PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K:/Proj/Acc&Pres/CIRC/DaIeDue Indd AQUEOUS PHASE CATALYTIC HYDROGENATION AND DEUTERIUM EXCHANGE OF a-SUBSTITUTED ORGANIC ACIDS USING Ru/C By Jennifer Elizabeth Jacobs A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2009 ’ ABSTRACT AQUEOUS PHASE CATALYTIC HYDROGENATION AND DEUTERIUM EXCHANGE OF a-SUBSTITUTED ORGANIC ACIDS USING Ru/C By Jennifer Elizabeth Jacobs The biomass refinery of the future will require a wide range of catalytic conversions, largely aqueous phase reductions. This dissertation explores the use of a 5% ruthenium on carbon catalyst in hydrogenation and C-H activation studies of carboxylic acids, a significant class of biogenic feedstocks. Several two- and three-carbon organic acid substrates were studied to probe (a) the effects of a- substituents on hydrogenation and H/D exchange reactions and (b) the relationship between these two processes. Substrates included the following groups: -H, -CH3, -OH, -N+H3, -OCH3, -N+H2CH3, -N+H(CH3)2, and -N+(CH3)3. Most reactions were performed at 100 °C or 130 °C in D2O solution but ranng from 50 °C to 150 oC and 1000 psi H2 or D2. As expected, both catalyst and H2 gas are required to convert organic acids to the corresponding alcohols. Simple alkanoic acids (acetic, propanoic, and isobutyric) and betaine, ((CH3)3N+CH2COOH-Cl'), a quatemized glycine, are very unreactive. For the other methylated glycines, the rate of hydrogenation falls off by a factor of three with each additional methyl, and reduction requires acid, as found earlier by Jere. Use of water as reaction medium appears optimal; addition of organic co-solvents such as ethanol, ethylene glycol, THF, or ethyl lactate led to reduced hydrogenation rates. This apparent inhibition by organic additives may reflect their tendency to tie up catalytic sites via metal insertion into C-H bonds, especially those neighboring heteroatoms. Another feature of both the solvents and the substituents is their capacity for hydrogen bonding, which may play a role in modulating hydrogenation rates. Similarly, steric crowding may also affect the substrates’ ability to access the surface in productive orientations. To explore this issue, activation of OH bonds next door to amine N sites was probed via H/D exchange studies performed on the N- methylated glycine series. Reactions run in D20 with either H2 or D2 and catalyst were effective at exchanging deuterium into the CH2 and CH3 sites of the amino acids. The Ru catalyst also exchanges H (from H2) for D (from D20) to form HOD and HD, independent of substrate. Analysis by internal standard-calibrated 1H NMR allowed complete isotopomer-specific inventory of H concentration in substrates and water. Simple sequential replacements of H by D were observed in the amino acids. Two kinetic models were developed to determine the rate constants for H/D exchange at both the methylene and methyl positions. Both included possible time delays due gas/liquid isotopic equilibration (kH) and/or a catalyst induction (ktum_on) period; ultimately the latter was discarded as redundant. The first model treated each isotope replacement step as a process unrelated to the others. Thus, between forward and reverse replacements in CH2 and CH3 sites and k”, eleven rate constants were freely varied in the fitting. The second model used a single rate constant for each site, further modified by primary and secondary isotope effect values, resulting in six values + kg for this fitting. The latter, more economical and interpretive model gave fits nearly as good as those from the eleven-variable one. Somewhat surprisingly, these analyses found H/D exchange rates to decrease in the following order: D2O > sarcosine (CH3NH2+CH2CO2') > glycine (CH2 site) > N,N-dimethy1glycine. TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. vi LIST OF FIGURES .......................................................................................................... vii LIST OF SCHEMES ........................................................................................................ xiv LIST OF EQUATIONS ................................................................................................... xvi Chapter 1. INTRODUCTION ............................................................................................ 1 1.1 Background ............................................................................................................... 1 1.2 Significance .............................................................................................................. 1 1.3 Literature Review ..................................................................................................... 4 1.3.1 Hydrogenation/Dehydrogenation ...................................................................... 4 1.3.2 H/D Exchange ................................................................................................. 18 1.3.3 Alkynylation of sp3 C-H Bonds Adjacent to a Nitrogen Atom ...................... 33 1.3.4 Secondary Deuterium Isotope Effects for Enolization Reactions ................... 35 1.3.5 Catalytic Hydrogenation of on-Amino and a-Hydroxy Esters ........................ 37 1.3.6 Catalytic Properties of Ru/C ........................................................................... 39 1.3.7 Hydrogenolysis of Glycerol to Propylene Glycol ........................................... 40 1.3.8 Deuterium NMR ............................................................................................. 43 1.4 Summary of Previous Work ............................................................................... 45 Chapter 2. EXPERIMENTAL ......................................................................................... 68 2.1 Hydrogenation Studies ............................................................................................ 68 2.1.1 Materials .......................................................................................................... 68 2.1.2 Batch Reactor System ..................................................................................... 69 2.1.3 Reaction Procedure ......................................................................................... 70 2.1.4 Organic Acid Analysis .................................................................................... 71 2.1.4.1 HPLC Analysis ........................................................................................ 71 2.1.4.2 pH Analysis .............................................................................................. 72 2.2 H/D Exchange Studies ........................................................................................... 72 2.2.1 Materials ......................................................................................................... 72 2.2.2 Batch Reactor System ..................................................................................... 73 2.2.3 Reaction Procedure ......................................................................................... 73 2.2.4 Amino Acid Analysis ...................................................................................... 74 2.2.4.1 HPLC Analysis ........................................................................................ 75 2.2.4.2 'H NMR ................................................................................................... 75 Chapter 3. CALCULATIONS ......................................................................................... 76 3.1 1H NMR Peak Fitting and Deconvolution ............................................................. 76 3.1.1 NMR Calibration Calculations ....................................................................... 76 3.1.2 Peak Fitting Equations .................................................................................... 77 3.1.3 Peak Integration Calculations ......................................................................... 80 3.2 H-D Content Inventory .......................................................................................... 82 iv 3.2.] Percent H in Water .......................................................................................... 82 3.2.2 Percent H in Sample ........................................................................................ 83 3.3 Kinetic Rate Data Determination ........................................................................... 88 3.3.1 Concentration Calculations ............................................................................. 88 3.3.2 Kinetic Model A: 10-variables ....................................................................... 91 3.3.3 Kinetic Model B -— 6 variables ........................................................................ 93 Chapter 4. RESULTS AND DISCUSSION .................................................................... 98 4.1 Control Experiments .............................................................................................. 98 4.1.1 Hydrogenation Studies .................................................................................... 98 4.1.2 H/D Exchange Studies .................................................................................. 102 4.2 Substrate Adsorption on the Catalyst/Carbon Support ........................................ 113 4.2.1 Adsorption on Ru/C ...................................................................................... 113 4.2.2 Adsorption on Carbon Support ..................................................................... 115 4.3 Effect of HCl ........................................................................................................ 116 4.3.1 Hydrogenation Studies ................................................................................... 116 4.3.2 H/D Exchange Studies .................................................................................. 122 4.4 Effect of H3PO4 .................................................................................................... 125 4.4.1 Hydrogenation Studies .................................................................................. 125 4.4.2 H/D Exchange Studies .................................................................................. 129 4.5 Solvent Effects on the Hydrogenation of Organic Acids ..................................... 143 4.6 Organic Acid Hydrogenation Studies .................................................................. 156 4.6.1 Effect of the a-Substituent on the Hydrogenation Reaction ......................... 156 4.6.2 Amino Acid Substrates ................................................................................. 162 4.7 H/D Exchange Studies ......................................................................................... 166 4.7.1 Amino Acid Substrates ................................................................................. 166 4.7.1.1 H/D Exchange at the Methylene Position .............................................. ' 166 4.7.1.2 I-I/D Exchange at the N-Methyl Position ............................................... 216 4.7.2 Other Substrates ............................................................................................ 237 4.8 Competition Experiments .................................................................................... 242 4.8.1 Hydrogenation Studies .................................................................................. 242 4.8.2 H/D Exchange Studies .................................................................................. 246 Chapter 5. SUMMARY ................................................................................................. 251 Chapter 6. REFERENCES ............................................................................................. 261 Table I. Table 2. Table 3. Table 4. Table 5. Table 6. Table 7. Table 8. Table 9. LIST OF TABLES Carbon characterization by N2 adsorption at 78 K ............................................ 53 Hydrogenation of C3 and C2 organic acids to alcohols after six hours (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................................... 161 Hydrogenation of glycine, sarcosine, N,N-dmg, and betaine at 100, 130, and 150 °C after six hours (0.3 M H3PO4, H2O, X °C, 1000 psi H2, 1g dry 5% Ru/C) ......................................................................................................................... 162 Rates of reaction for hydrogenation of glycine, sarcosine, and N,N-dmg (0.3 M H3PO4, H20, 100, 130, and 150 °C, 1000 psi H2, 1g dry 5% Ru/C) ............... 163 Rate constants for glycine (100 °C, 1000 psi H2 or D2) using the original kinetic model, k10 + kH, k6 + kH ................................................................................. 190 Rate constants for sarcosine (100 °C, 1000 psi H2 or D2) using the original kinetic model, k10 + k”, k6 + kH ..................................................................... 193 Rate constants for N,N-dimethylglycine (100 °C, 1000 psi H2 or D2) using the original kinetic model, k10 + k“, k6 + kH ....................................................... 196 pKa values for glycine, sarcosine, and N,N-dimethylglycine (25 °C) ............ 210 Summary of the rate kinetic rate constants and primary and secondary isotope effects for glycine, sarcosine, and N,N-dimethylglycine ................................ 236 Table 10. Summary of rate constants for glycine, sarcosine, and N,N-dimethylglycine ......................................................................................................................... 258 vi LIST OF FIGURES Figure 1. Hydrogenation of 0.25 M refluxed lactic acid (0.3 M H3PO4, H20, 130 0C, 1000 psi He, 1 g dry 5% Ru/C) ......................................................................... 98 Figure 2. Hydrogenation of 0.25 M refluxed lactic acid (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, no catalyst) ................................................................................... 99 Figure 3. Hydrogenation of refluxed lactic acid (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ...................................................................................... 101 Figure 4. Hydrogenation of 0.25 M glycolic acid in the absence and presence of 0.3 M H3PO4 (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................................... 102 Figure 5. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D2O, 130 °C, 1000 psi H2, no catalyst) .................................................................... 103 Figure 6. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 130 °C, 1000 psi H2, no catalyst) .................................................................... 104 Figure 7. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (D2O, 130 °C, 1000 psi H2, no catalyst) .................................................................... 104 Figure 8. Fit of the kinetic rate data for 0.25 M sarcosine (D2O, 130 °C, 1000 psi H2, no catalyst) ........................................................................................................... 106 Figure 9. 1H NMR of H/D exchange at the methylene position next to the —OH of 0.25 M choline bicarbonate (D20, 130 0C, 1000 psi H2, no catalyst) ......................... 108 Figure 10. lH NMR of H/D exchange at the methylene position next to the —N+(CH3)3 of 0.25 M choline bicarbonate (D2O, 130 °C, 1000 psi H2, no catalyst) ............. 108 Figure l 1. 1H NMR of H/D exchange at the methylene position of 0.25 M betaine (D2O, 130 °C, 1000 psi H2, no catalyst) .................................................................... 110 Figure 12. 1H NMR of H/D exchange at the methylene position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, 1g dry 5% Ru/C) ........................................................... 110 Figure 13. 1H NMR of H/D exchange at the methyl position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, no catalyst) ........................................................................... 111 Figure 14. 1H NMR of H/D exchange at the methyl position of 0.25 M betaine (D2O, 130 °C, 1000 psi H2, 1g dry 5% Ru/C) .................................................................. 111 Figure 15. Concentrations of CH2, CHD, and CD2 of 0.25 M betaine with and without 1 g dry 5% Ru/C catalyst (D2O, 130 °C, 1000 psi H2) ....................................... 113 Figure 16. 0.25 M Refluxed lactic acid concentrations measured (bench and reactor). 114 Figure 17. Substrate adsorption on the carbon support .................................................. 116 Figure 18. Hydrogenation of 0.25 M betaine hydrochloride (0.3 M H3PO4, H20, 100 °C and 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................................... 118 Figure 19. Hydrogenation of 0.25 M betaine after 6 hours (0.3 M H3PO4, H20, 100 °C, 130 °C, and 150 °C, 1000 psi H2, 1 g dry 5% Ru/C) ...................................... 118 vii Figure 20. Hydrogenation of 0.25 M glycine in the absence and presence of 0.3 M HCl (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ......................... 120 Figure 21. Hydrogenation of glycine alone and in the presence of NaCl or HCl (H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................. 121 Figure 22. 1H NMR of H/D exchange at the methyl (3.2 ppm) and methylene position (4.1 ppm) on 0.25 M betaine hydrochloride (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ......................................................................................................... 123 Figure 23. Fit of the kinetic rate data for 0.25 M betaine (D20, 130 °C, 1000 psi H2. 1 g dry 5% Ru/C) .................................................................................................. 124 Figure 24. Hydrogenation of 0.25 M glycolic acid in the absence and presence of 0.3 M H3PO4 (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................... 125 Figure 25. Hydrogenation of 0.25 M glycine (no H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................................................................................. 126 Figure 26. Hydrogenation of 0.25 M glycine (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............................................................................................... 127 Figure 27. Hydrogenation of 0.25 M glycine in the absence and presence of 0.3 M H3PO4 (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................................... 128 Figure 28. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (0.3 M H3PO4; D20; 50 °C; 1000 psi D2, 1 g dry 5% Ru/C) ...................................... 130 Figure 29. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D20; 50 °C; 1000 psi D2, 1 g dry 5% Ru/C) ............................................................ 130 Figure 30. Concentration data for 0.25 M glycine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) ...................................................................................... 131 Figure 31. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................. 133 Figure 32. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................. 133 Figure 33. Concentration data for 0.25 M sarcosine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) ...................................................................................... 135 Figure 34. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) ....................................... 135 Figure 35. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................. 137 Figure 36. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ....................................... 137 Figure 37. Fit of the kinetic rate data for 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ........................................................................ 138 Figure 38. Fit of the kinetic rate data for 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ........................................................................ 138 viii Figure 39. Rate constants for 0.25 M sarcosine in the absence and presence of 0.3 M H3PO4 (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ....................................... 140 Figure 40. 1H NMR of H/D exchange at the methylene position of 0.25 M betaine (0.3 M H3PO4, D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................................... 141 Figure 41. 1H NMR of I-I/D exchange at the methyl position of 0.25 M betaine (0.3 M H3PO4, D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................................... 141 Figure 42. Fit of the kinetic rate data for 0.25 M betaine (0.3 M H3PO4. D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ........................................................................ 143 Figure 43. .Hydrogenation of 0.25 M glycolic acid after 6 hours (THF/H2O, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................................................. 145 Figure 44. Hydrogenation of 0.25 M glycolic acid (0.3 M H3PO4, CH3CH2OH, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ........................................................................ 148 Figure 45. Hydrogenation of 0.25 M glycolic acid (CH3CH2OH, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................................................................................. 148 Figure 46. Hydrogenation of 0.25 M glycolic acid in the absence and presence of 0.3 M H3PO4 (CH3CH2OH, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ........................ 149 Figure 47. Hydrogenation of 0.25 M ethyl lactate (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............................................................................................................... 150 Figure 48. Hydrogenation of 0.25 M glycolic acid after 6 hours (CH3CH2OH/H2O, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................................. 152 Figure 49. Hydrogenation of C3 organic acids to alcohols (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ........................................................................ 159 Figure 50. Hydrogenation of mostly C2 organic acids to alcohols (0.3 M H3PO4, H20, 130 0C, 1000 psi H2, 1 g dry 5% Ru/C) .......................................................... 160 Figure 51. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D2O, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .......................................................... 167 Figure 52. Stacked lH NMR of H/D exchange at the methylene position on 0.25 M glycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................... 168 Figure 53. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................................................................................. 169 Figure 54. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................ 171 Figure 55. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............................................................................................... 171 Figure 56. lH NMR of H/D exchange at the methylene position of 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................ 173 Figure 57. Fit of the kinetic rate data for 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............................................................................................... 173 ix Figure 58. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) .......................................................... 178 Figure 59. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................................................................. 178 Figure 60. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ................................................ 180 Figure 61. Fit of the kinetic rate data for 0.25 M sarcosine (D2O, 100 0C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................................................... 180 Figure 62. 1H NMR of H/D exchange at the methylene position of 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ................................................ 182 Figure 63. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................................................... 182 Figure 64. Rate constants for H/D exchange at the methylene carbon (D2O, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ................................................................................. 183 Figure 65. Kinetic rate constant and primary and secondary isotope effects for glycine, sarcosine, and N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ........ 185 Figure 66. Percent H in D2O with either 1000 psi H2 or D2 .......................................... 186 Figure 67. Fit of the kinetic rate data for 0.25 M glycine (D2O, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 191 Figure 68. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 0C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 191 Figure 69. Kinetic rate constant and primary and secondary isotope effects for glycine with 1000 psi H2 or 1000 psi D2 or H2 and D2 ................................................ 192 Figure 70. Fit of the kinetic rate data for 0.25 M sarcosine (D2O, 100 0C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 194 Figure 71. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 194 Figure 72. Kinetic rate constant and primary and secondary isotope effects for sarcosine with 1000 psi H2 or 1000 psi D2 or H2 and D2 ................................................ 195 Figure 73. Fit of the kinetic rate data for 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 197 Figure 74. Fit of the kinetic rate data for 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 197 Figure 75. Kinetic rate constant and primary and secondary isotope effects for N,N-dmg with 1000 psi H2 or 1000 psi D2 or H2 and D2 ................................................ 198 Figure 76. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................ 199 Figure 77. Fit of the kinetic rate data for 0.25 M glycine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................................................................. 199 Figure 78. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................. 201 Figure 79. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................................................................. 201 Figure 80. ‘H NMR of H/D exchange at the methylene position of 0.25 N,N-dmg (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................ 202 Figure 81. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................................................................. 202 Figure 82. Rate constants for H/D exchange at the methylene carbon (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ................................................................................. 203 Figure 83. Kinetic rate constant and primary and secondary isotope effects for glycine, sarcosine, and N,N-dmg (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) .......... 205 Figure 84. Rate constants for H/D exchange at the methylene carbon at 100 °C or 70 °C (D2O, X °C, 1000 psi D2, 1 g dry 5% Ru/C) ................................................... 206 Figure 85. Fit of the kinetic rate data for 0.25 M glycine (D20, 50 0C, 1000 psi D2. 1 g dry 5% Ru/C) .................................................................................................. 207 Figure 86. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................................................................. 208 Figure 87. Kinetic rate constant for glycine, sarcosine, and N,N-dmg (D2O, X °C, 1000 psi D2, 1 g dry 5% Ru/C) ................................................................................. 209 Figure 88. Fit of the kinetic rate data for 0.25 M glycine (D20, 90 °C, 1000 psi D2, 1 g dry 5% Ru/C) .................................................................................................. 213 Figure 89. Fit of the kinetic rate data for 0.25 M betaine (D20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................................................................................. 214 Figure 90. 1H NMR of H/D exchange at the methylene positions of 0.25 glycylglycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C ................................................. 215 Figure 91. IH NMR of H/D exchange at the methylene position of hydrolyzed glycine from 0.25 M glyclglycine (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ...... 216 Figure 92. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) .......................................................... 218 Figure 93. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............................................................................................... 218 Figure 94. 1H NMR of H/D exchange at the methyl position of 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) .......................................................... 219 Figure 95. Fit of the kinetic rate data for 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............................................................................................... 219 xi Figure 96. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (D2O, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) .......................................................... 221 Figure 97. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................................................... 221 Figure 98. 1H NMR of H/D exchange at the methyl position of 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) .......................................................... 223 Figure 99. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................................................... 223 Figure 100. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 226 Figure 101. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) .............................................................. 226 Figure 102. Kinetic rate constant and primary and secondary isotope effects for sarcosine with 1000 psi H2 or 1000 psi D2 or H2 and D2 ................................................ 228 Figure 103. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) .......................................................... 228 Figure 104. Fit of the kinetic rate data for 0.25 M N,N-dmg (D2O, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) .......................................................... 229 Figure 105. Kinetic rate constant and primary and secondary isotope effects for N,N- drn with 1000 si H2 or 1000 si D2 or H2 and D2 ........................................ 229 8 P P Figure 106. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................................................... 230 Figure 107. Fit of the kinetic rate data for 0.25 M N,N-dimethylglycine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) ........................................................................ 231 Figure 108. Kinetic rate constants and primary and secondary isotope effects for sarcosine and N,N-dimethylglycine; 70 °C, 1000 psi D2 ................................ 232 Figure 109. Rate constants for H/D exchange at 100 °C or 70 °C (D20, X °C, 1000 psi D2, 1 g dry 5% Ru/C) ...................................................................................... 233 Figure 110. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) ............................................................................................... 234 Figure 111. Rate constants for H/D exchange of sarcosine (D20, 50, 70, or 100 °C, 1000 psi H2 or D2, 1 g dry 5% Ru/C) ....................................................................... 235 Figure 112. lH NMR of H/D exchange at the methyl position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) .......................................................... 237 Figure 113. 1H NMR of H/D exchange at the methylene position next to the —OH of 0.25 M ethanolamine (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................... 239 Figure 114. 1H NMR of H/D exchange at the methylene position next to the —NH2 of 0.25 M ethanolamine (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............. 239 xii Figure 115. Concentration data for 0.25 M ethanolamine (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) .................................................................................................. 240 Figure 116. 1H NMR of H/D exchange at the methylene position of 0.25 M glycolic acid (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................ 241 Figure 117. 1H NMR of H/D exchange at the methyl positions of 0.25 M tert—butanol (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................................................ 242 Figure 118. Hydrogenation of 0.25 M betaine hydrochloride and 0.25 M glycine (0.5 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ..................................... 243 Figure 119. Hydrogenation of 0.25 M glycine alone and 0.125 M glycine and 0.125 M betaine (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............ 245 Figure 120. Hydrogenation of 0.25 M betaine alone and 0.125 M glycine and 0.125 M betaine (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............ 246 Figure 121. Fit of the kinetic rate data for 0.25 M glycine alone and 0.25 M glycine and 0.25 M sarcosine (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ................... 247 Figure 122. Fit of the kinetic rate data for 0.25 M sarcosine alone and 0.25 M glycine and 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) ............. 247 Figure 123. Rate constants for H/D exchange of 0.25 M glycine alone, 0.25 M sarcosine alone, 0.25 M glycine in the competition experiment, and 0.25 M sarcosine in the competition experiment (D2O, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C)... 249 xiii LIST OF SCHEMES Scheme 1. Biomass based organic acids in the production of commodity chemicals ....... 3 Scheme 2. Dehydrogenation of l-methylcyclohexene onto Pt(111) surface ..................... 5 Scheme 3. Suggested molecular structure of methylcyclohexenyl species (C2Hlo) .......... 6 Scheme 4. Hydrogenation of glycolic acid ........................................................................ 9 Scheme 5. Hydrogenation of 2-ethylheptanoic acid ........................................................ 10 Scheme 6. Hydrogenation of 2-ethylheptan-5-olide ........................................................ 11 Scheme 7. Hydrogenation of ethyl lactate ....................................................................... 15 Scheme 8. Hydrogenation of n-valeraldehyde ................................................................. 16 Scheme 9. Reaction scheme for generation of D2 gas ..................................................... 19 Scheme 10. 10% Pd/C H/D exchange on 5-phenyl-l-valeric acid .................................. 21 Scheme 11. Hydrogenation of L-alanine .......................................................................... 22 Scheme 12. Generalized amino acid hydrogenation sequence ........................................ 22 Scheme 13. Ion exchange with water of 2-acetylthiamin pyrophosphate and 2-acetyl-3,4- dimethylthiazolium ............................................................................................ 23 Scheme 14. Calculated acidities for the zwitterions of glycine and its derivatives ......... 26 Scheme 15. H/D exchange at the a-carbon on glycine .................................................... 28 Scheme 16. Hydrolysis of N-protonated glycine methyl ester ........................................ 29 Scheme 17. Sodiated glycine in the charge-solvated (CS) and zwitterionic (ZW) forms 31 Scheme 18. H/D exchange on cyclododecene ................................................................. 32 Scheme 19. Proposed model for adsorption of ethylene glycol ....................................... 41 Scheme 20. Langmuir-Hinshelwood kinetic expression for competitive degradation of ethylene glycol or propylene glycol .................................................................. 42 Scheme 21. Langmuir-Hinshelwood kinetic expression for hydrogenolysis of glycerol 43 Scheme 22. Langmuir-Hinshelwood model of hydrogenation of lactic acid .................. 48 Scheme 23. Langmuir-Hinshelwood kinetic expression of hydrogenation of lactic acid 48 Scheme 24. Deuterium incorporation in D2/D2O using Ru/C .......................................... 50 Scheme 25. Langmuir-Hinshelwood model of hydrogenation of protonated alanine ..... 52 Scheme 26. Heats of formation (kcal/mol) for lactic acid hydrogenation ....................... 57 S cheme 27. Catalytic hydrogenation of lactic acid and propanoic acid .......................... 59 Scheme 28. Rate expressions for single acid hydrogenation ........................................... 61 Scheme 29. Structure of choline bicarbonate ................................................................ 106 xiv Scheme 30. Proposed mechanism for Ru/C catalyzed HD generation from H2 and D20 ......................................................................................................................... 187 Scheme 31. 1H NMR positions for the methylene hydrogens on glycyl glycine ........... 214 Scheme 32. 1H NMR positions for the methylene hydrogens on ethanolamine ............ 238 XV Equation 1. Equation 2. Equation 3. Equation 4. Equation 5. Equation 6. Equation 7. Equation 8. Equation 9. Equation 10. Equation 11. Equation 12. Equation 13. Equation 14. Equation 15. Equation 16. Equation 17. Equation 18. Equation 19. Equation 20. Equation 21 . Equation 22. Equation 23. Equation 24. Equation 25. Equation 26. Equation 27. Equation 28. Equation 29. Equation 30. Equation 31. LIST OF EQUATIONS NMR tube ratio: outer-to-insert volume ...................................................... 76 Lorentzian lineshape ..................................................................................... 77 Singlet peak fit .............................................................................................. 78 Triplet peak fit .............................................................................................. 78 Quintet peak fit ............................................................................................. 79 Total H concentration for each spectrum at time t ....................................... 80 Singlet area fraction ...................................................................................... 81 Triplet area fraction ...................................................................................... 81 Quintet area fraction ..................................................................................... 81 CH2 concentration ...................................................................................... 81 CHD concentration ..................................................................................... 81 CD2 concentration ...................................................................................... 81 Peak area of individual isotopomer ............................................................ 82 Conversion .................................................................................................. 82 Yield ........................................................................................................... 82 Selectivity ................................................................................................... 82 Peak area of water in sample ...................................................................... 83 Peak area of water in internal standard ....................................................... 83 Moles of deuterium in D20 (1) .................................................................... 84 Moles of hydrogen in H2 (g) ....................................................................... 84 Substrate — NHx contribution ...................................................................... 84 Substrate - methylene (CH2) and methyl contribution y*(CH3) ................. 85 Time zero H—D inventory ........................................................................... 85 Molarity of H at time zero .......................................................................... 85 Molarity of D at time zero .......................................................................... 85 Moles of D (l) at time zero ......................................................................... 86 Moles of H (1) at time zero ......................................................................... 86 Moles of H (g) at time zero ........................................................................ 87 Moles of D (l) at time t ............................................................................... 87 Total moles of H (l) and H (g) .................................................................... 87 %H at time t ................................................................................................ 87 xvi Equation 32. Molarity of H at time t ................................................................................ 87 Equation 33. Molarity of D at time t ................................................................................ 87 Equation 34. AH2 concentration at time zero .................................................................. 88 Equation 35. AH2 concentration at time t ........................................................................ 88 Equation 36. AHD concentration at time t ....................................................................... 89 Equation 37. AD2 concentration at time t ........................................................................ 89 Equation 38. CH3 concentration at time zero ................................................................... 89 Equation 39. CH3 concentration at time t ........................................................................ 89 Equation 40. CH2D concentration at time t ..................................................................... 89 Equation 41. CHD2 concentration at time t ..................................................................... 89 Equation 42. CD3 concentration at time t ........................................................................ 89 Equation 43. Aqueous hydrogen concentration in H20 and HOD .................................. 90 Equation 44. Aqueous deuterium concentration in D2090 Equation 45. Sum of the H area .......................................................... ............................ 90 Equation 46. Aqueous hydrogen concentration equation of fit ....................................... 90 Equation 47. Rate equations for H/D exchange at the methylene position ..................... 92 Equation 48. CH2 concentration at time t ........................................................................ 92 Equation 49. CHD concentration at time t ....................................................................... 92 Equation 50. CD2 concentration at time t ........................................................................ 92 Equation 51. Rate equations for H/D exchange at the methyl position ........................... 92 Equation 52. CH3 concentration at timet ........................................................................ 93 Equation 53. CH2D concentration .................................................................................... 93 Equation 54. CHD2 concentration at time t ..................................................................... 93 Equation 55. CD3 concentration at time t ........................................................................ 93 Equation 56. Forward and reverse rate constants for H/D exchange at the methylene position .............................................................................................................. 94 Equation 57. Forward and reverse rate constants for H/D exchange at the methyl position ........................................................................................................................... 95 Equation 58. Expression for H on the surface at time t ................................................... 96 Equation 59. Expression for D on the surface at time t ................................................... 96 Equation 60. Equilibrium expression for hydrogen and deuterium in the liquid and on the surface ............................................................................................................... 97 Equation 61. Rate of deuterium on the surface versus time ............................................ 97 xvii Equation 62. Conversion of substrate in the hydrogenation reaction ............................ 158 Equation 63. Yield of desired product in the hydrogenation reaction ........................... 158 Equation 64. Selectivity of desired product in the hydrogenation reaction ................... 158 xviii Chapter 1. INTRODUCTION 1.1 Background The Jackson/Miller group at Michigan State University is focused on emerging plant/crop-based renewables technology, where these renewables can serve to replace or complement conventional feedstocks to meet the ever-growing need for chemicals and other materials. The group is working in an interdisciplinary and collaborative effort using mechanistic organic chemistry and chemical reaction engineering. The studies described in this dissertation explore the aqueous phase catalytic hydrogenation reaction of organic acids and their derivatives to alcohols. In addition a detailed H/D exchange study was conducted on glycine and its N-methylated derivatives. Though several effective procedures have been developed, the mechanistic details of such reactions are far from understood. By examining the effects of structure and medium on organic acid reductions and H/D exchange, an improved understanding of some of the mechanisms involved in heterogeneous catalysis can be obtained. This knowledge will prove to be far reaching in that it helps map out the chemical transformations that will be needed for the renewables-based “refinery of the future”. 1.2 Significance Green chemistry, also known as sustainable chemistry, is an umbrella concept that has grown substantially since it fully emerged in 1990. By definition, green chemistry is the design, development, and implementation of chemical products and processes to reduce or eliminate the use and generation of substances hazardous to human health and the environment. Many of the concerns green chemistry is trying to address are not laboratory curiosities or individual research projects,1 but global issues such as climate change, energy consumption, pollution, and management of water resources. Green chemistries can be used to help reduce our dependence on petroleum and natural gas. The supply of petroleum is limited and finding an alternative to crude oil is increasingly important.2 Petroleum has been the most used energy source due to automobiles since the 1950’s due to its energy content, ease of transport, and abundance (albeit finite). It is used as a raw material for nearly all organic chemical products, including pharmaceuticals, solvents, fertilizers, etc. Identifying renewable feedstocks and green pathways to producing the same high-valued chemicals is required to ensure a sustainable future. Use of renewable feedstocks also has the potential to dramatically reduce U.S. dependence on foreign oil and allow for domestic growth of a bio-based chemical industry. Research at Michigan State University seeks non-fossil based alternative pathways for the production of a variety of commodity and specialty chemicals. The research presented in this dissertation is aimed at the development and characterization of novel pathways to produce high-valued chemicals using a “green chemistry” approach. Central to this objective is the need for insight into the reactions’ mechanisms as well as a broad descriptive knowledge of their variability as a function of input variables such as substituents, solvent, pH, temperature and reaction pressure. In this work, the main tools employed to probe the hydrogenation pathways of organic acids have been isotopic deuterium exchange and kinetic modeling. The ability to successfully and selectively hydrogenate biomass derived organic acids will open pathways to the use of renewable starting materials in polymers, foods, and pharmaceutical applications, all at costs competitive with existing petroleum based routes. In evaluating the possible renewable “green” feedstocks, it is evident that carbohydrate polymers (starch, cellulose) and monomers (glucose, fructose, mannose, etc.) and sugar alcohols will be major contributors. Currently, bioconversion (fermentation) of carbohydrates can selectively produce a large variety of products. The direct chemical conversion of these. downstream fermentation products can in turn result in a plethora of high-valued products. The combination of chemical- and bio-processing, to date, offers the most promising and flexible strategy for the development of renewable resource refining. Such a partnership is illustrated in Scheme 1 for the net transformation of glucose to propylene glycol, a commodity chemical with US production of ca. 5 x 108 kg/yr. CHO H——OH H0—— H fermentation A OH H2: catalyst _ OH H——OH ' Acoou ' /|VOH Hj—OH Lactic acid Propylene CH2OH glycol D-glucose Scheme 1. Biomass based organic acids in the production of commodity chemicals An additional advantage of this “environmentally friendly” approach is the domestically centered economic potential of converting simple sugars to high-valued and large volume commodities. In the US in 2006, 9178 thousand tonnes of oil equivalent ethanol were produced, up 24% from 2005 (7389).2 A tonne of oil equivalent is equivalent to the amount of energy given off by burning one tonne of crude oil (~ 42GJ). The US. Departments of Energy and Agriculture determined that 1.3 billion tons of US. biomass feedstocks are potentially available for the production of biofuels. Thus, there is no shortage of bio-based raw materials for chemicals production. The development of low- cost, high efficiency hydrogenation routes to exploit this opportunity is a major goal of the work described herein. In addition, a series of methylated amino acids was evaluated in terms of H/D exchange to better understand the mechanism of hydrogenation and possibly build on the opportunity for stereoretentive reduction to high-valued amino- substituted alcohols. 1.3 Literature Review 1 .3.1 Hydro genation/Dehvdro gen_ation Hydrogenation refers to a general chemical reaction that results in the addition of H2 to either double or triple bonds. Hydrogenolysis (to be discussed later, section 1.3.7) refers to a general chemical reaction that involves breaking bonds and the addition of hydrogen. In the work described here, the hydrogenation reactions that are being considered are mostly the reduction of X=Y double bonds. The exact ways in which the hydrogen atoms are delivered are varied depending on how polar the substrate is, etc. The largest and best studied class of hydrogenation includes hydrogenation of hydrocarbons. The first paper described below discusses the effect of hydrogen pressure in a reaction with 1- methylcyclohexene and following this is a brief discussion of the effect of hydrogen pressure on the hydrogenation rate of organic acids (our research). Hydrogen pressure is a critical parameter in catalytic hydrocarbon-refining processes, controlling reaction rates and selectivity. Traditionally the rate of dehydrogenation and fragmentation of hydrocarbons is slowed by the presence of hydrogen adsorbed on a catalyst surface. Surprisingly, Yang et al. found that in the presence of excess hydrogen at 0.3—0.7 psi (reported as 15-40 Torr) and at temperatures between 290-480 K, dehydrogenation of l-methylcyclohexene on Pt(1 1 1) to form toluene dominates over hydrogenation to form methylcyclohexane (Scheme 2). \ ...__.5 5 “d? ——> + 0 i—aiw—Cw Scheme 2. Dehydrogenation of l-methylcyclohexene onto Pt(1 l 1) surface The authors acknowledge that this dehydrogenation is somewhat counterintuitive and suggest that perhaps the excess hydrogen might assist in displacing adsorbed hydrocarbon fragments, which inhibit the chemisorption of the reactant. Hydrogenation to 1-methylcyclohexane is temperature dependent reaching a maximum at 370 K, whereas dehydrogenation to form toluene increases, starting around 350 K. In general, as temperature increases, n-bonded alkenes, which are weakly bonded to the surface, easily desorb, or they are converted into strongly di-c-bonded species. Taking a SFG (sum frequency generation) spectrum of l-methylcyclohexene at 403 K in 0.03 psi (reported as 1.5 Torr) of substrate and 0.3 psi (reported as 15 Torr) of hydrogen pressure and again at 300 K and below 9.7x10'll psi (reported as 5 x 10'9 Torr), Yang et al. found no change in the SFG spectrum (two bands at 2845 and 2925 cm'I). SFG spectroscopy is a second-order nonlinear optical process for analyzing surfaces and from the output, structural information regarding the molecules on the surface can be determined. The process involves an infrared laser beam combined with a visible laser beam to generate a sum frequency output. Other surface analytical techniques exist, such as thermal desorption spectroscopy (TDS), high-resolution electron energy loss spectroscopy (HREELS), and reflection absorption infrared spectroscopy (RAIRS). The results from Yang indicated that the species responsible for the spectrum was 0- bonded to the surface (Scheme 3). Scheme 3. Suggested molecular structure of methylcyclohexenyl species (C7Hlo) They suggest that excess strongly chemisorbed hydrogen molecules react with weakly 1t- bonded l-methylcyclohexene to form the hydrogenated methylcyclohexyl. C7H13 which is o-bonded to the surface. The methylcyclohexyl intermediate can undergo a disproportionation reaction with a weakly n-bonded l-methylcyclohexene to form methylcyclohexane, and a dehydrogenated species C7H1 1, which can continue losing hydrogen to form toluene, C7Hg.3 The effect of pressure has been studied in our research group on the hydrogenation of organic acids (X=Y). Jere et al. the Jackson/Miller group determined the kinetics of aqueous-phase hydrogenation of L-alanine to L-alaninol over 5 wt% Ru/C. Reaction temperatures were between 353—398 K with 0.22-0.46 M feed concentration, 0-1.2 M phosphoric acid concentrations and hydrogen pressure between 1.7-13.8 MPa (6.9 MPa = 1000 psi). Results indicated only a mild pressure dependence of the hydrogenation rate above 1000 psi H2 partial pressure. Jere et al. found that at pressures above 1000 psi H2, the rate was independent of the hydrogen concentration concluding that the surface was saturated with hydrogen.4 Literature describing hydrogenation of organic acids to high-valued chemicals exists; however, the fundamental chemistry involved at the catalyst surface has been little discussed, especially for processes in aqueous phase. For producing chemicals that replace traditional petroleum-derived species, hydrogenation or other reductions must play a key role to reduce the overall oxidation state of the oxygenated feed material. The ability to successfully and selectively hydrogenate organic acids will open pathways to the use of biomass-derived materials in polymers, foods, and pharmaceuticals applications. The development of low-cost, high-efficiency hydrogenation routes is the goal of our research group. Research conducted on hydrogenation of organic acids to alcohols will be discussed first, followed by a discussion on the research conducted on hydrogenation of esters and aldehydes. Somewhat recently (1998), Antons et al. described the preparation of optically active alcohols from optically active carboxylic acids using a ruthenium catalyst. For the hydrogenation of lactic acid the enantiomeric excess of propylene glycol dropped with increasing temperatures (from 383 K to 413 K).5’6 Recent patents describing heterogeneous catalysis for the hydrogenation of carboxylic acids to alcohols provide yields and selectivities, but do not go into the mechanistic aspects of the chemistry involved. Carnahan and co-workers examined hydrogenation of carboxylic acids using a ruthenium catalyst. They used both ruthenium dioxide and ruthenium-on-carbon at temperatures of about 423 K (reported as 150 oC); platinum and palladium under similar conditions were not successful. They obtained an 80% yield of ethylene glycol from 20 wt% glycolic acid (hydroxyacetic acid) after a reaction time of 1 hour using 10 g of 10% Ru on carbon and extremely high pressures. The hydrogenation was performed at a temperature of 418-423 K (reported as 145-150 0C) and 9,500-10,400 psi (reported as 650-710 atm) in water (Scheme 4). 10% Ru/C; 0 H2O (“\‘m 420 K; (\0“ 0” 9550-10,400 psi 0” t = 1hr yield = 80% Scheme 4. Hydrogenation of glycolic acid The yield was reduced to 68% ethylene glycol when using 3 g of 10% Ru on carbon at the same reaction conditions. In an experiment with glycolic acid and 10% Ru on carbon at 523-528 K (reported as 250-255 °C) and 7,900-10,300 psi (reported as 540-700 atm), Carnahan et al. analyzed the gas in the head space using mass spectrometry and reported the following distribution: 78.3 mole % hydrogen; 1.8 mole % methane; 0.8 mole % ethane; 0.04 mole % carbon dioxide; 0.8 mole % oxygen; 0.09 mole % argon; 18.2 mole % nitrogen (they mention that the last three constituents may have resulted from contamination of the sample with air and with nitrogen used to purge the sample lines). The yield of ethylene glycol from this reaction was 32% with side products methane and ethane, lower than the 80% obtained at the conditions in scheme 4. The lower yield was a result of reducing the catalyst amount from 10 to 3 g and increasing the reaction temperature from 420 K to 523 K. The yield of 32% was obtained after 15 minutes compared to the yield of 80% after a reaction time of 1 hour.7 Behr et al. reported homogenous bi-metallic catalyzed reduction of carboxylic acids to alcohols with hydrogen. One drawback of traditional heterogeneous hydro genations using metal catalysts is the use of high pressure, namely ~2900-4300 psi and expensive high pressure reactors. Thus, homogeneous catalysis can often be advantageous in that it can be performed at more moderate reaction temperatures and hydrogen pressures. Both homogeneous and heterogeneous catalysis were investigated in the reduction of 2- ethylheptanoic acid to 2-ethylheptanol. Catalysts included [Mo(CO)6], [Rh(acac)(CO)2], [Ru(acac)3], Rh/C, Rh/Al2O3, and Ru/C (Scheme 5). 0 1 mol% catalyst; dioxane OH 9' OH 473 K; V\/\l:\ ~2200 psi t = 2 hrs Scheme 5. Hydrogenation of 2-ethylheptanoic acid Yields of 10% (for the heterogeneous case of 1 mol% Ru/C) or less were obtained for all catalysts, however, nearly quantitative yields were possible with homogeneous bimetallic catalysts. An alcohol yield of 99% was found with Mo(CO)6, a group 6 early transition- metal carbonyl, and Rh(acac)(CO)2, a group 8 late transition-metal complex, at 473 K (reported as 200 °C), 2200 psi (reported as 150 bar) H2 pressure in dioxane for 4 hours. They suggest that the acetylacetate anion, C5H7O' (acac) oxygen-donor ligands in general show the highest activity for the reduction of carboxylic acids. Behr and co-workers also looked at the reduction of 2-ethylheptanoic acid with homogeneous/heterogeneous bimetallic catalysts. A yield of 69% alcohol was achieved using the homogeneously dissolved Mo(CO)6 and heterogeneous Ru/C at 473 K (reported as 200 °C), 2200 psi (reported as 150 bar) H2 pressure in dioxane for 2 hours. Although the yields are not as high as in the case of the homogeneous bi-metallic catalysis, there seems to be a collaborative effect. In the case of Mo(CO)6 or Ru/C alone 10 yields of only 2 and 10%, respectively were reported after 2 hours. It may be that the main catalytic reaction is occurring on the Ru surface because in the case of reduction of 2-ethylheptan-5-olide to 2-ethyl-1,5-heptanediol (Scheme 6) with Mo(CO)6 and [Rh(acac)(CO)2] reducing the amount of rhodium (reduction of ~66%) leads to a greater loss of activity (from 92% to 19% yield), however, less molybdenum (reduction of ~66%) has less of an effect on the reaction (from 92% to 53% yield) (Scheme 6). Mo(CO)6/ Rh(acac)(CO)2; dioxane > OH 0 o 473 K; 2200 psr OH t = 2 hrs _ _ yield = up to 92% Scheme 6. Hydrogenation of 2-ethylheptan-5-ollde Behr also expanded the work to evaluate reduction of commercially available lactones with the rhodium/molybdenum catalyst. They found that the reduction of lactones was more difficult if the compound was branched in the a-position to the carboxylic acid and suggested perhaps it is because the compound is more sterically hindered.8 A disadvantage of this chemistry is that homogeneous bimetallic catalysts can be difficult to prepare and often cannot be reused because they are not separated from the substrate, etc. Mao and co-workers found that a magnesia-supported poly-y-aminopropylsiloxane-Ru complex (MgO-NH2-Ru) was extremely effective in the aqueous catalytic hydrogenation of several carboxylic acids (acetic, propanoic, lactic, and isobutyric acid). Optimal reaction conditions included 513 K (reported as 240 °C), 725 psi (reported as 5.0 MPa) 11 H2 in 5 mL of water for 18 h. Upon examination of the hydrogenation of 0.1 mL (1.35 mmol) propanoic acid, results revealed that the amount of Ru in the MgO-NH2-Ru complex, reaction temperature, and the amount of water significantly affected the yield of desired alcohol. An increasing Ru content up to 0.12 mmol/g (from 0.048 mmol/g) increased the yield of propanol formed. As the temperature increased from 433 K to 523 K (reported as 160 °C to 250 °C), the yield of propanol went from 0% to 95%. Increasing the water amount from 1 to 5 mL increased the yield of propanol to 100% (from 5%), however, more than 5 mL (up to 8 mL) reduced the yield of propanol (to 28%). In summary, the optimal conditions for 0.1 mL of propanoic acid in water is 513 K, 725 psi H2 pressure, 5 mL of water, and a Ru content of 0.12 mmol/g in the MgO- NH2-Ru complex.9 F uchikami and co-workers discussed transition-metal catalyzed hydrogenation of pentadecanoic acid using 1.0 mL of dimethoxyethane as a'solvent. Optimum reaction conditions included 1.0 mmol substrate and 0.01 eq. 5% [Rh/A1203] and [Mo(CO)6] at 423 K (reported as 150 °C), 1470 psi (reported as 100 atm) H2 pressure for 16 h. The reaction in the presence of only Mo(CO)6 resulted in no conversion to the desired alcohol indicating that bimetallic catalysts consisting of Group 8 to 10 late transition-metal compounds and Group 6 or 7 early transition-metal carbonyls were extremely effective in the hydrogenation of organic acids. '0 Summarizing the hydrogenation of organic acid reactions, Antons et al. found that with a ruthenium catalyst the enantiomeric excess of propylene glycol from hydrogenation of 12 optically active lactic dropped with increasing temperature (similar to results observed by our research group, to be discussed later). Carnahan obtained yields of 80% ethylene glycol from glycolic acid after 1 hour using Ru/C but at high H2 pressures (~10000 psi). Research in this paper presents yields up to 50% for ethylene glycol from glycolic acid using Ru/C at similar conditions except for lower H2 pressures (1000 psi) and 1/3 the catalyst : substrate amount. Behr reported low yields (less than 10%) for hydrogenation of 2-ethylheptanoic acid to the desired alcohol using 1 mol% Ru/C at conditions similar to what our research used, but again Behr used higher pressures (~3000-43 00 psi) compared with our research at 1000 psi H2. Behr expanded his work to lactones and found they were more difficult to hydrogenate than organic acids. Expanding the use of ruthenium to a magnesia-supported poly-y- aminopropylsiloxane-Ru complex (MgO-NH2-Ru), Mao et al. found this catalyst very effective in hydrogenating several carboxylic acids in water. Reaction temperatures were about 100 °C higher than what we used in our research (to be discussed later). Fuchikami et al. found that bimetallic catalysts (similar to the results from Behr) exhibited extremely powerful reducing ability for the hydrogenation of organic acids focusing mostly on [Rh/A1203] and [Mo(CO)6]. It may be suggested to expand the use of heterogeneous Ru/C hydrogenations to include bimetallic catalysts. Hydrogenation of esters and aldehydes has also been studied. Adkins et al. looked at optically active esters, the effect of , free acids on the hydrogenation of esters, and also (1- and B-oxygen substituted esters. Vaidya et al. looked at hydrogenation of n- 13 valeraldehyde and fit a power law model to the experimental data in order to generate a rate expression for the reaction. We will first consider the hydrogenation of esters then follow this discussion with Vaidya’s work with aldehydes. Adkins and Bowden described the hydrogenation of optically active esters to alcohols using copper-chromium oxide at a temperature of 523 K (reported as 250 oC) and pressures of 2200-2900 psi (reported as 150-200 atm). A yield of 81% 1,2-propanediol from butyl lactate (91% conversion) was obtained after 2 hours at these conditions.ll Earlier (1932) studies by Adkins and coworkers also showed that hydrogenation of esters over copper-chromium oxide at 523 K (reported as 250 °C), 2900-4400 psi (reported as 200-300 atm) H2 in ethanol is inhibited in the presence of free acids, presumably due to competitive adsorption resulting in less desired product formation. 12,13 In 1947, Adkins and Pavlic studied alpha amino ester hydrogenation over Raney nickel in a non-aqueous solvent, namely dry ethanol, using lower temperatures (25-75 °C) and high pressures (150-200 atm or 2200-2900 psi). The reaction was carried out in a 96 5 M stainless steel reactor with about 20 to 35 mL of solution. The hydrogenated products were isolated by fractional distillation. Yields of desired alcohol were as high as 98% for hydrogenation of a-ethylpiperdinoacetate. ‘4 Adkins and Billica reported effective hydrogenations of esters to alcohols at modest temperatures (298-423 K reported as 25-150 °C), yet with the requirement of higher 14 catalyst loadings. They suggest using similar mass amounts or greater of catalyst compared with the mass used of ester. In most cases, they used 10 g of the ester substrate (roughly 0.085 moles for the case of ethyl lactate) and 15 g of catalyst. When using an excess of Raney nickel the hydrogenation at 373 K (reported as 100 °C) and 5000 psi of a-amino esters was quantitative and reasonably rapid (e. g. ethyl piperidinoacetate was complete in 1.7 hrs). One limitation with Raney nickel is that hydrogenation of phenyl substituted esters unavoidably results in benzene ring reduction to form the cyclohexyl substituted amino alcohol. Raney nickel was also used for the hydrogenation of or and B-oxygen substituted esters. A quantitative yield of 1,2-propanediol from ethyl lactate in dry ethanol at 423 K (150 °C) and 5000 psi was obtained after 1.5 hrs. The research presented in this thesis includes a-oxygen substituted organic acids and found an enhanced hydrogenation rate with these substrates compared with the unsubstituted substrates (to be discussed later). 0 Raney Ni; V/k Et0H 2 DHO + D2 (55 mmol) (~ 83 mmol) Scheme 9. Reaction scheme for generation of D2 gas After 48 hours, 87 mL of D2 gas was generated from 1.0 mL of D2O (TON = 829). In the absence of 10% Pd/C or H2 gas (under Ar) results indicated a negligible increase in the HOD signal over time. Sajiki et a1. expanded this work and examined the amount of deuterium exchange into organic substrates in a closed system, in the presence of a limited amount of H2 gas. After 24 h of stirring, deuterium was incorporated via reduction of 0.5 mmol cinnamic acid (C6H5CHCHCOOH) into both methylene positions of dihydrocinnamic (3- phenylpropanoic) acid. The reaction consisted of 7.4 mg of 10% Pd/C in 5.0 mL or 3.0 mL (277 mmol or 166 mmol) of D2O in a 100 mL hydrogen-charged sealed flask at room temperature. The actual internal volume of the hydrogen-charged sealed flask (100 mL) was 160 mL. They also found that the efficiency of incorporation was greatly reduced in reactions which used a larger flask. The mechanism of D2 gas generation is unclear. However, Sajiki et al. propose oxidative insertion of Pd(0) activated by H2 or HD as a ligand into D20 to give Pd(II) species followed by the key step, a ligand exchange between H2-D2O, in the catalytic isotope exchange reaction. Subsequent reductive elimination gives the corresponding Pd(0)-HD or D2 complex. At the end of the reaction D2 gas is liberated.'7 19 Performing the reaction under an atmosphere of D2 and in dry EtOAc as an inert solvent as opposed to H2 and D20, resulted in significantly less deuterium exchange, even at a temperature of 180 °C. Thus, Sajiki, et a1. concluded that D20 is required for efficient isotopic exchange. Deuterium-labeled organic compounds can be prepared using two main strategies: (a) a multistep synthetic method starting from deuterium-labeled synthons and (b) selective catalytic replacement of carbon-bound hydrogen with deuterium. Both homogeneous and heterogeneous catalytic conditions have been employed, with a wide variety of - - I8 19 20 21 22 23 catalytically active metals (e.g. Ir , Rh , Co , Pt , Ru , and Mn ). However, because they are easily separated from reaction mixtures by filtration, heterogeneous catalysts are most convenient, where they work for these applications. The work described (later) in this thesis discusses aqueous-phase heterogeneous catalytic HfD exchange. Sajiki et al. using heterogeneous catalytic H/D exchange evaluated the effect of temperature and also described the H/D exchange observed at carbons alpha to an acid, phenyl ring, primary alcohol or ether. The work discussed in this thesis centers around the H/D exchange at the alpha carbon of glycine and its derivatives. A maximum amount of deuterium incorporation into 5-phenylvaleric acid occurred at a temperature of 160 °C. Deuterium incorporation decreased with increasing distance from the benzene ring. Over the time course of the reaction, rapid exchange was observed at the benzylic carbon (C1) as well as at the beta carbon (C2) and also at position C3 with respect to the phenyl ring, however, the carbon alpha to the acid (C4) and the aromatic carbons show only a gradual 20 increase of deuterium exchange (Scheme 10). At a temperature of 110 °C, there was no measurable incorporation at the carbon alpha (C4) to the acid. 0 D D o 10% Pd/C, H2 A P, 002” COH 7 Ph/W 2 D2O,24h,heat o D o D Scheme 10. 10% Pd/C H/D exchange on 5-phenyl-l-valeric acid Sajiki et al. found that the hydrogens on the alpha carbon adjacent to a primary alcohol or ether as a substrate exchanged with deuterium to a lesser extent as compared with the methylenes closer to the phenyl ring. Reactions with substrates which lacked an aromatic ring, such as octanoic acid, led to almost no deuterium incorporation (at reaction temperatures ranging from 110-160 °C). In addition, substrates with an oxygen between the phenyl ring and the alkyl chain, such as 5-phenoxyvaleric acid underwent almost no deuterium incorporation. Thus there is apparently a need for an aromatic ring to be directly connected to the alkyl chain for H/D exchange using 10% Pd/C. The exact mechanism of the exchange is not clear, but the authors proposed that the reaction goes via Pd/C-catalyzed C-H bond activation combined with deuteration from D2O. The results presented provide a deuterium gas-free, totally catalytic and postsynthetic deuterium labeling method in D2O.24 Our research group, specifically Jere et al. report aqueous stereoretentive hydrogenation of L-alanine to L-alaninol using a 5% Ru/C catalyst at 100 0C and 1000 psi in the presence of added acid (Scheme 11). Alanine had become a compound of interest because an enhanced rate of hydrogenation of a-substituted organic acids such as alanine 21 and lactic acid (to be discussed in section 1.4 summary of previous work) had been noted. A slight excess, 0.29 M H3PO4 (pKa = 2.15) was added to protonate 0.22 M alanine (pKa = 2.34) and afford a 91% yield of alaninol after 6 hrs. 5% Ru/C; D20; equimolar H3PO4 NH OH 373K; f0” 2 1000 psi D2 2 t=6hrs yield = 91% Scheme 1 l. Hydrogenation of L-alanine To track the reaction’s stereochemistry, D2O was used as the solvent at 1000 psi D2 and 100 °C. Surprisingly, although rapid deuterium incorporation at the a-carbon was observed for alanine and alaninol, little or no racemization was seen. OH H 100°C OH OH cam-I3 @NH3 Scheme 12. Generalized amino acid hydrogenation sequence 22 Scheme 12 shows a generalized amino acid hydrogenation sequence which includes an aldehyde and two potential enols as intermediates. Jere et al. suggest that the enhanced rate of hydrogenation with lactic acid and protonated alanine is due to the electron withdrawing a-OH and a-N+H3, enhancing the carbonyl group’s preference for sp3 hybridization.” Looking again at enolization, Frey et al. using 2-acetylthiamin pyrophosphate (AcTPP, 1) and 2-acety1-3,4-dimethylthiazolium (2) performed an H/D exchange kinetic analysis of sequential reactions. Frey et al. found that an electrophilic thiazolium ring enhances the ionization rate and that the acetyl methyl hydrogens of AcTPP, l and 2 ion exchange with water at a pD of ~4.0 (Scheme 13). "'3C OH O OH CH3 8 s \ \ = N\ R2 \ g\R1 R2 \ <9 R1 CH3 CH3 H2N /NIYCH3 \ N 3- 12 R1 = HZC R2 = CH2CH20P206 22 R1 = CH3 R2 = H Scheme 13. Ion exchange with water of 2-acetylthiamin pyrophosphate and 2-acetyl-3,4- dimethylthiazolium 23 The rate of ionization for AcTPP is on the order of 107 faster than acetone. The following kinetics were employed to describe the exchange of the acetyl methyl hydrogens (R-CH3): R—CH3 —-) R—CDH2 —> R—CDzH Defining a = [R-CH3] and b = [R-CDH2], at time t = 0, a = a0 = [R-CH3]0, and b = b0 = 0. Describing consecutive (pseudo) first-order reactions the differential equations and solutions are given by: 3$=kl*a=kl*ao*eXP(Tk1*t) a:aO’I‘CXP(—kl*t) 512:,“ *ao *exp(—ki *t)-k2 *b b = 1" *ao *lexP(—ki *1)“CXP(“ k2 *0] (11 (k2 _k) The rate constant k is defined for exchange of a single proton, thus the values of k1 and k2 are 3k and 2k, respectively. The substitution of these variables into the above equations can be expressed as: a ==a0 *exp(—3*k*t) and b =3*a0 *[exp(—2*k*t)—exp(—3*k*t)] From these equations, the following ratios can be expressed. _fl_ [3a+b) 3 Using lH NMR, normalized integral ratios were measured as a function of time to the iR—CH3] [[3*(R-CH3)+R—CDH2]:| 3 = exp(—- k * t), thus = exp(—k * t) extent of 70-80% exchange of one proton. Frey et al. found that the rates of exchange with water were the same (within error) for the hydrate and ketone forms and thus the interconversion between them is fast compared with the exchange reactions. The rate of exchange was not observed at pD S 2, thus the reaction is not acid catalyzed but base 24 catalyzed and presumably occurs via an enolization mechanism. Using the iodine scavenging method, rate constants for the uncatalyzed and base catalyzed enolization were determined (koE = (3.8 x 0.5) x 10'6 s" and (6.73 r 0.22) x 10'5 s" at pD 4.17 (46 mM Na oxalate)). Frey et al. suggest that the exchange reaction occurs via the rate-limiting loss of a proton to produce an enolate, which then can be rapidly deuterated in the presence of D20. The rate of enolate formation from a carbonyl compound is related to the carbon acidity. In the case of AcTPP, Frey et al. believe the largest contributor to the fast enolization is due to through-space electrostatic effects of the positively charged thiazolium ring. Fast rates of enolization have been reported for other cationic ketones, namely PhCOCH2N+C5H5 (kOHE = 1.8 x 105 M'ls'l). The rate enhancement in the ionization of carbon acids and the enolization of carbonyl compounds is poorly understood, but Frey et al. suggest that rates of base-catalyzed enolization are accelerated by the through—space electrostatic effect of a nearby positive charge.26 Proteins have stereochemical integrity due to the weakly acidic nature of the a-protons on the repeating amino-acid monomers. Richard et a1. calculated aqueous carbon acidities for the zwitterions of glycine and its derivatives: H3N+CH2COO', pKa = 28.9 :1: 0.5; H3WCH2CO2Me, pKa = 21.0 :1: 1.0; Me3N+CH2CO2Me, pKa = 18.0 i 1.0; Me3N+CH2CO2', pKa = 27.3 i 1.2 (Scheme 14). 25 0CD O G 0 CH3 0 (+3 pKa=28.9 H3N\/“\ pKa=27.3 H3C\rL H3C/@ 0 CH3 0 HC pK,=21.o Hf} pKa=18.0 3\| N OCH3 H3C/® OCH3 Scheme 14. Calculated acidities for the zwitterions of glycine and its derivatives The ionization state of the amino acid relates to the acidity of the or-proton. A decrease in pH results in protonation of glycine and an increased reactivity towards enolization, whereas enolization of the anionic form of glycine is far less favorable. The pH is mathematically related to pKa by the following general equations. The acid dissociation constant, Ka, is a quantitative measure which defines the equilibrium for a chemical reaction. It can be written in terms of equilibrium concentrations, Ka = [H+] [A-]/ [HA]. Thus, [H+] = Ka[HA]/[A-] and due to the large orders of magnitude spanned by Ka values, the log is of Ka is commonly used. The logarithmic measure of the acid dissociation constant is written as, log[H+] = log Ka + log[HA] - log[A-]. Since pH = - log[H+] the equation can be rewritten in the form of the Henderson-hasselbalch equation, pH = pKa + log[A-]/ [HA]. In glycine and its derivatives, Richard et al.28’29 found that modifying the a-substituent from N+H3 to NH2 increases the pKa values of the a-proton by 5 pKa units, and changing the a-substituent from COO' to COOCH3 decreases the pKa of the oc-proton by 8 pKa units. Malthouse and co-workers used the above values and estimated the pKa values and 26 tryptophan-synthase-catalyzed exchange rates of a-protons of alanine, serine, aminobutyrate, norvaline, norleucine, and tryptophan. They assumed a linear relationship and estimated that ApKa= -oer*, where 0t=l.6-2.2 and 0* represents the appropriate Taft substituent constant. From the calculations, the pK,1 values of alanine, aminobutyrate, norvaline, norleucine, and tryptophan are all raised by about 1 pKa unit, namely ~35, relative to glycine and serine, ~34.27 Richard et al. performed kinetic studies in D2O at 25 °C using a substrate concentration of 20 or 40 mM glycine with a constant ionic strength of 1.0, maintained with potassium chloride. In the 1H NMR spectra, they monitored the replacement of the or-CH2 group’s singlet resonance with the triplet due to a-CHD sites. This mono-deuterated product signal exhibited an upfield shift of 0.014 ppm for glycine and its derivatives or 0.016 ppm for betaine and its derivatives (J HD = 2.5 Hz). To quantify the progress of deuterium ' exchange, values of R, reaction progress, were calculated: R = (Acriz/(ACH2+ACHD))- Note: There is a mistake in this journal in that there is a two missing in front of the ACHD term to represent the fact that for CH2 there are two hydrogens available for exchange and for CHD there is only one hydrogen. ACH2 and ACHD are the integrated areas of the singlet due to the oc-CH2 of the substrate and triplet due to the a-CHD of the monodeuterated product. They did not account for the oc-CD2 group, because it was their assumption that no dideuterated glycine was formed. For the early reaction time samples, the ‘3 C satellite of the signal due to the a-CH2 group was not resolved from the most downfield peak of the 27 triplet due to the or-CHD group. Due to this, the total integrated area of the triplet due to the or-CHD group, ACHD was calculated by multiplying the integrated area of the most upfield peak of the triplet by three. The area of the singlet due to the a-CH2 group, ACH2, was calculated by subtracting the calculated value of ACI—ID from the total integrated area of all signals due to the or-protons. Semi logarithmic plots (In R = -k0b5dt) of R vs time were linear with negative slopes equal to kobsd (5"), which is the rate constant for exchange of a single proton of the or-CH2 group. The fraction of deuterated glycine, fD, was calculated: fD = 2ACHD/(ACH2 + 2ACHD). Their assumption was that there was essentially no formation of dideuterated glycine, oc-CD2 during the hydrolysis of glycine methyl ester in D2O. They suggested a mechanism whereby the zwitterionic glycine enolizes to form the anionic intermediate which they believe is not stabilized via protonation of the substrate carbonyl oxygen (pKa < 7.0). This intermediate quickly forms the carboxylate compound and undergoes deuteration at the a-carbon (Scheme 15). O 0 D20; 298K 3 1=1.o KCl 9 M < >, OAK/K I O 3 3 o / o e e G) H D Scheme 15. H/D exchange at the or-carbon on glycine Calculating rate constants, Richard et al. found that simple amino acid zwitterions (glycine and betaine) are weaker carbon acids than the simple oxygen ester ethyl acetate (pKa = 25.6). In addition, there is about a 2-3 unit increase in acidifying effect of an or- N+Me3 as compared to an or-N+H3 substituent on the carbon acidity, corresponding to the difference in polar effects of these substituents. There is more of a polar effect with the 28 N+Me3 due to the formation of hydrogen bonds between the (x-N+H3 and the solvent, which moves the positive charge away from the nitrogen and onto the solvent, increasing the effective separation between the interacting cationic and anionic centers at the enolate. There is greater stabilization of negative charge by electrostatic intramolecular interactions with the (l-N+M€3, wherein the methyl groups reduce the effective dielectric constant of the local medium.28 Richard et al. found that hydrolysis of N-protonated glycine methyl ester is 10-times faster than deuterium exchange of the first a-hydrogen at a neutral pD (pD = 7.4). The glycine obtained after the completion of the hydrolysis reaction (>10 half-times) is enriched with 0.10 atom of deuterium (Scheme 16). N-protonated glycine methyl ester has a second-order rate constant of km = 6.0 M'ls'l for DO' catalyzed. exchange of the first a-proton, whereas the rate constant for H/D exchange of acetone is kDO = 0.3 M'ls'l (20-fold difference). O H @ OCH3 OCH3 D3N H In D20; \ 25°C H G H G o 0 e O D3N D3N H Scheme 16. Hydrolysis of N-protonated glycine methyl ester 29 The fast H/D exchange of N-protonated glycine methyl ester is presumably due to polar stabilization by interactions between the or-NH3+ group and the developing enol’s partially negative carbon next door. The oc-protons of N-protonated glycine methyl ester (pKa “~" 21) show roughly the same acidity as the corresponding protons for the thiol ester and mandelic acid.29 In general, the extent of H/D exchange for a given species increases with the basicity of the exchange reagent. Cox et al., using an external ion source 7-T FTICR (Fourier transform ion cyclotron resonance) mass spectrometer, found that charge-solvated sodiated glycine oligomers (Gllea+ see Scheme 18 to GlysNa+) are of the lowest energy (gas phase). Both experimental and computational methods were employed. An F TICR is one of the highest performance mass spectrometers and measures the mass-to-charge ratio of ions based on the cyclotron frequency of the ions in a fixed magnetic field with excellent resolution and mass accuracy. Using a low level of theory calculations, 3-21G. and comparing and confirming with higher theory calculations (Jensen used 6-3 10* see JACS 1992, 114, 9533, Moison and Armentrout used MP2/6-31G* see J. Phys. Chem. A. 2002, 106, 10350, and Wyttenbach et al. used B3LYP/6-311++G** see JACS 1998, 120, 5098) in the literature, the most stable structure was the one in which the glycine was not in the zwitterionic form but rather when the sodium cation was bound to both the nitrogen and the carbonyl oxygen (Scheme 17). 30 8,2;- Ho CS C2H5NO2Na+ ZW Czl‘lsl\102l\lf:l+ Scheme 17. Sodiated glycine in the charge-solvated (CS) and zwitterionic (ZW) forms Experimentally, sodiated peptide ions were generated by matrix-assisted laser desorption/ionization (MALDI) and samples were deposited onto a stainless steel probe tip. The MALDI solutions consisted of a mix (623:2) of 1 M 2,5-dihydroxybenzoic acid in ethanol, 0.03 M peptide in water/acetonitrile (3:7 v/v), and 1 M D-fructose in water. A static pressure of the H/D exchange gas, ND3 (7 x 10'8 Torr) was maintained. In the gas phase for sodiated pentaglycine they experimentally found the rate constants for deuterium exchange with ND3 were 3.0 x 10'”), 2.5 x 10'”), and 1.5 x 10"0 cm3 molecule‘l s'1 for the three observed hydrogen exchanges (nitrogen and oxygen hydrogens). The exchange rate ratios were about 3:2: 1, thus three equivalent hydrogens are presumed to be exchanging with ND3. They also suggest that the mechanism requires a carboxylic acid hydrogen due to the fact that no exchange occurs with methyl esters of glycine oligomers.30 Matsubara et al. examined deuterium exchange of cyclododecene in hydrothermal or subcritical water (523 K reported as 250 °C, 5 MPa 3 725 psi), in the absence of hydrogen. At these conditions the pKw of water is lower (ca. 11) and thus the water ionizes more to form OH' and H3O+ than water at ambient conditions. They propose that 31 It in these conditions the Pd(0) might oxidatively insert into O-H bonds of water to form the Pd(II) species, H-Pd-OH. The key species is the palladium hydride that exchanges with the solvent water’s hydrogens. Using cyclododecene (5.0 mmol), 10 wt% Pd on active carbon (100 mg, 0.1 mmol Pd) and deuterium oxide (20 g) in a 30 mL Teflon vessel under hydrothermal conditions for 12 h, cyclododecene was converted into the fully deuterated product (E/Z = 72:28) (Scheme 18). 10 wt% Pd/C (2 mol%) L D2O, 250 °C, 12h I E/Z=33:67 E/Z= 72:28 Scheme 18. H/D exchange on cyclododecene The product was extracted with hexane and analyzed by GC, 1H NMR, 2H NMR, and mass spectra. Matsubara et al. explain the deuteration of alkene containing compounds via an interaction of the alkene and the palladium hydride. However, they also report full deuteration of saturated hydrocarbons, and in this case suggest that direct C-H activation may be the probable route to forming the fully deuterated alkanes. T err-butyl benzene was subjected to the above conditions and no exchange was observed at the sterically hindered ortho positions, however full deuteration occurred at the meta- and para- positions. Nearly 45% of the C-H bonds on the tertiary butyl group were exchanged with deuterium after 10 h (10 wt% Pd/C (2 mol%), D2O , 250 °C). Using the same conditions, biphenyl was fully deuterated. Mechanistic studies have not yet been examined. It is 32 worth noting that hydrogen gas was needed for the 10% Pd/C-catalyzed deuterium exchange reaction below 180 oC.31 1.3.3 Alkynylation of sp3 C-H Bonds Adjacent to a Nitrogen Atom A challenging, yet attractive organic synthesis reaction is C-C bond formation via C-H bond activation. The possibility of enantioselective catalytic C-C bond formation via this method could enhance asymmetric synthesis both theoretically and practically: In order to construct a chiral carbon center, a prochiral sp2 carbon center is generally necessary as the precursor; however, Li et al. discuss the synthesis of enantioselective C-C bond formation based on activation of a sp3 C-H bond. Three papers by Li et al. will be discussed, starting with alkynylation of sp3 C-H bonds adjacent to a nitrogen atom followed by enantioselective alkynylation of prochiral sp3 C- H bonds adjacent to a nitrogen atom, and finally cross-dehydrogenative coupling between sp3 C-H bonds and sp3 C-H bonds. In the first paper, Li et al. report a copper-catalyzed (5 mol%) oxidative cross-coupling of amines with various alkynes via a direct sp3 C-H bond alkynylation in the presence of tert-butyl hydroperoxide (2:1 :1 ratio, respectively) at 100 °C for 3h with yields as high as 74%. The reaction did not take place in the absence of the tert-butyl hydroperoxide. There was less desired product formed in reactions with aliphatic amines, as opposed to aryl amines. Interesting to note is that N,N-dimethylbenzylamine produced alkynylation of the methyl group rather than the benzyl position. 33 Li et al. propose a tentative mechanism for the direct oxidative coupling suggesting the formation of an imine-type intermediate (coordinated to copper) through activation of sp3 C-H adjacent to nitrogen. Copper also activated the terminal alkyne and coupling of the two intermediates resulted in the desired product and regenerated the copper catalyst.32 Li et al. expanded on this work to include tetrahydroisoquinoline alkaloids which exist widely in nature and are of biological and pharmacological interest. Specifically, they discuss the development of a novel catalytic asymmetric l-alkynylation of tetrahydroisoquinoline derivatives via activation of the sp3 C-H bonds of prochiral CH2, yielding optically active C-l-substituted tetrahydroisoquinoline derivatives. Li et al. found that the optimal conditions were with CuOTf (as opposed to CuBr seen in their previous work), at 50 °C (as opposed to 80 °C) in THF with a chiral ligand, yielding 63% of the desired product. Asymmetric alkynylation also proceeds in water or without solvent but the yields are lower. Aromatic substituted alkynes provided good yields and enantiomeric excess in most cases, whereas aliphatic substituted alkynes resulted in fair or low enantiomeric excesses. The factors which influence the enantioselectivities of the reaction are still being investigated.33 Vicinal diamines are found in biologically active natural products and are used in medicinal chemistry as well as in chiral ligands for asymmetric catalysis. Li et al. previously discussed both alkynylation of sp3 C-H bonds adjacent to a nitrogen atom and enantioselective alkynylation of prochiral sp3 C-H bonds adjacent to a nitrogen atom. 34 They expand on this work and discuss what they consider to be the first simple and efficient cross-dehydrogenative-coupling reaction between two sp3 C-Hs. Due to the importance of vicinal diamines they explore an alternative approach to these compounds. Typically a nitro-Mannich (aza-Henry) reaction to afford the new C-C bond B-nitroamine derivative might be employed. Following the formation of the B-nitroamine derivative, reduction of the nitro group to give vicinal diamines can be obtained. Using tetrahydroisoquinoline and nitromethane, Li et al. report a 90% yield of the C-C coupled product at room temperature with 2 mol% of CuBr catalyst with 1.0-1.2 equivalents of tert-butyl hydroperoxide, 0.1 mmol tetrahydroisoquinoline, and 1.0 mL of nitromethane. The exact mechanism of this coupling is unknown. However, Li et al. propose three possible intermediates: 1) an imine-type intermediate formed via coordination to copper through H-abstraction of the sp3 C-H adjacent to nitrogen. 2) In addition to the previous intermediate, the copper catalyst activates the nitroalkanes to form an oxygen coordinated copper species which couples with the previous intermediate to form the product, while regenerating the copper catalyst. Lastly, 3) an intermediate, in which tart-butyl hydroperoxide is involved, which further is converted into the cross-coupling products catalyzed by CuBr.34 The mechanism of this reaction is still under investigation. 1.3.4 Secondary Deuterium Isotope Effects for Enolization Reactions Murray et al. report a— and B-deuterium isotope effects for the enolization reaction and equilibria for acetone or acetophenone in deuterium oxide using IH NMR spectroscopy, triton exchange kinetics, and ab initio calculations. For the C-H,D bond adjacent to the carbonyl in the ionization of protonated acetone, an inverse secondary isotope effect of 35 0.97/D was calculated. This isotope effect is consistent with the decrease in hyperconjugation upon ionization. Upon further loss of hyperconjugation in the out-of- plane CH in the enolization reaction, a secondary a-deuterium kinetic isotope effect of 1.06 i 0.02 was calculated. Murray et al. described the kinetics according to the following first order equations for a-CH3 singlet (P), triplet a-CH2D (Q), and quintet or- CHD2 (R): P = P0 =I‘exp(—7tl *t) 9‘1 =1: :1: _ :1: Mi— :1: _ =1: Q-[m] P0 exp( )‘1 t)+|:(Kl-X2)+Q0:l exp( )‘2 t) R=[( AlarkzakPo )]*exp(—kl*t)+[[( klakkzakPo ]+ kzrth ]*exp(—?t2*t) rpm-(2.4. t,_r,)r(2.3_22) ((3%) Apfltzirpo 12*Q0 *ex — * +ii[(x1_)‘3)*(}‘2 ‘x3)i+ (A, *l3):i+Roi p( l3 t) P0, Q0, and R0 correspond to the initial peak area, and M, A2, and A3 are the pseudo-first- order rate constants for deuteroxide ion catalyzed exchange of protons. Murray et al. display the 1H NMR spectra for acetophenone illustrating the incomplete resolution of the multiplets for the a-CH2D and a-CHD2 signals (J: 2.2 Hz for both). The data were fit to first-order consecutive process equations with rate constants 3kHHH [OD-] = 4.20 x 10“1 s", ZkHHD[OD-] = 2.52 x 10'4 s", and kHDD[OD-] = 1.25 x 10“ s". A secondary isotope effect of kHHH/kHHD = 1.11 and kHHD/kHDD = 1.01 was calculated. (kLij where L is the primary hydron, hydron meaning hydrogen cation, that is cleaved and i and j refer to the remaining secondary hydron sites). The isotope effects determined by 1H NMR for proton exchange of CHH and CHD are not significantly different from one another. Murray et al. examined several methods of accounting for the lack of complete resolution 36 of the multiplets including adding 0.1 Hz of line broadening, Gaussian weighting of the F IDs, and using peak heights instead of integrals.35 1.3.5 Catalytic Hydrogenation of a-Amino and a-Hydroxy Esters Carboxylic acids and esters are among the least reactive functional groups in catalytic hydrogenation reactions. They usually require higher temperatures (>200 °C) and higher pressures, and thus can result in racemization of stereogenic centers. However, Broadbent et al. demonstrated with rheniurn heptoxide reduced in situ as catalyst, propylene glycol yields as high as 84% from ethyl lactate at 150 °C, 250 atm, and neat.36 Studer et al. report using 10% Rh/Pt oxide (N ishimura catalyst 45.9% Rh, 19.9% Pt) for the mild reduction of various oc—amino and a-hydroxy esters in methanol at 100 bar (1450 psi) hydrogen pressure and 25 °C while maintaining the integrity of the a-stereogenic center. Antons et al. report using Ru oxide to hydrogenate alanine at 100 °C and 200 bar with moderate to good yields and some racemization (93-98.5% ee).37 In looking at the available catalyst systems for the hydrogenation of acids and esters, many of them are not commercially available, catalyst loading can be high, and racemization can be a problem at higher temperatures. Studer et al. found that the Nishimura catalyst was the only catalyst effective (compared with Ru and Pt oxide which showed weak activity; and Rh, Pd, Ir, and Re oxide as well as 2% Pt, 4% Rh/C were completely inactive) in the reduction of or-substituted esters at room temperature, and methanol and ethanol as solvents yielded the highest amount of desired product. Monomethylation of the amino group resulted in ~17-25% reduction in yield of 37 hydrogenated product. In addition, the alanine derivatives, N(n-Pr)2 or an NH-BOC group did not react at all. Studer et al. also found that Nishimura catalyst can selectively reduce an activated ester in the presence of an unactivated ester, as is the case with the dimethyl ester of glutamic acid. The non-activated ester forms a lactam, while the or-functionalized carboxylic acid derivative is hydrogenated to the corresponding alcohol. Studer et al. also found that or- methoxy and B-hydroxy esters did not react over the Nishimura Catalyst at these conditions. In preparative-scale experiments with alanine methyl ester and methyl lactate (2 g substrate and 6 g substrate; 10:30 catalyst to substrate ratio) yields of greater than 80 and 90%, respectively were obtained and a distinct induction period of 4 minutes was observed after hydrogen saturation. The ease of ester reduction decreased in the order of a-NH2 > oc-OH, NH-alkyl > B-NH2 using these mild reaction conditions and the oxidic, bimetallic Nishimura catalyst. The addition of acid lowered the hydrogenation rate and deactivation of the catalyst was observed at higher temperatures. Studer et al. propose that pre-hydrogenation of the catalyst in the absence of substrate reduces the catalytic activity. They suggest a working hypothesis of a multi-functional active center, both basic and multi-metallic sites in which the substrate is adsorbed, then transformed in several steps, and the desired product is desorbed. Studer et al. think the a-NH2, a-NHR, or or-OH functionalities serve to interact with the basic (oxidic) sites, thus anchoring the activated ester near the metallic centers. This interaction would not be possible with the OR or NR2 groups and less favored with B-functionalities. 38 A drawback of this research is that the Nishimura catalysts are very expensive and these reactions require high catalyst loadings.38 In terms of direct relevance to green chemistry goals, the use of esters as substrates and alcohols as solvents imply additional steps and disposal requirements for any practical implementation of these processes. 1.3.6 Catalytic Properties of Ru/C Ruiz and co-workers explored the interactions between Ru nanoparticles and carbon support.39 Carbon materials for support in heterogeneous catalytic reactions such as hydrogenations (and hydroprocessing of petroleum fractions) have several advantages. They are chemically inert, stable (in the absence of oxygen), have high surface area and optimum porosity. In addition to a large surface area, the carbon-support surface must also be chemically accessible so that active sites (unsaturated valences) must be present. Ruiz et al. discuss that the narrow microporosity of activated carbons can hardly participate in the adsorption process of the metallic precursors from solution, so there is no simple relationship between the surface area of the activated carbon and the final metal dispersion of the carbon-supported catalysts. They prepared Ru catalysts with three different carbon material supports to evaluate how the physical and chemical properties of the support and the Ru precursor on the surface affect the Ru nanoparticles generated after reduction using hydrogen. The three carbon materials were 1) powdered carbon black (Seer = 135 m2/ g); 2) mesoporous high surface area graphite (SBET = 297 mz/g); and 3) microporous carbon molecular sieves (SBET=1500 mz/g). All were prepared using the incipient wetness impregnation technique, using 39 metal precursor solutions with the proper precursor quantity in order to obtain a 2wt% Ru in the final catalyst. The reduction of Ru is from Ru(III) to Ru0 via a continuous flow of 20 cm3/min of a H2/He gas mixture (10% H2) with 200-300 mg of prepared sample. The reduction started at room temperature and continued to 800 K at 2 K /min. In the presence of chlorine, a higher reduction (470 K as compared to 400 K and 440 K for RuS and RuNH, respectively) temperature was used due to the chlorine chemisorbing on the metal particles decreasing their reducibility. Furthermore, Ruiz and co-workers found that residual chloride species reduce absorption capacity of the sample by poisoning the surface of the Ru particles. 1.3.7 Hydrogenolysis of Glycerol to Propylene Glycol About 1 kg of crude glycerol by-product is produced for every 9 kg of biodiesel produced. Nearly 1 billion pounds of propylene glycol is generated annually in the United States40 and retails for about $0.71 per pound.41 The commercial route of preparing propylene glycol is the hydration of propylene oxide derived from propylene by either the chlorohydrin process or the hydroperoxide process. Dasari et al. discuss the selective hydrogenolysis of concentrated glycerol to propylene glycol at lower temperatures and pressures using various catalysts. All catalysts were pre-reduced at 300 °C for four hours under a constant stream of hydrogen.42 Shanks et al. discuss the kinetics in the hydrogenolysis of glycerol to form either propylene glycol or ethylene glycol. Looking first at the degradation rate of the products, they found the rate was zero order (independent of initial concentration) for the 40 concentration levels of 2.5 — 10 wt% (0.3-1.5 M) in water. The catalyst is saturated with the glycols even at very low concentrations. The calculated degradation rates for ethylene glycol and propylene glycol were 40 and 50 mol/(kg of catalyst)-s, respectively. Therefore, the degradation of products needs to be included in the model for glycerol hydrogenolysis. Shanks et al. also examined the competitive adsorption of ethylene glycol and propylene glycol. Results indicated that even though propylene glycol degraded slightly faster than ethylene glycol, it was less competitive for active sites. Previous work suggested that the polyols adsorb to the ruthenium through oxygen, thus the non-oxygenated end of propylene glycol may be responsible for a lower binding energy compared with ethylene glycol.43 The following model (Scheme 19) was used for ethylene glycol (EG): EG+S EG-S (fast) EG-S = EG'OS (slow) EG'°S = EG’ + S (fast) EG’ ——* X (pH dependent) Scheme 19. Proposed model for adsorption of ethylene glycol A similar model was proposed for propylene glycol, where S is the catalytic site, EG' is the aldehyde formed via dehydrogenation in the first step, and X is the degradation products. From Gibbs free energy data, the equilibrium concentration of glycol is strongly favored over the formation of the aldehyde (second step) under these conditions. 41 Assuming Langmuir-Hinshelwood kinetics, and assuming competitive adsorption between the glycols, the following rate equation (Scheme 20) was initially derived: k'iG *(iG) kEG *EG+kpG *PG+1) "r'r‘G=( Scheme 20. Langmuir-Hinshelwood kinetic expression for competitive degradation of ethylene glycol or propylene glycol where i = E for ethylene glycol and P for propylene glycol, k',-G is the degradation rate constant, iG is the respective concentrations of the glycols, and km and kPG are the adsoption constants. Experiments were conducted at two pH levels, 11.7 and 8.0. The glycols had about twice the degradation at the higher pH, even though, the adsorption constants did not vary significantly. Another observation was that as the pH increased, the amount of product degradation increased, but presumably due to ethylene glycol degradation. At the higher pH, a 5-fold increase in the glycerol reaction correlated to a 5 times higher production in propylene glycol, but less than two times higher production in ethylene glycol. Including glycerol in the competitive adsorption, it was found that ethylene glycol had an inhibitory effect, whereas propylene glycol did not. Fitting experimental data to the proposed reaction rate model, a modified equation was obtained which included a reaction order of 1.5 for the glycerol. The mechanistic reason for this 1.5 order is not clear, but may be explained by a reaction intermediate. The modified equation (Scheme 21) is shown as: 42 _rr. _ k10 *(IG)—SiG *k’G *(Glj) 10 (k0 *G'l'kEG *EG+kPG*PG+1) Scheme 21. Langmuir-Hinshelwood kinetic expression for hydrogenolysis of glycerol They performed the reactions at 500 rpm, although data was also obtained at 1000 rpm. No difference in reaction rates was found between the two mixing speeds, confirming that the system is not limited by external diffusion. 1.3.8 Deuterium NMR There are several differences between lH and 2H including spin, natural abundance, frequency, shape, quadrupolar relaxation, isotope shift, and conventional coupling. Unlike the familiar spin 1/2 proton (1H), deuterium (2H) with its extra neutron has a nuclear spin of 1 and a natural abundance of 0.01 5% (proton 99.985%), and an NMR frequency (at 7.05T) of 46.05 (as compared to the proton NMR frequency of 300). When I=1, as in 2H, there are three stable nucleus orientations in the magnetic field. The orientations are parallel, orthogonal, and antiparallel. The mechanism is distinguished from dipole-dipole relaxation in two ways. First, it does not require a second nucleus in motion; the quadrupolar nucleus experiences its own fluctuating field due to its motion in the unsymmetrical electron cloud. Second, because the mechanism is extremely effective when the quadrupolar moment of the nucleus is large, T; can become very short (milliseconds or less, in contrast to multisecond T; values typically seen for 1H). When the lifetime of the spin state, as measured by the relaxation time, is very short, the larger 43 uncertainty in energies implies a larger range of resonant frequencies, or a broader signal in the NMR spectrum. Deuterium has a very weak quadrupole moment, with the result that neighboring protons exhibit normal couplings to 2H. Thus nitromethane with one deuterium (CH2DNO2) shows a 1:1:1 triplet because the protons are influenced by the three spin states (+1, 0, - l) of deuterium. Nitromethane with two deuteriums (CHD2NO2) shows a 1:2:3:2:1 quintet from coupling to the various combinations of the three spin states (++; +0, 0+; +-, 00, -+; -0, 0-; --).44 Lambert et al. summarized the general isotopic shift observed for hydrogen exchanged with deuterium. As the number of bonds increases between the perturbing and resonating nuclei, the magnitude of the isotOpic shift decreases. Typical shifts for hydrogen perturbed by deuterium are on the order of 0.04 ppm, 0.015, and 0.01 for 1, 2, and 3 bonds between the participating nuclei.45 Isotopic proton shifts have been observed in several pairs of compounds. Gutowsky reports an upfield proton shift for CH2D in Ph- CH2D of 0.015 t 0.002x10'6 with respect to CH3 in Ph-CH3. Gutowsky suggests that differences in vibrational amplitudes of D and H atoms are the primary causes of such isotopic shifts. The difference between the electronic wave functions of the isotopic compounds shift the protons in CHD upfield relative to CH2. Since C-D has a smaller vibrational amplitude, it is closer to the nucleus and imparts an enhanced shielding effect.46 44 A coupling interaction between a spin 1/2 and spin >'/2, as in the case of a deuterium, may lead to broadening of the individual components of both multiplets assuming a relatively slow rate of relaxation. If the rate of relaxationbecomes fast enough it may lead to a broadening that results in a merging of broad signals. Pople discusses the theory behind broadening of a component with spin 1/2 coupled to a nucleus with a spin of I = 1. His results revealed that for slow quadrupole relaxation there is individual and unequal line broadening and as the lines continue to broaden, the outer peaks move inward until they merge with the center to become a single signal.47 1.4 Summary of Previous Work Over the past several years, the J ackson/Miller group has probed the Ru/C-catalyzed aqueous-phase reductions of several organic acids, focusing on lactic acid hydrogenation to propylene glycol and amino acid hydrogenation to the corresponding amino alcohols. Related aqueous hydrogenolysis reactions of glycerol and higher sugar alcohols to high- valued polyols have also been studied, providing valuable additional insights. In 2001 , Zhang et al. examined several factors including catalyst, effect of pressure and temperature, feed concentration, and catalyst loading in the aqueous-phase hydrogenation of lactic acid to propylene glycol.48 Reactions were carried out in a 300 ml stirred batch reactor (Parr Instruments) with a typical reaction time of 5 hours. Zhang found that the RWC catalyst achieved nearly 100% conversion of 0.55 M lactic acid at 423 K (reported as 150 °C), 2100 psi (reported as 14.5 MPa) H2, and 1.0 g of catalyst within 5 hours. In comparison at the same conditions, Ru/Al2O3 achieved ~60 % conversion, Ru/TiO2 45 achieved ~38 %, Raney Nickel achieved ~20 %, and CuCrOx, Pd/C, and Ni/Al2O3 achieved conversions all less than ~5%. Zhang also characterized the product mixture from hydrogenation of lactic acid using Ru/C < 170 °C the major byproducts at < 170 °C are light hydrocarbons (methane, ethane, and propane). For reactions at temperatures > 170 °C the byproducts are 1-propanol and ethanol, with trace amounts of 2-propanol. An optimum reaction temperature of 150 0C was determined; at higher temperatures (170 °C) cracking and excessive byproduct formations become an issue. Due to low solubility of hydrogen in water, higher pressures are required to increase the hydrogen concentration in the liquid. Pressures between 500 psi and 2,000 psi (reported as 3.5 MPa and 13.8 MPa) were evaluated. Higher pressures resulted in an increased yield of propylene glycol. Zhang calculated the absolute reaction rate (moles lactic acid/h/g catalyst) for three different feed concentrations (0.55, 1.15, and 3.6 M) and found that the reaction rate increases at higher lactic acid concentrations, but leads to a lower conversion. These results suggest that the reaction may not be first-order in lactic acid and a Langmuir-Hinshelwood rate expression might better describe the reaction kinetics. Zhang compared three different catalyst loadings and found that with higher catalyst loadings, conversion rates increased, yet selectivity to propylene glycol slightly decreases due to the increased exposure of propylene glycol to the catalyst. Zhang also looked at the addition of potassium and calcium salts on the effect of lactic acid hydrogenation. Typically calcium lactate is the usual product of fermentation and Zhang found that it can 46 be easily converted to propylene glycol by adding a stoichiometric quantity of sulfuric acid followed by hydrogenation of the free acid. Zhang et al’s achieved 90-95 % lactic acid conversion to propylene glycol at optimal conditions (423 K and ~2100 psi). Zhang et al. continued with this work and determined the kinetics of aqueous-phase hydrogenation of lactic acid to propylene glycol (423 K, 2,000 psi, 1.5 g of catalyst, and 1.15 M lactic acid in water).49 Mass transfer was considered in a three-phase reaction. For the hydrogenation reaction to occur several physical and chemical steps need to occur: (1) transport of hydrogen from the gas phase to the liquid phase; (2) mass transfer of both hydrogen and lactic acid from the liquid phase to the catalyst surface; (3) diffusion of both hydrogen and lactic acid within the porous catalyst to the metal surface sites; and (4) hydrogenation of the acid to the desired alcohol via a sequence of surface chemical reaction steps. To calculate the reaction rate of lactic acid hydrogenation to propylene glycol, a fourth order polynomial was fit to lactic acid conversion vs time data. The equation was then differentiated and evaluated at various times during the reaction. Using stirring rates between 400 and 1200 rpm, Zhang et al. evaluated the influence of gas-liquid mass transfer on the rate. They found a weak dependence of reaction rate on hydrogen concentration. And although at lower stirring speeds (400 rpm) there is a reduced concentration, as the reaction proceeds, the reaction rate declines and mass transfer has less effect and the result is no observable influence of stirring speed on the reaction rate over the course of the conversion. 47 Zhang et al. used Langmuir-Hinshelwood kinetics and described the hydrogenation of lactic acid by the following steps (Scheme 22), where lactic acid (LA) and hydrogen (H2) adsorb on the catalyst surface (S) and react to form propylene glycol (PG). The exact mechanism of the surface reactions are unknown. (1) LA + S = LA-S (fast) (2) H2 + S = H2°S (fast) (3) H2°S + LA-S —> P1°S + S (slow) (4) PI'S + H2°S = PG-S (fast) (5) PCS = PG + S (fast) Scheme 22. Langmuir-Hinshelwood model of hydrogenation of lactic acid Zhang et al. found that the above model which suggests molecular adsorption of both lactic acid and hydrogen (as opposed to dissociative adsorption), the rate-limiting surface reaction step between lactic acid and hydrogen, and desorption of propylene glycol gave all positive kinetic and adsorption constants with the correct temperature dependence. The rate-limiting step was found to be irreversible. The effect of propylene glycol on the reaction kinetics was evaluated by adding propylene glycol in the amount of 2.5 times that of lactic acid feed concentration and the rate was unaffected. So, propylene glycol was not used in the final Langmuir-Hinshelwood model. The final form of the kinetic expression is given in Scheme 23. kCLAPHZ (1+KH2PH2 +KLAC1..4 )~ -R G“(kmol/kg of cat/s) = Scheme 23. Langmuir-Hinshelwood kinetic expression of hydrogenation of lactic acid 48 Also, when Zhang et al. used only the initial rate data, the Langmuir-Hinshelwood model gave similar regression coefficients thus the catalyst deactivation was not a major factor at these experimental conditions. This work was later re-visited by Chen (2007, discussed later). Chen found that using a multivariable regression on the initial rate data resulted in error from an unrealized interdependence of the kinetic parameters. But, when using the initial rate data from Zhang et al, an activation energy of 46 kJ mol was calculated which agreed with later results. From the model, heats of adsorption of lactic acid and hydrogen were estimated based on the adsorption constants (KLA and KH2) at the two temperatures. Lactic acid had a AHLA = 47 kJ/mol and hydrogen had a AHH2 = 79 kJ/mol. They also looked at the relative surface concentrations of hydrogen and lactic acid on the catalyst. They found the ratio of empty, hydrogen-occupied (KH2PH2), and lactic acid-occupied (KLACLA) sites is 1 : 0.05 : 0.6 at ~2000 psi hydrogen pressure, 403 K and 1.13 M lactic acid. This supports the finding that increased hydrogen pressure increases the reaction rate. There is significant lactic acid adsorbed but relatively little hydrogen. In summary, Zhang developed a useful kinetic model which could be used a tool for further investigation into the mechanism of hydrogenation of lactic acid to propylene glycol. Jere et a1. expanded on this work by including hydrogenation of the amino acid L-alanine ((5)-2—aminopr0panoic acid) to L-alaninol (S-(+)-2-amino-1-propanol).25 This simple chiral amino acid was chosen to further probe the mechanism and stereochemistry of the hydrogenation by using isotopic labeling and reaction rate data. It was found that the addition of phosphoric acid (slight excess) is required to pronate alanine (pKa = 2.34) 49 enabling the carboxylic acid functionality to be undissociated. Jere et al. found that with sufficient acid (0.29 M), hydrogenation of protonated L-alanine (0.22 M) to protonated L-alaninol at 398 K (reported as 125 °C) and 1800 psi D2 in D2O occurred with more than 95% selectivity and 99% enantiomeric excess (ee). At 448K (reported as 150 °C) and 1000 psi hydrogen pressure, the product L-alaninol was found to racemize to D- alaninol, however, at no point during the hydrogenation reaction was D-alanine detected. At lower reaction temperatures, less racemization is observed with essentially no product racemization detected at 100 °C. Also at 100 °C they found incorporation of deuterium at C2 without loss of optical purity for both alanine and alaninol (greater than 90% D incorporation at six hours reaction time). At higher temperatures (150 oC), H/D exchange is observed at the C3 positions with racemization of alaninol (65% D incorporation; 81% ee alaninol) (Scheme 24). O O 100 °C 150 °C —_'> H3C DH2C 5 ‘ OD Hs NH3 Us N03 100 °C , 100 °C D D D o D 0 DH C H3C 100 C H30 150 C 2 ‘. OD .. OD s‘ s‘ s‘ s‘ H D ND3 ND3 Scheme 24. Deuterium incorporation in D2/D2O using Ru/C Jere et al. also found that the catalyst binding sites were saturated with hydrogen above 1000 psi and that at pressures at or above this neither hydrogen nor substrate 50 concentrations limit the reaction rate. From their results, they postulate a general amino acid hydrogenation sequence (Scheme 12). The scheme does not explicitly represent the role of the catalyst. It is expected that hydrogenation of amino acid is fast relative to dehydration of the 1,1- diol and reduction of the aldehyde. The aldehyde has not been detected but it is postulated that the hydrogenation reaction does go through this intermediate as opposed to enol which is catalyst-independent and not seen in the conditions investigated (Scheme 12). The H/D exchange reaction presumably occurs separately from hydrogenation. One reason for this is that in the H/D exchange reaction there is little or no loss of chirality at C2, despite exchange. So, a free enol, spz hybridized at C2 cannot play a role in this reaction. Also, in the absence of additional acid, hydrogenation does not occur but H/D exchange at C2 occurs rapidly. These results suggest that deuterium is incorporated via the catalyst directly removing hydrogen from the arnine-bearing carbon yielding a surface-bound intermediate that retains the original C2 configuration. Jere et al. continued with this work and developed a Langmuir-Hinshelwood kinetic model for predicting alanine conversion over Ru/C catalyst.50 Reaction conditions included alanine concentrations between 0.22 to 0.46 M, phosphoric acid concentrations between 0 and 1.2 M, 80 °C to 125 °C (reported as 353 K to 398 K), and hydrogen pressures from 250 psi to 2,000 psi (reported as 1.7 MPa to 13.8 MPa). Jere et al. made some new assumptions than what Zhang had previous proposed. Jere modeled hydrogen as dissociatively adsorbing on the Ru metal surface and also 51 suggested that the hydrogen and the acids (protonated alanine and phosphoric acid) absorb on different sites. The hypothesis that protonated alanine and phosphoric acid compete for the same sites is based on experimental data. The reaction rate of hydrogenation of protonated alanine starts to decrease upon excess phosphoric acid addition, which can not be explained unless competitive adsorption is at work. The assumption that hydrogen dissociatively adsorbs comes from recent work done by both Vannice and Neurock.51’52 Jere et al. proposed the following series of elementary reactions to model the hydrogenation of alanine (Scheme 25). (1) A+ + S, = A+-S1 (fast) (2) P + S1 = P-S. (fast) (3) H2 + 2S2 = 2H-S2 (fast) (4) 214-82 + A+-Sl —> A1+°S1 + 282 (slow) (5) 2H-S2 + Af-Sl = Alol+-Si + 2S2 (fast) (6) Alol+°S1 = Alol+ + S1 (fast) Scheme 25. Langmuir-Hinshelwood model of hydrogenation of protonated alanine Earlier research in the J ackson/Miller group looked at aqueous-phase catalytic hydrogenolysis of glycerol to propylene glycol.53‘54’55‘56’25’57‘58 Glycerol is a by-product of biodiesel formation, however, there is more of an industrial demand for propylene glycol than glycerol. Propylene glycol is used as a solvent, coolant, moisturizer in medicines and cosmetics, and chemical intermediate. Peereboom et al. sought to gain a deeper understanding of glycerol hydrogenolysis by studying the activated carbon support used in the catalytic reaction.59 Two carbon supports were used, a 0.8 mm extrudate activated 52 carbon referred to as ROX and a power activated carbon designated as 3310. The properties of the carbon supports are presented in Table 1. Table 1. Carbon characterization by N2 adsorption at 78 K Carbon type 3310 ROX BET surface area (mz/g) 716 834 Micropore area (mz/g) 375 586 Total pore volume (cm3/g) 0.654 0.536 Micropore volume (cm3/g) 0.173 0.272 Typically there are local concentrations of the substrate and product (glycerol and propylene glycol) in the activated carbon micropores. These concentrations are presumably different than those in the bulk solution phase. In an effort to gain a bigger picture regarding the reaction kinetics it is necessary to try and understand the difference in concentration between solution and pore. Typically only the concentration in the bulk solution is being measured and used for the chemical kinetics. Peereboom et al. performed carbon adsorption experiments using propylene glycol and glycerol (0.01 M to 2.0 M each) at room temperature (298 K reported as 25 °C) and modeled the results using both F reundlich and Langmuir isotherms. Both models are relatively simple and describe non-competitive adsorption. The F reundlich is based on a least-squares linear regression of experimental data from ln(CAS) vs ln(CA), where C AS is the concentration of A adsorbed onto the activated carbon and C A is the observed A concentration in solution in equilibrium with carbon. The data is fit to the equation C A3 = KanA to give SIOpe, n, and intercept ln(Kp). The Langmuir is based on a plot of (C A/C A5) vs C A and fit to the 53 equation CA3 = KACACTA/(HKACA). This gives a slope of l/CTA and intercept, (l/KACTA), where CM is the maximum concentration of A in activated carbon absorbent. In general for ROX, the Langmuir isotherm gave the best fit of the data and the indicated a similar value for the maximum concentration of absorbed glycerol and propylene glycol (CTA = 1.77 and 1.64 mol/kg, respectively). However, at concentrations below the Langmuir maximum, the quantity of propylene glycol adsorbed is greater than that of glycerol. This is deduced from the equilibrium constants for each, KA is 0.0163 M'l for propylene glycol and 0.00475 M'1 for glycerol (3.5 times more for propylene glycol). The enhanced adsorption for propylene glycol is due to the less hydrophilic nature of propylene glycol favoring it presence in the carbon micropores rather than the bulk solution. There is also a lower extent of solvation of propylene glycol via hydrogen bonding. Peereboom et al. also studied elevated temperature adsorption (313 to 433 K) with substrate concentrations of 0.1 to 0.5 M using the assumption that all adsorption sites have equal binding energies. The Van’t Hoff equation is described as ln(KA) = -(AH/R)*(1/T)+ln(Ko) and the heat of adsorption (AH) is the slope calculated from a plot of ln(KA) vs l/T. The heats of adsorption for glycerol and propylene glycol were calculated to be about 15 kJ/mol each (3.6 kcal/mol); suggesting similar van der Waal’s type forces and weak hydrogen bonding. 54 They also studied two-component adsorption and used the extended Langmuir model to describe the adsorption of the two species into the activated carbon micropores. Aqueous solutions of propylene glycol and glycerol with a total species concentration of 0.05 to 0.5 M were evaluated. From the data, the results show a clear preference for propylene glycol adsorption, by about a factor of about 3.5 times. Using a total species concentration of 0.5 M, the heat of adsorption was calculated and verified to be constant at 353 K, 393 K, and 433 K (reported as 80 °C, 120 oC, and 160 °C). Using the measured micropore volume at a temperature of 353 K for different bulk glycerol and propylene glycol concentrations it was found that the glycerol solution concentration is not very different than the pore concentration, however, the pore concentration of propylene glycol is about five times larger than the corresponding glycerol solution concentration. The presence of propylene glycol lowers the glycerol pore concentration significantly. In conclusion, they found that the pore concentrations for both propylene glycol and glycerol deviate from the bulk solution concentrations enough that they should be accounted for in the kinetic modeling of catalytic glycerol hydrogenolysis. Dalavoy et al. recognized the importance of the commodity chemical propylene glycol and decided to explore electrocatalytic hydrogenation (ECH) as a mild alternative to the traditional chemical catalytic method (involving higher temperatures 423 K, and pressures 1200 psi H2).60 In this method, the acid is reduced by hydrogen formed from electrolytic acid decomposition at the surface of a catalytic metal electrode. The hydrogen is produced in situ on the catalyst surface and this presents some challenges. Due to the low hydrogen gas solubility, there is a need for high pressures in order to 55 achieve adequate concentrations of the gas. Also mass transport and the splitting of hydrogen atoms on the catalyst surface can be a challenge. A series of ECH experiments using lactic acid and a RVC electrode suffused with 5% Ru/C powder catalyst (same catalyst as used in the chemical catalytic hydrogenation studies) were conducted. Using a two-compartment cell, the anode and cathode compartment were separated by a glass frit preventing the diffusion of gaseous products. The electrode was an Ag/AgCl and the counter electrode was Pt wire. Depending on the electrolyte and the temperature, the potential of the electrolyte was -100 to -400 mV (vs Ag/AgCl). The final concentration of lactic acid in solution was 11.1 mmol (0.1 wt%) and the total volume was 75 mL. The final electrolyte concentration was about 0.01 M and argon gas was bubbled through the solution to remove oxygen and other dissolved gases. Water was circulated through a glass ATR cell (~20 mL in volume) to help maintain a constant temperature. At a temperature of 363 K (reported as 90 oC), yields of 5.7% and 9.8% lactaldehyde and propylene glycol were achieved by ECH in 0.01 M H2SO4 after 9 h and current = 40 mA. Increased yields of lactaldehyde were achieved using HCl instead of H2SO4. Again using 11.1 mmol of lactic acid but at 343 K (reported as 70 °C), 0.0] M HCl and with a current = 100 mA a yield of 80.3% lactaldehyde was achieved. Only 2.9% of propylene glycol was formed at these conditions. The G3 composite ab initio method (in Gaussian 98 code) was performed to determine the heats of formation calculated from atomization enthalpies (Scheme 26). 56 OH OH OH OH OH H H -1420 —-2—H—> F: "2H ""> /|\/OH OH O OH O lactic acid 1,1,2-propanetriol lactaldehyde 1,2-propanediol (lactaldehyde hydrate) Gas -l47.9 -151.5 -142.8 -161.0 Phase Liq. -158.8 -163.5 -156.7 -177.3 phase Scheme 26. Heats of formation (kcal/mol) for lactic acid hydrogenation To confirm the formation of the intermediate aldehyde, glyceric acid, glycolic acid, and alanine were subjected to the same conditions and were all converted to the intermediate aldehyde (glyceraldehydes, glycolic aldehyde, and 2-amino propanol, respectively). The confirmation of the aldehyde intermediate was an interesting discovery because it was a proposed intermediate in the catalytic hydrogenation experiments and proposed reaction mechanism but was not confirmed. Dalavoy et al. looked at several reaction conditions to try and optimize the efficiency of the reaction. Results indicated that with increasing temperature, yield of lactaldehyde and propylene glycol increased. Using identical conditions (11.1 mmol lactic acid in 0.01 M H2SO4 for 9 h at 363 K) currents between 10 and 100 mA were examined. They found that with increasing current, the propylene glycol yield increases and the lactaldehyde yield decreases, while the total conversion remains nearly constant. A simple interpretation of this based on scheme G is that propylene glycol is formed by two 57 discrete intermediates (lactaldehyde hydrate and lactaldehyde). At low currents, the lactaldehyde hydrate is weakly bound and easily replaced by the lactic acid substrate. The, replacement occurs faster than the dehydration and capture by surface hydro gens. However, at higher currents the lactaldehyde would more quickly be converted to propylene glycol due to the availability of surface hydrogens. In addition, the increased amount of propylene glycol formed at higher currents may be due to the fact that there is higher electrode surface energy at more negative potential, this results in a higher mobility of H atoms. Three different electrolytes at concentrations of either 0.01 or 0.1 M were studied, HCl, HClO4, H2S04, using 11.1 mmol of lactic acid, 343 K and 100 mA current. In general, the rate of formation for lactaldehyde was found to decrease in the order HCl > HClO4 > H2SO4. The differences in ECH reaction rates are correlated to the extent of surface poisoning by the anions of the electrolyte. Chen et al. took the aqueous phase Ru/C catalytic hydrogenation reactions a step further by looking at competitive reactions with lactic acid and propanoic acid to their respective alcohol products, propylene glycol and 1-propanol.6' The reactions were carried out using several feed concentrations (0.05 to 5 M), temperatures (343 to 423 K), and hydrogen pressures (500 to 1500 psi reported as 3.4 to 10.3 MPa). However, for most reactions a pressure of 1000 psi and temperature of 403 K were used for most experiments. Catalyst loading was 0.5 g (dry basis) in 50 mL of aqueous solution. Scheme 27 shows the general hydrogenation reaction, conditions, and products. 58 ‘ f... OH + 2H2 + H O . . 5% Ru/C 403 K /l\/0H 2 OH Lactrc Acrd ? ’ g < 1000 psi H2 1,2-Propanediol o '1' 2H2 \/U-\0H J KH°\/\ + H2O Propanoic Acid l-Propanol Scheme 27. Catalytic hydrogenation of lactic acid and propanoic acid Chen et al. found that the reaction rate of lactic acid was nearly 4x faster than that of propanoic acid and that the conversion rates of both increased dramatically with temperature. These results support previous findings (discussed earlier) in our research group. In addition the selectivity to propylene glycol from lactic acid ranged from 80 to 99%, while selectivity to l-propanol from propanoic acid was less than 60% for most conditions. The lower selectivity to l-propanol is due to a faster degradation rate of l- propanol and also the slow hydrogenation rate of propanoic acid. Also, it was found that l-propanol adsorbs more strongly than propylene glycol on the catalyst Ru metal and carbon support, creating locally high pore concentrations of l-propanol near the Ru sites which contribute to its elevated degradation rate. At four feed concentrations (0.1, 0.5, 2, and 5 M) and otherwise identical concentrations. initial lactic acid and propanoic acid hydrogenation rates were calculated. A nonlinear dependence of the initial rate on the initial concentration indicates that the reactions are not first order and as previously found in our research group, suggests that a Langmuir- 59 Hinshelwood rate expression would best describe the kinetics. Activation energies were also calculated from plot of the natural log of the initial rate vs UT. The average activation energy for lactic acid was 52 kJ/mol and 68 kJ/mol for propanoic acid. In competition experiments, it was found that the addition of a second acid has an inhibitory effect on the conversion and this effect increases with concentration. However at concentrations over 2 M the catalyst is saturated with acids and thus the inhibitory effect becomes constant. The influence of the product alcohols were also and it was found that the presence of either propylene glycol of l-propanol in the starting solution had little effect on the lactic acid hydrogenation rate. However, l—propanol did have an effect on the hydrogenation rate of propylene glycol. Together these results suggest that the kinetic model should include these inhibitory effects in order to properly describe the reaction. Chen et al. performed a mass transfer analysis on the three-phase catalytic hydrogenation reaction and used these data and experimental reaction data to develop a kinetic model. Again using a Langmuir-Hinshelwood kinetic model and the same assumptions from Jere (discussed previously) a set of elementary reactions were used to describe the lactic acid and propanoic acid hydrogenation reactions. Again the rate limiting step is the irreversible surface reaction of the adsorbed acid and the hydrogen to form a surface bound intermediate. The resulting rate expression for single acid hydrogenation is presented in Scheme 28. 60 k . C . __ rAcid (kmol/kg of catalyst/s) = ACld AcrdPH 2 2 (1 + KAeidC Acid + KA1coho1CA1cohollll + KHz PH2 ) 2 kAcid = ksAKAcidKH2 Ct] (Ct2) Scheme 28. Rate expressions for single acid hydrogenation Performing a least squares regression analysis on 35 experiments at 403 K the adsorption constant for lactic acid was slightly more than propanoic acid (1.75 vs 1.36 m3/kmol) and both significantly more than that of propylene glycol (0.063 m3/krnol). However, as noted in the competition experiments, the adsorption constant of l-propanol is much higher (2.41 m3/kmol) and it’s presence in the starting solution decreased the propanoic acid hydrogenation rate. The rate constants for lactic acid and propanoic acid were calculated to be 1.03 x 10'5 and 1.66 x 106, respectively. The Langmuir-Hinshelwood model was extended to include adsorption of additional species for the mixed acid/alcohol hydrogenation reactions and worked reasonably well. Using the model, it was found for 0.5 M concentrations of lactic acid and propanoic acid at 403 K resulted in site occupancies of 39.1% vacant, 34.3% lactic acid, and 26.6% propanoic acid. However, with 0.5 M l-propanol added the fractions change to 26.6% vacant, 23.3% lactic acid, 18.1% propanoic acid, and 32% l-propanol. Propylene glycol was not included in the study due to its weak adsorption. Also, at the typical hydrogen concentration of 1000 psi, the fraction of occupied sites is 60%. These results suggest that at high concentrations (>2 M), the surface is saturated with reactants. 61 Zhang et al., who originally (2001, discussed above) studied aqueous phase Ru/C catalytic hydrogenation of lactic acid to propylene glycol expanded this work to. understand the effect of biogenic fermentation impurities on the hydrogenation reaction.62 The work done in our research group consisted of using reagent grade lactic acid, which allowed us to study different catalysts and reaction conditions as well as measuring intrinsic reaction kinetics. However, in order for propylene glycol production from lactic acid to compete with existing petroleum-based routes it is necessary to consider the use of lower cost, less pure lactic acid feedstock. The current petroleum-based routes produce propylene glycol from propylene involving either hydroperoxide intermediate of direct hydration of propylene oxide. Conversion of propylene to propylene glycol requires only 0.6 kg propylene/kg propylene glycol compared with 1.25 kg lactic acid/kg propylene glycol using lactic acid as the feedstock. So, in order to compete, the renewable feedstock must be on the order of one-half that of propylene. Currently, lactic acid is made from glucose via fermentation with Lactobacillus microorganisms. The yields obtained are nearly quantitative and the fermentation media has a complex mixture of nutrients, minerals, and base (Ca(OH)2) to neutralize the forming acid. After fermentation lactate is formed and is acidulated with H2SO4 to form lactic acid, which is then purified (~88% wt acid in solution). Characterizing the key biogenic impurities present in the fermentation process and their effect on catalyst performance in producing propylene glycol was the key objective in this work. 62 Three different lactic acid samples from along the purification train of a demonstration- scale lactic acid production facility were studied. The three samples were afully refined lactic acid (1 ), lactic acid that was withdrawn after the extraction/back extraction step (light brown color) (2), and lactic acid that was not refined beyond acidulation and filtration (dark brown color) (3). The catalyst used in this studied was 5 wt% Ru/C and was prepared in-house. Experiments were conducted in a three-phase trickle bed reactor using the following general conditions: 403 K (reported as 130 °C), 990 psi (reported as 6.8 MPa), 10 g catalyst bed mass, 1.0 mL/min feed rate, 1.35 M feed concentration, and H2: LA molar feed ration of 3.5 : 1. Results from the refined lactic acid (1) indicated that increased pressure (range studied: 200 to 1200 psi) improves both conversion and selectivity. In addition, a temperature of 373 K (reported as 100 °C) gave the best combination of activity and selectivity (range studied: 353 to 423 K). A long-term catalyst evaluation was performed (conditions: 120 kg refined lactic acid/kg catalyst at 403 K and ~1000 psi H2) over 370 h of operation and the conversion was found to be nearly constant at 90%. The liquid by-products identified included ethanol, n-propanol, and i-propanol in decreasing amount and together accounted for 3.5% of the lactic acid fed. The gas by-products identified included methane, ethane, and a small amount of propane in decreasing amounts and together accounted for ~12% of the lactic acid fed. To study the conversion of the partially refined lactic acids (2 and 3), a steady-state lactic acid conversion was established using the refined lactic acid (1) followed by switching 63 the feed to one of the partially refined lactic acids (2 or 3). For partially refined lactic acid 2, the conversion drops rapidly from 55% (established steady-state conversion using 1) to 30% for the first 5 h and slowly continues to decrease thereafter. For partially refined lactic acid 3, the conversion drops again from 55% to 10% in the first 5 h then essentially levels off. Switching back to the refined lactic acid feed results in a partial recovery and stabilization of the lactic acid conversion. In order to try and improve the conversion using the partially refined lactic acid (2), both heat treating (273 K at 5 h under helium) and bubbling ozone gas through the sample was tried, but there was no improvement in conversion. In an effort to model the effect of the impurities, the various impurities (salts, organic acids, extraction solvents, proteins, etc) were added to the purified lactic acid solution and the results interpreted. Three different salts (N aOH, Na2SOa, Na2HPO4) were added to the refined lactic acid feed. When adding NaOH, conversion of lactic acid decreased by an amount proportional to the amount of sodium lactate formed. These results are supported by previous work in our research group (Zhang et al., 2001). However, when Na2SO4 and Na2HPO4 were added there was little reduction in lactic acid conversion. So, addition of salts (PO43' and SO42) did not appear to poison the Ru/C catalyst. Several organic acids (glucose, sorbitol, succinic acid, or propanoic acid) and extraction solvents (dodecane, l-octanol, or trioctylamine) were added to the lactic acid feed and the results for each indicated no effect on the conversion of lactic acid. 64 To study the effect of protein and protein fragments, 0.12 wt% bovine albumin was added to the lactic acid solution. Results indicated a steady decline in conversion before essentially leveling off. The behavior observed partially matches the results from the partially refined lactic acid (2 and 3), indicating that proteins are partially responsible for the decreased activity for those feeds. Using nitrogen adsorption at 78 K the pore volume of the protein deactivated catalyst was studied. Results indicated that the deactivation was partly caused by a combination of pore plugging and sulfirr poisoning (sources: cysteine and methionine). The main sources of the sulfide were studied and when adding cysteine or methionine to the feed solution, the conversions declined nearly linearly over time. It was found that both cysteine and methionine strongly adsorb on the Ru/C and it results in poisoning of the catalyst. Due to previous interest in the substrate alanine, it was chosen to be studied as a non- sulfur-containing amino acid impurity. Results showed a significant drop in lactic acid conversion with the addition of alanine. However, when the feed stream was replaced by pure lactic acid, the conversion returned and remained at its original value. It was postulated that alanine reversibly and more strongly adsorbs on the catalyst surface, blocking the adsorption and reaction of lactic acid. These results are consistent with Jere et al.’s results (discussed earlier). Based on all of the studies, it appears that partial plugging of the pores and reversible amino acid (sulfur containing) adsorption are the main contributors to the reduced catalyst efficiency. Initial design considerations have been identified from this work and depending on purification requirements and cost, a more efficient biorefinery could be pursued. 65 Most recently published, Pimparker et al. studied the hydrogenation of three amino acids (alanine, serine, and valine) and their mixtures to their respective alcohols.63 The hydrogenation conditions were at 403 K (reported as 130 0C), 1100 psi (reported as 7.0 MPa H2), with 0.12 to 0.5 M feed concentration and 5-20% excess H3P04. Results indicated nearly comparable hydrogenation rates for alanine and serine (nearly complete conversion after 3.5 h), with significantly lower reactivity for valine (~10% conversion after 3.5 h). However, a selectivity of ~90% was achieved with both alanine and valine, compared with 70% for serine. A second-order polynomial was fit to the concentration vs time data and differentiation of the curves allowed rate data to be calculated. Initial rate data was used to estimate an activation energy and as seen previously in our research group, a nonlinear dependence on initial concentration was observed. The hydrogenation reactions appear to not be first order with respect to the amino acid concentration. The activation energies were calculated: alanine, Ea = 88.5 kJ/mol; serine, E5 = 83.1 kJ/mol; and valine, Ev = 95.5 kJ/mol and were comparable to values reported earlier by our research group (Jere et al. reported alanine, E,l = 86.4 kJ/mol). Mixed amino acid studies were conducted and the results indicated that as serine or valine is added to alanine, the alanine conversion decreases. The conversion continued to decrease as the amount of competing amino acid is increased (similar results as Chen et 'al., discussed earlier). Using an equimolar amount of competing amino acid (either serine or valine), the alanine conversion drops from nearly 100% to ~40% after 5 h of reaction time. A kinetic model was again developed to describe the hydrogenation reactions. Using the model developed by Jere et al. (discussed above) a two-site 66 Langmuir-Hinshelwood kinetic mechanism was proposed. Again, it was assumed that the protonated amino acids adsorb on one type of surface catalytic site (8,) and the hydrogen dissociatively adsorbs on a second type of site (S2). The optimized rate constants at 403 K for serine, alanine, and valine were 0.018, 0.016, and 0.0015 kmol/(kgcat h), respectively. The adsorption coefficients were calculated to be 56, 24, and 124 M‘1 for serine, alanine, and valine, respectively. The larger adsorption affinity for valine (as evidenced also by its essentially zero-order behavior) is not completely understood. Lastly, the Langmuir-Hinshelwood rate expression allows for insight into the relative surface concentrations of hydrogen and amino acid on the catalyst. For a single amino acid the relative concentrations (conditions: 0.22 M amino acid concentration; 0.3 M H3PO4) of vacant, H3PO4 occupied, and amino acid-occupied surface sites was 9%, 52%, and 39%. In experiments with two amino acids the site distribution was 4% vacant, 35% H3POa, 15% alanine, and 46% serine. Thus, the amount of active sites occupied by the original amino acid (alanine) decreased from 39% to 15%, resulting in a decreased conversion rate. 67 Chapter 2. EXPERIMENTAL 2.1 Hydrogenation Studies The catalytic studies were conducted in water, a “green” solvent, as opposed to the more traditional gas-phase hydrogenation widely used in the petrochemical industry. Advantages of developing aqueous-phase conversion reactions include the following: a) most bio-based feedstocks are produced and readily handled in water solution; b) nonvolatile organic acids can be converted without having to be esterified, as required for vapor-phase conversion; c) energy costs associated with vaporizing and condensing the process streams are reduced; d) feed purity is less critical; e) the low concentration of non-volatile residuals from fermentation need not be removed in liquid-phase processing as they must be in vapor phase conversion; and 1) conversion temperatures well below the boiling point of the acid or its ester are possible. 2.1.1 Materials Several three-carbon and two-carbon organic acid substrates were chosen in order to evaluate the extent to which the alpha substituent affects the hydrogenation rate. Substrates included the following substituents: -H, -CH3, -OH, -N+H3, -OCH3, - N+H2CH3, -N+H(CH3)2, and -N+(CH3)3 (note: purchased as commercially available HCl salt (CH3)3N+CH2COOH). Feeds were purchased from Aldrich Chemical Co. The acid used in the reaction was ACS grade 85% phosphoric acid (J .T. Baker). Water used was of HPLC grade purity (J .T. Baker, USA). Ultra high purity hydrogen (AGA Gas 99.999%) was used for all runs. One wt% dry 5% Ru/C, pre-reduced under H2 with 0.25 M substrate and 0.3 M H3PO4 in water was used. These quantities corresponds to a 50:1 68 substratezRu ratio (5% x 1 g /101.07 = ~0.5 mmol oftotal Ru and 100 mL x 0.25 M = 25 mmol of substrate). As detailed by Jere, the acid was needed for successful hydrogenation of the amino acids due to the zwitterionic nature of amino acids and the requirement of a protonated carboxylic acid group for hydrogenation. In order to be consistent, the H3P04 was added to all experiments. A 5% ruthenium on carbon powder catalyst obtained from PMC, Inc. (Severville, TN) was used for all experiments. The particles had a N2 BET surface area of 716 m2/ g, porosity of 0.6, a mean diameter of 150 microns, and a bulk density of 800 kg/m3. The as-received ruthenium metal dispersion was 8.8% (measured by volumetric hydrogen chemisorption in a Micromeritics ASAP 2010 instrument). This corresponds to a metal surface area of 1.6 mZ/g catalyst. Upon receiving the catalyst it was 50.9 wt % slurry in water. For all experiments and associated reaction rates a dry catalyst basis was used. 2.1.2 Batch Reactor System A 300 mL continuously stirred steady-state batch reactor (Parr Instrument Co, USA) equipped with a gas entraining impeller was used for all of the reactions. The temperature was maintained to $1 °C using a temperature controller that would either heat or cool the vessel in order to ensure the temperature set point. The reactor was constructed of 316 stainless steel and rated to a maximum pressure of 3000 psi at 350 OC. Analyses were performed during the course of the reaction via a liquid sampling system. The experimental conditions include reaction temperatures between 70 °C and 150 °C with the majority of the experiments conducted at 130 °C, 1000 psi H2, and six hours 69 reaction time. Hydrogen was added throughout the reaction, as needed, in order to maintain the pressure. The sample composition was analyzed every 30-60 minutes using high-pressure liquid chromatography, for which ~1-2 mL of solution was removed from the reaction vessel. 2.1.3 Reaction Procedure The reaction procedure consisted of pre-reducing 1 g dry weight (2.036 g wet) of Ru/C catalyst under 250 psi hydrogen at 250 °C, overnight. The reactor vessel was purged three times with nitrogen to flush out oxygen, then heated to the desired temperature. The impellor was set to ~100 rpm to facilitate a consistent temperature throughout the vessel. Following the reduction the reactor was cooled to room temperature and the hydrogen vented. A carrier vessel was attached to the reactor and a 0.25 M aqueous organic acid solution was charged into the reactor using hydrogen. After charging the reactor it was again purged three times with hydrogen. After the final purge, the heating and stirring (~1000 rpm) were initiated. When the reactor vessel reached the desired temperature, it was pressurized with hydrogen and a sample was taken. This sample was known as time zero, or the start of the reaction. Reaction temperatures were between 70 and 150 °C. Hydrogen pressure was set to 1000 psi for all reactions. Reaction times usually were six hours with samples being taken and analyzed every 30-60 minutes. At the end of the experiment, the vessel was cooled and the pressure was released. 70 2.1.4 Organic Acid Analysis 2.1.4.1 HPLC Analysis High-pressure liquid chromatography (HPLC) was used in the analysis of the hydrogenation reactions. Two analytical methods for each substrate and product set evaluated were required. A meticulous protocol of analytical standards and techniques was used. Internal standards were used for all experiments and the calibration of each compound was verified before and after all samples for a particular experiment. These methods allowed the absolute error of each data point to be assessed. In addition, several control experiments were designed to evaluate both the accuracy and repeatability of the analysis method. Method 1 A Bio-Rad Aminex HPX-87H column was used for all substrates, excluding the amino acids. The mobile phase consisted of 5 mM H2SO4. A Waters 410 Differential Refractometer and Hitachi UV Detector set to a wavelength of 210 were used for the detection. To ensure a constant injection volume a LDC/Milton Roy Autosampler was employed. The pump was a Spectra-Physics SP8810 Precision Isocratic Pump. The sample loop was 10 mL and the flow rate was optimized to 0.6 mL/min. Before each analysis a standard mix was run to evaluate the daily performance of the instrument with respect to the calibration curve. If a measurement error of more than 2-5% was obtained, new calibration standards were prepared and a new calibration was performed. The calibration curves were 3-point curves for all compounds (0.02, 0.06, and 0.1 M). The mobile phase consisted of 5 mM H2SO4. 71 2.1.4.2 pH Analysis The pH of the liquid samples was measured and recorded using a bench top pH meter. The meter was calibrated before each use using solutions of pH 4, 7, and 10. 2.2 H/D Exchange Studies Valuable information related to retention of stereochemistry during hydrogenation, has been ascertained by our group using H/D exchange studies. To build upon the knowledge gathered from the hydrogenation of (Jr-substituted organic acids, an extensive H/D exchange study was done with glycine and its N-methylated analogs and will be discussed-in this thesis. 2.2.1 Materials All substrates were reagent-grade, including glycine (crystalline, 99%), sarcosine (98%), N,N-dimethylglycine (99%), betaine (98%), betaine hydrochloride (99%), ethanolamine (99%), N-methylethanolamine (98%), N,N-dimethylethanolarnine (98%), choline chloride (98%), and choline bicarbonate (80%) were obtained and used as purchased from Aldrich Chemical Co. HPLC grade purity water was used for all aqueous phase experiments (J .T. Baker, USA). The hydrogen used was ultra high purity (Aga Gas, 99.999%) and the deuterium was ultra high purity (Spectra Gases). For deuterium exchange experiments, deuterium oxide (Sigma Aldrich, 99.9%) was used in place of water. 72 2.2.2 Batch Reactor System Paramagnetic broadening was observed in early 1H NMR studies, presumably caused by solution contact with the stainless steel reactor and its parts. Thus, extensive effort was put forth to minimize the amount of stainless steel exposed to the reaction mixture. The 300 mL stainless steel batch reactor (Parr Instrument Co.) was custom coated with a Teflon based material by Unconventional Solutions (Milford, MI). The stainless steel dip tube used for liquid sample removal was replaced with a glass dip tube. A Teflon tube was placed over the stirring shaft. The stainless steel impellor was replaced with a titanium impellor. The stainless steel cooling loop was removed from the system. This eliminated the opportunity for both heating and cooling; therefore, in order to compensate for the lack of dual-control, the rate of heating was slowed. A glass tube was placed over the stainless steel thermocouple. The glass tube was filled with a small amount of water to facilitate the heat transfer. 2.2.3 Reaction Procedure The reaction procedure consisted of pre-reducing 1 g dry weight (2.036 g wet) of Ru/C catalyst under 500 psi hydrogen at 150 °C, overnight (although the hydrogenation studies above had used 250 °C, it was necessary to reduce the temperature from 250 °C to 150 °C in order to maintain the integrity of the Teflon lining). The lower temperature of 150 °C proved to be sufficient; however, a longer reduction time (overnight) was used (as opposed to 4 h). Several experiments were conducted using either a catalyst reducing temperature of 250 °C or 150 °C to evaluate whether there was an effect on the hydrogenation reaction. For the substrate glycine (0.25 M) with 0.3 M H3PO4 at 130 °C 73 and 1000 psi H2, the hydrogenation rate was nearly identical for both reducing conditions. The reactor vessel was purged, heated to the desired temperature, and stirred using the same method as described previously. The carrier vessel was attached to the reactor and the 0.25 M aqueous (D2O) solution was charged into the reactor using hydrogen or deuterium. The charged reactor was then again purged three times with hydrogen or deuterium. After the final purge, the heating and stirring (~1000 rpm) were initiated. When the reactor vessel reached the desired temperature, it was pressurized with hydrogen and a sample was taken. This sample was defined as time zero, or the start of the reaction. Reactions were usually conducted at temperatures between 50 and 100 °C. Hydrogen or deuterium pressure was set to 1000 psi. Reaction times usually were six hours with samples being taken and analyzed every 30-60 minutes. At the end of the experiment, the vessel was cooled and the pressure was released. 2.2.4 Amino Acid Analysis Both HPLC and lH NMR were used in the analysis of the I-I/D exchange studies. Results from the HPLC allowed for quantitation of the various substrates and products for each sample (samples collected every 30-60 min for six hours). Results from the 1H NMR allowed for individual isotopomer ratios to be calculated for each sample. Using these data, kinetic models were developed in order to model and understand the deuterium exchange behavior. 74 2.2.4.1 HPLC Analysis Method 2 A Hewlett Packard Series II 1090 liquid chromatograph instrument with a Waters 410 differential refractometer and Hitachi UV detector model L-4000H was used for the amino acid analyses. The mobile phase was 0.01% acetonitrile in 0.05 M phosphoric acid. 2.2.4.2 ‘H NMR Liquid samples were analyzed using a Varian 500 MHz superconducting NMR- spectrometer operating at 499.738 MHz interfaced with a Sun Microsytems Ultra 5 UNIX console. NMR tubes with an outer diameter of 5 mm were used (Wilmad- Labglass, Buena, NJ). A glass insert (2 mm outer diameter; 1.25 mm inner diameter) filled with 0.25 M acetic acid internal standard was used for all experiments. The chemical shift of the internal standard was set to 1.9008 ppm. The integration area of the internal standard was set to 30 for all experiments and based on calibration curves, the concentrations of the observable amino acid isotopomers were calculated. 75 Chapter 3. CALCULATIONS 3.1 1H NMR Peak Fitting and Deconvolution 3.1.1 NMR Calibration Calculations The slope of the area versus the concentration for all compounds was determined both experimentally and mathematically. Experimentally, three-point calibration curves were constructed for all amino acid and amino alcohol compounds. On average, the slope of the calibration curves was 9. Mathematically, the ratio of outer NMR tube to inner insert tube volume was calculated based on the dimensions of the tubes. Equation 1 below was used for the calculation. Equation 1. NMR tube ratio: outer-to-insert volume z ((oo — wt)2 - (013(3)) 2 , where IDin OD = outer diameter (mm) of outer NMR tube wt = wall thickness (mm) of outer NMR tube OD = outer diameter (mm) of NMR insert ID,n = inner diameter (mm) of NMR insert Using the low and high values specified by the manufacturer for the dimensions, an average of 9.008 :t 0.298 was calculated for the volume ratio. The prepared concentration of the internal standard, acetic acid, was 0.249 M. The peak integral for the internal standard, acetic acid, was set to 30. Thus, the peak area per hydrogen for the internal standard was 10. 76 3.1.2 Peak Fitting Ecmations The data from the IH NMR was exported into Excel and using a Lorentzian function, the peak integrals were determined by fitting the data and using deconvolution. The general equation for a Lorentzian function is found in equation 2: Equation 2. Lorentzian lineshape = (g) * - 1 ), where g = half-width and x0 = unshifted peak center 7‘ ((x-xo)2 +g2 The integrated area of this function is unity. Global values and local values were defined. For each H/D exchange reaction, after Fourier transform and phasing, the NMR spectra were exported into Excel as x,y data where y represented the peak height and x represented the peak position, ppm. The data were then analyzed in parallel as a set. For each such set, the fitting procedure defined quantities that were global (i.e. common to all spectra) and those that were local (i.e. specific to a given spectrum). Global values included isotope shifts, coupling constants, and multiplet proportions, while local values included chemical shift, baseline offset, linewidth, and of course, the individual deconvoluted peaks' intensities. Least-squares fitting to all datapoints was used to align and quantitate the Lorentzian functions. The respective fractional areas were then sealed to molarities using the ratios of measured total substrate peak integral to that of the internal standard. Modeled spectra and their integrals were plotted and compared to the originals to ensure accuracy. In the case of a triplet, a triplet scaling factor was defined. And in the case of a quintet, a quintet outer and inner scaling factor was defined. 77 The equation used for fitting the data is below (3, 4, 5). It is broken into three parts representing the singlet (CH2 or CH3), triplet (CHD or CH2D) and quintet (CHD2) resonances. In the case of fitting the data which represented the methylene, only the first two parts of the equation (3 and 4) were used. The portion of the equation used to fit a quintet was omitted due to the fact that for the methylene the singlet represented CH2, and the triplet represented CHD. However, in the case of the methyl, all three parts of the equation were used (3,4,5). The singlet represented CH3, the triplet represented CH2D, and the quintet represented CHD2. Equation 3. Singlet peak fit r (a0*[§] \ = baseline offset + 2 (X-Xo)2+g \ J Equation 4. Triplet peak fit / + a3 * {—g—D ( 31! s3 1 53 + + + ((x—xo +.IS+HI))2 +g 2) ((x—xo +IS)2 +g2) ((x—xo +IS—HD)2 +g 2) 78 Equation 5. Quintet peak fit (sell SSout 2"‘S‘Sin + 2 2 U ~x0+2*IS+2*HD)2+(350ut*g)2) (x—x0+2*IS+HD) +(55in*g) :1: + 3 ((x—x0+2*IS)2+(g)2) ((x—xo+2*IS—HD)2+(35in*g)2) (x—XO+2*IS—2*HD)2+(SSOut*g)2 Samples were analyzed using 1H NMR every 30-60 minutes of the reaction. Using the above models to fit the full time point sets of NMR spectra, the global values described above were optimized for all time points. The local values were individually optimized for a particular time point (i.e. NMR spectrum). The local linewidth was defined by the global linewidth multiplied by the local broadening. The global linewidth for the singlet was usually in the range of 0.0006 to 0.001 ppm. The global linewidth for the triplet was usually slightly larger and in the range of 0.0009 to 0.002 ppm. The global isotope shift was between 0013-0015 ppm. The global H-D coupling constant was 0.003 to 0.006. The local values were dependent on the individual experiment. Starting with a good estimate of the global and local values, the sum of the squares error was minimized for each spectrum using Solver. Once the fits were optimized for individual spectra, Solver was run on all spectra to obtain one set of values for each experiment. The sum of the integrals was calculated for both the actual data and the 79 fitted data and compared. Using the amplitudes (a1, a3, and a5), the mol% of singlet, triplet, and quintet was calculated. 3.1 .3 Peaflrtegration Calculations MestReC is a powerful NMR software package used for analyzing high-resolution NMR data (Mestrelab Research, Spain www.mestrec.com). The raw F ID data was exported to Excel, then to calculate the individual concentration of isotopomers, the below equations were solved. Equation 6. Total H concentration for each spectrum at time t [IS]* (areasfl ) area [SJ #of H 1H1= , where 9.008 *( [IS] = concentration of the internal standard, acetic acid (0.24881) area s,t = area of the sample at time t area 13,1 = area of the internal standard, acetic acid, at time t (set to 30) # of H = 3, for the 3 methyl hydrogens in acetic acid (internal standard) The value 9.008 is the slope of the calibration curve for the sample versus the standard (see NMR Calibration Calculations 3.1.1). The area fractions as [H] were determined using the following equations (7,8,9). The variables a1, a3, and a5 represent the mol% of the singlet, triplet, and quintet, respectively. For example, to calculate the singlet mol%, al, for methylene the amplitude of the singlet is divided by 2; divided by the sum of the amplitude of the singlet divided by 2 plus the amplitude of the triplet. 80 Equation 7. Singlet area fraction al = H =1: __ . [ ]t(a1+a3+a5) Equation 8. Triplet area fraction a3 = H * ____..__ [ ]t(a1+a3+a5) Equation 9. Quintet area fraction a5 = H * .______ [ 1‘ (al+a3+a5] The concentration of isotopomers at each time point can then be calculated using the equations below. Equation 10. CH2 concentration Equation 11. CHD concentration [CHD] t: (area) CHD Equation 12. CD2 concentration [CD21t=(lCHzlt0+lCHDItol—(ICHzlt+lCHDl t) The equations for the methyl are analogous to the above equations. Next, the peak areas of each of the isotopomers can be derived using the following equation. 81 Equation 13. Peak area of individual isotopomer [s] ”(Till ]*9.008 0 [IS] (areaS )t = , where S = individual isotopomer The conversion, yield and selectivity can be calculated from the concentrations of the individual isotopomers. Equations 14, 15, and 16 were used. Equation 14. Conversion = [S] to- [S] t 13] t0 Equation 15. Yield _ [Slt ’1—81—5 Equation 16. Selectivity , where S = individual isotopomer and to = time zero _ yield conversion 3 .2 H-D Content Inventory 3.2.1 Percent H in Water The molecular weight of H20 is 18.015 g/mol and the molecular weight of D20 is 20.028 g/mol, and at 25 °C, their respective densities are 0.9971 and 1.104 g/mL. Thus their concentrations are 55.34 and 55.14 M, respectively. Their equivalent hydrogen or deuterium concentrations are 110.68, [H]H20, and 111.28, [D]D2o. The internal standard, acetic acid, in the insert had a bench concentration of 0.248 M (at 25 °C), thus the equivalent hydrogen concentration was 0.746 M. The peak area of the internal standard 82 was set to 30. Using the ratio of the outer-to-insert volume of 9.008, the concentration in the insert was calculated to be 0.083 M (not the true concentration, but would be equivalent to this if it had been in the whole volume). During all calculations the untruncated values were used until the end. If the water was H2O (no D20 or HOD) then the water peak should integrate to 40,173. Equation 17. Peak area of water in sample [1H] H20 — * (area),S = 40,173 [15} in J The acetic acid was prepared in deuterium oxide and the 1H NMR of acetic acid was taken in the outer NMR tube (not the insert). If the water in the internal standard, acetic acid, was composed of H20 (no D20 or HOD) then the water peak should integrate to 4,462. Equation 18. Peak area of water in internal standard [M US] )*(area),5 = 4.462 When measured, the peak area of the water in the acetic acid was 52.52, thus the %H in the deuterium oxide of the internal standard was taken as 52.52/4,462, thus 1.2%. 3.2.2 Percent H in Sample The molarity of hydrogen and deuterium could be determined experimentally from the NMR integrals and also mathematically by the following equations. 83 Mathematical: The initial liquid volume in the reactor, VL, was 100 mL. The pressure, P, was 1000 psi (68.046 atm). The temperature, T, was 100 °C (373.15K). The initial gas volume, Vg in the reactor, was 200 mL. Using the below equations, an example calculation for sarcosine will be carried out. Equation 19. Moles of deuterium in D20 (1) * V moles of DL 2 [p L j , where mw = molecular weight mw The moles of DL are 5.55, thus in D2O there are 11.11 moles of D. In practice, we will assume [H] and [D] are equal to ~111 (used 111.11). Equation 20. Moles of hydrogen in H2 (g) , P * Vg . moles of Hg = 2 * R T , where R = 1deal gas constant * The moles of Hg are 0.438, thus in H2 there are 0.876 moles of H. Equation 21. Substrate — NHx contribution moles of HN = x * [substrate]* VL , where [substrate] is the prepared bench concentration The typical bench substrate concentration is 0.25 M, thus for sarcosine the moles of HN are 0.05 (x=2). 84 Equation 22. Substrate - methylene (CH2) and methyl contribution y*(CH3) moles of HC = (2 + y * 3)* [substrate]* VL , where y = number of methyls Again, using the typical bench substrate concentration of 0.25 M, the moles of Hg in sarcosine are 0.125 (y=1). Using equation 19 to calculate the moles of deuterium and equations (20, 21, and 22) for the sum of the moles of H in the different isotopomers, the H-D inventory can be determined. Equation 23. Time zero H-D inventory ZmolesofH ]*100 Zmoles of H + Zmoles of D o/°Ht0 =( The sum of the moles of H at time zero is calculated from equations 20, 21, and 22 and equal 1.05 for sarcosine. The sum of the moles of D at time zero was calculated from equation 19 and equal 11.11. Thus, the %Hto in the sample (sarcosine) is 8.6%. Equation 24. Molarity of H at time zero 2 moles of H j H =H * [ ]t0 [ ]H20 l2molesofH+ZmolesofD Based on the above calculations, the molarity of H at time zero for sarcosine is 9.602 mol/L (again where [H] H20 = 111.11). Equation 25. Molarity of D at time zero [DI t0= [H] H20 — [HI t0 85 The molarity of D (for 0.25 M sarcosine in D2O) at time zero is calculated to be 101.51 mol/L. Once the reactor has reached the reaction temperature, the vessel is pressurized and the reaction is started. A sample (~2.5 mL, Vremoved, which is equal to 0.14 moles) is removed and is referred to as the “time zero” sample. Then, the hydrogen and deuterium composition of the remaining solution is calculated from the NMR analyses using the below equations. Equation 26. Moles of D (l) at time zero moles of D = 2 * ( Vremoved * p :1: 2 moles of D to mw (2 “10165 0f H)’ “10135 0f He + 2 moles of D Equation 27. Moles of H (l) at time zero V * molesof H10 = 2*( removed p]*[( (2 moles of H) — moles of HC mw 2 moles of H)- moles of HC + Z moles of D The number of moles of Hc are subtracted out from the total moles of H at time zero because at time zero a negligible amount of methylene and methyl hydrogens are exchanged, thus more deuterium is removed. However, after time zero a H-D equilibrium is assumed. After the sample (~2.5 mL) is removed, the reactor is re- pressurized to 1000 psi. The amount of hydrogen re-introduced to the reactor is ~100 psi. Thus the moles of hydrogen added at that time can be calculated from the equation below. 86 Equation 28. Moles of H (g) at time zero P*Vg R*T moles of H gto = 2 *[ J, where R = ideal gas constant From this point, it is an iterative process whereby the composition can be re-calculated after each sample (~2.5 mL) is removed. Equation 29. Moles of D (l) at time t moles of D” = moles of DL — moles of Dto Equation 30. Total moles of H (1) and H (g) moles of H” = (Zmoles of H)— moles of Hto + moles of Hg,t0 Equation 31. %H at time t moles of Ht] °/H = 0 t1 molesothl +molesof D” Equation 32. Molarity of H at time t [HItl = lHlHZO * (%H)tl Equation 33. Molarity of D at time t lDlt1 = [HIH20 -[HItl After this point, the process repeats. The time 2 sample is taken (~2.5 mL) and the new composition of the solution is calculated. After the reaction is complete, the final sample composition (%H, [H]tf, [D]tf) can be compared with the experimental values obtained from the NMR integrals. The results from both methods were in agreement. 87 3.3 Kinetic Rate Data Determination 3 .3 .1 Concentration Calculations The peak areas from the 1H NMR fitting are used to calculate the concentration of the individual isotopomers for each sample time. The methylene carbon is referred to as “A” and the methyl carbon is referred to as “C”. The concentrations at time zero were determined using the internal standard, acetic acid, and the ratio of area to concentration. Again, the slope was determined to be 9.008, the [IS] = 0.25 M, and the peak area was set to 30. All concentrations after time zero were based on the initial concentration of the individual isotopomer and ratio of the area at time, t, to the area at time zero. However, the total concentration of all isotopomers was verified using HPLC to ensure the starting material was not consumed before time zero (once the reactor has reached reaction temperature and pressure) of the experiment. The below equations were used. Equation 34. AH2 concentration at time zero (area AH2//) 2 - . t0 :1: [[5] (area 1.8/3')t0 9.008 IAHzltO = Equation 35. AH2 concentration at time t (area AH2/j t l H I l H l /2 A 2 t — A 2 t0*(areaAH2/) t0 /2 88 Equation 36. AHD concentration at time t * [area AH%) t t0 (area AHD/2 ) t0 Equation 37. AD2 concentration at time t [AHD], =1AH21 [AD2], =1AH21 ,0 —11AH21.+IAHD1,) Equation 38. CH3 concentration at time zero (areaCHg, ) 3 - . ‘0 *[IS] (area 1% )t0 9.008 [CH3]t0 = Equation 39. CH3 concentration at time t (area CH 3 /) I / [CH3], =ICH31,0.( 3 area CH3/j / 3 t0 Equation 40. CH2D concentration at time t area CH2D I /21. area CH3 ( /3th Equation 41. CHD2 concentration at time t [€11let = [CH3lt0- (area CHD2 )t t0* areaCHy 3 t0 Equation 42. CD3 concentration at time t [CHD2]t = [CH3] [CD3], =[CH31t0-IICH3I ,+[CH2DI t+lCHD21 t) 89 Equation 43. Aqueous hydrogen concentration in H20 and HOD average([AH2],[CH3]) average((area AH2/2),(al’€a CH3/3)) , [H]L =(area HL )*[ where [ average([AH2 ],[CH3]) is equivalent to a calibration based average((area AH2/2),(area CH3/3)) on the internal standard (IS) acetic acid concentration to area and equals 9.008. Equation 44. Aqueous deuterium concentration in D20 1000 D = --— — H The sum of the hydrogen area can be calculated for each time point, and during the course of the reaction the areas for C-H sites decrease as hydrogen is replaced with deuterium. Equation 45. Sum of the H area 2 Harea = (area AH2 ) + (area AHD) + (area CH3 ) + (area CH2D) + (area CHD2) The aqueous hydrogen concentration data was fit to the following equation over time. Equation 46. Aqueous hydrogen concentration equation of fit [HlL = A * exp(- B * t)+ C * t + D , where A, B, C, and D are fitted variables and t = time The sum of the squares error is minimized to determine the fitted variables. 90 3.3.2 Kinetic Model A: 10-variables The time zero concentrations for the isotopomers of the methylene and methyl sites were determined from the time zero areas for each. The kinetic model assumed that the adsorption/desorption of H20 (or D20) is fast and that the ratio of H:D in the solution represents the H:D on the surface. This assumption proved to be adequate for the experiments with 1000 psi D2 and D20. However, after reviewing the results from the experiments with 1000 psi H2 and D20 a modified kinetic model was used (to be described later). Using the original kinetic model, for experiments with 1000 psi H2, the initial hydrogen on the surface was assumed to be 111.11. The initial deuterium on the surface was assumed to be zero. For experiments with 1000 psi D2, the initial deuterium on the surface was set to the amount of deuterium in the liquid (usually less than 3%) and the hydrogen was equal to the difference between 111.11 and the amount of deuterium. The hydrogen and the deuterium in the liquid at any time, t, was calculated using the rate equations in 47. The experiment describing the hydrogen-deuterium exchange on the methylene carbon is depicted by the following two reversible, consecutive first-order reactions. Again the role of the catalyst, including the adsorption/desorption events are represented in the aqueous hydrogen (HL) and consequently the aqueous deuterium (DL). 91 Equation 47. Rate equations for H/D exchange at the methylene position km AH2+DS (TIA—2 AHD+HS sz AHD + Ds <__—-> AD2 + HS 2A Hence, doing a numerical step-by-step fit, where the time-step increment (At) was 0.5 min, the equations become: Equation 48. CH2 concentration at time t [AHZL = [AHZL—At _]k [A *[AH211-AI*[D]L,t.—At _ k _]A *[AHDlt—At *[H]L,I—Al )* At Equation 49. CHD concentration at time t lk-IA * [AHDIt—At * [HlL,t—At + sz * [AHDlt—At * [D]L,t-At) * At - (km * [AWL—At * [D]L,t-At 3‘ k—ZA * IADzlt_At * [H]L,t—At) [AHDIt : [AHDlt—At '— Equation 50. CD2 concentration at time t [AD2]. await... — It-.. wins]... 41312.2, - t2. *[AHDI._..*1D1,,_,.1*At The experiment describing the hydrogen-deuterium exchange on the methyl carbon is depicted by the following three reversible, consecutive first-order reactions. Equation 51. Rate equations for H/D exchange at the methyl position I 5 CH3 '1' DS k C $—— CHzD + H8 ‘IC k2C CH2D+DS (‘15—'— CHD2 +HS '2C k3C CHD2 + Ds 61:— CD3 + HS ~3c 92 Hence, doing a numerical step-by-step fit the equations are defined as follows. Equation 52. CH3 concentration at time t [CH3]t = [CH3]t—At ‘ 1le *[CH31t—At * [D]L,t-At — k-IC * lCHlet—At * [HIL,t-At )* At Equation 53. CH2D concentration lk—1C * [CHlet—Ar * [Hint-At + kZC * [CHZDL—At * [DlL’t—At )] * A = D — [CHZDL [CH2 L—At i'lklc * lCHslt—At * [D]L,t—At + k~2C * [CHDllt—At * [mm-Al) Equation 54. CHD2 concentration at time t (k-ZC * [CHD2 lt-At * [H]L,[—At + k3C * [CHD2 lt-At * [D]L9t_At )] * A _ = H _ [CHD2 L [C D2 L—At |:— (kzc * [CHlet—At * [D]L,t—At + k—3C * [CD3 lt—At * [H]L,t—At ) Equation 55. CD3 concentration at time t [CD3 ]t = [CD3 lt—At - lk—3C * [CD3 lt—At * [HIL t-At _ k3C * [CHD2 lt—At * [D]L, t-At )* At The 10-variables: km, k-1 A, k2A, k-2A, k1c, klc, k2c, k-2c, k3c, and k-3c were calculated by minimizing the sum of the squares error between the experimental and the calculated concentrations for all measured time points. To verify that this was a short enough time increment for direct analysis of these multi-hour reactions we confirmed that the resulting k values were not significantly affected by further 2-fold shortening or lengthening of the time increment. 3.3.3 Kinetic Model B — 6 variables The second kinetic model used to fit the data incorporated primary and secondary isotope effects. All of the rate variables above (Model A) represent the same hydrogen trading 93 processes except for the isotopes involved. Thus it should be possible to simplify the model. The equations from Model A were used in Model B, but here, the 10 rate variables were re-defined as isotope-perturbed instances of the same events in order to incorporate the possibility of a primary and or secondary isotope effect. The six rate variables used to optimize the data fit were: kiA, pA, SA, kic, pc, and sC. The kinetic 2 isotope effect at the methylene carbon and methyl carbon(s) is represented by km and kic, respectively. A primary kinetic isotope effect arises when the bond to the isotopic substitution is broken at or before the rate-limiting step and for the methylene carbon and methyl carbon(s) is represented by p A and pc, respectively. When hydrogen-deuterium substitution is not involved in the bond breaking or bond making step, an isotope effect (albeit smaller) may still be observed and is represented by SA and sc for the methylene carbon and methyl carbon(s), respectively. Equation 56. Forward and reverse rate constants for H/D exchange at the methylene position kIA = 2 * kiA k. k-IA =45- PA k. sz =4}: SA k 2*kiA -2A =———— (PA *SA) 94 Equation 57. Forward and reverse rate constants for H/D exchange at the methyl position le=3*kiC k. 2*k'C k2C= SCI k _ 2*kiC 'ZC (PC*SC) k. k =_£_ 3C (SC*SC) k 3*kiC '3C :(PC *SC ”CI The time step used in calculating the kinetic rate variables was 0.5 min. Kinetic Model Modification The kinetic model for isotope exchange was varied to explore the effects of various assumptions with the intent of optimizing the fit to the experimental data. Experiments were conducted with either 1000 psi H2 or 1000 psi D2 and for reactions at the same temperature, the calculated rate constants should be the same. Using the kinetic model discussed above, the H/D exchange for the experiments with 1000 psi H2 and D20 appeared to be going to fast for some of the early experimental points. This is partly due to the fact that in the kinetic model, the hydrogen and deuterium available for exchange with the substrate was assumed to be in equilibrium with the concentration of hydrogen or deuterium in the liquid. But from the experiments with H2 and D20, it appears that 95 there is a delay in the substrates' H/D exchange reaction due to the time it takes to reach equilibrium among liquid, gas, and surface hydrogens. For experiments with 1000 psi D2, the aqueous hydrogen concentration is almost an order of magnitude lower than the typical 15% or so that is observed with H2 (Figure 67). This time evolution is important, but also reflects the gas/liquid exchange. To attempt to address this, the hydrogen and deuterium available were defined by equations below. The equations for calculating the hydrogen and deuterium on the surface, assuming the amounts are in equilibrium with the liquid concentrations, at any time are below. Equation 58. Expression for H on the surface at time t Hat = [Hlear +(2*IAH2].-a +[AHDIt_At +3*[CH31._A. +2*[CH2DI..ai +lCHDzl.-At) — (2 *[AH2]1+[AHD]t + 3 * [CH3]t + 2 - [CHZD]t + [CHDZM - k H * llD]L,t—At * “st—At — [H]L,t-At * Ds,t—At) Equation 59. Expression for D on the surface at time t D“: 111.11—H53 The kinetics for the surface hydrogen incorporate the hydrogen-deuterium surface-liquid exchange represented by kH. At any point in time, in the solution the sum of hydrogen and deuterium concentrations equals the total concentration [Ct]. During the reaction, we will assume the hydrogen on the surface and the deuterium concentration in solution is in equilibrium with the deuterium on the surface and the hydrogen concentration in solution. 96 Once the equilibrium between the hydrogen and deuterium is reached, their concentrations in the liquid and their respective amounts on the surface can be expressed by the equilibrium equation below. Equation 60. Equilibrium expression for hydrogen and deuterium in the liquid and on the surface [DIL + H5 ‘—— lle. + Ds Let HS + DS = Ct, the total catalyst site density (Ct) for hydrogen adsorption in the reactor. From this, the rate of deuterium on the surface versus time can be derived. Equation 61. Rate of deuterium on the surface versus time “135%” = k a.r([[)]L :1: H5 ”l-Hll. =I= D5), thus = k*([D]L -(c. -D.)-D. *(lctl—[Dlr)) Assuming the total concentration in solution is in equilibrium with the surface, then the total catalyst site density for hydrogen adsorption, Ct, equals the concentration of hydrogen and deuterium in the solution, [Ct]. Factoring this out of the equation, the effective rate constant can be described as k*Ct and is defined as kH in the rate equations. D, is a variable and represents the fraction of total sites occupied by deuterium. This variable can be used in all kinetic expressions recognizing that all the rate constants will then also be effective rate constants with C, in them, kH=k*Ct. 97 Chapter 4. RESULTS AND DISCUSSION 4.1 Control Eigeriments 4.1 .1 Hydrogenation Studies Early research in our group highlighted the need for both catalyst and hydrogen pressure in the hydrogenation of organic acids.25’48’49‘50 Several control experiments have been performed and results confirm that molecular hydrogen and catalyst are required for the hydrogenation reaction. Using otherwise identical procedures, a reaction was performed replacing H2 with helium. The catalyst was exposed to 1000 psi He pressure. No product formation was observed for the reaction of 0.25 M refluxed lactic acid at these conditions (H20, 130 °C, 1000 psi He, 1 g dry 5% Ru/C). 0.30 + Lactic Acid +1.2 Propylene Glycol A o.25~——- :g— -— v o/T— l l l l 41 0 l Concentratlon (moIlL) O a l l l l l l l l l l l g: 8 i l l l l l l l l l l 0.05-~——-—~—-»e- -- . f. - , ~ ~ ~ ._ 0.00 FRY“. I I r . r I I l I . 0 50 100 150 200 250 300 350 400 Tlmo (mln) Figure 1. Hydrogenation of 0.25 M refluxed lactic acid (0.3 M H3PO4, H20, 130 °C, 1000 psi He, 1 g dry 5% Ru/C) 98 The concentrations of lactic acid and product alcohol, 1,2-propylene glycol, are shown in Figure 1. The purchased (Sigma-Aldrich) lactic acid was 85% (w/w) in H2O, so before using, the water was removed by refluxing the lactic acid. A reaction with 0.25 M refluxed lactic acid in the absence of 5% Ru/C catalyst was performed and the results indicated no product formation after six hours (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, no catalyst). The concentration data for lactic acid and the desired alcohol product, 1,2—propylene glycol are shown in Figure 2. 0.30 -~—-~—---r, A A 9 ¢ e e 4 02516:-“ L- L L- a- +Lactic Acid -—I— 1.2 Propylene Glycol 5' 0.20 1%.. ~- —— ~A ~~ - ----- - --~ 0 .5. 8 3015 — r - 74 i ' c i e . O l e . 8 0.10 - - - ~— -— — , — t 0.05 -..._._- , , - ~~ -72 77~ , 7 ,, , ~ emu-g i 0.00 I———I———-r—I . I . I r I . I . I i 0 50 100 1 50 200 250 300 350 400 Time (min) Figure 2. Hydrogenation of 0.25 M refluxed lactic acid (0.3 M H3P04, H20, 130 0C, 1000 psi H2, no catalyst) For all experiments discussed in this thesis, the reaction procedure involves pre-reducing the 5% Ru/C catalyst under H2 pressure (150 °C overnight with ~300 psi H2). The sample is then injected and the reactor is heated to the reaction temperature, 130 °C (one- 99 step introduction of the solution). One-step introduction involves adding the substrate (0.25 M) at the point in which the reactor has reached the desired temperature. The next step is to immediately pressurize the vessel with H2. The point when the H2 is introduced is the point that the reaction is considered to have started (t = 0 min). In some experiments a small amount of conversion was observed at time zero. So, a ' control experiment was designed to evaluate the effect of adding the substrate after the reaction vessel was at temperature and pressure. This was done in an effort to avoid any pre-reaction conversion that can occur during the heat up phase of the reactor. The refluxed lactic acid was mixed with 25 mL of water and set to the side. The other 77 mL (all H2O) of the 102 mL solution was injected into the reactor and allowed to heat to the reaction temperature. Once at the desired temperature, the lactic acid solution was injected (two-step introduction of the solution). Results indicated no significant difference between this and introducing the sample at room temperature and allowing it to heat up (in the absence of H2) in the reaction vessel to 130 °C (Figure 3). The bench concentration for the two-step was 0.32 M (1.3 M in 25 mL) and slightly higher than the traditional experiment in which the bench concentration is 0.25 M (one-step introduction). 100 0.40 a —— _____-_.__. .. ..-.---_.-._ .- —I-— Lactic Acid -—I~-~~ 1.2 Propylene Glycol - A— - Two-Step Lactic Acid — -A~- - Two-Step 1.2 Propylene Glycol .0 N or Concentration (molIL) p o —b N 0’! O 0 50 100 150 200 250 300 350 400 Time (min) Figure 3. Hydrogenation of refluxed lactic acid (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) All hydrogenation experiments were carried out in aqueous solvent in the presence of 0.3 M H3PO4. A control experiment using 0.25 M glycolic acid (GA) in the absence of additional acid was conducted to determine the effect of equimolar H3PO4 on the hydrogenation rate. No significant difference in the conversion of glycolic acid and the yield of ethylene glycol (EG) was observed over six hours (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C). The concentrations of glycolic acid and product alcohol, ethylene glycol, are presented in Figure 4. 101 025 W ______.-._. ....._-_._..- .. .........-... . . —I— GA 4 - EG ~A~ GA with addtl acid A EG wlih addtl acid 0.20 -. ~ , , _.._-_ -L "L, 3 =6 E 0.15 c 0 =1 5 c 8 0.10 - _ , ,2/ lf’rd/l ,/"l/ in I 0.05 - ————— 7- / ’14- 7 47 if // I . I . ,-/ 0.00 I r i a . . T . . 0 50 100 150 200 250 300 350 400 Time (min) Figure 4. Hydrogenation of 0.25 M glycolic acid in the absence and presence of 0.3 M H3PO4 (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) It is known from previous research in our group that the catalyst surface interacts with the solvent, substrate, and presumably the acid.25’59‘63 Thus, the presence of additional acid in the case of glycolic acid may not have a significant effect due to a high affinity of the glycolic acid for the catalyst surface, but this may not be true for every substrate and all conditions. The effect of H3PO4 on hydrogenation and H/D exchange will be discussed further in section 4.4. 4.1.2 H/D Exchange Studies Several control experiments were conducted in the absence of the 5% Ru/C catalyst. An experiment with 0.25 M glycine in D20 at 130 °C and 1000 psi H2 pressure showed no deuterium incorporation after five hours of reaction. The 1H NMR specifics for samples 102 taken at time zero and after five hours are shown below (Figure 5). Based on the 1H NMR integrations using acetic acid as the internal standard, the concentration, at time zero and after five hours, was essentially unchanged. A very small, but visible amount of deuterium incorporation was observed, partly due to a small amount deprotonation at the a-carbon. 9000 _-._-__ ._ -___-__. .---__- ., _ , .- 8000 —I—Time: 0 Hours + Time: 5 Hours 8000 .. -—- i # 4 — a , — - - A f e , 7000 7000 -_ 4 4 e — - - - a» , , ~ 6000 6000 ~»— — - - —- — - - -~ - e 5000 5000 .. — i i A A A” - — --------- e - . 4000 4000 a» — — — — —— --— l —— — —— -— .. 3000 3000 4» e i - A 7 ~— ' e -F V» -~ 2000 2000 --————- I 1 -~ - - 1000 J k . 0 V -1000 -1000 ----—----——-—— —- ---__-__-- -2000 3.42 3.40 3.38 3.36 3.34 3.32 3.30 Figure 5. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D2O, 130 °C, 1000 psi H2, no catalyst) An experiment with 0.25 M sarcosine in D2O at 130 °C and 1000 psi H2 pressure was allowed to react for six hours in the absence of the 5% Ru/C catalyst. About thirty percent of the methylene was partially deuterated during the reaction. The 1H NMR of time zero and after six hours is shown in Figure 6. No observable deuterium incorporation was obtained at the N-methyl position on sarcosine after six hours. The 1H NMR of time zero and after six hours is shown in Figure 7. 103 2250 -- »~--——-~—~~-~--~-~-—~~- ----_____--_____-- i 2000 +Time: 0 Hours + Time: 6 Hours 2000 ~ 1750 - 1500 3 1250 ~ 1000 ~ 750 * 500 ‘ 250 3.47 3.45 3.43 3.41 3.39 3.37 Figure 6. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 130 °C, 1000 psi H2, no catalyst) 3500 _..-.. _..__..A_“_-.__-_- .. --------.-.-- - - W - - -..._ .. _ W.---.-- - ...----_n_,_ 3000 +Time: 0 Hours a» Time: 6 Hours 3000 -—- -—— -— — —-#—- -#---~— i ~#r ~ — -—-——#--- - ~ 2500 2500 «- ___,--- - # -# - ## ##n --- ,__ - 2000 2000 --—- __-,- ##w .# -## . __-_,, - # 1500 1500 ---—— --— —~- # -# —---— - ##- # 1000 1000 ~- — # #- # ,k-- #- ‘ ' -—-- ~ # - - ~ ~ 500 500 0 0 - -~ «- -- -500 -500 -1000 2.58 2.57 2.56 2.55 2.54 2.53 1 2.52 2.51 2.50 Figure 7. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (D2O, 130 °C, 1000 psi H2, no catalyst) 104 The data were fit to the kinetic models discussed in chapter 3 (Figure 8). The H20 and HOD experimental (symbols) and fitted (line) data refers to the percent H in water and is on the right y-axis (secondary y-axis). The experimental (symbol) and fitted (line) concentration of CH2 (referred to as AH2), CHD (referred to as AHD), and CD2 (referred to as AD2) are on the left y-axis (primary y-axis). The rate constants for the hydrogens on the methylene carbon were calculated using the IO-variable model. The forward rate constant for the first hydrogen exchanged with deuterium was found to be k1A=9.8x10'6 L/mol/min, but the forward rate constant for exchanging the second hydrogen with deuterium was found to be k2A=O (both reverse rate constants were zero k- 1 A=O, k-2A=O). All rate constants have the units L/mol/min. There is an appreciable amount of H/D exchange at the methylene position (Figure 8). Results from the H/D exchange studies (to be discussed later) have shown that sarcosine is the most reactive substrate (as compared to glycine and N,N-dimethylglycine). So, the H/D exchange observed in this control experiment (no catalyst) may be partly due to some residual catalyst being present in the reactor lines or reactor wall. However, the majority of the H/D exchange is presumably due to the non-catalytic process of deprotonation of the weakly acidic zwitterion. The compound is most stable in the zwitterionic form. But, again at 130 °C there is a larger population of the higher energy form that can undergo deprotonation. 105 030 -_.-__.._._.___._..-._-.___-._..._-.-...._...__--_.-_.... ...__..-..._.__-_._.-_.-__._-..___-_-_-.._W- ..,,......_..._.-......_...._-_.._...----.-..-.. “WWW-T 42 ——AH2 ——AHD —ADZ I AH2_exp A AHD_exp o A02_exp —H20 and HOD 0 H20 and HOD_exp 0.25 —* ~— - - *—~ ,. — i 4- 3.5 g . + 2.8 .E. x a o c g - + 2.1 g 0 g o. 5 0 0.10m - — ~ — — —— — - 7 ~~1.4 A A /l/ A ./’”"’ 0.05 « — / -. 0.7 //,.W 0.00 —/. +1 . . c . c . .s o O 50 100 150 200 250 300 350 400 TImo(mln) Figure 8. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 130 °C, 1000 psi H2, no catalyst) A control experiment with 0.25 M choline bicarbonate in D20 at 130 °C and 1000 psi H2 was allowed to react for four hours in the absence of 5% Ru/C catalyst. Again, there was no observable deuterium incorporation at either of the two methylenes (3.88 ppm next to -OH and 3.32 ppm next to -N+(CH3)3) or on the nitrogen methyl groups (3.02 ppm) ' (Scheme 29). This is an expected result and differs from the case of the amino acids in that there is not an activating carboxylate group to facilitate non-catalytic H/D exchange. l/ 0 Ho/\/ \ HO 0' Scheme 29. Structure of choline bicarbonate 106 The 1H NMR taken at time zero and after four hours for the more downfield methylene position (3.88 ppm) corresponding to the methylene next to the -OH is shown below (Figure 9). The 1H NMR of time zero and after four hours for the more upfield methylene position (3.32 ppm) corresponding to the methylene next to the -N+(CH3)3 is shown in Figure 10. An experiment with choline bicarbonate was performed with 1 g dry 5% Ru/C (D20, 130 °C, 1000 psi H2). Hydrogen exchanged with deuterium at all three positions on choline bicarbonate, however, the 1H NMR peak data was too convoluted to distinguish between the CH2 and CHD for the two methylene positions and the CH3, CH2D, and CHD2 for the three N-methyl groups. So, the overall integrals were compared. From the results, the methylene hydrogens adjacent to the -OH group exchanged the most quickly reaching a total conversion of 58% after six hours (sum of CH2 and CHD). The methylene hydrogens adjacent to the -N+(CH3)3 and the N-methyl hydrogens exchanged at nearly an identical rate reaching a total conversion of 35% each after six hours. These results confirm that H/D exchange occurs faster at positions adjacent to an oxygen as opposed to a nitrogen. Further evidence for this will be discussed in the H/D exchange of ethanolamine. 107 —l—- Time: 0 Hours + Time: 4 Hours 150 #7 7 100 -» 7 3.96 3.94 3.92 3.90 3.88 3.86 3.84 3.82 3.80 Figure 9. 1H NMR of H/D exchange at the methylene position next to the —OH of 0.25 M choline bicarbonate (D20, 130 °C, 1000 psi H2, no catalyst) 250 2.00 -I— Time: 0 Hours + Time: 4 Hours 200 -— 7 . . . » 150 150 .- 100 100 ~~ * 7" - _ 5o 50 w y 0 0 WWW L—» -50 -50 -100 3 40 3 37 3.34 3 31 3 28 3 25 Figure 10. 1H NMR of H/D exchange at the methylene position next to the —N+(CH3)3 of 0.25 M choline bicarbonate (D20, 130 °C, 1000 psi H2, no catalyst) 108 A control experiment with 0.25 M betaine ((CH3)3N+CH2COO') at 130 °C with 1000 psi H2 in the absence of catalyst was compared with the identical experiment in the presence of l g dry 5% Ru/C. Figure 11 depicts the 1H NMR for the methylene position at both time zero and after five hours for the experiment conducted in the absence of catalyst. The H/D exchange reaction occurs at a comparable rate in both the presence and absence of catalyst. Figure 12 shows the 1H NMR for the methylene position at both time zero and after five hours for the experiment conducted in the presence of catalyst. In the experiment with catalyst, after only three hours nearly 50% of the CH2 is partially or fully deuterated. After six hours, 86% of the CH2 is partially or fully deuterated. In contrast, the CH3 peaks show no evidence of H/D exchange (in either the absence or presence of catalyst). The 1H NMR of the nitrogen methyls after five hours of reaction time are presented in Figure 13 (no catalyst). The integration area and peak shape is unchanged throughout the course of the reaction. The lH NMR at time zero and after five hours for the methyls on nitrogen for the experiment with betaine in the presence of l g dry 5% Ru/C are shown in Figure 14. Again, there is no HfD exchange on the nitrogen methyls even in the case of catalyst. The peak shape is the same for time zero and after five hours. 109 1000 800 +Time: 0 Hours +Time: 5 Hours 800 ~ - 600 600 <~7 , 7 ~ 400 400 -» - 7 . 200 200 O 0 -200 3.80 3.78 3.66 3.64 Figure 1 l. 1H NMR of H/D exchange at the methylene position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, no catalyst) 900 ._.__._____ 700 '— 120l ,. Time: 0 Hours .{- Time: 5 Hours 800 ~ .. .. 7 , . - . ~ » 600 700 . O ~ 500 600 - » 400 500 7777777~ 300 400 i~ , ~ 200 300 - ,W- 7 W — 100 200 0 100 - - -100 0 -200 3.77 . . . . . . 3.63 3.61 Figure 12. 1H NMR of H/D exchange at the methylene position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, lg dry 5% Ru/C) 110 4000 ~~~~7 7~-77—---~-~-- 3500 +Time: 0 Hours ~I— Time: 5 Hours 3500 77#-- ---#7 # #7 # # # — #7- #7 #7 #7 # # #7 3000 3000 77— #- 7- ——--# #7 77 # 7# —- l ~# # # 77# # -# 7 2500 2500 77 #- 7777777 7 # 77# 7# I 7# #7 7# #7 # # 7 2000 2000 77 -—# # # # 7 7 ~ — 7 1500 1500 77-——#— --# -— #7 7# 7 # 77 7 1000 1000 77 #7 # # # # 500 500 o 0 ,_ - 7500 3.16 3.14 3.12 3.10 3.08 3.06 3.04 3.02 3.00 Figure 13. 1H NMR of H/D exchange at the methyl position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, no catalyst) + Time: 0 Hours 7:7— Time: 5 Hours 300077-7—#—# 77 --—# -- -#77 -—-——##7 . ##7## 77 # 7 72500 I I 2500 777- ## 7#7 #7 ##7-7 ' --7777— 77 ##7##7- 7 2000 200077 # # #7 # # : 777## 7 7 --7-77771500 u I 1500 #- # ## ##7 #77# = — #### 7777 71000 ' 1000 7 #-#~—-7 -— #7 # 7 . ——-- 500 s; 500 7 3~ o o 7500 3.20 3.15 3.10 3.05 3.00 2.95 2.90 Figure 14. 1H NMR of I-I/D exchange at the methyl position of 0.25 M betaine (D20, 130 °C, 1000 psi H2, 1g dry 5% Ru/C) 111 Figure 15 shows the concentrations of CH2, CHD, CD2 and the 3 CH3’s over time. The hydrogens on the methyls of the positively charged nitrogen do not exchange. Although the peaks areas not the same height, the peak areas after—six hours are about 6% less than the peak areas at time zero. This percentage loss is consistent across all experiments and may represent initial adsorption of the substrate into the catalyst. The rate constants for both experiments were calculated and using kinetic model A (Chapter 3.1.1). The forward rate constants for the experiment with no catalyst were: k; A = 3.5x10'5 and sz = 2.6x10'4 L/mol/min. The forward rate constants for the experiment with 1 g dry 5% Ru/C catalyst were: km = 4.1x10'5 and kg», = 1.6x10'2 L/mol/min. The rate of km is very similar in both reactions suggesting that betaine is undergoing H/D exchange at the same rate irrespective of whether catalyst is present. This suggests that betaine is undergoing a non-catalytic mechanism for H/D exchange. This is supported by the work of Richard et al. They evaluated the acidity of the or-proton of betaine at 25 °C in D20 and observed nearly 90% of exchange of the first or-proton via deprotonation by the hydroxide ion and formation of the intermediate enol. 112 +0112 —I—CHD +CH3 - t - - CH2_no cat - A ~ - CHD_no cat 7 A CDZ_no cat — t - - CH3_no cat ‘ 0.30 ~- - — - —# —77 — #77 777— _. 77-. -. . -, 7 0.35 “7' l 0.25 .0 N O i Concentration (moIIL) O a: 0.10 .._ — i — 2 7 0.0577 -— -- 7- ,777/5 " -- 7 , 77 -- g ' ‘Wfl7l" I ' T T 2 -_. ”“5 i 0.00 1 i 150 200 250 300 350 Tlmo(mln) Figure 15. Concentrations of CH2, CHD, and CD2 of 0.25 M betaine with and without 1 g dry 5% Ru/C catalyst (D20, 130 °C, 1000 psi Hz) 4.2 Substrate Adsorption on the Catalyst/Carbon Support 4.2.1 Adsorption on Ru/C An early observation in this research was the initial disappearance of substrate in the reaction vessel before reaction conditions were achieved. The concentration of the substrate was measured after preparing the solution on the bench (bench concentration). Again, the concentration was measured after it was exposed to the catalyst (reactor concentration). It was proposed that the substrate was being adsorbed (possibly irreversibly) onto the carbon support. When doubling the catalyst loading using refluxed lactic acid as the substrate, the initial concentration of substrate was reduced by more than half (Figure 16). This provides further evidence that the reaction is kinetically limited by catalyst sites and not by interphase mass transport. In addition, a control was 113 conducted with no H2 pressure at room temperature, knowing that the hydrogenation reaction would not be occurring at these conditions, and an initial conversion of about 3.5% was observed, yet no products were detected. The amount “lost” or converted after the bench concentration is exposed to the catalyst is fairly consistent across all experiments. 0.30 I Bench Concentration I Reador Concentration 0.28 -# # # 7 777 7 7 7- S3 M a) i Concentration (mol/L) _o o M M m A I l i 0.207# . ,, ,, 0.18 ~ 19 Dry Catalyst 29 Dry Catalyst 2 9 Dry Catalyst - Repeat Figure 16. 0.25 M Refluxed lactic acid concentrations measured (bench and reactor) It is not clear yet, whether all of the initial material adsorbed is desorbed and allowed to react or whether some amount of the substrate is permanently lost on the carbon support. In our research group, Peereboom et al. performed carbon adsorption experiments at room temperature (25 °C) using propylene glycol and glycerol (0.01 M to 2.0 M each) at room temperature and modeled the results using both F reundlich and Langmuir isotherms. Peereboom et al. observed that substrates adsorb on the catalyst and that propylene glycol has a greater preference for the carbon micropores over the bulk 114 solution than glycerol, due to its less hydrophilic nature. Using the Langmuir model, they found that the bulk glycerol concentration is not very different from the pore concentration. However, the pore concentration of propylene glycol is 5x larger than the glycerol concentration.59 In general, it appeared that after time the molecules were getting torn apart and did not desorb. In order to further follow up, a Ru-free sample of the carbon support used in our Ru/C catalyst from PMC was obtained and adsorption studies with different substrates were conducted. 4.2.2 Adsorption on Carbon Support Six substrates were evaluated to determine the amount of substrate adsorbed onto the carbon support. The carbon was 3310 lot #28850 from PMC. The percentage of water in the carbon support was 8.81%. Three-point calibration curves were nm for each substrate using the methods described in chapter 2. The six substrates studied were: lactic acid, glycolic acid, propylene glycol, ethylene glycol, phosphoric acid, and n- butanol. Normally, in the experiments discussed in this thesis the ratio of moles of substrate to grams of dry catalyst has been 0.025 mol/g dry catalyst. Thus, in this work 5 mL of a concentration of 0.25 M substrate was combined with 0.05 g of dry catalyst resulting in the same ratio as most experiments. The results indicated that 7% of lactic acid adsorbed onto the carbon support, more than any other substrate. Figure 17 presents all of the substrates (lactic acid = 7.1%, butanol = 5.7%, glycolic acid = 4.8%, propylene glycol = 3.4%, phosphoric acid = 1.0%, ethylene glycol = 0.8%). The error bars represent the measurement error as determined by the 115 relative standard deviation from the calibration curve for each substrate. The amount of lactic acid adsorbed on the carbon support is similar to the amount of lactic acid lost after being exposed to the 5% Ru/C catalyst. It can thus be assumed that most of the substrate lost is being adsorbed on the carbon support. 100% _ 90% 7# # #7 #7 # # 80% 7 #7 #77 # # # # #7 #- #7 77 7 # 70% 7 ##7 # 7 #7 ##7 77# 77# 7# #77 # 77 #7 7 # i 60% 7- #7 7# # ## #7 #7 7 # #- 7# #7 50%. WW W W, W., --W W- . W. ___W_ - 40% _ ____ ____ _W_ W_ .W W _ _W_ __ 30% 7# 77 # #77 #7 #7 #7 20% 7-—--#-— ## #77 7 -#-~-- —~ #77 # 7## 7##7 Percent Adsorbed on Carbon Support We) 10% 0% - i 7 Lactic Acid Butanol Glycolic Acid Propylene Phosphoric Ethylene Glycol Acid Glycol Figure 17. Substrate adsorption on the carbon support 4.3 Effect of HCI 4.3.1 Hydrogenation Studies A precedent for chloride inhibition has been observed when using a ruthenium catalyst. In our research group, Dalavoy et a1. performed electrocatalytic hydrogenation experiments using 11.1 mM lactic acid and a RVC electrode suffused with 5% Ru/C powder catalyst. The electrolytes examined were H2SO4, HCl, and HCIO4. At 90 °C in 0.01 M H2804 a yield of 9.8% propylene glycol was achieved after 9 hours and at the same conditions but with 0.01 M HCl instead, only 2.9% of propylene glycol was 116 achieved. The major product was lactaldehyde. Increasing the electrolyte to 0.1 M HCI yielded more propylene glycol after 48 hours (6.6%) than 0.1 M H2SO4 (2.7% yield of propylene glycol) (11.1 mM lactic acid, 100 mA, and 70 °C). We continued this work and performed a pair of experiments with 0.25 M betaine hydrochloride ((CH3)3N+CH2COOH-Cl') in the presence of 0.3 M H3PO4 at both 100 °C and 130 °C, 1000 psi H2, and 1 g dry 5% Ru/C. The results were compared to identical experiments using 0.25 M betaine ((CH3)3N+CH2COO' (chloride-free» at 100 °C, 130 °C, and 150 oC. None of the desired product, choline chloride, was formed from betaine hydrochloride hydrogenation. The effect of a slight excess of HCl may be to inhibit the hydrogenation reaction. Figure 18 shows the hydrogenation reaction of betaine hydrochloride. Results for chloride—free 0.25 M betaine ((CH3)3N+CH2COO') at the same conditions (130 °C) indicated 41% conversion and formation of 7% product, choline, after six hours. When the reaction was run at 150 °C the reaction was faster and yielded a conversion of 74% and 17% choline yield. As the temperature was increased, more conversion of the substrate was observed. At a reaction temperature of 100 oC there is minimal conversion and no product formation. The results for all three temperatures (100 0C, 130 °C, and 150 °C) are presented in Figure 19. 117 —I— Betaine'HCl 130C --I-- Choline Chloride 1300 k 7 Betaine‘HCl 1000 A Chioline Chloride 1000 0.25 I 7 # 7 # 7t A A A A ' £0.20 7 7# 7 7 # # # # # # e .E- S = 0.15 7 7# # # 7 7 .2. B 8 fl 8 0.10 - 7 # 7 # # # 0.05 7 # 7 # # # # # 7 7 0.00 7 = i = i = r : . ---------- [I ----------- . -l 0 50 1 00 1 50 200 250 300 350 400 Time (min) Figure 18. Hydrogenation of 0.25 M betaine hydrochloride (0.3 M H3PO4, H20, 100 °C and 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) 100% I 100C 121130C B1500 w#'w- 80%-7 70% 17 60% 7 50% 4 40%7 7 30% - 20%- 7 10%- - Choline Yield Betaine Conversion Figure 19. Hydrogenation of 0.25 M betaine after 6 hours (0.3 M H3PO4, H20, 100 °C, 130 °C, and 150 °C, 1000 psi H2, 1 g dry 5% Ru/C) 118 When comparing the experiments using betaine hydrochloride and betaine at the same conditions, it appears that the presence of the chloride is inhibiting the hydrogenation reaction presumably due to the competitive interaction of the chloride with the catalyst surface. There is an unknown peak that grew in at roughly the same rate as the choline. The unknown peak is most likely trimethyl amine. Due to the formation of the unknown peak in the experiment with 0.25 M betaine ((CH3)3N+CH2COO'), an additional experiment was run in the absence of H3PO4 (1 50 °C, 1000 psi H2, and 1 g dry 5% Ru/C) to evaluate whether the unknown peak was still formed. Results indicated no choline formation and no unknown peak formation after five hours. Part of the reason the unknown peak was not observed may be that if it is trimethylamine this good leaving group would be less stabilized in the absence of acid. No choline was formed because neutral betaine is in the zwitterionic form and hydrogenation of the carbonyl of the carboxylic acid will only occur if the carboxylic acid is protonated. An experiment with 0.25 M glycine was run at 130 °C, 1000 psi H2 in the presence of 0.3 M H3PO4 with 1 g dry 5% Ru/C in both the absence and presence of 0.3 M HCl. Addition of the HCl in proportion to the substrate was strongly inhibitory, cutting glycine conversion from 96% to 58% and ethanolamine yield from 80% to 25% after six hours. Figure 20 illustrates the results. In this experiment the amount of additional acid (0.3 M H3PO4 and 0.3 M HCl) is twice the amount (0.3 M H3PO4) used in our experiments with 0.25 M substrate and 1 g dry 5% Ru/C. Jere et al. examined the effect of excess H3PO4 119 and found a ~20% reduced alanine conversion (0.22 M alanine, 100 °C, 1000 psi H2, and l g dry 5% Ru/C) with 0.6 M H3PO4 (as compared to 0.3 M H3PO4).50 When further increasing the amount of H3PO4, the conversion continued to decrease slightly (730% reduced conversion with 1.2 M H3PO4 as compared to 0.3 M H3PO4). + G'Ycine Conversion I Ethanolamine Yield + Glycine Conversion - with H01 0 7 Ethanolamine Yield 7 with H01 100% 1—— 7 -—-7 777—77 —#7 #7 77—. _-. - W. -W- _W_ _ 80% ~ 60% ~ Conversion or Yleld (‘16) 40% 20%77 - 7, 0 50 100 1 50 200 250 300 350 400 Time (min) Figure 20. Hydrogenation of 0.25 M glycine in the absence and presence of 0.3 M HCl (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) An experiment was run with 0.125 M glycine at 130 °C with 1000 psi H2 and slight excess equivalents of both H3PO4 and NaCl (0.15 M each) and 0.5 g dry 5% Ru/C to determine if NaCl had the same effect on the hydrogenation reaction. The same volume of solution was used (100 mL) so the substrate concentration was more dilute (not allowing a direct comparison). It appears that the HCI exerted a larger effect than NaCl but a direct comparison using the same substrate concentrations and catalyst amounts should be done. Figure 21 displays the conversion of glycine in the hydrogenation 120 reaction with no chloride (0.25 M glycine, 1 g dry 5% Ru/C, and 0.3 M H3PO4), with 0.3 M HCI (0.25 M glycine, 1 g dry 5% Ru/C, and 0.3 M H3PO4), and 0.125 M NaCl (0.125 M glycine, 0.5 g dry 5% Ru/C, and 0.15 M H3PO4). In the experiment with HCl, additional inhibition may be occurring due to the fact that HCI is a very strong acid (pKa ~ -4) and may be changing the sample media enough to slow the hydrogenation reaction. —I— 0.25 M Glycine (0.3 M H3PO4. 1 g dry 5% RulC) 7&7 0.125 M Glycine with 0.15M NaCl (0.15 M H3PO4, 0.5 g dry 5% RulC) 7.7 0.25 M Glycine with 0.3 M HCI (0.3 M H3PO4. 1 g dry 5% RulC) 100% 80%1~ 60% + Conversion (7.) 40% 1 20% ~' 0% 0 50 100 1 50 200 250 300 350 400 11me (min) Figure 21. Hydrogenation of glycine alone and in the presence of NaCl or HCI (H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) In the hydrogenation experiments with amino acids, the role of the H3PO4 is to protonate the carboxylate, due to the preference of the amino acid to be in the zwitterionic form. The result is H2PO4'. In the experiment with both 0.3 M H3PO4 and 0.3 M HCl the HCI which is the stronger acid (pKa ~-4 as compared to ~2.15) is protonating the glycine carboxylate. Thus, H3PO4 is still in solution (as opposed to H2PO4'). The idea that perhaps it is the H3PO4 that is causing inhibition could be considered. However, in our 121 hydrogenation experiments with glycolic acid (discussed in section 4.4) we see no effect with the addition of an equimolar amount of H3PO4. From the experiments with betaine and glycine, it is evident that the presence of chloride slows or completely inhibits (as in the case of betaine hydrochloride) the hydrogenation reaction. Chloride inhibition is also observed in the H/D exchange studies with betaine hydrochloride (which will be discussed next). 4.3.2 H/D Exchange Studies As found in the hydrogenation experiments (Figure 18), betaine hydrochloride ((CH3)3N+CH2COOH-Cl') was also unreactive toward H/D exchange. There was no measurable amount of exchange of the methylene or methyl hydrogens with deuterium. An experiment with 0.25 M betaine hydrochloride in D20 was run at 100 °C with 1000 psi H2 and 1 g dry 5% Ru/C. The results were compared with an experiment using 0.25 M betaine ((CH3)3N+CH2COO') at 130 °C. The experiments with betaine hydrochloride showed no deuterium incorporation on the methylene or methyl carbon sites (Figure 22). This common inhibition by chloride could suggest that the hydrogenation and the H/D exchange catalyzing sites may be the same. But, we know from the experiment with betaine that the hydrogens on the methylene carbon undergo exchange with deuterium in the absence of catalyst due to the fixed cation promoting solution deprotonation (Figure l 1, 130 °C). So, if the chloride was deactivating the catalyst it should only affect the 122 hydrogenation reaction. The reason for the chloride in betaine chloride inhibiting the H/D exchange reaction is still unknown. 5000 T #7777~7~-77- 77777 4500 + Time: 0 Hours + Time: 6 Hours 4500 + -- _ 777 77 3500 4000 7 3500 7 #77 7 77 77 777 7 72500 3000 7 2500 7- 77 7777 ##7 777 7 77 777 7 7777 7 771500 2000 7 r 500 1500 77 1000 1* L ~ 7500 500 <- 0....1 4.2 4.1 4.0 3.9 3.8 3.7 3.6 3.5 3.4 3.3 3.2 3.1 Figure 22. lH NMR of H/D exchange at the methyl (3.2 ppm) and methylene position (4.1 ppm) on 0.25 M betaine hydrochloride (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) To try and push the H/D exchange reaction with betaine hydrochloride, an additional experiment was conducted at 130 °C and ran for 24 hours. During the course of the reaction, degradation of the betaine hydrochloride occurred yet still no observable H/D exchange. In the experiment with 0.25 M betaine at 130 °C with 1000 psi H2 and l g dry 5% Ru/C in D20 exchange of hydrogen with deuterium is occurring at the methylene position albeit relatively slow. Kinetic rate data were determined and using the kinetic model described in Chapter 3, the fit of the methylene data can be observed in Figure 23. The 123 H20 and HOD experimental (symbols) and fitted (line) data refers to the percent H in water and is on the right y-axis (secondary y-axis). The experimental (symbol) and fitted (line) concentration of CH2 (referred to as AH2), CHD (referred to as AHD), and CD2 (referred to as AD2) are on the left y-axis (primary y-axis). 0‘30 ,-W____-W..WW_-_ . -...W --- .._.. WW. -. . .. -,- W. -...-. --AH2 W--- _- .W, ..__ _. .. . ..., 18 AHD — A02 I AH2_exp A AHD_exp e ADZ_exp —— H20 and HOD 0 H20 and HOD_exp 0.25 ‘77 ~--#77 7 --# #7 ##7## # #7- ## ~- #7 #77 7 77 15 .. 0.20 # 12 i v I 8 E =3 0.15 77 9 o g s . m U c 8 0.10 i’ - W -. 6 0.05 77 i 3 0.00 1 T r r r T r 0 0 so 100 150 200 250 300 350 400 Time (min) Figure 23. Fit of the kinetic rate data for 0.25 M betaine (D20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) The rate constants for H/D exchange on the methylene carbon are: k. A = 4.1x10'5, k-l A = 0, k2A = 1.6x10'2, k-2A = 2.2x10'l L/mol/min; and the sum of the squares error (SSE) = 1.9x10'3 . At these low conversions, the back reaction rate constants are not likely to have much significance. A more detailed discussion considering the significance of the back reaction rate constants (km, k2A, k.1c, k-2c, and k.3c) will be discussed in 4.7 (as well as H/D exchange of betaine ((CH3)3N+CH2COO')). 124 4.4 Effect of H3PO§ 4.4.1 Hydrogenation Studies In the case of the amino acids, a slight excess amount of H3PO4 was required to protonate the carboxylic acid. Glycine and the N-methylated glycines are zwitterionic in nature and thus the carboxylic acid functionality is dissociated unless the additional acid is added. The H3PO4 is only required for the hydrogenation of the amino acids, but was added to all hydrogenation experiments in order to be consistent. To determine the effect of the H3PO4 on substrates that were not zwitterionic in nature, glycolic acid was studied. A pair of experiments with 0.25 M glycolic acid were conducted at 130 °C with 1000 psi H2 in both the absence and presence of H3PO4 (0.3 M). 100% , ' “‘1 + Conversuon + Yield [Conversion (with 0.3M H3PO4) .31 ie'dM-QQQMEE’QQ _ 90% ~— 80°/o - 70% -____-____ ___ g — 60% * 50% ~ 40% -~ Conversion or Yield 30% ~ 20% ~- 10% -- 0% 0 50 100 150 200 250 300 350 400 Tlmo (min) Figure 24. Hydrogenation of 0.25 M glycolic acid in the absence and. presence of 0.3 M H3PO4 (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) 125 The addition of H3PO4 did not significantly affect the hydrogenation of glycolic acid to ethylene glycol. The conversion and yield for the pair of experiments were essentially the same (Figure 24). A pair of experiments with 0.25 M glycine were conducted at 130 °C with 1000 psi H2 in both the absence (Figure 25) and presence (Figure 26) of H3PO4. In the presence of H3PO4, a conversion of 83% and a yield of 71% ethanolamine were obtained after four hours of reaction time. However, in the absence of H3PO4 much lower conversion and yield values, both only 15% were obtained (Figure 27), presumably due to the amino acid being in the zwitterionic form a majority of the time. Glycine 4" 240 min 180 min 120 min Ethanolamine 60 m!“ 30 mm l / 0 min i \2.‘ "‘N I -. .- . \ , , V. . -. _. 4, . .. 7 ._ . ... ..\.‘.. . J" ... 2;? . .A J: , .. . .. .1. {‘~.1»;~-~-- ;'*; ”_...—.--_._..,\g p , -. WI ,,,,,,, 4 \\;- —~ — a _ _ _ 2 3 4 5 6 minutes Figure 25. Hydrogenation of 0.25 M glycine (no H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) The H3PO4 is thought to protonate the carboxylic acid, allowing for the hydrogenation of glycine to form ethanolamine. In order to further determine whether or not the acid is competing with glycine for surface sites, H/D exchange studies can be done to see if the 126 presence of the additional acid slows or inhibits the H/D exchange. This is assuming that the hydrogenation and H/D exchange reactions are occurring on the same sites. The phosphoric acid is not necessary for H/D exchange and proved to be inhibitory (to be discussed next), but due to the need for the amino group to be free and electron donating. Glycine Ethanolamine 240 min 180 min 120 min 60 min 30 min 0 min 2 3 4 5 minutes Figure 26. Hydrogenation of 0.25 M glycine (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) In the experiment with H3PO4 present, glycine is essentially fully converted after three hours of reaction (as opposed to almost no reaction in the case of no H3PO4 added). The retention times in the HPLC analyses of glycine and ethanolamine are slightly shifted due to the H3PO4. It is assumed that the H3PO4 is adsorbing on the same sites as the glycine. So, competitive adsorption of the H3PO4 and glycine should be considered. Peereboom et al. did find that compounds that are more hydrophilic tend to stay more in the bulk solution and are less competitive for the catalyst due to the hydrogen bonding ability with 127 the solvent water.59 Competitive adsorption studies should be compared with the knowledge that H3PO4 was not significantly adsorbed on the Ru/C (Figure 17) and did not inhibit the hydrogenation of glycolic acid (Figure 24) to further determine whether the adsorption of H3PO4 on the catalyst should be included in the kinetic model. 1000/O -..._ _ _.- - . .. ..-..-__._.. ........-,_-.-_ __.--.-._ .. _ "._.-M..- _. . __ _...- .. ._.-._ _.- _... ..-_... , _ ._..- ...- . .- - . + Conversion (with 0.3M H3PO4) 90% w +Yield (with 0.3M H3PO4) + Conversion /. +Yield 70% . — 80% ~ 60% ~ 50% 4 40% «~ Conversion or Yleld 30% ‘- 20% r 10% « 0% 0 30 60 120 180 240 Time (mln) Figure 27. Hydrogenation of 0.25 M glycine in the absence and presence of 0.3 M H3PO4 (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) Another approach to evaluating the potential effect of H3PO4 on the hydrogenation reaction would be to perform an experiment with equimolar amounts of glycine, lactic acid, and a slight excess of H3PO4. There were several competition experiments performed by members in our research group.49’61’63 Results indicated that with equimolar amounts of alanine and lactic acid, the hydrogenation of alanine occurs but lactic acid is initially unreactive. One explanation for this may be that alanine binds stronger than lactic acid on the surface or has a greater affinity for the surface. But, a 128 second explanation may be that in the experiment with lactic acid and alanine, the lactic acid may serve to protonate the zwitterionic alanine resulting in deprotonation of the lactic acid thereby preventing the reduction. In the suggested experiment above, the H3PO4 would serve to protonate the alanine and allow a more direct comparison of alanine and lactic acid as well as evaluate the effect of H3PO4. 4.4.2 H/D Exchange Studies Several experiments were conducted in the presence of H3PO4 to try and determine if the additional acid caused an effect on the H/D exchange reaction. One goal was to try and establish a link between the hydrogenation and H/D exchange studies. A pair of experiments with glycine and sarcosine were run at 50 °C to try and slow or eliminate the competing hydrogenation reaction and isolate the effect of H3PO4 on H/D exchange. Glycine (0.25 M) with 0.3 M H3PO4 at 50 °C with 1000 psi D2 in D20 was allowed to react for six hours. Results indicated a negligible amount of H/D exchange. Figure 28 shows the 1H NMR at time zero and also at six hours. The yield of CHD and subsequently CD2 was zero. We can compare these results to the identical experiment in which no H3PO4 was added (Figure 29). 129 1250 -fi-_- - ~ 1000 +Time: 0 Hours -—I—— Time: 6 Hours 1000 .2- , ,. 7 ,, t. 7‘ ~ 750 750 «~ ~ 500 ~ 250 500 ~~ 250 0 .. -250 3.75 3.70 3.65 3.60 3.55 3.50 Figure 28. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (0.3 M H3PO4; D20; 50 °C; 1000 psi D2, 1 g dry 5% Ru/C) 4000 «~—-——-———~~ - e Wm ~ ~————— 3500 +Time: OHours ---l-Time: 6Hours 3500 ~~ ~ 3000 3000 ~ -~ 2500 2500 ~— ._ 2000 2000 ~ ,- 1500 1500 - ~1000 1000 _ 500 500 o - - -500 3.38 3.36 3.34 3.32 3.3 3.23 3.26 3.24 Figure 29. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D20; 50 °C; 1000 psi D2, 1 g dry 5% Ru/C) 130 The concentration data for the experiment with H3PO4 showed essentially no H/D exchange at the methylene carbon on glycine (Figure 30). The concentration of CD2 is not measured, but is the difference between the summed amount of AH2 and AHD at time zero minus the summed amount of CH2 and AHD at time t. Due, to some of the substrate being adsorbed on the surface ("lost") it may cause the calculated CD2 to be slightly higher. However, in this experiment the amount "lost" was equal to less than 3% (it is reflected as the CD2). The experimental percent H in water and the fitted line are also presented in Figure 30. I AH2_exp A AHD_exp O A02_exp —H20 and HOD 0 H20 and HOD_exp i I I I 0.25 — 77 7 7 77 7 —- I ———-7-- I . l - -- «7 2.5 g 0.20 — l» 2.0 o E ‘2' 2 a . - _ o g 0.15 15 § = I; § 8 0.10 - ° ° r A * 1-0 ’ o 0.05 -— _- 7 77 7 7 ~- 7 77777 + 0.5 ‘ A A A 0 000 e e . é . e T c c . A 00 0 50 100 150 200 250 300 350 400 Tlmo (mln) Figure 30. Concentration data for 0.25 M glycine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) The catalytic H/D exchange of the amino acids is promoted by having an electron-rich C- H bond available for insertion by the catalyst. When the amino group is positively charged (as in the case with H3PO4 added), the OH insertion by the catalyst is not as 131 likely to occur. This is also compounded by the fact that the experiment was run at such a low temperature (50 °C). A higher temperature reaction with sarcosine will be discussed later. The reason betaine (with a quartenary nitrogen) undergoes H/D exchange is because it is occurring non-catalytically through deprotonation of the or- carbon in solution. The same experiment but with 0.25 M sarcosine was performed (0.3 M H3PO4, D20, 50 °C, 1000 psi D2). Results again, as in the case of glycine, indicated a negligible amount of H/D exchange with H3PO4 added. Figure 31 shows the 1H NMR at time zero and also at six hours. The yield of CHD and subsequently of CD; was zero. There was a small amount of conversion of the CH2 peak area (2.4%), but is presumably due to degradation of the substrate over six hours and experimental error. When calculating the amount "lost" from the difference between the time zero and at six hours for the methyl concentration, the percent lost was zero. We can compare this reaction to the experiment (same conditions) in which no H3PO4 was added (Figure 32). The concentration data for the experiment with H3PO4 showed essentially no H/D exchange at the methylene carbon on sarcosine (Figure 33). The experimental percent H in water and the fitted line are also presented in Figure 33. 132 1000 -r__-_~__-_.- _ -...--.-..- -_ ... _..-. .---... _ .. -_.-. . ..-. - -....-.. -..... - .. ... ..--.--- .........-----..... .. _- ...- - _....-..._---_-. _ ._ -.. . _- _ _ . ._.-.k ... .... - .-.- . .. _ 900 + Time: 0 Hours 7'7 Time: 6 Hours 77—7- 77 7747 800 900 77- 800~ -- 700 '1 600 500 i 7 - 400 . 300 ~ 200 100 3.76 3.74 3.72 3.70 3.68 3.66 3.64 3.62 3.60 3.58 Figure 31. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (0.3 M H3PO4, D20, 50 °C, 1000 pSl D2, 1 g dry 5% RU/C) 2200 1800 "' _ +Time: OHours 47 Time: 6Hours 2000 H - 7+ - 7 1600 1800 J T" 77 777 1400 1600 17' i — 77777 —1200 1400 -- I * 7—7777 "r 1000 [3.45 3.43 3.41 3.39 I 0 1200 -_—_.-._.-.--__ .__— _ — _ 0 ._ _ _ -_____ __ 800 0 ' 1000 777—77—— - 777——— - x 7777 7 777777 - 7 600 0 0 800 4k 7—7- -~ —--~---— -7 ”i' -i 7777 7 Wfi 2. 400 0 600 - 200 400 0 200 ‘ ~ -200 400 o - 3.52 3.50 3.48 3.46 3.44 3.42 3.40 3.38 3.36 3.34 Figure 32. lH NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) 133 It is important to note that the amount "lost" after six hours with respect to time zero is small. This was confirmed in all experiments in which the desired reaction was not occurring. The catalyst and carbon adsorption studies indicated that up to 5-7% can be "lost" once the substrate is exposed to the catalyst. But, presumably this occurs immediately (at time zero) and does not increase with time. For the kinetic studies as well as conversion, yield, and selectivity the time zero concentration was used. It was assumed that the initial amount of substrate adsorbed on the carbon support does not eventually react but is degraded. As discussed above, the ability for the amino acids to undergo H/D exchange is affected by protonation of the amino group. Thus, it would be of value to do a detailed pH study to measure the extent of protonation with increasing concentrations of H3PO4. Slight changes in pH could drastically change the H/D exchange reactivity. In addition, it could be suggested to explore the effect of added base on the H/D exchange reaction to evaluate whether it would increase reactivity. We can also look at the rate of exchange on the methyl carbon (Figure 34). Again, there is essentially no I-I/D exchange occurring at the methyl position at these reaction conditions. 134 I AH2_exp A AHD_exp 9 AD2_exp --H20 and H00 0 H20 and HOD_exp 0.25 I7 .7 7. 7-7 77.7 7- -7 7777- 777 7 I 7 I 7 7 2.5 t g 0.20 7 7 77 7 7 7 7 77 77 2.0 E Ev J: E 30157 .- 77771—7 77.7 77 77- — - +150 g d 5 o 0.10 7 77 77 7 7 77 7 77 7 +1.0 0.05 77 — — 77 7 7 7 —7 7 77 77 0.5 000 {H r . r g I 3 T _e r z r z 00 0 50 100 150 200 250 300 350 400 Time (min) Figure 33. Concentration data for 0.25 M sarcosine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) 1 600 +Time: 0 Hours 7&— Time: 6 Hours 1400 7777—7 777- 1200 777 -7 7- "rioe '\ x 1000 777-77-7777 - 77 800 77 600 7 400 7777-- 200 o .-m_ .. 2.68 2.64 2.60 2.52 - 1400 7 1200 1000 r 800 600 7 7 400 77 200 77 7 -200 2.48 Figure 34. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (0.3 M H3PO4, D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) 135 Sarcosine is more reactive than glycine in the H/D exchange reactions, so a higher temperature, 70 °C, was also evaluated. It was desired to have some H/D exchange occurring so that the rate constants could be compared (absence and presence of H3PO4). In addition, higher temperatures promote dissociation and allow for more of the higher energy form (non-zwitterionic) to be present. Figure 35 shows the lH NMR for the methylene carbon at time zero and after six hours for 0.25 M sarcosine with 0.3 M H3PO4 at 70 °C with 1000 psi D2 in D20. A small amount of H/D exchange is occurring. H/D exchange is also observed at the methyl position. Figure 36 shows the 1H NMR for the methyl carbon of sarcosine at time zero and after six hours. The kinetics were determined using both kinetic models and the reaction data fit is shown below in Figure 37 and 38 (methylene and methyl). Again, by definition the rate constant klA should be twice that of km. The model defines the rate constants on a per hydrogen basis and in the case of converting CH2 to CHD there are two available hydrogens for exchange whereas in the conversion of CHD to CD2 there is only one. The same theory is applied to the methyls in that the rate of km should be 1.33 times that of kzc and 1.67 times that of k3c. This assumes there are no isotope effects (primary or secondary). This assumption as well as a detailed discussion on the rate constants will be included in section 4.7. 136 1800 7 +Time: OHours I Time: 6 Hours 1600777 777 7 77 7 77 7 1400 77 1200 77 777 77 777 10007777 77 7 1 800 77 777 77 77 7 777 -7 ...IL.‘_ ._A 600 +- 7 7 "T 1600 7 71400 7 ~ 1200 7 ~ 7-71000 77 — 800 7 600 7 400 400 — , 200 0 3.68 3.60 3.66 3.64 3.62 Figure 35. 1H NMR of H/D exchange at the methylene position of 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 3500 - —-—-7~ +Time: 0 Hours 7I-Time: 6 Hours 300077- 7 7—--——7-- 7 7-7 7 77 2500-7 7 7 3.58 2000 7 77 - 7 7 - 77 1500 77 7 77 7 1000777 7777 77 7-7...-- --___H 500 2.52 o . .... mun-._...." . . 2.60 2.58 2.56 2.54 Figure 36. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 137 2.50 777 200 -200 3.56 P--. 7.777 3000 7 7 2500 —- —- 72000 77 1500 7 7* 1000 77777 - 7500 7 -500 2.48 —— AH2 —— AHD — AD2 I AH2_exp A AHD_exp o ADZ_exp —H20 and HOD 0 H20 and HOD_exp 0.25 7. -—~~— — 7 - -— 77' --—-—— - v —* 7 —- - -— 2.0 9 N o [ i b 3 , , . 3 E v :r: S ‘a 3015 7 77 7 —- f \r *1-2 8 ir~7_\ . é 3 c 8 0.10 — —- — - - /1:’“/_,3_..77—1 .. 08 M” ‘ 0,05 7 77 7/ 7 7 7 7 7 7. 7777 7 7 7 777 0,4 “/A/ M/ 0.00 v I w ’ v r v . r 0.0 0 50 100 150 200 250 300 350 400 Time (min) Figure 37. Fit of the kinetic rate data for 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 0'30 .. _-....- -a- --_--.....-.-.--.__.._-...........-_ 3.0 ~-—-- CH3 W CHZD -— CHDZ m- CD3 I CH3_exp A CHZD_exp O CHDZ_exp O CD3_exp —H20 and H00 0 H20 and HOD_exp 0.25 <~———i- -- ~ _ > - , «- 2.5 g 0.20 77 2.0 8. V I C o-I 0 c E 015 * ~— 1 5 g ‘E E 8 8 o 0.107 77 7-- 7 — 7—7— 7 77 77 771.0 0.05 . + 0.5 4 0.00 7 . 0.0 0 50 100 150 200 250 300 350 400 Tlmo (min) Figure 38. Fit of the kinetic rate data for 0.25 M sarcosine (0.3 M H3PO4, D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 138 The rates for H/D exchange on the methylene carbon are: k1 A = 0.21x10‘4, k-1 A = 9.3x104, km = 0.18x10‘4, and k-ZA = 1.9x10'3 L/mol/min. The reaction rates for H/D exchange on the methyl carbon are klc = 0.18x10'4, k-1c = 1.1x10'3, kzc = 0.30x10'4, k-zc = 1.2x10'2 k3c = 0.31x10'4, and k-3c = 0 L/mol/min. We can compare the rate constant data for the experiment at 70 °C with 1000 psi D2 in D20 in both the absence and presence of 0.3 M H3PO4 (Figure 39). Again at these low conversions, the back reaction rate constants do not have much significance. In addition, these experiments were conducted with 1000 psi D2 and D20, so the available H in the water (which is in equilibrium with the surface) is only ~2%. Using a purely statistical approach and assuming the reaction has reached equilibrium the percent of CH2 would be (0.02 x 0.02) 0.04%, CHD would be (0.98 x 0.02 + 0.02 x 0.98) 3.92%, and CD2 would be (0.98 x 0.98) 96.04%. The large amount of deuterium available promotes the forward reactions to occur. The rate constant results show that the rate of H/D exchange is significantly suppressed in the presence of H3PO4 (Figure 39) Again, this is presumably due to the amino group being protonated and not in the free form. The ratios for the forward rate constants for both the methylene and methyl carbon are not equal to the expected ratios (as discussed previously), indicating that isotope effects may be playing a role in the H/D exchange reaction. Kinetic model B (chapter 3) incorporates the possibility of primary and secondary isotope effects and will be explored in section 4.7. 139 2.5E-04 7 ..._- ”MN“- “a-..” -_W, Iwithout 0.3 M H3PO4 lwith 0.3MH3PO4 2.05704 7 7 77 ---. __ _ ._ A i ’5‘ E 3 5 1.55704 «77 i , .3. 5 a 5 1.05-04 7 0 3 I n: 5.05705 77 o.oe+oo 7 k3_C Figure 39. Rate constants for 0.25 M sarcosine in the absence and presence of 0.3 M H3PO4 (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) An experiment was conducted with 0.25 M betaine ((CH3)3N+CH2COO') and 0.3 M H3PO4 at 100 °C with 1000 psi H2 in D20. Results indicated that even with the H3PO4 present, H/D exchange is still occurring at the methylene carbon. The 1H NMR of the methylene position on betaine is compared at time zero and at six hours of reaction time (Figure 40). The peak shape of the CHD (triplet) from the 1H NMR for betaine was not as resolved as compared to most of the other experiments. In Figure 42 the inset shows the CHD peak after six hours. At six hours of reaction, 100% of the CH2 had been converted. 140 300 7777777 7 . 250 —+7 Time: 6 Hours 250 1 ~ 200 200 3 - 150 150 1 ~ 100 100 ~ ~ 50 50‘ 0 = 7 :‘r . j -50 4.04 4.00 3.96 3.92 3.88 3.84 3.80 Figure 40. 1H NMR of H/D exchange at the methylene position of 0.25 M betaine (0.3 M H3PO4, D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 1500 _... _.... _-....- _-... “7“.-. -_..-- .7. . . .. 7.. ----.-m. _-.--~---....__ -.- _--.---“ .-...___.._..._.._. .--....--_.___._ .__..__ l" 1250 1250 —~ ~ 1000 1000 ~ ~ 750 750 7 - L 500 500 1~ ~ 250 250 o o . 77 77 77 7250 3.22 3.20 3.18 3.16 3.14 3.12 3.10 3.08 Figure 41. 'H NMR of H/D exchange at the methyl position of 0.25 M betaine (0.3 M H3PO4, D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 141 This is not surprising considering that betaine exhibits H/D exchange at the methylene position even in the absence of catalyst. Thus, if the H3PO4 is competing for the catalyst it still should not inhibit the H/D exchange in betaine, as the exchange does not involve the catalyst. Consistent with this notion, H/D exchange was not observed at the N—methyl position of betaine. The 1H NMR peak of the three methyls decreased in size over time due to the conversion of betaine to choline, but no H/D exchange occurred (Figure 41). In this experiment the presence of choline was also detected over time. From HPLC analyses, a 38% yield of choline was obtained after 6 hours of reaction time. A conversion of 100% of the methylene carbon was measured after 6 hours, due to both hydrogenation and H/D exchange. Using the N-methyls as a reference for the amount of hydrogenation that occurred, roughly 55% of the N—methyl 1H NMR peak was converted to choline. The HPLC analysis matched this, reporting 49% conversion after 6 hours. For this experiment, the percent H in water was much variable. So, the [HOD] fit parameters were modified to better represent the actual data. A 4th order polynomial was used to fit the aqueous hydrogen concentration. For all other experiments, equation 46 (Section 3.2.2) was used. The concentration data versus time and the fitted rate equations are presented in Figure 42. Betaine is undergoing uncatalyzed H/D exchange that does not appear to be inhibited by the H3PO4. 142 0.30 0.25 ~ ~—-— AH2 I AH2_exp _:'_H20 “(1-590 Concentration (mol/L) .0 a: .0 _L o 0.05 0.00 77 AHD A AHD_exp __°_._t'20 amLHQD—eXP Figure 42. Fit of the kinetic rate data for 0.25 M betaine (0.3 M H3PO4, D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) The experiment with 0.25 M betaine in the absence of H3PO4 at 130 °C with 1000 psi H2 50 100 150 200 250 Time (mln) and D20 will be discussed further in section 4.7. 4.5 Solvent Effects on the Hydrogenation of Organic Acids A key aspect involved in understanding the mechanisms controlling the catalytic hydrogenation of organic acids is the effect the solvent has on the hydrogenation. In catalytic hydrogenation of acids, it has been determined that catalyst and hydrogen are essential in order to convert the starting material. In addition it appears that the acids need to be in their protonated form. It is less clear what role, if any, the solvent plays in the overall catalytic hydrogenation scheme. 143 300 —36 ~~ 30 7- 24 7713 7— 12 400 Percent H There are several advantages of using aqueous-phase conversion routes including: 1) most bio-based feedstocks are produced and handled in water solutions and 2) feed purity is less critical. However, the question of whether water is involved in the hydrogenation mechanism has not been answered. It has also been suggested that a more "organic" environment might favor conversion of the more polar amino and hydroxyl acids to their less polar alcohol reduction products. To examine this issue, an experiment was conducted in which no water was present. Glycolic acid (0.25 M) in dry tetrahydrofuran (THF) at 130 °C with 1000 psi H2 revealed a suppressed hydrogenation reaction rate. The yield of ethylene glycol formed after six hours was about 14% compared with nearly 50% with water as the solvent. Though THF appeared to hinder the reaction some ethylene glycol was formed, indicating that the reaction did still occur. Additional experiments were conducted using ratios of 80/20 (v/v) water/THF and 20/ 80 (v/v) water/T HF . From this series, it is clear that as the amount of organic solvent is increased the reaction is further suppressed (Figure 43). To promote hydrogen solubility in water, high pressures are used (1000 psi H2). In the case of THF as the solvent, the solubility of hydrogen was examined to determine if less hydrogen was present in these reactions. 144 1000b 2.2.2-222 .-......__-... -..... -.. ”2-- . .. .... _-..“..-2... 2.... “-2... -... -_ _...- -..-2.-.“ . .. .. . -_......__ __ .. . .. _... . ,, ,_ , _ . . .. .. 2 ._ . . .. .. ., + Conversion of glycolic acid ~+» Yield of ethylene glycol 90% i 80%1----—-~77 777 7 7 7777 7- 7 77 7 77 094.777 777 7 7 7 7h 7 if a; 7 60%\—— 77 — 77777 7 77 --77 7 5 0% f: \777 -7- 7, fl "7’ 7 2-.-, ,,_ 7 7 fl \\ \ \\ 30% ‘ "** “W 7 7:1- -2 4f; ,__ _. J; -- '5'“"f“"""‘~\~-77777._: 20% 2. ,7 ,___ "7, , ”A7, - , . - W _ T. 7... 72,, 10%77 77 77 7 -- 7 7 77:: 0% 0 10 20 30 40 50 60 70 80 90 100 “/o THF In Water Figure 43. Hydrogenation of 0.25 M glycolic acid after 6 hours (THF/H2O, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) In contemplating the role of the solvent, the various reactions in which it can participate need to be considered. From previous work in our group it was known that catalytic H/D exchange is possible between solvents and both the organic acids and the alcohol products. To understand whether hydrogenation and H/D exchange are occurring at the same sites on the surface THF was chosen as a solvent. An experiment with 0.25 M THF at 130 °C with 1000 psi H2 in D20 and l g dry 5% Ru/C was run and results indicated that THF readily undergoes H/D exchange at these conditions. The H/D exchange was observed at both C1 (83% a-CH2 conversion) and C2 (42% B-CH2 conversion) carbons. The lH NMRs for the methylene sites were difficult to convolute once the exchange started due to the overlap of the CH2 and CHD peaks. So, the total amount of conversion was based on the disappearance of both CH2 and CHD. If H/D 145 exchange and hydrogenation were occurring on distinct sites then the rate of hydrogenation should not be slowed as much in the presence of a solvent that predominantly undergoes H/D exchange. As mentioned above, the rate was significantly slowed as more THF was added, suggesting the sites may not be distinct and that there is competition at the surface. Due to the a C-H activation present with the THF solvent, a solvent was chosen that was expected to be unreactive to H/D exchange or hydrogenation, namely tert-butanol. The results from 0.25 M glycolic acid hydrogenation in tert-butanol suggested that a competitive reaction is occurring involving the solvent on the catalyst surface, thereby possibly tying up the available sites for acid hydrogenation or perhaps simply excluding the water needed by the reaction. tert-Butanol in the presence of acid (0.25 M glycolic acid) at 130 °C with 1000 psi H2 may also undergo selective dehydration to isobutylene.“65 lsobutene could undergo hydrogenation on the surface of the catalyst. No significant yield of ethylene glycol was detected. H/D exchange seemed unlikely in this context, but an experiment with neat tert-butanol and D2 and D20 was run to evaluate the amount of deuterium incorporation at the methyl groups of tert-butanol. After four hours, little H/D exchange was observed using lH NMR, eventhough there was some loss (~25%) of the CH3 as assessed by the 1H NMR (discussed further in section 4.7.2). From previous research in our group and the work presented in this thesis, we know that alcohols can undergo C-H activation. But, we also know that, in general, they have less of an affinity for the Ru/C catalyst as compared to the organic acid substrates. So, 146 ethanol and ethylene glycol were both explored as solvents for the hydrogenation reaction of glycolic acid. Hydrogenation experiments with glycolic acid (0.25 M) in ethanol were run in both the absence and presence of H3PO4. The presence of H3PO4 promoted esterification (9%) at t=0 min, defined as the point where the reaction temperature reaches 130 °C and the H2 pressure (1000 psi) is introduced. Figure 44 shows the HPLC chromatogram (mV versus time (min)) for glycolic acid in ethanol (with 0.3 M H3PO4) for the pre-reaction solution, time zero, and at 6 hours. In the absence of acid, the ester is not present until after 5 hours of reaction (Figure 45). However, there is some conversion of glycolic acid that is occurring over time. This may be due to the glycolic acid adsorbing into the catalyst. In either case, nearly the same amount of product, 8% of ethylene glycol, was formed. The conversion and yield over time is presented in Figure 46. This was an interesting result in that it also begs the question of whether the ester formed is able to undergo hydrogenation itself, albeit at a slower rate. 147 ii \ Glycolic \ acid . , Ethyl ti 8 2“ I I. l colate Ethylene t' rs. I, g y glycol Pre—reactlon Figure 44. Hydrogenation of 0.25 M glycolic acid (0.3 M H3PO4, CH3CH20H, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) Glycolic acid t= 6 hrs t= 0 hrs Ethyl Ethylene Pre-reaction glycolate glycol ii \i\ ._-_ -7 ----- , 777.--} . /‘\ _ -,__- i \\ * ' r ~-:--~~~~--- _...—._.J/ \777777— 5/ ¥n V Figure 45. Hydrogenation of 0.25 M glycolic acid (CH3CH20H, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) The literature suggests that esters undergo hydrogenation more rapidly than the corresponding acids with certain catalysts (Raney Ni, Copper/chromium oxide, rhenium 148 black). However, in the case of 100% ethanol there is little desired product formation. This result is not unexpected; early studies by Adkins (Raney Ni and Copper/chromium oxide) showed that esters hydrogenate poorly in the presence of free acids. Thus, the presence of both the organic acid and ester leads to competitive adsorption resulting in less desired product formation. If the ester formed does not undergo hydrogenation, less hydrogenation product is expected in the presence of acid due to more ester formed and depletion of free acid substrate available. However, these results indicate the same amount of ethylene glycol yield which suggests that some amount ester hydrogenation may be occurring (albeit small). In addition, in the presence of ethanol (and H3PO4) esterification is occurring faster than the hydrogenation of glycolic acid. 100% ~7 + Conversion of Glycolic Acid (with 0.3M H3PO4) 90% _ I ~ Yield of Ethylene Glycol (with 0.3M H3PO4) r-A— Conversion of Glycolic Acid 80% +__ + Yield of Ethylene Glycol 70% _.7 ___, ".7 —' - ‘ .... ' ' 'u7i734" " -:::’ 477:;_:_4 *7’ "" M,,7~k——————O“*"’ 4* 60% .. 7- . _./.V:__ . _ -7 777-7 .. . ,. 777-777 500A, _,_,_ _/__ _ ._ __ . _ __ ___..__._ _ _..-...7_ _ .. __.__ . . _ . . .. . ___ __.___ 400/0 "‘ / ““ _ ,7 ’ ‘ 77,77, ’ " " ‘ r 77__..‘ "— ‘‘‘‘ 30% .. 7 7 77 "—7777 7A — 7 7 20% WP — ~~7 — 77 7" 77 7 - 77 7 7 77 10% 7 _,...7A 77- 7 7 0 50 100 150 200 250 300 350 400 Time (min) Figure 46. Hydrogenation of 0.25 M glycolic acid in the absence and presence of 0.3 M H3PO4 (CH3CH20H, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) 149 An experiment with 0.25 M ethyl lactate in water was exposed to the hydrogenation reaction conditions (130 °C; 1000 psi H2). The ethyl lactate quickly hydrolyzed to form lactic acid. The lactic acid peak is overlapped by the ethyl lactate peak (Figure 47). The product propylene glycol was not observed until lactic acid was formed. However, the yield of propylene glycol at 6 hours was comparable to the yield using lactic acid as the substrate, nearly 72%. .1": 1’2- i l Propylene lllglycol l 1 “‘1 l. l i l Ethanol Lactic Et h l i l. in . '7’ l "~ / ac1d 1 ityt {It i, \\ /\~. r. acae 1' ix r/ x I; I ‘1 \\ a“. ‘-- [1" .. “M w ‘___' 7...-'7'7“77:7r714.17-2317: 1;": /7' 1‘7" m;.T.i.;;'.;;_;7‘,:;;.\;;.;._;7;i _-///\\ l¥uvi7lJiub t = 6 hrs "' .( t: 1 hrs 77 Pre-reaction T?" Figure 47. Hydrogenation of 0.25 M ethyl lactate (H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) This experiment suggests that the hydrolysis of ethyl lactate is faster than its hydrogenation, and that lactic acid hydrogenation is occurring with little hindrance from the ester or the product alcohol. In the case of glycolic acid in ethanol, the hydrogenation is suppressed due to the solvent (solvent/catalyst interactions) and it appears that hydrogenation of ethyl glycolate is slow. 150 The reduced amount of hydrogenation observed with the ester be due to the ester functionality or may be more of a result of the overwhelming presence of the alcohol as the solvent. In the case of ethyl lactate in water, ethanol is formed as the ethyl lactate is hydrogenated. However, the amount of alcohol present is small compared with the solvent water and thus in this case the alcohol is not inhibiting the hydrogenation reaction. An experiment that would help verify whether ester hydrogenation is possible is ethyl lactate in ethanol. However, as mentioned earlier, the solvent ethanol significantly suppresses the hydrogenation reaction and would presumably suppress it for an ester in the same way as it does for an acid. Our group has shown, using isotopic labeling experiments, that both glycerol and sorbitol exhibit rapid deuterium incorporation at C1 without reaction, suggesting reversible dehydrogenation on the metal surface as the first step in hydrogenolysis. This previous work sets a precedent for ethanol to undergo the same mechanism. Indeed, early qualitative studies by Kovacs et al. found deuterium incorporation at the 1-position of 1-propanol and 1,3-propanediol sites are very similar to the 1-position of ethanol.53 The dehydrogenation/hydrogenation of ethanol is a possible reaction competing with the glycolic acid hydrogenation for catalyst sites. Additional experiments using 0.25 M glycolic acid at 130 °C with 1000 psi H2 in 80/20 (v/v) water/ethanol and 20/80 (v/v) water/ethanol mixtures were conducted and revealed that the yield of desired glycol continued to decrease as the organic solvent proportion was increased. However, net acid conversion was relatively the same (Figure 48) due to 151 increasing esterification and less reduction product formation as the solvent’s ethanol fraction increased. The ethanol is known to undergo C-H activation and may be undergoing H/D exchange on the catalyst surface sites thereby reducing the amount of glycolic acid hydrogenation. 1000/0 “F” __-__....... ...-_...-..._..--- ._-- ...... . .. .. 7l— Conversion of glycolic acid 79— Yield of ethylene glycol 90% 777 7 - 7 7 77 7 7 800,6 ..7- . _77 - -7 7 .. 7777.. 77-7- 77 7 77 777 7.7 7 70% 77 7 - .77-- 7 7777 600/0 -\\ _ 7 7 7 \\ 50% K \ 7 7 — — é _ 7 I 40% ...17-77\'“——~“~77777:77 ._.-7-7-7’“ \ \ i 30% +7 777 77“ 7.7 7 7 777 7 7 20% ~ 7 77 - — — 7 ~ T-\~ "‘ “r 7- 7— 77 77 -—7 -—7 777:1 ‘ r ‘7 7- 10% — 7-— - ——-7-—7--- —-—“—‘-=-77-. 4‘ 0% 1 i 0 20 40 60 80 100 % Ethanol In Water Figure 48. Hydrogenation of 0.25 M glycolic acid after 6 hours (CH3CH20H/H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) Previous work in the J ackson/Miller group revealed that lactic acid had a higher surface affinity than propylene glycol.66 These results were deduced from H/D exchange studies in which the rate of exchange and chirality loss in propylene glycol decreased as lactic acid was introduced to the system. Using this logic, ethylene glycol was chosen as a solvent. If glycolic acid has a higher affinity for the catalyst than ethylene glycol then some hydrogenation product might be expected. Studies of 0.25 M glycolic acid in ethylene glycol (130 °C; 1000 psi H2) revealed that the dominant reaction occurring was 152 the esterification of the acid. The ester peak overlapped on the HPLC chromatogram with the glycolic acid peak and it was not possible to quantify the conversion of glycolic acid to the ester. Thus, the response factor was not determined for this suspected ester. Qualitatively, the amount of glycolic acid converted (based on manual integrations) and the amount of ester formed indicates that not much, if any hydrogenation product was formed. However, a small amount of ethanol was formed. In the presence of hydridic hydrogens on the Ru surface, the mechanism to form ethanol as a side product from the degradation of ethylene glycol may involve a hydrogen from the surface displacing a water molecule.67 To further examine the possible degradation of ethylene glycol a control experiment in which ethylene glycol alone is subjected to the reaction conditions should be run. However, we have seen ethanol formed after six hours in experiments with glycolic acid that resulted in high conversion (62%) yielding ethylene glycol (49%). Up to this point, it seems apparent that the solvent may be occupying the surface and undergoing either H/D exchange or hydrogenation/dehydrogenation thereby competitively preventing the hydrogenation of the organic acid. If an organic acid is chosen as the solvent, then one would hypothesize that the product alcohol from the solvent would be formed. Glycolic acid has a much higher rate of reaction than acetic acid, 49% yield compared with 5% yield after six hours (H20, 130 °C; 1000 psi H2, 1 g dry 5% Ru/C). This may be partly due to the difference in adsorption affinity as well as the activation due to the or-substituent. 153 From the present work, results with 0.25 M glycolic acid at 130 °C with 1000 psi H2 in acetic acid as the solvent indicate that neither ethylene glycol nor ethanol is produced. The acetic acid forms esters with the glycolic acid’s —OH side chain, and there is 0% yield of either alcohol products. If the solvent, acetic acid, were saturating the surface then some ethanol would be expected. However, the pKa of acetic acid is about 4.75 and it may be that the acid is reacting with the surface and causing deactivation of the catalyst due to the amount of acetic acid present in the system. Though less reactive than lactic or glycolic acid, propanoic acid is more reactive and slightly less acidic (pKa = 4.85) than acetic acid. After six hours of reaction, nearly 13% of propanol is formed from 0.25 M propanoic acid (H20, 130 °C; 1000 psi H2, 1 g dry 5% Ru/C). 0.25 M glycolic acid at 130 °C with 1000 psi H2 in propanoic acid revealed about 3% propanol and detectable ester. Ethylene glycol elutes at the same time as the solvent propanoic acid and therefore could not be quantified, although 70% of the glycolic acid was converted after six hours. Some of the conversion may have yielded ethylene glycol and the ester of propanoic acid. Again, because there was hardly any propanoic acid hydrogenation product formed it begs the question of what reaction is taking place to prevent the acid hydrogenation. Competitive adsorption is a key component in trying to understand these phenomena. Due to the fairly unreactive nature of the unsubstituted organic acids, it may be that the competitive reactions or esterification and self dimerization are enough to suppress the 154 hydrogenation of the propanoic acid. In order to further explore this, an experiment in which there is an equimolar amount of propanoic acid and glycolic acid in water was conducted. The hydrogenation of glycolic acid may be slightly suppressed if there is competitive adsorption with an equimolar amount of propanoic acid, but production of the alcohol products would be expected. In this case, the esterification and self dimerization are not expected to be dominant. Competition experiments will be discussed in section 4.8. To summarize the solvent studies, it is evident that if the solvent is able to undergo reactions on the catalyst surface, these reactions compete with the organic acid hydrogenation. Solvent H/D exchange and hydrogenation/dehydrogenation, and catalyst deactivation (whether temporary or permanent) can almost completely stop the organic acid hydrogenation reaction. Thus, the water is playing an important role in the overall catalytic hydrogenation of organic acids in that it is not competing with the reaction. The water may be playing an additional role of assisting H2 cleavage into hydrogen atoms on the surface. The potential role of water in the overall hydrogenation mechanism using 5% Ru/C will continue to be explored via detailed H/D exchange reactions and discussed later. Again, there are several advantages of aqueous phase processing. The nonvolatile organic acids can be converted without having to be esterified, as would be required for vapor-phase conversion. And energy costs associated with vaporizing and condensing process streams is reduced. 155 4.6 Organic Acid Hydrogenation Studies 4.6.1 Effect of the a-Substituent on the Hydrogenation Reaction One of the key findings from the previous research done in the Jackson/Miller research group was that simple carboxylic acids such as propanoic acid and acetic acid are substantially more difficult to reduce than lactic acid and L-alanine. Thus, the presence of the heteroatomic side chains modify reactivity. Specifically, protonated alanine and lactic acid undergo aqueous-phase hydrogenation over a carbon-supported ruthenium catalyst ca. 10 and 3-5 times as fast as their unsubstituted analogue, propanoic acid. The present work explores the substituent effects on organic acids' reactivity and selectivity toward hydrogenation by comparing reduction rates over a wide range of electron- withdrawing, hydrogen-bonding, or sterically bulky vicinal substituents. Several three—carbon and two-carbon organic acid substrates were chosen in order to evaluate the extent to which the oc-substituent affects the hydrogenation rate. Substrates included the following a-substituents: -H, -CH3, -OH, -N+H3, -OCH3, -N+H2CH3, - N+H(CH3)2, -N+(CH3)3CI', -N+(CH3)3 .7 The experiments were conducted at 130 °C with 1000 psi H2 using 1 gram of pre—reduced 5 wt% dry Ru/C with 0.25 M substrate and 0.3 M H3PO4 in water. Amino acids represent an attractive group of compounds for liquid- phase hydrogenation, both because of the diversity of substituents available and because of the recognized importance of optically active amino alcohol products as intermediates in pharmaceuticals. 156 Amino acids typically form dipolar ions in aqueous solution at their isoelectric point (pH ~5-7); the actual form is thus determined by the pH of the solution. The reason the H3PO4 was required (as discussed previously) for the successful hydrogenation of the amino acids was to protonate the zwitterionic carboxylic acid, thus enabling hydrogenation. In order to be consistent, the H3PO4 was added to all experiments. Three carbon organic acid substrates: 0 o 0 . OH propanoic acid isobutyric acid pivalic acid 0 0 0 OH OH HO OH OH NH2 OH lactic acid alanine glyceric acid Two carbon organic acid substrates: 0 o o o O\/“\ HO\/“\ HZN \i iOH / OH OH OH acetic acid methoxyacetic acid glycolic acid glycine O ’ O I O H \ +\/U\ /”\/lkOH /”\)]\OH /” O- N-methylglycine N,N-dimethylglycine N,N,N-trimethylglycine (betaine) 157 Glycine and the N-methylated glycine series were protonated due to the presence of H3PO4. If additional acid is not present, they are in zwitterionic form. In addition, the hydrogenation of both betaine with and without the chloride counterion was compared. A set of optimum reactor conditions was selected for this study based on previous work done by our research group.25‘48'49‘50 The results are presented in the graphs (Figures 49 and 50) and table (Table 2) below. Table 2 presents the overall data after six hours of reaction time and details the conversion based on the initial concentration; selectivity, in units of mole product per mole reactant converted, and yield, in units of mole of product per mole of reactant fed. Carbon balances were calculated to account for most products (>90%). Equation 62. Conversion of substrate in the hydrogenation reaction _[ reactant moles consumed J reactant moles fed Equation 63. Yield of desired product in the hydrogenation reaction _ [ product moles formed J reactant moles fed Equation 64. Selectivity of desired product in the hydrogenation reaction _{ product moles formed ] reactant moles consumed 158 100% . Alanme 90% 7 7 - 7 777—7 777 7 o 80% 7 —- 7—7—7 77777 7-~—7- - ~77 ’ Lactic Acid 70% . 7-7- 7 7 -- —7777 : Glyceric Acid . I A 60% 777—7- - - -7 7 7 -- 7- 7777—7377 -— o O a: 0 IE 50/0 4L*—— _____7 if 7 _. ”if " ' a i 7 >. 40% 7777 — --—-7—---7777—— —---————77— 77777 7 7 7 7 — 7 — I o 30% -7—— — 77777 O n u 20% 2 777—l -— 7777777 777 77 7 7 7777777 — - Isobutyrlc Acrd o . . ‘ A Propanorc Acrd o . . . 10% ,, ._. - 7 7 7 ‘6 . .777— § 87.77 Plvallc Acrd A 0 x 4 9 00/0 ‘ 5 g T r I r r . O 50 IOO ISO 200 250 300 350 400 Time (min) Figure 49. Hydrogenation of C3 organic acids to alcohols (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) The various a-substituents were chosen in order to evaluate the effect of hydrogen bonding, electron withdrawing ability, and heteroatom coordination to the surface. In acid, the OL-substituent if it is an amino group is presumably protonated. This allows for two types of hydrogen bonding. The hydrogen can interact with the hydridic hydrogens on the surface thereby increasing the reactivity by promoting interaction between the substrate and the catalyst surface. The other type of hydrogen bonding is intramolecular hydrogen bonding with the free electrons on the carbonyl oxygen, stabilizing the partial positive charge on the carbonyl carbon. In the case of the unsubstituted alkanoic acids 159 (acetic, propanoic, isobutyric, and pivalic acid) and the trimethylated amino species ((CH3)3N+CH2COOH'C1’) the hydrogen bonding ability is eliminated and it was observed that they are very unreactive. In addition a comparison between glycolic acid and methoxyacetic acid indicates a reduced rate of hydrogenation in the case of the methoxy (it-substituent. Thus, hydrogen bonding may be playing a role in accelerating the hydrogenation reaction. 100% 90%7777— 7 77 -7 80% - 7 7 77 7 7777. —+—— — Glycine 700/0 _,-,,__._____v,m_ W, 77#, a , ,. fi,~m ,9 __d W A 60% - 7— . 7 7 7-777777-7 777777- e: E 50% ‘* , ‘ * i _ A * i ‘ Glycolic Acid .2 O >-' o ' - 40% ‘ .____g_*___ __ _ o___ _‘ Methortyacetic Acrd Sarcosme o I A 30% 7 -- -- 77 7777 - 7 77 - 7 O O l . N,N-dmg 20% ———--— — “—6 7} — 7 7 7* x . ° = A Propanoic Acrd X . . 10% 7— ——— — 7—7— 77 -- X777 7 —— 7— A 77 7 —- Acetic Acrd 9 f . 8 0 Betaine (w/o Cl-) 0% i a I I a 1 I; 01 I? To Betaine (w Cl-) 0 50 100 150 250 300 350 400 Time (min) Figure 50. Hydrogenation of mostly C2 organic acids to alcohols (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) It was observed that (CH3)3N+CH2COOH°C1' was not very reactive (<2% yield of choline chloride) and the extent to which the chloride counter-ion was playing a role was not 160 fully understood. So, betaine ((CH3)3N+CH2COO') was also run under the same conditions (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C). A conversion of 41% was obtained, but only 7% yield of the choline product of carboxylate reduction was achieved. Thus, the presence of the chloride counter-ion appears to be inhibiting the reaction. No other products were detected in the HPLC. Table 2. Hydrogenation of C3 and C2 organic acids to alcohols after six hours (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) (Substrate) a-substituent Conversion (%) Yield (%) Selectivity to Alcohol (c3) NH3 94 86 91 (c2) NH3 96 80 83 (C3) OH 93 79 85 (c3) 61,13 (0H)2 76 68 100 (c2) OH 56 47 82 *(c3) OCH3 48 47 98 (C2) OCH3 53 35 67 (c3) N+H2(CH3) 37 35 96 (c3) N+H(CH3)2 4o 20 51 (c3) (CH3)2 46 12 27 (c3) CH3 39 16 40 (c2) CH3 / (c3) H 33 13 39 (C3) N+(CH3)3 41 7 17 (C2) H 19 7 37 (c3) N+(CH3)3Cl' o 3 0 *Data from Norbert Varga’s thesis (Michigan State University) Looking at the low selectivities to the desired alcohol for the unsubstituted alkanoic acids, it is assumed that some of the by-products that are formed include methane, ethane, and propane.”49 Zhang et al. found that at temperatures below 170 °C these were the 161 typical by-products and at temperatures above 170 °C, degradation to l-propanol, ethanol, and 2-propanol were observed. 4.6.2 Amino Acid Substrates Additional experiments at temperatures of 100 °C and 150 °C were conducted with glycine and the N—methylated glycine series. The results for conversion, yield and selectivity are presented in Table 3. In general, the selectivity decreased at higher temperatures indicating increased degradation. Table 3. Hydrogenation of glycine, sarcosine, N,N-dmg, and betaine at 100, 130, and 150 °C after six hours (0.3 M H3PO4, H20, X °C, 1000 psi H2, 1g dry 5% Ru/C) 100 °C Conversion (%) Yield (%) Selectivity Glycine 31.4 28.2 89.8 Sarcosine 10.2 11.7 100 N,N-dimethylglycine 13.8 11.2 81.1 Betaine 4.1 0 0 130 °C Conversion (%) Yield (%) Selectivity Glycine 89.8 82.8 92.2 Sarcosine 40.5 30.4 75.1 N, N-dimetMglycine 29.4 18.6 63.2 Betaine 40.6 6.8 16.7 150 °C Conversion (%) Yield (%) Selectivity Glycine 98.5 79.9 81.1 Sarcosine 71.6 50.7 70.8 N,N-dimetrflqucine 57.1 41.1 72.0 Betaine 73.7 17.4 23.6 The initial rates were calculated from plots of concentration (mol/L) versus time (min) and are reported in units of kmol/kg of cat/min (Table 4). The hydrogenation rate of reaction is reduced by a factor of ~3 as each methyl group is added. The activation 162 energy was calculated for each substrate and was 41, 56, and 50 kJ/mol for glycine, sarcosine, and N,N-dimethyglycine. Table 4. Rates of reaction for hydrogenation of glycine, sarcosine, and N,N-dmg (0.3 M H3PO4, H20, 100, 130, and 150 °C, 1000 psi H2, 1g dry 5% Ru/C) -Ri (kmol/kg of cat/s)* 100°C 130°C 150°C Glycine 6.2 16.7 30.9 N-methylglycine 1.6 6.9 13.3 N,N-dimethylglycine 0.6 3.9 10.8 Betaine 0.0 0.0 0.0 *Rates are multiplied by 10"7 It is not clear whether the heteroatom at the oc-position is adsorbing on the surface. There are a few different mechanisms to consider. It may be that the a-carbon adsorbs, or if the acidic proton on the heteroatom is able to dissociate, then a pi bond can be formed and interact with the surface. The ability to form this pi bond is only present in the hydroxy, amino, methylamino and di-methylamino substrates. In traditional C-H activation, it is the lone pair on the N that is active, so varying the amount of acid might lead to an understanding of how the OH activation is affected. We know that for hydrogenation to occur, acid must be present to ensure the carboxylic acid is protonated. Looking at the amine species, the acid amount can be varied to slow or prevent hydrogenation. Jere et al. examined the effect of phosphoric acid concentration on conversion of 0.22 M alanine (100 °c and 1000 psi H2).50 The conversion was ~90% with 0.29 M H3PO4, ~65% with 0.14 M H3PO4, and 45% with 0.07 M H3PO4. H/D exchange will be exploited and discussed in section 4.7 to examine how the reactivity of each of the amine species is affected. 163 The methoxy substituent evidently benefits the reaction over the similar sized methyl substituent. From the data, it appears that the hydrogenation reaction is not as sensitive to sterics, with respect to a methyl group. In fact, the methylated substrate (isobutyric acid) appears to be slightly more reactive than the simple carboxylic acid substrates (propanoic acid, acetic acid). But, pivalic acid is about the same in reactivity as propanoic and acetic acid suggesting that at some point the addition of another methyl (as compared to isobutyric acid) may start to slow the reaction. In theory, the relatively bulky methoxy or methyl group may adversely influence the adsorption of the substrate on the catalyst surface as free rotation around the sigma bond between the C1 and C2 carbon is possible. The presence of a bulky substituent close to the sp2 carbon of the carboxyl functionality could hinder the delivery of the hydride. However, this steric effect does not appear for the simple alkanoic acids. The methylated amine species have provided more information on the sensitivity of the hydrogenation reaction to steric modifications at the a—amino position. Looking at the trend for —N+H3, -N+H2CH3, -N+H(CH3)2 the reactivity falls off by a factor of three as each hydrogen is replaced by a methyl. However, these replacements also represent changes in the amino groups’ H-bonding abilities. To determine whether the addition of a methyl(s) is hindering the substrate interaction with the catalyst surface, H/D exchange studies can be done on this series. If less exchange is observed as methyl groups are added, it may support the decrease trend in reactivity, partly due to a steric effect. 164 In the hydrogenation of organic acids, electronics may be more important than sterics due to OH activation. To understand the relative rates of OH activation as well as gaining more mechanistic understanding of the reaction occurring at C2 in the organic acids, a series of substrates has been studied; Direct C-H activation has been observed with an ether (THF) and direct C-H activation or indirect (via dehydrogenation) C-H activation in terminal alcohols (ethanol, glycols) is commonplace with the heterogeneous Ru catalyst systems studied here. For future work, it is suggested that (S)—2-hydroxybutanol, (S)-2- methoxybutanol, and (S)-2-aminobutanol be evaluated because they are secondary alcohols (similar to the organic acid substrates). It would be interesting to measure the rate of deuterium incorporation at C2 as well as observe whether deuterium is incorporated at either C1 or C3 in the hydroxyl and amino substrates, indicating the enol and enamine intermediates. The rate of deuterium incorporation at C2 can be compared to the amount of deuterium incorporated at C2 of the corresponding acid. If deuterium is incorporated in (S)-2-methoxybutanol it may occur via direct or-C-H activation. Another factor that is inevitably affecting the hydrogenation reaction is the adsorption affinity of each of the substrates. Competitive adsorption studies with the four aniino species (N+H3, N+H2CH3, N+H(CH3)2, and N+(CH3)3)C1' should also be carried out to evaluate their relative adsorption affinity. If the trend follows their reactivity trend then a deeper understanding of the decreased reactivity can be realized. If they have similar adsorption affinities; however, it would suggest that the sterics or electronics are the controlling factors. Understanding the product distributions in the amino species and the reaction kinetics will help fill in details regarding their relative reactivities. 165 4.7 H/D Exchange Studies 4.7.1 Amino Acid Substrates 4.7.1.1 H/D Exchange at the Methylene Position The catalytic hydrogenation of the methylated glycine amino acids to amino alcohols has been discussed. In general, amino alcohols can serve as building blocks for numerous products in pharmaceutical and fine chemical applications. Research conducted in the Jackson/Miller group by Jere et al. and Pimparkar et al. has shown the stereoretentive reduction of alanine, serine, and valine.50,63 In an effort to build upon this research and probe the reactivity differences in hydrogenation of the methylated glycine series discuss above, H/D exchange studies were conducted on this series. The amino acid substrates that were studied included glycine, N-methylglycine (sarcosine), N,N-dimethylglycine, betaine, and glycylglycine. Again, the rate equations (Chapter 3) can be used to calculate rate constants k1 A, km, km, and km. The reactions described consider the back reactions (as opposed to the work by Frey et al.26 and Richard et al.28’29). Although, the back reactions were not expected to have much significance, especially in the reactions with 1000 psi D2 and D20. This will be discussed more as the individual experiments are considered. An early experiment with 0.25 M glycine at 130 °C with 1000 psi H2 and D20 for two hours showed that the H/D exchange reaction occurred very quickly at this temperature. A CH2 conversion of 99% was found and 18% and 80% (calculated) yield of CHD and CD2, respectively. The 1H NMR spectra taken of time zero and after two hours of 166 reaction time are compared in Figure 51. Conversion of CH2 is already at 71% at time ZCI'O. 110 ---“- n— w—-~~——--—--—--——v-——————~—~—-- ———~——~—---——-——-~~ 90 +Time: 0 Hours —+—- Time: 2 Hours 100 ~- — , ,, _._____.___ _, #-- v __V,, — 80 90 ~~ .,,,, 7— "—9 «W 7* A 7777777 . . ~~ r 70 80 ~—~ ~— — 777w, , ~ — WWW, - _--VVLW. . ~ 60 70 -.,___-_ 7* 777w , - - ,, , ~ 50 60 w - i—— - . r 40 50 - ———~~ ~77 — . ,Aifl ~ - , —~ ———~~We~ 30 40 . 20 30 ~ 10 20 0 10 ~——~—————————— ———————7—- —- -— - — m» -10 0 = = - ~ , -20 3.33 3.32 3.31 3.30 3.29 3.28 3.27 3.26 3.25 Figure 51. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) H/D Exchange Studies at 100 °C with 1000 psi Hg In order to better understand and model the reaction rate kinetics, the reaction was subsequently run at lower temperatures (100, 90, 70 and 50 0C). An experiment with 0.25 M glycine at 100 °C with 1000 psi H2 in D20 was allowed to react for five hours, the singlet 1H NMR peak representing the methylene (CH2) position continued to decrease in size over time, the triplet peak representing CHD grew in reaching a maximum peak height at 60 minutes then started to decrease as the hydrogen at that position was further exchanged with deuterium. The stacked 1H NMR spectra of time 0, 1, 2, 3, 4, and 5 hours of reaction time is compared in Figure 52. 167 Reaction " ” “ A time My Figure 52. Stacked lH NMR of H/D exchange at the methylene position on 0.25 M glycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) Figure 53 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methylene carbon are: klA = 2.3x10‘4, km = 1.2x10'4, kZA = 1.2x10'4, k-2A = 3.2x10'4 L/mol/min and the sum of the squares error (SSE) = 3.9x10'5. The rate constants for glycine, sarcosine, and N,N-dimethylglycine at these reaction conditions (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) will be summarized after each experiment is discussed. The sum of the squares error (SSE) is minimized for AH2 and AHD; thus it does not include AD2 because AD2 is not directly measured. The plotted points as shown 168 also represent residuals, differences between the known initial glycine concentration and the observable CH2 and CHD amount. 0.25 _..--.---- ..--.-.._..-__---.- -- _-_. ...-._- _ _.... _...--_--._...----._._-_F 20 l VAH2 —AHD —AD2 I AH2_exp A AHD_exp o AD2_exp —H20 and HOD 0 H20 and HOD_exp 0.20 « L, . -----___ - - -7 9—— #4 4» 16 -. . a 3 = g 015 -. 12 E’ ' : ° ‘5 a E 010 g 8 IL : o O 0.05 s» .. 4 0.00 V I T r T 7 0 0 50 100 150 200 250 300 350 Time (mln) Figure 53. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) For this experiment, glycine reaches equilibrium at about 4 hours, thus the statistical equilibrium end points can be calculated and compared to the experimental data. The statistical concentrations of CH2, CHD, and CD2 based on the actual percent hydrogen in water at five hours (14.5%) and the concentration at time zero (0.23 M) are 0.0048, 0.057, and 0.168 M, respectively. The actual concentrations of CH2, CHD, and CD2 at five hours of reaction time are 0.0054, 0.075, and 0.166 M, respectively. Thus, the values are in good agreement. 169 The same experiment using sarcosine (N -methylglycine) was performed at 100 °C with 1000 psi H2 and D20 for six hours. The 1H NMR spectra of time zero and after six hours of reaction are compared in Figure 54. Figure 55 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The kinetics were determined using the models described in Chapter 3. Using kinetic rate model A (Chapter 3.2.1) the rates for H/D exchange on the methylene carbon are: km = 1.6x104, km = 0, k2A = 1.6x10'4, k-2A = 2.6x10'4 L/mol/min and the sum of the squares error (SSE) = 4.5x10'3. Again, we can compare the statistical equilibrium end points and the experimental data. The statistical concentrations of CH2, CHD, and CD2 based on the actual percent hydrogen in water at six hours (15.0%) and the concentration at time zero (0.24 M) are 0.0054, 0.062, and 0.176 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 0.004, 0.054, and 0.199 M, respectively. 170 3500 3000 +Time: 0 Hours 7I7 Time: 6 Hours 3000 7 0 7 7 7 2500 1 1 2500 77 2 77 77 777 77 2000 1 4 2000 <— 7 77 7 1500 q I 1500 77 7 777 7 4 1000 1000 77 7 7 77 7 500 500 0 0 —500 3.47 3.45 3.43 3.35 3 33 Figure 54. IH NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 0.30 18 7AH2 7AHD 7AD2 I AH2_exp A AHD_exp o AD2_exp —H20 and HOD 0 H20 and HOD_exp 0.25 i 7777777 7 77 7 77 7 15 g 0.20—7 -7 777 7 .77 r 7 « 12 o o E v 3: C a 3 015- 7 7 7 7 7 7 77 7779 c s - 7 g = a. 3 S o 0.10 —7 77 7 7 7 7 77 7 77 7 77 77 6 \74 0.05 i 77 77 777 77 7—777—‘ 7 3 A I \\.K 0.00 7 1 . I , Y Q r I 0 0 50 100 150 200 250 300 350 400 Time (min) Figure 55. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 0C, 1000 psi H2, 1 g dry 5% Ru/C) 171 From the data, it appears the rate constants are too large or perhaps the fitted line should be starting at some time, t, after time zero. In experiments with 1000 psi H2 and D20 there may be two reasons for the apparent need for a time offset or something equivalent. The first reason is the slight delay in equilibrium between the liquid concentration and the amount of available hydrogen or deuterium on the catalyst surface. And in the beginning of the reaction, the amount of deuterium available on the surface is less so hydrogen exchange at this point may be occurring but with the hydrogen on the surface. The second reason may be a time delay due to the catalyst ramping up in activity. This will be discussed further. The same experiment using N,N-dimethylglycine (N ,N-dmg) was performed at 100 °C with 1000 psi H2 and D20 for six hours. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 56. Figure 57 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. Using kinetic rate model A (Chapter 3.2.1) the rates for H/D exchange on the methylene carbon are: km = 7.1x10'5, km = 0, k2A = 6.4x10'5, k-2A = 1.6x10'4 L/mol/min and the sum of the squares error (SSE) = 2.5x10'3. Again, we can compare the statistical equilibrium end points and the experimental data. However, for the experiment with N,N-dimethylglycine the equilibrium is not reached at six hours. The statistical concentrations of CH2, CHD, and CD2 based on the actual percent hydrogen in water at six hours (15.3%) and the concentration at time zero (0.24 M) are 0.0057, 0.063, and 0.175 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 0.018, 0.084, and 0.142 M, respectively. 172 1400 1 1200 1 1000 1 800 1 600 7 400 7 200 3.59 |3.53 3.52 3.51 3.50 3.49 3.48 ———__—_—‘ 3.55 3.57 +Time: 0 Hours 7| Time: 6 Hours 1 ------ A v vvvv'- A 1 v -A‘ vav' [ 1200 . 1000 w» 800 7 600 400 7 200 -200 3.45 Figure 56. 1H NMR of H/D exchange at the methylene position of 0.25 M N,N-dmg (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 0.30 -_.__....._._---._-___-.. ..-.-. __ .- _---.-. -.. . -____ _...-..._.-.______-_-___ ..--.... 18 7AH2 7AHD 7A02 I AH2_exp A AHD_exp O A02_exp —H20 and H00 0 H20 and HOD_exp 0.25 77777 7 7- 77 7—77 -—7777 777 7 7“ 15 g 0.20 7 77777 W~7~ 12 O f 3: o E 0.15 7 77 9 E i 8 8 0 0.10 7 ~ 6 0.05 .. 3 0.00 r , 0 0 50 100 150 200 250 300 350 400 11me(mln) Figure 57. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 0C, 1000 psi H2, 1 g dry 5% Ru/C) 173 It is important to note, that we need to be skeptical of the rate constants for the back reactions (km and k-2A, deuterium to hydrogen) because these rate constants depend so heavily on small variations in the “main event” of replacing the deuterium with hydrogen and its rate variation with time and how well we are modeling the aqueous [H] concentration. Unlike the hydrogen to deuterium exchanges, there is no initial rate behavior we can easily compare to the fitted constants, because deuterium to hydrogen only happens after substantial deuterium incorporation has occurred, and even then, it’s a minor process because of the relatively low hydrogen concentration in the water. In addition, our water hydrogen and deuterium concentration versus time function is an empirical fit to fairly spread out data points and if it were off a bit, that could throw off the back reaction rate constants significantly. In order to partially address this, a more constrained model using essentially one k value for each site and then modified via isotope effect numbers was also explored. A kinetic isotope effect is observed when the reaction rate is affected as one atom in a compound is replaced with its isotope. For the case of hydrogen and deuterium, the K”; is expressed as KIE=kH/kD. All of the H/D exchanges are assumed to be occurring through the same process, but now instead of four independent variables, the system is redefined and has three independent variables, kiA, pA, and SA. As discussed in Chapter 3, k1 A is equal to 2*kiA indicating that there are two hydrogens on the methylene carbon and subsequently two chances for H/D exchange. The equations for how the k, A, etc rate constants could be effectively calculated from the kiA, pA, and SA values are in Chapter 3 (equations 56 and 57). It was assumed pA and SA are independent. 174 Looking at the same three experiments glycine, sarcosine, and N,N-dimethylglycine at 100 °C with 1000 psi H2 and D20 the chemical equations which incorporate the kinetic isotope effects (6-variable kinetic model) were solved. For the forward reactions, the rate constants using both models are similar. For the back reactions the rate constants are somewhat different. Again we need to be skeptical of their meaning. When constraining the model to use only one rate constant, ki and incorporating primary and secondary isotope effects, the rate constant for glycine is kiA = 1.2x10“1 L/mol/min and the isotope effects are pA = 0.76 and SA = 0.98. For sarcosine kiA = 0.82x104 L/mol/min, pA = 1.25, and SA = 0.51. For N,N-dimethylglycine kiA = 0.36x10'4 L/mol/min, pA = 2.12, and SA = 0.69. The rate constants, ki, for glycine, sarcosine, and N,N-dimethylglycine are about half of k1 A and k2A that were calculated using the lO-variable model. This makes sense based on the fact that k1 A = 2*kiA and the klA as computed from both the 10 and 6-variable models are nearly the same. In addition, the fact that the kiA values are similar to k2A for sarcosine and N,N-dimethylglycine makes sense because k2A = kiA/SA and SA for those two substrates is close to 0.5. For glycine, the secondary isotope effect is essentially one, so kiA is essentially the same as k2A and half the value of k1 A. It is important to note that it is the value of p A that is most sensitive to the back processes, so the SA values should be more reliable than the pA values. For each N-methyl group added, the rate drops off by about one-third. 175 To further verify whether isotope effects are playing a role in the IUD exchange reaction at 100 0C with 1000 psi H2 and D20, the primary and secondary isotope effects for the experiment with 0.25 M sarcosine were fixed to 1. The rate constants km and kic were then solved. The newly solved rate constants km and kic were equal to 1.1x10“4 and 0.51x10‘4, respectively. These were compared to the original km and kiC which were equal to 0.82x10'4 L/mol/min and 0.40x104 L/mol/min, respectively. The quality ofthe fits indicated that when excluding the potential isotope effects, the rate constants are too large with respect to the experimental data. Due to the knowledge of a delay in the available deuterium on the surface in the reactions with 1000 psi H2 and D20 a modified kinetic model was developed. In addition, the same set of experiments with glycine, sarcosine, and N,N-dimethylglycine were performed with 1000 psi D2 (and otherwise identical conditions). The rate constants for these pairs of experiments should be the same. The comparison of these rate constants as well as the modified kinetic model will be discussed following the experimental results with 1000 psi D2 (100 °C). H/D Exchange Studies at 100 °C with 1000 psi D2 As mentioned earlier, it is assmned that the equilibrium between the hydrogen/deuterium on the surface and the deuterium/hydrogen in the liquid at a temperature of 100 °C is fast with respect to the experimental data points. It is expected that the equilibrium is reached within the first hour. In the experiments with 1000 psi D2 and D20 there is only ~3% or less H in the solvent water and the gas is D2, so the equilibrium is essentially reached at 176 time zero. In the experiments with 1000 psi H2 and D20 there is about ~15% H in the solvent water and the apparent delay in equilibrium needs to be addressed. The delay in equilibrium and its potential affect on the H/D reaction will be explored in this section. The same three experiments were performed at 100 °C with 1000 psi D2 and D20. For the experiment with 0.25 M glycine, the 1H NMR spectra of time zero and after six hours of reaction time are compared in Figure 58. The singlet 1H NMR peak representing the methylene (CH2) position decreased continuously while the CHD triplet grew in over time, reaching a maximum peak height at sixty minutes of reaction time, and then decreasing as hydrogen at that position was further exchanged with deuterium (Figure 59). Figure 59 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The forward rate constants for the reaction with glycine at 1000 psi D2 are fairly similar to those with 1000 psi H2. The rates for H/D exchange on the methylene carbon are: 1(1A = 1.7x104, k-1A = 0, k2A = 1.2x10‘4, k_2A = 7.0x10'5 L/mol/min and the sum of the squares error (SSE) = 7.9x10'5. 177 5000 4000 _ "" "" "" '_ — — "'25" + Time: 0 Hours 7!- Time: 6 Hours 4500 .J ‘ 3500 4000 77' 77 7 7 3000 3500 71 7 2500 3000 "I 3.30 .7 . id. 2000 I— —— —— —— —— — _ — — 2500 77 - 1500 2000 77 7 1000 1500 77 77 7 7 7 7 500 1000 0 500 7- 7 77 7 7 77 7 -500 f"""""'_'--"'1 o __________________ 1 4000 3.34 3.33 3.32 3.31 3.30 3.29 3.28 3.27 3.26 3.25 Figure 58. lH NMR of H/D exchange at the methylene position of 0.25 M glycine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 0.30 3.6 AH2 7AHD —AD2 I AH2_exp A AHD_exp O A02_exp —H20 and H00 0 H20 and HOD_exp 0.25 7 7 77 77 3.0 I .0 N o Concentration (mol/L) O a? 3 Percent H 0.10 - 17 1.2 0.05 — «- 0.6 0.00 0.0 o 400 Tlmo (min) Figure 59. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 178 As in the experiments with 1000 psi H2 and D20, we can compare the statistical equilibrium end points and the experimental data. The statistical concentrations of CH2, CHD, and CD2 based on the actual percent hydrogen in water at six hours (2.2%) and the concentration at time zero (0.24 M) are 1.1x10'4, 0.0101, and 0.227 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 4.7x104, 0.012, and 0.24 M, respectively. For this experiment, it appears that glycine has not quite yet reached the equilibrium at six hours. The concentration of CHD at time zero was compared in the reactions with 1000 psi H2 and with 1000 psi D2. For both experimental conditions, the presence of AHD was noted. This suggests that even in the case with 1000 psi H2, there is some deuterium available on the surface for exchange. In general, the experiments that were conducted in 1000 psi D2 and D20 resulted in larger concentrations of CHD at time zero. The same experiment with 0.25 M sarcosine was performed at 100 °C with 1000 psi D2 and D20 for six hours. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 60. Figure 61 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The reaction of sarcosine started out quickly and the peak concentration of CHD occurred at about 60 minutes of reaction time but then rapidly fell off as the CD2 grew in. The H/D exchange reaction rate constatns on the methylene carbon are: k1A = 2.8x10'4, k1 A = 6.9x104, k2 A = 2.9x10'4, k-2A = 8.3x10'4 L/mol/min and the sum of the squares error (SSE) = 5.7x104. 179 4500 3500 '— _ _ — _ _ _ "£01 +Time: 0Hours 7I-7T1me: 6Hours 4000 ..I 7 716 L 777 7 - 3000 7 7 7 12 I 3500 4 7’ 7 7 ' 8 777 77777 7- 2500 1 ’ ‘4 ' . r 1 0 3000 77 7 777 7 7 7 777 2000 I 3.40 3.39 3.38 3.37 3.36 2500 .-_-- _ — — .'._ f. :-: -, Li --- -.. - . 1500 200077 7 7 7 7 77 7 777777777 1000 1500 -7 7 7 77 500 AAA 1000 1 r , 0 500 «77 7 777 77 7 7 7 7 - 7500 l"""""""_-—__'_'-_l 0 _____________________ -1000 3.43 3.42 3.41 3.40 3.39 3.38 3.37 3.36 3.35 Figure 60. lH NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 0.30 3 0 7 AH2 7 AHD —AD2 I AH2_exp A AHD_exp o AD2_exp -—H20 and HOD 0 H20 and HOD_exp 0.25 77 77 777 7 77 77 77 2.5 g 77 2.0 O E V I C a O C E 0.157 7 o 77 7 7 7 7 77 7 7 777 7 7 715 3 O C n. 8 S o 0.1077 77 77 777 7 77 77 77 7 7 7 7 771.0 0.054 7 7‘77 7 7 7 7 7 7 7 7 7 7 7 7 7 77- 0.5 A NEW—... 0.00 r r r i r ' -— 00 0 50 100 150 200 250 300 350 400 Time (min) Figure 61. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 180 The experiment with sarcosine occurred very quickly, reaching an equilibrium at or before 4 hours. The statistical equilibrium end points were again compared to the experimental data. The statistical concentrations of CH2, CHD, and CD2 based on the actual percent hydrogen in water at six hours (1.95%) and the concentration at time zero (0.23 M) are 7.3x10'5, 0.0081, and 0.226 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 1.2x104, 0.009, and 0.225 M, respectively. The same experiment using 0.25 M N,N-dimethylglycine was performed at 100 °C with 1000 psi D2 and D20 for six hours. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 62. Figure 63 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The reaction of N,N-dimethylglycine again started out quickly and the peak concentration of CHD occurred at about 60 minutes of reaction time but then rapidly fell off. The rate constants for H/D exchange on the methylene carbon are: k; A = 1.4x10'4, kl A = 0, k2A = 1.1x10'4, k-2A = 2.2x10'4 L/mol/min and the sum of the squares error (SSE) = 6.3x10‘4. The experiment with N,N-dimethylglycine was slower and did not reach equilibrium after six hours. The statistical equilibrium end points were again compared to the experimental data. The statistical concentrations of CH2, CHD, and CD2 based on the actual percent hydrogen in water at six hours (2.7%) and the concentration at time zero (0.20 M) are 1.5x10'4, 0.0105, and 0.188 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 5.3x10'4, 0.0114, and 0.187 M, respectively. 181 1200 1000 +Time: 0 Hours 7'7 Tlme: 6 Hours 0 3.49 3.48 3.47 3.46 3.45 3.44 3.43 3.42 3.41 Figure 62. ]H NMR of H/D exchange at the methylene position of 0.25 M N,N—dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 0.20 4 —AH2 —AHD —ADZ I AH2_exp A AHD_exp o ADZ_exp . —-H20 and HOD 0 H20 and HOD_exp I 0.15 7 7 7 3 1 o 3 6 . E o ‘ I 8 a E 010 7 7 7 7 7 7 7 7 2 § .- O 5 n. 8 A o I o 0.05 - . 7 7 7 7 7 7 7 7 7 77 1 1 A \\ I I ‘ 0.00 . 7 . 7 j 7 Y ._. 0 0 50 100 150 200 250 300 350 400 Tlme(m|n) Figure 63. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 182 Figure 64 compares the rate constants for glycine, sarcosine, and N,N-dimethylglycine on the methylene carbon at 100 °C, 1000 psi D2, and D20 (reaction time = 6 hrs). For the first hydrogen exchanged, sarcosine reacts the most quickly and for glycine and N,N— dimethylglycine the rate falls off by roughly one-half. After the first hydrogen is replaced with deuterium the rate of exchanging the second hydrogen with deuterium is essentially the same for each compound as compared with the first hydrogen that is exchanged (kl A compared to k2A). 3.5E-04 I Glycine I Sarcosine I NNdmg 3.0E-04 7 7 7 . i7 2.5E-04 7 2.0E-04 7 1 .5E-04 7 77 Rate constant (leollmin) 1.0E-04 77 7 5.0E-05 7 0.0E+00 7 Figure 64. Rate constants for H/D exchange at the methylene carbon (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) The sensitivity of the kinetic model to the back reactions is greatly reduced in the experiments conducted with 1000 psi D2 and D20 at 100 °C. To verify this, the kinetic model was modified and the back reaction rate constants were set to zero. So, the number of fitted variable was decreased from 10 to 5 (Model A). The fits were compared 183 and for the experiment with D2, setting the back reaction rate constants to zero had essentially no effect. Using the more constrained model (6-variable) the rate constants k. A, k-1A, k2A, and k-2 A were compared to the computed rate constants obtained from using the 10-variable kinetic model. And, for the forward reactions the rate constants k1 A and k2 A are essentially identical using either model (as previously mentioned). For the experiments with 1000 psi D2 (D20, 100 °C, 1 g dry 5% Ru/C) the calculated secondary isotope effects are comparable and less than one for glycine, sarcosine, and N,N-dimethylglycine (Figure 65). For glycine kiA = 0.84x10“1 L/mol/min, pA = 4.4, and sA = 0.72. For sarcosine ki A = 1.3x10‘4 L/mol/min, pA = 0.76, and sA = 0.46. For N,N-dimethylglycine kiA = 0.71x104 L/mol/min, pA = 1.1, and SA = 0.65. The rate constants, ki, for glycine, sarcosine, and N,N-dimethylglycine are roughly half of klA using the 10-variable model. Again this makes sense because k, A is defined to be half of k1 A. And the k] A from the 10- and 6-variable models are nearly the same. In addition, the secondary isotope effects, 3A, are between 0.5 and 0.7, thus the calculated k2 A from the 6-variable model are all about twice the value of k, A. This is again similar to the values computed for k. A. ' 184 1.5E-O4 7 7 7 5.0 DKi i 17315-04 Oprimary isotope effect , 4'5 . A secondary isotope effect 1.2E—04 «7 4.0 g 3.5 E ‘E’ 9.0E705 7 8.4E-05 ' 3.0 ,- 2 ‘ E: E 1 7.1E-05 .. 25 g g 6.0E-05 — p . 2,0 3‘ 2 ’3 g 1.5 3.0E-05 . V .. , 7 1,0 7 ‘ -. O7 I . A . I, 0 5 0.0E+00 V 0.0 Glycine Sarcosine NN—dmg Figure 65. Kinetic rate constant and primary and secondary isotope effects for glycine. sarcosine, and N,N-dmg (D20, 100 °C, 1000 psi D2. 1 g dry 5% Ru/C) It is interesting to note that in the case of 1000 psi D2 and D20 the rate constants for glycine and N,N-dimethylglycine are comparable and both about one-third to one-half that of sarcosine. Kinetic Model Modification — Linking experimental data 1 1000 psi H2 and D3) The kinetic model for isotope exchange was varied to explore the effects of various assumptions with the intent of optimizing the fit to the experimental data. Experiments at 100 °C were discussed for glycine. sarcosine, and N,N-dimethylglycine with either 1000 psi H2 or D2 and D20 as the solvent. The rate constants calculated for these pairs of experiments should, in theory. be the same. Using the kinetic model discussed above. the H/D exchange rate constants for the experiments with 1000 psi H2 and D20 appeared to 185 be too large. This is partly due to the fact that in the kinetic model the hydrogen or deuterium available for exchange with the substrate was assumed to be in equilibrium with the concentration of hydrogen or deuterium in the liquid. But from the experiments with H2 and D20, it appears that there is a delay in the substrates' H/D exchange reaction due to the time it takes to reach equilibrium among liquid, gas, and surface hydrogens. Note that for the experiments with 1000 psi D2, the aqueous hydrogen concentration is almost an order of magnitude lower than the typical 15% or so that is observed with H2 (Figure 67). This time evolution is important, but also reflects the gas/liquid exchange. The rate of exchange for the second hydrogen is essentially the same in both conditions (H2 or D2) which suggests an equilibrium is reached and at that point the reaction is not limited by the presence of available deuterium. 0 H20 and HOD_1000psi H2 9 H20 and HOD_1000psi DZ ' 16 -~— 14« 12 «- 10- Percent H —-—-'—‘f 4" “v- _A» . -L 0 50 100 150 200 250 300 350 400 Time (min) Figure 66. Percent H in D20 with either 1000 psi H2 or D2 186 In terms of the delay of the deuterium available on the catalyst surface, this may be a result of not having deuterium available to get on the surface until there is enough HD formed. Scheme 30 shows the proposed mechanism for exchanging H and D between the various isotopomers of H2, H20, and Ru-bound H atoms. Previous work in our research group has shown that a significant amount of HD is generated before D2 is formed. According to this mechanism, it is not until HD is formed (in the experiments with H2 and D20) that the catalyst surface is deuterated so that the substrate can start undergoing H/D exchange. ,0 o D 1: H D x’ \ / Dx‘ / H O I ‘0 R. + H ——> ——» l RU \RU H—H H-D ,H D H D 3i: H D H \0/ I ‘O/ , / «— ‘__ Ru ”—0 H‘\ l RU ‘Ru Scheme 30. Proposed mechanism for Ru/C catalyzed HD generation from H2 and D20 Note that exchange between H2 and D20 takes place without isotopic exchange on the Ru surface. Various kinetic model modifications were employed to address this background equilibrium process. To start, the corresponding H2, D20 and D2, D20 experimental datasheets were linked and a common set of rate constants was defined. The hydrogen and deuterium available were defined by HS and DS (Chapter 3) as opposed to using the 187 liquid hydrogen and liquid deuterium concentrations. It would be ideal to assume that the total concentration in solution is in equilibrium with the surface, but since this can not be confirmed, it has to be treated simply in terms of rates. DS is a variable and represents the fraction of total sites occupied by deuterium. For the experiments conducted with 1000 psi H2, the initial H on the surface was assumed to be 111.11. This is because at time zero it is assumed that the amount of hydrogen on the surface is more similar to the gas (which in this case is H2) than the aqueous D20. For the experiments conducted with 1000 psi D2, the initial H on the surface was assumed to be equal to the amount of H measured in the liquid (usually less than 3%). In addition, a catalyst tum-on function was included, as defined by: 1 -(tdep__coeft)*exp(tdep__exp*t), where tdep_coeff is constrained to range 0-1 and can describe the fact that part of the catalyst is fully active and there is a fraction that is not. This variable will ramp up to full activity on a timescale described by the tdep_exp variable, an effective rate constant for catalyst recovery. Numerous fitting attempts were performed with the catalyst tum-on function included. The results for both the lO-variable model and the 6-variable model (both now with kH as defined in Chapter 3) indicated that with tdep_coeff and tdep_exp there was little to no improvement of the fits. This finding is understandable in that much of the mathematical flexibility the turn-on function offers is quite similar to that conferred by the kH variable. 188 The main focus was on using the 10- or 6-variable model with the addition of kH (as defined by the equilibrium between the available hydrogen and deuterium on the surface and in the liquid). It is important to mention that we do not know the exact amount of available hydrogen and deuterium on the surface as it is not directly measured; only its ratio is truly meaningful. Table 5 presents the rate constants for glycine at 100 0C with 1000 psi H2 or 1000 psi D2 and the linked H2 and D2 results for the 10- or 6-variable model plus kH. It also includes the rate constants derived from the kinetic model presented earlier in this thesis (based on the aqueous hydrogen and deuterium concentrations). The rate constants for the case of D2 and D20 are not affected much by the kinetic model modification, due to the fact that this system essentially starts at equilibrium, so there is not a delay in building up available deuterium on the surface (as in the H2, D20 case). And kH was calculated to be zero with this model, again indicating that for the experiments with D2, the assumption of using the aqueous H and D concentrations in the kinetic model was adequate. For the experiment with 1000 psi H2 and D20, the experimental (symbols) and fitted (line) concentrations of CH2 (referred to as AH2), CHD (referred to as AHD), and CD2 (referred to as AD2) are presented in Figure 67. The fitted lines are a result of the 6- variable model plus k“. This figure also shows the fitted (line) concentrations that resulted from linking the experiment conducted with H2 with that with D2 and solving for 189 one common set of rate constants. Figure 68 presents the analogous results from the experiment with 1000 psi D2 and D20. Again, there is not much difference in the fits. Table 5. Rate constants for glycine (100 °C, 1000 psi H2 or D2) using the original kinetic model, k10 + k“, k6 + kH Glycine Glycine Glycine Rate Constant 100 °C 100 °C 100 °C (leollmin) H2 02 Linked H2 and D, ' k10 k10+kH k10 k10+kH k10+kH k1_A 2.3E-04 2.7E-03 1.7E-04 1.7E-04 1.9E-04 k-1_A 1.2E-04 1.8E-03 0.0 0.0 1.3E-05 k2_A 1.2E-04 1.4E-03 1.2E-04 1.2E-04 1.2E-04 k-2_A 3.2E-04 4.0E-03 7.0E-05 1.2E-04 3.1E-04 kH(10) 8.0E-05 5.6E-05 1.0E-02 k6 k6 + kH . *k6 k6 + kH k6 + kH ki_A 1.15E-04 2.20E-03 8.45E-05 8.49E-05 9.88E-05 p_A 0.76 0.69 4.39 2.36 0.80 s_A 0.98 2.10 0.72 0.71 0.81 kH(6) 8.20E-05 0.00E+00 8.22E-03 The apparent "blip" at the start of the curve fitting is a result from the initial H being defined as all hydrogen (111.11), so that there is essentially no forward rate occurring, k1 A but k.1A can take CHD back to CH2 (this is the same in the methyl case for sarcosine and N,N-dimethylglycine for k2 A and k-3A). The rate constants and primary and secondary isotope effects were calculated and presented in Figure 69. 190 - CH2-_exp A CHD_exp o c0235” ——CH2_H2 —CHD_H2 —c02_H2 0 25 7 7— -~ CH2_Iinked H2 and 02 CHD_|inked H2 and 02 ‘CDZ_linked H2 and 02 g 0.20 7 7 77 7 7 7 O E _ g . _ -7 7'1?“ —# E 0.15 7 7777— ,. 7”“ — — 777 — g ,, 3 0.10 — 7 ‘7 7 /‘““~ 7 7h 7 —77 \\ a“ _ a. _ g ‘\ 0.05 7 — 7 7 >7 —7 7 — 7 —— 77 — 7 A" // 7 at. 0.00 . T . . ' . ‘. 7 o 50 100 150 200 250 300 350 400 Time (min) Figure 67. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) 0.30 - i - CH2_exp A CHD_exp o CDZ__exp —-—CH2_02 ——CHD_DZ —coz_02 0 25 ~ CH2_Iinked H2 and D2 CHD_Iinked H2 and DZ 77 CDz_linked H2 and D2 ' l §0.2077-7—~777—- —— — — 7 ,7 ' — 7 7 --7 7 o .2: //’,/ g '7‘ //'/ 3 /' 30.1577 77 77 / 7 7 77 77 7 7 7 7 7 77 8 ‘ / c . '”"‘—"‘"\\ . 80407—7577 77 7 7 7 7 7 7 7 7 7 7 7 / ‘ i 005-" /7—\ 7777 7777777777 * / \ 4 7‘ ”T" e _ -.h 0.00 7 7 . 7 7* , u o 50 100 150 200 250 300 350 400 Time (mln) Figure 68. Fit of the kinetic rate data for 0.25 M glycine (D20, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) 191 2.5E—03 ._- 2.5 . E1 KiA 0 primary isotope effect A seconda isoto effect 2.0E-03 — ry pe 2.0 E E g 1.5E—03 — 1.5 A n 3 i‘ 3. I '= L“, g y 3 1.05-03 7 1.0 “ ti o “5 A 5.0E-04 7 - 0.5 0.0E+00 7 - 0.0 02 Linked H2 and DZ Figure 69. Kinetic rate constant and primary and secondary isotope effects for glycine with 1000 psi H2 or 1000 psi D2 or H2 and D2 Table 6 presents the rate constants for sarcosine at 100 °C with 1000 psi H2 or 1000 psi D2 and the linked H2 and D2 results for the 10- or 6-variable model plus kH. It also includes the rate constants derived from the kinetic model presented in this discussion prior (based on the liquid hydrogen and deuterium concentrations). Figures 70 and 71 are analogous to 67 and 68 (glycine) and show the experimental and fitted concentration data for the methylene carbon. The rate constants and primary and secondary isotope effects were calculated and presented in Figure 72. 192 Table 6. Rate constants for sarcosine (100 °C, 1000 psi H2 or D2) using the original kinetic model, k10 + kH, k6 + k” Sarcosine Sarcosine Sarcosine Rate Constant 100 °C 100 °C Linked (leollmin) H2 02 Linked H2 and 02 k1_A 1.6E-04 4.6E-O4 2.8E 04 3.0E 04 2.7E—O4 k-1_A 0.0 1.3E-O4 7.0E-04 9.4E-O4 2.4E-05 k2_A 1.6E-04 9.5E-04 2.9E-O4 3.2E-04 2.9E-04 k-2_A 2.6E-O4 1.6E-03 8.3E-O4 7.8E-O4 5.3E-O4 k1_C 1.1E-04 2.1E-04 1.6E-04 1.6E-04 1.6E-04 k-1_C 0.0 0.0 0.0 0.0 0.0 k2_C 1.4E—04 3.5E-04 1.9E-04 1.9E-04 2.0E-04 k-2_C 2.5E-06 4.0E-04 0.0 0.0 1.2E-O4 k3_C 8.3E-05 2.2E-O4 1.4E-04 1.5E-O4 1.3E-04 k-3_C 2.0E-04 8.5E-04 8.5E-04 8.1 E-04 4.7E-O4 kH(10) 7.1E-05 5.2E-05 1.2E-O4 2.9E-04 1.5E-04 _A 1.25 1.34 0.76 0.80 1.08 s_A 0.51 0.16 0.46 0.44 0.49 ki_C 4.0E-05 9.4E-05 5.5E-05 5.6E-05 5.8E-05 _c 0.71 0.75 0.45 0.49 0.82 s_C 0.47 0.40 0.59 0.59 0.63 mg 6.3_E-05 505-05 1.3E-04 Comparing kiA using the linked experimental data and that obtained in the case of 1000 psi D2 and D20, the values are not much different. This is to be expected because the rate constants that are derived from the experiments with D2 and D20 are the " gold standard". Figures 70 and 71 compare the fitted concentration versus time data using the k6 + kH rate constants as well as the linked H2 and D2 k6 + kH rate constants. The fits are extremely close for both the individual data fit and the linked data fit. The results from the H/D exchange on the methyl carbon of sarcosine will be discussed in section 4.7.1.2. 193 I AH2_exp A AHD_exp o AD2_exp -—AH2_H2 AHD_HZ ADZ_H2 0 25 AH2_Iinked H2 and DZ AHD_Iinked H2 and DZ ADZ_linked H2 and D2 . l' "7’ ”” ' ’77””' g 020 . a - .2./,. , artif—7Q a a /” / é // S o '5 0.15 ~7 *5 , i ' ’ ’ 2 ‘0 E d) 0 g _ o 0.10 7777 / \ 7 7 777 7 / F x / \\ ‘\\. I, \ ‘‘‘‘‘‘‘ 7““ 771i _ _. 0.05 - 1" i , "r - A , , 0.00 7 7 . , ‘ . T , " 0 50 100 150 200 250 300 350 400 Time (min) Figure 70. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) 0.30 7 I AH2_exp A AHD_exp o AD2_exp j -—-AHZ_DZ ---AHD_D2 —AD2_DZ Q 025 7 :fl-12_IlnkedH2 anngZ 77 7—AHD:linkefl2 and DZ 7: ADZTlinkediHZ and D2 // / £0.20. 7 77 ’77 7 7 7 7 7 o E 8 fiO.15-~77 i 777 77 7 7 7 7 .‘5 E 8 I: 80.1077; 777 7 77 7 7 77 7 7 /' \. ‘\ (105! .\ wt 7 “#7, r.- i a a 7 a, \. A \\7._ 8“ \ 7 77.77 . _ *1 0.00 . r . - . - , = . = 0 50 100 150 200 250 300 350 400 Tlme(min) Figure 71. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) 194 3.5E-04 1.6 El KiA 3 0E 04 0 primary isotope effect -. 14 ' ' A secondary isotope effect 1.2 E 2.5E-04 r E 3 1.0 A 5 205-04 7 g E - 0.3 i g 155-04 7 g 0 1- 0.6 3 fl 9‘ 1.05-04 - o 77 .4 5.0E-05 ‘ -e 02 0.0E+00 . > 0.0 H2 DZ Linked H2 and D2 Figure 72. Kinetic rate constant and primary and secondary isotope effects for sarcosine with 1000 psi H2 or 1000 psi D2 or H2 and D2 The analogous data is presented below for N,N-dimethylglycine (Table 7, Figures 73-75). The calculated fits for glycine, sarcosine, and N,N-dimethylglycine using the k5 + k“ kinetic model and linking the experimental data from the experiments with 1000 psi H2 and D2 were qualitatively very good and similar to the fits obtained when not linking the experimental data. This indicates that incorporating kH into the kinetic model and using the initial hydrogen and deuterium on the surface (inferred) is the best approach for experiments conducted with 1000 psi H2 and D20. The original kinetic model that does not include kH is adequate for experiments with D2 and D20. 195 Table 7. Rate constants for N,N-dimethylglycine (100 °C, 1000 psi H2 or D2) using the original kinetic model, k10 + k“, k6 + kH N,N-dimethylglycine N,N-dimethylglycine N,N-dmg Rate Constant 100 °C 100 °C Linked (L/mollmin) ‘ 33:11:10.1: -'fli3:.."-I:|EH23'E-‘i.€3.72:538.179 .- . ..ti..-. 3. 1 E.’ .- ' _ k1_A 7.1E—05 1.1E-04 1. E-04 k-1_A 0.0 1.1E-04 0.0 k2_A 6.4E-05 8.6E-05 1.1E-04 k-Z_A 1 .6E-04 3.15-04 2.2E-04 k1_C 1.0E-04 1.4E-04 1.9E-04 k-1_C 0.0 0.0 0.0 k2_C 1 .4E-04 1.8E-04 3.2E-04 . k-2_C 1.8E-04 3.4E-04 0.0 . 0.0 4.8E-04 k3_C 1.5E-04 1.8E-04 2.8E-04 2.95-04 3.3E-04 k-3_C 5.8E-04 7.5E-04 8.8E-04 8.7E-04 1.3E-03 kH(10) 1.5E-04 3.0E-05 7.9E-05 2,, .. """" [(8 7 ‘ as + $7? k1_A 3.6E-05 7.1E-05 7.2E-05 7. E-05 p_A 2.12 1.14 1.13 0.52 s_A 0.69 . 0.65 0.64 0.60 ki_C 3.5E-05 4.6E-05 6.9E-05 7.0E—05 6.6E-05 _C 0.83 0.81 0.87 0.93 0.68 s_C 0.52 0.58 0.47 0.46 0.48 kH(6) 1.7E-04 3.2E-05 8.2E-05 The rate constants for glycine, sarcosine, and N,N-dimethylglycine at 100 °C (k6 + k”, linked H2 and D2) are kiA = 0.99x10'5, 1.5x104, and 0.74x10'4 L/mol/min, respectively. The trend in reactivity (sarcosine > glycine > N,N-dimethylglycine) does not match the reactivity observed in the hydrogenation of these same substrates (glycine > sarcosine > N,N-dimethylglycine). This may be due to the fact that for H/D exchange an electron- rich C-H bond promotes the reaction. Whereas, in the hydrogenation reaction, it is presumed that steric effects are playing more of a role. It is observed that for each N- methyl group added, the rate of the hydrogenation reaction falls by a factor of 3. 196 0.30 - I AH2_exp" . .. .. . A . AHD_epr .. -.. .. ... _ _. ;-.. AD2_exp - .. - ,_ - .. .. . .. _ , AH2_HZ AHD_HZ ADZ_HZ 0 25 ~ AH2_Iinked H2 and DZ ~ AHD_Iinked H2 and DZ ADZ_linked H2 and DZ _\ g 0.20 7- -- —7 — 77— —7— — 77 O \ g . g t '3 0.15 T ’7 ' F;— ‘ i —i T T “ _ é \ 2..., . g \ ..77 ~ 7 ~r__2‘ _/ 3 0.10 7 —7 , 7 77-7 , , . I A o ,// Q /// \- 0.05 7— -- —- —7 /F 7 // v7-:‘~ i, A ' / /. / o / ' . / /” ‘ 7/ i 0.00 I V Y T T 1 0 50 100 150 200 250 300 350 400 Time (min) Figure 73. F it of the kinetic rate data for 0.25 M N,N—dmg (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) 0.30 I AH2_exp A AHD_exp o ADZ_exp — AH2_DZ AH D_DZ ADZ_DZ 0 2, ___:Afiyinked H2 and- 9.2; 211977.912 and. 92 :_:_603—J"_“$Ed 32am 02 £02077 7 22 - _ _ i ..... - * o E 8 E 0.15 — —7 B 8 c 8 0.10 7 {a 0.05 7 ’ 7‘ 1 0.00 . 0 50 1 00 1 50 200 250 300 350 11me(mln) 400 Figure 74. Fit of the kinetic rate data for 0.25 M N,N—dmg (D20, 100 0C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) 197 1.0E-04 ___ wwww we 1_4 E! KiA 0 primary isotope effect Aseconda isoto effect 8.0E-05 ~ W pe 0 6.0E-05 * 1 r .o m KIE (kHIkD) Rate constant (leollmln) 4r > .~ } if; _ O O) 4.0E-05 ‘ “*3; . .ri: : ,. ' 1. . . ‘ tr , ¢ ‘ "'_,{ . ,7 , , a "72“.: ‘7 “:55, ‘2 ‘1‘. , ,A , .m'r‘i; 2“. Tfizdh.‘ ‘_ Fifi-g" , 22::‘F‘L‘3 rift-3' . . .“fltr‘ fir E'Ti'fiyifiifi-‘Jhr .. ' ‘ x. 3&3?“ 2 “153's 0 4 , , , , A . V _r,_ .:.,‘J;. 'b "7 ‘h ‘ ‘ 7' .‘ I t r t‘ , _' ‘532 2 0E'05 ‘ ‘i i ' ' ‘1 '. ; . .t .t 411-2.: - ‘3‘ i 7 -' f 7.9.5 ,2‘»;'~‘.-, 1;; g: ' i "X"? ‘ “l " + 0.2 0.0E+00 ,...‘...’_ I 71:32:53.". "$331: ' I “_..ALZ‘L‘ 00 H2 DZ Linked H2 and 02 Figure 75. Kinetic rate constant and primary and secondary isotope effects for N,N-dmg with 1000 psi H2 or 1000 psi D2 or H2 and D2 H/D Exchange Studies at 70 °C with 1000 psi D2 The reactions at 100 °C were occurring quickly reaching a maximum CHD concentration at sixty minutes. Therefore, the reaction temperature was reduced to 70 °C in order to obtain additional reaction rate data. The same three experiments were individually performed at 70 °C with 1000 psi D2 and D20. For the experiment with glycine (0.25 M) the 1H NMR of time 0.5 hours and 6 hours is compared in Figure 76. The singlet 1H NMR peak representing the methylene (CH2) position continued to decrease in size over time, while the CHD grew in, reaching a maximum peak height between 60 and 180 minutes of reaction‘time, and then started to decrease as the hydrogen at that position was further exchanged with deuterium (Figure 77). 198 2500 _________ 1750 _ 600 +Time: 0.5 Hours -—I—Time: 6 Hours 2250 J 7 7 »- 1500 2000 «7| 7 1250 1750 71 7 7 1000 ‘1500 4 3.30 3.29 3.28 3.27 3.26 , 77W t 750 1250 " , 77 7 _ 500 1000 77 7 7 77 77 77777 250 750 o 500 777 """"""" 717 777777 - -250 I I 250 -7 7 1 7 - -500 I I o . -750 3.35 3.33 3.31 3.25 3.23 Figure 76. 1H NMR of H/D exchange at the methylene position of 0.25 M glycine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 0.30 4.8 _ —AH2 77AHD 7—A02 I AH2_exp A AHD_exp o ADZ_exp —H20 d H D 0 H20 dHOD 0.25 3 ,, an 0 an 7 _expr , 0 4 I Q 0.20 - 7 7 7 7 77 3.2 3 E z 5 E = 0.15 < 7 7 7 7 7 77 7 77777 77 2.4 w 8 A M C A _.A. 8 0.10 7 7 7 7'7 77 7 ‘7“1 7 7- 16 I o 0.05-77 7 47 7 77 7 7 7 7 7 7 770.8 . I 0.00 c v t t . , , , 0 0 50 100 150 200 250 300 350 400 Tlme (min) Figure 77. Fit of the kinetic rate data for 0.25 M glycine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 199 Figure 77 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange at the methylene carbon are: km = 1.5x10'4, km = 2.8x10'3, k2A = 4.1x10'5, k-2A = 2.5x10' 3 L/mol/min and the sum of the squares error (SSE) = 2.9x104. The same experiment using N—methylglycine (sarcosine) was performed at 70 °C with 1000 psi D2 and D20. The 1H NMR of time zero and after four hours of reaction time is compared in Figure 78. Figure 79 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. A maximum concentration for CHD was realized in less than sixty minutes of reaction time. The rate constants for H/D exchange on the methylene carbon are: km = 1.9x10'4, k- 1 A = 0, k2A = 1.4x10'4, k-2A = 1.2x10'3 L/mol/min and the sum of the squares error (SSE) = 6.4x104. The same experiment using N,N-dimethylglycine was performed at 70 °C with 1000 psi D2 and D20. The 1H NMR of time zero and after four hours of reaction time is compared in Figure 80. Figure 8i shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. A maximum concentration for CHD was realized after about 180 minutes of reaction time and was much slower than either glycine or sarcosine at 70 °C. The rate constants for H/D exchange on the methylene carbon are: km = 1.3x10'5, km = O, k2A = 2.2x10'5, k-2A = 0 L/mol/min and the sum of the squares error (SSE) = 1.2x10'3. 200 1500 77--___ 7 - ___ 1250 :— ——————— "" +Time: 0 Hours 'Tlme 74 Hours 30 1 | 1250 "I - 7 -- 7*» 20 l ”_-... 77777 77 r 1000 I l 77“- r“ 10 I , 100074 . to ‘ ‘ i750 |3.43 3.42 3.41 3.4 3.39 3.38 I " 750 77 - 77 0777- 7 7 500 500 7 250 3.47 3.46 3.45 3.44 3.43 3.42 3.41 3.40 3.39 3.38 3.37 Figure 78. lH NMR of H/D exchange at the methylene position of 0.25 M sarcosine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Rqu) 0.30 777-7- 777-7777 7777 3.0 7AH2 7AHD 7ADZ I AH2_exp A AHD_exp o ADZ_exp —H20 and HOD 0 H20 and HOD_exp 0.25 77 7 777- - - 7 - 7 777—77 7 77 7 2.5 51 0.20 7 2.0 2 .. z: 5 a '- 7— 77 0 g o 15 1-5 2 8 8 S o 7 7 7 7 7 771.0 7 A — 7 77 0.5 A N v 7 l 1 00 150 200 250 300 Time (min) Figure 79. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 201 3000 777 777-77 7777 7777 7- 77 2500 +Time: 0Hours --+- Time: 4 Hours i 2500 ~ 7 2000 2000 4* 7~ 1500 1500 ~ 7 1000 1000 ~* ~ 500 500 0 -~ 7 7 - » -5oo 3.58 3.56 3.54 3.52 3.50 3.48 Figure 80. 1H NMR of H/D exchange at the methylene position of 0.25 N,N-dmg (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) 7AH2 7AHD —ADZ I AH2_exp A AHD_exp e ADZ_exp -9- H20 and HOD 0 H20 and HOD_exp 0.25 7 7 ~77 7 7 - 7 ~ 1.5 g 0.20 7 7 1.2 2 V I C a o c E 015 — — 7 7 7 77 0.9 3 h o C n. 8 E o 0.10 7 7 7 7 7 7 rt 0.6 A 0.05 7— 77 -- 7 7 7 7 M777: 7 77 7* 0.3 K/ e 4/ __fl/// 0.00 3 r v 7 .7 , O r I o 0 50 100 150 200 250 300 Time (min) Figure 81. Fit of the kinetic rate data for 0.25M N,N-dmg (D20, 70 0C, 1000 psi D2, 1 g dry 5% Ru/C) 202 Figure 82 compares the rate constants for glycine, sarcosine, and N,N-dimethylglycine on the methylene carbon at 70 °C, 1000 psi D2, and D20 (reaction time = 6 hrs). For the first hydrogen exchanged, the rate constant, km, for sarcosine is the largest, but is comparable to glycine. The km rate for N,N-dimethylglycine is significantly less (more than a factor of ten) than the rate for either glycine or sarcosine. After the first hydrogen is replaced with deuterium the rate constants representing exchange of the second hydrogen with deuterium are less for both glycine and sarcosine but are comparable for N,N-dimethylglycine (k; A compared to k2A). 2.5E-04 I Glycine I Sarcosine NNdmg 2.0E-04 7 1.5E-04 3 1.0E-O4 - 7 7 Rate constant (Umollmln) 5.0E-05 7 0.0E+00 7 Figure 82. Rate constants for H/D exchange at the methylene carbon (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) A comparison of rate constants for the temperatures 100 °C and 70 °C with 1000 psi D2 for glycine, sarcosine, and N,N-dimethylglycine will be presented at the end of this section. 203 Using the more constrained model (6-variable) the rate constants k1 A, k-1A, k2A, and k-2A were compared to the computed rate constants obtained from using the lO-variable kinetic model. And, for the forward reactions the rate constants k] A and k2A are essentially identical using either model (as mentioned previously). The rate constant and primary and secondary isotope effects were calculated for the experiments with 1000 psi I)2 (D20, 70 °C, 1 g dry 5% Ru/C) and for glycine km = 0.74x104, pA = 0.03, and sA = 1.5. For sarcosine kiA = 9.9x10'5 L/mol/min, pA % 0.24, and sA = 0.72. For N,N- dimethylglycine kiA = 5.9x10'6 L/mol/min, pA = 1.0, and sA = 0.40 (Figure 83). The rate constants, ki, for glycine, sarcosine, and N,N-dimethylglycine are roughly half of k1 A using the lO-variable model. Again this makes sense because kiA is defined to be half of k1 A. The secondary isotope effects, SA, are between 0.4 and 1.5. 204 1.2E—04 1.8 EIKi Oprimary isotope effect __ 1 6 _ Aseconda isoto effect ' 1.0504 _ A 9.9E 05 W P6 E g 8.0E-05 - E A a E E 6.0E-05 - 5 5 E g ._ o X o 3 4.0505 - E 2.0E-05 7 0.0E+00 7 Glycine Sarcosine Figure 83. Kinetic rate constant and primary and secondary isotope effects for glycine, sarcosine, and N,N-dmg (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) The rate constants for glycine, sarcosine, and N,N-dimethylglycine can be compared for the reactions at 100 °C and 70 °C with 1000 psi D2. Additional experiments were conducted at 50 °C and 90 °C, but not for all three substrates. So, the individual substrates will be compared at each temperature at the end of this section. The comparison data of the rate constants is in Figure 84. All of the rate constants are lower at 70 °C, which was expected. 205 3.5E-04 7 IGlycine 1OOC_DZ DGlycine 7OC_DZ lSarcosine 1OOC_DZ E Sarcosme 7OC_D2 3.0E—04 — mNNdmg1ooc_02 DNNdmg 7OC_D2 W i 7— g E 2.55-04 7 E a o 5 2.0504 « 2' S u g 1.5E-04 , o 999 . 999 999 999 999 E 999 999 1 ()5 04 . :3: :3: ' ' 999 999 999 , 999 999 999 999 999 3:: 3:: 5.0E—05 i 999 999 999 999 999 999 999 999 999 999 999 999 . 0.0E+00 7 Figure 84. Rate constants for H/D exchange at the methylene carbon at 100 °C or 70 °C (D20, X °C, 1000 psi D2, 1 g dry 5% Ru/C) H/D Exchange Studies at 50 °C with 1000 psi D_z The reaction temperature was lowered to 50 °C and the same experiments were individually performed using glycine and sarcosine with 1000 psi D2 and D20. The substrate N,N-dimethylglycine was not evaluated at this temperature because of the low reactivity observed at 70 °C. The singlet lH NMR peak representing the methylene (CH2) position continued to decrease in size over time, while the CHD grew in albeit slowly. This reaction was significantly slower than at 70 °C and 100 °C. Figure 85 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methylene carbon are: klA = 5.5x10'5, 1cm = 8.1x10'3, k2A = 3.9x10'5, k-2A = 5.6x10'3 L/mol/min and the sum of the squares error (SSE) = 6.0x10'4. 206 —— AH2 —— AHD — ADZ l AH2_exp A AH D_exp 9 ADZ_exp --H H D o H H 0,25 . -flggand 0 20 and 003w ._ 25 5‘ 0.20 « 2.0 o E V :I: 8 9a 3 0.15 i 1 5 § In 5 & u 8 o 0.10~~-—7v 7- 1.0 0.05 7- — 7» 0.5 A 0.00 l I T r Y Y 0.0 0 50 100 150 200 250 300 350 400 Time (mln) Figure 85. Fit of the kinetic rate data for 0.25 M glycine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) An experiment with sarcosine at the same conditions (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) was performed and results again showed a slower rate of H/D exchange. Figure 86 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methylene carbon are: km = 5.8x10'5, k-1A = 6.4x10'4, k2A = 4.3x10'5, k-2A = 3.4x10'3 L/mol/min and the sum of the squares error (SSE) = 4.9x104. 207 0.25 »-- - ..._.__._~__.s-.~-.___--_.....-c__....- .W- _. --- _ 20 —AH2 ———AHD ——A02 i I AH2_exp A AHD_exp 9 ADZ_exp —H20 and HOD 0 H20 and HOD_exp 0.20 7— — ————- —— — ., J7 __,__ 7 a 1.6 3 = g 0.15 .- . .— ~- 1.2 I g A A g f— ' 1 0. § 0.10 4 - l» 0.8 c o O \ 9 0.05 7 ——~— J —— 7— 0.4 ’ \- . 9 0,00 I I I I I I I 0_0 0 50 100 150 200 250 300 350 400 Time (mln) Figure 86. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) The kinetic rate constant for the H/D exchange on the methylene carbon will be compared to the rate constant for the H/D exchange on the methyl carbon at all three reaction temperatures in Figure 111. In summary, we can compare the rate constants for glycine, sarcosine, and N,N- dimethylglycine at 100 °C, 70 °C and 50 °C with 1000 psi D2. The rate constants at 100 °C were calculated from the linked H2 and D2 experimental data (and the kinetic model includes kH). Using this set of data, activation energies were calculated. The BA for the methylene carbon of glycine and sarcosine was 45 kJ/mol and 36 kJ/mol, respectively. The activation energy for N,N-dimethylglycine was calculated based on two 208 experimental points only, and was equal to 88 kJ/mol. Figure 87 presents km for glycine, sarcosine, and N,N—dimethylglycine at 100, 70, and 50 °C. 1 .8E-04 I Glyci ne__kiA_C H2 Sarcosine_kiA_CH2 1 .5E-04 1.25-04 ~ 9.0E—05 “-7— 6.0E-05 7r“ , Rate constant. klA (Umollmln) 3.0E-05 7— ' 0.0E+00 - .5... I N,N-dmg_kiA_CH2 L-_,_ -. J—d 50 °C Figure 87. Kinetic rate constant for glycine, sarcosine, and N,N-dmg (D20, X °C, 1000 psi D2, 1 g dry 5% Ru/C) Again, it was observed that the rate constant for sarcosine is the largest and it falls off by about one-third for glycine and one-half for N,N-dimethylglycine at 100 °C. The same general trend is observed at lower temperatures. The H/D exchange reaction is presumably promoted by having an electron-rich C-H bond. If the pKa's are examined for the three substrates (Table 8), the carboxylic acid of sarcosine is the most acidic and may result in sarcosine favoring the zwitterionic form more than glycine favoring the zwitterionic form. For H/D exchange to occur, the amino must be free, so this may suggest a slower rate for sarcosine than that of glycine. But, 209 sarcosine has a more electron-donating amino group as compared to that of glycine which would promote the H/D exchange reaction. Table 8. pKa values for glycine, sarcosine, and N,N-dimethylglycine (25 °C) Substrate 'NHX(CH3)Y(+) -COOH Glycine (x=3, y=0) 9.6 2.34 Sarcosine (x=2, y=1) 10.01 2.23 N,N-dimethylglycine (x=1, y=2) 9.8 2.08 Mass Transport Considerations It was noticed that for the H/D reactions at 100 °C, the exchange is occurring quickly. So, a brief discussion on mass transport limitations is included. From our hydrogenation experiments it was determined that the overall rate of reaction was proportional to the catalyst loading (based on experiments with half and double the catalyst amount) and thus there are no external mass transport limitations in the system. This can be assumed to be true for the H/D exchange reactions as well. For our system with H2 and D20, the reaction rate should be relatively insensitive to catalyst loading because the flux of H2 from the gas to liquid would define the reaction rate. But, it is still possible to have internal mass transport limitations (even if the rate is proportional to the catalyst loading). This is because the catalyst will have the same rate per unit mass even if there are concentration gradients within the catalyst particles. lntemal mass transfer limitations occur when the Thiele modulus for the reaction is greater than 0.3. For an n-th order reaction, the Thiele modulus, (D, is equal to 210 L(k*Co(n+l)/2/Dc)0'5, where L = catalyst dimension (W3 for a sphere), k = rate constant, Co = bulk fluid concentration, and D6 = effective diffusivity of reactant in the catalyst. The H/D reactions are essentially second-order but will be treated as first-order in the D2/D2O case due to the large molar excess of deuterium. So, the equation is simplified to (D = L(k/De)0'5. The effectiveness factor for intraparticle resistance is the ratio of the actual rate to the true kinetic rate. The rate constant in the equation for (I) is the true or intrinsic rate constant, which cannot be determined experimentally until we are sure there are no mass transfer limitations. It can be determined using the observable modulus. The effective diffusivity of glycine can be found from the Wilke-Chang equation. For the experiment with glycine at 70 °C in water, the diffusivity was equal to 3.5x10'5 cmZ/s. The effective diffusivity in the catalyst pore takes into account the pore fraction, so Defflgly = epsz"'Dg|y = 1.2x10'9 mz/s (based on eps = 0.59). The initial bulk concentration is 0.25 M = 0.25 kmol/m3 but will vary throughout the reaction. To get the observable modulus, the rate of reaction Rg = k*gly must be expressed in units mol gly/cm3 catalyst/s (rate per unit catalyst volume). Using the data for the experiment with glycine with 1000 psi D2 and 70 °C, the slope from the concentration (mol/L) versus time was multiplied by the volume of the reacting phase and dived by the volume of catalyst in the reactor (volume = mass catalyst/density of catalyst). For this experiment the rate equaled 5.3x10'3 kmol/m3*cat/s. 211 The best way to assess the presence of mass transport is to use the observable modulus (11*(D2 = Rg*L2/Co/De) and if this (11* (D2) is greater than about 0.3, then mass transport limitations in the catalyst particles during exchange need to be considered. For the experiment with glycine the observable modulus is equal to (5.3x10'3 kmol/m3/s)*(2.5x10'5 m2)/(0.25 kmol/m'i')/(1.2x10'9 mz/s) = 0.01. So, the conclusion is that there does not appear to be serious mass transfer consideration in the H/D exchange experiments. This is primarily because we are using such a small catalyst particle. An additional experiment with 0.25 M glycine was performed at 90 °C with 1000 psi D2 and D20. The singlet lH NMR peak representing the methylene (CH2) position continued to decrease in size over time, while the CHD grew in, reaching a maximum peak height between 30 and 60 minutes of reaction time. Figure 88 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methylene carbon are: k1A = 1.9x10'4, in“. = 1.0x10'5, k2A = 1.3x10'4, k2,. = 4.8x10‘4 L/mol/min and the sum of the squares error (SSE) = 3.2x10'4. 212 025 _.______-__.___._ _ .-. -_-" _ __.__. _ _. -___ _-..._....-__...-..._ .. p 5 —AH2 -—-*AHD ——ADZ I AH2_exp A AHD_exp 9 ADZ_exp —H20 and HOD 0 H20 and HOD_exp ’ 0.20 ~ ~ - -, ~ ~ ,-, r 4 Q E 0.15 « — — .. 3 I g e E E v 9 . § E 010 ‘ g e O 4 \ A \ \ 0.05 ~ . -—- \\ _,__ ~ , _J .- 1 I \\ A 0.00 r r T _F‘M. . 0 0 50 100 150 200 250 300 350 Time (min) Figure 88. Fit of the kinetic rate data for 0.25 M glycine (D20, 90 °C, 1000 psi D2, 1 g dry 5% Ru/C) An experiment with 0.25 M betaine ((CH3)3N+CH2COO') at 130 °C was conducted to evaluate the extent of H/D exchange at the methylene position (Figure 12). Results at 130 °C with 1000 psi H2 for six hours revealed 86% conversion of CH2 to both CHD and CD2. Figure 89 shows the experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methylene carbon are: k; A = 4.1x10'5, k-.A = 0, k2A = 1.6x10'2, k-2A = 2.2x10'l L/mol/min and the sum of the squares error (SSE) = 1.9x10'3. 213 0.30 ~-—----——--- —« --- . -_._ .. -.____..._. ....___.___ . __ . -.__-__.______._ --...-Hm _-_ 18 —- AH2 —— AHD — A02 I AH2_exp A AHD_exp 9 ADZ_exp —H20 and HOD 0 H20 and HOD_exp 0.25 . 777 , .___ — d 15 .2: 0.20 12 .. :r: 3 e .l O B 0 15 9 2 o E :L 8 0.10 , J~ 6 0.05 T 4.- 3 0.00 . r . . . . . e o 0 50 100 150 200 250 300 350 400 Time (min) Figure 89. Fit of the kinetic rate data for 0.25 M betaine (D20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) Isotopic exchange at the two methylene positions on glycylglycine was evaluated using 0.25 M glycylglycine at 100 °C with 1000 psi H2 and D20. Position A, the more downfield 1H NMR signal, exchanged much more rapidly than position B (Scheme 31). B o 3 63 ppm H H2N/\”/ A OH O 3.67 ppm Scheme 31. lH NMR positions for the methylene hydrogens on glycyl glycine From the lH NMR results, it appeared as if both hydrogens at the methylene position exchanged at the same time, rather than step-wise. In most of the other isotopic exchange 214 experiments, the NMR results indicate the disappearance of the singlet and the appearance of the triplet over time. However, in this case the singlet (3.67 ppm) disappears quickly and without detectable triplet formation. The methylene position corresponding to the singlet at 3.63 ppm does appear to form CHD, as illustrated by the triplet formation. A . i Reaction time 2L - Figure 90. 1H NMR of H/D exchange at the methylene positions of 0.25 glycylglycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C Again, the amount of triplet formed is quite small. Presumably much of the reason for the decrease in both 3.67 and 3.63 ppm methylenes is due to the formation of glycine. The results every hour over a six hour period are displayed in Figure 90. Hydrolyzed glycine is formed over time (3.36 ppm) and undergoes H/D exchange as it grows in (Figure 91). 215 i Reaction time n Figure 91. 1H NMR of H/D exchange at the methylene position of hydrolyzed glycine from 0.25 M glyclglycine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 4.7.1.2 H/D Exchange at the N-Methvl Position In this section H/D exchange at the N-methyl positions will be examined. At the end of this section a comparison will be made between the rate constants at the methylene and methyl carbons. Again, the following chemical reactions (Chapter 3) can be used to calculate rate constants klc, k-1c, k2c, k2c, k3c, and k-3c. As mentioned in Chapter 3, the equilibrium between the hydrogens on the catalyst surface and the hydrogen concentration in the 216 liquid is not reached at time zero for the experiments with 1000 psi H2 and D20, so the hydrogen concentration in the liquid is used. The experiment describing the hydrogen-deuterium exchange on the methyl carbon is depicted by three reversible, consecutive first-order reactions (Chapter 3). The amino acid substrates which were studied included sarcosine (N -methylglycine), N,N- dimethylglycine, and betaine. An experiment with 0.25 M sarcosine at 100 °C with 1000 psi H2 and D20 was run for six hours. The singlet 1H NMR peak representing the methyl (CH3) position continued to decrease in size over time, while the triplet peak representing CH2D grew in, reaching a maximum peak height at 120 minutes of reaction. A quintet peak representing CHD2 grew in, reaching a maximum peak height at 180 minutes of reaction time. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 92. Figure 93 shows the experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: krc = 1.1x104, k-1c = 0, k2c = 1.45.10“, k-2c = 2.5x10'6, k3c = 8.3x10'5, and k.3C = 2.0x10'4 L/mol/min. An experiment with 0.25 M N,N-dimethylglycine at 100 °C with 1000 psi H2 and D20 was run for six hours. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 94. Figure 95 shows the experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: klc = 1.0x10'4, lclc = 0, k2c = 1.4x10'4, k-2c = 1.8x10‘4, log = 1.5x10‘4, and k-3c = 5.8x10'4 L/mol/min. 217 4000 f— _ _ _ _ _ _— __ __.__ 3000 25 +T'Ime: 0 Hours 7G77TIme: 6 Hours 7 2500 7 7 2000 12.54 2.53 2.52 2.51 2.50 2.49 "‘ “___ ___“ * 150° 2000 7—————— - 7 - 77 7 7~— - —— ~ 1000 1500 77 7r 500 1000 0 50077—777 7w—7—7 — ———— ——-——- ——--~-5oo : ------------------- -| 0 i ___ —-———-——L-— _________________ ’ b -1000 2.58 2.56 2.54 2.52 2.50 2.48 Figure 92. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 0.30 18 — CH3 —— CH2D — CHDZ —CD3 I CH3_exp A CH20_exp 9 CHDZ_exp e C03_exp —H20 and HOD 0,25 . e H20 and HOD_exp 1 15 2 0.20 77—7- —- 7 —— 7777 —— —7777 77 - 7 7 77 12 o 5 = as o c E 015 7 —— 7 7 7 — — 7 — . 7 7 9 § § ’ a 8 0.10 - e — ° .. 6 ‘ \i 0.05 7 7 777 7I-7 ~ - 7 - - 7 7 ~— 77 77 3 O I N 0.00 - e r r r l . 4 . I o 0 50 100 150 200 250 300 350 400 Tlme(m|n) Figure 93. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 218 4000 3500 l" — _ _ _ "‘ " “'5'“ +11me: 0Hours +Time: 6Hours 3500 J " 23 l 7 7 3000 1_. 3000 7| 7 77 2500 250° ii 2.73 2.71 2.69 W ‘“ 2000 2000 7777 7 1500 1500 77 7 77 7 7 1000 1000 77 7 — 500 500 . o I ------------- 'I o l ' -5oo 2.82 2.80 2.78 2.76 2.74 2.72 2.70 2.68 2.66 2.64 Figure 94. 1H NMR of H/D exchange at the methyl position of 0.25 M N,N—dmg (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 0.30 18 777- CH3 777— CHZD — CHDZ —coa l CH3_exp A CH2D_exp 9 CHD2__exp e CDa_exp --H20 and HOD 025 l 0 H20 and HOD_exp —“ 15 e 3 e = 0.20 -- 77 7 7 77 7 7 7 7 7 7712 E .. I 8 ‘e =1 0.15 7 7 7 7 7 77 9 3 E o 3 : n. E . - ,0 .- 8 0.10 6 O 0.057 7A7 7 7777 7 777 7773 l I 9 I 0.00 . . . . E . - o 0 50 100 150 200 250 300 350 400 Time (min) Figure 95. Pit of the kinetic rate data for 0.25 M N,N—dmg (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 219 The same pair of experiments were performed at 100 °C but with 1000 psi D2. An experiment with 0.25 M sarcosine was run for six hours. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 96. Figure 97 shows the experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: kic = 1.6x10‘4, km 7 0, k2c = 1.9x10'4, k-2c = o, k3C = 1.5x10'4, and R. 3c = 8.5x104 L/mol/min. The H/D exchange on the N-methyl of sarcosine was slower than the methylene (in general, for all conditions) and did not fiJlly reach equilibrium after six hours. The statistical equilibrium end points were again compared to the experimental data (for the experiment with 1000 psi D2). The statistical concentrations of CH3, CH2D, CHD2 and CD3 based on the actual percent hydrogen in water at six hours (1.76%) and the concentration at time zero (0.22 M) are 1.2x10'6, 2.0x10'4, 0.0113 and 0.210 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 2.1x104, 1.0x10'3, 0.027 and 0.22 M, respectively. 220 6000 777-777777 777----7-77 777—7777r 5000 +Time: 0 Hours +11me: 6 Hours 5000 7 7 7 4000 4000 77 7 7 77 7 3000 3000 7 7 2000 2000 7 -~ - 7 1000 1000 0 0 I __________ f: f _______ . -1000 2.55 2.54 2.53 2.52 2.51 2.50 2.49 2.48 2.47 Figure 96. 1H NMR of H/D exchange at the methyl position of 0.25 M sarcosine (D20, 100 °C,'1000 psi D2, 1 g dry 5% Ru/C) 0.30 777777 777777-777-77-77- 7 3.0 777-CH3 777CHZD —CH02 —CDB I CH3_exp A CHZD_exp I CHDZ_exp e CDS_exp —H20 and HOD 0-25 e H20 and HOD_exp ‘“ 2-5 E 0.20 7 --77 77 ' —- 7 77— 77 7777 7 77 —7 77 2.0 e 5 x 5 .. 27: E «h i 2 e e. 8 7 77 1.0 77 77 0.5 “fig-"__n 00 400 Time (min) Figure 97. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 221 An experiment with 0.25 M N,N-dimethylglycine was run for six hours. The 1H NMR of time zero and after six hours of reaction time is compared in Figure 98. Figure 99 shows the'experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: kic = 1.9x10‘4, kic = 0, k2c = 3.2x104, k-2c_ = O, k3c = 2.8x104, and k-3c = 8.8x10'4 L/mol/min. The H/D exchange on the N-methyl of N,N-dimethylglycine appeared to just approach equilibrium after six hours. The statistical equilibrium end points were again compared to the experimental data. The statistical concentrations of CH3, CH2D, CHD2 and CD3 based on the actual percent hydrogen in water at six hours (2.7%) and the concentration at time zero (0.15 M) are 2.9x10'6, 3.1x104, 0.011 and 0.134 M, respectively. The actual concentrations of CH2, CHD, and CD2 at six hours of reaction time are 6.7x10'6, 1.3x104, 0.0041 and 0.196 M, respectively. 222 3500 -~— 7 ._.T 3000 +Time: 0Hours I'— +Timez 4 Hours 3000 77 7—7 777777- - 7 I 7 2500 2500 777—7 7 77 - -——7 -- I r 2000 1 I268 2.67 2.66 2.65 2.64 I 200077- 77 77‘ 7777-— L __ _ _ _ _ _ _ _ __ 71500 1500 7* 1000 1 500 2.71 2.70 2.69 2.68 2.67 2.66 2.65 2.64 2.63 Figure 98. 1H NMR of H/D exchange at the methyl position of 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Rqu) 777-CH3 —-—CHZD —-CHDZ i ——CDB I CH3_exp A CHZD_exp 9 CHDZ_exp e CD3_exp —H20 and HOD 0.20 g 0 H20 and HOD_exp e g 4.0 2.7 a ~€0.15777—77—777777 7 77 7 7 7777777— 7 77 3'0: e = e. 8 0.10 7 -———————— 7 7 77 777 —- 77—7 7777777 77 77 -7—— 77 — —— — 7777 2.0 C o o 0.0577 77 - 777 -—- 77 77 -77 7777 7 7 7 7 7 771.0 4 9 9 . A 1 v 0.00 . . L 7W. . I 00 O 50 100 150 200 250 300 350 400 Tlme(m|n) Figure 99. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2, 1 g dry 5% Ru/C) 223 Again, as in the case of the methylene carbon, the kinetic model can be redefined using three independent variables, kic, pC, and SC instead of six. As discussed in Chapter 3, kic is equal to 3*kic indicating that there are three hydrogens on the methyl carbon and subsequently three opportunities for H/D exchange. The equations for how the k 1c, etc rate constants could be effectively calculated from the kic, pc, and SC values are discussed in Chapter 3. It was assumed pc and sc are independent. For the reactions at 100 °C with 1000 psi H2 one k (ki) value for each site that is modified via an isotope effect was calculated. The results show that sarcosine and N,N- dimethylglycine the primary isotope effects are similar and about 0.7-0.8. For sarcosine kiC = 0.40x10'4 L/mol/min, pc = 0.71, and sc = 0.47. For N,N-dimethylglycine kic = 0.35x10‘4 L/mol/min, pC = 0.83, and SC = 0.52. The secondary isotope effects for sarcosine and N,N-dimethlyglycine are similar and equal to about 0.5. The trend for the rate constants, kg, for sarcosine and N,N-dimethylglycine match well with be using either the 10 or 6-variable model. The ki values for both sarcosine are N,N-dimethylglycine are about 1/3rd that of km, respectively. Again this makes sense based on the fact that kic is defined as being equal to 3*kiC- For the reactions at 100 °C with 1000 psi D2 one k (k,-) value for each site that is modified via an isotope effect was calculated. For sarcosine kg = 0.55x10'4 L/mol/min, pc = 0.45, and sc = 0.59. For N,N-dimethylglycine kic = 0.69x10'4 L/mol/min, pc = 0.87, and sc = 0.47. The secondary isOtope effects for sarcosine and N,N-dimethlyglycine are similar 224 and equal to about 0.5. Again, the ki values for both sarcosine are N,N-dimethylglycine are about 1/3rd that of kic, respectively. As discussed in the IUD exchange on the methylene carbon section, the data from the experiments with 1000 psi H2 and D2 can be linked in order to define a common set of rate constants. Figure 100 presents the comparison of the kinetic fits using the k6 + kH model with the data from the sarcosine experiment with 1000 psi H2 and also from the linked data set (1000 psi H2 and D2). Figure 101 is the analogous plot showing the comparison of the kinetic fits with the data from the experiment with 1000 psi D2 and also from the linked data set. 225 0.30 Concentration (moIIL) 0.05 0.20 777- - I CH3_exp 9 CHDZ_exp -—CH3_H2 —CHDZ_H2 7 CH3_linked H2 and 02 77 7 CHDZ_Iinked H2 and 02 a CH20_exp e CD3_exp —CHZD_H2 -—CD3_H2 CHZD_Iinked H2 and 02 7 7C03_linked H2 and 02 0.00 Time (min) Figure 100. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) 0.30 0.25 7— 9 N o J I CH3_exp 9 CHDZ_exp CH3_DZ CHDZ_02 7 CH3_linked H2 and DZ ----- 7 CHDZ_Iinked H2 and DZ Concentration (mol/L) O a a CH20_exp l e CD3_exp i CHZD_DZ coa_02 ~ 7 CHZD_linked H2 and 02 ..—7 -'“" ~~~~ .th.’ o ‘1 ’- , 200 Time (min) 250 300 400 Figure 101. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) 226 The fits from the linked experimental data (1000 psi H2 and D2) is comparable to those from the individual experiments. For the experiment with 1000 psi H2, the fit is slightly faster using the linked data. This result again gives credence to the common rate constants. Figure 102 shows the kinetic rate constant and primary and secondary isotope effects for the H/D exchange on the methyl carbon of sarcosine. The rate constants for the H/D exchange on the methyl are slower than those for the methylene (kiA = 1.5x10“1 versus kiC = 0.58x10'4 L/mol/min). The primary and secondary isotope effects are fairly close pA = 1.08 versus pc = 0.82 and SA = 0.49 versus sc = 0.63. Again, there is not much significance to the primary isotope effect because the back reaction is not occurring under these conditions. Figures 103 through 105 present the results for N,N-dimethylglycine. The rate constants for H/D exchange on the methyl are slower than those for the methylene (km = 7.4x10'5 versus kg = 6.6x10'S L/mol/min). The primary and secondary isotope effects were compared for the methylene and methyl carbon. As with sarcosine, they are very close pA = 0.52 versus pC = 0.68 and SA = 0.60 versus SC = 0.48. The secondary isotope effects for both sarcosine and N,N-dimethylglycine at the methylene and methyl carbon are about 0.6 (note: glycine SA = 0.81). 227 3.515704 16 EKiC oprimary isotope effect ._ 1 4 3'0E'04 l Asecondary isotope effect 7* 1.2 E 2.5E-04 7 E 3 77 1.0 5 2.05704 - ’9‘ g . 9 71 0.3 g 8 155-04 7 '5‘": o O % 0‘ 1.05-04 7 5.0E-05 ' 0.0E+00 7 H2 DZ Linked H2 and DZ Figure 102. Kinetic rate constant and primary and secondary isotope effects for sarcosine with 1000 psi H2 or 1000 psi D2 or H2 and D2 0.30 I CH3_exp A CH2D_exp 9 CHDZ_exp o CDa_exp ——CH3_H2 CHZD_H2 0-25 I —CHD2_H2 —CD3_H2 * \ - CH3_linked H2 and DZ . CHZD_Iinked H2 and DZ \\ 77 7 CHDZ_Iinked H2 and D2 - CD3_linked H2 and 02 g 0.207 7 7 ‘777 7 7777 777 7777 777 7- 7 7 7 7 77 7 o 7. E \ s 320157—77 1-7 7 7 7 7 7 7 7 7 7 7 7- S ,9 C x 0 \‘ ’ _____ 0 \ s 72 . - 7 00107777 777 \7 7 77:— 77 7 _fl 7 /’ I. I — K >“ it”- I / / F \77\___ 0.05-7777 7 A I 7 -7 7 7 77.7.; 77 77 7 7 7 7 // I”, i FFFFFFF xx ’ . 7" F \ \\\\\\ - F“‘7777 7 __ ...A 0.00 . . . . I: . . 0 50 100 150 200 250 300 350 400 Tlme(m|n) Figure 103. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi H2 or linked H2 and D2, 1 g dry 5 % Ru/C) 228 0.30 I CH3_exp A CHZD_exp o CHDZ_exp O CDS_exp 0 25 ‘_ CH3_02 CHZD_DZ ' —CHDZ_DZ —CDS_DZ 7 7 CH3_llnked H2 and DZ 77 7 CHZD_linked H2 and 02 . 7 7 CHDZ_linked H2 and Dz 77 CDa_linked H2 and DZ 5! 0.207 7 7 7777 77 777 77 77 7 77 7 . 7 o g - u s E 0.15i77 7 77 7 . 7 7 77 7 7 7 7 E *’ ' 8 7’ C / 80.107 7 7 7 77 7 7 7 777 7 77 7 7 ./ / 0.05 77 7 7 i 7777\\ ‘ 7 7 77 7 77 77 77 7 7 7 4 77\_ “*7 ,/ '7. \\“797\ 1 x “7 ¥ ~ , 0.00 r , - . - 7 _ , - .7. i 0 50 1 00 1 50 200 250 300 350 400 Tlme(m|n) Figure 104. Fit of the kinetic rate data for 0.25 M N,N-dmg (D20, 100 °C, 1000 psi D2 or linked H2 and D2, 1 g dry 5 % Ru/C) 1.0E-04 14 El KIC 0 primary isotope effect 1 2 A secondary isotope effect ' 8.0E-05 4 7.0E-05 -t ,5 6.6E-05 1'0 E 3 g 6.0E-05 7 A J -- 0.8 g 2' :I: g 5. g 7— 0.6 E 8 4.0E-O5 7 2 N I! 7 0.4 2.0E-05 7 - 0.2 0.0E+00 7 - 0.0 H2 DZ Linked H2 and D2 Figure 105. Kinetic rate constant and primary and secondary isotope effects for N,N- dmg with 1000 psi H2 or 1000 psi D2 or H2 and D2 229 A third pair of experiments was conducted at 70 °C with 1000 psi D2 and D20. An experiment with 0.25 M sarcosine was run for four hours. Figure 106 shows the experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: km = 7.4x10'5, k-1c = O, k2c = 9.4x10'5, k-2c = 2.1x10'3, k3c = 1.7x10’4, and 1m =6.Ox10'3 L/mol/min. 0.30 7777-- 3.0 7CH3 7CHZD 7CH02 1 7003 I CH3_exp A CH20_exp 0.25 _ o CHDZ_exp o CDS_exp —H20 and H00 _2_ 2.5 0 H20 and HOD_exp ’57 0.20 J7 2.0 E v :I: 5 E 1 A 7 O E O 5 15 g = o. 8 S 0.05 7 77 0.5 0.00 r r w r 0.0 0 50 100 150 200 250 300 Time (min) Figure 106. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) An experiment with 0.25 M N,N-dimethylglycine was run for four hours. Figure 107 shows the experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: klc = 3.0x10'5, k-lc = 0, k2c = 1.1x10'4, k-2c = 1.1x10'2, k3c = 5.9x10'5, and k-3c = O L/mol/min. 230 0.2 5 -....- WW .. __ . _ . ._-- «WWW WW .... .. ...___- ___-..__.__--_....._.. _. ._ .. . __._~W___.-____.-.__.._. .. . .. T 2_ 5 —CH3 ——CHZD -—CHDZ 7C03 I CH3_exp A CHZD_exp o CHD2_exp e CDB_exp -—H20 and HOD 0 H20 and HOD_exp 0.20 i _s 0 Concentration (molIL) .0 a: / 3. Percent H 0.05 7 7 77 77 0.5 0.00 1 r 1 . . 0.0 o 50 100 150 200 250 300 Time (min) Figure 107. Fit of the kinetic rate data for 0.25 M N,N-dimethylglycine (D20, 70 °C, 1000 psi D2, 1 g dry 5% Ru/C) For the reactions at 70 °C in 1000 psi D2 we calculated one k (ki) value for each site that is modified via an isotope effect (Figure 108). The results show that for sarcosine and N,N-dimethylglycine the secondary isotope effects are both about 0.5 (similar to the experiments discussed above). For sarcosine kg = 0.24x10“1 L/mol/min, pC = 1.5, and sc = 0.56. For N,N-dimethylglycine kic = 8.7x10'6 L/mol/min, pC = 1.0, and SC = 0.38. Again, the ki values for sarcosine and N,N-dimethylglycine are about one-third of their respective klcs. 231 3.0E-05 2.0 EIKi 0 primary isotope effect Aseconda isoto effect 2.5E-05 - ”I ”6 1.6 .‘s‘ g 2.0E—05 - 3 E 1.2 A a 2 E 1.5E-05 7 g g “2‘ u -7 0.8 § 1.0E-05 7 I! 77 0.4 5.0E-06 7 0.0E+00 7 - 0.0 Sarcosine NN-dmg Figure 108. Kinetic rate constants and primary and secondary isotope effects for sarcosine and N,N-dimethylglycine; 70 °C, 1000 psi D2 We can compare the reactions in 1000 psi D2 at both 70 °C and 100 °C for the H/D exchange on the methylene carbon and N-methyls of sarcosine and N,N-dimethylglycine (Figure 109). Again for sarcosine the rate of exchange for the methylene hydrogens is slightly larger than for the methyl hydrogens. Whereas in the case of N,N- dimethylglycine, the methyl hydrogens exchange at a slightly faster rate than the methylene hydrogens. Again, this is partly due to the fact that with N,N—dimethylglycine there are twice as many N-methyl hydrogens available for H/D exchange. As seen previously, the rate constants calculated in the reaction at 70 °C are slower than the experiment with otherwise identical conditions at 100 °C. 232 1.6E-04 I kiA CH2_70 1.45704 I kiA CH2_100 u kiC CH3_70 1.2E-04 ? akic CH3_1OO :5 -° 1.05-04 E a ‘E 8.0E-05 a n I: 8 6 0E 05 8 , N 0: 4.05705 2.05705 o.oe+oo Sarcosine NN-dmg Figure 109. Rate constants for H/D exchange at 100 °C or 70 °C (D20, X °C. 1000 psi D2, 1 g dry 5% Ru/C) An additional experiment was done with sarcosine at 50 °C with 1000 psi D2 and D20. Figure 110 shows the experimental and fitted concentrations of CH3, CH2D, CHD2 and CD3 over time as determined by the equations in Chapter 3. The rate constants for H/D exchange on the methyl carbon are: klc = 1.6x10'5, k-1c = 3.lx10'5, k2c = 1.6x10'5, k-2c = 1.6x10'6, k3C = 3.2x10'5, and k_3c = 2.6x10'3 L/mol/min. 233 0.30 71.8 -- CH3 --- CHZD — CHDZ 7003 I CH3_exp A CHZD_exp 025 . o CHDZ_exp e CD3_exp —H20 and HOD __ 1 5 ' 0 H20 and HOD_exp ' g 0.20 . — .77 1.2 v 0 I c a o c E 0.15 i 7 ~77 7 7 09 § :: \ 8 § \- 3 0.10 -77 —7 7—7- 7 7 77— 7 7 777 7 - 0.6 0.05 7 -7 - 7 7 7 0.3 W 0_00 ‘ r r r f r 1 r 0 “me (mln) Figure 110. Fit of the kinetic rate data for 0.25 M sarcosine (D20, 50 °C, 1000 psi D2, 1 g dry 5% Ru/C) We can compare the rate constants for the H/D exchange at both the methylene carbon and the N—methyl carbons for the experiments with sarcosine at 100 °C, 70°C, 50 °C with 1000 psi D2. Figure 111 compares km and he The activation energy was calculated for this set of experiments and determined to be EA = 67 kJ/mol (note the EA for the H/D exchange on the methylene carbon was calculated to be 36 kJ/mol). 234 1.8E-04 I kiA CH2 I kiC CH3 1.5E-04 7 1.2E-O4 7 9.0E-05 7 6.0E-05 7 Rate constant (Umoiimin) 3.0E-05 7 0.0E+00 7 100 70 Reaction Temperature (’C) Figure 111. Rate constants for H/D exchange of sarcosine (D20, 50, 70, or 100 °C, 1000 psi H2 or D2, 1 g dry 5% Ru/C) In summary, the kinetic rate constants from the 6-variable model can be compared for the three substrates at the reaction temperatures of 50, 70, and 100 °C (Table 9). An experiment with 0.25 M betaine ((CH3)3N+CH2COO') at 130 °C was conducted to evaluate the extent of H/D exchange at the N-methyl positions. Results at 130 °C, 1000 psi H2 for six hours reveal no H/D exchange at the N-methyl positions (Figure 112). The integration value for the N-methyls differed by about 6% comparing time zero and after six hours. This is within the expected value for adsorption onto the Ru catalyst. 235 Table 9. Summary of the rate kinetic rate constants and primary and secondary isotope effects for glycine, sarcosine, and N,N-dimethylglycine Glycine Sarcosine N,N-dmg 100 °C 100 °C 100 °C Rate constant Dz D2 Dz " '0 kiA 0.99 1.48 0.74 , ' DA 4.39 0.76 1.14 sA 0.72 0.46 0.65 kiC V 0.58 0.66 90 0.45 0.87 SC 0.59 0.47 aycine Sarcosine N,N-dmg 70 °C 70 °C 70 °C Rate constant Dz Dz Dz pA 0.03 0.24 1.00 SA 1.54 0.72 0.40 pC 1.50 1.00 5C 0.56 0.38 aycine Sarcosine 50 °C 50 °C Rate constant Dz Dz " kiA 0.09 0.25 pA 0.10 0.35 5A 0.42 1.00 kiC 0.05 PC 1.00 sc 0.70 All rate constants are multiplied by 104 236 3500 --.- Time: 0 Hours +Time: 5 Hours 30007777 -~ —- 2500 7777—777 7W- WWW,” 2000 777- 500 7 O 3.20 3.15 3.10 Figure 112. 1H NMR of H/D exchange at the methyl position of 0.25 M betaine (D20, 3.05 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) 4.7.2 Other Substrates Several additional substrates were examined. There were specific reasons for evaluating each of the substrates. Ethanolamine was chosen in order to determine whether the hydrogens on the methylene next to the oxygen or the methylene next to the nitrogen would exchange faster and it was hypothesized that the hydrogens on the methylene next to the oxygen would exchange faster. Using 0.25 M ethanolamine at 100 °C with 1000 psi H2 and D20 and l g dry 5% Ru/C for four hours, the hydrogens on the methylene adjacent to the oxygen exchanged more quickly than the hydrogens adjacent to the nitrogen (Scheme 32). 237 3000 77 2500 ~ 2000 7 1500 7 1000 77 500 -500 2.90 l 2.6 ppm HO\/\ NH 2 3.4 ppm T Scheme 32. 1H NMR positions for the methylene hydrogens on ethanolamine It was not possible to deconvolute the CH2 from the CHD due to the fact that the splitting was overlapping. Thus, the concentrations are expressed as CH2 plus CHD. The 1H NMR of both the methylene next to the oxygen (Figure 113) and the methylene next to the nitrogen (Figure 114) are shown below. After two hours of reaction time, the deuterium exchange at both positions reached equilibrium and was essentially identical (Figure 115). The concentrations expressed in Figure 115 are the sum of CH2 and CHD. 238 2250 2000 2000 7- 7 1750 1750 7 1 500 - 1250 1250 71— 77 - 1000 7- 750 7 500 - 250 Figure 1 13. 1H NMR of H/D exchange at the methylene position next to the 70H of 0.25 M ethanolamine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 2250 2000 7 +Time: 0 Hours +Time: 4 Hours 20007 7 7 7 77 7 7 7 7 77 7 7 771750 17507 . H 7 777777 7 7 771500 1500 7 7 W .- 7 77777 7 7 — 1250 1250 77— 7 1000 1000 7 i 750 750 - 7 7 500 500 7 7 7 L250 250 o 0 7250 2.61 2.59 2.47 Figure 114. 1H NMR of H/D exchange at the methylene position next to the —NH2 of 0.25 M ethanolamine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 239 0.25 + CH2-OH 7'7 CH2-NH2 : l I 0.20 77 -— -- l S. g 0.15 — S a . S ; § 0.10 77 7 c i o l o a 0.05 77 -7— i 0.00 i i + 4 + l pre-rxn 0 60 120 180 240 Tlme(m|n) Figure 115. Concentration data for 0.25 M ethanolamine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) An experiment with 0.25 M glycolic acid at 100 °C with 1000 psi H2 and D20 was run for 6 hours. This experiment was conducted in order to evaluate the extent of H/D exchange at the methylene carbon. Results indicate that very little H/D exchange . occurred at the methylene carbon and no measurable amount of CHD was observed. The concentration after six hours of reaction time was 0.23 M. However, the peak did become broader over the course of the reaction. Figure 116 compares one hour and after six hours of reaction. Glycolic acid will undergo hydrogenation under similar conditions, however, at 100 °C the hydrogenation rate is slowed significantly and no hydrogenation products were detected. 240 150 _...- _-..-.. -_.- was".-. .---W-mv.-.-._--w M, _-...-.“ _-.-.-._-.-.--,. 125 + Time: 1 Hour + Time: 6 Hours 1257777 777 7 —— 7— 777—7— - A7 7777 i 777 7100 100 4777 7 75 ~ 50 75* 50 r— ,p 25 25 4.24 4.20 4.16 4.12 4.08 4.04 4.00 3.96 Figure 116. 1H NMR of H/D exchange at the methylene position of 0.25 M glycolic acid (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) An experiment with 0.25 M tert-butanol at 100 °C for with 1000 psi H2 and D20 for four hours was run to evaluate the extent of H/D exchange at the methyl carbons, although little was expected. Results indicate that little H/D exchange occurred at the methyl carbons and a nominal amount of CH2D and CHD2 was observed. The conversion after four hours of reaction time was 26%. However, the peak did become broader over the course of the reaction. Figure 117 compares one hour and after four hours of reaction. In the solvent study work tert-butanol was examined as a solvent and it was hypothesized that H/D exchange was not occurring on the methyl carbons due to an inability for OH activation and this experiment confirmed this. However, there is still some solvent degradation and presumably the interaction with the solvent and the catalyst is preventing the substrate (in the hydrogenation experiments) reduction to the alcohol product. 241 5500 ~- 77777777777777 5000 -O-Time: 0 Hours -..-.. Time: 4 Hours 5000 «777—- —---- - 77 — --7 77 —7—7—7- 7 77 — 7 7 4500 4500 «7 — —7 7-77 7 77 — i - 77 7 — 7 7 4000 4000 +7 7— 7777 7777 7 77 77 7 7 77 7 3500 3500 17 -77 7 777 77 7— 7 ~— 7—7 7 7 7 7 7 777 i 3000 3000 + 7 - -- 7 7 77 7—- 7 7 7 —- 7~ 7 7— 7 77 7 2500 2500 7777 77 7 7 7 77 7 77—7 777 7 77 - 7- 7 2000 0 2000 «~— 7 7 7 - — 7 777—- — 77 7 ——— 7777 7— 71500 n 1500 ~7 ~77 - 7 77 7 77 77—7 ’. —7 7 77 7 —77—7- 7 71000 1000 77—7— - 777—- 77— 7777- 7777— 7 $2 77 77777 77 7—7 77—7 7 500 i 0 7- -500 1.06 1.05 1.04 1.03 1.02 Figure 117. 1H NMR of H/D exchange at the methyl positions of 0.25 M (err-butanol (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 4.8 Cometition Experiments 4.8.1 Hydrogenation Studies Betaine hydrochloride did not undergo either hydrogenation or H/D exchange and it was not clear if the substrate was interacting with the surface and whether the chloride ion was poisoning or deactivating the surface. To evaluate the hypothesis that betaine hydrochloride was not interacting with the surface (possibly due to sterics) a competition experiment with glycine was conducted. Presumably, if betaine hydrochloride is not competing for the surface then the hydrogenation of glycine should be unaffected. However, if the chloride ion is reducing the efficiency of the catalyst, then the glycine hydrogenation rate would be affected. 242 Two competition experiments with betaine and glycine were performed. The first was hydrogenation of 0.25 M betaine hydrochloride, (CH3)3N+CH2COOH°C1' and 0.25 M glycine in the presence of 0.5 M H3PO4 at 130 °C with 1 g dry 5% Ru/C catalyst. The ratio of both the substrates and the additional acid to the catalyst was roughly double with respect to the baseline reaction of glycine alone. The results of the reaction compared with the baseline reaction are presented in Figure 118. 100% ".--“..- - 7 I - - Glycine alone conversion 90% ._. --0--Ethanolamine alone yield .. ______ i —I— Glycine conversion _ . I‘ ' ' 80% ,_ +Ethanolamine yield -' . 4 ........ 9 —I- Betaine conversion g, .- " _._ . . _.7 f 70% d» g”, Choline yield, __7 _ _7 ,_ _ —,—.-.7"..,.7' _ . ._I ' 60%77 ~77 - 77 — - - 7 — —. 500/017 77 - 7 7 77 7 -— III-77,7- 7 7 77 77 40% 77 ~ " 30% 77 77 20% «7 -— 10% ‘77 0% 3‘- 0 3O 60 120 180 240 300 360 Time (min) Figure 118. Hydrogenation of 0.25 M betaine hydrochloride and 0.25 M glycine (0.5 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) The conversion of glycine is significantly suppressed by the presence of betaine hydrochloride, however, the amount of substrate and additional acid present is double with respect to the amount of catalyst typically used. So, the increased substrate-to- catalyst ratio may have contributed partly to the reduced rate. In addition, chloride was 243 present as a complex with betaine. To examine the effect of the counter-ion chloride, an experiment without it was conducted. The second experiment with glycine and betaine (chloride-free) was performed using 0.125 M of each and 0.3 M H3PO4. The results of 0.25 M glycine alone versus 0.125 M glycine and 0.125 M betaine are presented in Figure 119 (130 °C, 1000 psi H2, 1 g dry 5% Ru/C). After six hours of reaction time, the conversion of glycine alone is 90% compared with 71% when in the presence of an equimolar amount of betaine. The yield of ethanolamine is 83% compared with 59% when in the presence of an equimolar amount of betaine. Thus, it appears that the betaine (chloride-free) is competing for surface reaction sites. This is consistent with the early findings that betaine will undergo hydrogenation at 100, 130, and 150 °C. So, it was expected that betaine would be reacting on the catalyst surface, thus a potential competitor to glycine. 244 100% 7- ‘ - ~ Glycine alone conversion ‘ 0-7 Ethanolamine alone yield .. ....... 17—1 “mi—- Glycine conversion . \O— Ethanolamine yield 90% 80% 70% 60% 50°/° 40 t>4, 30 “lb ROD/O fl 1 00A) w 0% , r 30 60 120 180 240 300 360 Time (mln) Figure 119. Hydrogenation of 0.25 M glycine alone and 0.125 M glycine and 0.125 M betaine (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) The results of betaine alone and betaine in the presence of an equimolar amount of glycine are presented below (Figure 120). Betaine is much less reactive than glycine so presence of glycine appears to have less of an effect on the betaine. The experiment with betaine hydrochloride resulted in a negligible amount of betaine hydrochloride conversion and the glycine conversion was reduced from 90% to 21%. Presumably this is mostly due to the presence of chloride and also partly to the substrate- and additional acid-to-catalyst ratio. 245 100% 7 ---"- ‘ ‘ h - - Betaine alone conversion 90% ‘ ‘ --- Choline alone yield “~3— Betaine conversion #— Choline yield 80% 70% 7 -7 —- 0 3O 60 120 180 240 300 360 11mo (mln) Figure 120. Hydrogenation of 0.25 M betaine alone and 0.125 M glycine and 0.125 M betaine (0.3 M H3PO4, H20, 130 °C, 1000 psi H2, 1 g dry 5% Ru/C) 4.8.2 H/D Exchange Studies A competition experiment was run with 0.25 M glycine and 0.25 M sarcosine at 100 °C with 1000 psi H2 and D20. These results were compared to the results from the H/D exchange studies using the individual substrates. The experimental and fitted concentrations of CH2, CHD, and CD2 over time as determined by the equations in Chapter 3 are presented in Figures 121 (with respect to glycine alone) and 122 (with respect to sarcosine alone). When looking at the rate of H/D exchange for glycine at the methylene position, the rate is slightly suppressed in the competition experiment compared with the experiment of glycine alone. 246 Figure 121. Fit of the kinetic rate data for 0.25 M glycine alone and 0.25 M glycine and ! 7— Gly comp AH2 —— Gly comp AHD — Gly comp AD2 ‘ I Gly comp AH2_exp A Gly comp AHD_exp 0 Gly comp ADZ_exp -—Gly alone AH2 —Gly alone AHD —G|y alone AD2 0-20 ‘ D Gly alone AH2_exp A Gly alone AHD_exp 0 Gly alone ADZ_exp ' g o E 0.15 7 7 7 77 7 7 v o g Q fill E g 0.10 l 7 ‘ 7 8 o 0.05 i 7 7 7 7 7 i 0.00 . , . . . f 7 0 50 100 150 200 250 300 350 400 Time (mln) 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) Concentratlon (mol/L) Figure 122. Fit of the kinetic rate data for 0.25 M sarcosine alone and 0.25 M glycine 0.30 77— Sarc comp AH2 ~77 Sara comp AHD — Sarc comp A02 I Sarc comp AH2_exp A Sarc comp AHD_exp o Sarc comp ADZ_exp 0.25 i— —Sarc alone AH2 --Sarc alone AHD —Sarc alone A02 [3 Sarc alone AH2_exp A Sarc alone AHD_exp o Sarc alone AD2_exp 0.20 4 77 7 7 7 7 7 7 7 7 7 7 7 7 7 7 0.157 7 777 777777 77 A o 0.10 77 7 7 7 7 e 7 7 7 7 o 0.05-7 7 7 7 77 77777777 A 0.00 . l . . , , . 0 50 100 150 200 250 300 350 400 Tlme (mln) and 0.25 M sarcosine (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) 247 The conversion after 4 hours of reaction time is 97% in the experiment of glycine alone and 95% in the competition experiment. The yields of CHD and CD2 after 4 hours of reaction time are 30% and 68% in the experiment of glycine alone and 35% and 59% in the competition experiment. However, the rate of H/D exchange for sarcosine at the methylene position appears to be much slower than in the experiment of sarcosine alone. The conversion after 4 hours of reaction time is 98% in the experiment of sarcosine alone and 94% in the competition experiment. The yields of CHD and CD2 after 4 hours of reaction time are 21% and 75% in the experiment of sarcosine alone and 37% and 48% in the competition experiment. The rate constants for glycine alone, sarcosine alone, and in the competition experiment are presented in Figure 123. The kinetics were determined using the models described in Chapter 3. The rate constants for H/D exchange on the methylene carbon of glycine in the competition experiment (and glycine alone) are: k1 A = 1.6x10'4 L/mol/min; (2.3x10'4), k1,, = 4.8x10"5 L/mol/min; (1 .2x10'4), k2A = 8.3x10'5 L/mol/min; (1.2x10'4), k-2A = 1.6x10'4 L/mol/min; (3.2x10'4). The rate constants are about 30% larger in the case of glycine alone for the exchange of both the first and second hydrogen. 248 2.5E-04 I Glycine alone BGlycine and sarcosine competition _glycine fl Sarcosine alone 2 0504 _ , Elisaircosine and glycine competflion_§arcosine , E E . 'o' E 1.5E-04 7 3g ; :l ,. I.» v a 8 {gig '_’ = 3 i3 {0 i § 1.05-04 «7 7 o » . xiii". :00; “I: Q. E 3: ' ,1 no ~. "’ 7] o: .. .. 5.05-05 7 g " 9% 3° : z: 2: $07,. '9 1.7 . it 0.0E+00 7 :3: k1_A k2_A k1_C k2_C k3_C Figure 123. Rate constants for PUB exchange of 0.25 M glycine alone, 0.25 M sarcosine alone, 0.25 M glycine in the competition experiment, and 0.25 M sarcosine in the competition experiment (D20, 100 °C, 1000 psi H2, 1 g dry 5% Ru/C) The rates for H/D exchange on the methylene carbon of sarcosine in the competition experiment (and sarcosine alone) are: klA = 1.5x10“1 L/mol/min; (1.6x104), k.1A = 7.8x10'5 L/mol/min; (1 .4x10'5), k2,, = 8.4x10'5 L/mol/min; (1.5x10'4), k2A = 2.3x10'4 L/mol/min; (2.5x104). The rate of exchange for the first hydrogen for sarcosine alone or in the presence of glycine is the same. Although the rate of the second hydrogen being replaced with deuterium is roughly almost twice as fast in the case of sarcosine alone as compared to sarcosine plus glycine. Again, using kinetic rate model A the rates for H/D exchange on the N-methyl carbon of sarcosine in the competition experiment (and sarcosine alone) are: klc = 6.5x10'5 249 L/mol/min; (1.1x10'4), 1m = 4.9x10'6 L/mol/min; (0), kzc = 8.0x10'5 L/mol/min; (1 .4x10' 4), 1m = 3.2x10“ L/mol/min; (9.7x10'7), k3c = 6.6x10'5 L/mol/min; (8.3x10'5), k-3c = 1.1x10'3 L/mol/min; (2.0x10'4). These results again illustrate a slower rate of deuterium exchange on the methyl hydrogens of sarcosine in the presence of an equimolar amount of glycine. The rate of H/D exchange for both the first and second hydrogen are nearly 2x faster in the case of sarcosine alone. The rate of exchange for the third hydrogen is comparable for both the competition experiment and sarcosine alone but still quicker for sarcosine alone. 250 Chapter 5. SUMMARY From the studies discussed it was confirmed that for hydrogenation of organic acids, the presence of hydrogen and catalyst (5% Ru/C) is required. Addition of equimolar H3PO4 resulted in a negligible effect on the hydrogenation reaction of glycolic acid, suggesting that phosphates, at least in protonated form, do not bind strongly to the catalyst. It was assumed that the additional acid would not affect the hydrogenation of the other substrates evaluated. However, from the competition experiments discussed in this paper as well as from the work by Chen, Jere, Peereboom, and Pimparkar et al. we know that the organic acid substrates have different binding affinities for the catalytic Ru particles and for the carbon support. Earlier amino acid reduction models included terms for H3PO4 competing with the same surface sites as the organic substrate. However, from the carbon adsorption studies it was found that H3PO4 has less of an adsorption affinity for the catalyst than the set of organic acids and alcohols examined, indicating that it is not in direct competition with the substrate for the actual hydrogenation sites on the catalyst. For the H/D exchange control studies it was observed that non-catalytic H/D exchange does occur in the methylene positions of glycine, sarcosine and betaine. For sarcosine, about 30% of the methylene was monodeuterated after six hours at 130 °C. This result can be compared to a similar experiment run with catalyst (5% Ru/C) at the slightly lower temperature of 100 °C (both in 1000 psi H2), where the methylene is ca. 30% monodeuterated after just one hour. Comparing rate constants, k1 A and k2A for the two experiments, the control experiment had a klA = 9.8x10'6 and k2A = 0 L/mol/min (i.e. CH; 251 and CHD peaks accounted for all starting substrate), whereas the catalyzed but 30 °C cooler reaction was ca. 15 times faster, k1 A = 1.6x10'4 and k2A = 1.5x10‘4 L/mol/min. For betaine the measured concentrations of CH2 and CHD are nearly identical in the experiment with or without catalyst (over 6 hours), showing little or no evidence of catalyzed exchange. H/D exchange in the absence of catalyst was not observed with glycine or choline bicarbonate (130 °C; 1000 psi H2; D20). Results from adsorption studies using 5% Ru/C illustrated that adsorption onto the catalyst is occurring and when the amount of catalyst is doubled, the amount of substrate lost to adsorption is also doubled. However, doubling the catalyst amount doubles the conversion rate for organic acid hydrogenations. Taken together, these facts indicate that the reaction is limited by the intrinsic kinetics and not by interphase mass transport. Peereboom et al. also confirmed and explored the adsorption of substrates on the catalyst surface and in the micropores. Results indicated that the less hydrophilic compounds have more of an affinity for the pores rather than the bulk solution and also the concentration of a substrate in the bulk solution could be very different from its concentration in the pores. Comparing six substrates and their relative adsorption onto the carbon support, it was found that lactic acid (7% adsorbed) was absorbed more than any other substrate. The other substrates examined were butanol (5.7% adsorbed), glycolic acid (4.8% adsorbed), propylene glycol (3.4% adsorbed), phosphoric acid (1.0% adsorbed), and ethylene glycol (0.8% adsorbed). Experiments evaluating the effect of additional acid, H3PO4 have 252 shown no effect on the hydrogenation rate and this is partly due to the fact that it may have less of an adsorption affinity for the catalyst and is more hydrophilic than for instance, butanol. It may thus prefer the bulk solution over the micropores. The amount adsorbed on the carbon support was similar to the amount of substrate typically "lost" after being exposed to the catalyst (5% Ru/C) at time zero of the experiment. The hydrogenation of betaine hydrochloride (CH3)3N+CH2COOH°C1' was compared to the hydrogenation of betaine (CH3)3N+CH2COO' (both experiments conducted with equimolar H3PO4). Results indicated that none of the desired alcohol product was formed using betaine hydrochloride. An experiment with 0.25 M glycine (130 °C, 0.3 M H3PO4, 1000 psi H2) in both the presence and absence of 0.25 M HCl showed a conversion drop (83% to 34%) of more than half for the experiment with HCI. The effect of an equivalent amount of sodium chloride was also evaluated. The hydrogenation rate dropped from 83% to 72% indicating that again the chloride ion may be partially deactivating the catalyst, affecting the substrate, or competing with the surface. However, in this experiment half the amount of glycine, NaCl, and additional acid, H3PO4 was added compared with the experiment with HCl (for the same amount of catalyst). So, the reduction in the amount of total acids may have reduced the effect seen by the NaCl. Additional insights arise from deuterium exchange studies comparing betaine hydrochloride (CH3)3N+CH2COOH°CI' and betaine (CH3)3N+CH2COO' substrates. In the former case, no deuterium was incorporated at either the methylene or methyl carbons. From the control experiments in which no catalyst was added, it was found that the same amount of deuterium incorporation was occurring in betaine in both in the presence and 253 absence of catalyst. This suggests that the lack of H/D exchange seen in the case of betaine hydrochloride is not due to a partially deactivated catalyst or competition by HCl for catalyst surface sites. The chloride ion in the solution appears to somehow inhibit methylene C-H activation in the bulk solution. It was already noted that the presence of an equimolar amount of additional acid, H3PO4, does not reduce the hydrogenation rate of glycolic acid. The need for the additional acid is to ensure that the carboxylic acid functionality in glycine and the N-methylated glycines is protonated. These amino acids are zwitterionic in nature and thus the carboxylic acid functionality is dissociated unless the additional acid is present. For glycine, a yield of 15% is obtained in the absence of catalyst, compared with 83% conversion with additional acid. To determine if the additional acid slows the H/D exchange reaction, betaine was initially chosen due to its slower competing hydrogenation reaction. Results showed that both with and without the additional acid present the amount of H/D exchange on the methylene carbons were essentially the same. However, betaine undergoes non-catalytic H/D exchange. The k; A values were 0.41x104 and 0.43x10'4 for the case of betaine (130 °C) and betaine with 0.3 M H3PO4 (100 °C), respectively. Presumably the rates should be different for the different temperatures, but at 100 °C with acid, some of the betaine is hydrogenated to form choline. After six hours a 38% yield of choline was detected by HPLC; thus, included in the H/D rates (methylene carbon) are both the exchange and also the conversion to choline. The yield of choline in this experiment is quite a bit higher than in the same conditions but at 130 °C. The 254 reason for this is that at higher temperatures more betaine degradation (presumably to trimethylamine and gases) is occurring. Glycine and sarcosine were also run at 50 and 70 °C in the absence and presence of 0.3 M H3PO4 to evaluate the extent of HfD exchange. The rate of exchange was significantly decreased in the presence of H3PO4; presumably protonation of the amine makes the neighboring C-H bond less electron-rich and therefore less activated for C-H attack by Ru. Various solvents were examined to explore their effects and the role of the water solvent in the hydrogenation reactions. The hydrogenation rate of glycolic acid was significantly depressed when using either dry THF or tert-butanol. Studies with THF indicate that it readily undergoes H/D exchange at both C1 (83% H replacement) and C2 (42% H replacement) positions over six hours. The fact that the H/D exchange reaction of THF slows glycolic acid hydrogenation supports the claim that the hydrogenation sites and H/D exchange sites are not distinct. Next, tert-butanol was tried based on the hypothesis that it would not undergo exchange or interact with the surface. Indeed, no exchange was observed in this solvent with or without added substrate, but neither was glycolic acid hydrogenated. Some loss of tert—butanol was seen, a finding likely due to dehydration of the solvent to isobutene and subsequent hydrogenation to isobutane. The solvent ethanol was used for the experiments with glycolic acid both with and without additional acid, H3PO4. As expected, the acid promoted a significant amount of 255 esterification, but even in the case with no additional acid, the ester ethyl glycolate was still formed. In both experiments, only 8% of ethylene glycol was formed compared with 47% in the solvent water under otherwise identical conditions. The literature suggests that esters undergo hydrogenation with certain catalysts (Raney Ni, copper/chromium oxide, rhenium black) but that there is competition for the surface between the acids and the esters, reducing the amount of ester hydrogenation. But, in this case, slow or no ester hydrogenation occurs over 5% Ru/C . To examine this further, ethyl lactate was subjected to the usual aqueous hydrogenation conditions. It quickly hydrolyzed to lactic acid (100% at one hour). The product propylene glycol, was not observed until after lactic acid formed, again supporting the claim that the ester reacts slowly or not at all. This experiment’s propylene glycol yield was nearly the same as with lactic acid itself, suggesting that in this case the lactic acid hydrogenation suffers little hindrance from the ester or dilute ethanol. The results from the organic acid hydrogenation studies indicate that simple carboxylic acids such as propanoic acid and acetic acid are substantially more difficult to reduce than lactic acid and alanine. The various substrates were chosen in order to evaluate the effect of hydrogen bonding, electron withdrawing ability, and heteroatom coordination to the surface. Hydrogen bonding was determined to be a key player in accelerating the hydrogenation reaction. Reactions with the (Jr-substituent -N+H3, -OH, -OCH3, - N+H2(CH3), and -N+H(CH3)2 were far more reactive than either the unsubstituted acetic acid or an a-substituent of -CH3 or -N+(CH3)3. Sterics were considered in the N- methylated glycine series (-N+H3, -N+H2(CH3), -N+H(CH3)2, and -N+(CH3)3) in that the 256 hydrogenation rate falls off by about 1/3 as each N-methyl group is added. However, there is a trend for hydrogenation in that the reactivity falls off by nearly half as a methyl is added. Also, when comparing the groups, -OH to a more bulky -OCH3 or -CH3, the substrates with the or-OH group (glycolic acid and lactic'acid) are far more reactive. This may be partly due to a less bulky group but also due to electronics (-OH is more electron- withdrawing than —OCH3) as compared to -OCH3 or 7CH3. Considering the role of electronics, it was found that the substrate with oc-OCH3 is much more reactive than on- CH3 (yield of desired product is 35% compared with 13%). In an effort to build upon the above research and probe the reactivity differences in hydrogenation of the methylated glycine series, H/D exchange studies were performed. Reactions run in D20 with either H2 or D2 and catalyst were effective at exchanging deuterium into the CH2 and CH3 sites of amino acids. The Ru catalyst also exchanges H (from H2) for D (from D20) to form HOD and HD, independent of substrate. Analysis by internal standard-calibrated 1H NMR allowed complete isotopomer-specific inventory of the H concentration in substrates and water. Simple sequential replacements of H by D were observed in the amino acids. Two kinetic models were developed to determine the rate constants for PUB exchange at both the methylene and methyl positions. Both included possible time delays due gas/liquid isotopic equilibration (kH) and/or a catalyst induction (ktum_on) period; ultimately the latter was discarded as redundant. The first model treated each isotope replacement step as a process unrelated to the others. Thus, between forward and reverse replacements in CH2 and CH3 sites and k“, eleven rate constants were freely varied in the fitting. The second model used a single rate constant 257 for each site, further modified by primary and secondary isotope effect values, resulting in six values + kH for this fitting. The latter, more economical and interpretive model gave fits nearly as good as those from the eleven-variable one. The H/D exchange took place at lower temperatures (SO-100 °C) than hydrogenation (IOO-150 °C) and therefore it was possible to study H/D exchange while minimizing the amount of competing hydrogenation reaction. Mapping out the optimal temperatures for both reactions allowed H/D exchange to be studied. Glycine, sarcosine, and N,N- dimethylglycine were run at 100 °C with 1000 psi H2 or D2 and the datasets from the pairs of experiments were linked in order to define a common set of rate constants. Experiments were also run at lower temperatures (70 and 50 °C). The rate constants for H/D exchange at the methylene and methyl carbons are smnmarized in Table 10. Table 10. Summary of rate constants for glycine, sarcosine, and N,N-dimethylglycine Rate Glycine Sarcosine N,N-dmg Glycine Sarcosine N,N-dmg Glycine Sarcosine constant 100 °C 100 °C 100 °C 70 °C 70 °C 70 °C 50 °C 50 °C (L/mol/min) Dz Dz Dz Dz Dz Dz Dz Dz ki_CHz 0.99 1.48 0.74 0.74 0.99 0.06 0.09 0.25 ki_CHa 0.58 0.66 0.24 0.09 0.05 All rate constants are multiplied by 104 Somewhat surprisingly, H/D exchange rates decreased in the following order: sarcosine (CH3NH2+CH2CO2') > glycine (CH2 site) > N,N-dimethylglycine. This trend in reactivity does not match that observed in the hydrogenation of these same substrates (glycine > sarcosine > N,N-dimethyglycine. This may be due to the fact that for H/D exchange an 258 electron-rich C-H bond promotes the reaction, whereas in the hydrogenation reaction, it is presumed that steric effects play more of a role. The I-I/D exchanges on the N-methyl(s) of both sarcosine and N,N-dimethylglycine were slower than on the methylene carbon (for all conditions) and in general did not fully reach equilibrium after six hours (Table 10) at 100 °C. In addition it was found that the H/D exchange occurs step-wise at both the methylene and methyl carbons (CH2 to CHD to CD2) and (CH3 to CH2D to CHD2 to CD3). Betaine was found to undergo non- catalytic H/D exchange at the methylene carbon. Glycine and sarcosine exhibited a small amount of H/D exchange in the absence of catalyst but only at higher temperatures (130 °C). With its fixed N+ next door, the acidity of betaine’s a-CH2 is enhanced, increasing the rate of deprotonation, enol formation, and the resulting H/D exchange. In summary, a detailed set of hydrogenation and H/D exchange experiments using a 5% ruthenium on carbon catalyst were conducted for several two- and three-carbon organic acid substrates and a kinetic model was developed for the H/D exchange. Simple alkanoic acids (acetic, propanoic, isobutyric) and betaine, quatemized glycine, are not very reactive. For the other methylated glycines, the rate of hydrogenation falls off by a factor of ca. three with each additional methyl, and reduction requires acid, as found earlier by Jere. Use of water as a reaction medium appears optimal. The H/D exchange studies indicated a different trend in reactivity for the methylated glycines (sarcosine > glycine > N,N-dimethylglycine) as compared to hydrogenation. The rate of H/D exchange appears to be promoted by an electron-rich C-H bond. The addition of H3PO4 259 significantly slows the H/D exchange while enabling the reduction, suggesting that the free amine is needed for the hydrogen replacement reaction. Future projects could include additional probes of the roles of steric effects, pH more precisely defined, additional catalyst site inhibitors, and stereoelectronic aspects of control. For instance, to explore whether the electrons on the amine are coordinating with the catalyst surface, an experiment with a t-butyl-locked piperidine (chair conformer) with various N-alkyl groups could be run to determine the preference for Hax versus Heq. The results would help to map out the elements controlling C-H activation alpha to amine sites. 260 Chapter 6. REFERENCES l Braun, B.; Charney, R.; Clarens, A; F arrugia, J .; Kitchens, C.; Lisowski, C.; Naistat, D.; O’Neil, A. J. Chem. Ed. J. Chem. Ed. 2006, 83, 1126 2 British Petroleum, BP Statistical Review of World Energy, 2007 3 Yang, M; Somorjai, G.A. J. Phys. Chem. B. 2004, 108, 4405 4 Jere, F. Ph.D. Dissertation, Michigan State University, 2003 5 Antons, S. In US Patent 5, 731,479; Bayer Aktiengesellschaft: USA, 1998 6 Antons, S. In US Patent; Bayer Aktiengesellschaft: USA, 2002 7 Carnahan, J .E.; Ford, T.A.; Gresham, W.F.; Grigsby, W.E.; Hager, G.F. J. Am. Chem. Soc. 1955, 77, 3766 8 Behr, A; Brehme, v. Adv. Synth. Catal. 2002, 344, 525 9 Mao, 13w; Cai, zz; Huang, M.Y.; Jiang, Y.Y. Polym. Adv. Technol. 2003, 14, 278 '0 He, D.I—I.; Wakasa, N.; Fuchikami, T. Tet. Lett. 1995, 36, 1059 “ Bowden, H; Adkins, H. J. Am. Chem. Soc. 1934,56, 689 ‘2 Adkins, H; Folkers, K. J. Am. Chem. Soc. 1932, 54, 1145—1154 '3 Adkins, H; Cramer, H.I. J. Am. Chem. Soc. 1930, 52, 4349 ‘4 Adkins, H.; Pavlic, A.A. J. Am. Chem. Soc. 1947, 69, 3039 '5 Adkins, H; Billica, H.R. J Am. Chem. Soc. 1948, 70, 3121 '6 Vaidya, P.D.; Mahajani, v.v. Chem. Eng. Sci. 2005,60, 1881 '7 Sajiki, H.; Kurita, T.; Esaki, H; Aoki, F.; Maegawa, T; Hirota, K. Org. Lett. 2004, 6, 3521 '8 (a) Faller, J.W.; Smart, C.J. Organometallics 1989, 8, 602. (b) Heys, R. J. Chem. Soc., Chem. Commun. 1992, 680. (c) Heys, J .R.; Shu, A.Y.L.; Senderoff, S.G.; Phillips, N.M. J. Labelled Compd Radiopharm. 1993, 33, 431. (d) Lukey, C.A.; Long, M.A.; Gamett, J .L. Aust. JChem. 1995, 48, 79. (e) Hesk, D.; Das, P.R.; Evans, B. J. Labelled Compd Radiopharm. 1995, 36, 497. (1) Chen, W.; Games, K.T.; Levinson, H.H.; Saunders, D.; Senderoff, S.G.; Shu, A.Y.L.; Villani, A.J.; Heys, JR. J. Labelled Compd 261 Radiopharm. 1997, 39, 291. (g) Klei, S.R.; Golden, J .T.; Tilley, T.D.; Bergman, R.G. J.Am. Chem. Soc. 2002, 124, 2092. (h) Skaddan, M.B.; Yung, C.M.; Bergman, R.G. Org. Lett. 2004, 6, ll ‘9 (a) Hesk, D.; Jones, J.R.; Lockley, W.J.S. J. Labelled Compd. Radiopharm. 1990, 28, 1427. (b) Hesk, D.; Jones, J.R.; Lockley, W.J.S. J. Pharm. Sci. 1991, 80, 887. (c) Lenges, C.P.; White, P.S.; Brookhart, M. J. Am. Chem. Soc. 1999, 121, 4385 20 (a) Weil, T.A.; Friedman, S.; Wender, I. J. Org. Chem. 1974, 39, 48. (b) Lenges, C.P.; Brookhart, M.; Grant, B.E. J. Organomet. Chem. 1997, 528, 199 2‘ (a) Garnett, J.L.; Hodges, R.J. J. Am. Chem. 1967, 89, 4546. (b) Gol’dshleger, N.F.; Tyabin, M.B.; Shilov, A.E.; Shteinman, A. A. Zh. Fiz. Khim. 1969, 43, 1222. (c) Buncel, E.; Clement, O. J. Chem. Soc., Perkin Trans. 2 1995, 1333 22 (a) Ogasawara, M.; Saburi, M. Organometallics 1994, 13, 1911. (b) Collman, J .P.; Fish, H.T.; Wagenknecht, P.S.; Tyvoll, D.A.; Chng, L.-L.; Eberspacher, T.A.; Brauman, J .I.; Bacon, J .W.; Pignolet, L.H. Inorg. Chem. 1996, 35, 6746 23 Perthuisot, C; Fan, M; Jones, W.D. Organometallics 1992, 11, 3622 24 Sajiki, H.; Aoki, F.; Esaki, H.; Maegawa, T.; Hirota, K. Org. Lett., 2004, 6, 1485 25 Jere, F .T.; Miller, D.J.; Jackson, J .E. Org. Lett. 2003, 5, 527 26 Haikides, C.J.; Frey, P.A.; Tobin, 1.3. J. Am. Chem. Soc. 1993, 115, 3332 27 Nibeilliu, M.B.; Malthouse, P.G. Biochem. J. 2004, 381, 847 28 Rios, A.; Amyes, T.L.; Richard, J .P. J. Am. Chem. Soc. 2000, 122, 9373 29 Rios, A; Richard, J.P. J. Am. Chem. Soc. 1997, 119, 8375 3" Cox, HA; Julian, R.R.; Lee, s.w; Beauchamp, J.L. J. Am. Chem. Soc. 2004, 126, 6485 31 Matsubara, S. Yokota, Y.; Oshima, K. Chem. Lett., 2004, 33, 294 32 Li, 2.; Li, C.J. J. Am. Chem. Soc. 2004, 126, 11810 33 Li, z; Li, C.J. Org. Lett. 2004, 6, 4997 34 Li, z; Li, C.J. J. Am. Chem. Soc. 2005, 127, 3672 262 35 Alston 11, W.C.; Haley, K.; Kanski, R; Murray, C.J.; Pranata, J. J. Am. Chem. Soc. 1996,118,6562 36 Broadbent, H.S., Campbell, G.C., Bartley, W.J., Johnson, J .H. J Org. Chem. 1959, 24, 1847-1854 37 Antons, s; Beitzke, B. Chem. Abstr. 1996, 125, 114175 38 Studer, M.; Burkhardt, s; Blaser, H.U. Adv. Synth. Catal. 2001, 343, 802 39 Ruiz, A. Guerrero; Alarcon, M. Cerro; Valiente, A. Maroto; Ramos, I. Rodriguez Carbon 2005, 43, 2711 40 Chemical Marketing Reporter, 1091-0110, 260 (11) (September 24, 2001) 30 4‘ Chemical Marketing Reporter, 1092-0110, 265 (17) (April 26, 2004) 20 (4) 42 Dasari, M.A.; Kiatsimkul, P.P.; Sutterlin, W.R.; Suppes, G.J. Appl. Catal. A: Gen. 2005,281,225 43 Muller, P.; Rimmelin, P.; Hindermans, J .P.; Kieffer, R.; Kiennemann, A. Transformation of Sugar into Glycols on a 5% Ru/C Catalyst. Stud. Surf Sci. Catal. 1991, 59, 237 44 Lambert, J .B., Mazzola, E.P. Nuclear Magnetic Resonance Spectroscopy: An Introduction to Principles, Applications, and Experimental Methods 45 Lambert, J.B., Greifenstein, L.G. J. Am. Chem. Soc. 1974, 96, 5120 46 Gutowsky, HS. J. Chem. Phys. 1959, 31, 1683f 47 Pople, J.A. M01. Phys. 1963, 7, 201 48 Zhang, Z., Jackson, J .E., Miller, D.J. Appl. Catal. A 2001, 89-98 49 Zhang, Z., Jackson, J .E., Miller, D]. Ind. Eng. Chem. Res. 2002, 41, 691-696 5" Jere, F.T., Jackson, J.E., Miller, D.J Ind. Eng. Chem. Res. 2004, 43, 3297-3303 5‘ Singh, U.K., Vannice, M.A. Appl. Catal. A 2001, 213, 1-24 52 Pallassana, v; Neurock, M. J. Catal. 2002, 209, 289-305 53 Kovacs, D.G., Jackson, J.E., Miller, D.J. Abstr Pap Am Chem. Soc. 2001, 221 U177-8 5“ Lahr, D.G., Shanks, B.H. J. Catal. 2005, 232, 386-394 263 55 Lahr, D.G., Shanks, B.H. Ind. Eng. Chem. Res. 2003,42, 5467-5472 56 Kusunoki, Y., Miyazawa, T., Kunimori, K., Tomishige K Catal. Commun. 2005, 6, 645-649 57 Chaminand, J ., Djakovitch, L., Gallezot, P., Marion, P., Pinel, C., Rosier, C. Green Chem. 2004, 6, 359-361 5“ Perosa, A., Tundo, P. Ind. Eng. Chem. Res. 2005, 44, 8535-8537 59 Peereboom, L., Koenigsknecht, 8., Hunter, M. Jackson, J .E., Miller, D.J. Carbon 2007, 45, 579-586 60 Dalavoy, T.S., Jackson, J .E., Swain, G.M., Miller, D.J., Li, J ., Lipkowski, J. J. Catal. 2007, 246, 15-28 6' Chen, Y., Miller, D.J., Jackson, J .E. Ind. Eng. Chem. Res. 2007, 46, 3334-3340 62 Zhang, Z., Jackson, J .E., Miller, D]. Bioresource Technology 2008, 5873-5880 63 Pimparkar, K.P., Miller, D.J., Jackson, J .E. Ind. Eng. Chem. Res. 2008, 47, 7648-7653 6“ Knifton, J.F., Sanderson, JR, and Stockton, ME. Catal. Lett. 2001, 73, 55 65 Abella, L.C., Gaspillo, P.D., Itoh, H., and Goto, S.J. Chem. Eng. Jpn. 1999,32, 742 66 WE. Smith. Tetrahedron 2002, 58, 2091 264