radiation and a 2 kW sealed-tube generator. Cell constants and an orientation matrix for data collection, obtained from a 41 least-squares refinement using the setting angles of 20 carefully centered reflections in the range 15.10 < 20 <19.85° corresponds to a monoclinic cell with dimensions: a=7.597 (6) A, b=6.454 (5) A, 0:20.255 (4) A, v=922.1 (9) A, 13:92.58 (3)°. For 2:4 and F.W.=l99.20, the calculated density is 1.334 g/em3. Based on systematic absences of hOl: 1:1th and 0k0: k¢2n and the successful solution and refinement of the structure, the space group was determined to be P2,/c. The data were collected at a temperature of -90 i 3°C using the 00-20 scan technique to a maximum 20 value of 500°. Omega scans of several intense reflections, made prior to data collection, had an average width at half-height of 032° with a take-off angle of 60°. Scans of (0.89 + 0.30tan0)° were made at a speed of 8.0°/min (in omega). The weak reflections (I <10.00(I)) were rescanned (maximum of 3 rescans) and the counts were accumulated to assure good counting statistics. Stationary background counts were recorded on each side of the reflection. The ratio of peak counting time to background counting time was 2: 1. The diameter of the incident beam collimator was 0.5 mm and the crystal to detector distance was 285.0 mm. Of the 2077 reflections which were collected, 1926 were unique (Rim = 0.023). The intensities of the three representative reflections which were measured after every 50 reflections declined by -0.60%. A linear correction factor was applied to the data to account for this phenomenon. The linear absorption coefficient for MoKa is 3.2 cm". An empirical absorption correction, based on azimuthal scans of several reflections, was applied which resulted in transmission factors ranging from 0.95 to 1.00. The data were corrected for Lorentz and polarization effects. A correction for secondary extinction was applied (coefficient = 0.6501 1x 10°). The structure was solved by direct methodsm'103 The non-hydrogen atoms were refined anisotropically. The final cycle of full-matrix least-squares refinement was based on 1170 observed reflections (I > 3.00 o(I)) and 162 variable parameters and converged 42 (largest parameter shift was 0.01times its esd) with unweighted and weighted agreement factors of: R = Ell Fol - chll / ZlFol = 0.034 Rw = [(2 w(l Fol - chl)2 / Z wFo2 )]”2 = 0.041 The standard deviation of an observation of unit weight was 1.49. The weighting scheme was based on counting statistics and included a factor (p = 0.03) to downweight the intense reflections. Plots of [(2 w(| Fol - chl)2 versus lFol, reflection order in data collection, sin0/A, and various classes of indices showed no unusual trends. The maximum and minimum peaks on the final difference Fourier map correspond to 0.21 and 0.26 e'lA3, respectively. Neutral atom scattering factors were taken from Cromer and Waber.‘°" Anomalous dispersion effects were included in Fcalcf"5 the values for Af’ and At” were those of Cromer.10° All calculations were performed using the TEXSAN crystallographic software package of Molecular Structure Corporation.107 A colorless plate crystal of bis(N,N-dimethylformamide)-sodium thiocyanate, C7HMN302NaS, having approximate dimensions of 0.1 x 0.22 x 0.38 mm, was mounted in a glass capillary. All measurements and data analysis was the same except for the differences described below. Cell constants and an orientation matrix for data collection, obtained from a least- squares refinement using the setting angles of 22 carefully centered reflections in the range 13.27 < 20 < 22.63°, correspond to an orthorhombic cell with dimensions: a=9.626:l:0.002 A, b=17.869i0.004 A, e=6.6701~0.007 A, v=1147:1 A3. For 2:4 and F.W. = 227.26, the calculated density is 1.315 g/cm’. Based on the systematic absences of Okl: k¢2n and h01: l¢2n packing considerations, a statistical analysis of intensity 43 distribution, and the successful solution and refinement of the structure, the space group was determined to be Pbcm. A total of 1211 reflections was collected. The intensities of the three representative reflections which were measured after every 150 reflections declined by 1.50%. A linear correction factor was applied to the data to account for the phenomena. The final cycle of full-matrix least squares refinement was based on 665 observed reflections and 102 variable parameters and converged with R = 0.065 and Rw = 0.072. A crystal of N a(12C4)2SCN, CI7H32NNaOSS, having approximate dimensions 0.17 x 0.35 x 0.40 mm, was mounted on a Riguaku AFC6S diffractometer with graphite monochromated MOK01 radiation and a 2 kW sealed-tube generator. Cell constants and an orientation matrix for data collection, obtained from a least-squares refinement using the setting angles of 20 carefully centered reflections in the range 15.10 < 20 < 19.85° corresponds to a monoclinic cell with dimensions: a =15.745 (3) A, b=l4.445 (3) A, e=19.454 (3) A, v=4411.2 (14) A3, 01:90", [1:94.47 (2)°. For 2:8 and F.W.=433.49, the calculated density is 1.305 g/cm3. The space group was determined to P21/n. The data were collected at room temperature (~24°C) using the 03-20 scan technique to a maximum of 500°. The weak reflections (I<10.00'(I)) were rescanned (maximum of 10 rescans) and the counts were accumulated to assure good counting statistics. Stationary background counts were recorded on each side of the reflection. The ratio of peak counting time to background counting was 2:1. The diameter of the incident beam collimator was 0.5mm and the crystal detector was 285.0 mm. Of the 6011 reflections collected, 5765 were unique (Rint = 0.0756). The intensities of the three representative reflections which were measured after every 50 reflections declined by 50%. A sixth order correction factor was applied to the data to account for this phenomena. The linear absorption coefficient for MoKor is 3.2 cm". An empirical absorption coefficient, based on azimuthal scans of several reflections. The data were corrected for Lorentz and polarization effects. A correction for secondary extinction was applied (coefficient=0.6501 1x10'°). The structure was solved by direct methods.1 '3" ‘4 The non-hydrogen atoms were refined anisotropically. The final cycle of the full-matrix least squares refinement was based on 5671 observed (I>2.000'(I)) reflections and 762 parameters and converged (the largest parameter shift was 0.01 times its esd) with unweighted and weighted agreement factors of: R=211F.1-1F.11/21F.1=0.0881 Rw =10: w(| F01 — | F, |)2/Zw F02]"2 = 0.2020 A crystal of Na(12C4)ZSCN-HOMe, CI8H36NNaO9S, having approximate dimensions 0.25 x 0.25 x 0.10 mm, was mounted on a Riguaku AFC6S diffractometer with graphite monochromated MoKor radiation and a 2 kW sealed-tube generator. All measurements and data analysis were the same except for the differences described below. Cell constants and an orientation matrix for data collection, obtained from a least-squares refinement using the setting angles of 20 carefully centered reflections in the range 15.10 < 20 < 19.85° corresponds to a monoclinic cell with dimensions: a =8.427 (3) A, b=20.695 (3) A, c=7.818 (2) A, V=1210.8 (6) A3, 01:90.01 (2)°, l3=117.37 (2)°, 7:89.99 (2). For 2:2 and F.W.=465.53, the calculated density is 1.277 g/cm3. The space group was determined to P2,/m. Of the 2395 reflections collected, 1176 were unique (Rim = 0.1064). The intensities of the three representative reflections which were measured after every 50 reflections. A linear correction factor was applied to the data to account for this phenomena. The structure was solved by direct methodsm'103 The sodium, oxygen, and carbon of the methanol were refined anisotropically. The C-O, C-C-O, C-O-C, C-H, H-H, C-C-H and C-C-H were constrained to have the same bond lengths for each respective 45 group. The final cycle of the full-matrix least squares refinement was based on 1166 observed (I>2.000'(I)) reflections and 282 parameters and converged (the largest parameter shift was 0.01 times its esd) with unweighted and weighted agreement factors of: R=211F.1-IF.11/21F.1=0.0656 Rw = [(2 w(| F01 - 1 Fc 1)2/2w F02]"2 = 0.1724 The largest residuals in the electron density was +0296 and -0. 147 e.A'3. C. Theoretical Calculations Ab initio calculations were performed by Gaussian 92/DFT (Gaussian, Inc)'°° on several Indigo workstations (Silicon Graphics, Inc.). The standard basis sets employed were as follows: d95v+* for Li, C, N, O, and F 109; d95+* for S”; and the 6—31+G*”° basis set for Na and Mg. The Watchers basis set with polarization functions was used for K and Ca' ' "' '3 and a Huzinaga basis set for Se.114 The four term Huzinaga basis set with a polarization function was employed for H.1 '5 Second order Moller-Plesset perturbation theory (MP2) with only the valence orbitals active and standard convergence criteria were used to optimize the geometry of the species under study. Vibrational frequencies and the infrared intensities of the normal modes were computed analytically except in a few cases where numerical computation was dictated by memory constraints. Atomic charges were determined by the Natural Bond Orbital (NBO) analysis algorithm.l '3 The electrostatic potentials for selected systems were generated by Gaussian 92 and stored in a cube file. Visualization of the electrostatic potential and the molecular orbitals were accomplished by Scianl ‘7 in order to create isosurfaces and contour plots. Isosurfaces of the molecular orbitals in selected species were visualized by importing the Gaussian output file into Scian. The potential energy distribution (P.E.D.) of the stretching modes was computed 46 for each normal coordinate, Q, by using the usual relation, equation 4.1 ‘2 The off diagonal elements were ignored since the Fii terms are larger than the Fij terms (i¢j). The Lik matrix roar-1,3,, P'E'D'diag = ——u-T% (4) Zr 111117: elements were determined by converting the normal coordinates from the Gaussian output, which are in cartesian coordinates, to internal coordinates. D. Data Treatment The number of vibrational components in a spectral envelope was determined by Fourier self deconvolution, second derivative analysis, and curve fitting. Fourier self deconvolution and second derivative analysis were performed via the PCIR program (N icolet, Inc.). In the deconvolution procedure, the width of the resolved peak was adjusted to 10-15 cm'l and the deconvolution factor was varied, 2-4. The deconvolution procedure was applied to small spectral regions (100 cm'l) having low noise (<0.0005 A), in order to minimize the noise in the deconvolved spectrum. Spectral envelopes were fitted to Gaussian/Lorenztian bandshapes in order to detennine the number of vibrational components and yield quantitative information about each component band. Sigma Plot (v. 4.14, Jandel Scientific, Inc.) was used to curve fit the infrared band envelopes. Spectral regions of interest, generally 100 cm", were converted to ASCH files and imported as text files into Sigma Plot. The Gaussian/Lorenztian bandshapes had the following form: A = Zeal—1.10M + l,L(v)] i (5) where 47 I L = (V) 1+(v_vi72 (6) W1 e-ln2(v-v,- )2/w,-2 and GM = (7) The calculated absorbance, A, is a sum of Gaussian, G(v), and Lorenztian, L(v) , lineshapes, where the contribution of each lineshape can be varied by the parameter lg. The independent variable v is the wavenumber(cm'1) of the spectrum. The parameters V), w, and ai are the peak position, width of the band, and intensity at the peak, respectively for a component band. The parameters are adjusted through nonlinear regression in order to minimize the difference between the calculated absorbance and the experimental absorbance. In addition to the Gaussian/Lorenztian sums, a linear offset is added to the regression model to compensate for subtraction errors between the sample and reference spectra. The initial guess used for the fitted parameters is derived from the Fourier self deconvolution and the second derivative analysis. These procedures yield the number of vibrational components, their position, and relative intensities. The uncertainties in the parameters and the residuals of multiple fits to a given experimental band were compared in order to determine which fits were statistically significant. Chapter III Theoretical Calculations A. Introduction The OCN', SCN' and SeCN' anions have been employed extensively as probes in the study of ionic associations and other interactions in solutions.75‘8"82'8°"19"20 The stretching modes of these linear anions are very sensitive to the local electrostatic environment.82 Ionic association markedly perturbs the electronic structure and bonding, and the vibrational modes are shifted (in either direction) from the frequencies of the ‘free’ anion, where the values of the ‘free’ anion are those of tetrabutylammonium cyanate, thiocyanate, or selenocyanide in a high dielectric aprotic solvent. Examples from the experimental literature are collected in Table 15.75'77'8°82‘84"18" '9 The identification of a number of solution species, such as ion pairs, dimers, tetramers, and triple ions has been accomplished by analyzing the vibrational perturbations. These investigations have been crucial in understanding the factors involved in ion pair, dimer and tetramer formation.”74 Ion pairing by alkali or alkaline earth metal cations with OCN', SCN‘ and SeCN' anions generally raises the frequency of both stretching modes (VCN and vex) with respect to the free anion.”82 The largest of these frequency shifts for alkali metal cyanate and thiocyanate ion pairs are observed for the lithium complexes. Likewise in alkaline earth metal thiocyanate ion pairs the greatest shift in the vibrational modes is found for the beryllium ion pairs. Unusually, the VCN stretching vibration (v3) of the Ba2*NCS' ion pair is red shifted from that of free SCN'.82 Dimerization of alkali metal thiocyanate and selenocyanide ion pairs, as well as tetramerization of lithium thiocyanate, reduce the frequencies of the vCN stretches with respect to those of the free anions, with the largest effect seen in lithium complexes?“ '8 The frequency of the ch (vl) stretching mode is higher than the ch stretch in the corresponding ion painm‘"82 It should be noted that the labels vCN for v3 and ch for V] have historically been employed as approximate descriptors of these stretching vibrations; the normal modes involve motions of all three atoms. 49 50 Table 15. Frequencies of the stretchingsrgiodes of OCN‘, SCN' and SeCN‘ species in solution 1,82,84,119.120 Species VCN (cm") ch" (cm") OCN’b 2138 1283, 1196h LIINCO'b 2172 1304, 1218h NaiNCO'b 2159 1298, 1212h ICNCO’b 2 148 1 290, 1204h LIINCO'Li+b 2203 i SCN" 2058 735 BCZINCSf 2 l 1 2 i MgeNcsr 2083 789 CaZINCS'c 2063 780 BaZINCS’c 2047 774 LiJ'NCS‘ 2072 765 Na’NCS‘ 2065 754 KINCST 2058 742 Ag+SCN4i 2080 721 (LIINCS')2c 2034, 2051(R)j 788, 813(R)l (N a”NCS')2° 2042, 2052(R)’ 769, 772(R)‘ (Li"NCS'),,c 1993 , 2022(R)l ' Li‘SCN'Li”r 2087 ' Na‘SCN'Na+r 2081 ' SeCN‘° 2066 54 l LAINCSC'c 2080 ' (Li +NCSe')2‘ 2040 ' ’X can be 0, S, or Se; °in DMSOm’ ; cin DMF”; °in dimethylthioforrnamideso"2°; cin THF“; rin CH3N028‘; l‘in diethylcarbonate'”; hthe VCO is in Fermi resonance with the overtone of the bending mode (280cm), resulting in two bands”; 'obscured by solvent; ’Raman bands denoted by (R) 51 Ion pairing interactions by a given cation produce substantially different perturbations in the frequency of the two stretching modes in these anions. In Li+NCO’, Li+NCS', and Li+NCSe' ion pairs, the frequencies of the vCN modes are higher than those in the corresponding free anions by 34 cm", 14 cm’1 and 14 cm", respectively. The frequency of the vCO stretch is increased by 21 cm’1 in LiNCO from OCN' and the frequency of the vCS stretch is increased by 30 cm“l in Li“ NCS' from SCN'. Moreover, the VCN stretch of Li*NCO‘Li" is blue shifted 65 cm’1 from that in OCN' whereas the vCN stretch of Li”NCS'Li+ lies 29 cm’1 higher than that in SCN'. It appears from these experimental studies that the VCN mode in OCN‘ is more sensitive to electrostatic interactions than the vCN mode in either SCN' or SeCN’. Additionally, the vCS mode in SCN' is more sensitive to an electrostatic interaction than is the vCO mode in OCN'. The basis for these frequency shifts in alkali and alkaline earth metal cyanate, thiocyanate and selenocyanide ion pairs and dimers are of interest since these characteristic shifts have been used to identify numerous inorganic complexes of these anions. A number of authors have discussed the basis for these frequency shifts.”75 318°” N akamoto proposed that 0’ donation from the highest occupied 0' orbital in the SCN‘ anion to the cation in an alkali thiocyanate ion pair raises the frequency of the vCN mode, since electrons are removed or depleted from the 0' orbital, which is a weakly antibonding orbital.88 Chabanel and coworkers suggested that the frequency shifts in ion pairs and dimers could be rationalized through a valence bond approach.”82 MNDO calculations of lithium thiocyanate ion pairs by the Chabanel group support their rationalization for the frequency shifts in alkali thiocyanates.89 These calculations show that ion pairing to the nitrogen end of the thiocyanate anion by lithium cations shifts the negative charge distribution of the anion to the nitrogen end. The molecular orbitals belonging to the thiocyanate anion in a Li"NCS' ion pair are lowered in energy by 5-6 eV, with the 40 orbital (valence orbital in SCN') showing the largest drop.” In dimers and tetramers, the energy of the 40 orbitals in the SCN' anion is slightly more negative than in 52 the ion pair, with an even greater charge transfer from the sulfur to the nitrogen end of the anion. Calculations for a Li”NCS' ion pair show that the ON bond is longer and that the GS bond is shorter than those in SCN'. Dimerization and tetramerization of the Li*NCS' ion pairs further shift the negative charge toward the nitrogen end of the anion, which lengthens the ON bond and reduces the GS bond, with only a small concomitant reduction in the energy for the molecular orbitals. The MNDO calculations of the vibrational frequencies for lithium thiocyanates show a balance between charge transfer and the stabilization of the molecular orbitals in ion pairs, dimers and tetramers and qualitatively reflect the experimental trends for the vCN mode. The frequency of the VCN stretch is predicted to decrease in the following order: LYNCS' > (Li"NCS')2 > (Li‘NCS'),. On the other hand, the calculated frequencies for the vCS stretch in these lithium thiocyanate species do not correlate well with experimental results. One peculiar result is that the empirical model predicts that ion pairing should reduce the ON bond length, yet the calculations show that the C-N bond in lithium thiocyanate ion pairs is lengthened. However, the lengthening of the C-N bond in Li‘NCS’ ion pairs does not reduce the frequency of the vCN stretch. Although the model proposed by Chabanel and coworkers to rationalize the frequency shifts in alkali thiocyanates is based on extensive experimental and theoretical studies, inconsistencies with the experimental observations remain. The practice of employing molecular orbitals to explain bond lengths in diatomics (ions and neutral molecules) has shown that the bond order (charge density) can be related '2' While this approach is empirical in nature, it to the bond length of a diatomic species. provides an intuitive understanding of how the degree of bonding relates to the bond length between two nuclei. The electrostatic theorem proposed (independently) by Feynman and Hellmann states that the buildup of charge density between two nuclei produces attractive forces between these two nuclei whereas electron density outside the internuclear axis produces forces that pull the two nuclei apart.122 Berlin equated the attractive forces to bonding regions and repulsive forces to antibonding regions.123 Bader has applied this 53 approach to explain bonding in both diatomic and polyatomic molecules and has shown a '24 Howard relationship between the electron density and bond lengths between two nuclei. has correlated the C-N bond lengths in OCN' and SCN' anions with the ON bond strengths of these anionsm He rationalized this correlation by showing that electron density around the ON bond is greater in SCN’ than in OCN'.‘25 An obvious criticism of using ab initio or semi-empirical computations to model ion pairing in solution systems is that these calculations lack the dielectric medium and solvation effects that are present in solutions. The vibrational modes of K*NCO' and K”NCS‘ ion pairs in the vapor and solution phases have been studied by Devore.12° He assigned a band at 2206 cm‘1 in the infrared spectrum to the vCN stretching mode of K”NCO' in the gas phase; this vibration is substantially blue shifted from its value of 2148 cm'1 in DMSO solutions and its value of 2124 cm" in the gas phase.”126 A band at 2185 cm‘1 that grows with higher K*NCO' densities is assigned to the infrared active vCN stretching mode in (K’NCO’)2.78 Similarly, Devore assigned a band at 2070 cm‘1 to the vCN stretch of the K"NCS' ion pair, and a band observed at 2055 cm" at higher K*NCS' concentrations to the infrared active vCN mode of (K"NCS')2.78 The value of 2070 cm’1 for the VCN stretch of KiNCS' in the gas phase is close to the 2058 cm’1 band attributed to the KNCS ion pair in DMSO solutions.82 These experimental studies suggest that the OCN' anion is much more sensitive to medium effects such as ion solvation and the dielectric environment than the SCN‘ anion. It is also clear that great care must be used when comparing theoretical predictions to experimental solution values. The OCN' and SCN' anions have also been employed in investigations of anion solvation, particularly in hydrogen bonding solvents, since the perturbation of the vibrations upon ionic association are well known, although the basis for the frequency shifts are not entirely understood.3°'°3"27"28 Several authors have assigned a blue shift in the VCN mode of SCN’ in protic solvents to a hydrogen bond at the nitrogen end of the thiocyanate anion.”43 In a recent study of the solvation of the thiocyanate anion in 54 methanol (MeOH), Gill and coworkers determined that the vCN absorption envelope of SCN' consisted of six component bands.43 Based on earlier observations on acetylenic and ethylenic stretching modes where hydrogen bonding to the it cloud of these bonds lowers the frequencies of the carbon-carbon stretching modes“, they assigned red shifted bands with respect to ‘free’ SCN' anion to complexes with MeOH hydrogen bonded to the it cloud of the C-N bond. Blue shifted bands were assigned to a single MeOH molecule hydrogen bonded to the nitrogen end and to one and two MeOH molecules hydrogen bonded to the 11: cloud of the ON bond in conjunction with a hydrogen bond to the nitrogen end of the thiocyanate anion.4° Much vibrational spectroscopic work has been devoted to the study and identification of ionic and hydrogen bonding complexes with OCN', SCN', and SeCN' anions. Unfortunately, the number of theoretical calculations to support these vibrational assignments and to understand the nature of the observed frequency shifts is very limited. With the exception of the Chabanel group’s calculation of lithium thiocyanates”, described above, theoretical studies of these anions have primarily focused on the structure and energetics of ionic and hydrogen bonded complexes, without much discussion of how these interactions perturb the anion.'29"°5 The purpose of our work is to explore the correlation between structure and vibrational frequencies. Towards this goal we will calculate the vibrational frequencies for alkali and alkaline earth metal XCN' complexes (X=O, S, or Se) that have been identified in solutions, ascertain that the level of theory is adequate to describe such electrostatic interactions, and apply the theory to hydrogen bonded complexes in order to assist in the assignments of these species in experimental solution studies. This investigation also provides insight into the basis for the frequency shifts observed for ionic and hydrogen bonded complexes by predicting the structural implications, changes in the force constants, shifts in the distribution of the negative charge, and perturbations of the molecular orbitals accompanying ionic and hydrogen bonding interactions. 55 B. Results 1. OCN, SCN', and SeCN' Ab initio calculations of the OCN‘, SCN', and SeCN' anions provide a basis for understanding how these anions interact with and are perturbed by a positive electrostatic interaction. The equilibrium bond lengths, vibrational frequencies, and atomic charges for these anions are summarized in Table 16. The C-N bond lengths decrease from OCN' to SeCN', which reflects the greater triple bond character of SeCN‘. The structural parameters for these anions are comparable to a number of previous theoretical calculations.”203733.3032 Table 16. Calculated equilibrium bond lengths, stretching frequencies, and atomic charges for OCN', SCN', and SeCN' Anions d(CN) d(CX)‘ VCN cha Charge Charge Charge A A cm" cm" @ X“ @ C @ N OCN' 1.21 8 1.244 2155 1205 -0.887 0.796 -O.909 SCN' 1 .208 1.671 1979 745 -0.514 0. 179 -0.665 SeCN' 1.205 1.770 2003 617 -0.350 0.009 -0.659 awhere X can be 0, S, or Se Interestingly, the frequency of the vCN stretching mode does not correlate with the length of the ON bond in these anions. The reason for the deviation from Badger’s rule136 is that the normal modes for the vCN stretch in these anions are not localized to the C-N bond, especially for OCN'. The vCN vibration in OCN' is similar to the asymmetric stretch in N3' or CO2 in that the normal mode involves significant displacements of all atoms, 56 whereas the VCN stretches of SCN' and SeCN' are more localized to the C-N bond. The dependence of the vCN and VCX modes of the OCN', SCN', and SeCN' anions on their respective bonds can be characterized quantitatively by the RED, which relates the contribution of each bond stretch to the potential energy of the normal mode.l '2 The results, summarized in Table 17, show that the stretching modes of the anions involve both the ON and C-X bonds. As the mass of the chalcogen increases, these modes more closely resemble a pure C-N or C-X stretch. Yet, this analysis shows that even in the extreme example of SeCN', the vCN and vcSe modes should not be thought of as a pure C-N or C-Se vibration. Table 17. Potential energy distribution and force constants for the stretching modes of OCN', SCN', and SeCN' Anions V”, v” Fa, (N/m) a“ (N/m) OCN' 1350 1113 %CN 62 41 %CO 38 59 SCN' 1400 502 %CN 88 8 %CS 12 92 SeCN' 1432 471 %CN 90 7 %CSe 10 93 “where X can be 0, S, or Se The predicted force constants for OCN', SCN', and SeCN' provide insight into the nature of the ON and C-X bonds. The C-N force constant, FCN, varies in the following order: OCN' < SCN' < SeCN’, showing that the SeCN' anion has the greatest triple 57 bond character of this series. The force constants for the C-X bond change in the following order: OCN' > SCN' > SeCN'. These trends are consistent with the resonance structures that can be drawn for these anions, such that greater triple bond character in the ON bond comes at the expense of the C-X bond order. The predicted frequencies of the VCN modes in OCN‘, SCN', and SeCN‘ correlate well with observed values for these anions in aprotic solutions. As in the experimental studies, the predicted vCN frequency decreases in the order OCN' > SeCN' > SCN’. The calculated values for vCN, shown in Table 16, are within 3% of the experimental solution frequencies (T able 15). The predicted frequency of the vCO mode in OCN' and the vCS stretch in SCN' also compare well with experimental solution values for these anions. No experimental comparisons can be made for the vcSe stretch due to the lack of experimental results. The atomic charges, as calculated by the NBO procedure and listed in Table 16, show that the negative charge is delocalized in these anions. In each case, the nitrogen atom has the largest concentration of negative charge. An indirect measure of the charge density around the nitrogen can be obtained experimentally by 14N NMR spectroscopy. The 1”N chemical shifts of OCN', SCN' and SeCN' are -288 ppm, —165 ppm, and -137 ppm, respectively.83 The large difference in the nitrogen atomic charge calculated between OCN‘ and SCN’ (Table 16) is supported by the dramatic change in the 14N chemical shift from OCN' to SCN'. The small predicted difference in the atomic charge on the nitrogen atom between SCN' and SeCN' is also concordant with experiment. The differences in the l“N chemical shifts of OCN', SCN', and SeCN' have been rationalized in terms of the change in the it cloud surrounding the nitrogen nuclei in these molecules.83 It is important to note that these calculated charges do not agree with the implications of the resonance model (Figure 15). Although the force constants and the bond lengths for these anions imply that the C-N bonds in the SCN' and SeCN' anions have a greater triple bond 58 character than the C-N bond in OCN', the calculated atomic charges on the chalcogens in SCN' and SeCN’ are less than those for the nitrogen atoms. The valence molecular orbitals for these anions, reproduced in Figures 13-15, show striking differences in shape across the series. The electronic configurations of the valence molecular orbitals are 102 20’2 30‘2 111:4 40‘2 21t4 for OCN', with the 11t° and 40’2 reversed for SCN' and SeCN'. The 10' and 20 orbitals are primarily bonding in OCN', SCN', and SeCN'. The 30' and 40 orbitals are primarily antibonding. There is a remarkable difference between the 11: orbitals in OCN’ and in SCN' and SeCN'. The maximum of the 11: orbitals shifts from the chalcogen end in OCN' toward the nitrogen end in SeCN' with an intermediate structure in SCN'. The 2n M08 in OCN' are slightly bonding for the ON bond and antibonding for the CO bond. On the other hand, the 211: M08 in SCN' and SeCN' are slightly antibonding for the ON bond and slightly bonding for the CS and C-Se bonds. 2. Electrostatic eflects on OCN', SCN' and SeCN The two resonance structures that can be drawn for OCN', SCN‘ and SeCN', diagrammed for SCN', suggest a number of implications that can be tested by ab initio calculations. This model predicts that electrostatic interactions to the nitrogen end of these anions should shift the negative charge density toward the nitrogen atom, with a consequent reduction of the C-N bond order and reduction of the C-X bond length, where X can be 0, S, or Se. The perturbations in the structure of these anions by electrostatic interactions should be reflected in changes in the force constants and bond lengths. ‘ Likewise, an electrostatic interaction to the chalcogen end should produce an opposite shift in the electron density, resulting in an increase in the strength of the ON bond and a reduction in the strength of the C-X bond. The shift in the charge density of these anions as the result of an electrostatic interaction can be followed by the NBC analysis, which 59 Figure 13. Contour plots for the valence molecular orbitals in OCN' Green to Blue (-.1 to -1.7), Green to Red (.1 to 1.7); stepsize 0.1 11:my 27‘x.y Figure 14. Contour plots of the valence molecular orbitals for SCN Green to Blue (-0.1 to -1.7), Green to red (0.1 to 1.7); stepsize 0.1 16 111:,“y 61 Q) 20 Figure 15. Contour plots for the valence molecular orbitals in SeCN' Green to Blue (-0.1 to -1.7), Green to Red (0.1 to 1.7); stepsize 0.1 @ 62 calculates the atomic charges and determines the occupancy of each atomic orbital in the complex. The effect of a purely electrostatic interaction with these species can be calculated by placing a bare proton, having no basis functions, at a fixed distance from the anion. The validity of the valence bond model can be tested by determining the effect of such an electrostatic interaction on the bond lengths. Here we limited the model to electrostatic interactions along the molecular axis, at a fixed distance of 3 A from either end of the anion, as shown in Table 18. An interaction to the nitrogen end of SCN' and SeCN' slightly increases the ON bond lengths and decreases the GS and C-Se bond lengths. The same type of interaction with OCN’ slightly reduces the ON bond length and decreases the length of the CO bond. Electrostatic interactions to the chalcogen end of these anions all Table 18. Effect of a positive charge on the bond lengths, force constants and vibrational frequencies for OCN', SCN‘, and SeCN‘, calculated upon fixing a bare proton 3 A along the internuclear axis from either end. Species d d a Vc~ vex“ FCN F ex“ XI 2 (cm") ( cm") (N/m) (N/m) OCN‘ 1.218 1.244 2155 1205 1350 1113 OCN""H+ 1.217 1.222 2244 1263 1414 1260 NCO""H" 1.208 1.258 2185 l 185 1423 1061 SCN’ 1 .208 l .671 1979 745 1400 502 SCN""H+ 1.21 1 1.636 1997 803 1429 586 NCS""l-I+ 1.201 1 .668 2032 756 1451 525 SeCN' 1.205 1.770 2003 617 1432 471 SeCN""H+ 1 .208 l .740 1999 652 1438 525 NCSe"“H* 1.198 1.754 2064 652 1489 534 'where X can be 0, S, or Se 63 produce a reduction in the length of the ON bond. On the other hand, the C-0 bond length in NCO""H+ increases whereas the C-S and C-Se bond lengths decrease in NCS""H+ and NCSe’"'H+ complexes. The results of these model calculations are in only modest accord with the predictions of the resonance structures and show that these anions respond in a more complicated manner to electrostatic interactions. The simple resonance model may have been successful in explaining structural and vibrational perturbations from metal-chalcogen interactions with SCN' and SeCN' because the metal cation does not approach along the molecular axis of the anion. For example, the crystal structure of Ag*SCN' shows that the AgSC angle is 104°.83 This nonlinear interaction lengthens the GS bond length and slightly shortens the C-N bond length. The perturbations to the stretching modes of OCN', SCN', and SeCN' by the same electrostatic interaction reveal the complicated nature of these vibrations. As discussed earlier, the vCN mode in these anions shifts by +89 cm", +18 cm", and -4 cm’1 for electrostatic interactions to the nitrogen end of OCN‘, SCN', and SeCN', respectively. An apparent dilemma is the predicted negative frequency shift of the vCN stretch for the SeCN""H+ complex relative to that in SeCN' despite an increase in both FCN and FCSe force constants. This negative frequency shift is the result of a change in the normal mode such that the ‘reduced mass’ of the mode is lowered. On the basis of the potential energy distributions, the greater sensitivity of the vCN mode in OCN' versus the SCN' and SeCN' anions to an electrostatic interaction at the nitrogen end is a result of the greater dependence of this mode on the CO bond. In the SCN' and SeCN' anions, the vCN mode is more dependent on the ON bond, which is only slightly increased in length by these interactions. The opposite trend is calculated for this mode when the interaction is at the chalcogen end of these anions. Here, the vCN mode shifts by +30 cm", +53 cm", and +61 cm". These trends are similar to those observed experimentally for lithium ion pairs of these anions. It is interesting to note that there are several examples where the frequency for the Vcn mode is predicted to shift to higher energies despite an increase in the calculated C-N bond lengths. The calculated ch frequencies more closely follow the calculated C-X bond length changes. Again, the greater sensitivity of the vCS and vCSc modes with respect to VCO to an electrostatic interaction at the nitrogen end is due to the larger dependence of these modes on the C-S and C-Se bonds, respectively. 3. Ionic and hydrogen bonding Ab initio calculations for cation or hydrogen bonded complexes with OCN’, SCN', and SeCN' can be grouped into four categories: MNCX, MXCN, triple ions and dimers, where M is an alkali or alkaline earth metal cation and X can be O, S, or Se. The equilibrium bond lengths, vibrational frequencies, and atomic charges for OCN', SCN' and SeCN‘ in ionic and hydrogen bonded complexes are summarized in Tables 19-21. Both ionic and hydrogen bonding interactions dramatically perturb these anions. The largest effects are predicted for the alkaline earth metal ion pairs, reflecting the large polarization in the anions. On the other hand, the hydrogen bonded complexes are calculated to have smaller changes in the bond lengths, vibrational frequencies, and atomic charges, owing to the weaker electrostatic nature of the hydrogen bond. Interestingly, similar electrostatic interactions with OCN‘, SCN' and SeCN‘ result in comparable structural changes. The strong electrostatic interaction by alkali metal and alkaline earth metal cations in MNCO, MNCS, and MNCSe ion pairs lead to predictions that can be rationalized by the resonance model. The C-N bond lengths are increased and the C-X bond lengths are decreased, with the notable exceptions of the K*NCO‘ and the K”NCS' ion pairs, where the C-N bond lengths are slightly decreased. Comparison of the bond lengths in the BehNCO', Be2*NCS', and Be21NCSe‘ ion pairs to their respective NCX’ anions shows ‘ that the ON bonds are lengthened by 0.022 A, 0.040 A, and 0.042A, respectively, whereas the C-X bonds are shortened by 0.077 A, 0.125 A, and 0.124 A, respectively. The same trend can be seen in the Li*NCO', Li”NCS', and Li“NCSe' ion pairs, where the 65 Hus—e we. GEE—58n— coaegmam. mqonoaam menace-=98. Waggon 583:3? 38o Sagas Ba ”2835 058mg 404 002. 5 mean 85 358mm: 338 8.258 83338 when-ma. A9 A8 <9. mm: <8 m8 kn: woo 9.9% gm figs 3 c» 2:; 535326. 2:; 5458? >3: >5: ©2 © 0. @Q 002- _.N_m Fun: Nam owq 50m um Sue :5 .908 chem 19me Wmu+200- ~.NAO H-Haq whom 5: ~23 #5 55m S—O LUG— —.~©~ touc— zm~+200. —.Nwo - — mN Nwoa Sou :uo um 7:00 awe L130 —. _ uN toduo 0m~+200. 5qu ~. Hem wwww Hqu 3.; Ho 73a Tam L .wmh foqw -9on 5+200. _ .NNO —.NO.~ Nwmm sow Sue m HANG 5.: .. H .NNN oboe touxm 23200. —.NNO —.N H m NNMO ~93 3A3 w 736 5 ~A .- — . fig 0b.: -94?“ “+200- 0&3 ~.N~o N~©u Sow ~Nqu 0 30m ~N83 Lwa ohm—O tau-mg mum—.200. Hboo —.Nuw NNma mma 32 Na gum _ Eu tobmm Chou 10.x“; =0:...200- _.N :4 ubwq NNOQ mi ~NAo wo T: — fl ~¢N 10.0% 9me to.ma~ 3003-200- TN ~h _.Nwm NNoa muw ~93 mo («8 — Go tobqw O.m©~ rob-No Pm+002- - ~ om fwow NNNm 0N0 sma Nam qu ~O~© tobac Chum L bub 2w+002. u .NON FNOG NNOQ ham 5cm Go Tgw Hows“ rename 9qu t ~ .omo az002. ~.NO© _.Nua N:— m¢© 29* ~@_ :30 25¢ 10.x; 9.3— .995 20:...002. .979 _.~u~ 23 a; :ou «N 53 Ewe rpmma Ouwmm tag-mow 30031002. _.N T¢ ENMN N _ mm 2: _ ~©u um 5.3 23m to.mam O.mom tob— _ ~1m+002-—L+ —. — cu _.Nww wwwu cuo waq .3 Gem ~Nwm raw-cow oqu tobau m:..002-zm m. —.No~ —.NAu NNaN cow 5% mm 3N4 — 73 to.mmm chew toaqm :0:...002-...:0= ~98 fuew NM _ .u w: SAN um Tgu fl 5m .90 ~ q 9me roach A5+2000~ FNNQ _.Noo qum N3 u 53 E. Hume :3 t _ bag :0: toQ — N wwoo 0 Gus o 66 Haw—a we. 823:3 5,.- ..200- rfiu :3 8.3 m3 5% 3 53 G _ u L .wwo 908 563 rm 1: . ..20O- Eu 5 rm 0 m 35 She Emu u :3 Eco - _ .8. 902 -943 mm- :03. ..Z00- rm 3 rmuo mmom mum 53 we $8 52 .98.» came 5&3 mom. 54.- ..002- :3 _.wow N _ 3 GM Ben E :33 mg 6.30 9qu - _ .o-E 67 Haw—a N0. 0&8580 dozen—firm. £88306 35:398. Sam—£8 020358. 380 88:58 25 39am 033mg «2 m02- 5 mean 25 308mg @308 "380.258 85293 M‘Nfimfih mg “a <02 N02 <0M m0“ *0? k0“ Q50 05“ GS“ x» L 2:; 03532». as; 03532... >3: >33 ©2 © 0 © 0. m02. _ .00m .03 030 uwu 03 N0 300 won -903 0. 0 ac -9u 3 wow-20m. _ .NAm _ .03 ~03 . mum :00 ~00 _ 00— 30 - 0 hmw 0.0g 98— Zm~+20m. 0N3 ruoq N000 :3 02 00 $00 30 -_ bum 0.000 90.0 09220” 0.000 fume pong 5.5 000 0 _ 0.00 .30 - 0 . 0 0.0 0. 30 0.00m E+Z0m. PM; 0.0; N03 :8 00$ 0 :5 000 L000 90?. 00% 2w+20w. 0.8 u 0.08 N000 03 wow 0 T: u 05 .9000 903 -903 520m. 0000 0.3m 500 m 0 0 000 u Ema mow -908 0.03 -9000 30-200. 0.~0~ .03 ~03 up: mom A 33 Km .930 9&3 -93» EOE-20m- 0 .000 0000 00. 0 A00 0.: _ _ 300 000 -9000 0.03 -950 5.002. 0.» 0 0 0.000 00....— ~00 0.2 mm 3mm moo -903 0.0m.» -905. 30-002. rm?» 03m N000 ~00 «mm .0 :00 $0 -903 9 En -930 ICE-002. 0.000 #03 $3 nmo .r: E Ewa .0on -900» 9 50 -9u _ 0 500200.. :8 0.0: ~03 m _ q mum _ Gum E -980 0.0 0 0 -92 0 “2.00255“ 0. Gm #000 00% #00 000 um 33 3A -908 09% -930 EOE-0023.20: 0.000 roam ~03 0mm 30 0 75m H: -9000 930 -930 E... inm- Ecv .03 00mm 0000 03 mmm u 300 .50 - _. am 9.03 -903 m: . ..z0m- 3+. 0.3 _ .600 N00— m3 mg 0m 33 umm -9000 903 -9000 :01. .28. 53 _.SN 3% E E a 3; E -3; 88 -930 :0: 68 0.9....” 0.. 9:858“. 00:05:03? £883.50 03:30.8. 580380 588.08. 38o 838:8 Ea 203.0 3808 000 0002. .: .200 BE 3.0302. 0308 8.380330 880.88 0.08.8 «.9 «.9 <9. mnz <8a m8. .82 .8. 0.0.9.0... QRSR QB»? x. 0. as; .3330? as; .3530? 20: >3: ©.< ® 0. 000... 0002. . .000 . .000 0000 000 0 . 0 0. . .000 A0. 0.000 0.000 0.000 0034600. ..000 ..000 0000 .0.# 000 A00 .00A 000 .. . .000 0.000 0.000 2.08.2000. ..00. ..00. 0000 .000 000 00.0 .000 000 -..0.0 0.0.0 0..00 Om~+200o. ..00A . .000 .000 .000 000 00.0 .000 000 - . .000 0.000 0. . 0A ....+an0. ..0.. ..0 .0 0000 ..00 000 00k .t0 000 -.000 0. .00 0. . 00 20+2000. . .000 ..00. 00.0 0. 0 000 0.0 .000 00. 0.000 0. .00 0. . 00 .A+ZOmo. . .0 .0 ..000 0000 000 000 0. . .00. 000 0.000 0. .00 0.0.0 0..-2000- .. .00 . .000 0000 A00 000 .00. .0 .0 #00 0.000 0. .00 0.000 10......2000. ..000 ..00. 00.... 000 00. 0.0 .A00 “«00 0.000 0.000 0.00. F3082. . .000 . .000 .000 . 00 000 #0. 0 .000 B«00 0.000 0.00 . 0.00 . 0..-.0002- ..00. . .000 0000 . 00 000 0.0 7900 B00 0.000 0.00 . 0.0.0 20......0002. ..000 . .000 0000 .00 0 . 0 0.. :00 h000 0.0.0 0.000 0.000 500.025.. .. . 00 . .000 0.00 000 000 00.0 .000 #00 0.0 .0 0.000 0.000 .uI...moOZ-.......u .. .0A ..000 0000 00.. 000 0.0 .050 A00 0.000 0. .00 0.000 ICE-0002-.....O: ..00. . .000 0000 000 0&0 P0 .000 R00 0.0.0 0.00. 0.000 5+ ”.280. E“... ..0A0 ..00. .000 0A0 0A0 00.0 .0.. 000 -...A0 0.000 0. .00 m: c ..ZOMo- 3.... . .000 . .000 00.0 000 00A 0.0 :00 000 0.00A 0. .00 0.000 :02. ..z8o- 30*... . .000 . .000 0000 000 000 0.0 it 00. 0.000 0.000 0.000 69 C-N bonds are lengthened by 0.002 A, 0.008 A and 0.006 A, with the ox bonds decreased by 0.037 A, 0.057 A 0.051 A, respectively. The opposite trend in the bond lengths is seen in Li*OCN' ion pairs. The C-N bond length is reduced by 0.020 A and the C-0 bond length is lengthened by 0.025 A with respect to OCN'. Unfortunately, the same comparisons cannot be made for the MSCN and MSeCN ion pairs, because the lowest energy structures of these species are not linear, so that the simple resonance model would not apply. In general, the calculated bond lengths for the linear MOCN, MNCO, MNCS, and MNCSe ion pairs satisfy the predictions of the resonance model. A given cation perturbs bond lengths of the anions in metal thiocyanate and selenocyanide ion pairs more than those of the corresponding metal cyanates. This effect, illustrated by the Be2+NCX’ and Li"NCX' ion pairs, can be rationalized in terms of the simple valence bond model. Since the ON bonds in the SCN‘ and SeCN‘ anions have greater triple bond character than that of the ON bond in OCN', the SCN' and SeCN' anions are more easily polarized by an electrostatic interaction to the nitrogen end. A single hydrogen bond to the nitrogen end of OCN', SCN' or SeCN', by HF (hydrogen fluoride) or HOH and MeOH in the case of OCN’, shortens both the C-N and C- X bond lengths in these anions. A hydrogen bond to the chalcogen end shortens the ON bond and lengthens the C-X bond. The reduction of the C-N bond lengths in XCN""HD complexes, where HD is a hydrogen bond donor, suggests that the hydrogen bonding interaction cannot be interpreted as a simple electrostatic interaction as in ion pair complexes. The MZNCX* triple ions and (DH)2"‘NCX’ doubly hydrogen bonded complexes lengthen the C-N bond and shorten the C-X bond. The (Li*)2NCO' triple ion is qualitatively similar to the (Li“NCO‘)2 dimer in terms of the calculated bond lengths, stretching frequencies and charge distribution, and thus serves as a good model species for the dimer. Again, the response of these anions to two ionic or hydrogen bonding interactions to the nitrogen end results in changes in the bond lengths that are qualitatively 70 predicted by the resonance model. The (Li*)2OCN‘ triple ion shows the opposite effect from (Li”)2NCO', which also is consistent with the resonance model. Both the ON and C-X bond lengths are calculated to shorten in the Li‘NCX'Li” triple ions and DH"'NCX""HD hydrogen bonded species shorten. The largest changes are seen in the Li+ complexes where the C-N bond is shortened by ~0.1 A and the C-X bond by ~0.02 A. 4. Atomic charges in ionic and hydrogen bonded complexes The shift in the distribution of the negative charge in OCN', SCN' and SeCN' in response to an electrostatic interaction can be described by the atomic charges. The N BO atomic charge analysis shows that the cyanate anion responds uniformly to the strength of the Coulombic potential applied. The atomic charges on the nitrogen, carbon, and oxygen atoms of OCN', when it forms a hydrogen bonded or ionic complex at the nitrogen end of the anion, are plotted as a function of the Coulombic potential generated by the hydrogen bond donor or the metal cation in Figure 16. Here, the Coulombic potential is defined as the atomic charge of the metal cation or the hydrogen bond donor divided by the hydrogen- nitrogen or cation-nitrogen distance. The linear correlations between the atomic charge on the oxygen, the carbon or the nitrogen atoms and the Coulombic potential are 0.989, 0.976, and 0.960, respectively, reflecting the predominantly electrostatic nature of the interaction. The largest change in the atomic charges is calculated for the nitrogen atom, which draws electron density from both the oxygen and carbon atoms. The atomic charges in OCN' also correlate well with the Coulombic potential when the ionic or hydrogen bonding interaction is on the oxygen end. In this case the distribution of the negative charge shifts towards the oxygen end. The change in the atomic charges in response to a Coulombic potential is consistent with the resonance model for these interactions, where an increase in the negative charge at the nitrogen atom in OCN' will come at the expense of the negative charge on the oxygen atom. The atomic charges in MeOH, HOH and HF 71 complexes to the nitrogen end of the anions do not correlate as well with the Coulombic potential generated by the proton of hydrogen bond. Indeed, the correlation between the atomic charges and the Coulombic potential is improved when the hydrogen bonded complexes are removed from the set. The difference is a reflection of the more complex nature of the hydrogen bonding interaction. As discussed earlier, the ion pairing interaction is primarily electrostatic in nature, with a small amount of a charge transfer involved for the alkaline earth metal ion pairs. On the other hand, hydrogen bonding can result not only from electrostatic and charge transfer interactions, but also may have some covalent character. The sums of the atomic charges on the cyanate anion in the MeOH'"NCO', HOH"'NCO' and FH'"NCO' species are 0952, -0.958 and-0.899, respectively, suggesting the importance of charge transfer in these hydrogen bonded species. The same trends are found for the atomic charges on the atoms in metal thiocyanates, metal selenocyanides, hydrogen bonded thiocyanates, and hydrogen bonded selenocyanides, where the interaction is to the nitrogen end of these anions. Correlations similar to those in Figure 16 are also observed for the ionic and hydrogen bonded SCN‘ and SeCN‘ complexes. For these complexes the slopes of plots of the atomic charges on the nitrogen atoms versus the Coulombic potential are larger than that for the OCN' system. The calculated atomic charges in Li*NCX'Li+ triple ions and analogous hydrogen bonded complexes, DH"'NCX""HD, show only small differences from their respective anions (Tables 19-21). Electrostatic interactions to either end of the anion produce opposite results, and the two effects approximately balance. The largest changes are predicted for the lithium triple ions. The (Li+)2NCX' triple ions and analogous hydrogen bonded species, (DH)2"'NCX' display effects similar to those calculated for ion pairs, where the negative charge on the anion is substantially shifted to the nitrogen end. The opposite effect is predicted for (Li+)2OCN' ion pairs, where the negative charge on the cyanate anion is shifted to the oxygen end. Alr- 72 1.500 .. r = 0.989 (r= 0.987) * BeNCO‘ 1.000 - CaNCO’ MgNCO HOH”‘NCO OCN' 0.500 .. 2.“? a 0.000 .. 3 l 9‘0 | g r = 0.976 .2 (r = 0.995) g ‘0.500 dr- 2 ' BeNCO‘ rNCO OCN' HOHNCO KNCONaNOOL ‘: MeOH...NCO" . FH'NCO‘ '1 .000 db MCOHMNCO'. O muNCO- OCN HOH‘"NCO‘ r = 0.960 . Charge @ N (r = 0.998: -l.500 - I Charge @ C BCNCO. A Charge @ 0 -2.000 : : : : 0 5 10 ,5 20 Coulombic potential (J/C) Figure 16. Atomic charges at O, C, and N in MNCO species, where M can be a hydrogen bond donor, an alkali metal cation, or an alkaline earth metal cation, versus the Coulombic potential provided by the cation or hydrogen bond donor; r is the correlation coefficient with and without (in parenthesis) the hydrogen bonded species 73 The resonance model predicts that the polarization of the charge distribution of OCN‘, SCN' and SeCN‘ is primarily in the 11: system. The calculations reveal that the ls and 23 core orbitals are little affected by an electrostatic interaction, whether on the nitrogen or chalcogen end. The pz orbital, which lies along the molecular axis, is slightly perturbed by an ion pairing or a hydrogen bonding interaction. The greatest changes are seen in the pK and py orbitals, as exemplified for the OCN’ species in Table 19. Ion pairing and hydrogen bonding to the nitrogen end of these anions shift the pK and py electrons from the chalcogen atoms to the nitrogen atom. The opposite trend is predicted when the ion pairing or the hydrogen bonding interaction is at the oxygen end. Table 22. Occupancies of 2px, 2p,, and 2pz atomic orbitals in alkali metal, alkaline earth metal, and hydrogen bonded cyanates Species Oxygen Carbon Nitrogen 2p... 2p. 2p. 2p” 2p. 2p, 2p... 2p. 2p, OCN' 1.80, 1.80 1.53 0.79, 0.79 0.81 1.39, 1.39 1.46 Nitrogen species Be2*NCO' 1.60, 1.60 1.55 0.64, 0.64 0.79 1.70, 1.70 Li‘NCO' 1.73, 1.73 1.54 0.72, 0.72 0.80 1.53, 1.53 FH'NCO' 1.78, 1.78 1.53 0.76, 0.76 0.80 1.44, 1.44 1.48 HOH"'NCO’ 1.77, 1.79 1.53 0.77, 0.77 0.80 1.42, 1.42 1.48 MeOI-PNCO' 1.77, 1.79 1.55 0.77, 0.77 0.80 1. 3, 1.42 1.49 Oxygen species Li*OCN' 1.87, 1.87 1.58 0.83, 0.83 0.79 1.27, 1.27 1.46 FH"'OCN‘ 1.76, 1.83 1.58 0.81, 0.80 0.80 1.36, 1.35 1.43 HOH"‘OCN' 1.80, 1.80 1.52 0.80, 0.80 0.80 1.35, 1.37 1.46 MeOH"'OCN‘ 1.71, 1.81 1.63 0.81, 0.80 0.80 1.40, 1.37 1.43 74 5. Molecular orbitals in ionic and hydrogen bonded complexes Evidence that the polarization of these anions in ion pairs is mainly through the 1: system is provided by the visualization of the molecular orbitals, shown in Figures 17-21 for OCN', Li*NCO', Li*OCN', Be2"NCO' , LizNCO", and L120CN“. The 16 and 20' valence molecular orbitals in OCN‘, SCN' and SeCN‘ in MNCX ion pairs are primarily bonding in nature and are not substantially affected by an electrostatic interaction (Figures 22-27). The 36 and 4G valence orbitals, which are primarily antibonding, become slightly bonding for both the C-0 and C-N bonds when perturbed by electrostatic interactions to the nitrogen end. In Figure 28 which compares the 1: orbitals in OCN', LiNCO, LiOCN and BeNCO‘, the extreme positions perpendicular to the intermolecular axis for the i0.26x10‘5 m3” and :1: 2.1x10'5 m‘3’2 contours of the 11: orbital and 21: orbital of OCN’ are referenced by bars for each species shown. The Li+ cation primarily compacts the 11: orbital around the C—0 bond in the Li"NCO‘ ion pair. The same electrostatic interaction pulls electron density away from the ON bond in the weakly bonding 21: orbital. Electron density is slightly pulled into the CO bond in the 21: orbital from the antibonding region. These changes are more pronounced for Be2*NCO'. The 11: orbital for Be2+NCO' is substantially compacted around the CO bond and is reflected by the substantial structural changes for the anion in this ion pair. The shift of the 21: orbital to the Be2+ cation is more exaggerated than for the Li” cation thereby removing additional electron density surrounding the C-N bond and increasing the electron density around the C-0 bond over that of the Li+ interaction. In the Li‘OCN’ ion pair, the opposite effect is seen. The 11: and 21: orbitals are shifted toward the oxygen end, which reduces the electron density surrounding the C-0 bond and increases the electron density around the ON bond. Similar shifts in the electron density of the 11: and 21: molecular orbitals are also calculated for the MNCS and MNCSe ion pairs. The maximum electron density of the 11: orbital shifts toward the nitrogen. These interactions increase the bonding character of the GS and C-Se bonds by pulling electron density from the chalcogens onto the OK bond. Likewise, the antibonding 75 3o 46 11:,"y 21:” Figure 17. Contour plots for the valence molecular orbitals LiNCO Green to Blue (-.1 to -1.6), Green to Red (.1 to 1.6); stepsize 0.1 76 lo 20 30 4o 11:”, 21:” Figure 18. Contour plots for the valence molecular orbitals of LiOCN Green to Blue (-.1 to -1.6), Green to Red (.1 to 1.6); stepsize 0.1 77 @D‘ 11:,Ly 211:,“y Figure 19. Contour plots for the valence molecular orbitals in BeNCO+ Green to Blue (-.1 to -1.6), Green to Red (.1 to 1.6); stepsize 0.1 78 21:x 21:}, ft» ' C0) Figure 20. Contour plots for the valence molecular orbitals in LizNCO” Green to Blue (-.1 to -1.7), Green to Red (.1 to 1.7); stepsize 0.1 79 26 36 4o 3) 21k zuy @ Figure 21. Contour plots for the valence molecular orbitals in LiZOCN“ Green to Blue (-.1 to -1.7), Green to Red (.1 to 1.7); Stepsize 0.1 80 26 ”@j 11:x Figure 22. Contour plots for the molecular orbitals for LiNCS Green to Blue (-0.1 to -1.7), Green to Red (0.1 to 1.7); stepsize 0.1 81 @500 00 @= Figure 23. Contour plots for the molecular orbitals BeNCS+ Green to Blue (-0.1 to -1.7), Green to Red (0.1 to 1.7); stepsize 0.1 82 83 16 26 36 46 @ - @8 11:,“y 27R)! 25. Contour plots for the valence molecular orbitals for LiNCSe Green to Blue (-0.1 to -1.7), Green to Red (0.1 to 1.7); stepsize 0.1 11:,“y 84 26 @.\\\ g 21:Ly Figure 26. Contour plots for valence molecular orbitals for BeNCSe" Green to Blue (-0.1 to -1.7), Green to Red (0.1 to 1.7); stepsize 0.1 85 9%) @V e 27% Ky 27. Contour plots for valence molecular orbitals in LizNCSe+, Green to Blue (-0.1 to -1.7), Green to Red (0.1 to 1.7) stepsize 0.1 86 11: OCN' 21: OCN‘ 2n OCN‘Li+ eeee 11: OCNBe2+ 11: ocmse2+ %@ 11: Li+OCN' 21: 1.1+OCN‘ 0 I e 28. Contour plots of the 11: and 21: Molecular orbitals in OCN‘,OCN‘Li*, OCN‘Bez”, and I.i+OCN‘. (Green to purple: 0261th15 to -2.6x1015 m'm, green to red: 0.26x10‘5 to 2.6x10‘5 m‘m; stepsize 0.26x10‘5 m'm) 87 character of the 21: orbital around the ON bond in these complexes is increased by pulling electron density off the nitrogen atoms. Similar arguments can be made for the MzNCX+ triple ions. The non-axial interaction of the lithium cations has a dramatic effect on the 1: orbitals in the plane of the complex, reducing the bonding character of these orbitals around the C-N bond in these anions. Hydrogen bonding interactions at the nitrogen end show similar trends in the atomic charges and occupancies of the px, py and pZ atomic orbitals as the MN CX ion pairs, yet the C—N bonds in these hydrogen bonded complexes are shorted relative to their free anions rather than lengthened the ON bonds as in the ion pairs. The difference between hydrogen bonding and ionic interactions is revealed by visualization of the molecular orbitals, shown for the OCN""HF complex in Figure 29. In the hydrogen bonded anion complexes, the molecular orbitals of the anion and the hydrogen bond donor are not separately localized, as iS true for the ion pairing interactions. Rather, the molecular orbitals of both the anion and the hydrogen bond donor are delocalized throughout the adduct. The dramatic reduction of the C -N bond length in FH'"NCX' complexes results from the delocalization of the p" and Py orbitals from HF into the C-N bond of the anions. The reduction of the ON bond length in (Li*)2NCX' triple ions and doubly hydrogen bonded adducts, (DH)2"'NCX', is 3130 reflected by the molecular orbitals. The electrostatic interactions to the 1: cloud of the C‘N bond pull electron density away from the anion thereby lengthening the bond. 6- Electrostatic potential in ionic and hydrogen bonded complexes The electrostatic potential in OCN’, SCN' and SeCN' shows that the negative charge is distributed throughout the anions. The cyanate anion has the most negative electrostatic potel‘ltial, the selenocyanide anion the least, shown in Figure 30. This is consistent with ion pairing studies of these anions, where OCN' forms stronger ion pair complexes than SCN‘ and SeCN'.70 Ion pairing and hydrogen bonding dramatically alter the electrostatic potential in the OCN‘, SCN' and SeCN' anions. The perturbation in the electrostatic 88 Figure 29. Contour plots of a 1: molecular orbital in HF extending into the C-N bond of OCN‘, thel‘eby increasing the electron density surrounding the C—N bond in the OCN'MHF complex (Green to Purple: -0.26x10‘5 to -1.6x1015 m'm, Green to Red: mama” to 1.6x1015 m‘m; stepsize 0.2mm15 m'm) 89 potential by ionic and hydrogen bonding interactions to OCN' are shown in Figure 31. In this series, the largest perturbation from the OCN' electrostatic potential is seen in the lithium cyanate ion pair. The most negative electrostatic potential in Li"NCO' ion pairs is on the oxygen; the same result holds for the Li+NCS' and the Li+NCSe‘ ion pairs. The most negative electrostatic potential in OCN‘"HF is also on the oxygen, yet the nitrogen end also has a substantial negative potential. The weakest hydrogen bonded complex, OCN""HOH, retains significant negative electrostatic potential on the nitrogen end of the anion. The electrostatic potential in SCN' and SeCN' anions respond to hydrogen bonding interactions similarly. The -1.36 J/C isosurface of the electrostatic potential in Li+NCO', shown at the bottom in Figure 31, illustrates a peculiarity found in experimental studies of alkali cyanates. No ion pair with the alkali metal bound to the oxygen end of the cyanate anion has been identified experimentally. Although the calculated atomic charges suggest that ion pairing to either the nitrogen or oxygen end of the cyanate anion are almost equally likely, the electrostatic potential of OCN', shown at the top in Figure 30, predicts quite correctly that the nitrogen end is preferred for electrostatic interactions. The calculated electrostatic potential for Li”NCO' also demonstrates that the interaction with another Li+ cation will occur at the oxygen end of the complex, thereby forming the Li*NCO'Li+ triple ion. Chabanel and coworkers have observed this complex experimentally.75 The sensitivity of these anions to the strength of a hydrogen bonding interaction provides important implications for solution complexes, especially for weak hydrogen bonded complexes. The electrostatic potential correctly predicts that a 1:1 complex will be formed at the nitrogen end by either hydrogen bonding or ionic interactions. More importantly, these calculations provide predictions for the 1:2 complexes. Unlike the situation for cations and strong hydrogen bond donors, for weak hydrogen bond donors a second interaction at the nitrogen end will be more favorable than an electrostatic attraction 90 Figure 30. Electrostatic potential for OCN‘, SCN', SeCN’ Red to Green: 0125 to -0.275, stepsize 0.015 91 Isosurface .680 J/C Isosurface -6.26 I/C OCN‘-“HOH Isosurface -5.85 J/C Isosurface -1.36 J/C Li*NCO‘ Figure 31. Molecular electrostatic potential of OCN‘, OCN""HOH, OCN'"'1-IF, Li*NCO‘ 92 to the chalcogen end of the atom. Experimental investigations by Schultz et al. have confirmed this prediction.137 Studies of the OCN' anion in methanol, formamide, and N - methylformamide have shown that the methanol can hydrogen bonds to both ends of the cyanate anion, but the amide solvents do not, even though the extent of solvation of the cyanate anion is greater in formamide than in methanol.137 7. Vibrational Frequencies in ionic and hydrogen bonded complexes Comparison of the predicted to the observed stretching frequencies in the alkali and alkaline earth metal cyanates and thiocyanates, where possible, shows that there is good agreement between the theoretical predictions and experimental results, as shown in Figures 32 and 33. The calculations tend to overestimate the shifts in the vibrational modes of the alkali cyanates by a factor of four. This difference has also been seen experimentally by Devore, who found that the K*NCO' ion pair is very sensitive to the medium differences between the gas and solution phases.78 Indeed, the predicted value for the frequency of the Von stretch for K*NCO' is 2193 cm“, which is very close to the gas phase value of 2206 cm‘1 reported by Devore.78 The difference between predicted and experimentally observed stretching frequencies for the thiocyanate ion pairs and dimers is much smaller than that form OCN' systems, which is not surprising in knowing that the thiocyanate anion is less sensitive to medium effects. Due to computational costs, the (Li“NCS')2 and (Li”NCSe')2 complexes were not studied. Rather, the (Li*)2NCS' and (Li*)2NCSe' triple ions were used to model the frequency perturbation in the lithium thiocyanate and selenocyanide dimers. The calculated frequencies of the stretching modes in (Li+)2NCS‘ and (Li+)2NCSe' qualitatively reproduce the negative frequency shifts reported for the vCN mode of the dimers.“1 '9 Interactions with the nitrogen end of OCN‘, SCN', and SeCN' by an alkali metal, alkali earth metal or hydrogen bond donor increase the frequency of both stretching modes. The largest perturbations are predicted for the Be“ complexes, reflecting the large change 93 2350 l 2325 o LiNCOLi 2300 v: o LiNCO o NaNCO 2250 - / 2225 Ab initio v01 frequencies (cm") OCN' r = 0.95 O 21ml a . e + 2125 2150 2175 2200 2225 Experimental frequencies for the v(CN) stretch (cm") Figure 32. Ab initio frequencies of the VCN stretch in alkali cyanates versus experimental frequencies of the Von stretch of alkali cyanates in DMSO75 2100 2075 l . LiNCSLi/ o BeNCS’ 2050 -. 2025 - o CaNCS‘ /MgNCS yLiNCS 2000 > e aNCS 1975 SCN’ o KNCS Ab initio vm frequency (cm") 1950 ‘- 1925 o LizNCS’ r = 0.876 1900 i 1 ~ -- 1 1 4 2025 2050 2075 2100 2125 Experimental v(CN) frequency (cm") Figure 33. Ab initio frequencies of the VCN stretch in alkali thiocyanates versus experimental frequencies of the vCN stretch of alkali thiocyanates in aprotic solvents (LizNCS* is compared with the experimental value for (LiNCS)2)"'82 94 in the bond lengths of the anions in Be2+NCX' ion pairs. Interestingly, perturbations by the I—IF hydrogen bond donor produce frequency shifts as large as those calculated for the alkali metal cation ion pairs. The perturbation of OCN‘, SCN' and SeCN' frequencies caused by interactions of alkali metal cations, alkaline earth metal cations and hydrogen bond donors at the chalcogen end of the anion is more complex. For the OCN' anion, these interactions raise the frequency of the vCN stretch, although the increase is not as large as those seen in analogous complexes where the interaction is at the nitrogen end. Alkali metal cyanates have v00 stretching frequencies at higher energies with respect to OCN’. However, hydrogen bonding at the oxygen of OCN’ red shifts the frequency of the VCO stretch. Ion pairing to the sulfur atom in SCN' and to Se in SeCN‘ by Li+ shifts the vCN modes in these anions to lower frequencies and the VCX modes to higher frequencies with respect to their free anions. These results for the stretching modes in Li+SCN' and Li+SeCN' ion pairs are substantially different from the experimental observation for the Ag‘SCN' ion pair.”120 The reason is presumably the difference in the geometries. The LiSC and LiSeC angles in the Li+XCN’ ion pairs are acute (< 70°) and the anions are bent. Hydrogen bonding by HF, HOH or MeOH with SCN‘ and SeCN’ shift the vCN stretch to higher frequencies and lower the frequency of the vCX stretch with respect to the free anions. Triple ions of the form Li”NCX'Li+ show the highest frequency stretching modes. Indeed, the calculations predict that these complexes can be easily distinguished spectroscopically from the free anion, ion pairs, dimers and tetramers, since they have the highest frequency VCN stretch for ionic complexes of a given metal cation. The triple ions of the form (Li+)2NCX', which we employ to approximate the dimer species, red shift the . fre’quency of the VCN mode and blue shift the frequency of vCX with respect to Li*NCX' ion Pairs - The red shifts in the vCN mode of (Li*)2NCS’ and (Li")2NCSe' are so large that the VCN frequency in these anions is lower than that of their respective anions. The use of these 95 triple ions to approximate dimers was confirmed by the (Li*NCO')2 dimer, where the calculated stretching modes are close to those of the (Li*)2NCO' triple ion. The frequency shifts in the OCN', SCN', and SeCN' anions can be understood by relating the frequencies to the force constants for the bonds. Selected examples of changes in the bond lengths, force constants and vibrational frequencies are collected in Table 23, where the complicated nature of the stretching modes for these anions is revealed. In the OCN‘ complexes, there is little correlation between the frequency shifts of the vCN mode and changes in the FCN force constants. For example, the FCN force constant in (Li*)2NCO' is reduced by 93 N/m with respect to OCN', yet the frequency of the vCN mode is increased. The greater dependence of the vCN mode on the ON bond stretching force constant for selenocyanide complexes is illustrated by the Ca2*NCSe' and (Li*)2NCSe' species. In Ca2+NCSe‘, the frequency of the VCN mode is slightly decreased with respect to SeCN’, reflecting a balance of the reduction in the FCN and growth in the FCSe force constant. In (Li+)2NCSe', the dramatic reduction in the FCN force constant is not similarly offset by the che force constant, so that the predicted frequency of the VCN mode in this complex is StIIbStantially lower than that in SeCN'. The vCO and vCSC modes in OCN' and SeCN' c0111plexes exhibit similar complexities. For example, the reduction in the FCO force constant for Li”OCN' is more than balanced by an increase in the FCN force constant, reSulting in an increase in the frequency for the Vco stretch. As illustrated in Figure 34, the predicted changes in the FCN and ch force constants in the OCN', SCN', and SeCN' anion complexes can be related to the calculated ON and C‘x bond lengths via Badger’s rule: F,j(rc - dij)3 = C; where F”. is the force constant, rc is the equilibrium bond length, dij is an adjustable parameter and C is a constant.136 The FCN force constants for the OCN', SCN', and SeCN' complexes have correlations of 1=0.809, 1:0-904, and r=0.963, respectively with the ON bond lengths in these species. The relationships between the ch and the C-X bond lengths in these species are much better: I:0-995, r=0.998, and r=0.999, respectively. The good empirical correlation between the 96 force constants and the bond lengths in these species provides an indirect connection between the vibrational frequencies and structural parameters for these anions through the force constants. Table 23. Calculated changes in bond lengths, force constants and frequency shifts for selected ionic cyanates and selenocyanides, with respect to the free anions Species Ad(CN) Ad(CX)“ A Va: A vex“ AF (N AFCX" A A (cm") (cm") (N/m) (N/m) OCN' Li‘NCO' +0.002 -0.037 +133 +145 +76 +258 Ca2*NCO' +0009 0051 +178 +17 1 +66 +385 OCN"“HF -0.009 -0.01 1 +101 +76 +125 +80 Li‘OCN' -0.020 +0.025 +73 +3 1 +147 —84 (Li*)2NCO' +0.025 -0.055 +1 19 +84 -93 +402 Li‘ NCO‘Li+ -0.021 -0.01 1 +176 +132 +242 +125 SeCN’ Li‘NCSe‘ +0.006 -0.051 +19 +250 +13 +1 1 l Ca2*NCSe‘ +0.019 -0.074 -8 +217 -125 +184 SeCN""HF -0.008 —0.014 +65 +62 +82 +24 (Li*)2NCSe' +0.037 -0.079 -124 +125 -221 +206 Li‘NCSe'Li” -0.011 -0.017 +104 +146 +154 +23 “X can be 0 or Se The hydrogen bonding and electrostatic interactions with the OCN', SCN‘ and SeCN' anions perturb not only the vibrational frequencies of the stretching modes but also the infrared molar absorptivites (see Tables 19—21). It is difficult to obtain a simple model to easily explain these absorptivity changes since this property is related to the dipole moment derivative. Nonetheless, these predictions for the absorptivities provide additional clues for the identification of solution structures. The relative intensity predictions compare well for previously identified solution species. For example, ionic interactions to the 97 1800 - E Z V 1600 - 8 >< .2 Q: E O U 'z’ _ U 1400 O .5 :2 5 53 U) S 1200 — U Q 8 c9. 8 u. '2 1000 - N 5 1.1.. 800 1 1 1 1 1 1 1.150 1.175 1.200 1.225 1.250 1.275 1.300 Bondlengths for the C-N and C-0 bonds in OCN‘ complexes (ml 10"”) Figure 34. Plot of the FCN and FCO force constants versus their respective bond lengths in ionic and hydrogen bonded OCN' complexes 98 nitrogen end of either the cyanate or thiocyanate anions by alkali metal cations increase the absorptivity of the VCN stretch, with the largest changes Li+ cations. An important application of these calculations is the assignment of vibrational bands where the identification is ambiguous. For example, the 2100-2300 cm'l infrared spectrum of Li*OCN' in DMSO where LilClO‘,‘ has been added consists of three bands, at 2203 cm", 2189 cm'1 and 2172 cm".75’77 The highest frequency band has been assigned to the Li“OCN‘Li+ triple ion and the lowest to the Li+NCO‘ ion pair.”77 The 2189 cm‘1 absorption could be assigned as either Li“(OCN‘)2 or (Li+NCO')2.75'77 Of these two possibilities, the (l.i"NCO')2 is more likely since the stretching modes of OCN' in this complex are shifted to higher frequencies than LilOCN' ion pairs (Table 19). 8. Cation Salvation As discussed earlier, the above calculations neglect the effects present in a dielectric medium and the solvation of the cation and anion. Chabanel’s MNDO calculations show that solvation of the lithium cation by water reduces the perturbation of the Von vibration in lithium thiocyanate ion pairs, dimers and tetramers.89 Likewise, solvation of the cation in the LiNCO ion pair by water, shown in Figure 35, does not shift the frequencies of the stretching modes in the cyanate anion as great as that in the unsolvated ion pair (Table 24). These ion pair models, which are calculated in a gas phase medium, exaggerate the . perturbations of the OCN' and SCN' anions. Inclusion of cation solvation markedly reduces the perturbation and should result in more quantitatively reliable results. C. Conclusions Ab intio calculations of ionic and hydrogen bonding interactions to XCN‘ anions at the MP2 level using large basis sets such as the d95v+**, correctly predict a number of trends seen experimentally in solution. The calculations successfully reproduce several important experimental observations: (1) the vCN mode for the OCN‘ anion is more 99 Table 24. Calculated bondlengths and vibrational frequencies at the HF/d95v+** level for OCN', LiNCO, and Li(HOH)3NCO Species dc” dd ch Vm OCN' 1.171 1.215 2380 1361 LiNCO 1.182 1.175 2420 1518 Li(HOH),NCO 1 . 178 l . 185 2412 1474 sensitive to ionic interactions than the VCN modes for SCN' and SeCN‘; (2) the vCS stretch in SCN’ complexes is more perturbed by ionic interactions than is the vCO mode in OCN' species; (3) the perturbation of the VCN mode of the OCN’ and SCN‘ anions in M+NCX‘ complexes increases as the charge density of the cation increases; (4) ionic interactions to the nitrogen end of SCN' increase the absorptivity of the vCN mode. Generally, the calculated frequency shifts for the stretching modes are much larger than those seen experimentally. This discrepancy results from not incorporating the solvation of the ions and the neglect of the dielectric medium surrounding the complex. The somewhat better agreement of the frequency shifts for the VCN modes in SCN' and SeCN’ complexes, compared to those of OCN', is due more to the nature of the RED. for this vibrational mode than to the level of theory. (This phenomenon is supported by the small difference between the experimental gas phase and solution phase values for the VCN stretch for K"NCS' ion pairs.) Compared to the vCO mode of OCN', the more dramatic frequency shifts calculated for the VCS and VCSC modes in SCN' and SeCN’ complexes are the result of the greater dependence of these modes on the C-X bonds. Despite the absence of a simple model that correlates the vibrational frequency shifts in ionic and hydrogen bonded complexes, it is important to note that similar electrostatic 5» Figure 35. Pictoral view of the Li(HOH)3NCO ion pair lOl interactions result in similar perturbations to the vibrational modes. For example, a single ionic or hydrogen bond along the molecular axis to the nitrogen end of these anions will increase the frequency of the vCN mode relative to the free anion. Two ionic or hydrogen bonds to the nitrogen end of these anions will decrease the frequency of the VCN mode below that of the single interaction. On the other hand, an interaction to each end of the anion results in the greatest blue shift of the VCN mode among all the species studied. The calculated frequency shifts strongly support previously assigned solution structures and provide an important resource in assigning hydrogen bonded complexes with these anions. The additional support from the predicted molar absorptivities provides added confidence in the assignment of solution structures. The direction and magnitude of a frequency shift for the stretching modes is dependent on the RED. for a given mode. The vCN mode for OCN' is always blue shifted in ionic complexes since this mode is very dependent on the C-0 bond. In the SCN' and SeCN‘ ions, negative frequency shifts are observed for dimer and tetramer structures since the VCN mode in these anions is more dependent on the ON bonds. The calculations reveal the limitations of the valence bond model in understanding the effect of electrostatic interactions on the OCN', SCN' and SeCN‘ anions. Although this model successfully predicts the structural changes and shifts in the electron distribution for ionic interactions in MNCX species, it is unable to explain many other interactions. For example, the charge density in (Li+)2NCX' triple ions is not as greatly shifted to the nitrogen end as in Be2"NCX' ion pairs, yet the ON bond in each complex is almost equally weakened. For LilNCX'Li+ triple ions, the nitrogen atom has a slight increase in negative charge compared to the free anion, yet the C-N bond is shortened, rather than slightly lengthened. Finally, hydrogen bonding interactions to the nitrogen end of these anions shorten the C-N bond rather than lengthening it, as ionic interactions do. A better understanding of how these ionic and hydrogen bonding interactions affect the electronic 102 structure of the XCN' anions is obtained by following the perturbations to the molecular orbitals, especially the 11: and 21: orbitals. The shifts in the electron density of the 11: and 21: orbitals correlate well with the structural changes in OCN', SCN', and SeCN' that result from electrostatic interactions. The 11: and 21: orbitals are somewhat localized to the ON and C—X bonds in these anions. For MNCO ion pairs, the 11: orbital is more bonding and the 21: orbital is less antibonding around the C-0 bond relative to OCN'. The maximum electron density of the 21: orbital in these ion pairs is pulled off the nitrogen atom, thereby reducing the bonding nature of this orbital. These trends in electron density correlate well with the shortening of the C—0 and lengthening of the ON bonds for these ion pairs with respect to OCN'. In Li*OCN’, the electron density shifts towards the oxygen end, with the opposite effect on the C-0 and C- N bond lengths relative to OCN'. Although all of the molecular orbitals are affected to some degree by electrostatic interactions, the structural changes as a result of electrostatic interactions are consistent with perturbations in the 11: and 21: valence molecular orbitals of the anions in these complexes. Additional support for this model is found in the population analysis. The occupancies of p, atomic orbitals are not substantially altered by electrostatic interactions. Rather, the largest change in the atomic orbital occupancies are found in the px and py orbitals which comprise the 11: and 21: molecular orbitals. This is consistent with perturbations in the 11: and 21: valence orbitals of these anions having the greatest impact on the structure changes in these anions We have also applied the results of this theoretical investigation to our study of solute-solvent interactions of OCN', SCN‘ and SeCN' anions with hydrogen bonding solvents.'38"39 The unique frequency shifts and intensity changes predicted for the stretching modes in these anions upon hydrogen bonding complexation have enabled us to identify several specific hydrogen bonded complexes. Chapter IV Salvation of OCN' A. Introduction The cyanate anion has been used to probe ionic associations in alkali cyanate solutions because its vibrational modes are accessible and sensitive to electrostatic interactions.”77 The OCN‘ anion has three fundamental vibrations, all of which are infrared active.88 The highest frequency mode, v3, which occurs at ~2140 cm", is commonly labeled the vCN stretch.78'127 The lowest frequency mode, v2, which occurs at ~625 cm", is the doubly degenerate bending mode.”127 The vCO mode, V], which should be observed at ~1250 cm", is in Fermi resonance with the first overtone of the bending mode (250m)."‘88‘78'm The Fermi resonance couples and splits the VCO and 250CN absorptions, resulting in two bands of similar intensity in DMSO solutions that are centered at 1240 cm'l and displaced by 44 cm" on each side.77 Hereafter, absorption in this region will be termed VCO & 2500,. It should be noted that the labels vCN and VC0 are convenient, but somewhat misleading, since these vibrations are not localized at either the ON or CO bonds.88"39 Rather, the vCN mode resembles the antisymmetric stretching vibration in CO2 or N 3188 Previous studies of alkali metal cyanates have shown that the stretching modes in the cyanate anion are very sensitive to electrostatic interactions. For example, in lithium cyanate ion pairs (LiNCO) in DMSO solutions the vCN mode increases by 34 cm‘l and the vCO and 250CN absorption increases by 21 cm'1 from their respective absorptions in ‘free’ OCN'. Moreover, a number solution species (ion pairs, triple ions, and dimers) have been identified by their characteristic vibrational frequencies. 77'88'78"27 Because of these unique frequencies the factors involved in ionic association have been studied by vibrational spectroscopy. Therefore, the cyanate anion is a good probe of hydrogen bonding interactions since the stretching modes in this anion are very sensitive to electrostatic interaction. A further advantage of this anion is that each of the nuclei in the OCN' anion have magnetically active isotopes that can be studied by multinuclear N MR spectroscopy. These 104 105 two complementary methods will be employed to obtain information about solute-solvent interactions. B. Results 1. Infrared Measurements The VCN stretching vibration of the OCN' anion is sensitive to interactions in solution. Hydrogen bonding broadens the band and shifts the peak frequency of the ch band with respect to this mode in solutions of OCN’ in aprotic solvents, as shown in Figure 36 and summarized in Table 25. The bandshape of the spectral envelope is complex and results from several overlapping component bands that can be resolved by techniques such as Fourier self deconvolution. Quantitative information about the component bands was obtained by fitting the observed spectral envelopes with Gaussian-Lorenztian bandshapes. The resulting peak positions, bandwidths and absorbances are collected in Table 26. In the aprotic solvents DMF and NM, the vCN mode consists of two component bands. The more intense one is assigned as the ‘free’ vCN vibration, whereas the weaker component (4—5% of the main band) is attributed to a hot band transition emanating from the first bending vibrational excited state (v3 + v2 - v2), located below the fundamental.77 Dramatic changes are observed in the 2200-2100 cm‘1 spectral envelopes for ~0.1M NBu4OCN in the hydrogen bonding solvents: NMF, FA, and MeOH, with respect to the VCN mode in DMF and NM. In NMF, the vCM spectral envelope is comprised of two bands, each about a factor of two broader than the absorption in aprotic solvents. Only the higher frequency component is observed in FA solutions. Interestingly, the vCN mode of OCN' in MeOH consists of three constituent bands, two of which are similar to those seen in the amide solvents, plus a new band at a higher frequency. Unfortunately, the vco & 250CN and 5001 vibrations are obscured by the hydrogen bonding solvents, so no information could be obtained about these modes from the neat solution spectra. Table 25. Infrared parameters for the vCN mode of ~0.1 M NBu4OCN in several solvents 106 VCN (cm") FWHH (cm‘l) Solvent system $0.5 cm" 10.5 cm" DMF 2136.5 8 NM 2142.0 9 NMF 2155.5 18 FA 2159.0 16 MeOH 2161.5 21 Table 26. Peak position, band width, and molar absorptivity parameters of component bands of the vm spectral envelope of ~0.1M OCN' in DMF, NM, NMF, FA, and MeOH Solvent system Peak Position vm (F WHH) Molar ( cm" ) (cm") absorptivity (M"cm") DMF 2136.5 10.1 7.9 :t0.1 21992125 2124.3 :1: 0.1 6.5 :1: 0.1 125 :1: 5 NM 2142.5i0.1 8.81201 211121:4 2130.4 :1: 0.2 8.1 i 0.2 92 :1: 4 NMF 2151.7i0.5 18.4:t0.1 1121 :75 2157.3 :1: 0.2 16.3 :1: 0.1 593 :1: 75 FA 2158.5i0.1 15.6iO.I 1652i4 MeOH 2151.8:t0.1 13.2:l:0.l 233i28 2160.6i0.2 12.1 21:07 418212189 2166.4 :1: 1.0 18.4 i 0.8 456 :1: 94 Molar absorbtivity (M"cm'l) 107 2500 — 2000 -1 FA NMF 1500 .. I 1000 _ MeOH - \ DMF 500 _ \ O "" 1 r 1 l 1 l 1 l 1 l l l L 2190 2180 2170 2160 2150 2140 2130 2120 2110 Wavenumbers (cm") Figure 36. VCN stretch of ~0.1 M NBu4OCN in DMF, NM, NMF, FA and MeOH 108 A clearer picture of the formation of anion-solvent adducts can be obtained by adding either the protic solvent or the cyanate anion to a dilute solution of the other in a weakly solvating (‘inert’) solvent, such as nitromethane (NM), has a low donor number.9 Infrared studies of the C-H stretching modes for NM show that this solvent only weakly solvates anions.”0 Thus, the solvation of the cyanate anion by methanol can be followed by measuring the 2200-2100 cm’1 spectral envelope of NBu4 OCN in NM as a function of added MeOH. When no methanol is present, the vCN stretch of OCN' is a sharp band at 2142 cm“. As MeOH is added, the intensity of the 2142 cm'I band falls, with the spectral envelope broadening and shifting to higher frequencies, as shown in Figure 37. Likewise, similar curves are seen for the FA and NMF titration series, except that the extent of complexation is lower (Figures 38 and 39). At intermediate concentrations of hydrogen bonding solvent, an isobestic point is formed owing to the formation a 1:1 complex of the OCN' anion with MeOH, FA or NMF. With higher concentrations of protic solvent, the spectral envelope continues to broaden and shift to higher frequencies and the isobestic point is lost. These results imply that the cyanate anion initially forms a 1:1 complex and that an additional complex is formed at higher protic solvent concentrations. In order to address the underlying composition of the spectral envelopes, the lineshapes were Fourier deconvolved (Figure 40). The deconvolution procedure reveals that the absorbances of the v3 fundamental and the (v3+v2-v2) mode of the OCN' anion dramatically decrease as the concentration of MeOH increases. Concurrently, a new band at 2154 cm‘1 grows. Toward the highest MeOH concentrations employed in this titration study (>2 M), the intensity of the 2154 cm" band decreases and a new band at 2163 cm" grows. Similar results are seen in the titration studies for NMF and FA. In these titrations, the free OCN’ absorption does not decrease as fast with the amide solvents as with MeOH. The new bands that appear in the amide titration series are not as blue shifted as those for the MeOH series Quantitative information can be obtained by fitting the absorption envelopes to Gaussian-Lorenztian sum components, which yields the absorbance, bandwidth, and peak 109 0.7 — 0.6 — 0.5 -~ 0.4 A 0.3 — Relative Absorbance 0.2 — 0.1 — 0.0 — I I I I I I F T I I I 17 I I I I I I I I 2200 2190 2180 2170 2160 2150 2140 2130 2120 2110 2100 Wavenumbers (cm") 2.00 —— 1.75 — _, 2142 cm / 1.50 — 2155 cm'1 1.25 — Relative Absorbance 8 1 2163 cm" l I I I I I I I I 2200 2190 2180 2170 2160 2150 2140 2130 2120 2110 2100 Wavenumbers (cm") Figure 37. vCN stretch of 0.1M NBu4OCN in NM (top) and the Fourier deconvolution of the vCN stretch (bottom) as a function of MeOH added 110 0.7 — 0.6 —1 0.5 —+ 8 4 5 I— o (I) _ f2 0 0 3 — .2 E —1 3; 0.2 - 0.1 — 0.0 I I I I I I I I I I I I l I I I I I I I 2200 2190 2180 2170 2160 2150 2140 2130 2120 2110 2100 Wavenumbers (cm“) 2143 cm‘1 21512 cm’1 0-75 ‘ 2157 cm" Relative Absorbance 2131 cm'1 I I I I I I I I 2200 2190 2180 2170 2160 2150 2140 2130 2120 2110 2100 Wavenumbers (cm“) Figure 38. VCN stretch of 0.1M NBu4OCN in NM(top) and the Fourier deconvolution of the VCN stretch (bottom) as a function of formamide added Relative Absorbance Relative absorbance 0.6 — 0.5 -— 0.4 - 0.3 -— 0.2 —< 0.1— 111 0.0 l ' l ' l ' l ' l ' l ' I ' l ' I ' l 2200 2190 2180 2170 2160 2150 2140 2130 2120 2110 2100 Wavenumbers (cm") / 2142 cm'1 Tm 1‘1 2200 2190 2180 2170 2160 2150 2140 2130 2120 2110 2100 Wavenumbers (cm") Figure 39. vcN stretch for 0.1M NBu4OCN in NM (top) and the Fourier deconvolution of the VCN stretch (bottom) as a function of N-methylformamide added 112 0.6— O U C 4 3 2154cm" ‘5 0.4— (I) .D 4: - 7; 2163cm" a: 0.2— 00‘ 2143cm" l'l'l'l‘l'l‘l‘l 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Concentration of MeOH (M) 0.6— 0 o .1 t: :6 -1 'E 0.4+ 215lcm 8 .D a < 7; 0,2- 2159cm" a: 0.0— - 2142cm" l‘l‘l'l'l'l'l'l 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Concentration of FA (M) 0.6— 0 E) 0.4—1 2 s - 2151cm'l (I) .0 <11 02— B 0‘ - 2159cm" 0.0— O O 2142cm" I‘TFTfil'l‘l'lTl 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Concentration of NMF (M) Figure 40. Absorptions for the component bands in the 2200-2100 cm’1 region for ~0.1M NBu4OCN/NM solutions as a function of MeOH (top). FA (middle) or NMF (bottom)concentration. 113 position of each constituent band. The absorption curves in the methanol series can be well simulated with four components: at 2130 cm" for the hot band, at 2143 cm’l for the free anion, at 2154 cm‘1 for the first solvated complex, and at 2163 cm‘l for the second solvated complex. The intensities of the fundamental constituents, plotted in Figure 41, are qualitatively similar to those obtained by the Fourier self deconvolution of the observed spectral envelopes. Similar titration experiments were performed by adding NMF and FA to ~0.1M NBu4OCN in NM. Curve fitting and Fourier self deconvolution of the 2200- 2100 cm‘l spectral envelopes also show that the lineshape in each case is a convolution of four component bands: the hot band (v3+v2-v2), the ‘free’ VCN band, a 1:1 hydrogen bonded complex, and a 1:2 hydrogen bonded complex (from lowest to highest frequency). The 1350-1180 cm’1 spectral region contains the vco & 2800., bands. In a 0.1 M NBu4OCN solution in NM, these bands are centered at 1286 cm‘l and 1198 cm". Addition of MeOH broadens the absorptions and shifts the maxima of the spectral envelopes to higher frequencies. Moreover, new bands grow in with the increasing methanol concentration, as illustrated in Figure 42. Fourier deconvolution is not necessary to determine the positions of these new bands: at 1309 cm’1 and 1297 cm‘l for the higher frequency component of the vcJO & 250CN absorption, and at 1219 cm" and 1209 cm’l for the lower frequency component. The concentration dependence of the 1286 cm" and 1198 cm‘1 peaks is qualitatively similar to that of the 2142 cm’1 ‘free’ vcN stretch. Due to a strong absorbance of the amide solvents, only the lower frequency component of the VCO & 280m absorption is measurable. As in the studies of the vCN mode, the new bands that are a result of amide-OCN' interactions are not as blue shifted as those found in the OCN""HOMe complexes (Figure 41). In the FA titration, two new bands are observed at 1207 cm’1 and 1219 cm". Due to spectral interferences in the NMF titration, only a portion of the titration could be collected which shows that a new band forms at 1207 cm‘1 as the NMF concentration increases. 114 Absorbance (rel.) 0.08 -— <'—‘ 1286 cm'1 0.06 ‘1 1297 cm'l a ‘—- 1198 cm" 0.04 — 1209 cm‘l 0.02 — 0.00 W 1 l l l I T j 1320 1300 1280 1260 1240 1220 1200 l 180 Wavenumbers (cm") Figure 41. vCN and 250014 stretch for ~NBu4OCN in NM as a function of MeOH added (a. 0 M, 0.04 M, C. 0.08 M, d. 0.12 M, and e. 0.16 M MeOH) Rel. Absorbance Rel. Absorbance 0.020 0.016 — 0.012 —« 0.008 0.000 115 1207 cm'1 1219 cm" 0.10 T 0.08 I 0.06 - Wavenumbers (cm") (a. 0 M, b. 0.04 M, c. 0.13 M, d. 0.17 M, 3.1 M FA) 1207 cm’l 0.04 — 0.02 - 0.00 — l l I 7 1220 1210 1200 1190 1180 Wavenumbers (cm") (a. 0 M, 0.04 M, 0.08 M. d. 0.51 M) Figure 42. vCN and 2500: stretch for ~0.1 M NBu4OCN in NM as a function of FA added (top) or NMF added (bottom) 116 The spectral envelopes of the vCO & 2500,, bands also can be modeled by Gaussian- Lorenztian sums. The lower frequency band (1230—1180 cm") is more amenable to curve fitting, since this spectral region has fewer interferences from the protic solvents. The concentration dependence of the three bands in the 1230-1180 cm'1 region is similar to that of the component bands in the VCN stretch, in that the absorbance of the band attributable to ‘free’ OCN’ plunges with increasing MeOH concentration, with the concomitant growth of new bands. Qualitatively similar results are obtained in this spectral region from the NMF and FA titrations. The results in both stretching spectral regions for the three protic solvents are summarized in Table 27. Table 27. Summary of peak positions and band widths of the curve-fitted component bands of the vCN stretch of 0.1M NBu4OCN/NM solutions with NMF, FA, or MeOH added. VCN V10 86111 V60 & 250011“ Via Eco Solvent Peak Position F WHH Peak Position F WHI-l flstem ( cm") ( cm") (km/mo!) ( cm" ) (cm") (km/mag NMF 2141.8 i 0.8 10.4 i 0.8 309 :1: 7 1198.2 i 0.5 11.2 i l 5.4 :1: 0.6 2150.8 :t 0.8 13.5 i 0.2 358 i 7 1207.2 i 0.2 5.0 i l 1.8 i 0.4 2159.2 :1: 0.2 13.9 i 0.8 320 i 20 b b h FA 2142.0 i 0.5 10.6 i 1.2 322 :1: 7 1199.1 :1: 1.0 10.0 :1: 1 4.8 i 0.6 2150.6 :1: 0.9 12.9 :t 0.6 371 i 7 1207.1 1: 0.5 6.6 i l 2.4 i 0.6 2158.9 :1: 0.4 12.8 :t 1.0 330 :1: 20 1215.4 i 0.4 6.2 :t 1 0.8 i 0.2 MeOH 2142.8 :1: 0.7 10.5 i 1.6 311 i 4 1198.3 :1: 0.5 9.8 i 0.4 5.0 i 0.4 2154.0 :1: 0.9 12.3 i 1.2 399 i 4 1209.0 i 0.5 8.2 :t l 3.6 i 0.6 2163.3 :1: 0.4 14.0 i 0.3 450 :1: 20 1219.4 :1: 0.4 8.2 :1: 1 2.2 i 0.4 “only values for the lower frequency component are reported bobscured by an NMF vibration 117 2. Assignments The structure of the primary solvation sphere surrounding the cyanate anion in neat solutions of NBu4OCN in MeOH, FA, and NMF can be understood by identifying the component bands of the VCN stretch. The concentration behavior of the infrared absorbances of the component bands in the titration studies of the cyanate anion with protic solvents clearly shows that the OCN' anion is involved in two equilibria. Moreover, the increase in the absorbance of the 2163 cm‘1 component in the MeOH titration comes at the expense of the 2154 cm‘1 absorption (Fig. 41), suggesting that the 2154 cm’1 band can be attributed to a 1:1 methanol:OCN' adduct and the 2163 cm‘1 band arises from a 2:1 MeOH:OCN' complex. Since the negative charge is distributed between the oxygen and nitrogen ends of the anion, the methanol in these 1:1 complexes might be hydrogen bonded to either the nitrogen or oxygen atom. The OCN""HOMe and NCO""HOMe complexes can be easily distinguished from one another by their infrared spectra. Ab initio calculations of these complexes show that the frequency of the vCN mode will increase relative to free OCN' in either case. However, these calculations predict that the frequency of the VCO stretch will increase in a OCN""HOMe complex and decrease for a NCO"“HOMe adduct.I39 With respect to free cyanate, it is also predicted that the absorptivity of the Vcn mode in OCN""HOMe will 139 increase and that of the vCO mode will decrease. The opposite trend is predicted for the molar absorptivities of the stretching modes in a NCO""HOMe complex. From the observed spectroscopic behavior, the 1:1 complex can be assigned with confidence as OCN""HOMe. This assignment is supported by experimental studies of alkali cyanates, where positive frequency shifts in the vCN and v00 & 250CN modes of OCN' with respect to OCN‘ have been assigned to an alkali metal ion pairing to the nitrogen end of the cyanate anion.”77 Likewise, the infrared spectra of OCN' in the NMF and FA titration series are also consistent with the amide solvents hydrogen bonding to the nitrogen end of the cyanate anion. 118 There are three possible ways that two MeOH molecules can interact with OCN’: OCN""(HOMe)2, NCO“"(HOMe)2, and MeOH"'OCN""HOMe. Of these possibilities, the NCO'"'(HOMe)2 complex, where two methanol molecules form a hydrogen bond to the oxygen end, can be easily dismissed since no NCO""HOMe adducts were observed. Moreover, the accompanying red shifts in the frequencies of the vCO & 280CN modes relative to those in OCN' are not observed experimentally. Ab initio calculations for 1:2 complexes of cyanate with water or hydrogen fluoride indicate that it is difficult to distinguish the three 139 structures based on the frequency shifts in the vCN mode alone. Again, the frequency of the VCO mode and the molar absorptivities of the vCN and vCO modes provide the best way to discriminate between the MeOH'"OCN"“HOMe and the OCN"“(HOMe)2 complexes. The experimental frequency shift values and absorptivities in the MeOH, FA, and NMF titration series suggest that both hydrogen bonding interactions are with the nitrogen end of the cyanate anion in all three 2:1 solvates. The vibrational assignments from the titration studies provide a firm foundation for assigning the vibrational components of the vcN mode of 0.1M OCN' in neat solutions of the anion in MeOH, FA, and NMF. They are collected in Table 28. The 2152 cm'l and 2157 cm’l components observed in neat N MF are similar to those seen in the titration experiment. In neat FA solutions the 2151 cm‘1 band is absent; only the 2159 cm'1 component is observed. In neat MeOH solutions the component bands of the VCN mode are shifted by ~2 cm‘1 from the values determined in the titration studies. The 2152 cm‘1 component can be attributed to the OCN""HOMe complex and the 2161 cm‘1 component to the OCN""(HOMe)2. An additional component at 2166 cm’1 is observed in neat MeOH solutions which can be assigned to the MeOH“'OCN""HOMe complex. The assignment is based on ab initio calculations, which predict the frequency of the vCN mode will be higher than those of the OCN""HOMe and the OCN""(HOMe)2 species.I39 Additional support is provided by the experimental studies of alkali cyanates, where of the VCN frequency for the LiOCNLi+ triple ion is the highest of all of lithium cyanate complexes studied to date.”77 119 Table 28. Assignments for the VCN and VCO& 25OCN modes for ~0.1M OCN' in neat solutions with MeOH, FA, or NMF and in titration studies with these solvents T01 vent system VCN component Vc~ component Vco & 260m component bands in bands in titration bands in titration Assignment neat solutions studies studies NMF ,CH3 OCN'-"H" N 2152 cm'I 2151cm‘I 1207 cm‘l ‘CHo .CHs ,N-CHo ,H OCN' 1 2157 cm‘l 2159 cm'l ‘ N- CHO CH3 FA ,H OCN-...H-N , 2151cm" 1207 cm'1 ‘CHo V ’N-CHO .° H OCN 2159 cm" 2159 cm‘1 1215 cm'1 ' H‘ 1:1' CHO H MeOH OCN'---H- Q 2152 cm'I 2154 cm'1 1209 cm'1 CH3 ’O'CH3 ,H OCN- 2161 cm'1 2163 cm‘1 1219 cm‘I H\ O ' CH3 Ha; O‘H ”'0CN'WH'Q 2166 cm'1 b b CH3 aobscured by solvent; bnot observed at the studied concentrations 120 3. NMR Measurements Chemical shifts from the 14N and 170 NMR spectra of the cyanate anion in the five solvents employed in this investigation are summarized in Table 29. The information provided by NMR spectroscopy is less specific because the observed chemical shifts are a population average over all species involved in the solution equilibria. Nonetheless, these NMR results are consistent with the vibrational assignments. In NMF, the cyanate anion is present in two hydrogen bonding environments. Relative to their values in aprotic solvents, the decrease in the 14N chemical shift (more negative) and the increase in the 170 chemical shift (more positive) suggest that the electron density of the anion is shifted from the oxygen end toward the nitrogen end. Ab initio calculations predict that a single or double hydrogen bonding interaction to the nitrogen end will shift the distribution of the '39 In FA, the further increase in the extent of the solvation negative charge in this direction. of the cyanate anion suggested by the infrared results is reflected by the increased positive shift of the ”0 NMR chemical shift. On the other hand, the 14N chemical shift becomes more positive with respect to the value in NMF solutions, which is not predicted by the model ab initio calculations. This discrepancy shows that care must be employed in correlating chemical shifts with atomic charges. The l“N chemical shift of the OCN' anion has the most negative value in MeOH solutions, where the '70 chemical shift has the smallest positive value among the protic solvents. This pattern also fits well with the infrared results. The predicted atomic charges on the anion in the MeOH"'OCN""HOMe species are similar to those of the free anion. The increase in the negative charge on the oxygen end is reflected by a decrease in the ”0 chemical shift with respect to the values for the OCN' anion in the amide solvents. The large negative |“N chemical shift of OCN' in MeOH solutions can also be explained through the three hydrogen bonded species described earlier. The MeOH hydrogen bond is stronger than those of NMF and FA, which results in a greater 1“N chemical shift in 121 OCN""HOMe and OCN""(HOMe)2 relative to the analogous NMF and FA species. Although the l“N chemical shift in MeOH"'OCN°"'HOMe should be more positive than that in OCN""HOMe and OCN""(HOMe)2, the observed chemical shift is dominated by contributions from the OCN""HOMe and OCN""(HOMe)2 complexes. Table 29. l“N and 170 NMR chemical shifts of ~0.1M NBu4OCN in DMF, NM, NMF, FA, and MeOH Solvent system "N chemical shift (10.02) ”0 chemical shift (10.1) DMF -290.76 ppm 39.0 ppm NM -293.28 ppm 38.4 ppm NMF -301.25 ppm 42.9 ppm FA -295.61 ppm 46.9 ppm MeOH -310.65 ppm 41.1 ppm 4. Thermodynamic Parameters The '4N and l7O chemical shifts of the cyanate anion were measured by NMR spectroscopy for the MeOH, NMF, and FA titration series in NM in order to complement the infrared studies. The temperature dependence of the l“N chemical shifts was also determined. The concentration dependence of the l“N chemical shift for OCN’ upon addition of MeOH, representative also of the NMF and FA systems, is shown in Figure 43. The concentration dependence displays a classic saturation curve for associating complexes. The temperature dependence of the NBu4OCN/MeOH/NM dilution series shows that the solvated complexes are more stable at lower temperatures. l7O chemical shift (ppm) 14N chemical shift (ppm) 122 -302 — 1: -25.0°c v -12.5°C -300 _ v 00°C -298 0 125°C 0 250°C ~296 — -294 — '292 l l l l 1 l l l 1 111 1.1 l Concentration of MeOH (M) 42.5 — 42.0 _. 41.5 — 41.0 e 40.5 -— 40.0 e 39.5 e 39.0 — 38-5 1 1 1 1 1 1 1 1 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Concentration of MeOH (M) Figure 43. l“N and 170 NMR chemical shifts for 0.1 M NBu4OCN as a function of MeOH added and temperature in the case for the l“N resonance. 123 In order to extract thermodynamic parameters from the concentration dependence of the l“N and '70 chemical shifts or the intensity of the component infrared bands of the vCN or VCO & 500,, absorption, an appropriate model must be applied to the N MR and infrared results. The concentration dependence of the intensities of the 2154 cm'l and 2163 cm'1 bands in the methanol dilution series implies that methanol initially forms a 1:1 complex with the cyanate anion. At higher MeOH concentrations the 1:1 cyanate-methanol complex associates with an additional methanol molecule, forming a 1:2 cyanate-methanol complex. Thus the components of the VCN absorption provide an important clue to the equilibria responsible for the concentration dependence of the N MR and infrared measurements. In addition to multiple associations of the cyanate anion with the hydrogen bonding solvents, both the self association of the hydrogen bonding solvent and ion pairing of the NBu4OCN salt must be considered and possibly incorporated into the model. Previous investigations of tetraalklyammonium salts in nitromethane have shown that ion pairing is negligible at 0.1M.2| Therefore, the NMR and infrared investigations have been modeled by three equilibria: (a) the self association of the hydrogen bonding solvent (limited to the dimerization), (b) formation of a 1:1 cyanate-hydrogen bonding solvent adduct, and (c) formation of a 1:2 cyanate-hydrogen bonding solvent complex, as illustrated in equations (8)-(10). Although these equilibrium expressions are written with the hydrogen bonding interaction at the nitrogen end of OCN', the determination of the equilibrium constants is independent of the site of the hydrogen bonding interaction. At the low concentrations of the cyanate anion employed in this work, the equilibrium constants can be written in terms of the concentrations of the species in the equilibria; the activity coefficients can be neglected since charge species are involved in both sides of the equilibria. Thus the equilibria can be expressed by equations (1 l)-( 13). 124 K1 K2 OCN""HOMe + MeOH Z OCN""(HOMe); (9) K1) 2MeOH ::_.> (MeOH)2 (10) where K1 = [OCN:"'HOMe] (11) [OCN ][MCOH] _ [OCN'"'(HOMC)2] (12) 2 [OCN""HOMe][MeOH] K1) = M1 (13) [MeOH] Equation (14) gives the total concentration of OCN ’ in terms of the concentration of the free OCN' anion, [OCN'], the concentration of the 1:1 complex, [OCN""HOMe], and the concentration of the 1:2 complex, [OCN""(HOMe)2]. Likewise, the total concentration of MeOH, CMcOH, can be written in terms of the concentrations of free MeOH, [MeOH], the 1:1 complex, the 1:2 complex, and the methanol dimer, [(MeOH),], equation (15). COCN = [OCN'] + [OCN""HOMe] + [OCN""(HOMe)2] (l4) CMeOH = [MeOH] + [OCN""HOMe] + 2[OCN""(HOMe)2] + 2[(MeOH)2] (15) 125 Substitution of equations (11)-(13) into equations ( 14) and ( 15) and solution of the resulting equations simultaneously for [OCN'] yields a quartic expression in [MeOH], equation (16). (2K,K,KD)[Me0H]4 + (K,1(2 + ZKIKD)[MeOH]3 + (K,1<,(2COCN - CMCOH) + K, + 2KD)[MeOH]2 + (K,(cOCN - CMcOH) + l)[MeOH] + CMeOH = 0 (16) The roots of equation (16) can be determined analytically, from which an expression for the concentration of free methanol can be derived in terms of the association constants (K,, K2, and KD), CMCOH, and COCN. An expression for [OCN'] can then be written in terms of the association constants and the total MeOH and OCN' concentrations by substituting equation (11), equation (12), and the solution for equation (16) into equation (14). The observed chemical shift of either the l“N or the 17O nucleus of the cyanate anion, 80b5, is a population average of those for the free anion, the 1:1 complex, and the 1:2 complex as shown in equation (17), where xfm is the mole fraction and 8f,” is the chemical shift of the free anion, x1 is the mole fraction of the 1:1 complex and 8, is its chemical shift, and x2 is the mole fraction of the 1:2 complex, which has the chemical shift 52. With the appropriate substitutions, equation (17) can be rewritten (equation (18)) such that the Bobs = Xfreesfree + XI51 + X292 (17) 60,, = K,[OCN‘][MeOH] (6, - 8m)/C0CN+ K,K2[OCN‘][MeOH]2 (8, — 51...) / COCN+ 5,,e (18) observed chemical shift is a function of the association constants (K,, K2, and KD), the free OCN’ concentration, the free MeOH concentration, and the limiting chemical shifts for OCN', OCN""HOMe, and OCN""(HOMe)2. Following substitution of the expressions for 126 [OCN'] and [MeOH], the values for the equilibrium constants, including KD, can then be calculated by fitting equation (18) to the chemical shifts observed at different methanol concentrations. These values are summarized in Table 30. The formation constants for the hydrogen bonded complexes can also be determined from the infrared band absorptivities. For example, the concentration of the 1:1 OCN""HOMe complex is proportional to the absorbance of the 2154 cm’1 component (A254) as shown in equation (19), where a is the molar absorptivity and l is the pathlength of the infrared cell. Upon substitution for [OCN'] and [MeOH], the absorbance of the 2154 cm‘l component becomes a function of the molar absorptivity, pathlength, equilibrium constants, and the total concentrations of OCN' and MeOH. The fitting of equation (19) to the absorbance of the component band resulting from the 1:1 complex also provides the formation constants of the solvated complexes and the self association of the hydrogen bonding solvent (K,, K2, and KD). Likewise, the KI association constant can be determined by relating the vCN and vCO & 280CN infrared bands that are attributable to the ‘free’ OCN' anion to the total methanol and cyanate concentrations, as shown in equation (21). This neglects the self association of the hydrogen bonding solvent, since this equilibrium is negligible when the protic solvent concentration is less than 0.5M, as shown by the small equilibrium constant, KD, for this association. The results are listed in Table 31. A2,54 = 82,54 6 [OCN""HOMe] = Kl £1154 l [OCN']f[MeOH]f (19) A2142 = 82.42 If [OCN']/(1 + K,[MeOH]) (20) The good agreement between the association values obtained from the NMR and the infrared measurements confirms that these spectroscopic measurements were modeled with the correct equilibria. The equilibrium constants for the 1:2 complexes in the amide 127 Table 30. Association constants, K,,K2 and KO, for the formation of OCN"“HB, OCN'"(HB)2, and BH'"BH where HB can be NMF, FA, or MeOH, determined by l“N and 170 NMR spectroscopy Solvent Kl (M4) K2 (M'l) K0 (Md) Temperature 1 i0. 2 ° C # NMF ”N NMR 25.0° C 4.8 :1: 0.1 0.26 i 0.04 0.06 :1: 0.01 12.5° C 5.7 i 0.2 0.28 i 0.09 0.07 i 0.02 0.0° C 6.8 i 0.2 0.28 :t 0.02 0.07 i 0.02 -12.5° C 8.3 :t 0.2 0.31 :1: 0.12 0.09 i 0.02 -25.0° C 10.0 i 0.1 0.34 :t 0.06 0.09 i 0.02 170 NMR 25.0° C 2.7 i l 0.33 i 0.20 0.04 i 0.02 FA l“N NMR 25.0° C 4.4 d: 0.1 ' 0.06 i 0.03 12.5° C 5.2 i 0.1 ‘ 0.07 i 0.03 0.0° C 5.9 :t 0.1 ' 0.11 i 0.03 -12.5° C 7.3 :1: 0.1 ' 0.10 i 0.03 -25.0° C 8.8 i 0.1 a 0.09 :1: 0.04 '70 NMR 25.0° C 3.2 i 1.0 0.53 i 0.20 0.06 i 0.03 MeOH 1“N NMR 25.0° C 8.8 d: 0.4 0.40 i 0.15 0.10 i 0.01 12.5° C 10.9 :1: 0.4 0.55 i 0.40 0.10 :t 0.01 0.0° C 13.5 i 1.0 0.71 :1: 0.17 0.16 i 0.02 -12.5°C 18.0: 1.0 0.98:0.17 0.15 $0.02 -25.0°C 21.41 1.1 l.24i0.13 0.17:t0.02 170 NMR 25.0° C 8.5 :t 3 0.47 :t 0.20 0.12 i 0.1 “not determinable 128 solvents should be considered approximate, since the measurements that are important in their determination lie near the detection limit. It is interesting to note that the 1:1 OCN‘ "'HOMe adducts, which have the largest vCN frequency shifts relative to those observed for the corresponding FA and N MF bands, have the highest formation constant for these complexes. A similar result has been obtained for alkali cyanates and alkali thiocyanates, where ion pairings that result in the largest frequency shifts of the Vcn mode, as in LiNCO ion pairs, also have the greatest ion pair formation constants.7°'75'77'32'85 Am, = 82154€[OCN""HOMe] = K,e,,5,t [OCN'][MeOH] (21) Table 31. Association constants, K1, K2, and KD, for the formation of OCN""HB and OCN"'(HB)2, and HB'"HB, where l-IB can be NMF, FA, or MeOH, determined by infrared spectroscopy Solvent K1 (Md) K2 (MJ) K1) (M'l) Temperature (24 :1: 1° C) NMF A“98 4.3 i 0.8 A2142 6.7 :1: 0.6 A2151 5.5 i 0.4 0.20 i: 0.05 0.10 i 0.01 FA AW, 5.8 :1: 0.9 2.42 5.6 i- 0.7 A215, 5.0 i 0.4 0.46 i 0.07 0.13 i 0.03 MeOH Am, 9.4 i 0.7 A2142 8.3 i 1.0 Am4 8.7 i 0.7 0.37 :1: 0.04 0.12 i: 0.02 129 From the temperature dependence of the association constants derived from the 1“N NMR measurements, the enthalpy, AH°, and the entropy, AS°, for these complexation equilibria were determined; they are summarized in Table 32. The negative enthalpy term is a result of the ion-dipole attraction between the anion and the hydrogen bond donor. The thermodynamic parameters for the complexation equilibrium in N MF, FA, and MeOH show that the entropy change is an important consideration in the formation of the complexes. There is little difference among the enthalpies and entropies of formation for the 1:1 complexes of OCN' with NMF, FA, and MeOH, although the MeOH complex has the highest AH,°. The dramatic reduction of K2 from K, appears to result primarily from the decrease in the entropy of the system as demonstrated by the methanol complexes. Table 32. Thermodynamic parameters for the formation of 1:1 and 1:2 OCN""solvent complexes in nitromethane solutions. Solvent system AH,° AS,° AH2° ASZ° (kJ/mol) (Jmol'l K") (kJ/mol) (.lmol"K°’) NMF -9.0 i 0.3 -17 :1: 1 -4.7 i 5.2 -27.8 :1: 17 FA -8.5 :1: 0.3 -16 i 2 ‘ a MeOH -12.1 i 0.7 -22i2 -12.1i0.6 47i2 'not determinable C. Conclusions The infrared and NMR spectroscopic studies of ternary nitromethane solutions provide a consistent and coherent picture of the structure and thermodynamics of anion- solvent complexes in the primary solvation sphere surrounding the cyanate anion by the hydrogen bonding solvents MeOH, FA, and NMF. Our investigations reveal that 130 hydrogen bonding interactions in the amide solvents occur at the nitrogen end of the anion. In moderately dilute, neat formamide solutions, each OCN' is hydrogen bonded to two FA molecules, whereas in MeOH, presumably a stronger hydrogen bonding solvent, the cyanate anion participates in both singly and doubly hydrogen bonded complexes. In neat methanol solutions, two distinct 1:2 cyanatezmethanol adducts are observed. An understanding of the difference between the hydrogen bonding interactions of the amide and methanol solvents is provided by ab initio calculations of the molecular electrostatic potential (MEP) of the cyanate anion in various electrostatic environments.’39 The MEP predicts the site of a potential electrostatic interaction.'“ For example, it predicts that a single ionic or hydrogen bonding interaction will occur at the nitrogen end of OCN', although population analysis suggests that the negative charge is almost equally distributed between the oxygen and nitrogen ends. In FH“'NCO‘, which models a strong hydrogen bond, the most negative MEP is found in a ring surrounding the oxygen end of the anion. The weaker hydrogen bond in an OCN""HOH complex only slightly perturbs the MEP in OCN' such that the minimum lies in an are that is slightly off the nitrogen end of the OCN' anion. These calculations demonstrate that the MEP in OCN' is very sensitive to the strength of an electrostatic interaction and that strength of the first hydrogen bonding interaction will determine the site of the second hydrogen bond.139 From the association constant values, it is clear that the amide solvents form weaker hydrogen bonds with OCN' than those by MeOH. Moreover, in the titration studies the perturbations of the VCN and Vco & 250e,,l frequencies in the cyanate-methanol complexes are larger that those observed for similar amide complexes, which also reflects the relatively greater strength of the cyanate-methanol interaction. The weaker amide interactions with the cyanate anion result in hydrogen bonding only to the nitrogen end, owing to the smaller perturbation in the MEP for the OCN' anion. The thermodynamic parameters obtained in the titration studies provide insight into the factors that are involved in the solvation of anions. Interestingly, the enthalpy and 131 entropy of formation for the hydrogen bonded complexes have similar values, which is probably a leveling effect due to the nitromethane solvent. The negative entropy values show that this term is dominated by the loss of freedom of the cyanate anion and the protic solvent upon formation of the complex, rather than by the release of nitromethane molecules that weakly solvate the complexation partners. In the MeOH titration, the entropy of formation for the second hydrogen bonded complex is more negative than that for the first complex. This is not surprising since in the 1:1 complex the nitrogen end of the anion will be less solvated by nitromethane than the free cyanate. Therefore, the entropy of formation for the second complex, A82, will be even more dependent on the loss of freedom of the complexation partners than on the nitromethane molecules liberated. The extent of solvation of cyanate in neat solutions of MeOH, FA, and NMF can be understood by considering the role of entropy in the formation of intimate ion-molecule complexes. The overall entropy loss when these solvents form a hydrogen bond with the cyanate anion will be smaller than for the association of two free particles, since these solvents are highly self-associated. This can be clearly demonstrated by comparing the degree of solvation of OCN' in NMF and FA solutions. Although the association constants for cyanatezamide complexes are comparable in NM, in neat solutions the cyanate anion is more solvated in FA than in NMF. Since FA is more self-associated than is NMF,” "'4 the entropy loss upon cyanate solvation will not be as great in FA as it is in NMF. Comparison of the solvation of OCN' in FA with that of MeOH also supports this hypothesis. The perturbations in the VCN stretch show that the OCN""HOMe hydrogen bond is much stronger than that with FA. Yet, the cyanate anion is more highly solvated in neat FA solutions than in MeOH. This seemingly inconsistent can be rationalized by the greater self association of FA relative to MeOH; the entropy loss upon hydrogen bonding with OCN' will be smaller in FA solutions. Chapter V Salvation of SCN' and SeCN' A. Introduction: The thiocyanate and selenocyanide anions have been used to probe ionic association in alkali and alkaline earth metal complexes in solution since the stretching modes for these anions are easily observed and are sensitive to electrostatic interactions.50'70'74'8'5M4“44 Both the SCN' and SeCN‘ anions have three fundamental vibrations, all of which are infrared active. The highest frequency mode (v3), which is commonly labeled the vCN stretch, occurs at 2058 cm" for SCN' and 2066 cm’1 for SeCN' in DMSO.83'88'”2 The lowest frequency mode (v,), which occurs at 465 cm’1 for SCN' and 420 cm" for SeCN', is the doubly degenerate bending vibration, labeled 55m and 8%, respectively.”13 The second stretching vibration (v,) occurs at 735 cm‘1 in SCN‘, commonly labeled VCS. The comparable vCSc vibration for SeCN' is easily observed experimentally and is predicted to '39 The justification for these labels was have a very low dipole moment derivative. demonstrated by ab initio calculations, from which, for example the potential energy distribution for the vCN modes of these anions was found to be about 90% C-N stretch.139 The perturbation of these stretching modes by electrostatic interactions have been ”'70'74'3"9"'42"“ From these investigations, the structures studied by an number of authors. of numerous solution species have been identified by their unique vibrational fiequencies. In general, alkali and alkaline earth metal cations interact with the nitrogen atom of SCN', which raises the frequencies of both stretching modes with respect to the unperturbed anions. Similar results have been reported for the limited studies of SeCN‘ systems. Interactions to the sulfur end of SCN', as exemplified by the Ag”SCN' ion pair, lowers the frequency of the VCN stretch and raises the frequency of the vCS mode relative to SCN'.l5 In dimer systems, such as (LiNCS)2 and (LiNCSe)2, each anion is bridged by a two lithium cations which reduces the frequency of the vCN vibration and increases the frequencies of the vCS and VCSe modesm" Likewise, hydrogen bonding interactions with these anions perturb the frequencies of the stretching modes, from which inferences about structures of the complexes that are 133 134 formed can be drawn. Perelygin observed that hydrogen bonding solvents shift the peak maximum and broaden the spectral envelope of the VCN stretch of SCN' compared to those seen in aprotic solvents.3s Corset assigned a blue shift of the vCN mode when he added phenol to a solution of NBu4SCN in CCl4 to the formation of hydrogen bond at the nitrogen of SCN'.38 Gill and coworkers measured the infrared spectrum of NBu4SCN in MeOH and deconvolved the Vet»: spectral envelope into six component bands.43 Another study of SCN' in methanol by Bencheikh showed that only four bands are necessary to simulate the VCN spectral envelope (2090, 2071, 2055, and 2037 cm").45 Recently, Hochstrasser and coworkers have studied the vibrational relaxation of SCN' in D20 and MeOH.“°’41 Their results are consistent with the solvent hydrogen bonding to the nitrogen end of the thiocyanate anion. X-ray diffraction studies of NaSCN in methanol imply that two methanol molecules are hydrogen bonded to the nitrogen of SCN‘, with a bond angle of 120°.23 As the perturbations in the vibrational modes reveal, electrostatic interactions with SCN' and SeCN' alter their electronic structure. All of the nuclei in the SCN' and SeCN' 9 and with the exception of 33S which has a large anions have magnetically active isotopes,5 quadrupole moment, the shifts in electron density can be easily studied by multinuclear NMR spectroscopy. l5N NMR spectroscopy has been employed to identify the isomers of organo—thiocyanates and isothiocyanates. RNCS compounds are characterized by a l5NMR shift of -275 ppm. whereas RSCN compounds have a l5N chemical shift of -100 ppm, which are both large shifts from that of the free anion at -165 ppm.145 Ionic associations of SCN' with Li” have been studied by l5N NMR spectroscopy, where interactions to either the sulfur or nitrogen atoms were identified by the paramagnetic or diamagnetic shifts, respectively.I46 Musikas et a1. studied the binding of SCN' to several lanthanide ions by the perturbations in the chemical shift and relaxation time of the nitrogen nuclei.147 Similarly, the 77Sc chemical shifts for SeCN' are extremely sensitive ionic interactions.‘48 Ionic interactions to the nitrogen end increase the shielding from -273 ppm for SeCN' in 135 DMSO to -318 in a zinc complex with SeCN'. Interactions to the selenium end, exemplified by a mercury complex, decrease the shielding to -191 ppm. In contrast to the large shifts observed for the nitrogen and chalcogen nuclei, the 13C resonances in SCN‘ and SeCN' are not very sensitive to bonding at either end of these anions.149 The quadrupolar relaxation of the 1“N resonance provides an additional dimension of information about solution equilibria. Formation constants of several ionic complexes have been determined from the change in the line widths of quadrupolar nuclei (Cl).'5‘“54 In the case of SCN' in H20, Au-Yeng found that the 14N line width is sensitive to the ion- solvent interaction; the concentration dependence of the 1"N line width is consistent with the electron distortion model, which relates the distortion in the paramagnetic shielding contribution of the chemical shift to the relaxation rate of the resonance.34 In this paper, we address the structure of the primary solvation shell surrounding the thiocyanate and selenocyanide anions in nonaqueous protic solvents. The protic solvents used in this investigation are methanol (MeOH), formamide (FA), and N-methylformamide (NMF); this series allows a variety of hydrogen bonding interactions to be studied. NMR and infrared spectroscopies are employed to probe the interactions in these solutions and provide data from which the thermodynamics of the equilibria which exist in them can be calculated. Spectroscopic measurements are reported for the thiocyanate and selenocyanide anions both in neat solutions of the appropriate salt and protic solvents, and in an ‘inert’ solvent where the concentration of the hydrogen bonding solvent and salt can be varied. B. Results: 1. Infrared Measurements The vCN stretching vibration of the SCN' anion is sensitive to electrostatic interactions in solution. Hydrogen bonding broadens the absorption and shifts the peak frequency of this mode compared to solutions of SCN' in aprotic solvents, as shown in 136 Figure 44 and summarized in Table 33. The bandshape of the spectral envelope, especially for methanol solutions, is complex and results from several overlapping component bands. Spectral enhancement techniques such Fourier self deconvolution and curve fitting can be applied to these spectral envelopes to obtain the underlying band structure, shown in Figure 45. The peak positions and band widths are collected in Table 34. Closer inspection of the VCN mode for SCN' shows that this mode is asymmetric (Figure 46.). Previous investigations of SCN' have shown that this spectral envelope in aprotic solvents can be simulated by a Voight lineshape or Gaussian-Lorenztian sums.“I44 A single Voight lineshape does not simulate this bandshape well. A modest improvement is obtained when a Gaussian-Lorenztian sum is applied, yet the large residuals suggest that this spectral envelope is not correctly modeled. The gas phase infrared spectra of SCN' and OCN' include a v3 + v2 - v2 transition (bending hot band) that is 6.5 cm’1 and 11.5 cm‘1 red shifted from the fundamental transition.‘2°"55"5° In solution studies of OCN’, the v3 + v2 - v2 mode is observed 10-12 cm'1 red shifted from the fundamental, with an absorptivity consistent with that predicted by the Boltzman distribution for this mode.‘""“0 The asymmetry of the VCN mode along with the gas phase studies of SCN' imply that the hot band transition (v3 + v2 - v2) should be incorporated in the model for the VCN stretching mode. This model (Figure 46) correctly simulates the vCN spectral envelope in DMF and NM solutions where we have assumed that the absorptivity for the hot band transition is given by the Boltzman distribution. The vCN mode of SCN‘ in the protic solvents is much broader than in the aprotic solvents. In the amide solvents, NMF and FA, the spectral envelope was found to consist of the three bands with and additional band at higher energy found in the methanol solution, as shown in Table 34. The vCN stretch for 0.025 M SeCN' in aprotic solvents is also asymmetric and is well modeled by a incorporating a bending hot band 6.5 cm‘l below that of the fundamental. The vCS stretching mode of SCN' is also perturbed by protic solvents when compared to this vibration in the aprotic solvents: DMF and NM. As seen in earlier Table 33. Summary of position and band widths for the VCN stretch of ~0.25M NBu,SCN 137 in NM, DMF, NMF, FA and MeOH Solvent system Vm (cm”) Avm 1cm") NM 2059.5 13.5 DMF 2056.5 12.5 NMF 2057.8 33.8 FA 2059.5 27.5 MeOH 2060.0 40.5 Table 34. Summary of the deconvolved parameters for the component bands in ~0.25M NBu4SCN in NM, DMF, NMF, FA, and MeOH Solvent system VCN (cm") Avm 1cm" ) NM 2059.2 :1: 0.1 12.2 i 0.1 2052.0 i 0.1 12.4 :t 0.3 DMF 2056.7 i 0.1 11.8 i 0.1 2050.0 i 0.1 12.3 i 0.2 NMF 2048.0 i 0.5 25.4 :1: 0.2 2058.2 :1: 0.2 15.1 i 0.8 2067.5 i 0.5 20.4 :t 0.2 FA 2050.1 :t 0.8 23.0 :1: 1.2 2057.8 :1: 0.3 14.0 i 1.8 2065.3 i 1.0 20.8 :1: 1.0 M80" 2043.9 :1: 0.7 29.3 :1: 0.3 2057.5 i 0.1 20.2 i 2.0 2072.8 it 0.2 24.8 i 2.0 2090.9 i 0.6 15.8 :1: 1.0 Molar absorptivity (M"cm") 138 1000— 800“ DMF NM 600— NMF 1‘ 400— FA ;l\ 1 M H 200_ e0 0 ..r l l 1 l I l l l 2150 2125 2100 2075 2050 2025 2000 1975 1950 Wavenumbers (cm") Figure 44. VCN for~0.25M NBu4SCN in DMF, NM, NMF, FA, and MeOH 139 0.4 — 0.3 — ES V 0.2 — 8 t: .8 ‘3‘ .o < 0.1 — 0.0 a l I l I l l l l 2150 2125 2100 2075 2050 2025 2000 1975 1950 Wavenumbers (cm' I) Figure 45. Fourier self deconvolution of vCN for ~0.25M NBu4SCN in MeOH Molar absorptivity (M"cm") 140 1000 — v, fundamental 800 — 600 — — Exp. lineshape 400 ._ — — Symmetric lineshape 200 — v_,+v,-v2 \ Residuals 0 _J . .-.. //V , \r "' a I l l l l l l l 2 l 00 2090 2080 2070 2060 2050 2040 2030 2020 Wavenumbers (cm") Figure 46. VCN stretch for ~0.25M NBu4SCN in DMF 141 studies, this mode is more sensitive to electrostatic interaction than the vCN vibration. It is clear that new bands are formed at higher frequencies in both the NMF and MeOH solutions, despite the low sensitivity for this mode. A clearer picture of the formation of anion-solvent adducts can be obtained by adding either the protic solvent or the chalcocyanate anion to a dilute solution of the other in a weakly solvating (‘inert’) solvent, such as nitromethane (NM), for which the donor number is low. Infrared studies of the OH stretching modes for NM show that this solvent only weakly solvates anions.I40 The low donicity of this solvent also precludes it from strongly interacting with electropositive species. Thus, the solvation of the thiocyanate anion by methanol can be followed by measuring the 2100-2000 cm" spectral enve10pe of NBu4SCN in NM as a function of added MeOH. When no methanol is present, the vCN stretch of SCN' is a sharp absorption at 2059 cm". As MeOH is added, the intensity of the 2059 cm‘1 peak falls, with concomitant broadening of the spectral envelope but virtually no change in the peak frequency. There are two isobestic points, observed above and below the peak maximum, which imply the formation of two thiocyanate-methanol complexes (Figure 47). In order to address the underlying composition of the spectral envelopes, the lineshapes were Fourier self deconvolved (Figure 48). The deconvolution procedures reveals that the absorbance of the central vCN component decreases as the concentration of MeOH increases. Concurrently, two new bands grow, at 2070 cm‘1 and 2045 cm". Similar results are seen in the titration studies with NMF and FA. Quantitative information can be obtained by fitting the absorption envelopes to Gaussian-Lorenztian sum components, which yields the absorbance, bandwidth, and peak position of each constituent band. The absorption curves in the methanol series can be well simulated with four components: at 2051 cm‘1 for the hot band, at 2059 cm" for the free anion, at 2070 cm'1 for a methanol-thiocyanate complex, and 2047 cm'l for a second methanol-thiocyanate complex. The absorptivities of the fundamental constituents, plotted Absorbance (rel.) 142 1.4 — ' l l l T l ' l ' l ' l ' l ' l 2100 2090 2080 2070 2060 2050 2040 2030 2020 Wavenumbers (cm") Figure 47. VCN stretch for 0.1M NBu4SCN in NM with MeOH added 143 Absorbance (rel.) l l l 1 l l l l 2100 2090 2080 2070 2060 2050 2040 2030 2020 Wavenumbers (cm") Figure 48. Fourier deconvolution of vCN stretch for ~0.1M NBu4SCN in NM as a function of MeOH added 144 in Figure 49, are qualitatively similar to those obtained by the Fourier self deconvolution of the experimental spectral envelopes. Similar titration experiments were performed by adding N MF and FA to solutions of 0.15 M NBu4SCN in NM and are summarized in Table 35. Curve fitting and Fourier self decovolution of the 2100-2020 cm’1 spectral envelopes also show that the lineshape in each case is also a convolution of four bands. The absorption envelope for SeCN' can also be well simulated by four bands: at 2162 cm" for the hot band, at 2169 cm’1 for the free anion, at 2078 cm‘1 for the a methanol- selenocyanide complex, and at 2061 cm‘1 for a second methanol-thiocyanate complex (Table 36). Table 35. Summary of the peak positions and bandwidths (cm"), and molar absorptivities from the curve-fitted component bands of the vCN stretch of 0.15 M NBu4SCN with NMF, FA, or MeOH added Solvent system Vc~ V,,2 VCS VCS Peak Position F WHH Peak Position F WHH NMF 2050.0 1 1.0 17.4 1 0.8 739.5 1 0.6 11.4 :1: 0.8 2059. 2 1 0. 2 12.2 1 0.2 751.0 1 0.4 10.3 1 2.6 2068.8 1 1.4 13.2 1 2.2 FA 2050.711.0 15.81 2.8 7398104 12.8112 2059.2.101 11.8103 7514116 13.011.6 2067..2120 14.3 :1: 1.8 MeOH 2047.0 1 1.0 20.4 1 3.0 739.5 1 0.3 11.4 1 0.4 2059. 4 1 0.1 12.8 :1: 1.2 751.7 1 0.8 14.5 1 2.0 2070.5 1 0. 4 13.2 1 4.0 Hydrogen bonding interactions with these anions not only perturb the frequencies of the stretching vibrations, but also absorptivity of these modes. The integrated absorptivity for the vCN stretch increases by 30% in the methanol titration series. Likewise, the integrated intensity for the VCN stretch in the NMF and FA titration series each increase 145 by 13%. Experimental studies have shown that ionic and hydrogen bonding interactions to the nitrogen of SCN' increases the absorptivity the vCN stretch.4M They are supported by a theoretical study by our group which predicts that electrostatic interactions to the nitrogen of SCN' and SeCN' increase the molar absorptivity of the vCN mode in these anions.157 Table 36. Summary of the peak positions and bandwidths from the curve-fitted component bands of the VCN stretch of 0.025 M NBu4SeCN with MeOH added. Solvent system VCN v,,2 Peak Position F WHH (cm" (cm") MeOH 2060.7 1 1 6 7.93 1 1 80 2068.8 1 0 1 5.30 1 O 52 2077.5 1 0 4 7.69 1 O 84 The 770-720 cm’1 spectral region contains the vCS mode. In a 0.15 M NBu4SCN solution in NM, this band is centered at 740 cm". Addition of MeOH broadens the absorption and shifts the maximum of the envelope, with growth of a band at 752 cm", as shown in Figure 49. This spectral envelope can be simulated adequately by two Gaussian- Lorenztian sums. 2. Assignments The structure of the primary solvation sphere surrounding the thiocyanate anion in neat solutions of NBu4SCN in MeOH, FA and NMF can be understood by identifying the component bands of the two stretching modes. The concentration dependence of the infrared absorbances for the component bands in the titration studies (Figure 49) shows clearly that the SCN' anion is involved in two equilibria that have 1:1 stoichiometry. The most straight forward assignments are those for the unperturbed or ‘free’ SCN'; these 146 1.4 4 1.0 - 0.8 - Absorbance (rel.) 2059 cm’1 0-4 ‘ 2070 cm" 2047 cm‘2 0.2 — 0.0 -— 1 1 l 1 l l l 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Concentration of MeOI-I (M) Figure 49. Absorptivities for component bands of the VCN mode of ~0.1M NBu4SCN in NM as a function of MeOH added 147 bands occur at 2059 cm" for the ch mode and 740 cm‘1 for the vCS mode. Positive frequency shifts for the VCN mode, upon complexation, can be due to interactions along the molecular axis at the sulfur or nitrogen in SCN'.74 These two types of interactions can be easily distinguished by looking at the perturbations of the vCS mode. An electrostatic interaction to the nitrogen atom in SCN' will blue shift this mode whereas an electrostatic interaction to the sulfur will result in a red shift of the VCS mode. Clearly, there is no band lower in energy than that for the ‘free’ SCN', which implies that the 2070 cm’1 band must be due to a SCN""HOMe complex. The red shifted bands in the vCN mode of SCN' have previously been assigned to nonaxial interactions to the thiocyanate anion, where one or two solvent molecules are associated with the complex. The concentration dependence of the 2047 cm’1 band suggests that the stoichiometry of the complex is 1:1. Thus, the lower frequency band is attributed to a nonaxial interaction to the nitrogen of the thiocyanate anion. Theoretical calculations support these assignments. They predict that hydrogen bonding to the nitrogen atom by hydrogen fluoride or water will increase the frequency of the stretching modes when the interaction is along the molecular axis.I39 On the other hand, a nonaxial hydrogen bonded complex will be characterized by a blue-shift for the VCN stretch and a slight red-shift in the VCS stretch.158 Additional support for this unusual hydrogen bonded complex is derived from the crystal structure of Na+(12C4)2 SCN'-HOMe, where the C-N-O angle is 108°.47 The infrared spectrum of the vCN mode in this complex is 10 cm‘1 lower than that of the VCN stretch in Na*( 12C4)ZSCN’. Likewise, the same assignments can be applied to the component bands seen in the selenocyanide titration. The low frequency band can be attributed to a non-axial interaction to the nitrogen end of the anion and the high frequency component can be assigned to a component that has an axial interaction to the nitrogen end. The assignments for the titration experiments can be applied to the neat protic solutions. In the amide solutions, the same number of vibrational components are found. As in the titration experiments, a band above and below the component at 2058 cm", which 148 can be assigned to the ‘free’ SCN', are found. Again, this can be attributed to interaction alone the molecular axis of the anion and a nonaxial interaction. It should be noted that our experiments probably cannot differentiate between one and two nonlinear interactions since the bands are so strongly overlapped. Similar structure is observed in methanol solutions where the band at 2058 cm'I can be assigned to ‘free’ SCN', the band at 2044 cm‘1 to a non-axial interaction, and 2073 cm". The additional band at 2091 cm‘l is probably due to at MeOH"'SCN""HOMe complex since a red-shift in the VCS mode is not observed. 3. Thermodynamics In addition to multiple associations of the thiocyanate and selenocyanide anions with the hydrogen bonding solvents, both the self association of the hydrogen bonding solvent and ion pairing of the NBu4SCN and NBu4SeCN salts in nitromethane must be considered and possibly incorporated into the model for the solution equilibria. Previous investigations of the tetraalklyammonium salts in nitromethane have shown that ion pairing is negligible at 0.1M.48 Therefore, the NMR and infrared investigations have been modeled by three equilibria: (a) the self association of the hydrogen bonding solvent (limited to the dimerization) and (b) formation of two vibrationally distinct 1:1 thiocyanate- hydrogen bonding solvent adducts, as illustrated for MeOH in equations (22)-(24). Although these equilibrium expressions are written with the hydrogen bonding interaction at the nitrogen end of SCN', the determination of the equilibrium constants is independent of the site of the hydrogen bond. At the low concentrations of SCN' and SeCN' employed in this work, the equilibrium constants can be written in terms of the concentrations of the species in the equilibria; the activity coefficients can be neglected since a negatively charged species is on each side of the equilibrium. Thus the equilibria can be expressed by equations (25)-(27). SCN + MeOH :22 SCN"’HOMe K2 SCN' + MeOH = scrxt .HOMe KD 2MeOH == (MeOH)2 where [SCN""HOMe]1 [SCN‘MMeon]f [SCN'°“HOMe]2 [SCN']f[MeOH]f 2 = [(MeOH)2] [MeOH]f2 (22) (23) (24) (25) (26) (27) Equation (28) gives the total concentration of SCN‘ in terms of the concentration of the free SCN' anion, [SCN'],, the concentration of the first 1:1 complex, [SCN""HOMe],, and the concentration of the second 1:1 complex [SCN""HOMe]2. Likewise, the total concentration of MeOH, [MeOH]T, can be written in terms of the concentrations of free MeOH, [MeOH],, the two 1:1 complexes, and the methanol dimer, [(MeOH),], equation (29). [SCN']T = [SCN'], + [SCN""HOMe], + [SCN""HOMe]2 [MeOH]T = [MeOH]f + [SCN""HOMe]l + [SCN"”HOMe]2 + 2[(MeOH)2] (28) (29) 150 Substitution of equations (25)-(27) into equations (28) and (29) and solution of the resulting equations simultaneously for [SCN‘]f yields a cubic expression in [MeOH],, equation (30). The roots of equation (24) can be determined analytically, from which an expression for the concentration of free methanol can be derived in terms of the association constants, KI+K2, and KD. It should be noted that due that KI cannot be separated from K2, only the sum can be determined. An expression for [SCN‘]f can be written in terms of the association constants and the total MeOH and SCN' concentrations by substituting equation (25), equation (26), and the solution for equation (30) into equation (28) {2KD(K. + K2)}[MeOH],3 + (KI + K, + l)[MeOH]f2 + {(K,+K2)([SCN']T-[MeOH]T)+1}[MeOH]f - [MeOH]T = 0 (30) The observed chemical shift of either the 14N or the 77Se (for SeCN') resonance in these anions, 60b5, is a population average of those for the free anion and both 1:1 complexes as shown in equation (31), where the xfm is mole fraction and of,“ is the chemical shift of the free anion, x, is the mole fraction of the first 1:1 complex and 5, is its chemical shift, and x2 is the mole fraction of the second 1:1 complex, which has the chemical shift 82. With the appropriate substitutions, equation (31) can be rewritten (equation(32)) such that the observed chemical shift is a function of the association constants (K1, K2, and K0), the free SCN' concentration, the free MeOH concentration, and the limiting chemical shifts for the free SCN' and the 1:1 complexes. Following substitution of the expressions for [SCN‘]f and [MeOH],, the values for the equilibrium constants, including KD, can then be calculated by fitting equation (32) to the chemical shifts observed at different methanol concentrations. These values are summarized in Table 37 and illustrated by the MeOH titration in Figure 51. 151 0.30 -— 0.25 A 0.20 0.15 Re]. Absorbance 0.10 0.05 Wavenumbers (cm") Figure 50. vCS stretch for ~0. 1M NBu4SCN in NM as a function of MeOH added 152 -25.0“C -176 — D-12.5°C 00°C -174- .12.5°C 25.0°C -172— g 470— a. 5' c ".5 468*— 5” § g -m 0 «Z -164- -162- ‘lwl'l'l‘l'l'l‘ll'l 00 05 1.0 15 20 25 3.0 35 4.0 Concentration of MeOH (M) Figure 51. 14N NMR chemical shift for ~0.1M NBu4SCN in NM as a function of MeOH concentration and temperature ' 153 5 = Xfreesfree + X151 + X252 (31) obs (K16,+K282)[SCN']f[MeOH],/[SCN‘]T + 51... (32) obs The line width of the 14N resonances can also be used to obtain association constants for these adducts. Under the condition of fast exchange between the free (1) and the bound (b) sites, and in the absence of exchange broadening, the spin lattice relaxation time (lff 101,3), equation (33) is the population average of the relaxation times for the free (l/T”) and the bound species (l/T 1b)' For quadrupolar nuclei, such as 1“N, in asymmetric environments the longitudinal relaxation rate is very fast and is the dominant relaxation pathway.159 In such systems, the approximation l/Tlobs = 1/1‘ 20115 can be made, thereby relating the line width of the MN resonance to the association constants for the complex formation since the line width of an NMR resonance (Av) is proportional to the reciprocal of the transverse relaxation rate, equation (34). Again, the relationships for [SCN']f and [MeOH]f discussed previously can be substituted for the concentrations of free SCN' and MeOH. Figure 52 shows the 14N line width dependence on temperature and MeOH for SCN'. 1 = .711 + x_b z 1 (33) Tlobs Tl f T1 b T20bs A _ 1 _ xf xb _ [SCN']f[MeOH]f[ 1 1 ] 1 V1/2 - — - — + —— — — - — + — “T2 “T1 f KT] b “[MeOHlT T] b Tl f KT] f (34) ”N linewidth (Hz) for SCN‘ Figure 52. 154 500 a -25.o°c 400 ~ 300 7 425°C ‘ o.o°c 200 — , . 125°C v e 0 , 0 ° - o 25.0°c ‘ . O 100 — H l l l ‘ l ' l ' l l l ' 1 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Concentration of MeOH (M) l4N line width of NBu4SCN in NM as a function of MeOH and temperature Table 37. Association constants, Kl + K2 and KD for the formation of the two 155 1 :1 complexes and self association of the protic solvent, determined by l“N NMR measurements Solvent system "N chemical shift ”N chemical shift "N line width l“N line width "(111(2) K1) (K1119) K1) MeOH 250°C 0.83 1 0.01 0.10 1 0.01 0.84 1 0.03 0.10 1 0.01 12.5°C 1.05 1 0.01 0.11 1 0.01 0.95 1 0.05 0.10 1 0.01 0.0°C 1.37 1 0.01 0.13 1 0.01 1.34 1 0.05 0.14 1 0.01 —l2.5°C 1.70 1 0.02 0.15 1 0.01 1.64 1 0.09 0.15 1 0.01 -250°C 2.07 1 0.02 0.17 1 0.01 2.12 1 0.07 0.14 1 0.01 FA 250°C 0.76 1 0.01 0.07 1 0.01 0.74 1 0.02 0.07 1 0.01 125°C 0.90 1 0.01 0.07 1 0.01 0.92 1 0.05 0.08 1 0.01 0.0°C 1.05 1 0.01 0.09 1 0.01 1.01 1 0.04 0.08 1 0.01 —12.5°C 1.27 1 0.01 0.10 1 0.01 1.25 1 0.05 0.09 1 0.01 -250°C 1.46 1 0.02 0.11 1 0.01 1.43 1 0.05 0.10 1 0.01 NMF 250°C 0.71 1 0.01 0.05 1 0.01 0.70 1 0.04 0.04 1 0.01 125°C 0.85 1 0.01 0.06 1 0.01 0.85 1 0.05 0.05 1 0.01 0.0°C 0.94 1 0.01 0.07 1 0.01 0.94 1 0.06 0.06 1 0.01 -12.5°C 1.00 1 0.01 0.07 1 0.01 0.99 1 0.03 0.06 1 0.01 -250°C 1.18 1 0.03 0.09 1 0.01 1.22 1 0.09 0.07 1 0.01 Similarly, the 77Sc and 14N resonances in NBu4SeCN are sensititive to electrostatic interactions and the concentration dependences can be fit to the models described earlier. Applying these association models, the K1+K2 parameter from the 77Se chemical shift is 0.68 1 0.06 M'l with a value of 0.69 1 0.01 M‘1 from the 1“N resonance. As seen in the previous l“N experiments for SCN’, the line width of the nitrogen resonance for SeCN' increases as a function of MeOH added, yet the line width for the 77Sc resonance of SeCN' remains constant throughout the titration series. This observation precludes the line broadening in the 1“N resonances SeCN' from being a result of exchange broadening. The formation constants for the hydrogen bonded complexes can also be determined in a number of ways from concentration dependence of the infrared band 156 absorptivities. The absorptivity of the 2059 cm‘1 component can be related to [SCN']f and [MeOH]f as shown in equation (23), where 82059, is the molar absorptivity and I? is the pathlength of the infrared cell. Upon substitution for [MeOH],, the absorbance of the 2059 cm" component becomes a function of the molar absorptivity, pathlength, equilibrium constants, and the total concentrations of SCN‘ and MeOH. The fitting of equation (35) to the absorbance of the v2059 band provides the formation constants of the solvated complexes and the self association constants for the solvated complexes. Likewise, the absorbances for the blue and red shifted component band in the methanol titration can also be related to the total SCN' and MeOH concentrations, shown in equation (36). Nonlinear regression of the band absorptivities to equation (36) yields the sum of the association constants (K,+K2), the self association constant for MeOH (KD), and the product of the molar absorptivity for that band and association constant for the complex (3207019)- Likewise, a similar relationship can be derived for the red shifted component, described in equation (37). The integrated intensity 82059€[SCN']T (35) 1+ (K, + K2)[MeOH]f A2059 = 320593150811 = 82070€K1[SCN-]T[MCOH]f 1 + (K, + K2)[MeOH]f A2070 3' €2070€[SCN-WHOMC]I = €2070£K1[SCN-]f[MCOH]f = (36) A _ €2047€K2[SCN-]T[MCOH]f 20‘” 1 + (K, +K2)[MeOH]f (37) 157 of the vCN mode provides fourth method of obtaining the association constants, since the molar absorptivity of the complexes are larger than that for the unperturbed anion. Here, the integrated absorptivity (fl) is the sum of all the component band absorbances, equation (38). Equation (38) can be rewritten so that the integrated absorbance is a function of the total concentrations for SCN' and MeOH by substituting the relationship for [MeOH],. It should be noted that the inclusion of the second solvate cannot be distinguished from only one 1:1 equilibria. The parameter obtained by fitting this model to the experimental data provides additional confidence that the vCN spectral envelope was correctly simulated since the model for the integrated absorptivity makes few assumptions about the equilibria involved in the vcN spectral envelope. If the vCN mode was simulated with only two bands Gaussian/Lorenztian bands, the association constant for this complex is substantially smaller than the results shown here. These results are collected in Table 38. fl = A... + 21, + A =efm€[SCN'], + e,€[SCN'"'HOMe], + e,€[SCN""HOMe]2 (38) _ ascmT A _ 1 + (K1 +K2)[MeOH]f(£°°° + (K,e, +K282)[MeOH],) (39) The vCS mode, in the titration series, was found to be a composite of two bands. The concentration dependence of these component bands can fit to several models. The model that best fits the experimental data and is consistent with previous results assumes that the 740 cm'1 is a composite band with contributions from the ‘free’ SCN' and one of the solvated complexes and the 751 cm" is a sum of a NBu,+ absorption and the other solvate complex. The absorbance for the 740 cm’l can be related to the total concentrations 158 of SCN' and MeOH, equation (40). The associations constants from this mode are in excellent agreement with those obtained for the VCN mode, Table 39. The infrared and N MR experiments only yield a sum of association constants (K,+K,) since molar absorptivities and chemical shifts for the specific complexes are convolved with the association constants in the equilibrium model. Theoretical calculations of a number of SCN'/HOH and HF adduct have been computed by our group, which provide predictions for the changes in absorptivity upon hydrogen bond formation.'39 It turns out, quite fortuitously, that the actual values for the molar absorptivities do not need to be known, as shown in equation (41). The ratio of the two adduct bands, A2070 and A204,, for the methanol series, can be related to the association constants for these complexes multiplied by their respective molar absorptivities. Likewise, a similar relationship can be derived for the vCS stretch. The individual association constants, using this assumption are collected in Table 40. 31308-11 A740 = EfrcchCN-]f + EZZISCN-WHOMC]2 = I + (K +K )[MCOH] l 2 T (Efrec + 82K2UVICOI-Hf ) (40) A2070 = 82070[SCI‘I_"'I'I()IVIC]1 = 82070K1 A2047 82047[SCN-WHOMC]2 82047K2 (41) Even though K, and K2 cannot be separated in the NMR experiment, it was found that the temperature dependence of these association constants fit the usual Gibb’s free energy relationship. The computed enthalpy and entropy values from these association constants are a convoluted from the two equilibria that they represent. Despite the difficulty in interpreting the meaning and significance of these parameters, they are fairly similar to results that we observed for complexes of the cyanate anion with these protic solvents.I37 Table 38. Association constants (K, + K2, and KB) of SCN’ with MeOH, FA, and NMF from the stretching modes of SCN' Solvent system A2053 Ablue shift Aredshift Integrated A740 absorptivity MeOH K,~1-K2 0.86 1 0.02 0.86 1 0.01 0.86 1 0.01 0.83 1 0.01 0.82 1 0.04 KD 0.09 1 0.01 0.10 1 0.01 0.10 1 0.01 0.09 1 0.01 0.10 1 0.01 FA K,+K2 0.71 1 0.01 0.70 1 0.01 0.72 1 0.01 0.72 1 0.02 0.68 1 0.09 KB 0.06 1 0.01 0.06 1 0.01 0.06 1 0.01 0.07 1 0.01 0.06 1 0.01 N MF K,+K2 0.67 1 0.01 0.62 1 0.01 0.71 1 0.01 0.71 1 0.02 0.67 1 0.06 KD 0.06 1 0.01 0.05 1 0.01 0.05 1 0.01 0.05 1 0.01 0.05 1 0.01 Table 39. Individual association constants using the extinction coefficients from ab initio calculations. VCN Vcs Vcs Solvent system K, (M") K2 (M") K2 (M”) K2 (M") MeOH 0.46 1 0.04 0.48 1 0.08 0.40 1 0.04 0.38 1 0.07 0.31 1 0.02 0.30 1 0.02 0.40 1 0.02 0.41 1 0.02 NMF 0.27 1 0.02 0.16 1 0.10 0.40 1 0.02 0.51 1 0.30 160 Table 40. Thermodynamic parameters for the association of SCN' with MeOH, FA and NMF from the NMR measurements "N chemical shift "N line width Solvent system AH“; (kJ/mol) ASH," (Jmol"K") AHM" (kJ/mol) AHM" (Jmol"K") MeOH-..114107 -4013-..118108 -4113 FA “82103 -3011 “80106 -2912 NMF -5.8107 -2212 -6.2108 2413 Here, the AS,+2 values are larger than those for the cyanate association and the enthalpy values for the thiocyanate complexes are smaller than those for the cyanate complexes. C. Conclusions The interactions of the thiocyanate and selenocyanide anions are weak. In fact, the strength of these interaction are an order of magnitude weaker than our previous investigation of the solvation of the cyanate anion by a number of protic solvents.137 Needless to say, the study of such weak complex is difficult. Here, we have studied anion-solvent complexes by a number of infrared and NMR techniques in order to be confident that the correct structure of the primary solvation sphere has been obtain. The concentration dependence of the integrated intensity for the VCN mode, the I4N chemical shift and the 14N line width all fit to a 1:1 equilibrium model. It should be noted that these models do not discriminate between two 1:1 equilibria and a single 1:1 equilibrium. The models that we applied to the infrared spectral envelopes are not without ambiguity. For example, the VCN mode in the titration experiments could have been modeled by only three bands (the hot band, the ‘free’ SCN‘ and a single SCN‘/MeOH complex) instead of four bands, but the fitted parameters from this procedure are not self consistent. The peak position and band widths of the component bands are not constant. Additionally, this spectral envelope may be fit with more than four components as seen by Gill and co- workers.43 Although four bands are more than adequate to simulate this spectral envelope, 161 due to the small residuals, a large number of bands could be convolved in this lineshape. The vCS stretch, which is very sensitive to electrostatic interaction, is very diagnostic for the number of vibrational components. This mode precludes a large number of vibrational components. Without applying a two 1:1 equilibrium model to deconvolve the stretching modes, the association constants would not be in agreement with those obtained from the integrated absorbance, the l“N chemical shift and the l“N line widths. Therefore, we believe that we have correctly determined the solution equilibria for our titration experiments by the harmony found between the various models. The SCN' and SeCN' anions are much less solvated than OCN‘. Whereas the OCN' anion has 1.5 to 2 solvent molecules associated with it in its ion pair, a significant proportion of the SCN' and SeCN' anions are not hydrogen bonded in neat solution of MeOH, FA and NMF. The titration studies show that the enthalpy is smaller and the entropy is larger for SCN' and SeCN' as compared to OCN'. Gas phase ab initio studies of these anions show that the cyanate anion has a more negative electrostatic potential off the nitrogen atom than SCN' and SeCN‘.4° The smaller electrostatic potential for SCN' and SeCN' results in a smaller hydrogen bond enthalpy formation. Even though the charge of the OCN', SCN' and SeCN' anion is -1, the distribution of this charge is also important in the how these anions order or structure the solvent molecules surrounding them. The smaller electrostatic potentials for SCN' and SeCN' do not order the nitromethane molecules surrounding them as well as OCN'. Thus upon formation of a hydrogen with SCN' or SeCN', fewer solvent molecules will be released thereby resulting in a more negative entropy than for OCN'. Chapter VI Crystal Structures A. Introduction The environment of a cation or anion in solution depends on many factors. A variety of experimental techniques have been used to determine or infer the structures surrounding anions and cations in electrolyte solutions. Of these, infrared spectroscopy has been widely applied to the study of solution equilibria, and has been used to identify such species such as ion pairs, dimers and tetramers.85 Ionic and hydrogen bonding interactions with the thiocyanate anion perturb its vibrational modes. The perturbation of the stretching modes, vCN and vcs, are unique for a given structure. For example, ionic interactions to the nitrogen end of the thiocyanate anion increase the frequencies of both the stretching modes, whereas an interaction to the sulfur end of the anion raises the frequency of the vCN mode and lowers the frequency of the vCS mode, with respect to the unperturbed anion.”144 Negative frequency shifts for the vCN vibration, relative to ‘free’ SCN‘, are the result of two electrostatic interactions to the It cloud of the C-N bond, as illustrated by dimers and tetramers.85 Positive vCN frequency shifts for SCN""HOMe complexes have been attributed to a hydrogen bond along the molecular axis of the anion and negative frequency shifts have been assigned to nonaxial interactions at the C-N bond of the anion.43 Our group has also attributed a red-shifted vCN vibration, as compared that of the unperturbed anion, to a single nonaxial interaction at the nitrogen of SCN'. Such an unusual interaction was first identified by Gill and coworkers in neat methanol solutions of tetrabutylammonium thiocyanate, where they attributed red-shifted bands to one and two hydrogen bonding interactions to the C-N bond.43 Infrared studies of solutions of alkali metal thiocyanates in NMF have shown that the cation and anion are not ion paired.160 Moreover, deconvolution of the vcN stretch for the thiocyanate anion shows that it is in three vibrationally distinct environments. Solid solvates of NaSCN with DMF and NMF were prepared in order to provide a pictorial understanding of potential solution structures and to determine what effect ionic and 163 164 hydrogen bonding interactions have on the vibrational modes of SCN'. Crown ethers bind to alkali metal cations. Solution studies have shown that crown ethers reduce ion pairing in electrolyte solutions by isolating the cation from the anion.3°'55'5°"°' Likewise, these ligands can be employed to minimize electrostatic interactions in crystal structures. B. Results 1. Crystal structures for NaNCSXDMF), and NaNCS-(NMF), The crystal structures for both bis(N,N-dimethylformamide)-sodium thiocyanate and bis(N-methylformamide)-sodium thiocyanate reveal linear chains of sodium cations bridged by the amide solvents and thiocyanate anions, as shown in Figures 53-56. The positional parameters for each structure are collected in Tables 41-42. Each sodium cation is coordinated by four amide molecules and two thiocyanate anions in a distorted octahedral '62 The amide solvents are coordinated orientation, similar to that seen for N aSCN-2H20. to the sodium cations through the carbonyl oxygen with Na*'"O distances of 2.39-2.43 A and 2.46-2.48 A for the NMF and DMF solvates, respectively. The intermolecular bond lengths are collected in Tables 43 and 44. The thiocyanate anions are coordinated through the nitrogen end of the anion with Na*"'N distances of 2.50-2.53 A and 2525-2526 A for the NMF and DMF solvates, respectively. In the N MF solvate the thiocyanate anions are in a trans orientation whereas the thiocyanate anions are in a cis configuration around the sodium cation in the DMF solvate. The bond angles for bridging ligands (Na“'"L"'Na*) are 80—90°. The bond angles between across a sodium cation for trans ligands (L,"'Na*"'L2) are about 160° for those in the DMF solvate and 170° for the NMF solvate. The greater distortion from octahedral geometry in the DMF solvate is reflected in the longer internuclear distance (Na“'"Na") of 3.335 1 0.005 A versus 3.235 1 0.003 A for the NMF solvate. 165 J A K 0 $05 N13 (2,4? / l 0.21 C22 0 Cfi>~w $81" 0? 04,9 M C240 C12 K Na] P394): (2 {Cb—(B .\ xii 25$ I. O D. Figure 53. Ortep structure for NaNCS'(DMF)2 166 I 1 I 1 _’ ‘ ' O y. B o . B o o . o o . e 0‘. o \. o .\,. e .‘ o .’ e . . ‘7 .. .‘o ../,l - O ‘ . . .ql. 0 Q . o. . e . . e .1 W .I' 0 O o“ < . C O T “‘0’ 6 0 0 o O . ‘0 O. . O ‘. 9 i .0 0 ’ 0A 0 ‘ Figure 54. Packing diagram for NaNCS-(DMF)2 167 1-. \V) V 1 .\ 0 v1 7 = &’ \ Hi2 0’9 3 03013/ H13 N1 14b \_ H23 1 e— 0 . 021 H12 m 0“ c1 \‘l N23 H14c N, \ ){1 C24 ‘. Na] H24d 112511 H246 I - \U 1" \:) Figure 55. Ortep plot for NaNCS-(NMF)2 168 Figure 56. Packing diagram for NaNCS'(NMF)2 169 Table 41. Positional parameters (A) and equivalent isotropic thermal parameters (A2), with estimated standard deviations in parentheses, for bis(N-methylformamide)—sodium thiocyanate atom x y z 3,,I S1 0.3166(1) 0.0840(1) 0.55631(3) 395(4) Nal 0.0024(1) -0. 1370(1) 0.75548(5) 225(4) 01 1 -0.2287(2) 0.1067(3) 0.731 1(1) 300(9) 021 0.1020(3) 0.1267(3) 0.83077(8) 308(9) N1 0.1589(3) 0.1030(4) 0.6784(1) 2.9( 1) N13 -O.4938(3) 0.1019(4) 0.7774(1) 3. 1( 1) N23 0.1678(3) 0.0798(5) 0.9398(1) 3.5(1) C1 0.2209(3) 0.0963(4) 0.6270(1) 2.4(1) C12 -0.3910(4) 0.1052(4) 0.7272(2) 2.9(1) C 14 -0.4270(5) 0. 1005(6) 0.8456(2) 4.2(2) C22 0.1273(4) 0.1929(5) 0.8874(1) 4.2(2) C24 0. 1904(6) -0. 1427(6) 0.9382(2) 4.2(2) H12 -0.455(4) 0.108(4) 0.684(1) 3.8(7) H 13 -0.603(4) 0.104(5) 0.771(2) 4.7(8) H 14a -0.479(5) 0.218(7) 0.870(2) 7(1) H14b -O.307(5) 0.107(5) 0.850(2) 5.7(9) H 14c -O.462(5) -0.019(7) 0.868(2) 7(1) H22 0.121(4) 0.342(5) 0.897(1) 3.5(7) H23 0.188(4) 0.140(5) 0.976(2) 4.0(7) H24a 0.l7(1) -0.19(l) 0.892(3) 3(1) H24b 0.12(1) -0.22( 1) 0.971(4) 6( 1) H24c 0.31(1) -0.17(1) 0.949(4) 5(1) H24d 0.08( 1) ~0.20( 1) 0.927(5) 5(1) H24c 0.23(1) -0.18( 1) 0.983(4) 3(1) H24f 029(1) -0. 18(1) 0.905(5) 6(1) 170 Table 42. Positional parameters (A) and equivalent isotropic temperature factors (A2), with estimated standard deviations in parentheses, for bis(N,N-dimethylfonnamide)-sodium thiocyanate atom x y z B(eq) S 1 0.8444(2) 0.0823(1) 1/4 3.9( 1) Na] 0.5073(3) 1/4 1/2 3.2(1) .01 1 0.3224(5) 0.2761(3) 3/4 3.9(3) O21 0.5478(5) 0.151 1(3) 3/4 3.2(2) N 1 0.6536(7) 0.1789(4) 1/4 4.5(4) N13 0.1279(5) 0.2049(3) 3/4 2.2(1) N23 0.6559(6) 0.0407(3) 0.7200 2.4(1) C 1 3/4 0.1387(4) 1/4 3.4(4) C12 0.1947(7) 0.2690(3) 3/4 2.4( 1) C 14 0.198(1) 0.1327(5) 3/4 4.0(2) C15 -0.0229(9) 0.2021(5) 3/4 3.6(2) C22 0.6429(8) 0. 1059(4) 0.807(2) 2.6(2) C24 0.772(1) -0.0093(5) 3/4 3.9(2) C25 0.558(1) 0.0139(7) 0.568(2) 3.5(2) H12 0.135(5) 0.313(3) 3/4 1(1) H14d 0.274(8) 0. 136(4) 3/4 3(2) H 14c 0.170(8) 0.1 1 1(4) 0.60(2) 14(2) H15d -0.048(8) 0.256(5) 3/4 6(2) H15e -0.048(5) 0.172(3) 0.639(9) 5(1) H22 0.720(8) 0.122(4) 089(1) 2(2) H24a 0.729(9) -0.062(5) 3/4 5(2) H24b 0.835(6) 0.005(3) 0.66( 1) 7(2) H25d 0.62(l) 0.000(5) 0.46(2) 4(2) H25e 0.55( 1) -0.038(7) 0.61 (2) 6(2) H25f 0.50(1) 0.041(5) 053(2) 2(2) 171 Table 43. Intermolecular distances (A), with estimated standard deviations in parentheses, for bis(N-methylformamide)—sodium thiocyanate atom atom distance atom atom distance 8 1 C 1 1.637(3) N23 C22 1.312(4) Nal Nal 3.235(3) N23 C24 1.447(5) Nal 01 1 2.392(2) N23 H23 084(3) Na] 01 1 2.392(2) C 12 H12 097(3) Nal 021 2.386(2) C14 H14a 1.00(4) Nal 021 2.425(2) C 14 H 14b 092(4) Nal N1 2.534(3) C 14 H14c 094(3) Nal N 1 2.502(3) C22 H22 098(3) 01 1 C12 1.232(3) C24 H24a 099(6) 021 C22 1.232(3) C24 H24b 1.00(7) N1 C 1 1.162(3) C24 H24c 094(7) N13 C 12 1.311(4) C24 H24d 095(8) N13 C14 1.448(5) C24 H24c 097(7) N 13 H13 083(3) C24 H24f l .05 (8) Table 44. Intermolecular distances (A), with estimated standard deviations in parentheses, for bis(N,N—dimethylformamide)-sodium thiocyanate atom atom distance atom atom distance S1 C1 1.640(5) C 12 H12 098(5) Nal Nal 3.335(3) C14 H14d 074(8) Na] 0]] 2.483(4) C14 H14c 1.1(1) Nal O21 2.462(4) C15 H15d 099(9) Nal N 1 2.526(5) C 15 H 15e 095(5) 01 1 C12 1.236(8) C22 H22 095(8) 021 C22 1.279(9) C24 H24a 1.02(8) N1 C 1 1.174(8) C24 H24b 088(6) N13 C12 1.314(7) C24 H25d 1.0(1) N13 C14 l.45(1) C24 H25e 1.0(1) N23 C22 1.308(9) C24 H25f 0.8( 1) N23 C24 1.45(1) N 23 C25 1.46(1) a. Amide Structure 172 In both solvates, one amide molecule is perpendicular to the N a“"'N a” axis. In the case of the DMF solvate, this DMF molecule is fixed perpendicular to the Na"'“Na+ axis in a mirror plane. The second DMF molecule is distorted out of the mirror plane. No such conditions are exists in the NMF solvate, yet one N MF molecule is roughly perpendicular and the other roughly parallel to the Na"'"Na+ axis. The carbonyl bond lengths, 1.232 (3) A in the NMF solvate and 1.238 (8) A and 1.279 (9) A for the DMF solvate, is slightly longer in the DMF solvate. The bond distances between the carbonyl carbon and the nitrogen are in much closer agreement of about 1.31 A. The alkyl carbon-nitrogen bond lengths are ~1.45 A, and the oxygen-carbon-nitrogen bond angles are ~125°. Each of the DMF and NMF molecules, in these solvates, are nearly planar. b. Thiocyanate The thiocyanate anion displays small differences between these two solvates. The anion is linear in the DMF solvate (ZSCN' 179.8 1 06°) and slightly distorted in the NMF solvate (ASCN‘ 177.5 1 02°). The C-N bond is slightly longer in the DMF solvate ( 1.174 1 0.008 A) than in the NMF solvate (1.162 1 0.003 A). There is no substantial difference between the C-S bond lengths for the NMF solvate (1.637 1 0.003 A) and the DMF solvate (1.640 1 0.005 A). c. Hydrogen bonding Owing to the acidic proton in NMF, a hydrogen bonding interaction with this solvate is possible. A search of intermolecular distances out to 3.6 A reveals a potential hydrogen bond to the sulfur end of SCN' from NMF. The hydrogen bond, S l"'H23, is 2.58 1 0.03 A lOng with a heavy atom bond (Sl'"N23) distance of 3.364 1 0.003 A. This interaction is distorted from linearity and has a bond angle of 157.5°, shown in Figure 54. Interestingly, this interaction links the sodium chains together. Such an interaction is not observed in the DMF solvate, nor would it be expected due to the lack of an acidic 173 hydrogen. This type of non-axial interaction to the sulfur end is observed for hydrates of various alkali metal thiocyanates. 2. Summary of NaSCN 0( DMF )2 and NaSCNO(NMF)2 crystal structures As with other solvates, the neutral solvent molecules form part of the coordination sphere surrounding the cation. Not surprisingly, the thiocyanate anion coordinates to the sodium cation through the nitrogen atom. The nonaxial cation interactions, in these solvates, lengthen the C-N bond and shorten the GS compared to SCN' in crystals having no electrostatic perturbation. These structures may represent the structure of solvated N aN CS dimers in NMF and DMF at extremely high electrolyte concentrations, where NaNCS will dimerize. 3. Crystal structures for Na(12C4)2NCSOH0Me and Na(12C4),NCS Previous investigations of 12—crown-4 have shown that the cavity is too small for Na” cations, yet solution and crystal structure studies have shown that this ligand can easily bind to Na“ with 1:1 and 1:2 (cationzligand) stiochiometry.'°3"°5 For the Na(l2C4),SCN and the Na( 12C4)4SCN0HOMe crystals grown in this work, each the sodium cation is surrounded by two 12-crown-4 ligands (Figures 57-60). The positional parameters for each structure are collected in Tables 45 and 46. In each case, the thiocyanate anion is not within van der Waal contact of the cation. The Na-O distances between the ligands and the Na+ cation are comparable, with an average value of 2.50 (1) A in the solvated complex and 2.50 (3) A in the Na( 12C4),SCN complex (Tables 47 and 48). The ligands surrounding the Na+ cation in the solvate complex are disordered shown in Figure 59. There are two conformations for the ligands, that are at 50% occupancy which are ~45° out of phase from each other. The l2-crown-4 ligands in the Na(12C4),,SCN complex are not disordered, 174 Figure S7. Ortep plot for Na(12C4)2NCS 175 Figure 58. Packing diagram for Na(12C)4NCS 176 Figure 59. Ortep plot for Na(l2C4)2NCS-HOMe 177 Figure 60. Packing diagram for Na(12C4)2NCS°HOMe Table 45. Atomic coordinates (10‘) and equivalent isotropic displacement parameters (A2 x 10“) 178 for the Na(12C4)2SCN complex. A tom x y z U(eq)‘ Occ. 8(1) 3486(3) 1228(3) 4913(2) 88(1) 1 8(2) 6448(3) 5939(3) 136(2) 97(2) 1 Na( 1) 8768(3) 6620(3) 2825(2) 49(1) 1 Na(2) 1274(3) 1620(3) 2186(2) 43(1) 1 0(101) 8139(5) 7140(6) 3889(4) 62(2) 1 0(104) 8639(5) 5327(5) 3708(4) 59(2) 1 0(107) 10178(5) 5968(5) 3256(4) 52(2) 1 0(1 10) 9665(5) 7815(5) 3439(4) 52(2) 1 0(201) 8778(5) 5373(5) 1927(4) 49(2) 1 0(204) 7303(4) 6157(5) 2425(3) 48(2) 1 0(207) 7943(5) 7970(5) 2252(3) 52(2) 1 0(210) 9410(5) 7164(5) 1771(4) 52(2) 1 0(301) 1842(5) 1 1 15(6) 1058(4) 56(2) 1 0(304) 1355(5) 2940(5) 1374(3) 51(2) 1 (X307) ~16 1 (5) 2283(5) 1873(4) 53(2) 1 (X310) 330(5) 461(5) 1566(3) 50(2) 1 0(401) 1262(5) 2788(5) 3157(4) 51(2) 1 0(404) 2732(4) 2083(5) 2596(3) 46(2) 1 (X407) 21 18(5) 272(5) 2669(3) 48(2) 1 0(410) 640(4) 980(5) 3212(3) 48(2) 1 N( 1) 4040(8) 2986(9) 4618(6) 88(4) 1 N(2) 6195(9) 78 1 8( 10) 378(6) 91(4) 1 C( 1) 3791(8) 2257(9) 4759(6) 54(3) 1 C(2) 6300(8) 7046(12) 269(6) 67(4) 1 C(102) 8145(12) 6521(12) 4454(8) 73(5) 1 C(103) 7997( 10) 5607(12) 4165(8) 72(4) 1 C(105) 9419( 10) 4995(12) 4021(8) 70(4) 1 C(106) 1009100) 5061(11) 3513(9) 67(4) 1 C(108) 10678(9) 6570(9) 3727(7) 57(4) 1 C(109) 10550(8) 7529( 10) 3465(8) 59(4) 1 C(1 1 1) 9404(9) 81 1 1(1 1) 4090(7) 67(4) 1 C(1 12) 8455(9) 8077(10) 4044(6) 66(4) 1 C(202) 7967(8) 5136(1 1) 1622(8) 57(4) 1 C(203) 7329( 10) 5248( 10) 2147(8) 61(4) 1 C(205) 6841(9) 6805(8) 1998(7) 48(3) 1 C(206) 7043(8) 7757(9) 2245(6) 50(3) 1 C(208) 8208(9) 8222(9) 1593(6) 57(4) 1 C(209) 9174( 10) 8107(10) 1643(8) 60(4) 1 C(2“) 9319(12) 6562(10) 1196(8) 63(4) 1 C(212) 9391(9) 5588(10) 1435(8) 57(4) 1 C(302) 1778(10) 1809(10) 541(7) 63(4) 1 C(303) 1936( 10) 2732(1 1) 864(8) 62(4) 1 C(305) 559(9) 3273(1 1) 1097(8) 63(4) 1 C(306) -69(9) 3205(9) 1647(7) 58(4) 1 C(308) -695(9) 1714(10) 1415(7) 54(4) 1 C(309) ~527(8) 749(10) 1621(7) 57(4) 1 C(31 1 ) 525(9) 245(9) 874(6) 51(3) 1 C(312) 1476( 10) 254( 10) 854(7) 63(4) 1 C(402) 2090(8) 2968(12) 3498(7) 62(4) 1 C(403) 2720( 10) 293 1 (10) 2959(7) 63(4) 1 C(405) 3240( 10) 1 385( 10) 2964(8) 62(4) 1 C(406) 2991(9) 464(10) 2663(7) 60(4) 1 C(408) 1861(9) -42(1 1) 3325(7) 58(4) 1 C(409) 923( 10) 56( 10) 3326(8) 57(4) 1 C(4“) 691(10) 1541(10) 3820(7) 60(4) 1 C(412) 2537(9) 3608(7) 642(9) 54(3) 1 'U(eq) is defined as one third of the trace of the orthogonalized U,,- tensor 179 Table 46. Atomic coordinates (10‘) and equivalent isotropic displacement parameters (A2 x 103), with estimated standard deviations in parenthesis, for the Na(12C4)2SCN0MeOH complex. x y z U(eq) Occ. Na( 1) 0 5000 5000 57(1) 1 8(2) 1283(3) 2500 3003(4) 120(1) 1 C(3) 3077(12) 2500 2863(12) 92(2) 1 N(4) 4409(12) 2500 2750(14) 140(3) 1 GOOD -7624(7) 5405(2) 299(6) 68(2) 0.50 C(102) 721 1 (10) 5428(5) -2292(10) 84(4) 0.50 C(103) 5459(11) 5741(5) -3487(14) 78(6) 0.50 O(104) 4070(6) 5395(2) 3324(7) -64(2) 0.50 C(105) 2400(8) 5723(4) 4004(9) -76(3) 0.50 C(106) 2418(12) 621 1(3) -2596(1 1) 75(4) 0.50 O(107) 2802(6) 5906(2) -823(6) 65(2) 0.50 C(108) 3349(9) 6336(4) 790(1 1) 86(3) 0.50 C(109) 5274(1 1) 6500(3) 1633(14) 76(4) 0.50 O( 1 10) 6324(6) 5926(2) 2168(7) 66(2) 0.50 C(1 l l) 8145(8) 5987(4) 2539(9) 65(3) 0.50 C( 1 12) 8305(13) 5993(4) 728(13) 86(5) 0.50 O(201) 6269(6) 5325(2) -2246(6) 57(1) 0.50 C(202) 5159( 10) 5783(4) -3661(12) 56(4) 0.50 C(203) 3254(9) 5631(4) -4295(8) 67(3) 0.50 O(204) 2821(7 5659(2) -2725(7) 63(2) 0.50 C(205) 2566( 1 l) 6297(3) -221 2(10) 64(4) 0.50 C(206) 2692(8) 6286(4) -258(10) 76(3) 0.50 O(207) 4403(6) 6060(2) 1 168(6) 68(2) 0.50 C(208) 581 1(10) 6526(3) 1821 ( 12) 68(3) 0.50 C(209) 7545(9) 6194(4) 2855(8) 66(2) 0.50 O(210) 7797(7) 5720(2) 1652(7) 63(2) 0.50 C(21 l) 8378(1 1) 5994(3) 364(1 1) 60(4) 0.50 C(212) 8102(8) 5523(4) -1 179(1 1) 70(3) 0.50 O(301) 7563(9) 2867(5) 6386(9) 185(5) 0.50 C(302) 8853(14) 2500 6936(14) 136(3) 1 U(eq) is defined as one third of the trace of the orthogonalized Uij tensor 180 Figure 47. Selected bond lengths, with estimated standard deviations in parenthesis, for the Na(12C4)2SCN complex Atoms Bond length N(1)-C(l) 1.163(14) N(2)-C(2) 1 . 15(2) S(l)—C(l) 1.598(14) S(2)-C(2) 1 .64(2) Na( 1 )-O(204) 2.469(8) Na(1)-O(1 10) 2.476(8) Na(1)-O(101) 2.479(9) Na(1)-O(210) 2.484(8) Na(1)-O(107) 2.496(8) Na(l )-O(201) 2.51 1(8) Na(1)-O(207) 2.551(8) Na(1)-O(104) 2.556(8) Na(2)-O(404) 2.463(8) Na(2)-O(410) 2.480(8) Na(2)-O( 307) 2.486(8) Na(2)—O(304) 2.486(8) N a(2)-O(3 10) 2.487(8) Na(2)-O(407) 2.499(8) Na(2)-O(401) 2.533(8) Na(2)-O(301) 2.541(8) O(101)-C(102) l.42(2) O(101)-C(112) 1.466(14) O(104)-C(105) 1.41(2) O(104)-C(103) l.45(2) O(107)-C(106) 1.413(14) O(107)—C(108) 1.451(13) O(110)—C(111) 1.428( 13) O(110)-C(109) 1.451(13) O(201)-C(202) 1.408(13) O(201)-C(212) 1.445(14) O(204)-C(205) 1.414(12) O(204)-C(203) 1.421(14) O(207)-C(208) 1.428(13) O(207)-C(206) 1 .449(13) O(210)-C(21 1) l.42(2) O(210)-C(209) 1.429(14) O(301)-C(312) 1.414(14) O(301)-C(302) 1 .419(14) O(304)-C(305) 1.410(13) O(304)-C(303) l.43(2) O(307)-C(306) 1.414(14) O(307)-C(308) 1.435(13) O(310)-C(309) 1.424(13) O(310)-C(3l 1) 1.439(13) O(401)-C(4l2) 1.409(13) O(401 )-C(402) 1 .440(14) O(404)-C(403) 1 .414(14) O(404)-C(405) 1 .443( 14) O(407)-C(406) 1 .404( 13) O(407)-C(408) 1 .442( 13) O(410)-C(409) 1 .420( 14) O(410)-C(41 l) 1.430(14) 181 Figure 48. Selected bond lengths, with estimated standard deviations in parenthesis, for the Na( 12C4)2SCN complex Bond Length (4) Na(1)-O(1 10)#1 2.460(5) Na(1)—O(1 10) 2.460(5) Na(1)-O(101) 2.475(5) Na(l)—O(101)#1 2.475(5) Na(1)-O(104)#l 2.483(5) Na(1)-O(104) 2.483(5) Na(l)-O(204)#l 2.485(5) Na(1)-O(204) 2.485(5) Na(1)-O(107) 2.503(5) Na(1)-O(107)#l 2.503(5) Na(1)-O(207)#1 2.513(5) Na(1)-O(207) 2.513(5) S(2)—C(3) 1.565(9) C(3)—N(4) 1.166(9) O(101)-C(1 12) 1.426(6) O(101)-C(102) 1.432(6) C(102)-C(103) 1.484(8) C(103)-O(104) 1.427(6) O(104)-C005) 1.427(6) C(105)-C(106) 1.489(8) C(106)-O(107) 1.419(6) O(107)-C(108) 1.436(6) C(108)-C(109) 1.483(8) C(109)-O(l 10) 1.424(6) O(110)-C(1 1 1) 1.430(6) C(1 1 1)-C(1 12) 1.483(8) O(201)-C(202) 1.430(6) O(201)—C(212) 1.437(6) C(202)-C(203) 1.480(8) C(203)-O(204) 1.433(6) O(204)-C(205) 1.424(6) C(205)-C(206) 1.482(8) C(206)-O(207) 1.436(6) O(207)-C(208) 1.429(6) C(208)-C(209) 1.474(7) C(209)-O(210) 1.441(6) O(210)-C(21 1) 1.424(6) C(2] 1)-C(212) 1.484(8) O(301)—C(302) 1.230(10) O(301)