:a‘a' -“-*< w.“ "71:11:;- ' w‘vv ) ‘ x— m“ v. I 1 ‘ 1' I I . . ‘ . 1 . l ' ugh; m‘ 3%}. 1:133:51, 3ft!“ :1:. .1 WM 3% . 9. _. @1535” 'I' ..j. E . . \I .‘ I 3;»: 4 ‘Aul ‘cwq‘g‘... \ “.l't\'1'\':n|"'|.'3\t‘ a}? ' ”My. ' ‘ .Ir‘ . , 1 M. 43.. “4:91.?“ J "n+1” f ‘ "I -'- vzwflr'auvfi': W'lbv , l"lfitv.. 'I'tfi 6‘11 ::|:‘-;-'1’d®1w17:1:£ ”at 'rr-." LII-‘1‘ U. ‘-_u :K' .310 .u ‘ 3:16 '1’ 1:01.}. .9}: ‘. ‘L‘|.VC1."I1 v .1». .t ‘5: ‘l.lh7’::;u.:' I‘ '1 . .4333. -. u . '.:~ . I cln‘ t ":h'uiiw “ ‘ O‘Q .. 31:91; :1.; ft} :19?” 33" 1-K? \ s ‘\ 5. .;—«~_._, ,,. .x- n.“ rm. ', *ffi’l’lfi‘v x... :33“ \’ '1' £9 “‘1 1 . ¢ Ib’: . ‘H.’ 0"1 “'11- ll 1'] r ctr, ‘- ':; .3;- {M $4 "1* 0‘“! b, .8" fr' _-. [( Hf»: (if T":"' This is to certify that the thesis entitled Part I) AN NMR ANALYSIS OF CRYPTAND 111. CONFORMATIONS AND RATES OF PROTON BINDING. Part II) AN NMR STUDY OF ALKALI Date METAL ANIONS IN SOLUTION presented by Patrick Bernard Smith has been accepted towards fulfillment of the requirements for Ph.D. Chemistry degree in of? Major professor November 10, 1978 0-7639 OVERDUE FINES ARE 25¢ PER DAY PER ITEM Return to book drop to remove this checkout from your record. PART I) AN NMR ANALYSIS OF CRYPTAND 111: CONFORMATIONS AND RATES OF PROTON BINDING PART II) AN NMR STUDY OF ALKALI METAL ANIONS IN SOLUTION By Patrick Bernard Smith A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1978 I‘\ ‘ \ f l\\ \ "J \ J \ ABSTRACT PART I) AN NMR ANALYSIS OF CRYPTAND 111: CONFORMATIONS AND RATES OF PROTON BINDING PART II) AN NMR STUDY OF ALKALI METAL ANIONS IN SOLUTION By Patrick Bernard Smith The conformations and sites of proton binding by cryptand 111 (C111) have been characterized in great detail. The free ligand exists in a conformation which has both ni- trogen lone pairs pointing inside its cavity (endo-endo), although external protonation studies indicate a degree of flexibility toward nitrogen inversion. The endo-endo form of the ligand in the absence of acid is preferred over the exo-endo form (one nitrogen lone pair pointing inside the cavity and the other outside) by a factor of about 50 in water at 298 K. External proton binding by 0111 occurs primarily at the lone pairs of the nitrogen atoms with no apparent par- ticipation in the binding by the oxygen atoms. C111 will protonate externally at a nitrogen lone pair (exo-endo form). The pKa value for external protonation in water at 298 K is 7.1 1 0.1. The pKa value for the second exter— nal protonation, which occurs at the other nitrogen atom (exo-exo form), is between 0.0 and 1.5 in water at 298 K. Patrick Bernard Smith The exo-endo form of Clll-H+ is preferred over the exo-exo form by a factor of 107 or greater. One or two protons may also be bound internally by C111. The first internal protonation is irreversible since the complex cannot be deprotonated unless the cryptand is decomposed. Internal proton binding also occurs primarily at the nitrogen atoms. The internal proton rapidly ex- changes between the nitrogen ione pairs, this motion being more rapid than the NMR time-scale even at 170 K. The internally monOprotonated complex exists predominantly in the endo-endo form, which is favored over the exo-endo con- figuration by a factor of about 108. The internally mono- protonated complex will externally protonate at a nitrogen atom (exo-endo form). The pKa value for this process in water at 298 K is roughly O.u. The internally diprotonated ligand, Clll-2H+,i-i exists in the endo-endo form. The internal protons of Clll-2H+,i-i either exchange very slowly or not at all. The temperature dependence of the NMR spectra of C111 and its internally protonated complexes indicate that as the temperature is lowered, one or more molecular motions are slowed down considerably. One such motion, which has been assigned to a vicinal carbon wagging motion, possesses 1 deg'1 for AH+ = 9.0 1 0.5 kcal mol-1 and AS+ = -6 cal mol- the internally monOprotonated species. A second type of molecular motion, which is only observed in the case of the internally monOprotonated ligand, is also slowed down at Patrick Bernard Smith temperatures below 190 K. This motion has AHI 8 16 t 2 kcal mol'"1 and AS4‘ = +39 cal mol"l deg"l and has been as- signed to a concerted torsional motion of the ligand. The rates of internal protonation and deprotonation processes of Clll are very slow, due to large activation barriers which are about 25 kcal mol'1 for each of the proc- esses. The rate of the first internal protonation exhibits a definite pH dependence which has been correlated with the pH dependence of the equilibria involving the externally protonated forms of Clll. It appears that two forms of the ligand are able to go to products with different rate con- stants, the first constant being 3.8 t 0.6x10'3 sec"1 and the second being 2.3 t 0.3x10'u sec-1. The rates for the second internal protonation and the removal of the second internal proton are also expected to exhibit a pH dependence although the rates of the second in- ternal protonation have not been studied as a function of pH. The rate of deprotonation of Clll-2H+,i-i by base has been shown to depend on the base strength. The rate data provide a means by which to estimate the thermodynamic stability of the various complexes. The pKa value for the first internal protonation of Clll in water at 298 K is 217.8 and AG° 5-5 kcal mOl-l. For the second in- ternal protonation, pKa 28.3 and AG° 2+8 kcal mol-l. NMR provides an excellent way to identify and charac- terize alkali metal anion solutions. NMR data reveal that the trend in thermodynamic stability of the alkali metal anions, relative to the solid metals is Na'>Cs'>>K',Rb'. to Karole ii ACKNOWLEDGEMENTS The author wishes to express special gratitude to Professor James L. Dye for his continual guidance and en— couragement. Professors A. I. Popov, J. F. Harrison, D. G. Farnum, and L. R. Sousa are also acknowledged for their many helpful discussions and insights. Appreciation is also extended to Mr. M. Yemen, Mr. M. DaGue, Dr. E. Kauffmann, Mr. B. Van Eck, and Dr. Harlan Lewis for their many hours of assistance. Mr. F. Bennis and Mr. W. Burkhardt must also be recognized for their assistance and support concerning the operation and main— tenance of the NMR spectrometers. Financial support from the Dow Chemical Company, the United States Energy Research and Development Association, and the General Electric Company is also acknowledged. Finally, my sincere thanks to my wife, Karole, and our families for their unending love, confidence, and sup- port. 111 LIST OF TABLES LIST OF FIGURES . . . . . . . . . . . . . . . LIST OF COMMONLY USED TERMS . . . . . . . . . PART I) AN NMR ANALYSIS OF CRYPTAND 111: CONFORMATIONS AND RATES OF PROTON BINDING Chapter 1. HISTORICAL . . . . . . . . . . . . . . I) INTRODUCTION . . . . . . . . . . II) GENERAL FEATURES OF NMR . . . . TABLE OF CONTENTS A) The Chemical Shift . . . B) Spin-Spin Coupling . . . C) NMR Relaxation . . . . . 2. EXPERIMENTAL I) II) III) IV) v) VI) VII) VIII) IX) GLASSWARE CLEANING . . . . SOLVENT PURIFICATION LIGAND PURIFICATION . . . SAMPLE PREPARATION . . . . A) Preparation of Anhydrous NMR Samples . . . . . B) Alkali Metal Anion Solution Preparation . . . . . Clll SYNTHESIS AND PURIFICATION METAL PURIFICATION VACUUM PROCEDURES . . NMR SPECTROMETER . . . . SAFETY . . . . . . . . . . . . iv vii ix xiii a: r4 t4 19 29 29 29 3O 31 31 33 35 37 MO “2 H3 3. CONFORMATIONAL ANALYSIS OF CRYPTAND 111 . . . I) II) III) INTRODUCTION . . . . . . . . . . . . . A) Clll Symmetry Considerations . NMR SPECTRA AT AMBIENT TEMPERATURES A) Clll in ’dQ-Acetone o o o o o o o B) Clll H+ ' in d -Acetone . . 0) 0111 2H+, 1- 1 201’ In du MeOH . . D) Musher's Theory of the Chemical Shift 0 O O O O O O I O O O O E) 011L H+,o in d6~Acetone . . . . F) C1110 0H+ ,o in Water . . . . . . G) The External Protonation of 0111 H+, 1 . . . . . . . . . . H) General Conclusions . . . . . . TEMPERATURE DEPENDENCE OF THE NMR SPECTRA O O O O O O O O O O O O O O O A) C111 in d6-Acetone . . . . . . . 8) C111 H+ 4i- oBr in d -Acetone . . C) Clll 2H ,1- i 2C1 In du-MeOH . . D) 0111- 2H+, i-o 1n du-MeOH . . . . E) Clllo H+ ,o in d6-Acetone . . . . F) summary 0 O O O O O O O O O O O A. THE KINETICS OF INTERNAL PROTONATION OF C111 I) II) III) IV) v) VI) VII) VIII) INTRODUCTION . . . . . . . . . . . . . THE FIRST INTERNAL PROTONATION OF Clll THE KINETICS OF THE SECOND INTERNAL PROTONATION OF C111 IN WATER . . . . . THE KINETICS OF DEPROTONATION OF Clll'2H+’1-i o o o o o o o o o o o o 0 METHOD OF ANALYSIS . . . . . . . . . . SUMMARY 0 o o o o o o o o o o o o o 0 THE REACTION OF Clll H+ ,1 WITH Na+ C222. Na- 0 I O O C O O O O O O O C THE REACTION BETWEEN C111-2H+,i-i AND es 0 O O O I O O O O O O O O O O O O O 5. DISCUSSION “5 A5 53 56 59 60 62 65 66 78 82 83 84 85 105 108 117 121 125 125 126 138 1A2 1A5 1“? 152 153 156 I) II) III) PART II) AN NMR STUDY OF ALKALI METAL ANIONS IN SOLUTION INTRODUCTION 0 O C O O O O O C O C O O O O O O O 16“ THERMODYNAMIC STABILITY OF THE ALKALI METAL ANIONS O O O O O O O O O O O . O O O O O O O O O O 172 IDENTIFICATION OF NEW SPECIES . . . . . . . . . 178 vi TABLE 1. 10. 11. 12. LIST OF TABLES Page Summary of relaxation mechanisms . . . . . . 27 The solubility of C111 and its complexes in various solvents . . . . . . . . . . . . . . 38 The chemical shifts of C111 and its complex- es at ambient temperature . . . . . . . . . 58 The dependence of the chemical shifts of Clll in d -acetone upon addition of trifluoro- acetIc acid (TFAA) . . . . . . . . . . . . . 72 The dependence of the chemical shifts of C111 in d -acetone upon addition of dichloro- acetIc acid (DCAA) . . . . . . . . . . . . . 75 The pH dependence of the chemical shift of the CH2N protons of C111 in water at 299°K . 80 The pH dependence of the chemical shifts of C111.H+ ’1 in water 0 O O O O O O O O O O O O 83 The temperature dependence of the NMR param- eters for a 0.02 M solution of Clll in d6- acetone 0 O O O O O O O O O O O O O O O O O 89 The temperature dependence of the NMR spec- tra of a 0. 0“ M solution Of C111-H+,i Br in d6-aCEt0ne o o o o o o o o o o o o o o o o o 91 The temperature dependence of the NMR+param- eters for a 0.0M M solution of Clll-H ,i-Br’ in du-MGOH o o o o o o o o o o o o o o o o o 10“ The temperature dependence of the NMR param- eters for a 0.04 M solution of 0111-2H+,1-1-201' in du-MeOH . . . . . . . . 107 The temperature dependence of the NMR param- eters for a 0.0u M solution of Clll-H+,i in du-MeOH, acidified with HCl . . 110 vii TABLE 13. 1A. 15. 16. 17. 18. 19. 20. 21. 22. 23. 2A. 25. Page The temperature dependence of the equilib- rium constant for the conversion of N Clll-H ,1 to 0111-2H+,i-o inidu-MeOH: K from the CH2-N protons and K from the in- ternal proton . . . . . . . . . . . . . . . . 112 The temperature dependence of the equilib- rium constant for the formation of 0111-2N+,o-o 1n d6-acetone with the addition of TFAA . . . . . . . . . . . . . . . . . . . 121 The dependence of the equilibrium constant for the conversion of Clll-H+,o to C111'2H+,0-0 on the concentration of TFAA . . 122 Buffer compositions and their ionic strengths in D20 I O O I O 0 O O O O O O O O O O O O O 127 The dependence of the rate of internal mono- protonation of C111 upon pH and temperature . 128 The dependence of the rate of the second in- ternal protonation of C111-H+,i upon tempera- ture in l M HCl . . . . . . . . . . . . . . . 142 The dependence of the rate of deprotonation of 0111-2H+,i-1 upon temperature in 5 M KOH . 1A5 The activation and thermodynamic parameters for internal protonation processes at 298 K . 1A8 A selected list of chemical shifts and line- widths in alkali metal NMR . . . . . . . . . 171 The products of some alkali metal solution preparations . . . . . . . . . . . . . . . . 173 The products of the reaction of Na and Cs metals and C222 in THF . . . . . . . . . . . 176 The temperature dependence of the chemical shift of Cs+0222 in EA . . . . . . . . . . . 179 The temperature dependence+of the EPR signal of a 0.08 M solution of Na l8-C-6-Na‘ in MA . 187 viii FIGURE 1. 10. 11. 12. 13. 1“. LIST OF FIGURES Structures of cryptand 222, 18-crown-6 and cryptand 111 O O O O C O O O C O O O O O C O The dependence of the formation constant on the ratio of the size of the ion to the ligand's cavity size . . . . . . . . . . . . The protonation scheme of C111 . . . . . . . Diagram showing the components of the Musher theory of proton chemical shifts . . . . . . The effect of unpaired electrons on the NMR spectrum 0 O O O O O O O O O I O O O O O O O The effect Of bond angle on scalar coupling, for A) vicinal protons and B) geminal pro- tons O O O O O O O O O C O O O O O O O O O 0 Apparatus used in the preparation of anhy- drous samples for NMR analysis . . . . . . Apparatus used in the synthesis of alkali metal anion solutions . . . . . . . . . . Steps in the synthesis of Clll . . . . . . "Trombone" apparatus for the storage Of alkali metals . . . . . . . . . . . . . . . Vacuum line arrangement used in sample prep- aration O O O O O O O O O O O O O O O O O O The conformations of 0222 . . . . . . . . . Space filling models of C111, A) D h sym- metry (large cavity) and B) compacé cavity The possible binding sites and conformations of C111, A) unprotonated, B) external pro- tonation, C) internal protonation, D) exter- nal protonation of the internally ix Page 1“ 16 20 32 3A 36 39 141 AS “9 FIGURE 15. 16. 17. 18. 19. 20. 21. 22. 230 2A. 25. 26. 27. 28. Page monOprotonated complex and E) internal dipro- tonation O O O O I I O O O O O O O O O I O I 50 Model showing the torsional angle of C222 . 53 Model showing the symmetry elements of C111 in the D3h configuration . . . . . . . . . . 55 The NMR spectra of C111 and its protonated complexes at ambient temperature . . . . . . 57 The dependence of the NMR spectrum of C111 in d -acetone upon the addition of trifluoro- acetéc aCid (TFAA) O I O O O O I O O O O O O 69 The dependence of the NMR spectrum of C111 in d -acetone upon the addition of dichloro- acetIc acid (DCAA) . . . . . . . . . . . . . 73 The dependence of the chemical shift of the CH O protons of C111 in d6-acetone upon ad- diéion of DCAA . . . . . . . . . . . . . . . 7A The NMR spectra of Clll in water at various pH values 0 O O O O O O O O O O O O O O O 0 77 The pH dependence of the chemical shift of the CH2N protons of C111 in D20 . . . . . . 79 The spectrum of C111 in CHF2Cl at 163 K at 250 MHz 0 O O O O O O C O O O I O O C O O O 86 The temperature dependence of the NMR spec- tra Of C11]. in d6-acetone o o o o o o o o o 88 The temperatu e dependence of the NMR spec- tra of Clll-H ,i in d6-acetone . . . . . . . 90 Computer-simulation of the temperature depen- dence of the spectra of the CHZO protons of C111’H ’1 in d6'acetone o o o o o o o o o o 93 The dependence of 1n(k) on 1/T for the tor- sional motion from the spectra of the CH2O protons of 0111-H+,1 . . . . . . . . . . . . 95 The temperature dependence of the internal proton resonance of Clll-H ,i in d6- acetone 0 O O I O I O O O O O O O O O O O O 97 FIGURE 29. 30. 31. 32. 33. 3M. 35. 36. 37. 38. 39. “0. A1. Computer-simulation of the temperature de- pendence of the internal proton resonance of 0111’ 11+, 1 in d6-acetone o o o o o o o o o o o The dependence of ln(k) on 1/T for the skele- tal locking process from the spectra of the internal proton of 01110 H+ ,i . . . . . . . . The temperature dependence of the NMR spectra Of C111. 11+ ,1 in du-MEOH o o o o o o o o o o o The temperature dependence of the NMR spectra Of C111. .2H+ ”1 1 in du-MGOH o o o o o o o o o The temperature dependence of the NMR spectra of C111-H,i in du-MeOH, acidified with HCl . The dependence of ln(K) on l/T for the equi- librium between Clll- H ,i and Clll 2H+, i-o in du-MGOH o o o o o o e o o o o o o o o o The variation in the NMR spectrum of the in- ternal proton of Clll H+, i in du-MeOH with the addition of HCl at 200 K . . . . . . . . The temperature dependence of the NMR spec- trum of the internal proton of Clll-H+,i in du-MeOH with about 50 mole % HCl added . . . The temperature dependence of the NMR spec— tra of Clll H ,o in d6-acetone, acidified "1 th TFAA O 0 O O O I C O O O I O O O O O 0 I The variation of ln(K2 ) with 1/T for the for- mation of Clll 2H+ ,o-o in d6-acetone upon the addition or TFAA O O O O O O O I O O O O I O The dependence of the equilibrium constant for the +conversion of 0111- H ,o to 0111- 2H+ ,o-o in d6-acetone on TFAA concentra- tion 0 O O O O O I O O O O O O 0 0 O O O O O A typical rate analysis for internal protona- tion of 0111 in water . . . . . . . . . . . . The dependence of ln(k) on 1/T for the inter- nal protonation of Clll in water at two dif- ferent pH values: squares; pH 8 7.5 and cir- cles; pH = “.9 . . . . . . . . . . . . . . xi Page 98 99 103 106 109 113 115 116 118 120 123 129 130 FIGURE A2. A3. A“. “5. A6. “7. A8. “9. 50. 51. 52. 53. 5A. The pH titration of Clll with H01 in water (only external protonation occurs) . . . . . The variation of the relative concentrations of the various forms of Clll in water with pH. 0 O O O O 0,. O O O O O O O O O O O I O The dependence of the rate of internal mono- protonation of 0111 on pH: dashed line; pK a 7.1 and solid line; pKa . 7.5 . . . . . 2 A typical rate analysis for the second in- ternal protonation of Clll H ,i in l M HCl . The dependence of ln(k) on l/T for+the sec- ond internal protonation of Clll-H ,i in l M HCl 0 O O O O O I O O O O O O O O O O C O O A typical rate analysis for the deprotona- tion of Clll 2H+ ,i- in 5 M KOH . . . . . . The dependence of ln(k) on l/T for the de— protonation of Clll 2H+ ,i- -i in 5 M KOH (crosses) and l M KOH (circle) . . . . . . . The potential energy diagram for the inter- nal protonation processes . . . . . . . . . The dependence of the total energy of the complex on the location of the H proton along the C3 axis for Clll H+ ,i in D3h sym- metry . . . . . . . . . . . . . . . The electronic spectra of the alkali metals in tetrahydrofuran (THF) . . . . . . . . . . The temperature dependence of the NMR spec- trum of Cs+ 0222 Cs” in MA . . . . . . . . . The temperature dependence of the NMR spec- trum of Cs +0222 Cs in THF . . . . . . . . . The temperature dependence of the NMR spec- trum Of Na+18-C-6.-.Na in MA 0 o o o o o o o xii Page 133 135 136 1A0 181 1u3 11m 151 159 165 180 183 186 0222 0111 18-C—6 Clll-H+,o Clll-H+,i Hi [.1 LIST OF COMMONLY USED TERMS cryptand 2-2-2 cryptand l-l-l 18-crown-6 externally complexed Clll internally complexed Clll inclusively bound proton magnetic moment magnetic field strength nuclear spin or moment of inertia magnetogyric ratio local magnetic field at the nucleus screening constant paramagnetic contribution to O diamagnetic contribution to 0 average excitation energy coupling constant spin-lattice relaxation spin-spin relaxation relaxation rate full linewidth at half height in Hz correlation time or exchange time Viscosity Boltzmann's constant xiii EA IPA DEE Na-K THF (DI U) KNEE Planck's constant divided by 2w temperature in °K nuclear spin Larmor frequency chemical shift methylamine ethylamine isopropylamine diethylether sodium-potassium alloy tetrahydrofuran solvated electron alkali metal anion alkali metal cation equilibrium constant rate constant xiv PART I AN NMR ANALYSIS OF CRYPTAND III: CONFORMATIONS AND RATES OF PROTON BINDING CHAPTER 1 HISTORICAL I) INTRODUCTION The complexation of metal cations in solution is of particular interest to chemists and biologists partly because of the mystery surrounding ion transport in biological sys- tems(l). Most attention has been given to transition metal complexes with ligands such as ammonia, carbon monoxide, EDTA, etc.(2), since these elements tend to form rather stable and long-lived adducts. This is true because the transition metal ions, in general, have valence deficiencies, and thus coordinate to electron donating ligands in order to satisfy the valence condition. In sharp contrast to the transition metals is another group of biologically signifi- cant ions, the alkali and alkaline earth elements which do not have this valence deficiency. Instead, they tend to be rather inert and very labile in solution. These ions com- plex with conventional ligands so inefficiently that they have often been used to maintain constant ionic strength in solutions where the complexation of other metal ions was be— ing studied. Alkali metal complexation has been shown to be impor- tant in the phenomenon of active ion transport across mem- branes. Biologically significant molecules have been observed (1’3). These molecules include to facilitate this process porphyrin and corrin ring complexes, phthalocyanines and l 2 valinomycin, each of which is a macrocycle that possesses a hydrophilic cavity and an organic backbone. The macro- cycle is thought to bind the ion in the cavity, while main- taining its organic nature, thereby dissolving the ionic species in the membrane. This process allows the complexing agent to "carry" the ion across an otherwise inaccessible membrane barrier with the expenditure of little energy. Synthetic macrocycles whose function mimics their naturally occurring analogues were introduced by Pedersen in 1967(8). These monocyclic polyethers were named "crown ethers" because Pedersen envisioned them as sitting atop the complexed ion in a fashion resembling a crown. Since the introduction of crown ethers, many more macrocyclic complex- ing agents have been synthesized, the designers of which em- ployed freedom of imagination and prowess in the field of synthesis to achieve very specific ends. Cryptands, bi- cyclic polyethers developed by J.-M. Lehn and co-workers, are the most famous Of these, since they combine extremely strong binding characteristics with a well-defined, three dimen- sional cavity. Thus, they possess selectivity in their bind- ing characteristics with various ions, based on the ratio of cavity to ionic size(5). Typical crown ethers and cryptands are shown in Figure l. The demonstrated ability to synthe- size these macrocycles, coupled with the unlimited freedom of their design has provided for great utility of these ligands in the aforementioned biological applications as well as in several other significant applications(6'12). KONG/W O 0 fl 0"‘\ N’VO’VO’VN [O O] N ’\,§’\,N L‘vx%__JSA,/j K\¢JD\’/J \\__/ \__J/ CRYPTAND 222 IB'CROWN'G CRYPTAND lll Figure 1. Structures of cryptand 222, 18-crown-6 and cryptand 111. It is not within the scope of this work to present a de- tailed review of their applications. Instead, let it suf- fice to say that macrocyclic polyethers hold potential in all areas of ionic separations and have, indeed, proven useful in many(l3'2u). As mentioned previously, cryptands differentiate between ionic species by binding most strongly those ions which are compatible with their cavity dimensions. This is portrayed in Figure 2 for several cryptands in which the binding constant of a particular ligand is plotted versus the ratio of the sizes of ion and ligand cavity. These plots peak as the ionic size equals that of the cavity of the ligand, and fall off considerably on either side of this value. The smallest cryptand, C111, is not shown because its cavity is too small to bind alkali metal ions efficiently<2u). Cryptand 111 does bind Li+ slightly with a pK of approximately 2.2 in water, but comparison with C211, which has a pK of 5.5, demonstrates that C111 is too small for .oufim huw>do m.ccmeH 029 On 20H on» no mean on» no oapmh on» so pcmpmcoo coaumsuom on» no mococcoqov 029 .m madman >260 mo 350.. :2 .206 B «:69. o.~ o... o u 030.. 5 efficient binding to occur. (From molecular models, the cavity radius of 0111 has been estimated to be about 1.0 A.) This size dilemma would render Clll rather mundane in the realm of ionic complexation except that it binds protons and does so with such efficiency that it is the strongest proton complexing agent yet reported. Clll binds one or two protons in a number of binding sites and conformations, thus qualifying it as a model Of the binding characteristics of the larger cryptands as well as biological systems<25). Lehn and cO-workers published two preliminary papers about 0111(26'27), in which they described the synthesis and unusual proton binding ability Of Clll. They also discussed the possibility of several different protonated complexes which might be formed as illustrated in Figure 3. In protic solvents or acidic media, Clll was shown to complex a proton externally, presumably at the nitrogen with its lone pair pointing outside the cavity. This complex behaved as a typ- ical protonated amine, the proton being readily exchangeable with protons in solution. Evidence for such complex forma- tion came from a rise in pH upon addition to aqueous solu- tions and from line broadening in the NMR spectra. Lehn estimated that the pKa for external protonation would be about 7.5-8.0 based on triethanolamine or N-ethylmorpholine if the ligand were present with one nitrogen lone pair point- ing outside the cavity and available for binding (exO-endo configuration). The actual pKa was found to be somewhat A lower (6.3) based on the pH of a 2x10' M solution of Clll 7 and therefore, an equilibrium between exo-endo and endo- endo forms was suspected. A second external protonation (at the other nitrogen) was thought to induce too severe a strain to be possible because both nitrogens would be forced into the exo conformation. This conclusion was based on molecular models and the fact that the reduction of the dilactam of C111 by diborane results in the monoborane ad- duct rather than the diborane adduct obtained in the case of larger cryptands (see Experimental section for details). Simultaneous with the rapid external protonation, a second, much slower process was observed by which a proton was very tightly bound. This process proceeded with a half- life of several days at room temperature and presumably cor- responded to inside protonation. The resulting complex was so stable that heating it to 60°C in 5 M KOH for days did not significantly deprotonate it! Even with excess sodium in liquid ammonia, the proton could not be removed without de- stroying the ligand. The internally monOprotonated ligand, Clll-H+,i, could be protonated a second time externally. The degree of exter- nal protonation was pH dependent, leading to the conclusion that there was again an equilibrium between the endo-endo and exo-endo configurations, the former being more favorable. As the pH was decreased, the ligand was forced into the exo- endo form but the exchange was always very rapid. A second proton could also be taken inside the cavity (endo-endo), but only under rigorous conditions (1 M HCl at 8 100°C for one hour). Again, this inclusive proton was found to be bound tightly and was very difficult to remove from the cavity, the complex being stable for days in 5 M KOH at room temperature. Above 60°C, however, one of the two protons could be removed slowly. Treatment with sodium in liquid ammonia caused reduction to the free amine in about 10% yield with about 90% decomposition. No mono- protonated ligand was ever obtained and thus an intermediate of the sort, Clll-H +, 1—1 was postulated, the reaction pro- ceeding via reduction rather than proton abstraction by base. In general it was thought that proton binding occurred mainly at the nitrogens and that external protonation was weak and rapidly exchangeable whereas internal protonation was difficult to achieve but very strong. Subsequent sec- tions of this discussion will deal with binding character- istics and rates of encapsulation of protons by 0111. II) GENERAL FEATURES OF NMR Nuclear magnetic resonance provides a direct probe of the electronic environment of the nucleus since the resonance frequency, or Larmor frequency, of that nucleus is greatly dependent on electronic shielding. The symmetry of the electron cloud and the electron density in the vicinity of a nucleus contribute to local magnetic fields which influence its chemical shift. These contributions will be discussed qualitatively in order to give significance to the Clll 9 spectra in the latter portions Of this work. For a more detailed discussion, see references 28-33. A) The Chemical Shift When a nucleus is placed in a magnetic field, Ho, its nuclear spin states are no longer degenerate but are separated by an energy difference given by: pH 0 AE I _I_'= YHOh (Hz) (1) where: u - Magnetic moment of the nucleus Ho - Magnetic field (gauss) I — Nuclear spin 7 - Magnetogyric ratio h - Plank's constant divided by 2 Transitions are induced between the levels by irradiation at the Larmor frequency, equal to the energy separation of the nuclear spin states (in the radio frequency region). The electronic distribution becomes important when deter- mining the true magnetic field at the nucleus, Hloc’ because the electrons produce magnetic fields of their own as they interact with Ho and may therefore cause considerable shift— ing in the resonant frequency from that given by equation (1). The electrons perturb the field at the nucleus in a number of ways. First, as was mentioned before, when an atom is placed in a magnetic field the nuclear spin states separate but the field also induces a precession of the elec- tron cloud about the field, Ho. The electronic precession in turn generates a magnetic field, which opposes HO so that 10 the nucleus is "screened" by these electrons according to the equation: H .. Ho(l-O) (2) 100 where: H10c - the field at the nucleus 0 - the screening constant This type of screening, which results in an upfield shift, is called diamagnetic screening and may be a sum of several contributions. Other effects, which cause downfield shifts due to the mixing of excited state molecular wave functions, are referred to as paramagnetic terms after the development of Ramsey<28), and hence: 0 = 0d + Up (3) Ramsey expressed these terms theoretically from perturbation theory as follows: AAA e2 rkzl-rkrk O = ‘—‘ + 2M02 k rk -1 A I MEG-EM) 21nxm> (5) kk’ r; where: A - average excitation energy When A is large, as in proton NMR, Op contributes little to the chemical shift. When A is small, excitation occurs readily and Up, which potentially can be very large, usually dominates. More specifically, 0d dominates in proton NMR even though it is only about 18 ppm, since op is negligible. In carbon and fluorine NMR, 0 is much larger, (about 260 d and 338 ppm respectively), but op may get as large as 1,000 ppm and is therefore the more dominant of the two. In some 12 heavy metal systems, the value of 0p may get as large as 10,000 ppm because of a high density of excited states and their ease of accessibility. The a term is also relatively d constant for a particular nucleus even though it may be large and its range of shifts is usually only about 20 ppm at most. The magnitude of O , on the other hand, varies p greatly from nucleus to nucleus, but more importantly, may change by several hundred ppm for the same nucleus in dif- ferent environments. For example, Cs+ in the gas phase and Cs+ inclusively complexed by the cryptand, 0222, are shifted about 500 ppm from one another<3u). Ramsey's equations provide a good theoretical start- ing point, but in practice are of little value in the pre- diction of chemical shifts due to the unavailability of molecular wave functions. More empirical approaches have found greater utility in this area and will be discussed. In order to be consistant with the literature, downfield shifts will be referred to as paramagnetic and upfield shifts as diamagnetic, but this terminology should not be confused with the two terms in the Ramsey equation. There are contributions to a which lead to paramagnetic shifts, d as we shall see, but which do not involve the mixing of ex- cited and ground electronic states. For example, the para- magnetic term in the Ramsey equation is extremely small for proton NMR and yet large downfield shifts from TMS are ob- served. These shifts are due to reduction of ad by the 13 decrease in electron density, etc., rather than to contri- butions from the 0d term. In general, the contributions to the chemical shift may be classified as either diamagnetic or paramagnetic, but there are several factors which may influence these terms. Four of the more important contributions in proton NMR include: a) anisotropic screening b) electric field effects 0) unpaired electrons d) solvent effects a) through d) contribute to °d whereas only d) may signifi- cantly contribute to O for nuclei with available excited P state orbitals. a) Anisotropic Screening Anisotropic effects are caused by an asymetrical elec- tronic distribution of neighboring substituents which does not average to zero with molecular tumbling. This asym- metry causes time-independent local fields at the nucleus which result in the shift. Perhaps the most familiar exam- ple of these effects is in relation to ring currents produced by aromatic systems, but all n electron systems have the po- tential to produce anisotropic screening. These shifts are normally only about l-A ppm, significant in proton NMR, but much less so in carbon NMR. b) Electric Field Effects A charged species or highly polar group may induce polarization of a bond and alter the electronic distribution 1A about a nucleus. This causes a two-fold change in the local field at the nucleus by altering the electron density around the nucleus and by initiating an anisotropic electronic en- vironment. The shift produced may be represented in the following way(30): qicosei Q1 0 = -l3.3 Z _______ -l7.0 Z __ (ppm) (6) e i r2 1r2 1 i where: q1 - the total charge of the polar group in units of one electron gisei - defined in Figure A This equation reduces to: o - -ll.2q - 12.0q2 (7) for axial symmetry assuming r = 1.09A, the typical hydrogen- carbon bond length. The former equation, called the Musher equation, provides the means by which to predict the shift expected from a change in the charge distribution at or near a nucleus. It should therefore prove amenable to the study of the protonation of Clll, since the dominant cause of change in the chemical shift of the backbone protons upon protonation is a result of this electric field effect. Figure 4. Diagram showing the components of the Musher theory of proton chemical shifts. 15 c) Unpaired Electrons The magnetic moment of an electron is 860 times as large as that of a proton and therefore, scalar coupling of nuclei with unpaired spins produces enormous local fields, as large as hundreds of thousands of ppm. Dipole - dipole relaxation via electrons is also very efficient and in many cases broadens lines to obscurity. This phenomenon is rep- resented in Figure 5 in which a) shows a line in the absence of scalar coupling. The normal chemical shift range in pro- ton NMR is about 18 ppm, so that scalar coupling to elec- trons with a magnitude of 100,000 ppm may put the resonance well out of the tunable range of the spectrometer. Figure 5b) shows the spectrum when electron relaxation is slow, and 50) when the relaxation is comparable to the hyperfine coupling, A. In the latter case the line is broadened over several kilohertz and in most instances it is not observable. When the relaxation is very fast compared to A, a narrow line occurs as shown in 5d) whose shift is the population average of the two lines as shown in 5b), when electron re- laxation is slow. This average position does not lie at the frequency average of those in 5b) because there is a large population difference between the levels as a result of the Boltzmann distribution and in fact, the line is usually shifted many kilohertz from the average frequency. (This shift also occurs in proton NMR with the collapse of multi- plets resulting from J coupling, but is not seen because the energy separation between spin states is three orders of 16 NMR Spectrum Electron ' Roloxotm 0) A00 I I l a) Ao-ZOMHz I I T,'>A I I A . c) Ao-mMHZ I 71" I l I l a) Ao-2OMH1 I T: 0mm OOH v as v OH poanm Hmofismno omo .maozo . .mmmm .mmz Asxauumv 0mm «OH v He v .-oa mdoassomsa sfioovoom .+:mz .uNmomm I_ I .omm .maomo.. .s\sm>H» 0mm ooa v a v H maoaaonmaoafia msopmam Hmofiome +om aowcmn He soauoomousH .msmficmnooe coaumxoaop ho mumEEsm .H canoe 28 Because T1 may not necessarily be equal to T2, a quantitative analysis using the pulsed-fourier transform (P - FT) method of data acquisition may be Jeopardized. Care must be taken in order to insure complete relaxation before pulse repetition or else standardization may be necessary. A detailed discussion of these problems will be presented in subsequent sections of this work. CHAPTER 2 EXPERIMENTAL I) GLASSWARE CLEANING All glassware used in alkali metal anion synthesis, solvent purification, or alkali metal storage was cleaned by first rinsing with an RF cleaner composed of 33% HNO3, 5% HF, 2% acid soluble detergent, and 60% water by volume. This rinsing was followed by several washings with distilled water and a 12 hour minimum soaking period in aqua regia. The ves- sels were then rinsed a minimum of five times with distilled water and five times with conductance water and dried at 110°C for at least eight hours. This procedure insured a fresh glass surface free of reactive species. II) SOLVENT PURIFICATION Tetrahydrofuran (THF) was dried over calcium hydride for at least 24 hours and then vacuum distilled over a sodi- um-potassium (Na-K) alloy. The resulting light blue color served as an indicator of solvent dryness. If this color faded, the solvent was transferred over a new metal alloy, but in most cases the solution was stable for several weeks. The Na-K alloy slowly attacked the THF to give a polymer, so the THF was transferred away from it for storage. Methylamine (MA), ethylamine (EA), and isopropylamine (IPA) were also initially dried over calcium hydride and then vacuum distilled over the Na-K alloy. If a blue 29 30 solution did not result, the amine was redistilled over a fresh metal film until the blue color was stable for two days. The solvent was then distilled into Pyrex storage bottles since the Na-K alloy also slowly decomposed the amines to yield ammonia and other decomposition products. Diethyl ether (DEE) was dried over calcium hydride and then vacuum distilled into a vessel containing benzophenone and excess Na-K alloy. A purple color developed with the formation of the benzophenone ketyl radical. If the purple color faded, the ether was redistilled onto a new mixture until the color became permanent. This solvent was stored over the metal since decomposition was negligible. Deuterated solvents were refluxed at least eight hours with barium oxide and then vacuum distilled onto freshly dried type AA molecular sieves (Linde, Union Carbide). The sieves were dried by passing dry nitrogen over them while baking at approximately 250°C. Final drying was accomplished by heating under vacuum. The solvent was stored over molecu- lar sieves to insure dryness and samples were made up by vacuum distilling into an NMR tube with dry solute already in place. This procedure provided samples with less than 100 ppm water. (usually about 50 ppm) with no noticeable- contamination or decomposition. III) LIGAND PURIFICATION C222 (E.M. Laboratories, Inc.) was recrystallized twice from hexanes and vacuum sublimed at approximately 31 110°C in the dark using a cold finger filled with a dry ice-isopropanol mixture. C111 was purified by filtering through a type F—2O basic alumina column with anhydrous ether. The liquid cryptands, C211 and C221 (both obtained from E.M. Laboratories) and C322 (synthesized in our labo- ratory at Michigan State University), were vacuum distilled. The purification of l8-C-6 (PCR, Inc.) was accomplished by following the procedure established by E. Mei(36). All ligands were stored under vacuum and in the dark. IV) SAMPLE PREPARATION A) Preparation of Anhydrous NMR Samples Figure 7 shows the apparatus used in the preparation of anhydrous samples for NMR analysis. The solvents were dried by using the previously mentioned methods and stored via vacuum distillation in Pyrex bottles equipped with high vacuum valves and Fischer-Porter Joints. The solvent was introduced into the NMR tube by distilling through a tee, the bottle being attached to one end and a valve, fitted with a ground glass Joint attached to the other end. An NMR tube, also equipped with a compatible ground glass Joint was affixed to the valve (using a minimum of high vacuum grease) and the entire system evacuated. Thus, any volatiles could be pumped from salts, ligands, etc. before distilling in the solvent. The solvent could then be distilled, the sample re- moved from the tee, and capped by a layer of argon using a 32 .mfimmawcw mzz now moHaEMm moonozncw mo cofiumnmoono on» Ca 0mm: maumpmqo< .5 onswfim 2:2 .35 p.52 9N1 an... EEG \< fin \J \r 2:: :23. 33 "Schlenk" line procedure(87) and the valve quickly replaced by a sealed ground glass top. This procedure allowed water content in the final sample to be kept well below 100 ppm (usually about SOppm) with no contamination. B) Alkali Metal Anion Solution Preparation The apparatus used in alkali metal anion synthesis for observation via NMR or EPR is depicted in Figure 8. The ligand was introduced from the top and evacuated to about 1 x 10-5 torr. Metal was introduced through a side-arm and distilled several times so as to obtain a pure metal mirror in the compartment before the frit. The side-arm was then removed by a flame seal-off at the constriction and discarded. The solvent was introduced as described earlier, by vacuum distillation through a tee and the apparatus transferred to a dry ice-isopropanol bath. The solution of ligand was trans- ferred over the metal as many times and at as high a tempera- ture as could be withstood without decomposition. When an equilibrium was achieved, the dark blue solution was poured into the NMR tube and flame sealed at the constriction. The prepared sample was stored at dry ice temperature until NMR analysis was conducted, usually within an hour after sample preparation was complete. The apparatus employed in the synthesis of solid powders and crystals of alkali metal anion salts was very similar to that previously described, except that the NMR tube was not incorporated in this process. The sample was initially made in an identical fashion by distilling metal and solvent, 3A .mCOHp: on :o p m w o m mm pcmm on» c on a no N we 3 mSuMHwaa H H H a x H a H c < .w whom rm T6807. 02252.3 EEC? \ V\ i . .502:- .QSE RN Em , .... an. 9.33 .....E. ....g 35 then washing over the metal several times. The deep blue solution was next poured away from the metal and cooled in a dry ice-isopropanol bath for about an hour and observed periodically for signs of precipitate formation. The cold solution was then decanted back over the metal and warmed as much as possible without inducing decomposition. This procedure was continued until either all the ligand was complexed or decomposition had begun. At this point, the cold solution was decanted back over the metal, leaving the powder behind, and the solvent distilled out of the vessel. The precipitate was either sealed under vacuum and stored in dry ice or washed with a nonpolar solvent, such as diethyl ether (if stable), and stored in a similar manner. v) 0111 SYNTHESIS AND PURIFICATION The synthesis of Clll follows the scheme depicted in Figure 9 and involves the condensation of the acid chloride, 1, with the monocyclic diamine, II, to give the bicyclic diamide, 111(26). This is reduced with diborane to give the unusual monoborane adduct, IV, rather than a diborane adduct, presumably because the ligand cannot extend both nitrogen lone pairs outside the ligand (exo) except with much difficulty. The borane adduct is then reduced with KOH in methanol to give the free diamine. Approximately 200 mg of C111 was received from Professor J.-M. Lehn and was purified by passing through an alumina column with anhydrous ether. 36 NH "* N’\’0 N L: OLEIV<0 Cl 0 Ci 0V.L‘o I Figure 9. Steps in the synthesis of Clll. Residues (8-10 grams) containing the bicyclic diamide of C111 and polymer by-products were also obtained from Lehn's laboratory. The residues were dissolved in CHCl3 and precipitated with diethyl ether. The polymer precipi- tated first and the diamide was left behind (about l.A grams of the impure diamide was obtained). The diamide was re— duced by suspending it in dry tetrahydrofuran (1A0 m1) and about 35 ml of diborane solution (1 M in THF, Alfa Divi- sion, Ventron Corporation) was added dropwise with cooling. The mixture was refluxed for two hours and cooled to room temperature upon which 7 ml of conductance water was added. This addition caused much effervescence as the diborane was decomposed. The solution was evaporated to dryness and ex- tracted with ether using a soxlet procedure. This procedure gave no monoborane product, but instead, about 0.7 grams of the monOprotonated, Clll-HX,i was obtained. The means by 37 which the borane adduct was reduced and the monoprotonated ligand formed is not known. The internally monoprotonated ligand with its un- identified anion was converted to the bromide by an ion exchange procedure using a Dowex l x 2 column. The resin was converted from the chloride to the bromide form by using an aqueous NaBr solution and testing the effluent with AgN03. 3 form by using precipitation followed by filtration through a small Clll-HBr,i was converted to the NO A N0 5 3 alumina column. Table 2 gives an indication of the solubility of several Clll complexes in many solvents. VI) METAL PURIFICATION Alkali metals were purchased as follows: Na: Alpha Division, Ventron Corporation (99.95%) K: Alpha Division, Ventron Corporation (99.95%) Rb: Fairmont Chemical Company Cs: Donated by The Dow Chemical Company All metals were stored under vacuum in pre-measured tubes using the "trombone" apparatus shown in Figure 10. The trombone consists of three Pyrex chambers separated by constrictions for sealing off, which allow the metal to be divided into three portions. These chambers also contain long glass tubes of premeasured inner dimensions into which the metals can be distributed. The volume of these tubes 38 Table 2. The solubility of C111 and its complexes in various solvents. Species Solvent Solubility Clll Water, MeOH, Acetone, Ether, Benzene v Clll°HCl,o Water v Clll-HCl,o Acetone i Clll°HBr,i Water, Acetone, MeOH v Clll-HBr,i Ether i Clll-HNO3,1 MeOH, EtOH s Clll-HNO3,i Acetone i Clll°2HCl,i-o Water, MeOH v C111°2HC1,i-o Ether, Acetone i Clll-2HCl,i-i Water, MeOH v Clll°2HCl,i-i Ether, Acetone i v - very soluble s - soluble i - insoluble is two to three times greater than that of the metal to be distributed. The metals were received in three to five gram am- poules under argon with a break seal for access. The am- poule was first flame sealed to the trombone into which a break bar had previously been placed. The entire system was then pumped to less than 1 x 10"5 torr, the break seal broken and the argon pumped out of the system. The metal was then distributed equally among the chambers which could be flamed sealed from each other to yield three separate systems. The metal was forced down the tubes by heating 39 .mampoe fiawxam no owmaopm on» non msumhmqam =mcon80he= .OH mhswfim A0 and gently shaking, then each tube was sealed off. In order to facilitate distribution of the metal into the tubes, it was imperative that the system had been pumped to high vacuum since any residual gas pressure would have blocked this transfer. Known volumes of metals were easily obtained from these tubes since their inside diameters had been measured. The proper amount of metal could then be isolated and sealed from the rest of the tube and introduced into the sample preparation vessel using heat-shrink tubing as shown in Figure 8. The tube was scratched with a glass knife and placed into a side-arm of the apparatus. The side-arm was then capped with a glass end made vacuum tight with flexible heat-Shrink tubing. The system could then be pumped to high vacuum (10-5 torr) and the metal tube broken by bending the heat-shrink tubing. The metal was then distilled into the vessel and the side-arm flame—sealed away. VII) VACUUM PROCEDURES Vacuum distillation provides a convenient method by which to purify and transfer air sensitive materials. Standard high vacuum procedures were employed throughout this work. Figure 11 Shows the type of vacuum line arrange- ment used. It was composed of two Pyrex chambers, one for high vacuum work and the other used exclusively for work which did not require high vacuum. These compartments were separated by a valve which permitted two preparations to be Al .COfiumpmamAQ mHQEwm CH poms ucmEmwcmLLm mafia Essom> .HH maswfim 9.55 fig on: corwaEu W 1 93:: am. m , Ii %% wxflaaman W 5w oz: 2:: mm a. mm 323. 0:30. A2 done simultaneously. Greaseless Teflon valves of the Kontes and Fischer-Porter type were utilized to eliminate the possibility of contamination by vacuum grease. Pres- sures of 10.6 torr were easily attained. The diffusion pump was separated from the manifold by a liquid nitrogen trap to prevent its contamination. All distillations were done through tees and not through the manifold in order to insure a clean and efficient system. VIII) NMR SPECTROMETER Two NMR spectrometers were utilized for the maJority of the analyses presented in this dissertation. Proton and carbon-13 NMR studies have been performed using a Bruker WH-180 super-conducting NMR spectrometer operating at a field strength of A.228 Tesla. The WH-180 operates only in the pulse-fourier transform (P-FT) mode and is equipped with quadrature detection which provides a A0% signal to noise enhancement. The spectrometer is completely computer controlled via a Nicolet 1180 computer package which provides versatile pulse and decoupling Options. Homo and hetero nuclear and gated proton decoupling, inter- nal deuterium lock, temperature control and multinuclear options are also capable with the instrument. The large field strength provides exceptional resolution and sensitiv- ity such that samples of less than 1 mg may be utilized for analysis. A3 The temperature meter of the WH-180 was calibrated on several occasions by placing an NMR tube of methanol with a thermocouple in the probe and allowing it to equili- brate. The temperature in the probe was measured by a calibrated Doric digital thermocouple and compared with the setting of the temperature meter on the spectrometer. A calibration chart was thus developed and periodically moni- tered for accuracy. The temperatures could be reliably measured to 13K with this procedure. Nuclei other than 13C and 1H have been analyzed on a greatly modified Varian DA-60 spectrometer which also oper- ates in the P—FT mode. It is equipped with external lock, a Nicolet 1080 computer system, a versatile multinuclear package patterned after the development of Traficante<37), and operates at a field strength of l.A09 Tesla. The tem- perature of the sample in the probe may be monitored to t2K by using a calibrated Doric digital thermocouple with the thermocouple affixed less than 20 mm from the sample tube. For further details of the design and operation of the DA-60 spectrometer see the Ph.D. dissertation of Joseph M. (38) Ceraso IX) SAFETY Many of the crown ethers have been reported to be highly toxic<39). l2-crown-A is absorbed through the skin and its vapors may be inhaled in sufficient quantity to cause severe inJury and death. Testicular atrophy, AA sterility, lack of coordination, convulsions and death have been observed in rats exposed to l to 6A ppm in a 6 hour day for one week. The toxicological studies of some higher crown ethers indicate that this entire class of mol- ecules should be treated with care. Caution should also be utilized when handling the cryptands although their toxicity has not been reported. The Chemistry Department safety guidelines should be consulted and understood before attempting laboratory work. These guidelines may be obtained from the Chemistry Business office. It should also be emphasized that safety glasses are a mandatory protection for work involving vacuum tech- niques. CHAPTER 3 CONFORMATIONAL ANALYSIS OF CRYPTAND lll I) INTRODUCTION The fact that cryptands bind ionic species is easy to demonstrate, but distinguishing how and where the interac- tion occurs is a much more subtle problem. Cryptands, for instance, may exist in a number of conformations and possess several potential binding sites. As shown in Figure 12, the free ligands are thought to exist as a mixture of rapidly interconverting isomers in solutionmo pomasoo Am smn A< .HHHO ho meUoE wcfiHHHM momam .MH whswdm one Azufi>Mo owuwav appoeszm C“) {if < /\Z $2) J C (V) 50 cm (CV/v b N J pd NEW 3 g S: CL C° Q,. (\o/N N-H' cw <\? l\ /\ Figure 1A. J C {i H ’\ Z 0 0 K4,: J _2 I «90 K {(43 C) N H N 3 A“)? O The possible binding sites and conformations of Clll, A) unprotonated, B) external protonation, C) internal protonation, D) external protonation of the internally monoprotonated complex and E) internal diprotonation. 51 6y oy Cc 9 6y 6 y CNN/y é my Figure 1A (cont'd.) 52 participate in proton binding. Figure 1A shows the more logical structures, although the exo-exo conformations may be too strained to exist. Protons probably bind externally at the nitrogens, since in solution the latter are much stronger bases than are the ether oxygens. The latter, how- ever, should not be reJected as viable sites, especially after one nitrogen has been externally protonated. Inside the cavity, the situation changes very drastically, because the ligand shields the inclusively bound proton from the sol- vent and the donicities of the polar groups would be expected to be very similar to those in the gas phase. Taft and co- worders have tabulated gas phase bacisities of several amines and ethers<57'58), which demonstrate the similarity in base strength of these groups in the gas phase. Therefore, both the nitrogen atoms and the oxygen atoms might be expected to contribute to the binding inside the cavity of Clll. Finally, torsional motions may also contribute to the strength of proton binding. The larger cryptands, for instance, conform to small ions bound inside, in order to optimize the interaction with the donor atoms. This optimization involves a concerted rotation of the nitrogens so that the angle, a, shown in Figure 15, can be large for small ions or small for the larger ions to provide conditions more conducive to bind- ing. Figure 13 depicts models of Clll in both configurations. The one in which a is large, B), might be expected to be most stable because in this conformation, the nitrogens are closer 53 Figure 15. Model showing the torsional angle of C222. to the Hi' This conformation would also be expected to greatly shield the inclusive proton from the solvent. A) Clll Symmetry Considerations Nuclear magnetic resonance spectroscopy is admirably suited for the identification of the nonequivalence of nuclei which develops due to the asymmetric configurations. This nonequivalence is observed in both the proton chemical shift and in spin-spin coupling constant. For example, geminal protons (those on the same carbon) have identical chemical shifts and do not couple with one another if they are chemi- cally equivalent, but if for some reason they become non- equivalent, they do couple to one another and their chemical shifts are no longer identical. A very frequent cause of geminal nonequivalence, and one which occurs often with the cryptands, is the slowing of molecular vibrations at low temperatures, which "locks" the geminal protons into slight- ly different chemical environments (on the NMR time-scale). Nuclear magnetic resonance also provides characteris- tic information about the electronic environment of the bind- ing site from the chemical shift and thereby facilitates the 5A identification of the functionality of that site. The relative stabilities of various sites may be obtained from the populations in those sites and the dynamics of exchange between sites may be obtained from line broadening. The theory of line broadening in NMR has been thoroughly devel- oped by Bloch<59'60), so that the NMR spectra may be math- ematically simulated in order to define the exchange rates associated with spectra at various temperatures (see the Appendix for a complete derivation). The temperature depen- dence of the exchange rates which, in principle, should ad- here to the Arrhenius theory, provides activation parameters for exchange. The interpretation of the NMR spectra of Clll and its complexes is greatly simplified if the symmetry of the mole- cule is understood. As can be seen in Figure 16, Clll is a highly symmetric molecule and belongs to the D3h point group. The most useful symmetry elements for interpretation purposes are the C3 rotational axis through the nitrogens, the Oh plane through the oxygens, and the three ov planes which in- clude the C3 axis, a polyether strand, and bisect the other two strands. It is simplistic to assume that the av symmetry exists, because gauch conformations are more energetically favored than are the eclipsed conformations which are neces- sary for 0 symmetry. When rapid methylene wagging motions v occur, so that the interconversion between the two gauche forms is rapid, the 0V symmetry will be presented if the "average" conformation preserves this symmetry. This 55 .coapmhswfimcoo nmo on» CH HHHO no mucmEon maumEEmm on» wcfizozm Hmooz NV (AN \ / .wa mhswfim 56 interconversion must proceed in milliseconds or faster in order to present an exchange averaged picture in the NMR spectrum, since the maximum exchange rate observable is on the order of the chemical shift difference of the two sites. This difference rarely exceeds 1,000 Hz. Therefore, if an exchange occurs, whether it be a proton Jump or interconver- sion of two conformations of the ligand, and is more rapid than milliseconds, the NMR spectrum will present a single line (or a multiplet if scalar coupling is present) which is the population average of those sites involved. The chemical shift of this exchange averaged line is defined by: 6obs = i xi 61 (22) where: 5obs - the observed chemical shift 61 the shift in site 1 x1 the mole fraction (the population) in site i The room temperature spectra of the Clll complexes Show that the av symmetry element is preserved because interconversion rates are faster than the NMR time-scale and the exchange averaged conformation preserves this symmetry. II) NMR SPECTRA AT AMBIENT TEMPERATURES At room temperature, molecular motions are rapid and the NMR spectrum of a given species usually reflects an ex- change average of several different conformations. This is true of the spectra of Clll and its complexes shown in Figure 17 and the chemical shifts which are given in Table 3. The 57 A B o b C o b “SOLVENT LkH D 4.00 3.50 3.00 2.50 6 ppm Figure 17. The NMR spectra of C111 and its protonated com- plexes at ambient temperature. 58 mm.» =m.m mm.m mommm moosazo mum.aomm.flmmo mm.m om.m mm.m mommm momz-:o “.mmm.mmmo mo.m :m.m mm.m mommm mcoumoaom momooam .onsumnmanu unmanem pm mmxoaasoo mum ond HHHU mo mpuficm awedscno one .m mHQMB 59 most striking feature of these spectra is the dependence of the chemical shift of the backbone protons on protonation. The CH20 proton lines shift downfield by about 0.3 ppm per proton added and the CH N proton lines shift by about 0.7 2 ppm per proton, more than double the CH O proton shift. This 2 downfield shift is caused by the delocalization of the positive charge of the bound proton throughout the backbone of the mol- ecule, which is an electric field contribution to the chemical shift. The fact that the CH2N protons show a greater shift is indicative of greater interaction of the internal proton with the nitrogens than with the oxygens. Assignment of the binding interactions should be done with care because these spectra represent the average of many rapidly interchanging molecular conformations. Nevertheless, some qualitative assignments may be made from them. Three of the four structures, Clll, Clll°H+,i and Clll-2H+,i-i, preserve the D symmetry since the spectra show equivalence of the 3h CH20 protons as well as of the CH N protons. The fourth struc- 2 ture, Clll-H+,o, has lost at least some molecular symmetry, because these two sets of protons are no longer equivalent. A discussion of the three symmetric complexes will be presented first, focusing on each structure individually, and the exter— nally protonated ligand will be discussed thereafter. A) Clll in d6-Acetone The NMR spectrum of Clll contains two sets of triplets, one set at 3.AA ppm (CH20 protons) and the other at 2.52 ppm (CH2N protons). The two sets of protons split each other 60 with a coupling of 5 Hz. Figure 1AA shows three possible conformations of the ligand. From the room temperature spectrum, one is not able to conclusively establish their presence except that, if exo-endo nitrogen inversion or any other conformational change is occurring, it is doing so rapidly on the NMR time-scale, and its average preserves D3h symmetry. Based on an examination of molecular models, the exo-exo form is not expected to exist, but a rapid exchange between the endO-endo and exo-endo forms would preserve the D3h symmetry. Rapid vicinal carbon wagging vibrations must also be occurring in order to retain the av plane through the polyether strand, which is necessary for the geminal CH2 pro- ton equivalence. This vibration entails exchange of the vic- inal carbons between the gauche conformations whose average is the eclipsed conformation. B) Clll-H+,i-Br- in d6-Acetone The internally monOprotonated ligand shows D3h symmetry at room temperature for the same reason as does Clll; namely, rapid interconversion between the conformers. The CH20 pro- tons are equivalent at 3.76 ppm as are the CH2N protons at 3.2A ppm. These two sets of protons again split each other into triplets of 5 Hz, and the inclusive proton, H , also 1 couples to the CH2N protons with a coupling constant of 2.5 Hz, resulting in a six line pattern. The H1 resonance is located at 9.07 ppm with a linewidth of about 15 Hz. No coupling is observed to the nitrogens or oxygens by any of 61 the protons or carbons, presumably because of very fast quadrupolar relaxation, but the breadth of the H1 signal (approximately 15 Hz) and the backbone proton lines (approxi- mately 1.5 Hz) probably originates from efficient scalar re- laxation by the nitrogens<62). No coupling of the R1 to the carbons is observed in the C13 NMR spectrum, but it would be expected to be about one fifth that Of the CH2O proton split- ting and would not be observable(61). The fact that coupling of the H1 proton is observed to the CHZN protons and not the CH20 protons is good evidence that the H1 proton interacts most strongly with the nitrogens and not the oxygens. Coupling is transmitted through a chemi- cal bond via Fermi contact interactions<62), and the magnitude of the coupling interaction is similar through the protonated ether oxygens and the protonated amines. Three bond coupling constants of about 5 Hz have been documented for ethers pro- tonated with super acids and for protonated amines(63'5u). Therefore, if the ether oxygens did interact appreciably with the Hi, we would expect to observe coupling to the CH20 pro- tons, although attenuated slightly since there are three oxy- gens, rather than only two nitrogens. The 2.5 Hz coupling of the H1 to the CH2N protons agrees with literature values for protonated amines. The observed coupling is half the litera- ture value since the H1 is shared equally between the nitro- gens. A structure which would be consistant with the nitrogen protonation as well as the preservation of D symmetry is 3h 62 shown in "a" of Figure lAC. The proton Jump between the nitrogen atoms is always rapid. The exo-endo inversion is probably hindered by coulombic interactions between the Hi and the nitrogen lone pair, and therefore occurs very infre- quently. The internal proton exchange phenomenon is very unusual, since it occurs much more rapidly than the NMR time-scale (since it preserves symmetry), and yet coupling to the CH2N protons by the H1 is not lost. Normally, when proton exchange is more rapid than the NMR time-scale, coupling between the exchanging protons and those attached near the binding site is lost. This loss occurs because protons of either spin state have an equal probability of binding at the site and the NMR spectrum presents a single line which is an exchange average of the two. With Clll-H+,i, the spin character of the exchanging proton is preserved between Jumps (ignoring Tl relaxation which is slow by comparison), because the same proton exchanges back and forth between the two nitrogens. The H1 is not able to ex- change with protons in solution, but is restricted to move only within the confines of the cavity of Clll. C) Clll-2H+,i-i-2Cl- in d“ MeOH The internally diprotonated ligand must exist in the endo-endo form as a prerequisite for D symmetry, due to 3h the size restrictions of the cavity. The H1 protons must be located on the C3 axis, each equidistant from the center of the cavity. Their distance from the nitrogens is probably 63 less than in the monoprotonated ligand, because of the coulombic repulsion of the two protons. As'a result of the symmetry, the spectrum, as seen in Figure 17D, presents the equivalence of all CH20 protons at 3.96 ppm and CH2N protons at 3.7A ppm, the two sets splitting each other into triplets with a coupling constant of 5 Hz. The H protons, 1 whose resonance is located at 7.33 ppm, couple to the CH2N protons (but not to the CH O protons) with a coupling of 2 5 Hz, producing two sets Of overlapping triplets which re- sult in a four line multiplet. Coupling, which originates from the Fermi contact interaction, is a function of the ability of the nitrogens to donate electron density to the H1. The introduction of a second proton into the cavity of Clll-H+,i would be expected to double the coupling through the nitrogen, because each nitrogen binds to one full proton rather than sharing one between them. Thus, the contribution of electron density from each nitrogen doubles. This is exactly what is observed, since the coupling to the CH2N protons increases from 2.5 Hz to 5.0 Hz with the addition of the second internal proton. When one Of the two internal protons is replaced by a deuteron which has a coupling constant one sixth Of that of a proton (less than 1 Hz for three bond coupling), the NMR spectrum of the CH2N protons indicates two overlapping pat- terns which may be distinguished by decoupling experiments. One pattern shows coupling of 5 Hz to the H1, whereas the other does not, a result of negligible exchange of the proton 6A and deuteron between the nitrogens inside the cavity. This complex probably exists in the configuration displayed in Figure 13A with the cavity as large as possible in order to allow maximum separation for the H1 protons. The upfield shift of the H protons in going from the i mono to the di internally protonated ligand is a consequence of two opposite contributions. The size limitations of the cavity of Clll tend to restrict the donation of electrons from the nitrogens to the internal protons because, as they gain electron density, they also increase in size. This contribution induces a downfield shift. The increase in charge density inside the cavity greatly overshadows the for- mer contribution, and induces a large electronic shielding, which in turn produces the Observed upfield shift. The linewidth of the H1 is dominated by scalar relaxa- tion to the nitrogen, and on the basis of a twofold increase in coupling constant from the mono to the di internally pro- tonated ligand, the linewidth of the Hi would be predicted to increase by a factor of four between the two complexes on the basis of the Fermi contact term. The observed line— width of the H1 in Clll-2H+,i-i is about 50 Hz, which is in the range of the predicted values (A8 to 60 Hz). The line- width decreases with decreasing temperature because the re- laxation rate of the nitrogen increases and thereby causes the scalar interaction to be less efficient<62). All assign- ments have been verified by decoupling experiments wherever possible. 65 The room temperature spectra indicate that the H1 protons interact at the nitrogens, and that the shifts pro- duced for both the skeletal protons and the inclusive pro- tons upon protonation are a result of the electric field contribution to the chemical shift. For this reason, Musher's equations, which predict the shifts resulting from changes in charge distribution, should prove valuable in understand- ing the observed shifts. D) Musher's Theory of the Chemical Shift<30> Musher's equations describe the chemical shift induced by the electric fields associated with charged or polar groups. The magnitude of the field is proportional to the charge density and the distance of the bond from the charge, as described in the Historical section. In order to calcu- late the shift expected upon protonation of 0111, a geometry for the ligand must be assumed. The nuclear coordinates have been calculated utilizing D symmetry, and the assumption 3h was made that the center of positive charge was at the nitro— gen. The theory predicts a downfield chemical shift of A.l ppm and 1.0 ppm respectively for the CH N and CH20 protons 2 upon the introduction of two protons into the cavity of Clll. The observed downfield shifts are l.A and 0.6 ppm, which are in the right direction and relative magnitude, but much smaller than the theory predicts. The theory predicts an upfield Shift of about 2.6 ppm for the H1 in going from Clll'H+,i to Clll°2H+,i-i, and again it is in the correct direction but too large (the real shift is 1.A ppm). The theory applies 66 to gas phase interactions, so that solvent interactions with the ligand may cause the discrepency between the theory and the observed shifts, but it lends credence to the belief that protonation occurs at the nitrogens, and is also able to account for the entire shift. E) Clll-H+,o in d6-Acetone The NMR spectrum of Clll-H+,o in d6-acetone acidified with trifluoracetic acid (TFAA) (Figure 178), reflects the loss of molecular symmetry, because four lines of equal area are present rather than two. (The multiplet at about 2.5 ppm is a C13 satellite of acetone.) There are two plausible sites for external protonation, shown in Figure lAB, one at an exo nitrogen and the other at an oxygen. Only nitrogen protonation, with the loss of Oh symmetry, could give rise to four lines of equal intensity. Protonation of the oxygen atoms would not give lines of equal intensity. Therefore, the resonances a and b correspond to CHZO and CH N protons 2 on the protonated side of the molecule and a' and b' to the unprotonated side. Assuming these assignments are correct, the chemical shifts of the skeletal protons are informative in determin- ing the contribution of exo-endo conformational changes to the chemical shifts of the Clll complexes. The Clll'2H+,i-i complex and C111 probably exist in the endo-endo form. Those resonances from the protonated side of Clll-H+,o, namely a and b, are very similar in their chemical shift to those of Clll-2H+,i-i, whereas those on the unprotonated side, a' and 67 b', are very similar to those of Clll. Therefore, even in the presence of a greatly different conformation, the chemi- cal shift of the backbone protons is dominated by nitrogen protonation and the contribution from the conformational change is less than 0.2 ppm. Perhaps the most striking feature in relation to exter- nal protonation of C111 by TFAA is the lifetime of the com- plex, being greater than milliseconds. The spectrum of this complex demonstrates that the exo-endo form of Clll exists in solution. The theory of Eigen concerning general proton-transfer processes predicts that when the strength of an acid is much different than that of the conJugate acid of the base used in proton-transfer reactions, the rate of transfer is fast. The transfer also is complete, favoring the formation of the species with greater pKa(65) and the reverse reaction is slow. If the relative strengths are similar, however, the proton is rapidly exchanged between the two, but the transfer is incomplete. Eigen has shown that a proton-transfer reac- tion may be written as follows: K1 K2 K3 AH+B:AH--~B:A~~HB:A+HB (23) where AH-ooB and A-ooHB are encounter complexes. When the reaction favors one set of products, say A + BH, as is the case involving protonation of Clll with TFAA, the proton is rapidly transferred and the resulting complex is long-lived. This can give a long enough value of the exchange time, T, to 68 permit observation of both HB and B when less than a stoichiometric amount of HA has been used. However, if the values of pKa and pr are similar, fast proton exchange and an incomplete transfer occurs. I The external protonation of Clll has been studied by NMR as a function of acid strength in d6-acetone and dA- methanol. Three acids were utilized; namely, HCl, trifluoro- acetic acid (TFAA), and dichloroacetic acid (DCAA), each of which appeared to interact differently with Clll. HCl was generated by addition of about 0.1 m1 acetyl chloride to about 10 ml of du-methanol and allowing time enough for the reaction to proceed to methyl acetate and HCl. This method provided an internal standard for a quantitative determina- tion of acid content. When one drop of this solution was added to a solution of Clll in du-methanol, a white powder immediately precipitated, and the NMR signal of Clll was no longer observed. Presumably the acid formed the monohydro- chloride salt of Clll, the proton being bound externally and the resulting complex was insoluble in du-methanol. This sample was immediately cooled to -70°C, after which some yel- lowing of the solution was noted. Addition of KOH cleared the solution and caused the precipitate to quickly redissolve. Addition of TFAA to a solution Of Clll in d6-acetone at 270 K caused no precipitation, but a dramatic change in the NMR spectrum was observed compared to that Of the free amine, as shown in Figure 18. Addition of less than one equivalent of acid caused immediate production of a 69 b+ b“ 1 1 1 oppm 4.00 3.50 3.00 Figure 18. The dependence of the NMR spectrum of Clll in d6-acetone upon the addition of trifluoroacetic acid (TFAA). 70 monOprotonated form of the ligand. This externally pro- tonated form exchanges only slowly with the free amine (as evidenced by slight line broadening) via the following scheme: 4. .. Clll + HTFAA:;::::Clll-H ,O + TFAA (2A) but the proton transfer greatly favors the C111. As TFAA is added, lines a', b', a", and b" grow as a and b (those of the free ligand), disappear (an impurity peak overlaps the signal, a'). The new lines are consistant with the structure "b" of Figure lAB. As mentioned previously, the a symmetry h is removed, thus the exchange scheme (2A) must be slower than the NMR time-scale and highly favor the monoprotonated form. As more than one equivalent of acid is added, the downfield CH2N line shifts farther downfield and eventually coalesces with the CH 0 line at A.00 ppm. Continued addition 2 of acid causes the upfield resonances to disappear and that at A.00 ppm to grow. At a mole ratio of about 50/1 acid to Clll, the sample is almost completely converted to this new form, since the resonance at A.00 ppm accounts for nearly the entire area of the resonances. This new form is probably a diprotonated species, which has protons bound externally at both nitrogens, the ligand being in the exo-exo form. A logical reaction scheme is shown below: + 0111 + H+-—.: Clll-H ,o (25) Clll-H+,o + I1+;_=-__ Clll-2H+,o-O (26) 71 Reaction (25) goes to products quickly and completely as mentioned earlier, but with excess acid, reaction (26) be- comes important. The first external protonation is stable and the transfer is fast and complete, whereas the second exo protonation is much more difficult to achieve. Therefore, the proton is exchanged rapidly and the equilibrium needs to be forced to the right by a large excess of acid. External oxygen protonation is ruled out as a maJor contributor to the second external protonation, because the CH O protons 2 would show a much larger downfield shift than the CH N pro- 2 tons. The fact that the CH20 and CH2N proton lines coalesce at A.0 ppm with excess acid indicates that the interaction involves exo-exo nitrogen protonation. External nitrogen protonation is also necessary to account for the very large shift in the CH2N proton resonances as acid is added. The identical chemical shifts of the CHZO and CH2 this externally diprotonated species seems to be accidental. N protons in The acid concentration was determined by the comparison of the exchangeable proton line of TFAA (the —OH signal) with those Of the ligand. Considerable error is inherent in this method, since at least two acid proton signals are present, that bound to the ligand and that bound to the TFAA anion. The latter line is very broad and the bound proton line often overlaps the ligand signals. Water also interferes. This makes the measurement Of the acid concentration very difficult, but general trends have been established. Table A lists the chemical shifts of the resonances as a function of TFAA present. 72 Table A. The dependence of the chemical shifts of Clll in dé-acetone upon addition of trifluoroacetic acid ( FAA). MR %§%%.(tuoz) ogggo 633 N A 0.0 3.u7 2.56 B 0.7 3.99, 3.65 3.76, 2.63 p 1.u u.01, 3.65 3.88, 2.63 E 1.7 u.o3, 3.66 3.96, 2.6A F 7.6 u.o2, 3.66 n.02, 2.6A 6 8.A u.01, 3.70 u.01, 2.67 H 29 3.98, 3.73 3.98, 2.77 I 50 3.91 3.91, 2.78 Dichloroacetic acid, DCAA, has been utilized to better determine the amount of acid added, since it has one non- exchangeable proton. As shown in Figure 19, the addition of DCAA gives very different NMR spectra than does the addition of TFAA. Proton exchange between the free and the monopro- tonated amine at 270 K is much more rapid. As DCAA is added, the lines broaden and shift downfield, the CH2N protons show- ing a more dramatic effect than the CH O protons. A plot of 2 the chemical shift of the CH O protons as a function of total 2 acid added, as shown in Figure 20 and Table 5, gives a break in the slope at a mole ratio Of 1:1 acid to Clll indicating efficient formation of the monoprotonated form ( as with TFAA) and a difficult second protonation. The break occurs at a chemical shift of 3.79 ppm, which is midway between the values 73 M. U13 M 01 M‘ a 1 6ppm 500 4100 3.00 4100 300 Figure 19. The dependence of the NMR spectrum of Clll in d6- acetone upon the addition of dichloroacetic acid (DCAA). 7n .< 6.5 L4 40 is 3'0 4'0 35 30 1 J 1 A 4.0 H 40 3.5 32) 40 3'6 3b 25 Figure 21. The NMR spectra of Clll in water at various pH values. 79 .O N a as ammo mo msoeoao z m so one mo smash Hoossoso on» mo oosoosoaoo ma one .mm osswam ccoamw 00.0 on.” Z~IU 00m ond. _ a 00 a /o( o d / o/ d o / H \- mom 0 I. 80 Table 6. The pH dependence of the chemical shift of the CH2N protons of Clll in water at 299°K.* pH? CHQO CH2N ppm ppm 9.0 ---- 2.53 8.7 ---- 2.69 7.5 3.78 3.05 6.5 3.85 3.21 A.9 3.82 3.13 A.0 3.81 3.19 3.2 ---- 3.23 2.6 ---- 3.27 1.8 ---- 3.A0 1.A 3.79 3.AA 1.1 3.81 3.A9 * Reference to DSS., those absent in the CH20 column are due to overlap. I Corrected for deuterium effects. the diprotonated species is not complete, even at a pH value of 1.0. Comparison with the shift in acetone indi- cates only about 50% conversion. The average shift of the CHZN protons of Clll-H+,o in acetone is very similar (10.1 ppm) to that in water (arrow "a" of Figure 22). Thus a chemical shift for the CH2N protons of Clll-2H+,o-o of 3.9 ppm (arrow "b") may be assumed, from which a pKa value lower than 2 is obtained for the second external protona- tion. The pH values of all buffers were measured (and 81 corrected for deuterium influence)(67) with a calibrated radiometer and microelectrode. A pH titration of Clll was also performed in water. Only one break was observed, yielding pKa = 7.1 i 0.1 for the first external protonation. The titration went only to a pH value of 1.0, which was probably still too high to per- mit observation of the second break. From the pH dependence of the NMR chemical shift, the second pKa is estimated to be between 0 and 1.5 in water. The pKa values Of C111 and its complexes reflect the presence of the exo conformation, since only this form may be externally protonated. 1,A diazabicyclo (2-2-2) octane, or Dabco, is a carbon analogue of the cryptands, whose small size requires both nitrogens to be in the exo conformation. The pKa values of the cryptands would be expected to be very similar to that of Dabco, 8.8 i 0.2 at 25°C, if the molecule had no endo conformation present. Proton transfer reactions of the cryptands actually involve a two step process: + Ka + K1 Clll-II ,0:H + Clll (endo-exo): Clll (endo-endo) (27) and the apparent Ka of C111 is equal to the following: app _ Ka - (Ka) (28) Assuming that the value of Ka for external protonation of Clll in the exo form is equal to that Of Dabco, 1.6 x 10'9, 8 the observed value of Kaapp, 7.9 x 10 M, gives K1 8 50. 82 Therefore, the endo-endo form is favored over the endo- exo form by about 2 kcal mol'"1 in the absence of acid and the endo-endo form accounts for about 98% Of the ligand at room temperature in the absence Of acid. For the second external protonation, if we assume that Kaapp is 10"2 or I less, we obtain K > 10+7, establishing a difference of at 1 least 6 kcal mol'1 between the two forms. The value of K1 represents the analog of K in equation (27) for the mono- 1 protonated ligand. 0) The External Protonation of Clll-H+,i In water and in methanol, the chemical shifts of the CH2O and CH2N protons are strongly dependent on the strength of the added acid and on the temperature, indicating that the exo-endo diprotonated complex is present in equilibrium with the internally monOprotonated form. The spectra in du-MeOH will be discussed in detail in the next section, but the results of this analysis show that the exo-endo inversion and the proton exchange rates are very rapid. A pKa value of 3 is obtained for external protonation in methanol. The pH dependence of the NMR spectra of Clll-H+,i in water indicates that an exchange similar to the one Observed in methanol is occurring. A compilation of the pH dependence of the chemical shifts Of the CH2O and CH2N proton lines of Clll-H+,i in water is shown in Table 7. It we assume that the chemical shift for the CH2N protons of C111°H+,i is 3.0 ppm and the shift of the same protons of Clll'2H+,i-O is 3.A5 ppm (from Table 7), a value Of 0.A i 0.5 is obtained for 83 the pKa of the process. This pKa value indicates that the preference for the endo-endo form is higher in water than in methanol by two orders of magnitude. Using the Ka of 8 in water Dabco, an equilibrium constant, Kg, of about 10 and about 106 in methanol is Obtained for the ratio of the endo-endo form to the endo-exo form. Thus the endo-endo form is the only significant species in solution in the ab- sence of added acid. The endo-exo form of Clll-H+,i seems to be as difficult to attain as the exo-exo form of Clll. Table 7. The pH dependence'og the chemical shifts of Clll-H+,i in water 5 . Acid Strength CH2N CHZO ppm ppm 3N 3.A5 3.82 1N 3.3A 3.76 0.AN 3.20 3.72 0.1N 3.10 3.70 0.01N ‘ 3.03 3.66 0.0 3.02 3.66 H) General Conclusions 1) External protonation occurs at the nitrogens in both C111 and Clll-H+,i. 2) Internal protonation would be expected to also in- volve oxygen interaction, but the NMR evidence suggests other- wise. 8A 3) The chemical shifts of the CH20 and CH2N protons upon protonation are determined mainly by electric field effects caused by the protonation at the nitrogen and not by conformational changes. A) Exo-endo inversion is rapid at room temperature, but favors the endo-endo form especially in Clll-H+,i. 5) Exo-endo inversion does not take place in Clll-2H+,i-i. 6) The solvent and acid used, greatly influence the extent of external proton binding and the exchange rate but have no observable effect on inclusively bound protons. III) TEMPERATURE DEPENDENCE OF THE NMR SPECTRA The time-scale of proton NMR is relatively slow since it is only able to distinguish between processes which occur more slowly than milliseconds and, in fact, is limited in the C111 systems to exchange times longer than about 0.01 seconds. Exchange phenomena which occur more rapidly than this yield only a population average signal, and the charac- ter of the individual site is lost. One of two things may be done in order to identify individual sites when exchange times are more rapid than 10'3 seconds. First, an analysis method which employs a faster time-scale may be utilized, such as IR, UV or EPR spectroscopy if applicable. The other alternative is to continue using NMR methods, but at lower temperatures. A change in temperature affects more factors than Just the rates of exchange processes, however. It also 85 alters the populations Of the sites involved in the exchange in accord with Boltzmann statistics, and changes such physi- cal properties of the medium as the viscosity and the di- electric constant. The enthalpy and entropy of a binding site are temperature invariant if external influences of the medium remain constant. Therefore, going to lower tempera- tures does not cause stronger binding (unless the site is perturbed externally) although the lifetimes at sites do become longer, as governed by the Arrhenius or Eyring equa- tions(68)- The equations which describe chemical exchange are well characterized for NMR. McConnell modified the Bloch equations to describe chemical exchange processes and ob- tained equations which are applicable to phenomena invOlving single lines and first order multiplets(59-6O). These equa- tions are developed in a simplified form in the Appendix, and may be easily utilized in processes which involve fewer than four or five sites. For systems of more sites, or second order spectra, a density matrix approach would be more suit- able(69). A) Clll in d6-Acetone The temperature dependence of the proton and carbon-l3 NMR spectra of Clll in CHF2Cl has been reported by Lehn and (26) co-workers at 250 MHz. The C13 NMR spectra were tempera- ture-independent down to 178 K except for slight line shifts, but the proton spectra at 163 K show a well resolved ABXY pattern for the backbone CH2N and CHZO protons with chemical 86 Ammv um: 0mm um x mmH um Ho N Emu :H HHHU ho Sahpomam 0:9 .mm enemas 87 shifts and coupling constants of 3.67 (A), 3.32 (B), 2.58 (X), and 2.A7 (Y); J = 10, JXY = 1A, J = 10, AB BX JAx = JBY AYmi 2 Hz as depicted in Figure 23. Since the pattern was of the ABXY form, the conformation at low 2 J temperatures did not involve exo-endo isomers, because the Oh symmetry element was still intact. Instead, the process was described as the slowing of a torsional motion, such as vicinal carbon wagging motions, concerted twists of the strands, etc. The rate of this torsional motion was obtained at the coalexcence temperature of -65°C, which yielded a AGI of 9.8 i 0.1 kcal molml (from the Eyring equation). The torsional motion was assigned to a possible concerted process, such as the passage of a strand through the intramolecular cavity, since the activation barrier for the process was rel- atively large. Typical vicinal carbon wagging motions possess activation enthalpies of 3 to 6 kcal mol-1 , depending on sub- stituents, whereas concerted processes normally yield an ac- tivation barrier which is roughly the sum of the individual components. Therefore, a concerted twist would be expected to have an activation barrier of 10 to 18 kcal mol-l, and the value reported by Lehn of 9.8 kcal mol”1 would put it in the range of concerted processes. (The free energy of activation has no direct physical meaning, since it is the sum of AH; and -TAS¥, both of which can be given physical meaning. There- fore, in order to equate the AG1 obtained form coalescence data with the activation enthalpy, AS; must be assumed to be zero.) 88 .m:ouoomnoo :« Haao mo onuOOQn mzz on» no consummate waspmnmaEOp 029 .:m oaswfim Edam MN om mm.” mm OM mm nm 0.4» mm om. CON CNN nNN : ))<fi 0.0 89 Figure 2A shows spectra for the same process in d6- acetone at 180 MHz which have very similar features, except that they are more poorly resolved, due to a lower field strength, higher temperature and a more polar solvent. The assignments are the same as those at 250 MHz (see Table 8), but the separations due to nonequivalence are not as pro- nounced in acetone as they are in CHF2C1. From the coales- cence temperature of -58°C for the process in acetone, AG* = 10.7 1 0.3 kcal mol’l, in agreement with Lehn's results. Table 8. The temperature dependence of the NMR parameters for a 0.02 M solution of Clll in d6—acetone. Temp (°K) 6C1120 6CH2N ppm ppm 290°K 3.AA 2.52 190°K 3.A9, 3.39 2.52 B) Clll-H+,i°Br' in d6-Acetone The temperature characteristics of Clll-H+,i (Figure 25 and Table 9) are more striking than for the free amine. As the temperature is lowered, both CH20 and CH2N proton lines broaden, the CH20 proton lines broaden more extensively be- cause their chemical shifts in the two conformers are farther apart than are the CHZN proton shifts. The CH20 lines split at about 215 K, and the CH2N lines at about 207 K. A very similar coupling pattern emerges with the spectrum at 195 K, compared to that of Clll at 163 K from Lehn's publication. This pattern is also an ABXY spectrum with lines and coupling 9O .o:oaoomloo :H H.+m.HHHo mo whuomdm mzz on» no cocoocomoo snowmaoqeou 0:9 .mm mpswfim Eoomomn on ed on _ . MW ndv 0m” _Mn 0d. mm. \\\\1 \llMMKY\/\>fi/i\\\. mm. 9 J 13 CON CNN CNN ¢ 91 Table 9’ 3h?.3fimfiefiii‘éiioffin‘éii‘fingfi1535-11343 323222303} H + Temp (°K) 6CH20* 5CH2N* 5H1* Av1/21 (hz) ppm ppm ppm 290 3.76 3.2A 9.06 10 270 3.7A 3.22 250 3.72 3.21 230 3.71 3.19 220 3.71 3.19 215 3.83, 3.59 3.19 210 3.85, 3.50 3.21, 3.13 200 3.86, 3.A6 3.2A, 3.11 9.0 70 195 3.85, 3.A6 3.23, 3.08 190 3.83, 3.A3 --, 3.07 9.1 150 185 .___ ___ 180 3.82, 3.36 -————, 3.09 8.3 150 175 8.0 110 170 A0 i Dashes indicate disappearence of no data available. Chemical shifts are reported i 0. Uncertainty in the linewidths is t 3 ppm- a signal, blanks indicate 92 constants of 3.85 (A), 3.A6 (B), 3.2A (X), and 3.08 (Y); JAB = 10 Hz, JXY BX = 10 Hz, JAX = JBY = JAY i 2H2. The preferential vicinal coupling, (JBX>>J = lO-lA Hz, J AX’ J JBY)’ explains the unusual breadth of X. These assign- AY’ ments demonstrate again that exo-endo isomerism does not contribute to the observed spectra and that the nonequiva- lence is caused by the slowing of torsional motions as the temperature is decreased. The fact that the CH20 protons broaden and split to a larger extent than do the CH2N protons signals indicates that the CH O protons are more nonequiva- 2 lent in the gauche conformation. The carbon-l3 NMR (CMR) spectra Of the same sample are temperature invariant, except for line broadening below 200 K. Carbon—13 NMR would be expected to be very sensitive to exo- endo nitrogen processes. Since the carbons retain their equivalence even at very low temperatures, the CMR spectra provide evidence that this inversion process either does not occur or remains very rapid. The lineshapes of the CH O protons and the CH2N protons 2 have been simulated using the modified Bloch equations. The simulation of the CH20 proton signals was done using line- width values of 0.3 Hz for both sites and 5.0 Hz for both sites in the absence of exchange. Either choice produced nearly the same calculated lineshape. A chemical shift dif- ference of 76 Hz between the two sites was utilized. The lineshapes were simulated for several values of 1, the ex- change time, which in this case, is the rate of the torsional 93 .ocouoomnmo :H H.+m.HHHo no mcouoaa ommo on» no opuooam on» no oococcoaoo onspwpooEop on» no coauwfisefimIAOpsoEoo .mm oaswam Edam on 0.... II .I. | o.» co can a: om... fl _ fl _ q _ l x mm. 0. com 58 x mm. xo- x OR 9A motion. The simulated lineshapes are shown in Figure 26. As this motion is slowed down, the lines go from a very sharp triplet to a broad singlet, and then split again when the exchange time is greater than 3 msec. The linewidths and general appearance of the experi- mental and calculated spectra were closely analysized in order to best match the simulated spectra with the real spectra shown in Figure 25. The rate of exchange, k, was then Ob- tained at various temperatures and an Arrhenius plot made by graphing 1n (k) versus l/T (Figure 27). This plot gives a straight line which yields Ea = 9.A t 0.5 kcal 1001""1 down to about 210 K. Then the slope deviates as another process be- gins to interfere. This latter process causes line broaden- ing in excess of that predicted for the slowing of a single exchange process. The exchange process which causes the line broadening above 210 K is probably due to the sloWing of a torsional motion. From the coalescence temperature (215 K and 207 K for the CH20 and CH2N proton signals), rate constants of 169 sec’l and 60 sec"1 are obtained, with A07 = 10.3 t 0.3 kcal -1 1 mol , AHI = 9.0 i 0.5 kcal mol' , and A87 = -6 cal mol-l deg‘l. These values are somewhat large for typical vicinal carbon wagging motions, but the bicyclic nature of Clll prob- ably causes these vibrations to be coupled throughout the rest of the molecule. The concerted exchange process which Lehn suggested for the free Clll, namely, the passage of a strand through the intramolecular cavity Of C111, could not 95 .H.+m.HHHo mo ecooosa cmmo one mo ‘ mnuooam on» anu coapoe HoseammOp on» non B\H co Axvca no cocoocoaoo one .nm shaman To. xi. 00.0 Omé 00.? Oman 7.2: .68. 3.6va 1.9: .8 c. n. om M 1 2. .. .106 A55 3 .1 0.0. 96 take place with the internally monoprotonated species. Thus, vicinal carbon wagging motions probably give rise to this phenomenon. Below 200 K, another process begins to slow to the millisecond time-scale as suggested by the temperature de- pendence of the resonance of the internal proton. The motion of the internal proton is very rapid at room temperature, giving rise to a single narrow line at 9.0 ppm. As the tem- perature is reduced, this line broadens greatly and splits within a very narrow temperature range, as shown in Figure 28. The two resulting peaks are located at 10.8 and 8.0 ppm with areas of roughly 1 to 3. The value of A01 at coales- cence (187 K) is 8.A i 0.3 kcal mol'l. These lineshapes were also simulated using linewidths of 30 Hz for both sites as well as 10 Hz for both sites in the absence Of exchange, the choice of which did not appre- ciably change the calculated lineshape. The chemical shift difference Of the two sites in the absence of exchange was taken as 50A Hz (2.8 ppm). The simulated lineshapes are given in Figure 29, and the plot of ln(k) versus 1/T is given in Figure 30. The Arrhenius plot (Shown in Figure 30) has two distinct slopes, a behavior similar to that of Figure 27 for the vicinal carbon wagging motion. Again, two processes seem to be involved, one which is operative at high tempera- tures, and one which is important below 190 K. Close compari- son of the two plots reveals that in the intermediate region, where the two plots overlap, the slope of the lines deviate 295°K 1 ZOO°K 97 L Aw U |85°K MN I80°K I 78°K M Figure 28. The temperature dependence of the internal proton resonance of Clll-H+,i in d6-acetone. 98 :ououa awakens“ on» no X05. v.00. cocoocogoo snapshanou on» no .ocouoomlmo CH a. m.HHHo mo mocmconOA +:o«umassfimlh093asoo .mm whswfim 0.0 O“ 0.0. 0.. . 0,0 cm 0.0. 9 . H V. CON xnm. xmmN 99 .H.+m.HHHo no cowona assumes“ on» no wpuooam can Bonn mmOOOHQ weaxooa ampoaoxm on» you B\H so Axvca no cocooconoo one .om shaman muo— x ... \ . 0.0 0.0 8.? COM QM 7.2: .68. o... om 0.0. m .IBE .00! N I DU $3.... l 0.0. 100 from that of the other temperature regions. In this inter- mediate region, both plots indicate the presence of two processes which causes line broadening or narrowing in ex- cess of that predicted by the computer simulation analysis. The high temperature process has already been assigned to the slowing of the vicinal carbon wagging motion. The simu- lation of the second process yields Ea = 16 t 2 kcal mol'l, AH+ = 16 t 2 kcal mol'l, and AS4‘ = +39 cal mol-1 deg-1. The large activation enthalpy necessitates a concerted proc- cess, if it involves only conformational changes of the ligand. The nature of the process is highly unusual, because any process with such a large value of AH+ would be expected to be slow even at room temperature. If the value of AH+ is correct, then it is the large positive AS+ which is respon- sible for the very rapid exchange rates at such low tempera- tures, but its origin remains a mystery. This exchange phenomenon might arise from one of three processes. First, a phase transition might have occurred at these temperatures. In order to rule out this trivial ex- planation, the sample was pulled from the magnet gap and in- spected. No abnormality of this sort was noted. In addition, the spectra were duplicated in methanol, and the same kind of splitting was also Observed. A second process which could give rise to the spectra would involve the exchange of the Hi between two sites inside the ligand, namely, the nitrogens (8.0 ppm) and the oxygens (10.8 ppm). The different pOpulations of these sites would 101 reflect their relative energies, the nitrogen site being more stable. One discrepancy in this mechanism involves the expected effect on the chemical shifts of the backbone pro- tons when the exchange of the H1 is slow. We have shown that the maJority of the shift of the backbone protons upon protonation is due to the location of the added proton rather than to a conformational change. Therefore, if the motion of the internal proton was slow, the CH2N proton sig- nals would be shifted about 250 Hz apart, corresponding to protonated and unprotonated sides. If the exchange of the Hi were slow enough (and simulation suggests that it indeed would be), the CH2N resonances should be vastly broadened at about 185 K and two or more lines shouId appear which are centered at 3.11 ppm, but separated by 250 Hz at 180 K and below. The spectra in Figure 25 indicate that this phenome- non is not occurring. Therefore, we must conclude that the exchange Of the H1 between the nitrogens is always rapid. The third explanation involves two conformations of the ligand, which possess different cavity sizes and donating abilities to the internal proton. The environment of this proton is determined by the topology of the ligand's cavity, and if it changes, the chemical shift of the H1 would also change. A process which involves an exchange between the two conformations of the ligand, similar to those shown in Figure 13, might well explain our observations. The cavity size would change by an Angstrom or more in going from one confor- mation to the other and the conformation with the smallest 102 cavity would be expected to be more stable. The line at 8.0 ppm would result from the proton in the small cavity, since this conformation would allow efficient electron donation, whereas that at 10.8 ppm would correspond to the conformation with the larger cavity. The populations of the two sites, reflected by the areas of the respective resonances, suggest that the former conformation is more stable. This exchange phenomenon could also reasonably explain the unusual thermodynamic parameters, because in a concerted process, the individual thermodynamic components of the + of 16 kcal mol"1 1 deg"1 is processes are additive. Therefore, the AH is unexpected, but the large 68* of 39 cal mol‘ still surprising. It may arise from several factors, such as an increase in molecular volume between the products and transition state, several pathways from transition state to products, a greater mobility Of the Hi in the transition state, etc. Proton tunneling might also effect the line- shape in a way which is not described by this classical ap- proach. Whatever its origin, the large value of A81 is the dominant factor which allows rapid exchange of this high energy process at such low temperatures. It must be stressed however, that the apparent values of the activation para- meters are fairly well determined experimentally. The large values of AHI and AS:'1 are not the result of large experimen- tal errors. The process which gives rise to these parameters must therefore be unusual. Of course, since the two-site 103 .mOmSlsfi CH Hn+$oHHHO .HO .mhpomnmw mzz mflu .HO mofimflfiwflmfl mLSQGLmQEmp 63H. .Hm mLSWHnH Edam con om.» 9...» can om.» 9...» can om.» 9...» mm. 0m. OON mON OVN OmN 10A Table 10. The temperature dependence of the NMR parameters for a 0.0A M solution of Clll°H ,i°Br' in dA' MeOH. CH 0 CH N H H Temp (°K) 6pp771 Gppg 6pém Avl/21(hz) 300 3.70 3.11 9.03 280 3.68 3.10 9.02 15 260 3.68 3.10 9.03 1A 2A0 3.67 3.09 9.0A 1A 220 3.67 3.10 9.10 20 210 3.79, 3.11 9.1A 30 3.56 205 3.82, 3.15, 9.16 30 3.50 3.0A 200 3.83, 3.18, 9.2 50 3.A8 3.01 190 3.82, 3.17, 9.1 130 - 3.A5 2.98 185 3.82, 3.16, -———— ———— 3.A5 2.96 180 3.83, 3-16: '— _— 3.A5 2.96 175 3.83, 3.16 8.0 200 3.AA 2.9A 170 3.82, 3.16, 7.9 150 3.A2 2.97 165 3°79, —_ 709 70 3.A0 3.00 105 mechanism was used in the simulation, the analysis could be in error if more than two conformations are involved. Figure 31 and Table 10 describe the exchange in dA‘ MeOH, and illustrate that the process is very similar in both solvents. The H1 splitting is also Observed in dA' MeOH, but the populations of the two sites are somewhat dif- ferent, probably as a result of different solvation of the ligand. C) Clll-2H+,i-i-2C1_ in du-MeOH The internally diprotonated cryptand, Clll-2H+,i-i, should have a very rigid structure in comparison to Clll or Clll-H+,i, because of the crowded nature of the cavity. Pro- ton-proton repulsion inside the cavity would force the ligand to expand to as great an extent as possible, and its tempera- ture characteristics should reflect this strain. Figure 32 shows that the temperature dependence of the NMR spectra of Clll-2H+,i-i in du-MeOH shows behavior similar to those of Clll and Clll-HI-Br‘. As the temperature is decreased, the CHZO lines broaden most rapidly and split into two lines, while the CH2N proton lines broaden, but neither shift nor split. This pattern is again consistent with an ABXY assign- ment, except that the breadth of the lines masks the coupling. The chemical shifts are assigned as follows (Table 11): A.15 ppm (A), 3.75 (B,X,Y). It should be noted that the CHZN proton and one of the CH20 proton resonances overlap, an assignment which is confirmed by the relative areas (1:3). 106 .moozlzo CH H1H.+mm.HHHo mo wnuooam mzz on» no cocoocoaoo ensuHLOQEOp 029 .mm oaswfim Edam 2am on? 2..» owe can 02» ON. ON ONN OnN OVN OmN 107 Table 11. The temperature dependence of thi NMR parameters for a 0.0A M solution of Clll-2H ,i-i-2Cl in dA’ MeOH. CH O CH2N H H Temp (°K) 6ppi7i ppm 6pém Avl/21(h2) 305 3.96 3.74 7.33 50 280 3.96 3.7A 7.33 50 260 7.31 37 2A5 3.96 3.76 7.30 2A0 3.96 3.7A 7.30 30 230 7.28 25 220 3.95 3.76 7.28 21 210 A.15, 3.75 7.2A 25 180 A.15, 3.75 7.21 30 170 A.15, 3.7A 7.20 30 The extreme breadth of the lines is probably caused by the rigidity of the molecule, which leads to more efficient dipole-dipole relaxation. The chemical shifts of the two Hi resonances are independent of temperature, which suggests that the topology of the cavity is also temperature indepen- dent. The linewidth of the H1 protons, however, is very temperature dependent, as mentioned in the section which considered spectra at elevated temperatures. This tempera- ture dependence results from efficient scalar relaxation via the nitrogens. This Complex provides supportive evidence that 108 the skeletal locking process described for C111 and Clll-H+,i is vicinal carbon locking rather than exo-endo isomerization, because the diprotonated species is not able to undergo nitrogen inversion. D) Clll-2H+,i-o in du-MeOH The inclusively monoprotonated ligand, Clll-H+,i has an NMR spectrum which is very pH dependent in water, as pre- viously attributed to the formation of the exo-endo dipro- tonated species, Clll-2H+,i-o. This complex is also pro- duced upon the addition of an acid to methanol and, upon the acidification of a solution of Clll in methanol and, as in water, external proton exchange is rapid on the NMR time- scale and average D symmetry is preserved. As acid is 3h added to the solution, the skeletal protons shift downfield, and broaden slightly (Table 6) in conformity with the fol- lowing rapid equilibrium: K + + + Clll-2H ,i-o —-‘___—-9—> Clll-H ,i + H (29) The temperature dependence of the process also indicates an equilibrium of this sort, with preference given to the di- protonated form as the temperature is lowered. This is shown in Table 12 and Figure 33 for a solution of 0.0A M Clll-H+,i-Br' in d -MeOH acidified with excess hydrochloric 0 acid (HCl). As the temperature is lowered, the CH2N proton resonance broadens more rapidly than that of the CHZO protons, both sets shifting downfield. The H1 resonance also shifts with temperature, but it moves upfield simultaneously with 109 : .Hom spfiz UOfiMHUHom .3002: o :H H. m.HHHo no mpuomam mzz 0:» mo mocoocoooo opSpMMOQEou 0:9 .mm mpswfim + Eammr can on...» one can 0m.» 03. on.» em...” ope iéj 110 Table 12. The temperature dependence of the NMR parameters for a 0. 0A M solution of 0111' H+ ,i in du-MeOH, acidified with HCl. CH2O CH2N H1 H1 Temp (°K) 699m 6ppm 6ppm Av1/2 (hz) 295 3.72 3.18 8.98 16 280 3.73 3.19 8.96 15 260 3.73 3 21 8.95 15 2A0 3.7A 3.25 8.93 18 230 3.75 3 28 8.91 22 220 3.76 3.29 8.89 A0 210 3.85, 3.3A 8.8A 66 200 3.89, obscured by 8.66 50 3.97 190 3.89, CH2O proton 8.6A A5 3.51 180 3.87, " 8.58 30 3.50 170 3.88, " 8.59 3.51 111 the downfield skeletal proton shifts. Below 200 K the CH2N proton resonances broaden extensively, and are ob- scured by the CH20 proton lines. The extreme width of all the lines and the shifts with temperature, which show a more pronounced effect for the CHZN protons, are consistent with external protonation at the nitrogen, which is fast at room temperature, but slow below 230 K. Comparison of these spec- tral features with those of the same sample prior to acidi- fication, shown in Figure 31 and Table 10, indicates that the two complexes are indeed different structures. Nitrogen inversion must be slow in order to facilitate this slow pro- ton transfer at low temperatures, thus the molecule is "locked" into the exo-endo isomer. A pseudo-equilibrium constant for protonation, defined by: (Clll-H+,i) K = (0111.2H+,i-O) ' (30) may be obtained from the shift of the CH N proton line and 2 the H1 line, by using equation (22) to obtain the mole ratio of the two species from the observed chemical shift. The CH2N proton chemical shift in the absence of exchange for the diprotonated (i-o) ligand was chosen as 3.7A ppm, the same as Clll-2H+,i—i in du-MeOH, and Table 10 gives a value of 3.10 ppm for the monoprotonated cryptand. The chemical shift of the internal proton of the diprotonated ligand (i-o) was taken as 8.58 ppm, that of Clll°H+,i was taken from Table 10 for each temperature because it shifted slightly with temperature. The values of ln(K) which were obtained 112 Table 13. The temperature dependence of the equilibrium constant for the conversion of Clll-H+,i to Clll-2H ,1-0 in du-MeOH: KN from the CH-ZN protons and K from the internal proton. Temp.(°K) (l/T)x10"3 KN ln(KN) K1 ln(xi) 295 3.39 7.0 1.95 8.0 2.1 280 3.57 6.1 1.81 6.3 1.8 260 3.8A A.8 1.57 A.6 1.5 2A0 A.17 3.3 1.19 3.2 1.2 230 A.35 2.6 0.96 2.1 0.72 220 u.55 2.A 0.88 1.5 0.39 210 A.76 1.7 0.53 0.87 -0.lA 200 5.00 0.15 -1.9 190 5.26 0.13 -2.0 (Table 13) from the two sets of resonances are plotted in Figure 3A versus the reciprocal of the temperature. Both sets of data fall on the same straight line, except at very low temperatures, where skeletal locking causes the H1 points to deviate. The slope of the line and the values of K give AHO = 2.0 i 0.1 kcal mol'l, and an apparent ASO = 11 i 1 cal mol-l deg_1 is obtained. Since the study was done only at one acidity, the true value of AS0 was not deter- mined. The two forms are very similar in energy, although the diprotonated complex is slightly preferred, but the monopro- tonated form has a more favorable entropy at this acidity. The increase in entrOpy might be attributed to greater 113 .moozuso cm oua.+mm.HHHo ocs .+:.HHH0 coozpon Edapnaaasco on» you B\H so AxvcH no occopsoaoo 039 .:m mpswfim n10_ X .—. \_ 0.0 0.? 0.? 0.» _ _ _ zmzo 68. _u _I Eat G Av: S 11A vibrational freedom for the endo-endo complex, both for the internal proton and the strands of the ligand. Much Of this vibrational freedom is lost in the exo-endo complex, as indi- cated by the very broad lines at low temperatures. This loss of vibrational freedom is also indicated by the H1 resonance, which does not split at low temperatures, in contrast to the monoprotonated ligand. 8 A pH titration of Clll-H+,i was performed in du-MeOH, using NMR techniques to measure the amount of diprotonated species formed. HCl was generated by preparing a solution of ~10% acetyl chloride in du-MeOH, which reacted with water to give HCl and acetic acid, or with methanol to give HCl and methyl acetate. This solution was added dropwise to a solution of Clll-H+,1'Br- (0.02M) in d -MeOH, and the acetyl u moiety served as an internal standard for acid determination (This titration was also attempted with TFAA, but it was not a strong enough acid to externally protonate the internally monoprotonated ligand.) Figure 35 depicts the behavior Of the H1 resonance at 200 K as acid is added. It goes from a single line in the absence of acid to two lines, when less than one equivalent of acid is added. These lines correspond to Clll-H+,i (9.2 ppm) and Clll-2H+,i-o (8.6 ppm). Figure 36 shows the temperature dependence of the system when the mole ratio of acid to ligand is about 0.5. It shows the same skeletal locking process which causes splitting of the H1 resonance in Clll-H+,i, while the H1 resonance of C111'H+,1-0 remains unchanged. These data confirm the conclusion reached 115 . +1CI mole r0110 0|er ()3) réflv-dfluflvflrJ/j/\\M/ <14£3 A JV I. 2 WA J W 10.0 9.0 so 80pm Figure 35. The variation in the NMR spectrum of the internal proton of 0111-H+,i in du-MeOH with the addition of HCl at 200 K. 116 128k l80 I l I 1 9.5 9.0 8.5 Sppm Figure 36. The temperature dependence of thg NMR spectrum of the internal proton of Clll-H ,i in du-MeOH with about 50 mole 2 H01 added. 117 previously, that the exchange between the mono and dipro- tonated ligand is fast at elevated temperatures, but slows greatly as the temperature is lowered and shifts toward the diprotonated form. The relative areas of the H1 resonances at 200 K provide an estimate of the equilibrium constant for the process: + Kg , + + Clll-2H ,i-o ‘____. Clll-H ,i + H (31) yielding xa - 7.0 s 1.0 x 10‘3 M. Using AH° = 2.0, the equilibrium constant may be extrapolated to 298 K, giving a value of about 0.06 M. The true AS° may also be calculated since the acid concentration is known, thus AS° = l cal mol"1 deg-1. The monoprotonated ligand, Clll-H+,i, may be recovered from Clll-2H+,i-o by evaporating the solvent and evacuating the ligand in a vacuum desiccator with the aid of a roughing pump, provided the acid is volatile and the concentration process does not cause cleavage of the ligand (as with HNO3). E) Clll-H+,O in d6-Acetone The temperature dependence of Clll-H+,o protonated with TFAA in d6-acetone gives rise to very complex patterns, due to the presence of two or more species in solution. As the temperature is lowered, the equilibrium tends to favor the mono and di externally protonated Clll, as shown in Figure 37, for a mole ratio of acid to Clll equal to 1.7. With increasing acid concentration, the equilibrium is also 118 0 M __J «1' b b 27 27.0"K ,J ' 240°K c 260°K 2 50°K 2|O°K d 4.0 3.5 3.0 2.5 Figure 37. The tempegature dependence of the NMR spectra of Clll°H ,o in d6-acetone, acidified with TFAA. 119 forced toward the diprotonated form in accordance with the following scheme: K + K C111 + 2H+--l-* Clll-H ,o + H+'—* 2 ? _— Clll°2H+,O-O (32) The first external protonation is very efficient and K1 is large, whereas K2 is small and the rate of deprotonation of C111'2H+,O-O is very rapid. Therefore, as the temperature is decreased, the line corresponding to the diprotonated species at 3.8-A.0 ppm grows in intensity. The other lines of Figure 37 correspond to the externally monoprotonated species. A rough calculation from the relative areas of the resonances provides the temperature dependence of the pseudo-equilibrium constant, K2([H+]), which is shown in Table 1A for mole ratios of 1.A and 1.7 acid/Clll. A plot of 1n K2 with l/T is presented in Figure 38, which gives a 1 1 value of AH° . -5 s l kcal mol‘ , AS° = -16 cal mol‘ 1 deg" . AG°298 = -0.2 kcal mol'l, K s 1.0 at 298 K and :20 at 220 K. 2 The same process may be followed at a single tempera- ture by addition of acid. For example, at 220 K the line at 3.8-A.0 ppm shifts from about 3.82 to 3.91 ppm upon addition of acid (Table 15). From this shift, the ratio of Clll-2H+,o-o to Clll-H+,o may be obtained, assuming that their chemical shifts in the absence of exchange are 3.99 and 3.82 ppm respectively. The plot of this ratio versus acid strength (Figure 39) gives a straight line of Slope = _l__ , where C K2Ct t is 0.005 M, the total concentration of Clll. 120 ueo ea ouo.+mm.HHHo mo sofiosssom one pom exa some A .<¢m9 mo COHWHUUm on» com: ocoumom mVCH mo cofipmfiaw> 0:9 .mm ogsmfim MIC—XF\_ nae nxum . o.. 121 Table 1A. The temperature dependence of the equilibrium constant for the formation of Clllo2H+,o-o in d6-acetone with the addition of TFAA. 3 Mole ratio %§%% Temp (°K) l/Txlo' K2[H+] K2 1n K2 1.A 220 A.55 0.16 23 3.1 1.A 230 u.35 0.10 1A 2.7 1.A 290 u.17 0.06 9 2.1 1.A 250 A.00 0.0u 6 1.7 1.7 210 A.76 0.30 35 3.6 1.7 2A0 A.17 0.08 9 2.2 1.7 250 . A.00 0.06 7 2.0 This yields a value of K2 8 5.0 at 220 K compared to 20 by the first method. The discrepency between the values Of K2 obtained by the two methods is probably due to large system- atic errors in both methods, especially the first, since the lines of the monoprotonated ligand are too broad to measure accurately. Therefore, the agreement is satisfactory and supportive of the previous assignments. F) Summary 1) C111 exists in the endo-endo configuration which is preferred by a factor of 50 over the exo-endo form in water at 298 K. 2) C111 protonates externally (exo-endo form) with a pKa Protonation occurs at the nitro- at 298 K in water of 7.1. gen. 122 :.s HH.H Hm.m om m.= H>.o mm.m mm o.» mm.o pm.m s.m m.m Hm.o mw.m p.m m.m see.e ms.m m.s H.m smo.o mm.m s.H m+muo.+m.aaao u ms o.+m.HHHo u mxm+xu . seam flwwwmm oases omoz +mm.HHHo ouo.+m~.HHHo +mm <coo on» now ucmumcoo Esfinnaafisvo onu no mocwocoaoo one .mH magma 123 .COfipmpusoocoo <coo on» now ucmumcoo Esfinnaaasoo MS» MO cocoocomoo 029 .mm madman 9.1.5 ..x o.o..&~.___o o... 010 o \Tle 0 711.711 mw . o . r0. 0 1 3 a .8 W. 0 111.111 Invnnwm V II Toe m. \\HWIQTI1 fifivm 12A 3) C111 will protonate a second time externally at the unprotonated nitrogen (exo-exo) with a pKa (in water) of between 0 and 1.5. A) Internal protons bind most strongly to the nitrogens and the proton exchanges rapidly between them. 5) The barrier (AHI) to vicinal carbon wagging motions 1 is about 9 i 0.5 kcal mol' with AS+ = -6 cal mol-l deg-1 for the internally monoprotonated species. 6) Another motion freezes out at temperatures below those which freeze out the vicinal carbon wagging. This process 1 has 60+ = 8.A kcal mol' at coalescence and apparent 1 values of AH+ of 16 i 2 kcal mol' and A81 of -39 cal -1 -1 mol deg . It probably originates from a concerted torsional motion of the skeleton of the cryptand. 7) The structure, Clll-H+,i, protonates externally at the nitrogen with a pKa value at 298 K in water of roughly 0.A. This low pKa value indicates that the endo-endo form of Clll-H+,i is favored by a factor of about 108 over the exo-endo form. 8) Clll-2H+,i-i exists in the endo-endo form with very slow or no exchange Of the H protons. i 9) For the external protonation of Clll-H+,i in du-MeOH, AH° = 2 i 0.1 kcal mol"1 and AS° = l. cal mol-1 deg-1. 10) For the second external protonation of Clll-H+,o in d -acetone, AH° = 5 i l kcal mol-1 and AS° = —16 cal mol"1 6 deg-1. CHAPTER A THE KINETICS OF INTERNAL PROTONATION OF Clll I) INTRODUCTION Cryptand 111 has the remarkable ability to irrevers- ibly bind protons. It does so by encapsulating them inside its cavity so as to exclude solvent interactions and to provide a very thermodynamically favored environment for the proton. The internally monoprotonated complex, for example, is not significantly deprotonated even upon heat- ing for days at 60°C in 5 M KOH! Neither do sodium in liquid ammonia or ion exchange methods accomplish deprotonation, except when the ligand is destroyed. The second internally protonated complex is also very stable to base at room tem- perature. At elevated temperatures, however, one of the two protons may be removed, but only at a slow pace. In light of the apparent stability of the internally protonated complexes, it is unusual that their rates of formation are slow. It was shown earlier that proton trans- fers between donors of quite different pKa values proceed very rapidly and favor the one of largest pKa. Yet the rate of formation of the internally monoprotonated complex pro- ceeds with a half-life of several hours at room temperature at a pH of 7 indicating that the mechanism is more complex than for normal proton transfers. This is confirmed by the disclosure that internal protonation occurs with an enormous 1 activation energy of 25 to 27 kcal mol' This large 125 126 activation energy causes the rates of proton transfer to be slow. Therefore, the long life-time of the complex is not completely thermodynamic in origin, but is also strong- ly affected by this kinetic phenomenon. The rates of internal protonation provide estimates of the thermodynamic stability of the internally protonated complexes. Their pH dependence also sheds light on the mechanism of proton transfer. These and other topics will be discussed in the ensuing sections upon presentation of the kinetic data. II) THE FIRST INTERNAL PROTONATION OF Clll The rate of the first internal protonation of Clll was studied as a function of pH and temperature in D O, 2 using buffers to obtain the desired pH. The pD value of the buffers was measured using a calibrated radiometer with a type GK2321C electrode, the values being shifted upward by 0.A pH units to give a correct pH reading<67). About 1 mg. of Clll was utilized for each sample, this amount being 5% or less of the buffer content and insignificant to the final pH. Table 16 presents the buffers utilized to achieve various pH values and their ionic strengths. The rates of internal protonation, which were measured at constant temperature (checked periodically) and pH pro- ceeded with first order dependences on the ligand in all cases and also depended heavily on both pH and temperature as shown in Table 17. A typical rate analysis is shown in 127 Table 16. Buffer compositions and their ionic strengths in D20. pH Composition Ionic Strength 8.65 KH2POu-NaOH 0.29 7.A7 KHZPOu-NaOH 0.37 6.A6 KHZPOh-NaOH 0.33 A.93 KHP-NaOH 0.3A A.0A KHTartrate 0.22 A.00 KHP-HCl 0.29 3.19 KHP-HCI 0.29 2.60 KCl-HCl 0.30 1.78 KCl-HCl 0.36 1.AA KCl-HCl 0.Al 1.25 KCl-HCl 0.38 1.11 KCl-HCl 0.55 Figure A0 in which the natural log of the reactant concen- tration is plotted versus time. It should be emphasized that these rates were extremely slow, each reaction requir- ing several hours to go to completion. The temperature dependence of the rate Of internal protonation has been studied at pH values Of 7.5 and A.9, both of which provide similar Arrhenius activation plots which are depicted in Figure Al. The temperature dependence of the rates at the pH values provide straight line plots of nearly equal slope, giving values for the activation parameters at 299 K of Ba = 26 i 2 kcal mol-l, 128 Table 17. The dependence of the rate of internal monopro- tonation of Clll upon pH and temperature. Temp. + pH (buffer)* (°K) k(sec"1) 0 1n k loglo k unbuffered (=9) 299 A.27x10’6 2.7::10"7 -12.36 -5.37 8.7 299 1.57x10“5 3.0::10"7 -ll.06 -A.8l 7.5 299 8.70x10'5 3.03:10'6 -9.35 -A.06 7.5 310 3.65x10'” 8.3x10‘6 -7.92 -3.AA 7.5 320 1.25x10‘3 5.0x10“5 -6.69 -2.90 6.5 299 1.1ux10’“ 8.0x10'6 -9.08 -3.95 A.9 299 1.70x10'” 1.2x10'5 -8.68 -3.77 A.9 320 2.08x10'3 2.0x10’“ l-6.l7 -2.68 A.0 (tartrate) 299 2.02x10'u 1.9x10'5 48.59 -3.73. u.o (HCl-KCl) 299 1.95x10'“ 1.2x10'5 -8.A8 -3.68i 3.2 299 2.25x10’“ l.0x10"5 -8.33 -3.62 2.6 299 A.22x10‘“ 7.3::10'5 -7.A7 -3.2u 1.8 299 6.A0x10'” 1.7x10"5 -7.35 -3.19 1.5 299 8.53x10‘" 1.3x10"5 -6.67 -2.90 1.3 299 9.u2x10‘“ 1.8x10'5 -6.65 -2.89 1.1 299 1.03x10"3 1.1x10‘5 -6.A9 -2.82 * pH values measured by calibrated radiometer and corrected up 0.A pH units for deuterium isotope effect. I Some values corrected to 299 K using Ea = 26.A kcal/mole. Linear estimate of the standard deviation on k. 129 .uoumz CH dado mo coaumcouOHQ Hmcnouca pom mamzamcm cusp Hecaozu < .o: oasmfim 65F; mit. 00 OM o N o. o . . AWN /O/ O O r O.” O ////// Amwbrh O /O/ 0 .0d. ///0 /O ///, 130 .m.= n ma ”moaonfio cum m.» n ma mmmnwsom umcsHm> ma pCOMOMMHU o3» um 900w: Ca HHHO mo COHuHCOpopQ assumpca on» pom B\H so Axvca no cocoocmaoo one MLU3§F\_ cum mm «Mn 3 AHN 2?... AH? \\ .H: mpswfim 131 A61 = 23 kcal mol"1 and AH+ s 25 kcal mol'1 and AS1’ - +7 cal deg"l mole-1(73). The large free energy of activation originates mainly from the enthalpy term. This is pre- sumably the energy required to distort the face of the ligand, which allows the proton access to the cavity. This topic will be discussed further when the activation param- eters of the other internal protonation processes have been documented. It is clear, however, that the slow rates of internal protonation are dominated by a large enthalpy bar- rier which results from ligand deformation. The large activation barrier is somewhat unexpected, since a proton possesses negligible size and should easily fit into the holes of the ligand's faces. However, in solution (as H3O+) the proton possesses electron density obtained from its hydrogen bonded counterpart and has a finite size. As the proton interacts with the nitrogen and the oxygen atoms of Clll, it accumulates considerable elec- tron density from this source which also imparts significant size to the proton. Therefore, the same interaction which induces proton binding also restricts the proton from enter- ing into the cavity of the ligand. Once inside, it accumu- lates even more electron density as it interacts with the nitrogen and oxygen lone pairs and its effective size be- comes larger. This phenomenon probably contributes most to the resistance of the complex to deprotonation. The conformation of C111 is very much influenced by the pH of the medium, since it may bind either one or two 132 protons externally. Such changes in conformation also seem to correlate with the pH dependence of the rate of internal protonation. Table 6 (Chapter 3) lists the chem- ical shifts of Clll as a function of pH in water, the CH N 2 proton shifts being plotted in Figure 22 (also in Chapter 3). A break is observed in this curve at a pH value of about 6, above which C111 and Clll-H+,o are the predominant forms of the ligand in solution. Below a pH value of 6, the ligand is fully monoprotonated externally and begins to add a second proton (also externally). However, even at a pH value of 1, only about half of the ligand is dipro- tonated externally, if it is assumed that the chemical shifts of these species in D20 are similar to those in d6- acetone (as they appear to be). The reversal of the CH2N proton shifts after complete formation of the externally monoprotonated form (at a pH value between 5 and 6) is also observed in d6-acetone upon the formation of the externally diprotonated complex. This reversal is probably a result of less effecient proton binding as the ligand must accomo- date a second positive ion. This NMR titration agrees qualitatively with a pH titration done in a conventional way with a calibrated radiometer, as shown in Figure A2. This titration gives an endpoint at a pH value of about 5 and a value of pKa = 7.1 i 0.3 at 299 K for the first external protonation. A second break has not been observed, even at pH values as low as 1.2, but the second pKa may be estimated at :_l.5. 133 901K 00 - '\ . ..\. 7o -1 \l O 0 pH \9, EK)‘ ‘\ . \fEND POINT 5,04 . 1 \o 4.0 \ C. . .\ .\.\. 310 . \.\o 2.0 . . , C O O I 0 2 0 3 0.4 0 5 O 6 07 Ml HCI Figure A2. The pH titration of Clll with HCl in water (only external protonation occurs). 13A (The NMR data indicate a pKa value of between 0 and 1.5.) The relative concentrations of the various forms of the ligand in solution may be calculated as a function of pH by using pKa values of 7.1 and l. for the two external protona- tions (Figure A3). The externally monoprotonated complex is the predominant structure over a very wide pH range, 2 to 7, above and below which the other forms are important. The rate of internal protonation correlates well with the form of the ligand in solution. Figure AA depicts the dependence of this rate upon pH. It shows a nearly linear curve between pH values of 3 and 7 where the monoprotonated species is predominant and in fairly constant concentration. Below a pH value of 3, the rate increases and approaches a first-order dependence on the hydrogen ion concentration simultaneous with the build-up of the externally diproton- ated complex. Above a pH value of 7, the rates are very slow, decreasing markedly as the mole ratio of Clll°H+,O decreases. The correlation Of the rates of internal protonation with the form of the ligand suggests that it is easiest to internally protonate the externally diprotonated ligand. The externally monoprotonated form, C111°H+,0 will inter— nally protonate, but with difficulty and the free ligand is even more difficult, if not impossible to internally pro- tonate. This observation is represented schematically as follows: 135 .ma nod: ponds CH Haao mo wagon nsoapm> on» no mcoaumhpccocoo o>HuwHop on» no coaumaam> one .m: shaman NOIiDVHd 310W 136 m .m.> n oxa ”mafia oaflom com a.» n ma «mafia cmnmmo "mg no Haao no coapocouopqocos Hmcp0p2H mo camp on» no cocoocoaoo one .3: mpswfim Ia cm 0.0 05 0.0 0.0 Q? on ON 0.. Dad ..on .0... . m. .0 no for. -00 137 0111 k + _1 1 Ii 1L Kal k Clll-H+,o 2 s: Clll-H+,i + -1 H H, K82 + k3 Clll-2H ,o-o (33) in which kl<fluom one .om manna 1U9 2.3xlO'u. k1 is the overall rate constant for the process given by: k1[Clll] = k2K [0111] (A2) [OH at a pH value of 9, where only Clll and the externally mono- protonated complex are present: a (2.3x10'“)(1o'7’1) (10‘5) = 1.8x10"6 (#3) k1 This value of k1 is very similar to that calculated by the other method. k_1 is too slow to be observed, but upper limits can be estimated from: + - k-2 + Clll-2H ,i-i + OH-—————a Clll-H ,i + H20 k_2 equals A.3x10'5 at 372 K for l M KOH, which may be extrapolated to 299 K using Ea = 24.8 kcal mol"1 to give k_2 = 1.Ax10'8. The formation of the free ligand must be slower than 5% of this rate or else it would be detected, so k_ may be estimated at less than 5% (k_2), giving 1 k_l:7.0xlO-lo. Therefore: Kb = -—— i _ z_6.lx10‘3 (8“) pr £_-3.8 and pKa 1 17.8. For reaction (43) as written above, AG° 5 -5. kcal mol"1 which indicates that because of encapsulation, the first internal protonation is 150 thermodynamically stable as opposed to possessing only kinetic stability; that is, C111 is a very good base thermodynamically but a poor one kinetically. For the second internal protonation, k Clll°H+,i +11 0 :32 Clll-2H+,i-i + on‘ (115) k-2 k2 . 3.1x10'7 at 300 K and [H+] = 1 M. Assuming a first order dependence on [H+], the rate may be extrapolated to pH . 7, giving k2 - 3.1x10‘1”. k - 1.!4x10‘8 at 300 K, -2 and [OH'] = l M as before. Therefore: -14 k 3.1xlO - sz : .3. a = 2.2x10 6 8 (46) k_2 1.14X10- pr2 = 5.7, pKa2 = 8.3, and AG° z +8.0 kcal mol-l. The stability of the second internal protonation is similar to that of Dabco, thus not as thermodynamically favored as the first. Its apparent stability, as indicated by the slow rates of proton transfer from the cavity, does not originate from a thermodynamic source. Rather, the slow proton transfer processes are due solely to the large acti- vation barrier which must be surmounted in order for the proton to escape from the cavity. A fairly complete potential energy diagram may be constructed from the kinetic and thermodynamic parameters which have been accumulated. This diagram, presented in Figure 49, utilizes free energies for consistency, since the AH values of some processes are not known. The 151 .mmmmoooao :ofipmcopona Hangoucfi IIO+_I_.+IN.=_U qu++r+ 2.2.6 \ on» non Ewhwwfio mmpocm o~:+_pz.=_o \l/ e x 22010.. 51...}, Hmfiucmuoq was .m: mpsmfim ll. oml l ON... o~z+=.o lo_+ ION+ .I On... 20:503. 0Q 152 direction of the plot, indicated by a solid line, has been determined experimentally, whereas the dashed line merely Joins the transition state with the products, and has not been measured experimentally. The diagram may be somewhat misleading because the pathway between reactants and transi- tion state is represented by a smooth curve, and is probably in error. The route more likely goes via a series of inter- mediates, but the one(s) of highest energy determines the rate of the reaction. The shape of the dashed line profile is also not known. It has been assumed that the reverse reaction follows the same profile as that of forward reac- tion in order to Join the reactants and products. The value of AGI has no physical interpretation since it is the difference of AH+ and TASI which do possess a physical interpretation. Rather, the value of AGI is only a measure of the reaction rate at a given temperature. Therefore, the difference in the values of AG+ for the first internal protonation of Clll by acid and by water merely reflects the pH dependence of the rates and does not sug- gest a different reaction profile for the two processes. The mechanism for this protonation has not been established but a definate pH dependence for the rate of internal pro- tonation has been observed. VII) THE REACTION OF 0111-H*,i WITH Na+C222 Na‘ The reaction between the internally monoprotonated 0111 (with a hydroxide counter-ion) and Na+C222-Na- was 153 studied by mixing approximately stoichiometric amounts (2.x10'5 moles) of Na+c222-Ma‘ and Clll-H+,i in methylamine. The sample was prepared by using the technique described in the Experimental section in a vessel similar to that shown in Figure 8. The vessel had two side-arms, one for the Na metal and one for the cryptand. The solution of Na+C222oNa- was prepared first and then was poured over the Clll-H+,i. The temperature was held at about 0°C. As the Clll-H+,i dissolved, the solution went from deep blue to clear, an evidence of decomposition. An NMR spectrum of the products of the reaction revealed complete decomposi- tion, except for some starting material which did not dis- solve. No free Clll was obtained. VIII) THE REACTION BETWEEN C111°2H+,1-1 AND ES The reaction between Clll-2H+,i-i and ES was studied by the addition of Na metal in liquid NH3 to a solution of =3.x10-5 moles of Clll-2H+,i-i in liquid NH The sample 30 was prepared in a vessel equipped with an EPR tube. The solution of Na+ and ES in liquid NH was prepared first (an 3 excess of about 3 to 1 metal was present) and the entire vessel cooled in a dry ice, isopropanol bath. The dark blue solution was then poured over the ligand upon which considerable effervescence was observed. The dark blue sam- ple was quickly transferred into a 3mm EPR cell, frozen in liquid nitrogen and its EPR spectrum taken. Only the signal of precipitated sodium metal with a linewidth of about 15A 20 gauss was observed. The sample was transferred back and forth over the undissolved ligand (probably decomposed), frozen and analysed several more times in like manner. Again, only the precipitated sodium metal line was observed. The NMR spectrum of the chloroform soluble portion of the resulting sample showed mostly decomposition, but some free 0111 (210%) was present. The reaction between 58 and the internally diprotonated ligand must have occurred very rapidly and the effervescence observed upon mixing was probably hydrogen gas being evolved. A rapid EPR mixing cell would be valuable for this experiment since electron transfer into the cavity of the ligand appears to be facile and very rapid. The speed of these deprotonation reactions indicates that a very different mechanism is operative than in the case of proton abstraction by base. The reaction products substantiate this belief since the free amine cannot be prepared from either of the internally protonated complexes by base in water. Therefore, the deprotonation probably proceeds via the production of a hydrogen atom inside the cavity of the ligand, which in the case of the diprotonated complex, may unite with the other proton to form H2+. The transfer of a second electron produces molecular hydrogen and the free ligand is generated. In the case of Clll-H+,i, the H' has nothing to react with except the ligand, which is therefore completely destroyed. In either case, H' is a very reactive species and causes considerable decomposition. 155 These experiments verify the findings of Lehn and Cheney (27) who conducted very similar reactions CHAPTER 5 DISCUSSION 156 Cryptand 111 has been shown to be unique in its ability to bind protons. It irreversibly binds protons by encapsulating them and isolating them from solution. The pKa for the first internal protonation is greater than 18, which illustrates the tremendous thermodynamic stability for a proton inside the cavity of Clll. The complex derives most of its stability, however, from a kinetic source, be- cause the activation barrier to internal protonation- deprotonation processes is at least 25 to 28 kcal mol. The activation barrier results from the difficulty in opening the face of the ligand so as to expose its cavity to the proton-donor. Thus, internal protonation proceeds slowly, with a half-life at 26°C of roughly 800 sec at a pH of 1. However, the half-life for deprotonation in 5 M KOH at 26°C is greater than six years. The internally monoprotonated complex, as mentioned, gains the bulk of its stability from two sources. First, the interaction of the donor atoms of the cavity of Clll with the H1 provide stability of a thermodynamic nature. Second, as the proton approaches the cryptand from a proton- donor in solution, it interacts with the nitrogen and oxy— gen atoms, thus accumulating electron density. Therefore its size increases as the interaction becomes stronger and the proton becomes sterically restricted from entering the cavity of the cryptand. Once inside, the proton attracts almost an entire electron from the donor atoms, and its size increases to about an Angstrom. This large increase in size 157 makes proton abstraction nearly impossible, because the ligand is physically unable to distort itself enough to allow the proton to exit. Proton acceptors are also ster- ically restricted from approaching the encapsulated proton. Therefore, it is likely that physical encapsulation provides the majority of the kinetic stability for internal complex- ation. Quantum calculations have been utilized to gain in- sight into these characteristics of proton binding by C111. 0111 is an excellent candidate for a theoretical analysis because it possesses properties which are similar to the larger cryptands, thus making the study of ion binding by this class of ligands possible. It is also a relatively small molecule whose topology and conformations are well characterized. C111 is a huge molecule for INDO and espe- cially ab initio calculations, but it fortunately possesses only atoms whose atomic wave functions are relatively well known. The calculations are based on the linear combination of atomic orbitals (LCAO) technique to construct the basis set. The energies of the ensuing molecular orbitals are determined by the self-consistent molecular orbital theory using some semiempiracal approximations which decrease the number of matrix elements to be evaluated. The INDO<8u) (intermediate neglect of differential overlap) method has been utilized for these calculations. This technique ignores interactions except between neighboring atoms and those two 158 bonds away. Clll possesses 93 atomic orbitals which are important to the calculation (only valence electrons are considered). The large number of matrix elements to be evaluated would render the calculation nearly impossible if these approximations were not made but in doing so, the accuracy of the energy values of the molecular orbitals is sacrificed. The INDO calculations are most valuable as models to guide our intuition and are only to be interpreted in light of experimental results. The calculations reveal that the inclusive proton's lowest energy position along the C3 axis is about 1.12 from the nitrogen (Figure 50). The charge on the proton is given beside each point. (Typical N-H bond distances are about 13.) The barrier to proton Jump from one nitrogen to the other, along the C 3 mol-l, but a known weakness of the INDO calculations is axis is indicated to be 31 kcal that it gives energy values which are much too large. The barrier may also be much lower if the proton leaves the C3 axis, because the oxygens may also participate. The sites of lowest energy seem to be nearest to the nitrogens and the INDO calculation shows that the proton in this position has accumulated 0.7 electrons. Therefore, its size is between 0.6 and 13, which is very large for the cavity. Its large size not only makes it impossible to re- move the proton from Clll, but it may also restrict partic- ipation of the H1 with the oxygens. The topology of the inside of the cryptand is such that there are two large 159 .znpoEEmm nmn :H fi.+m.HHHo Lou maxm mo on» mcoam sonopn m an» no coaumooa on» no meQEoo can ho mwhmcm Hmuou on» no mocoocmqmo one .om mpswfim N Illa? thO o\/ \/0 Ia 0_oc..\_Oux _m Enos/f. \ I I f- novd fl 2295.. 6525 cc 865 228d V4.89. \ 4 _ 88 catacoaom 2...... n “O... o: o cozoeooom z: Z. dean o L 0 +1 ——-°'——' 160 chambers between the nitrogens and oxygens, which are large enough to contain this ion. The cavity is restricted at .its center by the oxygens, and in order for the proton to exchange between the nitrogens, it must squeeze through this orifice at the center. The resulting steric barrier may explain why the Hi does not interact appreciably with the ether oxygens inside the cavity. If the energy barrier to proton exchange is substan- tial due to steric restriction by the ether oxygens, the exchange would be expected to be easily slowed to the NMR time—scale at low temperatures. Since this does not occur, a tunnelling mechanism is suggested for proton exchange. The expected tunnelling frequency may be calculated using the mass of the proton, a barrier height of about 30 kcal mol"1 and a proton Jump distance of 1.583 (3.0au). This frequency, which is typically 10’8 seconds or less for sym- metric potentials, is much faster than the NMR time-scale and is theoretically independent of temperature<86). If the cavity were smaller due to the concerted twistabout the nitrogens, the tunnelling rate would be expected to be even faster. The fact that C111 is small and rigid, and has a well defined cavity topology makes it a favorable model for test- ing several NMR theories. Its ability to shield solvent from its internally complexed protons magnifies this utili- (30) ty. iusher's theory , for example, was shown to predict shifts which were in the right direction and relative 161 magnitude although they were too large. However, if the correct charge density on the nitrogen is utilized for the calculation, 0.? (as obtained in the INDO calculations), rather than 1.0, the values are much closer to the observed shifts, -2.5 ppm for the CHZN protons, -0.9 ppm for the CHZO protons, and +1.6 ppm for the H The observed shifts 1. are -l.u, -0.6, and +1.A ppm respectively. The two former moieties are greatly influenced by solvation, whereas the H1 is not. Therefore it is interesting that the H1 shift is much closer to the predicted value. The theory of nuclear coupling via the Fermi contact interaction is supported by the behavior of the C111 com- plexes as is that of scalar relaxation between the H1 and the nitrogens. Both the increase in magnitude of the scalar coupling between the complexes, Clll-H+,i and Clll-2H+,i-i, and its temperature dependence are in close agreement with theory. Clll answers several questions concerning metal ion binding by the larger cryptands. The conformational analy- sis of Clll indicates that the free diamine exists substan- .tially in the endo-endo configuration, but has the flexi- bility of nitrogen inversion. Even when binding a proton, it may still invert a nitrogen lone pair to the exo config- uration, although this process occurs quite infrequently. The larger cryptands are much more flexible than C111 and would therefore be expected to more easily undergo exo- endo nitrogen inversion, even when complexed to a metal 162 cation. The acid catalyzed removal of metal cations com- plexes by the cryptands, observed by Cox and Schneider<25), probably does proceed by the mechanism which they suggest. Inference from the C111 binding conformations supports their belief that the exo conformation of the nitrogen should be fairly substantial and available for proton bind- ing. The binding characteristics of Clll also support the findings of Pizer, et. a1.(“7-“8), who attributed the slow rates of complex formation and dissociation of the cryptands to a steric phenomenon arising from the rigidity of the ligand. However, they also attributed part of the slowness to a very low percentage of the "reactive" endo-endo con- formation. The preference for this conformation with 0111 and the ease of nitrogen inversion (when unprotonated), suggest that this latter argument is not valid. Future Projects The most significant experiment left unfinished is the preparation and identification of H + inside the cavity 2 . of 0111 by adding es to Clll-2H+,i-i in liquid NH This 3. reaction was attempted once and although H + may have been 2 formed, the solution decomposed too rapidly to obtain its EPR spectrum. A different experimental design is necessary to accomplish this goal. The quantum calculations dealing with proton binding by C111 will also be very useful in identifying the most favorable conformation of the ligand, both when bound to a 163 proton and when free. This theoretical analysis may also help to determine the energy barrier to proton exchange inside the cavity of the ligand, the most favorable en- trance for a proton into the cavity from solution and many other important questions. The binding characteristics of C111 and Clll-H+,i with other ions would also be interesting. For example, +2 2 ions able to coexist in- are both protons and Be or Mg+ side the cavity of C111. If not, do any metal cations bind either internally or externally to the internally monopro- tonated complex? The kinetics of decomplexation of L1... or Mg+2 ions from C111 by protons may also prove interesting, since the kinetics of external protonation of Clll are now known. This would help to answer the question of whether acid catalyzed decomplexation of metal ions by other cryp- tands occurs by external or internal protonation at the nitrogen. Finally, the crystal structures of the protonated complexes of Clll would be of interest, especially C111°H+,i, Clll-H+,o and C111°2H+,i-o. PART II AN NMR STUDY OF ALKALI METAL ANIONS IN SOLUTION I) INTRODUCTION The properties of solutions of alkali metals in liq- uid ammonia (NH3) have intrigued investigators for well over one hundred years since the discovery of Weyl in 1863(7u) that sodium and potassium metals could be dissolved in liquid ammonia to give deep blue colored solutions. The solubility of alkali metals was later shown to be as high as 15 to 20 mole per cent in some cases, the resulting blue to bronze solutions being stable for days at room tempera- ture. Electronic and EPR spectra and electrical conduc- (75’78) indicate that the solutions contain a tance studies variety of chemical species, those of maJor importance being the metal cation (M+), the solvated electron (ES) and the neutral ion pair, M+-§S or "monomer". Other more exotic species such as e2 -, alkali metal anions (M') and higher order ionic aggregates have also been postulated to be pres- .ent(78). The study of alkali metal solutions has been limited almost entirely to liquid ammonia, since the alkali metals are quite insoluble and highly reactive toward most sol- vents. Exceptions to this rule include hexamethylphosphor- ic triamide (HMPA), ethylenediamine (EDA), and methylamine (MA), which dissolve at least some of the alkali metals to concentrations higher than 0.01 M. In most solvents, 16A 165 however, the alkali metals are quite insoluble or only slightly soluble without decomposition. The chemistry of solutions of the alkali metals in amines and ethers is superficially similar to that in ammonia, since deep blue, highly reducing solutions result<75’78). The maJor dif- ferences between metal ammonia and metal amine solutions are solubility and stability toward decomposition, both of which decrease in the amines, and the appearance of metal dependent bands in the electronic spectra of the metal amine solutions. Metal ammonia solutions show only one broad electronic band, regardless of the alkali metal, assigned to ES and aggregates containing this species. The elec- tronic spectra of metal amine and ether solutions are char- acterized by a band whose frequency maximum is greatly de- pendent on the metal as shown in Figure 51. These bands have been assigned to the alkali metal anions. |.O "’ 1 1 4 1 1 L L i 1 4 6 a lo I2 I4 I6 la 20 waveuwaen (car'aIO’3) Figure 51. The electronic spectra of the alkali metals in tetrahydrofuran (THF). 166 Electron paramagnetic resonance spectra of metal ammonia solutions in fairly dilute solutions (<10.1 M) show only the 8 line with a g-value near that of the free S electron. Metal amine solutions often show this line and an overlapping multiplet attributed to the solvated mono- mer. Thus, the solvent plays a maJor role in the character of species produced in solutions containing the alkali metals. The solubility of a particular metal is also greatly influenced by the prOperties of the solvent. Solubility seems to correlate well with solvent dielectric constant, which decreases in the progression NH3, HMPA, MA, EDA, ethylamine (EA), isopropylamine (IPA), tetrahydrofuran (THF), and dimethylether (DME). For example, the solubility of sodium in NH3 is about 15 mole per cent (=5 M), but in THF 6 M. it is less than 10- Solvent polarity may also be related to the following equilibrium: M ———-—-s-M+ + 258 (A7) The more polar solvents favor the formation of E , whereas, S those which are less polar tend to shift the equilibrium to M-. The character of the metal itself also influences the ratio of metal anion to 5 in solution. For example, sodium S has a strong preference to form the anion even in MA, but potassium seems to favor the formation of the metal cation and ES. kl is so large for lithium, that Li‘ has not yet 167 been observed in solution. In order of increasing k the 1, progression seems to follow Na, Cs, Rb, K, L1. The advent of crown ethers and cryptands has greatly extended the number of solvents in which alkali metal solu- tions may be studied<76). They provide a great increase in the solubility of the alkali metals for the previously men- tioned solvents, as well as for many other nonpolar sol- vents such as benzene and toluene<77). Cryptands, in par- ticular, fulfill another role in alkali metal solutions. They tend to isolate the cation from the anion, and thereby increase the stability of the anionic species by preventing recombination to form metal. The exchange rates from these ligands are slow, on the order of milliseconds and greater. The cations are therefore encapsulated and insulated from the alkali metal anion by the backbone of the ligand. This can lead to either thermodynamic or kinetic stabilization of M” (or both). The cavity dimensions of the ligand may also be impor- tant in determining the products of metal amine reactions and their stability. Since cryptands discriminate between ions by orders of magnitude in their binding constants, the products of mixed metal reactions may be perturbed by the proper choice of cryptand. For example, the reaction with excess metal produces one product via the following scheme: EA + _ 2Ha(s) + C222 —->Ha C222°Na (‘48) whereas: EA HaK(o) + 0222 —> K 0222-Na— (II9) L) 168 produces a different product<23). The replacement of the Na+ by K+ in the ligand occurs because C222 binds K+ more strongly than Na+ by a factor of about 100 in water and presumably also in ethylamine. The ligand may also influence the products in a dif- ferent way, namely by stoichiometry. If an equivalent amount of metal and ligand are introduced, the principal anionic species will be ES, whereas if a twofold or greater excess metal is utilized, the reaction will tend further toward the production of M’. For a more detailed discussion of the properties of these solutions, the reader is referred to the Ph.D. dis- sertation of J. M. Ceraso and the references listed in this chapter. The most definitive methods for characterizing alkali metal solutions in amines and ethers include EPR, NMR, opti- cal spectra, X-ray crystallography (if crystals can be ob- tained), and magnetic susceptibility. All the alkali metal anions, with the exception of lithium and francium, have been produced in solution or in films(23:75), and have been identified by their optical spectra. The optical spectra of the films produced by rapid evaporation of solvent (MA) from solutions of M+C222-M- or 3 give bands which are very similar to those in solution (Figure 51). The absorbance maxima for Na", K', Rb', Cs-, and 5 occur at 15,A00, 11,900, 11,600, 10,500, and 3,000 to 7,000 cm"1 respectively, in the film spectra. 169 Crystals of Na+C222 Na“ have been successfully grown and an X-ray structure performed<21'22). The crystal struc- ture indicates that Na- is a large anion, about the size of 1-, and that Na+ is trapped inside the ligand. The struc- ture is essentially a hexagonal closest-packed type with a closest (Na+) to (Na-) distance of 7.063 and (Na') to (Na’) distance of 8.83;. The films and crystals of Na+C222oNa- are diamagnetic, whereas those of K+C222-E produced by evap- oration of MA from a solution of K+C222 + E are strongly paramagnetic<76). Perhaps the strongest evidence that the alkali metal anions are indeed anions with electrons in the outer s- orbitals, rather than contact ion pairs between the cation and electrons, comes from the NMR spectra of these species in solution<20). The spectra present very narrow lines for the anions (except for potassium which has not been ob- served), which are diamagnetically shifted very near to the position of the gaseous neutral atom. This observation de- mands that the solvent be excluded from contact with inner p-electrons by filled outer s—orbitals. The extreme nar- rowness of the lines also indicates a very symmetrical elec- tron distribution, since the relaxation of alkali metals is governed by quadrupolar effects. Quadrupolar relaxation is very sensitive to the symmetry of the electronic environ- ment of the nucleus, thus the NMR spectra of the alkali metal anions provide strong evidence that they are indeed anions, with a symmetrical electron distribution, rather 170 than contact ion pairs between the cation and electrons which would produce large anisotropies in the electronic distribution. This chapter is devoted to the characterization of alkali metal amine and ether solutions with special empha- sis on the relative thermodynamic stability of the alkali anions and the preparation of new structures. The data which will be presented have been obtained from alkali metal NMR spectra using the DA-60 spectrometer described in the Experimental section. The NMR spectra provide a definitive method by which to characterize the species in these solutions. Table 21 gives the chemical shifts of several alkali metal cations and anions in solution, with each chemical shift value referenced to the free metal atom in the gas phase<79). Of particular interest is the fact that the Cs0222 chemical shift may vary from -590 ppm to -400 ppm depending on the temperature and solvent<55). This phenomenon has been ascribed to a rapid exchange be- tween inclusive and exclusive complexes (as described in the Introduction to Chapter 3), the shift of the inclusive com- plex being -590 ppm, and that of the exclusive crown-type complex being about -400 ppm. The samples have been synthesized by the procedure described in the Experimental section, except that metal alloys have been utilized sometimes rather than single metal films. Common characteristics of all samples are their blue to black colors, an evidence of the presence of M- and e 3;“ 171 882w 882I <2 mmI I88 .8+88 mo.o 8+88 882w oamI 222 omI I88. 8 88 888.8 8+M8 o2 m.mmI 228 28I :m 2.8 I 88 mv 8. memI 882 mm 288 2.8 I88 m2 8. 82 I 228 88I I82.o+sm 2.8 I82 omm 8.88 I <8 88I I82 .8+om 2.8 I82 8882 oomI 28m28 mm I2. 8+om 8.8 8+8m oom2 memI 882 mm I2 8 82 2.8 8+82 mm2 8.82NI 0mm mm 82 2.8 +82 mv m.m + 228 2 I I82.8+82 m2.8 I82 mIm 8.2 + <2 2+ o8 82I I82.8+82 m2.o I82 22 2.2 + <2 m2I I82.8+82 m2.o I82 2882888 8.m + 888 I82 2m 2.8m I 2mg 8 I I82.8 82 m2.o 8+82 882Iom2 8.8m I <8 2+ o8 82I I82.8+M2 82.8 8+82 8.8m 8.82 I <2 m2I I82. 8 m2.o 8+82 8.mm m.mm I 288 mm 822M82 8.8 +82 8.82 8.88 I <2 mm 282 mm.o +82 8.8 «.88 I <2 m2I 282 m.o +82 8.8 8.28 I 8mm mm 2882 888. +82 No.8 m.m w m.88mI 882 mm 2283 +88 8.2 8 8.22mI ow: mm 2283 +82 82.8 2 8 8.88 I o 2 mm 2283 +82 8 Sam 32 .238 Km: .m> >Homzvo acme/How Do .9289 E ..ocoo COH .822 28882 228228 :2 8888238822 888 888288 28822888 8o 8822 88888288 < .28 82888 172 solution, and considerable bronze to gold colored films upon solvent evaporation. M' and E absorb in the red, S therefore their solutions appear blue. Films of these spe- cies are gold and metallic-looking because they absorb strongly in the near-infrared and have a sharp cut-off in the visible, similar to the absorption spectra of metals. II) THERMODYNAMIC STABILITY OF THE ALKALI METAL ANIONS The measurement of the relative thermodynamic stabil- ity of the alkali metal anions involves several complex factors. It is very difficult to determine whether the products of these reactions are the most thermodynamically stable, or whether they possess kinetic stability, e.g., the kinetics of the reaction may favor one product which is much less thermodynamically stable than another. The lat- tice energy of the metal, alloying effects, the formation constant of the cation and ligand, the desire of the anion to dissociate to cation and electrons, etc., may each in- fluence the thermodynamic stability of the anion in solu- tion. The measurement of the relative thermodynamic stabil- ity of the alkali anions has been attempted in spite of the many pitfalls, by using mixed metal alloys. The metals were always in at least a twofold excess to C222. The prod- ucts of these reactions are given in Table 22. Tetrahydro- furan has been utilized because it tends not to support ES, and yet provides good solubility for the alkali metals in the presence of C222. Table 22. preparations. The products of some alkali metal solution Sample Conc. Identification by NMR NaK alloy (x8) 20.05 M Na- at+2 ppm 0222 inHtNH2 no NaC NaK alloy (xS) 20.07 M Ha’ at+2 ppm C222 in EtNH2 no NaC NaK alloy (xS) 20.07 M Na” at+fl ppm C222 in THF no NaC NaK alloy (xS) =0.05 M x0222+ at -7 ppm C222 in MeNH2 no K“, very broad K metal (xs) 20.08 M nothing observed 0222 in THF ... K, 0222 in THF 20.12 M K 0222 at 2-20 ppm very broad, no K‘ 03, 0222 in EtNH2 0.05 M 030+ at =-u33 ppm no Cs" Cs, 0222 in MeNH 0.06 M 030+ at -598 ppm, temp. dep. 2 no Cs , decomp. at -u30 0s, 0222 in THF 0.08 M 030+ at -58A and Cs’ at ~N3 ppm , + NaCs alloy 0.06 M CsC at -A18 ppm, C222 in EtNH2 no Cs or Na- NaCs alloy alloying effect to be C222 in THF discussed later KCs alloy 0.06 M Cs" at 2&7 ppm, C222 in THF ' no CsC222 Cst alloy 0.06 M Cs' at -u§ ppm, C222 in THF no CsC222 RbNa alloy 0.05 M Na' at =3 ppm, 0222 in THF no Mac222+ Na metal 15-0-5, MeNH2 Na 18-C-6 THF decomp., nothing observed very slow dissolution 17“ Table 22 (cont'd.) Sample Conc. Identification by NMR Na 18-0-6 Mac+ at -70 ppm, Na‘ at MeNH2 +1 ppm EPR gives temp. dep. elec- tron signal; g - 2.002A, linewidth - 0.2 gauss Na 18-C-6 dissolution or solubility EtNH2 problems; powder (metal + crown) observed Na C221 EtNH2 very slow dissolutiog, Na" at 3 ppm, no NaC Na C221 THF very slow dissolution, nothing observed NaCs 0221 THF Na“ at 2 ppm, 030* at -56A ppm; Cs' at -A3 ppm. Very slow dissolution of Na metal. Cs in EA nothing in NMR Cs in THF nothing in NMR Cs 18-0-6 in THF 0.11 M ? CsC+ at -322 ppm very broad at -100, disap- pears at higher temps. Na 0222+-Na’ solid Na metal observed Na metal film, Distilled over meta1--gave a blue solid. 4. .. Cs C322-Cs in THF Cs+C322 at -u39 ppm, Cs- at —50 175 An equimolar alloy of Na and K metals with C222 in THF gives complete conversion to Na“, but no Na+ is observed in the Na23 NMR spectrum of this solution. The K39 NMR spectrum has been taken of the sample in MA, which gives only one broad resonance, at -6.5 1 0.5 ppm (referenced to sat. KNO2 in H20), corresponding to K+C222. These findings suggest that the maJor product of the reaction is K+0222oNa'. Optical spectra of films from solutions prepared in the same (23). Rb-Na alloys give a manner support this conclusion similar product, Rb+C222oNa'. These two findings are entire- ly expected, since both K+ and Rb+ have a greater affinity for C222 than does Na+<8l). Also, Na“ has been shown to be more thermodynamically stable than the other anions, since Na+ is converted to Na” in the presence of K', Rb”, Cs", or 83(79)° The products of Cs-K and Cs-Rb alloys with 0222 in THF yield the production of Cs- as evidenced by their C8133 NMR spectra, but the NMR signal of 03+ is never observed in these solutions. This fact leads to the conclusion that the maJor products are K+C222-Cs" and Rb+C222oCs- respectively, again in line with expectation. The NMR spectra of K39 and both isotopes of Rb are difficult to obtain due to broad lines and low sensitivity. Thus, the measurement of their relative stability is not possible by this technique. An unusual dilemma is presented in the solutions from the alloys of Na and Cs, C222 in THF (Table 23). The compo- sition of the alloy tends to have a pronounced effect on 176 Table 23. The products of the reaction of Na and Cs metals and 0222 in THF. Sample in THF Products NaCs (excess metal) CsC+, 20% Cs-, 80% Na- NaCs (excess metal) CsC+, 20% 03', 80% Na- over Na metal first, + _ _ then over Cs metal CsC -Cs , slight Na over Cs metal first, + _ then over Na metal CsC -Na only over Na metal (gold powder), + _ _ then over Cs metal CsC °Cs , slight Na NaC+-Na' powder over Cs metal NaC+oNa-, slight Cs+C whether the maJor anion produced is C3" or Na”. This occurs because Cs and Na metals form a nonideal alloy, with the production of the compound, Na2Cs(80). The solution in THF is further complicated by the fact that the dissolution of Na is very slow, although it is enhanced by the presence of Cs. Several attempts have been made to overcome this alloy- ing effect, including the formation of Cs+C222-Cs- over Cs metal first and then taken over Na metal, and the use of Na+C222oNa' powder dissolved in THF and taken over Cs metal. The former experiment gives Cs+C222oNa- and the latter, most- ly Na+0222-Na', although some Cs+C222 is also produced. When these solutions are made up over the alloy, which is in great excess to C222, the products of the process usually are Cs+C222 and a mixture of 03' (lo-20%) and Na‘ (80-90%). The equilibria which control this process are: 177 Kl + _ 2 31(3): M + M (50) - + K2 - + C8 + Na :Na + CS (51) + K3 + M + C M C (52) The constant, K1, is not known because of the alloying effect and K2 is the one of interest, since it is deter- mined by the anion's relative stability. It appears that the equilibrium is not shifted too far to the right, but definite conclusions are difficult because of the slow dis- solution of Na in THF and by the alloying effect. Another very unusual consequence of this process is that the equilibrium constant, K3 (which is also influenced greatly by K1), tends almost exclusively toward the forma- tion of Cs+C222 rather than the Na+ complex. This is highly irregular because in water, the formation constant of 0222 with Na+ is more than 100 times greater than that of Cs+(81). This may be entirely a function of the alloying effect or possibly even a kinetic phenomenon. The formation constants of the complexes have not been measured in THF or the amines but that of Na+ would still be expected to be much greater than that of Cs+. This competitive complexa- tion should be explored in more detail in THF. The trend of thermodynamic stability relative to the solid metals from these NMR data indicate Na’>Cs’>>K',Hb'. 178 III) IDENTIFICATION OF NEW SPECIES Attempts to observe K“ by_NMR. The potassium anion has never been observed by NMR for two reasons, low sensi- tivity of the potassium nucleus and considerable affinity to dissociate to and interact with ES. Even a slight inter- action with electrons in solution tends to broaden the sig— ' nal out of existence (see Historical section). Four at- tempts have been made to observe K-, three in THF with ex- cess metal at concentrations varying from 0.01 M to 0.1 M. The samples were prepared at or near room temperature. very dark films were observed, but the NMR spectra showed only the complexed cation at about -20 ppm from saturated KNO2 in water, the line being 300 to 600 Hz broad, depend- ing on the temperature. The final attempt to obtain a K- signal in the NMR utilized IPA as solvent with ligand concentration at about 0.02 M. The sample was again made up at -10 to 0°C with no apparent decomposition. A black powder precipitate was ob- served which seemed to be somewhat crystalline (black nee- dles), as previously reported by Dye and co-workers(76). The NMR spectrum showed no peaks, an indication of either low solubility or the presence of solvated electrons which broaden both cation and anion peaks. Otherwise, the concen- trations should have been high enough to permit observation of the NMR spectra. Cs, 0222 in EA and MA. When Cs is dissolved in EA or MA in the presence of C222, only the complexed cation is 179 observed in the NMR spectrum at -A18 to -A33 ppm (depending on sample conditions and temperature) and -591 ppm respec- tively. The chemical shifts indicate that the complex in EA is an exclusive complex, whereas that in MA is inclusive. Both solutions are somewhat unstable, and decomposition peaks, presumably corresponding to Cs+ complexed by the de- composed ligand, occur at -384 and -A1A ppm in EA and MA. The linewidth of the complexed cation is somewhat peculiar in both solvents, since it broadens as the tempera- ture is raised in MA, while the opposite is observed in EA. Figure 52 shows this temperature dependence in MA. The chemical shift of this species is rather temperature in- variant in MA, but in BA the line shifts slightly downfield with decreasing temperature. When the ethylamine solution is frozen, the complexed cation resonance shifts to -580 t 30 ppm, corresponding to an inclusive complex (Table 2A). Table 2A. The temperature dependence of the chemical shift of Cs+0222 in BA. Temperature (°C) GPPm+ AV1/2* -9A -580 81000 Hz -69 -N33 108 -62 -u32 83 -52 -A28 3A -38 —A27 16 + i A ppm. * Full width at half height. 180 )\_jk-~ "HDB ' _] T I Sppm '800-600-400 Figure 52. The temperature dependence of the NMR spectrum of Cs+0222-Cs' in MA. 181 The EPR spectrum of a sample prepared in a similar way in EA gives a single, narrow, intense line with g = 2.002“ t 0.0002 between a temperature range of -32 to -70°C. This spectrum indicates that electrons are present, but that they do not interact to a large extent with the Cs nucleus. The electronic spectra of films from solutions in MA prepared in a similar way(23), show nearly complete con- version to Cs- with only slight presence of § The con- 8' centration of these samples was much lower than those for NMR analysis, however. All indications suggest that Cs- should be observed in the NMR spectrum in MA and EA, but it is possible that exchange between a paramagnetic species in minute concen- trations could account for its disappearance. This paramag- netic species could also be the source of the instability of these solutions. The strange Cs+C222 broadening with temperature is probably due to the rapid exchange between inclusive and ex- clusive complexes, the former being preferred at lower tem- peratures. In EA the exclusive crown-type complex is fa- vored, whereas in MA the opposite is true. The spectra in EA are especially supportive of this conclusion, because as the temperature is lowered, the Cs+C222 line broadens and shifts downfield. When the sample is frozen, the chemical shift of the complexed Cs+ is nearly identical to that of the inclusive complex. In methylamine, the equilibrium 182 seems to highly favor the inclusive complex, but above -90°C the exclusive complex is present in high enough con- centration to cause considerable exchange broadening of the Cs+C line. An interesting sidelight to these observations con- cerns the dissolution of Na metal and 0222 in EA which gives Na+C222oNa-. However, the solution prepared from the alloy of Na and Cs metals in BA with C222 gives a Cs+0222 signal at about -A18 ppm, but no anion or Na+ species is observed. Cs metal is soluble in EA to 10’3 or 10'” M(76), and gives rise to solvated Cs monomers. Since these species are para- magnetic, they would cause tremendous broadening and shift- ing the resonances of those ions with which they exchange or interact strongly. Only minute levels of these species would be necessary to induce the Cs- disappearance, because their chemical shifts are normally tens of thousands of ppm downfield. Cs and C222 in THF. Samples of Cs metal and C222 in THF seem to be much more stable than in MA and EA. An esti- mated half-life for decomposition in THF is approximately 12 hours. The NMR spectrum of these samples shows both Cs+C222 at -583 to -587 ppm and Cs— at -A6 to -Al ppm. The Cs+C222 line is about 100 Hz broad at -110°C, and broadens greatly as the temperature is raised (Figure 53). Meanwhile, the Cs- line remains very narrow (<10 Hz), and indicates no ex- change broadening, except at -A1°C, where slight broadening occurs. These spectra are reproducible upon raising and 183 THF C222 + Cesium /\ -74 4‘ - L I CsIC+ C5- .1. L 11 1 L. l L 11 -600 -590 -580 -570 -60 -50 -4O -30 ppm ppm Figure 53. The tgmperature dependence of the NMR spectrum of Cs 0222-Cs" in THF. 18A lowering the temperature, and are not symptons of decompo- sition. Above -A1°C, however, decomposition occurs rapidly. This exchange phenomenon is peculiar for two reasons. First, the chemical shift of the complexed Cs+ does not change appreciably throughout the temperature range. Sec- ondly, the exchange does not significantly involve the Cs' because it would be expected to broaden by exchange to the same extent as the cation resonance. The chemical shift of the Cs+C222 resonance demon- strates that it is an inclusive complex, and its temperature characteristics are similar to those of its counterpart in MA, which is also an inclusive complex. By inference from the exchange in MA, the line broadening of the Cs+C222 resonance in THF might be assigned to the rapid interconver- sion of inclusive and exclusive complexes with the latter present at lower concentrations which increase with increas- ing temperature. Above -A1°C, the Cs+C222 resonance is very broad indicating that the exclusive complex is appreciable. Since this complex allows more efficient interaction between Cs+ and Cs', the anion peak also begins to show evidence of the exchange phenomenon. In order to verify that the exchange process involving the complexed cation was the result of two rapidly intercon- verting conformations of the ligand, rather than an electron transfer reaction or related process, the temperature Char- acteristics of a simple salt, cesium octanoate, were studied. (Most cesium salts are insoluble in THF, and only cesium 185 octanoate has been found to be soluble enough for C8133 NMR analyses.) The chemical shift of the cesium cation in solution of cesium octanoate, C222 in THF comes at -403 to -A33 ppm, and indicates no complex formation. This obser- vation further demonstrates that complexation of Cs by C222 in THF is easily perturbed by solvent, temperature, and anion. Another possible explanation for the exchange broaden- ing of the Cs+C222 line would involve its interaction with ES or cesium monomers, both of which are paramagnetic. Ex- changes of this sort would also be expected to involve Cs- as in MA and EA, and since this is not observed, interac- tions of this kind are doubtful. Na and 18-0-6 in MA. Andrews, Ceraso, and Dye(3“) described a peculiar two electron exchange process between Na+ complexed by 18-crown-6(l8-C-6) and Na- in MA. The ex- change was followed by line broadening in the NMR, as shown in Figure 58. This process did not seem to originate from a simple two-site exchange, but a more complex reaction was indicated. In order to ascertain whether a paramagnetic species could give rise to a third site, an EPR analysis has been performed. A sample prepared from 0.08 M C222 in MA taken over excess Na metal showed a single narrow line with g = 2.002“. The intensity of the line was very temper- ature dependent (Table 25), indicating a process: Na-: Na+ + 2E8 or 52:28' (53) 186 Figure 5A. The t$mperature-dependence of the NMR spectrum of Na 18-C-6-Na in MA. 187 Table 25. The temperature dependence of the_EPR signal of a 0.08 M solution of Na 18-C-6-Na in MA. Temp, °K Intensity 269 18.1 259 16.2 246 10.5 237 5.1 227 $0.2 In either case, the presence of ES could greatly perturb the NMR spectra. The synthesis of crystals of Cs+C222-Cs-. Gold col- ored crystals, very similar in appearance to those of Na+C222oNa' were synthesized from a 20.1 M solution of 0222 in IPA, taken over excess metal at 0. to -20°C. As the solution was cooled to -78°C, the small crystals precipi- tated, and seemed to be quite stable if kept at ~78°C. The solution phase was less stable, and turned from deep blue to clear upon decomposition, but the crystals remained gold for several days at dry ice temperature. The crystals were never isolated from the solution phase because decomposi- tion occurred too rapidly. Attempts to grow crystals of other alkali metals and ligands were also made, but without success. They include Cs with 18-C-6 in MA and K+C222-E (1:1 stoichiometry of C222 and metal) from MA, with the addition of DEE. A black 188 powder precipitated from the latter when the MA was removed and DEE was added. Complexes With Cryptands Other Than C222. C221 has also been utilized to dissolve Na metal in EA and THF as well as Na-Cs alloy in EA. The rate of dissolution of Na metal by C221 is extremely slow in EA and THF and therefore, the products of these reactions may be greatly dependent on the kinetics of dissolution. The products of the reaction of Na-Cs alloy and C221 in EA are reflective of the slow dissolution of the Na metal since mostly Cs+C221 and Cs' are formed. The resonance assigned to the Cs+C22l complex at -56A ppm was very broad and may not have been real. A sample has also been prepared by distilling 0221 over a fresh film of Na metal. Upon contact with the metal, the C221 forms a dark blue crust which lasts for weeks at -78°C. The NMR spectrum of a solution from Cs metal and C322 in THF showed two resonances, one at -A39 ppm (1000 Hz broad) and the other at -50 ppm corresponding to Cs+C322 and 08' respectively. The chemical shift of the cation is normal since it is in the range of inclusive complexes of simple salts of Cs+ with C322<85). APPENDIX THE THEORY OF CHEMICAL EXCHANGE IN NMR APPENDIX The equations which describe the NMR lineshape of a system undergoing chemical exchange of the sort: kr A + L.¢——>.AL kb (A1) are those of Bloch, modified by McConnell to include chemi- cal exchange. These equations describe the X and Y compo- nents of the magnetization in the rotating frame of refer- ence as follows: dG A _ -1 -1 Et— + GAGA - -1YH1M0A 4' TB GB - TA GB (A2) d6 __5. - -1 -1 These equations are solved using the slow passage condition in which: dGA dGB —z—=O dt dt thus: -1 -l -(uA - TA )GA + TB GB (AA) = 17H (A5) lMOA 189 190 1 -(ab - ‘61)Gb + 1; Ca = -iYH1Mob (A6) The expressions are further simplified by: M0A = pAMo (A7) MOB 2 pBMO (A8) and by arranging in a matrix form: alch + 31203 2 CDA (A9) a21GA + a220B a CpB (A10) Using Cramer's rule (see any matrix algebra text): GA -.% :2 ::: 8 g-IEpAa22 - pBalz] (All) GB 3'% ::: E: ' % [pBall ' pAa21] (A12) D = :ll :12 (A13) 21 22 The total complex moment, 0, is the sum of the two: C The absorbtion lineshape function is given by the imaginary portion of 0, but for lines which are imperfectly phased: g(v) 2 (real G) sine + (imaginary G) cose (A15) 191 where 6 is the angle of zero order phase correction. These equations may be easily expanded to include three or more sites and are equivalent in both P-FT and CW modes‘7o’71’83) in most cases. They may be used to simulate or fit NMR lineshapes using KINFIT. (The structure of the input deck, subroutine EQN, may be obtained from my first notebook on pages 18A and 185, the entry dated August 25, 1977.) The mean lifetimes at sites A and B, TA and TB, are defined in the following way: the rate of removal of a nucleus from site i by exchange (A16) l/‘t1 = the total number of nuclei in site i and is related to the rate constant of reaction A1 in the following way: 1/rA = kaAJEL] (A17) 1/rB = kbEAL] (A18) TA and TB are not unique, but are coupled by the expression: TATB T ' ¥XI?§ (A19) and since: PA = I :fi (A20) 192 we have: t a p T a p 1 (A21) A B B A and only T is unique. The temperature dependence of the rate constant, k, may then be utilized to calculate an activation energy for the process, using the Arrhenius equation: k - Ae‘Ea/RT (A22) AH* is related to the activation energy for reactions in solution by: AH+ . Ea—RT (A23) The Eyring equation may be utilized to calculate AGIr and AS+: k a K[:%1] e'AG‘lI/RT (A2") where K is assumed equal to 1.0, and AS* a AH+ ; AG+ (A25) REFERENCES REFERENCES 1) 2) 3) A) 5) 6) 7) 8) 9) 1o) 11) 12) 13) 1A) 15) D. H. Haynes, B. C. Pressman and A. Kowalski, Biochem- istry, $2, 852 (1971). F. A. Cotton and G. Wilkinson, "Advanced Inorganic Chemistry," Interscience Publishers, New York, 1966; R. Winkler, Structure and Bonding, TR, 1 (1972). R. M. Izatt, J. D. Lamb, G. E. Maas, R. E. Asay, J. S. Bradshaw and J. J. Christenson, J. Amer. Chem. Soc., C. J. Pedersen, J. Amer. Chem. Soc., 29, 7017 (1967). J. M. Lehn, Acc. Chem. Res., 11, N9 (1978). E. V. Dehmlow, Angew. Chem., Int. Ed. Engl., 16, A93 (1977) M. Newcomb and D. J. Cram, J. Amer. Chem. Soc., 97, 1251 (1975) E. Blasius and P. G. Maurer, Makromol. Chem., 118, 649 (1977). E. Blasius, W. Adrian, K. P. Janzen and G. Klautke, J. Chromatog., 96, 89 (l97h). S. KOpolow, T. E. Hogen Esch and J. Smid, Macromole- cules, Q. 133 (1973). W. D. Curtis, D. A. Laidler, J. F. Stoddart and G. H. Jones, J. Chem. Soc. Chem. Comm., 833 (1975). D. J. Cram and J. M. Cram, Science, 183, 803 (197“). E. V. Kyba, R. C. Hegleson, K. Madan, G. W. Gokel, T. L. Tarkowski, S. S. Moore and D. J. Cram, J. Amer. Chem. 306-. 99. 256“ (1977). mm G. A. Rechnitz and E. Eyal, Anal. Chem., ii, 370 (1972). J. W. Mitchell and D. L. Shanks, Anal. Chem., A7, 642 (1975) 193 16) 17) 18) 19) 20) 21) 22) 23) 2A) 25) 26) 27) 28) 29) 3o) 31) 32) 33) 19“ G. Czerwenka and E. Schenbeck, Z. Anal. Chem., 276, 37 (1975). -~ B. Kaempf, S. Raynal, A. Collet, F. Schue, S. Boileau and J. M. Lehn, Angew. Chem. Int. Ed. Engl., 13, 611 (197“). ~~ C. J. Redersen and H. K. Frensdorff, Angew. Chem. Int. Ed. Engl., 11, 16 (1972). 'Vh R. M. Costes, G. Folcher, P. Plurien and P. Rigny, Inorg. Nucl. Chem. Lett., 12, H91 (1976). J. M. Ceraso and J. L. Dye, J. Chem. Phys., 61,1585 (197“) J. L. Dye, Scientific American, 237, 92, (1977). W F. J. Tehan, B. L. Barnett and J. L. Dye, J. Amer. Chem. Soc., 96, 7203 (197“). M J. L. Dye, M. R. Yemen, M. G. DaGue and J. M. Lehn, J. Chem. Phys., 68, 1665 (1978). M J. M. Lehn and J. Cheney, private communication. B. G. Cox and H. Schneider, J. Amer. Chem. Soc., 99, 2809 (1977) J. Cheney, J. P. Kintzinger and J. M. Lehn, to be pub- lished. J. Cheney and J. M. Lehn, J. Chem. Soc. Chem. Comm., #87 (1972). N. F. Ramsey, Phys. Rev., 7%, 699 (1950). Ibid, 86, 293 (1952)- Ibid, 2%, 303 1953). N” D. Shaw, "Fourier Transform N.M.R. Spectroscopy," Elsevier Scientific Publishing Co., New York, 1976. D. J. Farnum, Advances in Physical Organic Chemistry, 11, 129 (1975). J. A. Pople, W. G. Schneider and H. G. Bernstein, "High Resolution Nuclear Magnetic Resonance," McGraw-Hill, New York, 1959. T. C. Farrar and E. D. Becker, "Pulse and Fourier Transform NMR," Academic Press, New York, 1971. A. Abragam, "The Principles of Nuclear Magnetism," Oxford Univ. Press, New York, 1961. 3A) 35) 36) 37) 38) 39) A0) A1) A2) A3) AA) “5) A6) A7) A8) A9) 50) 51) 195 J. L. Dye, C. W. Andrews and J. M. Ceraso, J. Phys. Chem., 12, 3076 (1975)- P. Debye, "Polar Molecules," Dover Publications, New York, 19A5. E. H.)Mei, Ph. D. Thesis, Michigan State University 1977 . D. D. Traficante and J. A. Simms, J. Mag. Res., 15, A8A (197A). J. M. Ceraso, Ph. D. Thesis, Michigan State University (1975). B. K. J. Leong, T. 0. T. Ts'o and M. B. Chenoweth, Toxicol. Appl. Pharmacol., 27, 3A2 (197A). mm J. M. Lehn, J. P. Sauvage and B. Dietrich, J. Amer. Chem. Soc., 92, 2916 (1970). “A: B. Dietrich, J. M. Lehn and J. P. Sauvage, Tetrahedron, £2, 16A7 (1973). B. Dietrich, J. M. Lehn, J. P. auvage and J. Blanzat, Ibid, p. 1629. C. H. Park and H. E. Simmons, J. Amer. Chem. Soc., 9A, 718A (1972)- H. E. Simmons and C. H. Park, J. Amer. Chem. Soc, 90, 2A28 (1968). C. H. Park and H. E. Simmons, Ibid.,'W p. 2A29. A. C. Coxon and J. F. Stoddart, J. Chem. Soc., Perkin B. Metz, D. Moras and R. Weiss, Chem. Comm., AAA (1971). M. T. Lok, Ph. D. Thesis, Michigan State University (1973). V. M. Loyola, R. Pizer and R. G. Wilkins, J. Amer. Chem. Soc., 22, 7185 (1977). R. Pizer, J. Amer. Chem. Soc., 100, A239 (1978). E. Schori, J. Jagur-Grodzinski and M. Shporer, J. Amer. Chem. Soc., 95, 38A2 (1973). J. M. Ceraso, P. B. Smith, J. S. Landers and J. L. Dye, J. Phys. Chem., 81, 760 (1977). Y. M. Cahen, J. L. Dye and A. I. Popov, J. Phys. Chem., 79. 1292 (1975). 52) 53) 5A) 55) 56) 57) 58) 59) 6o) 61) 62) 63) 6A) 65) 66) 67) 68) 196 E. Schori, J. Jagur-Grodzinski, Z. Luz and M. Shporer, J. Amer. Chem. Soc., 23, 7133 (1971). E. Kauffmann, J. M. Lehn and J. P. Sauvage, Helv. Chim. Acta, 22, 1099 (1976). E. Mei, A. I. Popov and J. L. Dye, J. Amer. Chem. Soc., P. B. Smith and J. L. Dye, unpublished results. F. Mathieu, B. Metz, D. Moras and R. Weiss, J. Amer. Chem. Soc., 100, AA12 (1978). R. W. Taft and L. S. Levitt, J. Org. Chem, A2, 916 (1977). J. F. Wolf, R. H. Staley, I. Koppel, M. Taagepera, R. T. McIver, Jr., J. L. Beauchamp and R. W. Taft, J. Amer. Chem. Soc., 22, 5A1? (1977). F. Bloch, Phys. Rev., 12, A60 (19A6). H. M. McConnell, J. Chem. Phys., 28, A30 (1958). see also J. A. Pople, W. G. Schneider gand H. J. Bernstein, "High- Resolution Nuclear Magnetic Resonance," McGraw- Hill Book Company, New York, 1959, Chapter 10. Equa- tion (lo-25) of the previous reference is in error. G. J. Karabatsos and C. E. Orzech, Jr., J. Amer. Chem. Soc., 86, 357A (196A). see the Historical section. G. Binsch, J. B. Lambert, B. W. Roberts and J. D. Roberts, J. Amer. Chem. Soc., 86, 556A (196A). G. A. Olah and D. H. Obrien, J. Amer. Chem. Soc., 89, 1725 (1967). M. Eigen, Discussions of the Faraday Society, 39, 1965, pp. 7-11. A. Albert and E. P. SerJeaut, "Ionization Constants of Acids and Bases," Butler and Tanner Ltd., London, 1962, p. 12A. P. K. Glasoe and F. A. Long, J. Phys. Chem., 6A,188 (1960) W. J. Moore, "Physical Chemistry," Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1972, p. 383f. 69) 7o) 71) 72) 73) 7A) 75) 76) 77) 78) 79) 80) 81) 82) 83) 8A) 85) 86) 197. C. P. Schlicter, "Principles of Magnetic Resonance," Harper and Row Publishers, New York, 1963. R. K. Gupta, T. P. Pitner and R. Wasylishen, J. Mag. Res., 13, 383 (197A). D. E. Woessner, J. Chem. Phys., 35, A1 (1961). V. A. Nicely and J. L. Dye, J. Chem. Edu., A8, AA3 (1971) see Appendix for details of the calculations. W. Weyl, Ann. Physik., 121, 601 (1863). W J. L. Dye, Pure and App1. Chem., A9, 3 (1977) and references therein. J. L. Dye, submitted as a chapter for a Wiley- Interscience book on "Multidentate Macrocyclic Mole- cules," J. J. Christensen and R. M. Izatt, editors. B. Kaempf, S. Raynal, A. Collet, F. Schue, S. Boileau and J. M. Lehn, Angew. Chem. internat. edit., 13, 611 (197AL J. L. Dye, L Chem. Ed., 5A, 332 (1977) and references therein. J. L. Dye, C. W. Andrews and S. E. Mathews, J. Phys. Chem., 12, 3065 (1975). M. Hansen, "Constitution of Binary Alloys," 578, 1958. J. M. Lehn and J. P. Sauvage, Chem. Comm., AAO (1971). E. Mei, A. I. Popov and J. L. Dye, J. Amer. Chem. Soc., 22. 6532 (1977). D. Shaw, "Fourier Transform NMR Spectroscopy," Elsevier Scientific Publishing Company, New York, 1976, pp. 85- 87. J. A. Pople and D. L. Beveridge, "Approximate Molecular Orbital Theory," McGraw-Hill Book Company, New York, 1970. E. Kauffmann, J. L. Dye, A. I. Popov and J. M. Lehn, unpublished results. J. H. Busch and J. R. de la Vega, J. Amer. Chem. Soc., 22. 2397 (1977). 198 87) D. F. Shriver, "The Manipulation of Air-sensitive Com- pounds," McGraw-Hill Book Company, New York, 1969, pp. lu6‘15uo