. . .45. t! .u 1.7.1‘ , . 1:. . unblu t» (I 34.. 3. . . fligfifimrdu§fi a: P1!“- ... cl: 2. :Emfiuhww . . 31‘ . . I: . .X 9.7. II ¢ ‘4‘] i ‘ . .‘lvihhltiwmauni .‘DK I'VE-Ionith d I. 1.1!! but .nunnn . I v! ‘ , D . gr fidfiflni 81.81». mek.‘ - y B .1)?" n! aloud. 160- ‘5‘? 4.7.3:... I! 41!... dichuflvv‘hfl .. .l ..| ‘ ..‘ >Sliulil. Earls ‘(Ir . 1.1.3); .. kill] ‘ .0; .53.- 1... J IlllilWIHIlllHIlflHhIillmHHIMNIHWIHMHUI 3 1293 020611 This is to certify that the dissertation entitled Biomass Conversion: Sugar Hydrogenolysis, Glycerol Dehydroxylation and HF Saccharification presented by Keyi Wang has been accepted towards fulfillment of the requirements for Ph . D . degree in Chemical Engineering Major professor DateW7 MSU it an Affirmativtt Action/Equal Opportunity Institution 0-12771 IJBRAHY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE .gt'l .irblié OCquwifi W05 11/00 c/CIFICIDIuDqus-pu BIOMASS CONVERSION: SUGAR HYDROGENOLYSIS, GLYCEROL DEHYDROXYLATION AND HF SACCHARIFICATION By Keyi Wang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering 1999 BlOI In the l hydrogenolysi: sugar hydroger glycol; the glyc the HF sacchan The tuner reaction mechai mechanism stud; maul-Sh the me“ ABSTRACT BIOMASS CONVERSION: SUGAR HYDROGENOLYSIS, GLYCEROL DEHYDROXYLATION AND HF SACCHARIFICATION By Keyi Wang In the current research, three biomass conversion processes, namely sugar hydrogenolysis, glycerol dehydroxylation and HF saccharification, are investigated. The sugar hydrogenolysis process converts sugars into glycerol, propylene glycol and ethylene glycol; the glycerol dehydroxylation process converts glycerol to 1,3-propanediol; while the HF saccharification process converts lignocellulosic biomass into sugars. The current research on sugar hydrogenolysis has focused on understanding of the reaction mechanism and of the factors controlling the reaction selectivity. In the mechanism study, a series of 1,3-diol model compounds have been hydrogenolyzed, to establish the mechanism of C-C and CO bond cleavage in sugar hydrogenolysis. The experimental work continued our theoretical conception that the CC bond breaks via the retro-aldol reaction of a B-hydroxyl carbonyl precusor, and it also provided for the first time direct evidence that the C-0 cleavage is through dehydration of the same B-hydroxyl carbonyl precursor just mentioned. Establishment of the above mechanism makes it possible for one to adopt a more rational approach than the current practice in control of the selectivity of sugar hydrogenolysis. lire s in sugar hi study. has concentratl selectivity. In the Proposed rt 8T0Up. and Concept is tosylates at: In this Worl mp are l mihemum c2 The in: adsorption p] with a bencl ”actor. Two blSed 0n GXp‘ anilygis Of ti. Mum-Hg in l l“Sight into th C The selectivity study in the current work has focused on control of the C-0 cleavage in sugar hydrogenolysis. Specifically, an experimental study, as well as a simulation study, has been carried out in this work to investigate the effect of temperature, base concentration, hydrogen pressure, catalyst amount and catalyst type on the C—C vs 00 selectivity. These studies, both experimental and theoretical, yielded many useful results. In the research on conversion of glycerol to 1,3-propanediol, a new concept is proposed to selectively transform the second hydroxyl group of glycerol into a tosyloxy group, and then to remove it by hydrogenolysis. The technical feasibility of this new concept is verified experimentally in this work. Before this work, hydrogenolysis of tosylates are generally effected with base metal hydrides, such as LiAlH4 and LiHBEt3. In this work, for the first time, tosylates with a hydroxy group adjacent to the tosyloxy group are hydrogenolyzed with molecular hydrogen in the presence of a nickel or ruthenium catalyst. The mechanism for this hydrogenolysis reaction is also established. The work on vapor-phase HF saccharification has focused on modeling of the HF adsorption process in a packed-bed reactor. However, experiments were also conducted with a bench-scale reactor to study the behavior of HF adsorption in the packed-bed reactor. TWO models of HF adsorption have been developed. One is developed purely based on experimental observations, and the other one is developed based on a detailed analysis of the HF flow, the intrinsic HF adsorption and the heat transfer processes occurring in the reactor during HF adsorption. These models provide a great deal of insight into the HF adsorption process in the packed-bed reactor. First 0 guidance am he gave me d Dr. Kris A. r Dale for sen- Scot J. De Ari. Plant Biotec. Bloprocessmg Engineering a SUppOrL ACKNOWLEDGMENTS First of all, I wish to thank my academic advisor Dr. Martin C. Hawley for his guidance and support in this investigation. I am especially grateful for the encouragement he gave me during the final stage of my dissertation preparation. I would also like to thank Dr. Kris A. Berglund, Dr. Dannis J. Miller Dr. Rawle I. Howlingsworth and Dr. Bruce E. Dale for serving on my graduate committee, and to thank Mr. Todd D. Fumey and Mr. Scot J. DeAthos for their assistance in completing this investigation. Tire Consortium for Plant Biotechnology Research, the Michigan State University Crop and Food Bioprocessing Center, the Amoco Foundation and the Department of Chemical Engineering at Michigan State University are recognized for their generous financial support. iv LIST OF LIST OF LIST OF.‘ CHAP' PART I. c. CHIP] 2.1 f") h) TABLE OF CONTENTS LIST OF TABLES ................................................... viii LIST OF FIGURES ................................................... ix LIST OF SYMBOLS AND ABBREVIATIONS ............................ xi CHAPTER 1. GENERAL INTRODUCTION ............................ 1 PART I. CATALITIC SUGAR HYDROGENOLYSIS CHAPTER 2. INTRODUCTION: SUGAR HYDROGENOLYSIS .......... 10 2.1 Literature Review ........................................... 10 2.1.1 Studies on Process Development .......................... 11 2.1.2 Studies on Reaction Mechanism ........................... 14 2.1.3 Studies on Reaction Selectivity ............................ 17 2.1.4 Studies on Reaction Kinetics ............................. 18 2.2 Objective and Scope of Current Research ........................ 19 CHAPTER 3. EXPERIMENTAL SYSTEM AND METHODS ............. 21 3.1 Experimental System and Procedure ............................ 21 3.2 Reaction Regents and Product Analysis ......................... 25 CHAPTER 4. MECHANISM OF SUGAR HYDROGEN OLYSIS .......... 29 4.1 Theoretical Considerations .................................... 29 4.2 Experimental Verification .................................... 31 V CHAP 5.1 5.3 PART II. S CRAFT 6.1 6.3 t CHAPTER 5. SELECTIVITY OF SUGAR HYDROGENOLYSIS .......... 41 5.1 Selectivity Study Using 2,4-PD Model Compound: Rationale ....... 41 5.2 Experimental Study ........................................ 44 5.3 Simulation Study .......................................... 50 5.3.1 Model Development .................................... 50 5.3.2 Kinetic Parameter Determination ......................... 55 5.3.3 Simulation Results and Discussion ......................... 65 PART H. SELECTIVE DEHYDROXYLATION OF GLYCEROL CHAPTER 6. INTRODUCTION: GLYCEROL DEHYDROXYLATION . . . . 74 6.1 On-going Research on Glycerol Dehydroxylation .................. 74 6.2 A New Conversion Concept and Supportive Work ................. 77 6.3 Objective and Scope of Current research ......................... 83 CHAPTER 7. EXPERIMENTAL PROCEDURES ....................... 85 7.1 Acetalization .............................................. 85 7.2 Tosylation ................................................. 89 7.3 Hydrolysis ................................................ 89 7.4 Hydrogenolysis ............................................ 91 CHAPTER 8. RESULTS AND DISCUSSION .......................... 92 8.1 Acetalization .............................................. 92 8.2 Tosylation ................................................. 94 8.3 Detosyloxylation ........................................... 95 8.4 Design Concept of Industrial Conversion Process .................. 100 PART III. CHAP CHAR 10. 10.1 10.2 CRAP] 11.] 11.2 11.3 APPESDLX APPEXE APPEXE APPEND LIST OF REF PART III. VAPOR-PHASE HF SACCHARIFICATION CHAPTER 9. MODELING THE HF ADSORPTION PROCESS IN A PACKED-BED REACTOR ........................ 112 CHAPTER 10. SUMMARY AND CONCLUSIONS ..................... 116 10.1 Sugar Hydrogenolysis ...................................... 116 10.2 Glycerol Dehydroxylation ................................... 117 10.3 HF Saccharification ........................................ 118 CHAPTER 11. RECOMMENDATIONS ON FUTURE WORK ............. 120 11.1 Future Work on Sugar Hydrogenolysis ......................... 120 11.2 Future Work on Glycerol Dehydroxylation ...................... 121 11.3 Future Work on HF Saccharification ........................... 124 APPENDIX APPENDIX A. COMPUTER PROGRAMS FOR SIMULATION STUDY ..... 125 APPENDIX B. RETRO-ALDOL AND DEHYDRATION RATE CONSTANT DERIVATION PROCEDURE ........................... 131 APPENDIX C. EXPERIMENTAL DATA OF 2,4»PD HYDROGENOLYSIS . . . 134 LIST OF REFERENCES .............................................. 165 vii Table 3-1. Table 32 Table 4—1, Table 4-2~ Table 5-1_ Table 5-2_ Table 5.3_ Table 5-4_ Table 5-5. Table 5‘6. Table 5-7_ Table 5-3. Table 7‘ 1. Table D-1. Table D‘Z, Table 3-1. Table 3-2. Table 4-1. Table 4-2. Table 5-1. Table 5-2. Table 5-3. Table 5-4. Table 5-5. Table 5-6. Table 5-7. Table 5-8. Table 7-1. Table D— 1. Table D-2. LIST OF TABLES HPLC Methods for 1,3-Diol Hydrogenolysis Products ............... 26 GC Methods for 1,3-Diol Hydrogenolysis Products ................. 27 Structures of 1,3-diols Used in Mechanism Study .................. 33 Results of 1,3—Diol Hydrogenolysis ............................. 34 Effect of Temperature on Selectivity ............................. 46 Effect of Base Concentration on Selectivity ....................... 46 Effect of Hydrogen Pressure on Selectivity ....................... 47 Effect of Catalyst Amount on Selectivity ......................... 47 Effect of Catalyst Type on Selectivity ........................... 49 Equilibrium Constants for Relevant Dehydrogenation Reactions ...... 56 Summary of Product Concentration and Reaction Fluxes ............. 60 Summary of Determined Reaction Rate Constants .................. 64 GC Retention Tunes for Acetalization products .................... 88 Rate Constants Directly from Jenson and Hassan (197 6) .......... 158 Determined Rate Constants ................................. 159 viii Figure 3-1. Figure 3-2. Figure H. Figure 5-1, Fi{lure 5-2. F1Sure 5-3. Figure 54. Flgm‘e 5‘5 I"Tgure 5-6. Figure 5-8. Figure 5‘9 figure 5~10. Figure 3-1. Figure 3-2. Figure 4-1. Figure 5-1. Figure 5-2. Figure 5-3. Figure 5-4. Figure 5-5. Figure 5-6. Figure 5-7. Figure 5-8. Figure 5-9. Figure 5-10. LIST OF FIGURES Illustration of Hydrogenolysis Reactor .......................... 22 Schematic Illustration of Experimental System ................... 23 Mechanism of Sugar and Sugar Alcohol Hydrogenolysis ............ 30 Reaction Pathway of 2,4-Pentanediol Hydrogenolysis .............. 42 Alternate Hydrogenation Pathway of (LB-Unsaturated Ketone in Sugar Hydrogenolysis ..................................... 43 Product v.s Reaction Time Profiles. Experimental Conditions: 220 °C, 5.0 MPa H2 pressure, 0.1 g Ni/kieselguhr catalyst, 0.4 ml 1 N NaOH and 40 ml water solvent. Symbols: 0 2,4-PD; I EtOH; O acetone; A i-PrOH; CI 2-pentanone; A 2-pentanol. ......................... 45 Selectivity v.s. Reaction Time Profile Derived from Figure 5-3 ....... 45 Detailed Reaction Scheme of 2,4.PD Hydrogenolysis .............. 51 Simulated Concentration Profiles of Major Products. Simulation Conditions: 220 °C, [H]=100 mM, [M]0=28 mM, [OH]O=10 mM, [A]o=120 mM. ................................ 66 Simulated Selectivity Profile. S is calculated as ratio of the sum of [C] and [D] to the sum of [G], [I] and [J]. .............. 66 Simulated Concentration Profiles of Minor products ............... 67 Simulated Base Concentration profile .......................... 67 Simulated Effect of Base Concentration on Selectivity. Simulation Conditions: 220 °C, [H]=100 mM, [M]0=28 mM, [A]0=120 mM, varying [OH]0. ................................ 68 ix Figure 5-2 Figure 5-1 Figure 5-1 Figure 6-1 Figure 6-2. Figure 7-1. Figure 8-1. Figure 8-2. Figure 8-3, Fl{lure 8-4. Figure 8-5, Ifigure 86 figure 87. Figure 5-11. Figure 5-12. Figure 5-13. Figure 6-1. Figure 6-2. Figure 7-1. Figure 8-1. Figure 8-2. Figure 8-3. Figure 8—4. Figure 8-5. Figure 8-6. Figure 8-7. Simulated Effect of Catalyst Concentration on Selectivity. Simulation Conditions: 220 °C, [H]=100 mM, [OH]O=10 mM, [A]o=120 mM, varying [M]o. ................................. 68 Simulated Effect of Hydrogen Concentration on Selectivity. Simulation Conditions: 220 °C, [M]O=28 mM, [OH]0=10 mM, [A]o=120 mM, varying [H]. ................................... 69 Simulated Effect of k6 on Selectivity. Simulation Conditions: 220 °C, [H]=180 mM, [M]o=28 mM, [OH]o=10 mM, [A]o=120 mM, but varying k6. ................................ 69 A simplified Reaction Pathway of Glycerol Fermentation ........... 76 Illustration of the New Concept of Glycerol Dehydroxylation ....... 78 Experimental Set-up for Acetalization of Glycerol with Benzaldehyde .86 Reaction Pathway of TCH Hydrogenolysis ...................... 97 Reaction Pathway of TPD Hydrogenolysis ...................... 99 Glycerol Dehydroxylation Processes - Acetalization .............. 101 Glycerol Dehydroxylation Processes - Tosylation (Design #1) ...... 102 Glycerol Dehydroxylation Processes - Tosylation (Design #2) ...... 103 Glycerol Dehydroxylation Processes - Detosyloxylation ........... 104 Material Flow in Glycerol Dehydroxylation ..................... 109 Mathemat 1:1. It]. ~ . 3" ’V... RP! LIST OF SYMBOLS AND ABBREVIATIONS Mathematical Symbols k], k_,, ...... , k8, kg i» “ck a a [A10 reaction rate constants defined in Figure 5-5 forward reaction rate constants of Reaction No. i in Figure 5-5 backward reaction rate constants of Reaction No. i in Figure 5-5 reaction equilibrium constant equilibrium constant of Reaction No. i in Figure 5-5 equilibrium constant of dehydrogenation reaction equilibrium constant of dehydrogenation corresponding to Reaction No. i reversibility parameter of Reaction No. 2, defined by Eq.(37) reversibility parameter of Reaction No. 5, defined by Eq.(38) C-C v.s. C—O selectivity reaction time reaction flux rate of Reaction No.1 through No.8, respectively free energy change of dehydrogenation reaction free energy change defined in Page 57. concentration of 2,4-pentandiol initial concentration of 2,4-pentanediol [B] [C] [D] [E] [G] [H] U] [M] [MlO [MH] [on] {Dino Chemical [B] [C] [D] [E] [G] [J] [M] [M]. [MH] [0H] tomo concentration of 4-hydroxy-2-pentanone concentration of acetone concentration of isopropanol concentration of acetaldehyde concentration of ethanol concentration of 3-penten-2-one concentration of dissolved hydrogen concentration of 2-pentanone concentration of 2-pentanol concentration of metal catalyst initial concentration of metal catalyst concentration of metal hydride concentration of base or hydroxide ion initial concentration of base or hydroxide ion Chemical Abbreviations l ,2-PD 1 ,3-PD 2,4-PD HPA HPD HMPD PTD TPD 1,2-propanediol 1 ,3-propanediol 2,4-pentanediol 3-hydropropionaldehyde 5-hydroxy-2-phenyl- 1,3-dioxane 4-hydroxymethyl-2-phenyl-1,3-dioxolane 2-phenyl-5-tosy1- 1,3-dioxane 2-tosyloxy-1,3-propanediol xii CAT PART I CATALYTIC H YDROGENOLYSIS OF S UGARS Curren natural gas 1 resources. u 838 diminisl TUiure. Cor r35130031016 the biomag 01011133 Tag The si 01 311113;] C11 Declining tr “ESTES. III CHAPTER 1 GENERAL INTRODUCTION Currently, a large portion of the chemical process industry depends on petroleum and natural gas for feedstocks. Considering the finite and nonrenewable nature of these fossil resources, this dependence is a severe handicap. As the reserves of petroleum and natural gas diminish, the chemical process industry will possibly face feedstock problems in the future. Consequently, we will face a shortage of synthetic chemicals, which are largely responsible for our current living standard. In light of these possibilities, development of the biomass conversion processes that produce useful chemicals from the renewable biomass resource is of great significance. The significance of biomass conversion is also comprehensible from the perspective of enhancing biomass utilization. Biomass (lignocellulose) is an abundant material occurring mainly in the form of forestry residues, agriculture residues and municipal solid wastes. It is estimated that each year the US. alone generates about 1 billion dry tons of these residues and wastes (Barrier and Bulls, I992). The annual production of biomass in the world is estimated to be as high as 1011 to 1012 dry tons (Grohman et aL, 1993). At the current time, the greatest use of biomass is to burn it as a low-valued fuel, and a large portion of biomass is simply discarded as waste. Biomass conversion processes can convert biomass into a variety of value-added useful chemicals, and thus upgrade the use of biomass. In the hydrogenoly sugar hydro ; glycol; the g HF sacchar. propylene gl While sugar i In spite 10 sugar h) hi'drogenoly figDOCeIIUTOS as Sugar Crop is “OI neceg. which also a which Consis hydrogenglys condUCted is proYTde d in U‘ 2 In the current research, three biomass conversion processes, namely sugar hydrogenolysis, glycerol dehydroxylation and HF saccharification, are investigated. The sugar hydrogenolysis process converts sugars into glycerol, propylene glycol and ethylene glycol; the glycerol dehydroxylation process converts glycerol to 1,3-propanediol; and the HF saccharification process converts lignocellulosic biomass into sugars. Glycerol, propylene glycol, ethylene glycol and 1,3-propanediol are value-added useful chemicals; while sugar is an important intermediate in the conversion of biomass to useful chemicals. In spite of the fact that the HF saccharification process naturally provides feedstock to sugar hydrogenolysis, it needs to be noted here that the feedstock of sugar hydrogenolysis is not necessarily from HF saccharification. It may be from other lignocellulose saccharification processes. It may also be from other sugar sources, such as sugar crops and starchy materials. Similarly, the feedstock of glycerol dehydroxylation is not necessarily from sugar hydrogenolysis. It may come from sugar fermentation, which also converts sugars to glycerol, or it may come from processing of fatty materials, which consists of a different type of biomass. In this research, studies on the sugar hydrogenolysis, glycerol dehydroxylation and HF saccharification processes have been conducted as independent sub-projects. A brief discussion on each of these projects is provided in the following. W The purpose of sugar hydrogenolysis is to convert sugars to glycerol, propylene glycol and ethylene glycol. These chemicals have extensive uses. Glycerol is used in alkyd resin, ester gums, polyethers, pharmaceuticals, perfumeries, food additives and humectants. Ethylene glycol is used in antifreeze, hydraulic fluids, paints, deicers and alkyd and polyester resins. Propylene glycol has similar applications to 4 3.. .._——_4 ethylene gly cosmetic sot glycerol. pro At the as an intern chlorohydrin produce glyc epichlorohyd; gli'cidol as int as a b)’-ptodu< Pmpanedjol a4! Chemicals and able [0 prOdUCl Pmpylene g1.“ enl'imnmentauy methOds. A5 menu-0n fennentauOn, in . 3 ethylene glycol, and is additionally used as tobacco humectants, food additives and cosmetic softening agents, owing to its non-toxic nature. Because of their extensive uses, glycerol, propylene glycol and ethylene glycol are currently produced in large volumes. At the current time, ethylene glycol is produced from ethylene with ethylene oxide as an intermediate. Propylene glycol is produced from propylene with propylene chlorohydrin and propylene oxide as intermediates. Several commercial routes exist to produce glycerol from propylene. The major one has ally chloride, dichlorhydrin and epichlorohydrin as intermediates. The more recent one has acrolein, glycidaldehyde and glycidol as intermediates. A small portion of glycerol is also produced from fatty material as a by-product of soap production. Glycerol, propylene glycol, ethylene glycol and 1,3- propanediol are all petroleum-based chemicals. In view of the importance of these chemicals and the inevitability of petroleum depletion in the future, it is significant to be able to produce them from renewable biomass resources. Production of glycerol and propylene glycol by sugar hydrogenolysis also has potential to eliminate the environmentally unfriendly chlorohydrin intermediates from the current production methods. As mentioned previously, the conversion of sugars to glycerol can also be done by fermentation, in lieu of hydrogenolysis. Nevertheless, the hydrogenolysis process has some unique advantages. Theoretically, the hydrogenolysis reaction preserves all the carbon atoms in the original sugar molecules, and thus can potentially provide a high glycerol yield. The theoretical glycerol yield of fermentation, in contrast, is limited to 75% in aerobic fermentation and to 58% in anaerobic fermentation (Wise, 1983). This limitation results from the carbon loss as carbon dioxide in fermentation. The second advantage example. Consideri lignocellu pentose. I The disad‘ not yet 5: glycol. etl ensisionec process do The . rEaction n tnechamSn Plausible , Cleavage i] 1.3116ng Perfoxme d 4 advantage of the hydrogenolysis process is that it can also use pentose, xylose for example, as feedstock. In contrast, the feedstock of fermentation is limited to hexose. Considering that hemicellulose, which is one of the three major components of lignocellulose and accounts for 20 to 30% of its total weight, is composed mainly of pentose, this feature of sugar hydrogenolysis is of great importance to biomass utilization. The disadvantage of the hydrogenolysis process is that, at the current stage, this process is not yet selective. The current hydrogenolysis processes produce glycerol, propylene glycol, ethylene glycol, as well as other polyhydric alcohols, in a mixture. However, it is envisioned that the selectivity of sugar hydrogenolysis can be improved through proper process development. The current research on sugar hydrogenolysis has focused on understanding of the reaction mechanism and of the factors controlling the reaction selectivity. Reaction mechanism and selectivity control are two major issues of sugar hydrogenolysis. Several plausible mechanisms have been proposed by previous researchers to explain the C-C cleavage in sugar hydrogenolysis. In the current work, these mechanisms are critically reviewed on the theoretical grounds and against the hydrogenolysis experiments performed in this research using 1,3-diol model compounds. This leads us to identify the retro-aldol reaction of a B—hydroxyl carbonyl precursor as the CC cleavage mechanism of sugar hydrogenolysis. Other mechanisms are excluded based on theoretical considerations and experimental results. The hydrogenolysis experiments performed in this work have also provided for the first time direct evidence that the C-0 cleavage is through dehydration of the same B-hydroxyl carbonyl precursor just mentioned. Establishment of the above reaction mechanisms makes it possible for one to adopt a 5 rational approach in control of the selectivity of the sugar hydrogenolysis process. The selectivity study in the current work has focused on control of the C-0 cleavage in sugar hydrogenolysis. It is envisioned that, if the C-0 cleavage could be somehow controlled, the reaction products of sugar hydrogenolysis would be significantly simplified, and the reaction would be directed to produce the highest-valued product, glycerol. The selectivity study has also been carried out using a 1,3-diol model compound (2,4-pentanediol). The study on the competition between CC and C-0 cleavage in sugar hydrogenolysis is greatly facilitated by use of this model compound. Thus, an experimental study has been carried out in this work to investigate the effect of temperature, base concentration, hydrogen pressure, catalyst amount and catalyst type on the C-C v.s C-O selectivity, and a simulation study has been conducted to rationalize the experimental results and to look into ways to improve the selectivity. It is found that the temperature has little effect on the selectivity; increasing base concentration can increase the selectivity to certain extent; increase in the amount of catalysts used has an adverse effect on the selectivity; and the selectivity varies widely with the type of catalysts used. The effect of hydrogen pressure on the selectivity is the most interesting. At low pressure, the selectivity increases with the pressure; at medium pressure, the selectivity is not sensitive to the change in the pressure; while at high pressure, the selectivity decreases as the pressure is increased. Of the seven different types of catalysts that has been studied in our experiments, the best one is barium-promoted copper chromite, which has furnished a C-C vs. C-O selectivity of 2.5 in hydrogenolysis of 2,4—PD. Our simulation study also identifies a catalyst with strong capability to hydrogenate ketones and aldehydes but only moderate ability to hydrogenate (LB-unsaturated ketones as the one with high potential to 6 deliver improved selectivity. These results, both experimental and theoretical, point out the possible directions that further selectivity study may follow. W The purpose of glycerol dehydroxylation is to convert glycerol to 1,3-propanediol. 1,3-propanediol is a very high-valued specialty chemical, with a current marketing price of $12 per pound. Because of its high price, the use of 1,3- propanediol is limited to specialty polyester fibers, films and coatings. 1,3-propanediol does not impart the stiffness of ethylene glycol, yet avoids the floppiness of 1,4-butanediol and 1,6-hexanediol. Polyesters based on 1,3-propanediol have unique features that are difficult to obtain from other glycols. For instance, polytrimethylene terephthalate (P'IT), a new polyester based on 1,3-propanediol has the elastic recovery of nylon and the chemical resistance of polyester (Stinson, 1995). It is positioned by its manufacturers as a rival to tradition nylon in carpet area (Chapman, 1997). At the current time, the production scale of 1,3-propanediol is small, but is quickly expanding. Dedussa, a German-based company, has just installed 2.2 million lb per year capacity at its plant in Antwerp, Belgium (Stinson, 1995). Shell has also plan to install a 160-million-pound unit at Geiser, Louisiana by early 1999 (Chapman, 1997). The need for 1,3-propanediol is great in production of specialty polymers. Dedussa, Shell and Dupont all have plan to startup or expand their PTI‘ plants (Chapman, 1996). If a low cost process can be found, it is expected that 1,3-propanediol could quickly become a commodity chemical. Cunently, 1,3~propanediol is made by hydration of acrolein to B- hydroxypropionaldehyde, which yields 1,3-propanediol upon hydrogenation. Acrolein is a very dangerous chemicals. It must be synthesized on site. The yield from the acrolein 7 process is very low (Gunzel et al., 1991). The problem is with the first step of the process. Acrolein has a large tendency to polymerize through self-condensation, and hydration must compete with this reaction. 1,3-propanediol may also be made via carbonylation of ethylene oxide to B—hydroxypropionaldehyde. The yield of 1,3—propanediol in this process is also poor, however. Lack of conversion efficiency in the current commercial processes is the main reason for the current high price and small production scale of 1,3- propanedioL Effort to convert glycerol to 1,3-propanediol has been reported. These processes will be reviewed later in Part 2 of this dissertation. Briefly, these processes are still under investigation and are not yet commercializable. In the current research, a new concept to convert glycerol to 1,3-propancdiol is proposed. This new concept selectively transforms the second hydroxyl group of glycerol into a tosyloxy group with the help of a group protection technique, and then removes it by hydrogenolysis. The technical feasibility of this new concept is verified experimentally in this work. Before this work, hydrogenolysis of tosylates are generally effected with base metal hydrides, such as LiAlH4 and LiHBEt3. In this work, for the first time, tosylates with a hydroxy group adjacent to the tosyloxy group are hydrogenolyzed with molecular hydrogen in the presence of a nickel or ruthenium catalyst. W As previously mentioned, in addition to the HF process, other saccharification concepts, including dilute mineral acid, concentrated mineral acid and enzymatic processes, are available to convert lignocellulosic biomass into sugars. The research on HF saccharification in the current project is motivated by the potential advantages it has over other processes. As summarized by Hawley et al. (1986) and Rotter to 95" recon theli CXpe‘ are. 80m Hall bell the in ma: 8 Rorrer (1989), the HF saccharification processes is characterized by high sugar yield (85 to 95% of the theoretical yield), easy catalyst recovery (up to 99.6% of the HF may be recovered), and little pretreatment requirement (only drying and chipping). In addition, the HF process operates under ambient reaction time, needs short reaction time and can produce a lignin residue suitable for further conversion. The HF vapor used in the process is recycled internally. The HF residue left with the reacted substrates is dissolved in water and neutralized in post-hydrolysis. Both measures minimize the HF emission from the HF saccharification process. The current work on vapor-phase HF saccharification is a continuation of the research initiated by Rorrer et al (1988), Moring (1989), Rorrer (1989) and Reath (1989) at Michigan State University. It has focused on modeling of the HF adsorption process in a packed-bed reactor. However, experiments have also been performed with a bench-scale reactor to study the behavior of HF adsorption in the packed-bed reactor. Similar experiments has been done previously by Reath (1989), but the experiments in this work are conducted in a broader range of operating conditions, which leads to the finding of some behavior of HF adsorption that has not been observed by Reath. TWO models of HF adsorption have been developed. One is phenomenological nature, and is developed based on experimental observations of the HF adsorption behavior in a bench-scale, packed-bed reactor. The second model is more advanced than the first one. It is developed based on a detailed analysis of the HF flow, the intrinsic HF adsorption and the heat transfer processes occurring in the reactor during HF adsorption. This model provides a great deal of insight into the HF adsorption process in a packed-bed reactor, and it can be used to predict the HF loading distribution, as well as the overall HF loading. as modeling Wt adsorption p: Results Pan l preset dehydroxylal Pan lc Chapter 3 de 01 the mechar Pan llc concept as v documents th ”inpatient adiscllssionc Based 0: Published in me“ Papers To avoid m Therefore, or briefly descn' sacchaliftcatjc inchapmr 10$ 9 loading, as functions of the adsorption time and operating conditions. The above modeling work provides a theoretical foundation to design and operation of the HF adsorption process in a packed-bed reactor. Results from the current research are presented in this dissertation in three parts. Part I presents the work on sugar hydrogenolysis; Part 11 presents the work on glycerol dehydroxylation; Part III presents the work on HF saccharification. Part I contains four chapters: Chapter 2 is an introduction to sugar hydrogenolysis; Chapter 3 describes the experimental system and methods; Chapter 4 presents the results of the mechanism study; and Chapter 5 presents the results of the selectivity study. Part 11 contains three chapters: Chapter 6 presents the new glycerol dehydroxylation concept, as well as a review of the current glycerol conversion processes; Chapter 7 documents the experimental procedures used to carry out the conversion of glycerol to 1,3-propanediol; and Chapter 8 presents the experimental results. Chapter 8 also includes a discussion on the future work that may be done on glycerol dehydroxylation. Based on the research on HF saccharification, two papers have been written and been published in Chemical Engineering Science (Wang et al., 1994; Wang et al., 1995a). These papers document basically all the work that has been done on HF saccharification. To avoid rewriting, only abstracts of these papers are included in this dissertation. Therefore, only one chapter, Chapter 9, is allocated to Part III in this dissertation, to briefly describe the rationale, scope and accomplishments of the research on HF saccharification. Following Part III, a general summary of the current research is provided in Chapter 10 and recommendations on future work are made in Chapter 11. 2.1 CHAPTER 2 INTRODUCTION: SUGAR HYDROGEN OLYSIS 2.1 Literature Review Under high temperature and high hydrogen atmosphere, sugars can be catalytically hydrocracked into lower polyhydric alcohols in the presence of transition metal catalysts. This process is called hydrogenolysis, because the bond cleavage reactions are accompanied with hydrogen addition as shown below: R3C —CR’3 + H2 -—.-> R3CH + HCR’3 R3C —OH + H2 ——.> R3CH '1' H20 Sugars are available from a variety of biomass raw materials. Hydrogenolysis provides a potential process to convert sugars to synthetically important polyols, including glycerol, propylene glycol and ethylene glycol. In the literature, sugar hydrogenolysis is discussed indistinguishably from sugar alcohol hydrogenolysis, because of the close relationship between these two reactions. Sugar hydrogenolysis and sugar alcohol hydrogenolysis take place under the same conditions, yield the same products and have identical bond cleavage mechanisms and bond cleavage precursors (sugars), as will be shown later in Chapter 4. As a matter of fact, in sugar hydrogenolysis, a significant portion of the starting sugar will be hydrogenated to sugar alcohol before being finally hydrogenolyzed into lower polyhydric 10 alcohols I hydrogenolys understood a context indie The sug (Connor and sasstarted n Laboratory (1 been increas 511331 hydrug mitten in bi( hl‘drogenoly mLSPIOCcss, Studies 011 l: Sawfly ar 11.1 studie In the Purpose of . upenmenm’ nickel 0n he ll alcohols. In some occasion, sugar alcohols have been used to study the sugar hydrogenolysis process. In the following, references to sugar hydrogenolysis should be understood as the hydrogenolysis of both sugars and sugar alcohols, except where the context indicates otherwise. The sugar hydrogenolysis reactions have been performed since as early as the 1930’s (Connor and Adkins, 1932). However, the research for the purpose of biomass conversion was started much later. The first one was reported in 1950’s, at the US. Forestry Products Laboratory (Clark, 1958). Since then, the body of literature on sugar hydrogenolysis has been increasing steadily. Like many other biomass conversion projects, the research on sugar hydrogenolysis underwent a “boom” period in the 1980’s, as a result 'of the general interest in biomass conversion stimulated by the oil crisis in 1970’s. The research on sugar hydrogenolysis has been conducted in many countries, and has covered various aspects of this process. In the following, the available literature is reviewed according to four topics: studies on process development, studies on reaction mechanism, studies on reaction selectivity and studies on reaction kinetics. 2.1.1 Studies on Process Development In the 1950’s, Clark (1958) started the first research on sugar hydrogenolysis for purpose of biomass conversion at the US. Forestry Products Laboratory. In his experiments, sorbitol was reacted under the hydrogenolysis conditions in the presence of a nickel on kieselguhr catalyst. Calcium hydroxide was added to promote the reaction. The reaction was carried out in aqueous phase in a batch reactor. The temperature was at 215 to 240 'C, and the hydrogen pressures was at 2000 to 5600 psi. The identified products and (Boel ofcat and.B Pedon 12 included glycerol, propylene glycol, ethylene glycol, erythritol and xylitol. The product composition varied with the reaction time. The concentration of xylitol. erythritol, glycerol increased during the initial period of an experimental run, but after reaching a maximum, they started declining. The maximum yield of glycerol obtained in Clark’s experiments was 40%. At about the same time as Clark’s work, according to Boelhouwer er al., Geerards and co-workers investigated the hydrogenolysis of sucrose with HI as a catalyst (Boelhouwer er al., 1960). The products they obtained had very low oxygen contents. In their own work, Boelhouwer er al. (1960) hydrogenolyzed sucrose with a variety of catalysts, including nickel on kieselguhr, nickel-thorium oxide on kieselguhr, and MgO and BeO promoted copper chromite. Calcium oxide was also added. The reaction was performed in a rotating autoclave with methanol as the reaction medium. The temperature range was between 195 and 250 °C, and the hydrogen pressure range was between 150 and 200 atrn. The products were fractionated by distillation. In one experiment, the glycerol fraction was reported to account for 61% of the product. Since each fraction covered a broad boiling range, exact products were not determined. However, it was believed that glycerol, propylene glycol and ethylene glycol were included in the products. Van Ling and co-workers (van Ling et al., 1967; van Ling and Vlugter, 1969) hydrogenolyzed sucrose, glucose, fructose and inan with CuO-CeOz-SiOZ. The reaction was done in a 250 mL autoclave, at a temperature of 200-250 °C, a hydrogen pressure of 100-300 atm, with addition of caladium hydroxide. In the hydrogenolysis of sucrose, ethylene glycol, propylene glycol, glycerol, teritols, pentinols, hexitols and dehydrated hexitols were obtained. In hydrogenolysis of other sugars, the products were similar. Van Lll process [0 pi firStWaS a I“! m The teat“ Because 015 yield tbelow I [an]: reactors operating V'Oll. reanon syster (CuOCeOI-Si A great hydrogenolyzet hydrogenolyzet platinum on car carbonaceousp using a sulfur-: 64.6% in one e: mtheniurn and r 13 Van Ling er al. (1970) further developed a continuous pilot-scale hydrogenolysis process to produce glycerol from sucrose. Two reactor configurations were studied. The first was a tubular reactor, composed of a 2 m long stainless steel tube with a diameter of 2 cm. The reactor was filled with 1/8—inch Raschig rings to promote the gas-liquid contact. Because of sucrose decomposition, this reaction system furnished only a poor glycerol yield (below 20%). The second reaction system was composed of two continuous stirred- tank reactors (CSTR) in series. Each reactor had a total volume Of 500 ml and an operating volume of 250 ml. Sucrose was fed in 75% methanol-water solution. With this reaction system, a glycerol yield up to 30% was obtained. The catalyst used in this study (CuO-CeOz-SiOz) was the same as in their previous research. A great deal of work was accomplished in the 1980’s. Chao et al. (1982) hydrogenolyzed sorbitol with a supported nickel catalyst in aqueous phase. Arena (1983) hydrogenolyzed sorbitol, glucose and sucrose in the presence of nickel on kieselguhr, platinum on carbon, or metals (including palladium, rhenium and copper) deposited on a carbonaceous pyropolymer structure. Duback and Knapp (1984) hydrogenolyzed sorbitol using a sulfur-modified ruthenium catalyst, and achieved a propylene glycol yield of 64.6% in one experiment Ariono er al. (1986) hydrogenolyzed glucose with palladium, ruthenium and rhodium supported on carbon. Tanikella (1983a; 1983b) hydrogenolyzed sorbitol in both aqueous and non-aqueous solutions to produce ethylene glycol. In the work by the Motassier group (Sohounloue et al., 1983; Montassier et al., 1988, 1989 and 1991), various sugar alcohols, including sorbitol, xylitol, erythritol and even glycerol, were subject to the hydrogenolysis conditions. The catalysts used included RulSiOz, Ranney Cu and other metal based catalysts. In all t hgtgrpgelldtlu (Andrews an. and We} pyrrolidinone arm The It heterogenous lithlnrctose glycerol relat‘ hydrogenation to the previous favored than in Studies or al.1975;Piltlt llWage, lon pmdthlS. The “lililibn'um. 14 In all the work reviewed above, the hydrogenolysis has been carried out with heterogeneous catalysts, but a process was also reported using homogenous catalysts (Andrews and Klaeren, 1989). Andrews and Klaeren hydrogenolyzed fructose, glucose and manna-heptulose with H2Ru(PPh3). The reaction was done in N-methyl-Z- pyrrolidinone solvent, at a temperatures of 50 to 100 “C and a hydrogen pressures of 20 arm. The reaction conditions were significantly less severe than those used with heterogenous catalysts. In most of the experiments, KOH was added as a co-catalyst. With fructose as the starting compound, this process provided very good selectivity toward glycerol relative to other degradation products. However, because of the rapid simple hydrogenation of fructose (to sorbitol and mannitol), the overall glycerol yield was similar to the previous processes. With glucose as the starting compound, glycerol was much less favored than in the case of fructose. Studies on separation of the hydrogenolysis products were also reported (Sokolvo er al., 1975; Pikkov et al., 1973; Solkovo et al., 1977). Sokolvo et al., (1975) developed a two-stage, low pressure distillation process to separated glycerol from other reaction products. They also determined the activity coefficients for the glycerol-water vapor equilibrium. 2.1.2 Studios on Reaction Mechanism In sugar hydrogenolysis, both CC and 00 are subject to cleavage. These reactions are promoted by transition metals and base co—catalysts. Cleavage of C-0 bonds, as Montassier er al. (1988 and 1989) proposed, is through dehydration of a B—hydroxyl carbonyl: on or RCHCHCH On The structure molecule. and The direct pro carbonyl. whiej just discussed. hydrogenation s The origin “Plain the C{ reaction: 9H on RCHCHCHR'; ! 0H Wrens and K15 that the Primary Accord“ g to this Omit of this Timmy hl'drc Lam bOWeye 15 carbonyl: OH OH O OH O OH I I -H2 II 1 +120 ll +H2 | RCHCIZHCHR’ ———> RCCIZHCHR’ ——-> RCCII= CHR’ —-> RCHCHCHZR’ OH OH OH OH The structure of this B-hydroxyl carbonyl is already contained in an open-chain sugar molecule, and it may be generated from a sugar alcohol molecule by dehydrogenation. The direct product from dehydration of the B-hydroxyl carbonyl is an Ot,B-unsaturated carbonyl, which yields a polyhydric alcohol upon hydrogenation. In the reaction scheme just discussed, the dehydration step is catalyzed by bases, and the dehydrogenation and hydrogenation steps are catalyzed by transition metal catalysts. The original mechanism proposed by the Montassier group (Sohounloue, 1983) to explain the CC cleavage in sugar and sugar alcohol hydrogenolysis is the retro-aldol reaction: (I’ll 9“ I”) (RH O o H OH , 'H II II 1’ 2 l RCHCIZHCHR _.l Rcclzncrm' _. Rccnzom HCR’ -— RCHCH20H +HOCH2R’ OH OH Retro-aldol Andrews and Klaeren (1989) suggested the same mechanism, based on their observation that the primary C-C cleavage site is B to the carbonyl group in sugar hydrogenolysis. According to this mechanism, the QC cleavage precursor is again a B-hydroxyl carbonyl. Cleavage of this B—hydroxyl carbonyl leads to an aldehyde and a ketone, which are subsequently hydrogenated to alcohols. Later, however, Montassier er al. (1988) felt that it was difficult to explain the absence of“ and other 51 anc‘her mecl hydrogenoll'f on or cugcncn 1 OH This mechan: hydrogenolysi: allows format formaldehyde ; xylitol and sort the mechanism: 0H RCHCHCHCT ' l onOnor Thh mechanism The reactio hsuga; hldngc.‘ NIPPON am" of [hf 16 absence of methanol and the presence of C02 in the hydrogenolysis products of glycerol and other sugar alcohols with the retro-aldol mechanism. Therefore, they proposed another mechanism, namely the retro-Claisen reaction, for the C-C cleavage in glycerol hydrogenolysis: OH OH o O ” l -2H2 ll III-151V.“ Retro—Claisen HCOH CHZC'IHCHZ —> HCCIEHCH —> HO-C\r\ + J H/ CH OH H20 OH OHCH=CHOH OH This mechanism allows formation of formic acid which decomposes under the hydrogenolysis conditions to form C02. The mechanism based on the retro-aldol reaction allows formation of formaldehyde rather than formic acid. Upon hydrogenation, formaldehyde yields methanol. TO explain the CC cleavage in the hydrogenolysis of xylitol and sorbitol, Montassier et al. (1988) also proposed the retro-Michael reaction as the mechanism: ?“ 1’” 2n it it it it ' 2 RCHCIZHCIIHCHCHR’ —-- RCC|HC|HCIHCR’ ———- RC(|I = (IZH + OHCHZCR’ OHOH OH OHOH OH Retro-Michael OH OH This mechanism requires a S-dicarbonyl as the bond cleavage precursor. The reaction mechanisms just reviewed are all consistent with the products obtained in sugar hydrogenolysis. Other than that, no direct experimental evidence is available to support any of the bond cleavage mechanisms proposed. 2.1.3 Studi Studies Sohounloue noose, glu selectivity g representing Ell'COU in the the percentag Percentage. It Pmducts inch; that the CS ‘ Variables. such dipendgm Ont incheased. It yy matimplhmo Sbhounlo sorbitoh. He inchted. He; 17 2.1.3 Studios on Reaction Selectivity Studies on the reaction selectivity have been reported by van Ling er al. (1969) and Sohounloue er al. (1983). Ling er al. studied the effect—hf the starting sugars, including sucrose, glucose, fructose and inulin, on the reaction selectivity. He defined three selectivity parameters for sugar hydrogenolysis: 1) CS, or cleavage selectivity, representing the percentage of the C3-C3 cleavage products (glycerol and propylene glycol) in the total cleavage products; 2) HS, or hydrogenation selectivity, representing the percentage of glycerol in the total C3-C3 cleavage products; and 3) CR, or cleavage percentage, representing the percentage of the bond cleavage products in all reaction products including hexitols from the starting sugar via simple hydrogenation. He found that the CS varies with the starting sugar, but was independent of the other reaction variables, such as catalyst type and base concentration. The H.S, as he reported, is also dependent on the starting sugar, and in addition it decreases as the reaction temperature is increased. It was also found that the CF increases with decreasing hydrogen pressure, and that it is promoted by small addition of Ca(OH)2. Sohounloue et al. (1983) studied the reaction selectivity in the hydrogenolysis of sorbitols. He observed that the CC cleavage became random as the temperature increased. He also Observed that the CC v.s. C-O selectivity, as defined by the ratio of the glycerol to propylene glycol, decreased with temperature in a basic reaction medium but increased in a neutral medium. Both of the studies by Ling er al. and by Sohounloue et al. were experimental in nature, and the mechanisms controlling the reaction selectivity were not well reflected. 1.1.4 Studit The tir: (1958?). who hydrogen pre in a pressure activation ene determined to Chang 6. 5113th differ: Rho nickel a in the concentr mil11150.8 orde. am011111 01 cam exl‘el‘lrrlentg by m sufficienu) Chang ct al. (_ Wilding 01 5U. h, Merton-5'5 ‘ TWWMU 18 2.1.4 Studies on Reaction Kinetics The first empirical kinetic study on sugar hydrogenolysis was reported by Clark (1958), who found that the hydrogenolysis of sorbitol is first-order in the substrate. The hydrogen pressure does have effect on the reaction rate, but this effect has been considered in a pressures dependent activation energy. For a hydrogen pressure of 5600 psi, the activation energy was determined to be 44 kcal; for a hydrogen pressure of 2900 psi, it was determined to be 33 kcal. Chang et al. (1985) studied the kinetics of sorbitol hydrogenolysis also, but used a slightly different catalyst. Instead of nickel on kieselguhr as in Clark’ work, they used Raney nickel as the catalyst. They found that the hydrogenolysis of sorbitol is first order in the concentration of sorbitol, as Clark did. In addition, they found that this reaction is minus 0.8 order with respect to the hydrogen pressure and second order with respect to the amount of catalyst. The activation energy was determined to be about 22 kcal/mol. The experiments by Chang et al. were conducted in an one-liter autoclave. The agitation rate was sufficiently high to eliminate the effect of external mass transfer. In the work of Chang et al. (1985), the kinetics of glycerol hydrogenolysis was also studied. The reaction was also found to be first-order in the substrate. The Work by Tronconi et al. (1992) was the only one reportedly on mathematical modeling of sugar hydrogenolysis. This work provided a kinetic model for sorbitol hydrogenolysis which was different from the previous work. The authors claimed that this model had been developed based on their analysis of the whole reaction network of hydrogenolysis (which is not shown in the paper). The developed model was subsequently adopted in a simple pseudo-homogeneous non-isothermal plug flow reactor model. Corr and pilot-scz kinetic erpre A kinet (1973). Hon 2.2 Object A major control. As di generally 110n- ih‘tol are proc levels. It is cl: h«‘d”«‘t‘=’irlolysis Another r KI”white oil. imMince for Review, three Michaey have b fishl'drition mgc hhhr’cauy n0“ heCCcymgt3 vilified II] View of [m 19 model. Comparison of this reactor model against their experimental data from their micro and pilot-scale reactors allowed them to determine the model parameters present in the kinetic expression. A kinetic study on sugar hydrogenolysis was also reportedly done by Stefoglo er al. (1973). However, the detail of this study is not clear. 2.2 Objective and Scope of Current Research A major issue in development of the sugar hydrogenolysis process is selectivity control. As discussed in the Literature Review, current sugar hydrogenolysis processes are generally non-selective. In these processes, glycerol, propylene glycol and ethylene glycol are produced in a mixture, along with other polyhydric alcohols occurring at lower levels. It is clear that a more selective process needs to be developed, in order for sugar hydrogenolysis to be commercially viable. Another major issue of sugar hydrogenolysis concerns the reaction mechanism. Knowledge of how the CC and C-0 bonds are broken in sugar hydrogenolysis is of great importance for control of the reaction selectivity. So far, as discussed in the Literature Review, three competing mechanisms, namely retro-aldol, retro-Claisen and retro- Michael, have been proposed to explain the CC cleavage in sugar hydrogenolysis. The dehydration mechanism of C-0 cleavage is not disputed, but direct experimental evidence is basically non-existent. Before a useful strategy for selectivity control can be developed, the C-C cleavage mechanism must be clarified and the C-0 cleavage mechanism must be verified. In view of the above situation, the objectives of the current research are determined 20 as: 1) to carry out a mechanism study to discriminate and verify the existing bond cleavage mechanisms of sugar hydrogenolysis; and 2) to carry out a selectivity study to investigate the factors controlling the selectivity of sugar hydrogenolysis. In this capacity, a series of 1,3-diol model compounds are hydrogenolyzed, to establish the mechanism of CC and C- 0 bond cleavage in sugar hydrogenolysis. The experimental work confirmes our theoretical conception that the QC bond breaks via the retro-aldol reaction of a [3- hydroxyl carbonyl precusor, and it also provides for the first time direct evidence that the C-0 cleavage is through dehydration of the same B-hydroxyl carbonyl precursor just mentioned. In the selectivity study, the research focuses on control of the C-0 cleavage in sugar hydrogenolysis. It is envisioned that the reaction products of sugar hydrogenolysis would be considerably simplified, and the reaction would be directed to produce the highest- valued glycerol, if the C-0 cleavage can be controlled somehow. Thus in this work, the competition between the C-C and C-0 cleavage is studied again using a 1,3-diol model compound (2,4-pentanediol). Experiments, as well as computer simulation, are conducted to investigate the effects of temperature, base concentration, hydrogen pressure, catalyst amount and catalyst type on the selectivity, as well as possible ways to improve this selectivity to C-C bond cleavage. CHAPTER 3 EXPERIMENTAL SYSTEM AND METHODS 3.1 Experimental System and Procedure In this work, 1,3-diols with various structural features were hydrogenolyzed as model compounds to study the reaction mechanism and selectivity of sugar hydrogenolysis. These hydrogenolysis experiments were performed in a specially designed, stainless steel reactor of 50 mL capacity. The detailed design of this reactor is given in Figure 3-1. This hydrogenolysis reactor is basically a pressure vessel (note that hydrogenolysis is effected at high temperature and high hydrogen pressure), and is equipped with a magnetic stirring bar. Because of the need to regularly take samples during the reaction, a sampling port is provided with the reactor. This sampling system is composed of a sampling valve, a sampling tubing and an inlet filter emerged in the reaction medium. This filter has a 0.45 u rating. It can prevent solid catalyst particles from entering the sampling system and generate clean samples suitable for direct HPLC and GC injection. This filter also reduces the pressure exerted on the sampling valve, which levitates the requirement on the sampling valve. The sampling tubing is very thin, with an inner diameter of about 0.1 mm. The dead volumes of the sampling valve and the inlet filter are 0.2 and 0.4 uL, respectively. The whole sampling system has a total hold volume of only several microliters. This guarantees that the samples taken from this system always 21 ///% / / / w /< Z \[]\ Figure 3-1. Illustration of Hydrogenolysis Reactor 23 To vacuum H2 —’ Sampling valve Temperature controller _J¢ .— ..... - I O O Stirring gas (N2) ! 5' ' ' '— —-* i I 3 I l ; I I I Heating coil Oil bath\ __ . _|__ / x - O — . '- @ § @//Thermal couple 0 a @ o 32 @ o ' @ . @ ° . Magnetic stirrer / Figure 3-2. Schematic Illustration of Experimental System 24 represent the current condition in the reactor. The reactor is heated in a silicone oil bath. The heat is provided by an electric heating coil emerged in the oil bath, which is constantly stirred by bubbling a nitrogen flow into it. The temperature of the oil bath is controlled by a temperature controller, which controls the oil temperature within :l: 3 °C of the set point. The reaction medium is stirred by the magnetic stirring bar inside the reactor. To provide the magnetic force to the stirring bar, the whole oil bath is placed on a magnetic stirrer. The hydrogen pressure is provided to the reactor by a hydrogen cylinder with high pressure regulator for pressure control. To prevent over-pressurization, a pressure relief valve is installed in the H2 supply line as a safety device. A vacuum line is also connected to the reactor for the purpose of initial reactor purging. A schematic illustration of the whole experimental system is provided as Figure 3-2. To carry out the hydrogenolysis reaction, the starting diol, the catalyst, a certain amount of sodium hydroxide (if needed) and a proper amount of distilled water are placed in the reactor. The reactor is purged by alternately connecting it to hydrogen and vacuum. After purging, the reactor is pressurized with hydrogen, and heating and stirring (both inside and outside of the reactor) are started. Thereupon, the reaction starts. During the reaction course, reaction samples are taken regularly from the reactor and are analyzed for composition of reaction products. After the the desired conversion has occurred the reaction is stopped by shutting off the heater and hydrogen supply. 25 3.2 Reaction Regents and Product Analysis All the 1,3-diols and catalysts used in this work are purchased from Aldrich and are used as received. The hydrogen (99.9% pure) is obtained from Purity Cylinder Gases and AGA Gas Products. The hydrogenolysis products from the mechanism study are analyzed by GC and HPLC. The GC column used in this work is a non-polar DB624 column, supplied by J&W Scientific Company. With this column, samples can be injected in the aqueous phase. The HPLC column used in this work is a 25 x 0.46 cm C18 column obtained from Whatrnan (Cat.# 4621-1502). The detectors used with GC and HPLC are a FID detector and a differential refractometer, respectively. In both cases, samples taken from the reactor are injected directly. In the mechanism study, seven different 1,3-diols are hydrogenolyzed, giving seven different sets of products. GC and HPLC methods are developed for each set of these products, and are documented here in Table 3-1 and Table 3-2, respectively. However, the hydrogenolysis products from the selectivity study are analyzed by GC only. Since the selectivity study demands more accurate measurement of the product concentrations, calibration curves are prepared for all the products from 2,4-pentanediol hydrogenolysis (2,4.pentanediol is the model compound used in selectivity study). In the HPLC chromatograms of actual reaction products, a skewed peak has often been seen at a retention time well earlier than those for the expected reaction products. This peak is identified as the impurities introduced into the reaction medium with Raney Cu and Ni catalysts. An experiment without any 1,3-diol substrate was performed. Samples taken from this experiment are found to give the same skewed peak. Since no 26 Table 3-1. HPLC Methods for 1,3-Diol Hydrogenolysis Products . Conditions Retention Time Product Set #1: 2,4-dimethy1-2,4-pentanediol solvent: 25% acetonitrile/water flow rate: 1.5 mllmin (isocratic) temperature: 40 ° C 1. 2,4-dimethyl-2.4-pentanediol: 2.92 min Product Set #2: 2-methyl-2,4-pentanediol solvent: 42% methanol/water flow rate: 0.8 mllmin (isocratic) temperature: 40 °C 1. acetone: 3.98 min 2. isopropanol: 4.25 min 3. 2-methyl-2,4-pcntanediol: 4.91 min 4. 4-methyl-2-pentanol: 11.85 min Product Set #3: 2,2,4-trimethyl-1,3-pentanediol solvent: 25% acetonitrile/water flow rate: 2.0 mllmin (isocratic) temperature: 40 °C l. methanol: 1.30 min 2. isobutanol: 2.31 min 3. 2.2,4-trimethyl-1,3-pentanedi01: 3.62 min 4. 2,4-dimethyl-3-pentanol: 9.98 min Product Set #4: 2,2-dimethy1-1,3-propanediol solvent: 10% acetonitrile/water flow rate: 1.5 mllmin (isocratic) temperature: 60 ° C 1. methanol: 1.80 min 2. 2,2-dimethyl-l,3-propanediol: 2.70 min 3. isobutanol: 5.15 min Product Set #5: 2,4-pentanediol solvent: 42% methanol/water flow rate: 0.8 mllmin (isocratic) temperature: 40 ° C 1. ethanol: 3.62 min 2. 2,4-pentanediol: 3.84 min 3. isopropanol: 4.22 min 4. 2-pentanone: 6.54 min 5. 2-pentanol: 8.21 min Product Set #6: 1,3-butanediol solvent: 3% acetonitrilclwater flow rate: 1.2 mllmin (isocratic) temperature: 60 ° C 1. methanol: 2.58 min 2. ethanol: 3.15 min 3. 1.3-butanediol: 3.23 min 4. acetone: 4.50 min 5. isopropanol: 5.00 min 6. 2-butanone: 9.48 min 7. 2-butanol: 10.98 min 8. n-butanol: 13.85min Product Set #7: 1,3-propanediol solvent: 0.3% acetonitrilelwater flow rate: 1.5 mllmin (isocratic) temperature: 60 ° C 1. methanol: 2.04 min 2. 1.3-propanediol: 2.32 min 3. ethanol: 2.90 min 4. n-propanol: 5.51 min 27 Table 3-2. GC Methods for 1,3-Diol Hydrogenolysis Products. Conditions Retention Time Product Set #1: 2,4-dimethyl-2,4-pentanediol initial temperature/time: 150 °C/ 1 min temperature ramp: 40 ° C/min 1. 2,4-dimethyl-2,4-pentanediol: 2.10 min final temperature/time 260 °C/3 min Product Set #2: 2-methy1—2,4-pentanediol initial temperature/time: 80 °C/l min 1- isopropanol: 1-14 min 2. acetone: 1.20 min . 3. 4-methyl-2-pentanol: 2.50 min final temperature/time 260 010 min 4. 2-methyl-2,4-pentanediol: 3.76 min temperature ramp: 40 o C/min Product Set #3: 2,2,4-trimethyl-1,3-pentanediol initial temperature/time: 150 °C/l min 1- 1119111311013 091 min . o . 2. isobutanol: 1.11 min temperature ramp. 30 C/rrun 3. 2,4-dimethy1-3-pentanol: 1.57 min final temperature/time 260 010 min 4. 2,2,4-trimethy1-1,3-pentanediol: 3.35 min Product Set #4: 2,2-dimethyl-1,3-propanediol initial temperature/time: 150 °C/l min 1. methanol: 0.92 min temperature ramp: 30 °C/min 2. isobutanol: 1.11 min 3. 2,2-dimethyl-1,3-propanediol: 2.16 min final temperature/time 260 °C/10 min Product Set #5: 2,4-pentanediol . ethanol: 1.43 min . isopropanol: 1.69 min . 2-pentanone: 3.09 min . 2-pentanol: 3.19 min . 2.4-pentanediol: 4.68 min initial temperature/time: 4o °C/1 min temperature ramp: 40 °C/min final temperature/time 260 °C/3 min Ur«bU->NH Product Set #6: 1,3-butanediol methanol: --- min ethanol: 1.49 min 1,3-butanedi01: --- min acetone: -- min isopropanol: 1.69 min 2-butanone: --- min 2-butanol: --- min n-butanol: -- min initial temperature/time: 40 °C/1 min temperature ramp: 40 ° C/min final temperature/time 260 oC/3 min PSQEAPP'U‘N‘ Product Set #7: 1,3-propanediol methanol: 0.90 min ethanol: 0.96 min n-propanol: 1.09 min 1.3-propanediol: 2.32 min initial temperature/time: 120 °Cll min temperature ramp: 30 ° C/min final temperature/time 260 °C/10 min 99’3"!" 28 1,3-diol is added to the reactor in this experiment, this skewed peak can only result from the catalyst added. 4.] The: As It Clilsen ; hldfllgtn tiller tel amid require a mullet i “Klimt carbon) mom dehj'dtz fal'Orab fisultec d‘Omina re5101017 lbe deb“ CHAPTER 4 MECHANISM OF SUGAR HYDROGENOLYSIS 4.1 Theoretical Considerations As reviewed in the previous chapter, three competing mechanisms (retro-aldol, retro- Claisen and retro-Michael) have been proposed for the C-C cleavage in sugar hydrogenolysis. However, two theoretical considerations prevent us from believing that either retro-Claisen or retro-Michael is a dominating C-C cleavage mechanism over the retro-aldol in hydrogenolysis. First, both the retro-Claisen and retro-Michael mechanisms require a dicarbonyl precursor. The formation of a dicarbonyl presumedly occurs through further dehydrogenation of a mono-carbonyl. As dehydrogenation is thermodynamically unfavorable, the dicarbonyl formed is unlikely to be significant relative to the mono- carbonyl, the precursor of the retro-aldol reaction. Second, the dehydrogenation of the mono-carbonyl is in competition with dehydration, the C-0 cleavage reaction. Since the dehydration is both thermodynamically and kinetically (in presence of bases) much more favorable than the dehydrogenation, the hydrogenolysis products would have exclusively resulted from CD cleavage, had either the retro-Claisen or the retro-Michael been the dominating mechanism of CC cleavage in hydrogenolysis. In contrast, the retro-aldol reaction shares the same precusor as the dehydration, and has the ability to compete with the dehydration. Based on these reasons, we tend to believe that the C-C cleavage occurs 29 though ill lite. absence 0 )ltrtissle not have 1 reaction . by *ogen lormlc at “Miller catalysts 30 through the retro-aldol reaction. The need for a formic acid intermediate to explain the formation of C02 and the absence of methanol in the hydrogenolysis of glycerol was the direct motivation for Montassier et aL to propose the retro-Claisen mechanism. However, this formic acid does not have to be from the retro-Claisen reaction. It may be produced from the Cannizzaro reaction of formaldehyde, which is expected by the retro-aldol mechanism in the hydrogenolysis of glycerol. In the Cannizzaro reaction, formaldehyde is oxidized to formic acid while other carbonyl compounds are reduced to alcohols. This reaction is catalyzed by bases, but has been reported to be greatly enhanced by transition metal co- catalysts (Cook and Mailis, 1981). The Cannizzaro reaction enhanced by the transition OH II II +H2 | RCCH20H 'I“ HCR’ —-"’ RCHCHon 'I' HOCHzR’ III IV V VI retro-aldol T OH OH O OH I I ' H2 II I , RCHCIIHCHR’ —> RCCEHCHR OH 0H 11 I dehydration I - H20 0 OH II “*2 I RCC|3=CHR’ —'* RCHCEHCHZR' OH OH VII VIII Figure 4-1. Mechanism of Sugar and Sugar Alcohol Hydrogenolysis math in the hyd absence nl lne dehylntl. mechanisr processe i Inl-lgnre the bond alcohol I 4.2 E) Iound s Vellfica dchldr. that del ”natal; iICOIlQl 41%; the P111115 ”one Sllg HD‘VEVer, 31 metal is likely to be the side reaction competing with the hydrogenation of formaldehyde in the hydrogenolysis of glycerol and other sugar alcohols, which is responsible for the absence of methanol and the presence of C02 in the hydrogenolysis product. The retro-aldol mechanism needs a B—hydroxyl carbonyl precursor. Considering that dehydration is a well-known reaction of B-hydroxyl carbonyls. The dehydration mechanism of C-0 cleavage is very sound, at least in theory. Thus, the bond-cleaving processe in the hydrogenolysis of sugars and sugar alcohols can be pictured as Figure 4-1. In Figure 4—1, sugars are represented by H and sugar alcohols by 1. Based on Figure 4-1, the bond cleavage mechanisms in sugar hydrogenolysis ought to be the same as in sugar alcohol hydrogenolysis. 4.2 Experimental Verification The reaction mechanism described in Figure 4—1 can explain all the reaction products found so far in sugar hydrogenolysis. Nevertheless, it is still subject to experimental verification. The currently available data provide no clue that the CO bond is broken by dehydration of a B—hydroxyl carbonyl in sugar hydrogenolysis. Experimental evidence that dehydrogenation is a necessary step in the hydrogenolysis of sugar alcohols is also unavailable, and no sugar intermediate has been identified so far in any of the sugar alcohol hydrogenolysis experiments (Clark, 1958; Sohounloue et al. 1983; Motassier et al., 1988 and 1989; Chang et al., 1985). Andrew and Klaeren’s (1989) observation that the primary C-C cleavage site is B to the carbonyl group in sugar hydrogenolysis may be a strong suggestion that the C-C cleavage has occurred through the retro-aldol reaction. However, bond breakage at this site is also expected by the retro-Claisen mechanism. accomnh' dehydrng sugar alt pursued rnechml cnmpnu the Start I.3~dlnl underg: Is a re In the s have IC 32 Clearly, more evidence is desired to validate the mechanism described in Figure 4-1. In the current research, verification of the reaction mechanism in Figure 4-1 is accomplished using 1,3-diol model compound. According to Figure 4-1, a 1,3-diol or its dehydrogenation product is the basic structural unit giving the reactivity to sugar and sugar alcohol molecules. Therefore, a mechanism study of sugar hydrogenolysis can be pursued with 1,3-diol model compounds. Using 1,3-diol model compounds in the mechanism study has several advantages over direct use of the sugar or sugar alcohol compounds. First, the bond cleavage pattern is indicated by the reaction products, since the starting molecule undergoes only one bond-cleaving reaction in the hydrogenolysis of 1,3-diols. In the hydrogenolysis of sugars and sugar alcohols, the starting molecule can undergo a chain of bond-cleaving reactions before further hydrogenolysis is impossible. As a result, the bond cleavage pattern is difficult to recognize based on the final products in the sugar and sugar alcohol hydrogenolysis. Second, sugar and sugar alcohol molecules have too many functional groups, which makes it possible to interpret the experimental results in a variety of ways. In contrast, 1,3-diols possess only the functional groups necessary for the hydrogenolysis to occur, which minimizes the ways to interpret the results. The retro-Michael mechanism is excluded from the hydrogenolysis of 1,3-diols, for instance, because 1,3-diols are incapable of forming the 5—dicarbonyl as required by the mechanism. Last and the most important, 1,3-diols with special structural features can be used, to verify the role played by individual reaction steps in the hydrogenolysis. The 1,3—diols model compounds used in this work include 2,4-dimethyl-2,4— pentanediol (l), 2-methy1-2,4-pentanediol (2), 2,2,4-trimethyl-l,3-pentanediol (3), 2,2- dimethyl-l,3-propanediol (4), 2,4»pentanediol (5), 1,3-butanediol (6) and 1,3-propanediol III- lhe st 2.411131”: Z-ne'lzyl-I ‘l‘lr Jet-mm ziillllrt'l lipcnti Bhutan I.3~pt0p: \ hj'dltlgg lr‘ins'rtli Club's (101,. . *3. to. FIgure miss n Where imm- 10min AC4 Milk» 33 (7). The structures of these compounds are provided in Table 4-1. Table 4-1. Structures of 1,3-diols Used in Mechanism Study N 81116 R1 R2 R3 R4 R5 R6 1 ,3-(1101 2,4-dimethyl-2,4-pentanediol (1) CH3 CH3 H H CH3 CH3 2-methyl-2,4-pentanediol (2) CH3 CH3 H H H CH3 2,2,4-trimethyl-1,3-pentanediol (3) H H CH3 CH3 H CH(CH3)2 ‘EH I“ ?" “rec‘- ere-(R6 2.2-dimethyl-1,3-propanediol (4) H H CH3 CH3 H H R2 R, I, 2.4-pentanediol (5) CH3 H H H H CH3 1,3-butanediol (6) H H H H H CH3 1,3-propanediol (7) H H H H H H All the hydrogenolysis experiments in this work are performed at 210 °C and 5 MPa hydrogen pressure in the aqueous phase. NaOH is added to promote the reaction. The transition catalysts used in this work are Raney Cu and Raney Ni, two of the typical catalysts used in sugar and sugar alcohol hydrogenolysis (Montassier er al., 1988; Chang et al., 1985; Sohounloue et al., 1983). The experimental results are summarized in Table 4-2, together with a documentation of the expected products from each 1,3-diol based on Figure 4-1. The yields presented in Table 4-2 are based on the reacted 1,3-diols. The mass is not balanced perfectly, but is within the experimental error. In all the experiments where methanol is an expected product, it is found only in a trace amount. The reason is attributed to the transition-metal-catalyzed Cannizzaro reaction, which intercepts formaldehyde, the hydrogenation precusor to methanol, as discussed earlier. According to Figure 4—1, the hydrogenolysis of a 1,3-diol is initiated by dehydrogenating one of its hydroxyl groups. This postulation is strongly supported by the a. Hawk-OCUMe-bhfih .Cmcrflin he emu-Edi: o I.“ I‘II-lfl-k I l 2: 2.2. we: :3 a .5 2 a a. a 2 c2 2: | 2.3 2 l I a: 3.: 3: E: a $3 2 z a a 8 9: Ed I 3% N. | I I | 2.8 88 25.. ”em. a 5a 2.2. 25: a e a mm 2" 8.3 I 3.2.. : I . l I 58 . . I . a o 88 2.8 88 En. as. 2.8 25.. a 8 a a 9: a : S :n 2 I I | 2m 25.. 2.8 S: 2.8 E: 28 a 2.2. = : c2 36 I 8. S a I I I I one: one: one: 25: one: one: one: one: one: one: an v. _ ~ I sad: w I l c8 25.. as. «S a an S a a a a 3m | $6 8. S. A 2: i: one 98 mm” a: E: S fiv— 2 z a z .a 9 ca | m; 8.5. c I | S .o as: SE 3: a. 23 2 = a. a 3 5 I ”.2 3S m I I I I 0:0: 0:9— 25: an~ Q SE 060: 0:0: m— Q N ma Ow~ I 8.” W64; V I . E8 . . . so I I 2.2. 25.. 2.2. an. .5 25.. 2.2. a 8 a a 8“ I a a 8 S. m I l l 2: 85 2.8 E. 25.. E: 2.8 a 2.2. = : o3 I 3a 8.5 N I I I I 28: one: 28: 25: 28: one: one: one: one: 28: com I 85 c— . _ \n _ no: em mm «m .m meow Ed 930 {a cube oJ 9:6 .96 no.6 33 2:5 3 _z E 5 8.5 .oz . .250 08F >29 :3— . @3823. ES 209283.. 8558. e 388... 8886 seesaw 32m a fireman—i: .056.— ue $1.8m .né «Ema. F-‘I‘:281 50m 6 .48 o». .83 8:38 guano. on. go 982? 3.2 05. .aE :80 E 5:88 -8 a an 335 «a? $054 2 ~ 48 vd 49:88 >33. 25 2 .8333 E .583 2 8:3er Sweet»: 52 n 28 Do SN 3 con—coron— 803 8288.. =< .3 8.238 .3 2.3. results I10” and yea inactive ur to he that is a crush pmductsl used In th undergo hldluget Can not I IO be In Ilj‘tlrttge Carbon, 3and 9, ll a‘éitirtst rEactIOn dicarhor inhydro ”It of ; 510mb} Inc 36 results from hydrogenolysis of 1. Because of the absence of a hydrogen atom on both the or and y carbons, 1 is incapable of undergoing dehydrogenation, and thus is expected to be inactive under the hydrogenolysis conditions. As indicated in Table 4-2, 1 is indeed found to be inactive. The hydrogenolysis of 1 has been carried out for six hours with Raney Cu as a catalyst (Run 1) and for four hours with Raney Ni as a catalyst (Run 8). No reaction products have been identified at the end of either reaction. For comparison, other 1,3-diols used in this work are all found to be reactive under the same conditions, because they can undergo the dehydrogenation reaction. That dehydrogenation is a necessary step for hydrogenolysis is additionally supported by the results from hydrogenolysis of 2. As 2 can not be dehydrogenated at the or carbon, the C-C and C-0 bonds of 2 are not expected to be broken between the B and y carbons and at the 7 carbon, respectively, during hydrogenolysis, although they may be broken between the or and B carbons and at the or carbon, respectively. The experimental results turned out again exactly as expected (Runs 2 and 9). The results from hydrogenolysis of 2 (Runs 2 and 9) also provide good evidence against the retro—Claisen as the dominating C-C cleavage mechanism. The retro-Claisen reaction requires a dicarbonyl precursor. Since 2 is incapable of forming such a dicarbonyl, the retro-Claisen mechanism predicts that the CC bonds of 2 are unbreakable in hydrogenolysis. However, the experimental results show otherwise: the hydrogenolysis rate of 2 is not impaired by its incapability of undergoing the retro-Claisen reaction, as shown by the conversion data in Table 4-2. In contrast to the retrooCIaisen mechanism, the retro-aldol mechanism predicts that the CC bond of 2 may be broken between the or and B carbons to form two isopropanol molecule predicts i2. Isnl dnnutit Idenut‘u .Icetnn: Rises ‘ lnthel hydtug cnnsu again i CIiI they Our t 1in 37 molecules, as 2 can be dehydrogenated at the 7 carbon. The retro-aldol mechanism also predicts that the isopropanol is formed through an acetone intermediate. As seen in Table 4-2, isopropanol is not only found at the end of the reaction (Runs 2 and 9), but is a dominating product of 2-methyl-2,4-pentanediol hydrogenolysis. Furthermore, acetone is identified in the reaction when Raney Cu is used as the hydrogenolysis catalyst (Run 2). Acetone is not detected in the reaction when Raney Ni is used as the catalyst, because Raney Ni is a more efficient hydrogenation catalyst than Raney Cu. The acetone formed in the hydrogenolysis of 2 may be quickly hydrogenated to isopropanol. The results from hydrogenolysis of 2 are very supportive of the retro-aldo mechanism. The results from hydrogenolysis of 3, 4, 5, 6 and 7 are also supportive of the retro- aldol mechanism. In all these experiments, the C-C bond cleavage patterns are found to be consistent with the retro-aldol mechanism. In hydrogenolysis of 6 (Run 6), acetone is again identified as an intermediate as predicted by the retro-aldol mechanism. Hydrogenolysis of 2, 6 and 7 has been canied out by Conner and Adkins (1932) with a catalyst prepared by precipitation of Ni carbonates. However, in hydrogenolysis of 2, they identified only the end products, isopropanol (11) and 4-methyl-2-pentanol (19), as in our experiment with Raney Ni as the catalyst In hydrogenolysis of 6 and 7, they did not report the C-C cleavage at all. The successful identification of CC bond cleavage in hydrogenolysis of 6 and 7 and of the retro—aldo intermediates in hydrogenolysis of 2 and 6 in this work strengthens the theory that the C-C cleavage is through the retro-aldol reaction. Also according to Figure 4-1, the dehydration reaction is responsible for the C-0 cleavage in hydrogenolysis. This postulation is strongly supported by the results from hidt illit‘ suc't ncc' 0c the he Minds Its 38 hydrogenolysis of 3 and 4. Based on the dehydration mechanism, a hydrogen atom attached to the B carbon is necessary for the hydrogenolysis of C-0 bonds to occur. Since such a hydrogen does not exist in 3 and 4, cleavage of the C-0 bonds is not expected to occur in the hydrogenolysis of these compounds. Indeed, as indicated in Table 4-2, no C- O cleavage is found in hydrogenolysis of either 3 or 4. The dehydration mechanism is also supported by the C-0 cleavage pattern found in the hydrogenolysis of 2 (Runs 2 and 9). The dehydration mechanism expects the C-0 bond to be broken only at the or carbon in hydrogenolysis of 2, as 2 can be dehydrogenated only at the 7 carbon. The results from hydrogenolysis of 2 shows that this expectation is well founded. The additional products 2-pentanone (l8) and 2-butanone (16) identified in the hydrogenolysis of 5 (Runs 5 and 12) and 6 (Run 6), respectively, provide additional evidence to the dehydration mechanism. These compounds are not the direct products of dehydration, or VII in Figure 4-1, but intermediates from VII to VIII. The hydrogenation of VI] to VIII is believed to take place in two sequential steps, as shown in the following with hydrogenolysis of S as an example: 0 O H II + H2 II II CH3CH=CHCCH3 LL CH3CH2CH2CCH3 —. CH3CH2CH2CHCH3 VII = 3-pentene-2-one 2-pentanone VIII = 2-pentanol The accumulation of 2-pentanone in hydrogenolysis of 5 and of 2-butanone in hydrogenolysis of 6 suggests that, with Raney Cu and Ni as catalysts, the hydrogenation of C=C double bonds is faster than the hydrogenation of C=O double bonds. The mechanism pictured in Figure 4-1 is supported additionally by the selectivity 39 data collected from our experiments. In hydrogenolysis of a 1,3-diol, typically four bond cleavage selectivities can be defined, namely; Ca'Cp vs. Ca—O, Cfl-CY vs. CY-O, Ca—Cp vs. CB-Cy. and Ca-O vs. CTO. These selectivities are calculated for each experiment and are documented in Table 4-2. According to Figure 4-1, the retro-aldol reaction responsible for the CC cleavage and the dehydration reaction responsible for the C-0 cleavage share the same precursor. Therefore, the C-C v.s. C-O selectivities should be affected by relative rate of the retro- aldol reaction to the dehydration reaction. This expectation is justified by the CC v.s. C- O selectivity data in Table 4-2. In hydrogenolysis of straight-chain 1,3-diols (5, 6 and 7), the reaction is found to favor the C-0 cleavage products, because the dehydration is faster than the retro-aldo reaction in this case. In hydrogenolysis of a branched-chain 1,3-diol (2), the reaction is found to greatly favor the C-C cleavage products, because the retro- aldol reaction is faster than the dehydration in this case (Neilson and Houlihan, 1968). When the retro-aldol reaction occurs, the aldol reaction is also possible, which often leads to scrambling of the retro-aldol fragments. Thus, in the hydrogenolysis of 1,3-diols, one might expect 1,3-diols other than the starting ones to be formed as the result of the scrambling aldol reaction. However, no such by-products have been observed in any experiments carried out in this work. The reason is attributed to the fast hydrogenation of aldehydes intermediates under the hydrogenolysis conditions and the unfavorable condensation between ketone intermediates, which minimize the formation of any new 1,3-diols. As discussed above, the hydrogenolysis mechanism described in Figure 4-1 is strongly supported by the results from hydrogenolysis of 1,3-diol model compounds. To Slim and the rut res; .J sun arc “It‘tul 4o summarize the experimental results, the hydrogenolysis of 2,4-dimethyl-2,4-pentanediol and 2-methyl-2,4-pentanediol demonstrates that the dehydrogenation is a necessary step in the hydrogenolysis of sugar alcohols. The hydrogenolysis of 1,3-dimethyl-1,3— propanediol and 2,2,4-trimethyl-1,3-pentanediol demonstrates that the dehydration is responsible for the C-0 cleavage in hydrogenolysis. The hydrogenolysis of 2-methy1-2,4- pentanediol also provides good evidence against the retro-Claisen as the dominating C-C cleavage mechanism. The likelihood of the retro-Michael reaction being a dominating C- C cleavage mechanism is all but eliminated by the capability of all the 1,3-diols (except 2,4-dimethyl-2,4-pentanediol) to undergo hydrogenolysis. On the other hand, these experimental results are very supportive to the retro-aldol mechanism. In all the experiments but hydrogenolysis of 2,4-dimethy1-2,4-pentanediol, where no reaction is supposed to occur, the C-C cleavage is found to follow the patterns predicted by the retro- aldol mechanism. In addition, some ketone intermediates expected by this mechanism are identified in the hydrogenolysis of 2-methyl-2,4-pentanediol, 2,4-pentanediol and 1,3- butanediol. In no reaction, has a product other than the expected been found, and in no reaction, has an expected product not been found. The experiments have been performed with two different catalysts (raney Cu and Raney Ni), the reaction patterns are not affected by the change of catalysts. These results strongly support the mechanism described in Figure 4-1. Now that the mechanism of C-C and C-0 cleavage is known, it becomes possible for one to adopt a rational approach in control of the selectivity of sugar hydrogenolysis. 5.1 3. try: ('0 cl It can IIIliter I“ Intel CHAPTER 5 SELECTIVITY OF SUGAR HYDROGEN OLYSIS 5.1 Selectivity Study Using 2,4-PD Model Compound: Rationale Because of the multiple bond-cleavage reactions involved in sugar hydrogenolysis, three types of selectivities can be defined for this process, namely the CC vs. C—C, C-O vs. 00 and CC vs. 00 selectivities. In the current work, the CC v.s. C-O selectivity of sugar hydrogenolysis is studied using 2,4-pentanediol (2,4-PD) model compounds. The purpose is to find a way to minimize C-O cleavage in sugar hydrogenolysis. As the C-0 cleavage is to a great extent responsible for the complication of sugar hydrogenolysis products, minimizing C-O cleavage is expected to make the reaction more selective. Reduction of C-0 cleavage in sugar hydrogenolysis is also expected to increase the yield of glycerol. As mentioned in Chapter 3, glycerol is the highest-valued major product of sugar hydrogenolysis, and it preserves all the oxygen atoms in the starting sugar molecules. I Using 2,4-PD model compound in the selectivity study has a major advantage, that is, hydrogenolysis of 2,4—PD gives only one pair of CC cleavage products and one pair of C-0 cleavage products, and these products do not undergo further bond cleavage reactions to complicate the process. Figure 5-1 shows the reaction pathway of 2,4-PD hydrogenolysis, which is derived based on the hydrogenolysis mechanism established in our mechanism study. From this figure, it is seen that aldehyde, ethanol, acetone and 41 42 OH OH I CH3CHCH25HCH3 2.4-pentanediol In. I II I)” n o II ' 2 CH3CCH3 + CH3CHO ‘— CH3CCH2CHCH3 —-> CH3CH=CHCCH3 3061006 acetaldehyde 4-hydroxy-2-pentanone 3-penten—2-pentanone ¢ + H2 \I' H2 ¢ + H2 3H 0 CH CH CH II CH3 HCH3 3 2 isopropanol ethanol CH3C§§§E§§EH3 (In (H; CH3CH2CH2CHCH3 2-pentanol Figure 5-1. Reaction Pathway of 2,4-Pentanediol Hydrogenolysis isopropanol result only from the C-C cleavage, and that 3-penten-2-one, 2-pentanone and 2-pentanol result only from the C-0 cleavage. Therefore, the C-C cleavage in 2,4-PD hydrogenolysis can be quantified by the sum of acetone and isopanol, and the C-0 cleavage by the sum of 3-penten-2-one, 2-pentanone and 2-pentanol, which makes the selectivity study easy to carry out In contrast, in hydrogenolysis of real sugars, it is difficult to tie any reaction products to a specific bond-cleavage reaction, because of the general existence of more than one pathway to the same product. In spite of the convenience of using 2,4-pentanediol model in selectivity study, a difference between 2,4—pentanediol and sugars needs to be noted. Compared to 2,4-PD, 43 sugar and sugar alcohol are characterized by a third hydroxy group at the B position, which will slightly alter the hydrogenation pathway of the ocB-unstaurated ketone from the C-0 cleavage, or dehydration, reaction. In the presence of a hydroxy group at the B position, the or,B-unstaurated ketone can tautomerize to give a 1,2-diketone, as shown in Figure 5-2. Therefore, hydrogenation of the or,B-unstaurated ketone may occur via two pathways in the hydrogenolysis of real sugars. O o dehydration ll tautomerization II ——" RC? = CHR = RCCCHZR’ OH I hydrogenation I hydrogenation Figure 5-2. Alternate Hydrogenation Pathway of or,B-Unstaurated Ketone in Sugar Hydrogenolysis At 25 °C, the equilibrium ratio of the or,B.unstaurated ketone to the 1,2-diketone is about 1:99 (Alexander, 1954). At 220 °C, a typical temperature used in hydrogenolysis, this ratio still remains at about 1:20. However, because the hydrogenation of an or,B- unstaurated ketone is generally much faster than the hydrogenation of an 1,2-diketone, direct hydrogenation of the or,B-unstaurated ketone intermediate is still expected to be a major pathway, even in sugar hydrogenolysis. The 00 cleavage in sugar hydrogenolysis is expected to be less favorable than in 2,4-PD hydrogenolysis. The tautomerization reaction following the dehydration converts a better hydrogenation precursor to a worse one, and thus may reduce the hydrogenation rate of the dehydration product in sugar hydrogenolysis. Based on this reason, a catalyst Illljdj 44 which gives a good C-C v.s. C-O selectivity in 2,4-PD hydrogenolysis is expected to also deliver a good selectivity in sugar hydrogenolysis. 5.2 Experimental Study In the current work, 2,4-PD has been hydrogenolyzed under various reaction conditions to investigate the effects of temperature, hydrogen pressure, base concentration, catalyst amount and catalyst type on the C-C v.s C-O selectivity. Figure 5-3 presents the result from a typical experiment run, carried out at 220 °C, 5 MPa H2 pressure and with Ni on kieselguhr as the catalyst. In addtion to the end products, ethanol, isopropanol and 2- pentanol, two intermediate products, acetone and 2-pentanone, are normally observed at a significant level in 2,4-PD hydrogenolysis. Other intermediate products, including 4- hydroxy-Z-pentanone, acetaldehyde and 3-penten-2-one, are sometimes detected during the initial period of reaction, but generally exist only at low levels. These products are not quantified and therefore are not shown in Figure 3. Figure 5-4 shows how the CC v.s C-O selectivity changes during the hydrogenolysis. This selectivity is defined as the ratio of the mole of C-C bonds to the mole of C-0 bonds broken in the reaction. Since the amount of 3-penten-2-one is negligible in the reaction product, the selectivity is calculated as: _ [acetone] + [isopropanol] _ [2—pentanone] + [2-pentanol] S As seen from Figure 5-4, the selectivity is typically stabilized at a constant level after the initial period of the reaction. It is this stabilized selectivity that is compared in the following for different reaction conditions. 45 120 . - groo— - é a 80: _ E 8 cos . O C O 0 :1 j x U - 9 (1 20» « , /" i g . : 0 so 10 150 200 250 Time (min) Figure 5-3 Product v.s Reaction Time Profiles. Experimental Conditions: 220 °C, 5.0 MPa H2 pressure, 0.1 g Ni/kieselguhr catalyst, 0.4 ml 1 N NaOH and 40 ml water solvent. Symbols: 0 2,4-PD; I EtOH; O acetone; Ai- PrOH; D 2-pentanone; A 2-pentanol. 3 . . fl 2.5" 4 2h 4 2‘ E . “as K CD : 1t... . , , . .. 0.5" . 0 1 L 1 4 0 50 100 150 200 250 Time (min) Figure 5-4 Selectivity v.s. Reaction Time Profile Derived from Figure 5-3 Table 5-1 Effect of Temperature on Selectivity 46 Experiment Series # 1 “ Experiment Series # 2 b Temperature (°C) Selectivity Temperature (C) Selectivity 180 1.38 — — 200 1.42 200 1.06 220 1.42 220 1.02 240 1.45 240 1.10 0. Reaction Conditions: 5 MPa H2 pressure. 0.1 g Ni/kieselguhr. 0.4 ml 1 N NaOH. 0.5 starting 2,4- PD and 40 ml water. b. Reaction Conditions: 5.0 MPa H2 pressure, 0.25 g Raney Cu, 0.2 m1 1 N NaOH. 0.5 g starting 2,4-PD and 40 ml water. Table 5-2 Effect of Base Concentration on Selectivity ‘ Experiment Series # l b Experiment Series # 2 C NaOH Added (ul) Selectivity NaOHl Added (til) Selectivity 0 0.46 0 0.26 400 1.42 200 1.02 1000 1.56 500 1.73 a. In all cases, base is added as 1 N aqueous N aOH solution. b. Reaction Conditions: 200 °C, 5.0 MPa H2 pressure. 0.1 g Ni/kieselguhr, 0.5 starting 2,4-PD and 40 ml water. c. Reaction Conditions: 220 °C, 5.0 MPa H2 pressure, 0.25 g Raney Cu, 0.5 g starting 2,4-PD and 40 ml water. Table 5-3 Effect of Hydrogen Pressure on Selectivity Experiment Series # 1 0 Experiment Series # 2 b Pressure (MPa) Selectivity Pressure (MPa) Selectivity 3.0 1.50 3.0 0.80 5.0 1.42 5.0 1.02 7.0 1.49 7 .0 l. 15 0. Reaction Conditions: 200 °C, 0.1 g Ni/kieselguhr, 0.4 ml 1 N NaOH. 0.5 starting 2,4-PD and 40 ml water. b. Reaction Conditions: 220 °C, 0.25 g Raney Cu. 0.2 ml 1 N NaOH. 0.5 g starting 2,4- PD and 40 ml water. Table 5-4 Effect of Catalyst Amount on Selectivity Experiment Series # l 0 Experiment Series # 2 b Catalyst (g) Selectivity Catalyst (g) Selectivity 0.1 1.42 0.1 1.22 0.2 1.49 0.25 1.02 0.5 0.90 0.5 0.75 0. Reaction Conditions: 220 °C, 5.0 MPa H2 pressure, Ni/kieselguhr, 0.4 ml 1 N NaOH. 0.5 start- ing 2,4-PD and 40 ml water. 17. Reaction Conditions: 220 °C, 5.0 MPa H2 pressure. Raney Cu, 0.2 m1 1 N NaOH. 0.5 g starting 2,4-PD and 40 ml water. the rim hydro tnndir (Ibsen) CGIISlSle £ij Writs. 8let 0,, I) 48 Efleet of Temperature To investigate the effect of temperature, 2,4-PD is hydrogenolyzed under two sets of base conditions with varying temperature. The selectivity, as Table 5-1 shows, is found to barer change with the temperature, although increasing temperature greatly enhances the reaction rate. Eflect of Base Concentration The base concentration, as Table 5-2 shows, has a profound effect on the selectivity, as well as on the reaction rate. Increase in the amount of base added to the reaction solution increases the reaction rate, and favors the cleavage of CC bonds. This effect on the selectivity is most easily seen when the data from reactions with base addition are compared with those from reactions without base addition. One thing that needs to be pointed out is that the base concentration, as indicated by the PH of the reaction solution, is generally found to be lower than the initial concentration after the initial period of reaction. Eflect of Hydrogen Pressure The effect of hydrogen pressure seems to vary with the catalyst used. With Raney Cu as the catalyst, an increase in the hydrogen pressure increases the selectivity; while with Ni on kieselghur as the catalyst, the increase in the hydrogen pressure causes little change in the selectivity within the experimental conditions explored in this work In the case of Raney Cu, hydrogen pressure is also observed to have an inhibiting effect on the rate of 2,4-PD hydrogenolysis. This effect is consistent with the observation made by Chang et al. (1985) in hydrogenolysis of sobitol. Effect of Catalyst Amount The effect of catalyst amount again varies with the catalysts. However, the general trend is that high catalyst concentration has an adverse effect on the selectivity. With Raney Cu as the catalyst, the selectivity steadily decreases 49 with the catalyst amount. With Ni on kieselghur as the catalyst, the effect of catalyst amount only becomes significant at high level. Efiect of Catalyst Type A good catalyst is considered to be the key to achieve the control over the C-0 cleavage in sugar hydrogenolysis. In this current work, seven different types of catalysts have been investigated. The results are presented in Table 5-5. As can be seen from this table, the selectivity varies widely with the type of catalyst used. The best selectivity is furnished by barium-promoted copper chromite, and is as high as 2.5, which is yet to be optimized against other reaction conditions. In contrast to the barium-promoted copper chromite, the plain copper chromite catalyst only provides a selectivity of 1.21 in our experiments. It is interesting to see the difference made by the presence of barium oxide. The effect of the amount of barium oxide used in preparation of the catalyst may be another factor deserving study in the context of selectivity improvement. Table 5-5 Effect of Catalyst Type on Selectivity " Catalyst Type Selectivity nickel on silica-alumina 1.12 nickel kieselguhr 1.18 copper chromite 1.21 ruthenium on carbon 1.36 Raney nickel 1.46 Raney copper 1.73 barium-promoted copper chromite 2.50 a. All reaction are run at 210 °C, 3.5 MPa H2 pressure, with 0.05 g metal catalyst. 0.2 ml 1 N NaOH. 0.5 g starting 2,4-PD and 40 ml water. tea frnrr 5—5 at Iheret IIIISQIW‘ II'Ittde ICS‘tlIl of l 50 5.3 Simulation Study In order to rationalize the experimental results just presented, a simulation study has been carried out on the hydrogenolysis of 2,4wPD. In this study, the effect of base concentration, hydrogen pressure, catalyst amount and catalyst type are simulated and compared against the experimental results. The effect of temperature is not studied in this initial simulation effort, because first it is more difficult to do and second this effect is not as profound as other factors. Therefore, all the simulated reactions are assumed to be isothermal and take place at an arbitrarily chosen temperature of 220 °C. 5.3.1 Model Development At this time, the reaction pathway of 2,4-PD hydrogenolysis is clear to us. However, the pathway alone is not sufficient to allow us to develop the process model. It does not provide enough reaction details on how the H2 and the metal and base catalysts are involved in the reaction, which are necessary for process model development. For this reason, a more detailed reaction scheme of 2,4-pentanediol hydrogenolysis is developed from the reaction pathway in Figure 5-1. This detailed reaction scheme is shown in Figure 5-5 and is used as the basis to develop the process model. In the absence of transitional metal catalysts, molecular hydrogen is not reactive. Therefore, it must be activated by the metal catalyst before it can be used to reduce any unsaturated compounds. The result of the reaction of H2 with the metal catalyst is a metal hydride species, which has the ability to hydrogenate C=C and C=O double bonds. As a result of this hydrogenation reaction, the metal catalyst must be regenerated, so that it can be re-used in the next cycle of catalysis. The above mechanism has been used by many IESI Its'd; PDh IIIE me “thin- I IUllh: Winn womb: 51 2.4-pentanediol (A) M.0H' *1 MH H 0 ‘ r 2 I k- I OH‘ OH’ acetone + acetaldehyde 11,—— 4-hydroxy-2-pentanone :p: 3-penten-2-one (C) (E) k? (B) k (G) MH. H20 MH. H20 MB. H20 k4 k6 : b3 ”'0”. ‘ k1 M. OH' isopropanol ethanol 2-pentanone (D) (F) (1) MH, H20 ks H2+M+OH' —-:f> MH+H20 :k-7 It'ir 2-pentanol (D Figure 5-5. Detailed Reaction Scheme of 2,4-PD Hydrogenolysis researchers to explain the hydrogenation process of C=C and C=O bonds with molecular hydrogen, and is adopted in the current work to help develop the reaction scheme of 2,4- PD hydrogenolysis. The hydrogenation of carbonyl compounds by the metal hydride yields alcohol and the metal catalyst. Since this reaction is reversible, alcohols must be also able to react with the metal catalyst to get dehydrogenated and to release a metal hydride. Therefore, it is further pictured in Figure 5-5 that the hydrogenolysis of 2,4-PD is initiated by the reaction of this starting compound with the metal catalyst, and that the metal hydride works as a hydrogen carrier to transport hydrogen from 2,4-PD and molecular H2 to those nil hurl equil fitted “Inc on IIIII IIIIIS littered III he I 52 unsaturated intermediate products. For the convenience of later mathematical model development, the reactions shown in Figure 5-5 are all numbered (the reaction number corresponds to the subscript of the rate constants shown in Figure 5-5), and the products are also designated by single letters. From Figure 5-5, it may be noticed that the base catalyst (hydroxide ion in this case) is not only involved in the retro-aldol and dehydration reactions as a real catalyst, but also participates directly in the dehydrogenation of 2,4-PD and activation of H2. Participation of bases in H2 activation and alcohol dehydrogenation has been well documented in literature. As the dehydrogenation step is most likely to be the rate limiting step in 2,4-PD hydrogenolysis, the above postulation is in consistence with the fact that 2,4-PD hydrogenolysis rate is enhanced by the addtion of bases. It may also be noticed from Figure 5-5 that all the reaction step in 2,4-PD hydrogenolysis, except the hydrogenation of 3-penten-2-one, are assumed to be reversible. Retro-aldolization and dehydration are well—known reversible reactions. Hydrogenation of ketones and aldehydes and dehydrogenation of alcohols are also known to be reversible, but hydrogenation of alkene is generally considered to be irreversible, because the reaction equilibrium lies far to the product end. To derive the mathematical model from Figure 5-5, we assume that the reaction is carried out in a batch reactor, and thus the product concentrations are functions of reaction time only. We also assume that the reaction is carried out under isothermal conditions, and thus the reaction rate constants do not change with time. The concentration of H2 in the reaction solution may also be assumed to be time-independent and be in equilibrium with the gaseous phase, as the reaction is assumed to take place under a constant H2 In t. COIICE flirts Ctlllten andl0t 53 pressure and the solution is under constant violent stirring during the reaction. A mass balance on each of the species present in Figure 5-5 except H2 leads to: d d—I[A] = ‘XI (1) d $13] = Xr'xz‘xs (2) d d EID] = X3 (4) d EMS] = Xz-X4 (5) d Eur] = X4 (6) d are] = X5_X6 ‘7’ d EII] = X6‘X7 (8) d Em = X7 ‘9) d d d 25”“ =EIOH] =_-d—t[MI-l] =—X1+X3+X4+X6+X7—X8 (10) In the above equations, [A], [B], [C], [D], [E], [F], [G], [I] and [J] are the molar concentration of 2,4—PD, 4-hydroxy-2-pentanone, acetone, isopropanol, acetaldehyde, ethanol, 3-penten-2-one, 2-pentanone and 2-pentanol, respectively. [M] is the molar concentration of the metal catalyst, [MH] is the molar concentration of the metal hydride, and [OH] is the molar concentration of hydroxide ion in the reaction solution. X 1 to X8 54 are the reaction fluxes of Reactions No.1 to No.8 in Figure 5-5, respectively. Assuming all these reactions are first order in the substrates and catalysts, we further have X, = 1:1 [A] [M] [0H] —k_1 [B] [MH] (11) X2 = 1:213] [01!] —k_2[C] [El [0H] (12) X3 = k3IC] [MH] —k_3[Dl [M] [0H] (13) X. = ME] [MH] -k4[F] [M] [OH] (14) X5 = kslB] [0H] —k_5[G] [0H] (15) X(5 = 1:616] [MH] (16) X7 = k7rlt [MH] -—k., in [M] [OH] (17) X8 = kslfll [Ml [0H] —k_8IMHl (18) In the last equation, [H] is the molar concentration of H2 dissolved in the reaction solution. At the beginning of the reaction, we may assume that all the products, but the substrate (2,4-PD) and the metal and base catalysts, have a zero concentration. Therefore, [A] = [Ar,.rMi = [moron] =t0H1, and [B]=[Cl=[D]=[E]=IFI=[G]=III=[JI=[MH]=0 for t = 0. The above equations and initial conditions constitute a closed problem. If all the relevant reaction rate constants are obtained, the above equations can be solved to give product concentrations as functions of reaction time. 55 5.3.2 Determination of Kinetic Parameters As mentioned previously, the simulation study is to be carried out at a constant temperature to simulate the effects of reaction conditions other than temperature on the selectivity. Therefore, we need only to determine the reaction rate constants for one temperature at which we choose to run the simulation. This temperature is chosen here to be 220 °C. This way, we could use the experimental data presented in Figure 5-3 to help determine these rate constants. The data in Figure 5-3 are obtained at 220 °C. Among the eight reactions shown in Figure 5-5, reactions N o. 2 (retro-aldo) and No. 5 (dehydration) are, though at low temperature, well-studied. From the kinetic data published by Jensen and Hassan (1976), it is derived that k2=752 M‘lmin‘l, k_2=4480 M‘ 2min'1, k5=14100 M'lmin'1 and k,5=6590 M'lmin'l at 220 °c (see Appendix D). The rate constants (eleven in total) for the rest of the reactions shown in Figure 5-5 vary with the metal catalyst involved, and are not available from literature. Therefore, they are determined below, indirectly, from our 2,4-PD hydrogenolysis data The eleven rate constants to be determined are not all independent They are bound together by thermodynamic relations. Of these relations, an obvious one is that, for any reversible reaction, the ratio of the forward reaction rate constant to the backward reaction rate constant is equal to the reaction equilibrium constant. Thus, kr‘ I; = K i (19) for i=1, 3, 4, 7 and 8. In the above equation, K,- is the equilibrium constant of reaction No. i. The reaction is assumed to take place under dilute conditions. The relations which are not so obvious are that the above equilibrium constants are 56 also interdependent. These relations become apparent when the following two chemical equations are examined: AG”, K RICHOI-IRZ + M + OH“ 2;": RICORZ + MH + H20 AG°,r_r RICHOHRZ - H20) 2‘2 Rlcorz2 The first equation in the above is a general representation of the dehydrogenation/ hydrogenation reactions shown in Figure 5-5. This first equation is also the result of the second equation above added to Reaction No.8 in Figure 5-5. Therefore, we have A0” = A9" + A080 (20) where A60 and A9” are the free energy changes of the first and second reactions above, respectively. From Eq.(20), we further have K = I_(K8 (21) where K and K are the equilibrium constants for the first and second reactions, respectively. Unlike AG and K, ag" and K are well defined and are not dependent on the catalyst involved. Values of KS relevant to 2,4-PD hydrogenolysis have been estimated in this study from the free energy data documented in Guthrie (1977) for ketones and aldehydes and in Young (1981) for hydrogen dissolved in aqueous phase, and are listed in Table 5-6. Due to lack of free energy data at high temperature, these equilibrium constants have been calculated using free energies at 25 °C. Also because of lack of data, A9” '3 for Reactions No.1 and No.7 are assumed to be the same as Reaction No.3. The rationale is that these reactions are all about dehydrogenation of a secondary alcohol to a ketone. 57 Table 5-6 Equilibrium Constants of Relevant Dehydrogenation Reactions No. Alcohol Ketone/Aldehyde ag" (kcal) K 1 2,491) 4-hydroxy-2-pentanol 12.54 2.81x10‘6 3 isopropanol acetone 12.54 2.81x10‘6 4 ethanol acetaldehyde 16.59 4.52x10"8 7 2-pentanol 2-pentanone 12.54 2.81x10'6 Substituting Eq.(21) into Eqs.(19) for each i and rearranging gives k1 = Eth’ht (22) k—3 = §3K8k3 (23) ka=fiflh on k—7 = §7K8k7 (25) 18 = K814, (26) Elimination of K8 from the above equations leaves four relations relating the rate constants. Therefore, of the eleven rate constants to be determined, only seven are independent. The remaining four have to be determined from the seven independent rate constants chosen, in order to assure thermodynamical consistency. Besides the thermodynamic consistency, it is also important for the determined rate constants to fall within a reasonable range, so that the simulated processes are as close as possible to those actually occurring in our experiments. To achieve this, the best approach would be to accommodate as much experimental information as possible into the rate 58 constant determination process. Therefore, data from the experiments presented previously in Figure 5-3 are used in the following to help determine the rate constants. That experiment was carried out at 220 “C. Figure 5-3 shows how the concentration of 2,4-PD (A), acetone (C), isopropanol (D), ethanol (F), 2-pentanone (I) and 2-pentanol (J) changes with reaction time in the above experiment Concentration of 4-hydroxy-2-pentanone (B), acetaldehyde (E) and 3- penten-2-one (G) are low (estimated to be less than 1 mM) and are not determined in the experiment. The concentration of metal catalyst (M), metal hydride (MH) and hydroxide ion (OH) are also undetermined, but at beginning of the reaction they are calculated as 28, 0 and 10 mM, respectively, from the amounts of regents used in the experiment. The initial concentration of 2,4-PD (A) is 120 mM. The H2 concentration in the reaction solution is found to be about 0.1 M in equilibrium with 5 MPa hydrogen pressure at 220 °C (Young, 1981). From Figure 5-3, it is seen that, after 1:100 min, [C] and [I] are quite steady. Considering also that [B], [E], [G] and [OH] are low, we may assume that the reaction is at a pseudo-steady state after the initial period. Thus from Eqs.(2), (3), (5), (7), (8) and (10), it is derived that x2 = x3 = x4 (27) X5 = X6 = X7 (28) and X1 = X3 = X2+X5 (29) From Figure 5-3, x, and x7 can be determined to be 8.85 x 10-5 M min-1 and 6.07 x 10-5 M 59 min'l, respectively, at t=l75 min. All the other reaction fluxes at 1: 175 min are calculated with Eqs.(27) to (29) from X 3 and X7, and are listed in Table 5-7, together with the product concentrations at t=l75 min and t=0 min. Concentrations of ethanol (F) and 2,4»PD (A) in Table 5-7 have been re-calculated to make the mass perfectly balanced. These data are used below in conjunction with Eqs.(l 1) through (18) to determine the reaction rate constants. From Eq.(l l), we have x1 = k, [A] [Mi [0H1 —k_1[B] [MH] k 0 = a. tMHiIgj-IW [A] - [Bi] 0 = k_1[MHl(If1Kg%fl]-IA1 — 131) Let 0H [(0 = Kalb—{13%;}; (30) and substitute it into the above equation. We get X1 = k_1[MH] (15,19, [A] — [31) (31) Similarly, from Eqs.(13), (14), (17) and (18) we may get x, = k31MH1 ([Cl—I_<3KOID]) (32) X4 = k4 [MH] ([5] ‘E4K0 I”) (33) X7 = k7IMHI ([1] -I_{7KOIJ]) (34) and X8 = k_8[MH] (K0 [H] — 1) (35) Table 5-7 Summary of Product Concentrations and Reaction Fluxes t = 0 min t = 175 min Unit [A] 120 28.0 mM [B] 0 less than 1 mM mM [C] 0 5.89 mM [13} 0 48.0 mM [E] 0 less than 1 mM mM [F] 0 53.9 mM [G] 0 less than 1 mM mM [1] 0 6.26 mM [J] 0 31.7 mM [H] 100 100 mM [OH] 10 unknown mM [M] 28 unknown mM [MH] 0 unknown mM X1 _ 1.49 x 10“ M min“l X2 _ 8.85 x 10" M min'1 X3 _ 8.85 x 10" M min'l X4 _ 8.85 x 10'5 M Ma1 X5 _ 6.07 x 10'5 M min'1 X6 _ 6.07 x 10" M min‘l X7 _ 6.07 x 10'5 M min'1 X3 _ 1.49 x 10" M Ma1 61 Combining Eqs.(3l) through (35) with Eqs.(12), (15) and (16), there are now eight equations available in total for eleven unknowns. These unknowns include six rate constants (k_ 1’ lg, lg, k6, 1:7 and k8), four concentrations ( [B], [E], [G] and [OH] ) and a new variable K0. [M] and [MH] are related to [01-1] through the following equation: [MH] = [M],- [M] = rom,-10H1 (36) which is obtained by integration of Eq.(10). As the available equations are not enough for unknowns, three of them have to be assumed before others can be solved. In the following, this liberty is taken to make sure that the determined [B], [E], [G] and [OH] conform with our experimental observations and that the relative magnitudes of the determined rate constants are consistent with our knowledge of hydrogenation reactions. If the dehydration (Reaction No.5) were irreversible during hydrogenolysis, the maximum selectivity of 2,4-PD hydrogenolysis would be k2/k5 = 0.053 , or the CO bond cleavage would be tremendously favored over the C-C bond cleavage. The experiment shows otherwise. Therefore, the dehydration reaction must be reversible in reality in 2,4-PD hydrogenolysis. The retro-aldol reaction (Reaction No.2) may also be reversible in 2,4-PD hydrogenolysis, depending on the reaction conditions. In order to describe the degree that the above reactions are reversed, a reversibility parameter may be defined for each of these reactions, as the ratio of the forward reaction rate to the reverse reaction rate, or R _ k_2[Cl [51 [OH] _ k_2[Cl [£1 2" 1:218] [0H1 ‘ 1:213] (37) and 62 _ k-5IGI [0H] _ k_5 [G] 5 ’ kslB] [0H] ‘ k5 [B] (38) As later we will see, the reaction selectivity is closely related to these two parameters. With the newly defined reversibility parameters, we may rewrite Eqs.(2) and (5) as X2 1,181 [0H] (1 -112) (39) and X5 1:513] [0H1 (1-R5) (40) Dividing Eq.(39) by Eq.(40) and solving for R 5 yields: kZXS R5 = 1 — @(r—Rz) = O.97+0.03R2 (41) By definition, 0< R2 < 1. Hence, 0.97 < R5 < l, which implies that the dehydration reaction is highly reversible in the 2,4-PD hydrogenolysis process. Also derivable from the above equations are: B X2 I l ' k2(1—R2) [0H] (42) szs [E] = k_,(1-R,)1C110H1 ‘43) and ex x [G] - 5 2 5 (44) _ k2k_5(1—R2) [0H] _k__5[0H] Thus, [B], [E] and [G] can be determined if [OH] and R2 are assumed. The pH of the reaction solution has been checked during the experiment to be around 11. The retro-aldol reaction is speculated to be reversible but to a much less degree than the dehydration 63 reaction. Let [OH] = 1 mM and R2 = 0.2. It is obtained that [B] = 0.147 mM, [E] = 0.838 mMand [G] = 0.305 mM, allless than 1 mM. From Eqs.(36) and (41), it is also obtained that [M] = 19 mM, [MH] = 9 mMand R5 = 0.976. Now that all the product concentrations and reaction fluxes are known, all the concerned reaction rate constants are ready to be determined, providing that an appropriate value is assigned to K0 . The equations needed are derived from Eqs.(3l) to (35) and Eq.(16), and are given as follows: XI k = 4 -‘ [MH] (15X, [A] - [81) ( 5) X3 k = 46 3 [MH1([C]-I_(3KOID1) ( ) X4 = 47 "4 [MH]([E]—I_(4KOIFT) ‘ ’ X6 "6 = [MH] [Gr (48) X7 k = 49 7 [MH]([ll 45,19,111) ( ) and X k 8 (50) -8 "" [MH] (X01111 -1) Let K0 = 5000. k,,, k3, k4, k6, k7 and k_8 are calculated as 67.1, 1.89, 11.9, 22.1, 1.16 and 3.22x10'5 M" min", respectively. These rate constants satisfy that k_1 > k7 and that 64 k6 > k4 > k3 > k7 , and thus are consistent with our knowledge of hydrogenation reactions. With the above just determined rate constants, k1, k,3, k,.,, k; and k3 can be further calculated from Eqs.(22) through (26). Table 5-8 below summarizes all the rate constants obtained from the above procedure. Table 5-8 Summary of Determined Reaction Rate Constants In 16 k. k. In k6 k7 kt (M'Zmin‘l) (Mi-him") (M‘zmin'j) (M’zmin'l) (M‘zrnr'n'l) (M‘zmin'l) (M‘Zmin'l) (M'zmin'l) 435 752 1.89 11.9 14100 22.1 1.16 76.2 k_1 k_2 k,3 k4 k,5 k,6 k_7 k,8 (M‘Zmin‘I) (M‘Zmin‘l) (M‘zmin'l) (M’min‘l) (M‘znu'n‘l) (M‘zmin'l) (M‘zmin‘l) ("tin") 67.4 4480 12.2 1.24 6590 o 7.48 3.32x10'5 Although some arbitrary measure has been used in the above rate constant determination process to assign values to R2 and K0, the determined rate constants remain sensible in the sense that: a) they are thermodynamically consistent; b) they are within a reasonable range of absolute and relative magnitude; and c) most importantly, the simulation results obtained with these rate constants are comparable to those obtained experimentally, as will be seen in next section, which is the ultimate test for the determined rate constants. These reaction rate constants may not represent Ni on kieselguhr or any other catalysts 65 that have actually been used in our experiments, but they may represent a hypothetical catalyst which is not very different from them. If this is the case, the effects of reaction conditions on the selectivity should be similarly reflected on the actual and the hypothetical catalysts. Thus, the simulation study based on these rate constants would be useful at least qualitatively in interpretation of the experimental results, which is our goal in this study. 5.3.3 Simulation Results and Discussion Figures 5-6 to 5-9 present the results of a simulated run of 2,4»PD hydrogenolysis. The assumed reaction conditions for this simulated run are the same as those for the experimental run shown in Figures 5-3 and 5-4, except that the experimental run has undergone a temperature rise from ambient temperature to 220 °C during the initial period of reaction, while the simulated run is computed under isothermal condition. Because of this difference in temperature history, the concentration and selectivity profiles in Figures 5-6 and 5-7 do not match those in Figures 5-3 and 5-4 exactly. Nevertheless, they still bear the major features of the later profiles. For example, the selectivity profile in Figure 5-7 also becomes time-independent after the initial period of reaction. The concentration of acetone (C) and 2-pentanone (I) in the simulated run also undergo an up-and-down behavior during the reaction, as in the experimental run. In addition to the information on major hydrogenolysis products, the simulation study also reveals how the concentration of minor products and the concentration of the base catalyst (hydroxide ion) change during the reaction (Figures 5-8 and 5-9). All the minor products shown in Figure 5-8 seem at a higher level than that observed in the experimental run, which, however, may result from 66 3 : : : 55.100 --------- .i -------------------- : ---------- : --------- — § 5 ‘ E A 80 ------------------------ ‘7 -------- : --------- —t 9 ‘ : :' g eo---------3- ............................. 3 ......... - I . t , D I I / ’ IK 4o----—-----.+ ------------------ 12"": ----- ; --------- - I , /I l / l r I .. / ...... 1- , J 20I'"-—-‘:-:.-:€_V'5’-’:’-;‘-;“= ........ :___ _ — _"‘_-_._*____ Slur : ' C 0 ,4” ’ l . '. l A O 50 100 150 200 250 Time (min) Figure 5-6 Simulated Concentration Profiles of Major Products. Simulation conditions: 220 °C, [H]=100 mM, [M]o=28 mM, [OH]O=10 mM, [A]o=120 mM. (as) ____________________________________ N ([C]+[D])/II)GI+[|I+IJII S: in 0 50 100 150 200 250 Time (min) Figure 5-7 Simulated Selectivity Profile. S is calculated as ratio of the sum of [C] and [D] to the sum of [G], [I] and [J]. [B]. [E]. [G] (mM) [0H] (mM) 10 I I T I : r : I I | I I I I ‘ I 8------——----: ----------- : ----------- « I t I I r t t t t t r i J ' 6»-——--—------T -------------------------------------- -— .o—“fl‘x r ,,,,, x. l‘"\—-"" ..... I \‘\ 4.4 ____________________________ ‘34.; ———————————————————— _ I \5.‘ \- ' \'\\‘. I I \\\E ' 2IC".I..'1'¥"”""""‘~"’\"."_('37""\<~Z"'T”"‘ ‘ , \ ~ \‘xJ W i \ ~ :‘X 0 I r B 1 L ' O 50 100 150 200 Time (min) Figure 5-8 Simulated Concentration Profiles of Minor Products 10 _________________________________________ O 50 1 00 1 50 200 250 Time (min) Figure 5-9 Simulated Base Concentration Profile 68 Selectivity, S l O 5 10 15 20 25 30 35 40 Base Concentration, [OHlo (mM) Figure 5-10 Simulated Effect of Base Concentration on Selectivity. Simulation Conditions: 220 °C, [H]=100 mM, [M]o=28 mM, [A]o=120 mM, varying [OH]0. Selectivity, S o: '00 —L 1:. 1.2 1 1 1 1 1 10 1 5 20 25 30 35 4O 45 50 Catalyst Concentration, [M10 (mM) Figure 5-11 Simulated Effect of Catalyst Concentration on Selectivity Simulation Conditiom: 220 °C, [H]=100 mM, [OH]0=10 mM, [A]o=l20 mM, varying [M]..- Selectivity, S is a: in .NA O 50 Hydrogen Concentration, [H] (mM) 100 150 200 Figure 5-12 Simulated Effect of Hydrogen Concentration on Selectivity Simulation Conditions: 220 °C, [M]o=28 mM, [OH]0=10 mM, [A]o=120 mM, varying [H]. UT Selectivity, S h (A) L 100 150 Rate constant. k6 (1/M/min) 200 Figure 5-13 Simulated Effect of k6 on Selectivity. Simulation Conditions: 220 °C, [H]=180 mM, [M]o=28 mM, [OH]0=10 mM, [A]o=120 mM, but varying k6. 70 the temperature history difference that exists between the experimental and simulated runs. As in the experimental study, the effects of reaction conditions on the selectivity are also investigated in the simulation study. Figure 5-10 shows how the selectivity is affected by the change in base concentration. As observed in our experiments, the simulation study also indicates that the selectivity increases with the base concentration. This effect is most profound at low base concentration, but diminishes as the base concentration gets high. Because of the limited range of base concentration explored, the independence of the selectivity on base concentration at high base concentration is not revealed by the experiment Figure 5-11 shows how the selectivity is affected by the change in the amount of catalyst used in the reaction. Again as in our experiment, the selectivity is found to decrease with the catalyst amount. This effect remains over the whole range of catalyst concentration that has been explored. Lower catalyst concentration makes the reaction more selective, but it also makes the reaction slower. As for the effect of hydrogen pressure on the selectivity, one set of experimental data (with Raney Cu as catalyst) show that it increases with the hydrogen pressure, while the other set of data (with Ni on Kieselguhr as catalyst) show that it is basically independent of the hydrogen pressure. The simulation study reveals a more complete picture and provides a good explanation to the above apparently different behaviors of the selectivity. It shows that, for a given catalyst, the selectivity increases with the hydrogen pressure at low pressure, is indifierent to the change in hydrogen pressure within the medium pressure range, and decreases with the hydrogen pressure at high pressure (Figure 5-12). These 71 low, medium and high pressures are relative, and are catalyst-dependent. For Raney Cu, 7 MPa must be still in its low pressure range, while for Ni on kieselguhr, 3 to 7 MPa must happen to be its medium range. This explains why with different catalysts the selectivity behaves differently within the same hydrogen pressure range. The efiects of reaction conditions on the selectivity may be understood at a deeper level with the help of some mathematical equations relating the selectivity to base concentration, metal hydride concentration and the reversibility parameter for the retro- aldol reaction. These equations are derived as follows. To start, we have )12 k2(1—R2) [0H] [B] k2(1-R2) S = Y. = k,(1-R,)10H1 [B] = err-R.) ‘5” Assuming [G] is at a pseudo-steady state, we have £161 = "sIBI [0H] "£5101 [0H1 -k6[Gl [MH] so (52) which leads to [G] *5 [OH] 1731 = k_5 [0H] + 16 [MH] (53) Substituting the above equation into Eq.(38) and rearranging yields R [1 k6 [MH] I1 (54) 5 ’ 1 14510111 Further substituting Eq.(54) into Eq.(51), we obtain ".2 k-swm 55 S=k5(l-R2) l-l-W ( ) 01' 72 k2 k_5 [cm]0 s = E‘l’R2)[l+k_6(W—HT’1)] (56) Replace the rate constants in Eq.(56) with numbers. We get s = 0.053 (1 —R2) [1 +298( [5;];- 1)] (57) which provides additionally some quantitative ideas. Eqs.(56) and (57) clearly show that the selectivity will increase as [OH]0 increases, if other variables are kept unchanged. However, increase in [OI-I]o will inevitably lead to increase in [MH]. Increase in [MH], as well as increase in R2 , leads the selectivity to decrease. The level of R2 depends on levels of [OH] and [MH], as well as other factors. The simulation study reveals that, at low base concentration, R2 decreases as the [OH]0 is increased. The combination of the effects of [0H]oand R2 overrides the effect of [MH]. As a result, the selectivity shows an overall increase as the base concentration is increased. However, at high base concentration, R2 increases with [OH]0. The effect of [OI-I]o is offset by the combination of the effects of [MH] and R2. The overall result is a selectivity independent of base concentration. Increase in the amount of catalysts used simply increases the metal hydride level in the solution. Change in R2 is minor. Therefore, the selectivity decreases with the catalyst amount. Increase in hydrogen pressure decreases both [MI-I] and R2. At low hydrogen pressure, R2 is very sensitive to the change in hydrogen pressure. As the pressure increases, it becomes less and less sensitive, which explains why the selectivity responds to the change in hydrogen pressure diflerenfly in different pressure range. From Eq.(56), it is also seen that the selectivity should increase as the rate constant k6 73 decreases. This deduction is confirmed by the simulation results as shown in Frgure 5-13. Since different catalysts feature different k6 values, the dependence of the selectivity on k6 explains why it varies widely with the type of catalysts used. To improve the reaction selectivity, catalysts with small k6 value should be used. Figure 5-13 suggests that the effect of k6 on the selectivity is very profound. Therefore, the selectivity may be well improved by use of a good catalyst. A good catalyst implies that this catalyst possesses only a moderate capability to hydrogenate the cab-unsaturated ketone while still keeping a reasonable capability to hydrogenate saturated ketones and aldehydes. Lowering the reaction temperature may also lead to a reduced k6. However, it will also change other rate constants, including k2, k5 and k5 in Eq.(56). As a result, the selectivity may not be improved by reduction of the reaction temperature. As a matter of fact, our experiments indicate that the selectivity is not sensitive to the change in temperature, at least within the temperature range we have explored. PART II SELECTIVE DEH YDROX YLAT I 0N 0F GLYCEROL CHAPTER 6 INTRODUCTION: GLYCEROL DEHYDROXYLATION 6.1 On-going Research on Glycerol Dehydroxylation The purpose of glycerol dehydroxylation is to convert glycerol to 1,3-propanediol (1,3-PD) by removing the second hydroxyl group of glycerol. Research on this process has been motivated mainly by two considerations. First, the current commercial processes to produce 1,3-PD from acrolein and ethylene oxide are not efficient, as has been discussed in Chapter 1. Second, glycerol is a biomass-derivable chemical. It can not only be potentially produced from sugars through fermentation and hydrogenolysis, but also be obtained from processing of animal fats and vegetable oils. Considering the promise of 1,3-PD in polyester fibers, films and coatings, it is significant being able to derive this chemical from biomass sources. The glycerol dehydroxylation processes reported so far are mostly fermentation processes (Gunzel at al., 1991; Boenigk et al., 1993; Zeng er al., 1993; Biebl, 1995; Barbirato, 1996; Cameron, 1998; Reimann, 1998). Nevertheless, Che (1987) has invented an one-step catalytic process to convert glycerol to 1,3- and 1.2-PD’s. This process uses a synthesis gas co-feed, and is typically carried out under a temperature of 100-200 'C and a pressure of loo-10,000 psi, in the presence of tungsten and a Group VIII component. Like the current 1,3-PD production processes, however, this process furnishes only a low yield of 1,3-PD. The main products of this process are propanol and propylene glycol. 1,3-PD 74 75 accounts for only a small portion of the reaction product (generally less than 20%). The fermentation processes are based on the observation that, under appropriate conditions, certain bacteria, such as Colosm'um, K. pneumoniae and Citrvbacterfi'eundii, can convert glycerol to 1,3-PD. This conversion is generally accompanied with formation of by-product ethanol, acetic acid, butyric acid, lactic acid, hydrogen and carbon dioxide, but by careful control of the fermentation conditions the yield of 1,3-PD can be maximized. Research on glycerol fermentation has been started since the 1960’s. At the current time, this process is able to furnish a 1,3-PD yield as high as 0.66 mole per mole glycerol or 55% on weight basis (Zeng et al., 1993), which, unfortunately, is about the highest yield that one can obtain from the glycerol fermentation process. Having a low theoretical yield is one of the major disadvantages of the fermentation process. The reason for this low theoretical yield can be understood by examination of the reaction pathway of glycerol fermentation. Figure 6-1 shows a simplified reaction pathway of glycerol fermentation, which is derived from a more complex one in Zeng er al. (1993). From Figure 6-1, it is seen that the conversion of glycerol to 1,3-PD is completed through dehydration of glycerol followed by hydrogenation of the resulting 3- hydropropionaldehyde (HPA). Hydrogenation of HPA consumes NADHZ, which is regenerated in the oxidation of glycerol to pyruvate, the precursor to by—product ethanol, acetate, lactate, butyrate, H2 and C02 (the latter species are not shown in Figure 6-1). In other words, the conversion of glycerol to 1,3-PD is sustained by the formation of by- products, which inevitably reduces the theoretical yield of 1,3-PD. The theoretical yield that a fermentation process can provide varies with the by-product it produces. In the ideal case, only acetate would be produced as the by-product. As conversion of pyruvate to 76 Glycerol 1120 3-hydroxypropionaldehyde NADH2 NAD 1,3-propanediol 2NAD 2116113112 Pyruvate l Acetic acid 2mm2 ZNAD Ethanol 11.4113112 NAD Lactic acid Figure 6-1. A Simplified Reaction Pathway of Glycerol Fermentation acetate does not consume NADHZ, the theoretical yield of 1,3-PD would be 67% in this case. When other by-products are produced, the theoretical yield is expected to be lower than 67%. Another major disadvantage of the fermentation process is that this process is substrate-inhibited. The bacteria used in the fermentation are generally not able to stand a glycerol concentration above 17%. As a result, both the product concentration and the productivity are low. In a continuous process, Zeng et al. (1993) have obtained a 1,3- propanediol concentration of 30 g l‘1 with a productivity of 8.1 g 1'1 h'l, which are the best results reported so far in literature. 77 6.2 The New Conversion Concept and Supportive Work In the current research, we propose a new glycerol dehydroxylation concept to selectively remove the second hydroxy group of glycerol and thus to convert it to 1,3-PD. This new conversion concept has potential to provide a high 1,3-PD yield and does not have the disadvantages of the current glycerol dehydroxylation processes. The new glycerol dehydroxylation concept is presented in Figure 6-2. It consists of three steps, namely acetalization, tosylation and detosyloxylation. The idea is to selectively transform the second hydroxyl group of glycerol into a tosyloxy group (tosylation) and then remove it by catalytic hydrogenolysis (detosyloxylation). Compared to the hydroxyl group, the tosyloxy group is a better leaving group and is easier to replace with a hydride ion. In the following, a detailed discussion is made on each of the conversion steps. Wu, In order to selectively tosylate the second hydroxyl group of glycerol, the first and third hydroxyl groups of glycerol need to be protected. According to the proposed conversion concept, this is completed by acetalizing glycerol with benzaldehyde. Acetalization is an equilibrium reaction, but can be driven to completion by removing the water formed in the reaction. The difficulty with this step is that, besides the desired 1,3-product (S-hydroxyl-Z-phenyl-l,3-dioxane, or HPD), the undesired 1,2- product (4—hydroxylmethy-2-phenyl—1,3-dioxolane, or HMPD) may also be formed. Nevertheless, these products can be separated, and the separated 1.2-product can be returned to the acetalization reactor, where either it itself can be converted to the 1,3- product or it can help shift the reaction toward the 1,3-product. 78 Step 1: Acetalization OH + PhCHO H OH ’ + O OH H h h glycerol S-hydroxy-Z-phenyl- 4-hydroxymethy-2-phenyl- 1,3-dioxane (HPD) 1,3-dioxolane (HMPD) Step 2: Tosylation H S TsCl » h h HPD 2-phenyl-5-tosyloxy-1,3-dioxane (PTD) Step 3: Detosyloxylation and Group Protection Removal S + H20 01‘ng ———-> OIH !H - PhCHO - TsOH h 2-tosyloxy- 1 3-pl'0panediol 1.3-propanediol (TPD) PTD or S + H2 + H20 ———> —-——> ' T30“ - PhCHO OH H h h 1.3-propanediol pTD 2-phenyl-l.3-droxane Figure 6-2. Illustration of the New Concept of Glycerol Dehydroxylation 79 Besides benzaldehyde, many other carbonyl compounds could also be used as the hydroxyl group protection regent. The condensation between glycerol and various aldehydes and ketones has been extensively reviewed by Showler and Barley (1967). Generally, the reaction between glycerol and aldehydes gives more 1,3-product than the reaction between glycerol and ketones (Trosr; 1991). For our purpose, it is obvious that aldehydes are preferred to ketones. Besides the product yield, the ease of separating the 1,3- from the 1,2-product is another important factor that needs to be considered in selection of the group protection regent. In terms of the later aspect, benzaldehyde is superior to aliphatic aldehydes. The products from aliphatic aldehydes are generally high-boiling-point oils (Piasecki and Burczyk, 1980; Hilbbert and Carter, 1928), and the differences between the boiling points of the 1,2 and 1,3 products are usually small. For example, the condensation between glycerol and acetaldehyde gives four products: the cis- and trans- 5-hydroxyl-2-methyl- 1,3-dioxanes (1,3-product) and the cis- and trans- 4-hydroxylmethyl-2-methyl-1,3- dioxolanes (1,2-product). As documented in Beilstein, the boiling points of these products are 73, 100, 86, and 94 °C, respectively, under 18 torr. The 1,2- and 1,3- products from benzaldehyde have considerably different solubilities in benzene-ligroin mixture. The 1,3-product can be readily crystallized from the benzene-ligroin solution (Hill et al., 1928; Baggett er al., 1960). The reaction between glycerol and benzaldehyde has been carried out previously, in several ways. The early methods (Irvine et al., 1915; Baggett er al., 1960) allowed direct contact of glycerol with benzaldehyde. In some cases, acids, such as H2804, were added as catalysts. The late methods (Baggett et al., 1965; Szeja, 1983) are featured by use of a 80 benzene solvent. The advantage of doing the reaction in benzene is that the water formed in the reaction can be removed azeotropically, and thus the reaction can be driven to completion. This advantage can be further enhanced by using a Dean-Stark u'ap (Casey, et al., 1990), as Szeja (1983) has done. Acid catalysts were generally used in the late methods. The condensation between glycerol and benzaldehyde gives four products: the cis and trans- HPD’s and the cis- and trans- HMPD’s. As mentioned above, separation of the 1,3-product HPD from the 1,2-product HMPD has been achieved by Hill et al. (1928) and Baggett et aL (1960), through crystallization from benzene-ligroin solution. Although the trans—HPD (b.p. 63.5 'C) can also be potentially crystallized, this procedure yields mainly cis-HPD (b.p. 83.5 °C) . Cis- HPD is the more stable diastereomer of HPD (Showler and Darley, 1967). The equilibrium between the 1,2- and 1,3- products under the acetalization conditions has been studied. Hill et al. (1928) showed that the 1,2- product is favored over the 1,3-product by a ratio of 5.7: l in the absence of solvent. As stated by the authors, this result was more qualitative than quantitative, because only the fractionated part of the products was assumed as the 1,3-product in their study. The data from Baggett er al. (1965) also favor the 1,2-product, but only by a ratio of 1.9:1. The data documented in Iochims and Kobayshi (1974) show the contrast to the previous work: the 1,3-product is favored over the 1,2-product (by a ratio of 1.5:1). Both the data of Baggett er al. and of Jochims and Kobayshi were obtained by integration of the characteristic H1 NMR signals of the products, but the former was done in dioxane and the latter was done in chloroform. The solvents used may have effect on the position of the reaction equilibrium. Another 81 example that favors the 1,3-product is given by Juaristi and Antiunez (1992). They obtained 11 47:30:23 mixture of HMPD, cis- HPD and trans- HPD from the reaction between glycerol and benzaldehyde. Upon continued exposure to acid, this ratio changed to 22:34:44. Therefore, they concluded that the reaction thermodynamics favors the 1,3- products. No solvent was mentioned in their work. Clearly, more research is needed to accurately determine the equilibrium ratio of the 1,2- and 1,3- products. However, the previous work should have cleared any doubts about the feasibility to quantitatively produce the 1,3-product (mainly cis-HPD) from condensation of glycerol with benzaldehyde, and about the ready interconversion between the 1,2- and 1,3- products under acidic conditions. W In this step, the unprotected hydroxy group of glycerol is transformed into a tosyloxy group, so that it can be removed in the subsequent detosyloxylation reaction. This group transformation is completed by treating HPD, the 1,3—product from the first-step reaction, with tosyl chloride. Tosylation of alcohols is a well-known reaction and is generally carried out under basic conditions. Bases are needed to neutralize the hydrogen chloride formed in this reaction. If not neutralized, the chlorine ion can potentially attack the newly formed tosyl group (Tipson, 1953). A general procedure exists for tosylation of alcohols in pyridine (Fieser and Fieser, 1967). Pyridine functions both as a solvent and as a base in this reaction. The hydrogen chloride produced in tosylation may form a complex with pyridine, which crystallizes from pyridine at 0 °C. Based on this procedure, Juaristi and Antiunez (1992) have tosylated both the cis- and trans- HPD’s in 98% yields. In another example, Hesse] et al. (1954), by using a similar procedure, was able to obtain 2-phenyl—5-tosyl-1,3-dioxane 82 (PTD) in 94% yield in the tosylation of HPD. Besides the organic base, inorganic bases, such as sodium hydroxide and sodium carbonate, can also be used in alcohol tosylation, which has been reported in several incidents to tosylate other alcohols (Tipson, 1953; Flowers, 197 1). W In order to finally convert the selectively tosylated glycerol to 1,3-PD, the tosyloxy group, as well as the group protection, needs to be removed from PTD. As shown in Figure 6-2, there are two potential pathways to go from PTD to 1,3-PD. Either the hydroxy group protection can be removed, followed by removal of the tosyloxy group; or alternately the tosyloxy group can be removed, followed by removal of the hydroxy group protection. In either approach, removal of the group protection can be done with a hydrolysis reaction, and removal of the tosyloxy group is to be done with a hydrogenolysis reaction. The hydrolysis reaction regenerates benzaldehyde, which can be reused in the acetalization reaction. If the hydrolysis reaction is canied out first, 2-phenyl-l,3-dioxane results as an intermediate. Hydrolysis of this compound is expect to be easy, because no additional functional groups other than the ether linkages may be affected. If the hydrolysis reaction is carried out first, however, caution needs to be exercised to prevent the potential attack of solvents on the tosyloxy group of PTD and/or TPD. Although according to Flowers (1971) tosylates are exceedingly stable under the acid hydrogenolysis conditions, and many successful examples of tosylate hydrolysis have been documented by Tipson (1953), our experiments show that the tosyloxy group is subject to the attack of protic solvents at temperatures above 60 °C. Hydrolysis of PTD yields TPD. Although this reaction has never been reportedly 83 done, TPD has been synthesized by Chong and Sokoll (1993) through a different route, which makes both the H1 and C13 NMR data available for characterization of this compound. Removal of the tosyloxy group is expected to be the most difficult step of the proposed glycerol conversion concept. According to Figure 6-2, detosyloxylation is to be done through a hydrogenolysis reaction with molecular hydrogen as the reducing regent In the literature, hydrogenolysis of tosylates is generally effected with a lithium hydride, such as lithium aluminum hydride (LiAlH4) or lithium triethylborohydride (LiHBEt3). This reaction has never been reportedly done using molecular hydrogen as the reducing regent. Although LiAlH4 and LiHBEt3 are effective as detosyloxylation regents, however, they are too expensive to use on industrial scale. Therefore, it is up to us to develop a catalytic tosylate hydrogenolysis process with molecular hydrogen as the hydrogen source. In summary, the new glycerol dehydroxylation concept presented in Figure 6-2 has provided us with an alternate and promising route to convert glycerol to 1,3-propanediol. The first and second step reactions of the new concept have been successfully performed by previous researchers, although for different purposes. The third step reaction, detosyloxylation with molecular hydrogen as reducing regent, is the key to the success of this new conversion concept, which has never been reportly done in literature. 6.3 Objective and Scope of Current Research The objective of the current research is to explore the technical feasibility of the new glycerol dehydroxylation concept presented in Figure 6-2. As discussed above, removal 34 of the tosyloxy group from the selectively tosylated glycerol is expected to be the most difficult step of the new conversion concept. Therefore, our primary task in this work is to develop a catalytic hydrogenolysis process to remove the tosyloxy group from ether PTD or TPD. In this capacity, the two potential pathways shown in Figure 6-2 to carry out the conversion from PTD to 1,3-PD has been tested and compared. The reactions involved in the acetalization and tosylation steps have been performed by previous investigators for different purposes. The results from their work are promising. In this work, these reactions are re-performed to verify the published procedures and to provide the PTD needed in studying the detosyloxylation and protection removal reactions. CHAPTER 7 EXPERIMENTAL PROCEDURES 7.1 Acetalization The acetalization between glycerol and benzaldehyde is an equilibrium reaction. To drive this reaction to completion, the water formed in this reaction needs to be constantly removed from the reaction flask. In this experiment, water removal is done by performing the reaction in benzene solvent and removing the water as an azeotrope with benzene. Figure 7-1 shows the experimental set-up used in this experiment. This set-up includes a round-bottomed reaction flask, a condenser and a Dean-Stark trap. The function of the Dean-Stark trap is to trap the water formed in the reaction. The reaction takes place in the flask. The water formed in the reaction is boiled off from the reaction flask as an azeotrope with benzene (boiling point: 65 'C). The vaporized water and benzene flow condense in the condenser and flow down into the Dean-Stark trap. As water and benzene are immiscible and have different densities, they separate in the Dean- Stark trap with water forming the bottom layer. The Dean-Stark trap is pre-filled with benzene. The condensed benzene is returned to the flask by overflow. In a typical experiment, 100 g glycerol, 120 g benzaldehyde (c.a. 6% excess) and 300 mL benzene, together with l g p-toluene sulfonic acid catalyst are placed in a 500 mL round-bottom flask, equipped with a magnetic stirring bar. The reaction is initiated by bringing the reaction solution to a boiling state. Thereupon, vaporization and 85 86 Condenser _ \K W.../ Dean-Stark Thermometer Reaction flask / Sand bath 0 © Magnetic stirrer / with hot plate Figure 7-1. Experimental Set-up for Acetalization of Glycerol with Benzaldehyde 87 condensation commences in the reaction flask and the condenser, respectively, and the water is seen to accumulate at the bottom of the Dean-Stark trap. The Dean-ka trap is graduated. Therefore, progress of the reaction can be monitored by the volume of the water accumulated in the Dean—Stark trap. When the water in the Dean-Stark trap reaches a steady level, the reaction is completed. Completion of the reaction is also indicated by the temperature change of the reaction solution. Before completion, the temperature stays constant at about 80 °C. When the reaction is close to completion, the temperature starts to climb, until a new level (about 160 °C) is reached. After the reaction is completed, the valve with the Dean-Start trap is turned open, and the benzene is boiled off, leaving only the product in the reaction flask. The reaction product is a mixture of cis- and trans- 5-hydroxy-2-phenyl-1,3-dioxane (HPD, the 1,3-product) and cis- and tran- 4-hydroxy-methyl-2-phenyl-1,3—dioxolane (HMPD, the 1,2-product). A small amount of benzaldehyde may also exist in the product, because the excess benzaldehyde used in the reaction. To isolated HPD from this mixture, it is crystallized from a benzene-ligroin solution. The procedure is taken from Hill et al. (1928) and is documented in the following: 1) Dissolve the reaction product in a benzene-ligroin mixture. Typically, for an amount of 60 g of reaction mixture, about 90 mL of benzene and 150 mL of ligroin are needed. 2) Leave the above solution in a freezer for about one hour. At this time, crystals may not form from the solution. 3) Slowly add benzene while shaking, until the oil layer disappear. Crystals may or may not form at this time. Place the solution in the freezer, for about another hour. Crystals should form at this point. 88 4) Filter the solution with a Buckner filter under vacuum, and wash the crystals with a 50% benzene-ligroin solvent. Continue washing until the yellow color disappears. The yellow color indicates the presence of benzaldehyde in the crystals. At the end of this procedure, a white, wax-like crystal product yields, which is cis-HDP. The trans-HDP has a better solubility in the benzene-ligroin solvent and is more difficult to crystallize than the Cir-HPD. The above procedure can be repeated to further purify the HDP product obtained. Analysis of the raw reaction product is performed with GC. Characterization of the purified HPD is done with 1H and 13C nmr. The GC column used is a DB-S column from J&W. The carrier gas is hydrogen (at a flow rate of 1 mllmin). The oven temperature is programmed so that it stays at 150 °C for one minutes, then increases at a rate of 40 'C/ min and then stays at 320 °C for 3 minutes. The retention times for the relevant compounds are documented in Table 7-1. Table 7-1. GC Retention Times for Acetalization Products Compounds Retention time (min) benzene 1.30 benzaldehyde 1.42 cis-HPD 3.36 cis- and trans- HMPD 3.48 trans-HPD 3.83 89 7.2 Tosylation Tosylation of 5-hydroxy-2-phenyl-1,3-dioxane (HPD) is carried out in pyridine. To start the experiment, HPD and TsCl are placed in an Erlenmeyer flask. Then to the flask is added cold dry pyridine. The flask is shaken until both HPD and TsCl dissolve in pyridine. Then the flask is placed in a refrigerator (c.a. 5 °C) to allow the reaction to continue for about 12 hours. Typically, 5-10% excess TsCl is used in this reaction, and for each gram of HPD, about 20 mL pyridine is needed. Progress of this reaction can be monitored by formation of needle crystals. These crystals are pyridine-hydrochloride complex. When no more crystals are forming, the reaction is complete. At this time, the flask is removed from the refrigerator. The content of the flask is poured with stirring into a beaker containing about 6 g ice for each milli- liter pyridine. The tosylate product (2-phenyl-5-tosyl-1,3-dioxane, or PTD) precipitates immediately. Stirring is continued until all the ice melts. Then the solution is filtered with a Buckner funnel under vacuum, and the crystals is washed with water for three times. Finally, the crystals are dried in a desiccator under vacuum. After this procedure, PTD is obtained as white dry wax-like crystals. Characterization of PTD is accomplished using 1H and 13C nmr. 7.3 Hydrolysis Hydrolysis of PTD has been performed in both water and methanol. To carried out the reaction in water, PI'D from the previous reaction is placed in a glass vial equipped 90 with a magnetic stirring bar. Then to the glass vial is added a small amount of p-toluene sulfonic acid (T sOH) and a large amount of water. PTD is not soluble in water. Therefore, it floats on the water at beginning of the reaction. The glass vial is heated in a water or sand bath. The reaction solution is constantly stirred. After 4 to 6 hours' heating and stirring, the PTD floating on the water disappears, and yellow oil droplets are seen at the bottom of the glass vial. These yellow dmplets are the benzaldehyde formed in the reaction. Upon cooling and sitting, benzaldehyde separates from the aqueous phase. The benzaldehyde phase is removed from the glass vial. The aqueous phase is neutralized with NaOH, and is transferred to the hydrogenolysis reactor for further reaction. Possibly because of the two hydroxyl groups present in 2-tosyl-1,3-propanediol (TPD), most of the TPD produced in the hydrolysis reaction stays in the aqueous phase. To carry out the hydrolysis reaction in methanol, PTD is again placed in a glass vial. Then to this glass vial, the methanol solution of TsOH is added. The glass vial is simply placed in an oven at 40 °C. After 2 hours, the reaction is completed. The benzaldehyde formed in the reaction does not separate from the methanol phase. To isolate PTD, the reaction product is first extracted with hexane to remove benzaldehyde, and then it is dried at a temperature below 40 °C to remove the methanol solvent. Thus obtained PTD is re- dissolved in dioxane and then transfer to the hydrogenolysis reactor. Comparing the above two procedures, the later one is more successful than the former. The hydrolysis reaction in water is usually canied out at a temperature above 80 °C. At this temperature, tosylates are subject to the attack of protic solvents, as is discovered in this work. The hydrolysis reaction in methanol is carried out at a 91 temperature of 40 °C. At this temperature, the solvent attack on the tosyloxy group is insignificant. 7.4 Hydrogenolysis In this work, hydrogenolysis of tosylates is carried out under hydrogen pressure with either Ni on kieselguhr or Ru on carbon as the catalyst. The same experimental system and the same experimental procedure as described in Chapter 3 are used, but the reactor used in this experiment has a smaller capacity (10 mL). The 1,3-PD produced in TPD hydrogenolysis is quantified using GC. CHAPTER 8 RESULTS AND DISCUSSION 8.] Acetalization Acetalization with benzaldehyde protects the first and third hydroxyl groups of glycerol from being tosylated and removed in the subsequent reactions. This reaction has two important features: a) it is an equilibrium reaction, but can be driven to completion; and b) it is non-selective: besides the desired 1,3-product HPD, the undesired 1,2-product HMPD is also produced, which has to be separated from HDP and returned to the acetalization reactor. Because of the later feature, the percentage of HDP in the acetalization products and the ease to separate HPD from HMPD are our primary concern in this research. Three acetalization experiments have been performed. TWo are done in benzene, and one is done in a 1:1 benzene-ligroin mixture. In the first experiment, 50 g glycerol and 60 g benzaldehyde (c.a. % excess) are condensed in 300 ml benzene. The reaction takes about 3 hours to complete. GC analysis of the raw reaction product shows three product peaks. Of these three peaks, two are determined to result from HPD (cis— and tran- HPD), and one from HMPD. The two diastereomers of HMPD can not be separated with the GC column we were using. To separate HPD from the product mixture, it is crystallized from a benzene-ligroin mixture with the procedure described in Chapter 7. The crystallized product is 92 93 characterized with 1H and 13C nmr. The obtained spectra are found to be consistent with the structure of HPD and with those documented for this compound in the literature. Measurement of the melting point of this crystallized product (75-80 'C) further identifies it as the cis diastereomer of HDP. The previously reported melting points are 37 'C for HMDP, 65 ’C for trans-HDP and 83 'C for cis-HDP. The crystallized cis-HPD is not 100% pure. It contains a small amount of trans-HPD, but very little HMPD. HMPD is believed to not crystallize with cis-I-IPD, and therefore can be removed in the process of filtration and washing. Trans-HPD may crystallize with cis-HPD, although to a much less extent. Both cis-HPD and trans-HPD are desired in the acetalization reaction. From the first experiment, about 10 g cis-HPD is obtained after purification. The yield is calculated as 25%. Before purification, the yield of cis-HPD is estimated to be about 42%, based on the sizes of GC peaks. The ratio of cis-HPD to trans-HPD to HMPD is 42:33:25. In the second experiment, 100 g glycerol is condensed with 120 g benzaldehyde in 300 m1 benzene. The (cis-) HPD yield from this experiment is similar to that of the first experiment. From these results, it is seen that the yield of HPD from the acetalization reaction is in high percentage. The procedure to separate HPD from the product mixture is fairly simple and easy to carry out. It is also seen, however, that the undesired HMDP and unrecovered HDP in combination still account for about 75% of the reaction product. They need to be recycled back to the acetalization reactor, if an industrial process is to be designed. After crystallization, the undesired HMDP and unrecovered HDP are left in the benzene-ligroin mixture. If the acetalization is carried out in the same solvent, recycle of 94 this residue will be straightforward. It is motivated by this consideration that the third experiment is run in a 1:1 benzene-ligroin mixture to explore the feasibility to carry out the acetalization in such a solvent. This experiment turned out to be a success and furnished an HPD yield of 23%. 8.2 Tosylation The purpose of tosylation is to transform the unprotected hydroxy group of HPD into a good leaving group, so that it can be removed in the subsequent reactions. In the current research, this group transformation is accomplished by treating HPD with tosyl chloride (T sCl) in pyridine. TWo experiments have been performed. In one experiment, 2 g HPD is treated with 2.4 g TsCl (c.a. 12% excess), which gives 3.1 g PTD. The yield is calculated as 83%. In the other experiment, 5 g HPD is treated with 6 g TsCl (c.a. 12% excess), which furnishes 8.1 g PTD. The yield is 87%. In the literature, a yield as high as 98% has been reported in tosylation of HPD (Juaristi and Antiunez, 1992). The lower yield in our experiments may result from the unpurified TsCl and the undried pyridine used in this experiment, which are believed to affect the yield. However, no additional effort has been made in the current research to maximize the product yield. We are convinced that higher yield is achievable if more care is exercised in performing the reaction. The obtained PTD is characterized with 1H and 13c nmr. These spectra are consistent with the structure of PTD and with the data documented in the literature (J uaristi and Antiunez, 1992). 95 8.3 Detosyloxylation Detosyloxylation, or removal of the tosyloxy group, is the focus of this work. This reaction has never been done with molecular hydrogen as the reducing regent. In this work, we attempt to develop such a hydrogenolysis process to remove the tosyloxy group from either TPD or PTD, and to create a pathway from PTD to 1,3-PD. As discussed previously, the conversion of PTD to 1,3-PD can potentially be approached in two different ways. One approach is to first remove the hydroxy group protection and then remove the tosyloxy group. The other is to first remove the tosyloxy group and then remove the hydroxy group protection. Both approaches have been explored in this work. The second approach proves to be unsuccessful. PTD is found to be inactive under hydrogenolysis conditions. The experiment has been carried out in dioxane with Ni on kieselguhr as the catalyst. The reaction is run under 3 MPa H2 pressure. The temperature has been kept at 120 ’C for two hours, at 130 “C for two hours, at 140 °C for 10 hours and then at 150 “C for another 10 hours. No product is found at the end of the experiment The initial attempt to convert PTD to 1,3-PD via the first approach did furnish 1,3- PD. However, the yield was low (less than 15%). To start this experiment, PTD is hydrolyzed at 80 ‘C in 0.01 N aqueous p-toluene sulfonic acid solution. This reaction normally takes about 4 to 5 hours. After the reaction, the benzaldehyde released from the reaction, which separates itself from the aqueous phase, is removed. The aqueous phase, which contains most of the TPD resulting from the hydrolysis, is neutralized with NaOH and transferred to the hydrogenolysis reactor to remove the tosyloxy group. Three catalysts, including Ni on kieselguhr, Raney Ni and Ru on carbon, have been used in the 96 hydrogenolysis, and none of them has provided a 1,3-PD yield higher than 27%. A more detailed account of the about results may be found in DeAthos (1995), who has assisted this author in performing the above experiments. In spite of the low 1,3-PD yield delivered by this experiment, it shows that TPD, in contrast to PTD, is reactive under the hydrogenolysis condition. Comparing the structures of PTD and TPD, the tosyloxy group of PTD is isolated, while the tosyloxy group of TPD is next to a hydroxy group. Possession of this adjacent hydroxy group must have somehow facilitated the removal of the tosyloxy group in hydrogenolysis of TPD. To confirm the above speculation, experiments are carried out to hydrogenolyze 6- tosyloxy-hexanol (TH) and 2-tosyloxy-cyclohexanol (TCH). Like PTD, TH has an isolated tosyloxy group; while like TPD, the tosyloxy group of TCH is next to a hydroxy group. As expected, TB is found to be inactive under the hydrogenolysis conditions. TCH, on the hand, is hydrogenolyzed at a temperature above 140 °C. Below 140 °C, the reaction is slow. Both the hydrogenolysis of TH and of TCH have been conducted in dioxane. Hydrogenolysis of TCH has been conducted with both Ni on kieselguhr and Ru on carbon. Hydrogenolysis of TH, however, has been conducted only with Ni on kieselguhr. The results of these experiments confirm our speculation that the presence of a hydroxy group next to the tosyloxy group facilitates its removal in hydrogenolysis. Hydrogenolysis of TCH yields two products. One is cyclohexanol and the other is cyclohexanone, as found in our GC-MS analysis. With Ni on kieselguhr as the catalyst, the ratio of cyclohexanol to cyclohexenone in the product is about 1:1; with Ru on carbon as the catalyst, the ratio is about 3:1. No other products are found. Based on the observed reaction products, the hydrogenolysis of TCH may occur as shown in Figure 8-1, with 97 OH OH O OH on slow fast slow —> —> ——> - TsOH + H2 Figure 8-1. Reaction Pathway of TCH Hydrogenolysis cyclohexanone as the precursor to cyclohexanol. The function of transition metal catalysts in this reaction is not limited to catalyzing the hydrogenation of cyclohexanone to cyclohexanol in the last step. They must be involved in the initial detosyloxylation step as well, because, in the absence of transition metal catalysts, TCH is found to be inactive at 140 °C. . The assistance of the adjacent hydroxy group to detosyloxylation can be understood on the assumption that the initial detosyloxylation step in the above reaction scheme is thermodynamically unfavorable. Presence of the adjacent hydroxy group stabilize the initial detosyloxylation product through a subsequent tautomerization reaction. In the absence of such a hydroxy group, as in the cases of PTD and TH, the subsequent tautomerization is impossible and thus the initial detosyloxylation product is not stabilized. This explains the inactivity of PTD and TH under hydrogenolysis conditions. To find out the reason for the low 1,3-PD yield in our initial attempt to convert PTD to 1,3-PD, further experiments are done to hydrogenolyze TCH in 90% aqueous methanol. To our surprise, this reaction yields neither cyclohexanol nor cyclohexanone. The reaction does provide two products, but these products are from the reactions of TCH with water and with methanol, respectively. Further experiments show that the reactions of TCH with 98 water and methanol can both be effected simply by heating TCH in these solvents over 60 “C. At 80 'C, these reactions becomes fairly fast. It has been suggested in the literature that tosylate is stable under hydrolysis conditions (Flower, 1971). This statement appears not true to TCH, and is probably not true to PTD and TPD either. In hydrolysis of PTD, it has been observed that the reaction solution becomes substantially more acidic at the end of the reaction than at the beginning of the reaction. Increase in the acidity of the reaction solution is a strong sign that the attack of water on the tosylate has occurred. The reaction of water (and methanol) with tosylates releases TsOH, which increases the acidity of the reaction solution. Based on the above study on TCH hydrogenolysis, a new procedure is developed to convert PTD to 1,3-PD. This procedure starts with a hydrolysis reaction in methanol. PTD has a much better solubility in methanol than in water. Therefore, the reaction can be carried out at a temperature as low as 40 ’C, while still giving a reasonable reaction rate (normally complete in two hours). No acidity increase of the reaction solution is observed at the end of the reaction, which implies that the solvent attack on the tosyloxy group is insignificant at 40 °C. Hydrolysis of PTD yields TPD and benzaldehyde (in the form of acetal with methanol). To separate benzaldehyde from TPD, the raw product is extracted with hexane. Then the reaction solution is dried at the room temperature to remove methanol. The dried product is mostly TPD. To hydrogenolyze TPD, it is dissolved in dioxane, and then heated up to 140 'C in the presence of Ni on kieselguhr. Dioxane is an aprotic solvent, and it does not attack tosylate. The first experiment performed with the new procedure starts with 213 mg PTD. At 99 the end of hydrogenolysis, 27 mg 1,3-PD is obtained in the product. The yield is calculated as 56%. In addition to 1,3-PD, hydrogenolysis of TPD also gives n-propanol (24%). This is understandable, because hydrogenolysis of TPD has 3- hydropropionaldehyde (HPA) as an intermediate, which may also undergo a dehydration pathway to give n-propanol, as shown in Figure 8-2. L - TsOH -H O HOCHzCHzCHO —2—> CH2=CHCHO ¢+H2 ¢+H2 1,3-PD n-propanol Figure 8-2. Reaction Pathway of TPD Hydrogenolysis From Figure 8-2, the selectivity of TPD hydrogenolysis is controlled by the relative hYCII‘Ogenation and dehydration rates of HPA. Increasing the amount of the metal catalyst increases the hydrogenation rate. Therefore, the selectivity to 1,3-PD should increase as more catalyst is used. This proves to be true. When the mole ratio of Ni catalyst to subStl’ate is increased from 1:1 in the frrst experiment to 2:1 in the second experiment we Carried out, it is observed that the mole ratio of 1,3-PD to n-propanol in the product inc“? ases from 2.3:1 in the first experiment to 4:1 in the second experiment In the second exPetirnenr, 34 mg 1,3-PD has been obtained from 210 mg 1711). The 1,3-PD yield is 72%. The n—propanol yield is calculated as 18%. No further experiments are conducted to 100 maximize the 1,3-PD yield. Nevertheless, the above experimental results show that the new procedure is promising in carrying out the conversion of PTD to 1,3-PD. In addition to the 1,3-PD yield, the catalytic nature of TPD hydrogenolysis consists of another major concern of the current research. The stoichiometrical reaction between transition metal and tosylate has been previously reported, in spite of the fact that in the reported reaction it is the O-S bond, instead of the CO bond as in TPD hydrogenolysis, ' that is broken. Therefore, in the current work, the Ni on kieselguhr catalyst used in the second TPD hydrogenolysis experiment is collected after the reaction, and is reused in hydrogenolysis of TCH. It is found that the catalyst is still active and a turnover number of about 50 has been achieved in 24 hours, at 160 “C and 3 MPa H2 pressure. At this point, there should be no doubt that the hydrogenolysis of TPD in the presence of Ni on kieselguhr is a catalytic reaction in its nature. 8.4 Design Concept of Industrial Conversion Process As the goal of this research, the technical feasibility of the new glycerol dehydroxylation concept presented in Figure 6-2 is verified on the laboratory scale by the experiments just discussed. On the industrial scale, processes as shown in Figures 8-3 to 8-6 may be designed to produce 1,3-propanediol from glycerol. The acetalization process shown in Figure 8-3 features a continuous acetalization reactor, a continuous forced-circulation HPD crystallizer and a continuous vacuum filter. The reactor is equipped with a water trap of the same principle as the Dean-Stark trap used in our bench-top experiments. The purpose is to remove the water formed in the reaction. The design of this process has incorporated the following observations made in our 101 =c=3=58< : new . 832.. 5:22;:— EEG .3 2:»...— Lozi Esooa> SNESmbU Gm: 583% gang—88¢. 9:3 8:23;“. b5? 9:2: + 0m: + Echm 523-9822 n E923 L23; \ 537. OD \ H H ”0 .253 X r\/\/\/\/. Eulcm ,\/\/\/\/\/\/\/\( .3528 2:25:53 .1”..- >[# 23:. :8 M.“ g E an... 9:2: + Eelcm .59; — = = .553 m mam So: 3.932 ocxzofificum 102 El Ea case nausea. "a new . 885 5:33:2— _2c:_u .3 2:2..— .Bzm Ezooa> 553 + 2553 > Lei—Embu Q: \ 863m coca—snack \/ r @ e 5253 r 3. So E308 :_ 23—000 mm: 103 Q. $88 522?... a 35 - 922.. 5:3..an Esra .3 2&2. .§=m E:oom> .uNESmEU DE 955 5223.6 .05? .23: Emugm 2“ rm E928 8.03m cog—amok N. + 052.3 u E038 50:: \ 2.83.5 6:553: 523:3 Q Esq-05’ 00 \ So .5225 531% E 23.08 G: + Eolom ‘ ”1, a a 28> Em suing—mason "n 13m . mauve...— =c=ahchson 5.3.5.0 6.x 9.5»; 883% flmbocuwohgz SbQSEoma>m 5:200 conszio 5:230 8:233 .983— 3306»: 2:386 E E E + 359:; + 2:38.: u Eo>_:m \ _ 92m 2 3.93.: 51:33:95 A . .. . . . 2. 523252 + 9528... osxzuancom + 3n; + Eulom Baa—:22 n Evian @ 2a; ocmxo; I 3.: + Euicm @ E038 9:38». a 105 acetalization experiments: a) acetalization of glycerol with benzaldehyde occurs readily in the presence of acid catalyst, and can be driven to completion by using a Dean-Stark trap to remove the water formed in the reaction. b) the desired 1,2-product HPD can be readily purified from the product mixture by crystallizing it from benzene-ligroin mix. The unreacted glycerol and benzaldehyde and the undesired 1,3-product HMPD will stay in the solution. c) acetalization of glycerol with benzaldehyde can be carried out in benzene-ligroin mix, as well as benzene only, thus allowing direct crystallization of HPD from the reaction solution and direct recycling of the unreacted glycerol and benzaldehyde and the undesired HMPD back to the reactor, if benzene-ligroin is used as the solvent. According to the design, the HPD rich product will be directly fed into the crystallizer from the reactor. The desired HPD is crystallized from the solution and then separated from it by the filter. Owing to the nature of the selected crystallizer, some solvent will be vaporized in the crystallizer and then condensed in the condenser equipped with the crystallizer. This solvent will be collected and used to wash the HPD cake formed on the filter surface. The filter cake discharged from the filter may need further drying. The filtrate, composed of the solvent and the undesired HMPD and the unreacted glycerol and benzaldehyde dissolved in it, is lean in HPD and is recycled back to the reactor. The benzaldehyde used in acetalization will also be recycled, but from Step 3, the detosyloxylation process. Because of the recycling of HMPD, the process shown in Figure 8-3 produces no by- product. Under acetalization condition, HMPD and HPD are readily inter-convertible. As 106 a result, the HMPD formed in the reaction can only accumulate to a certain level in the reactor, in equilibrium with the HPD present. Once that level is reached, no HMPD will be further produced in the reactor. HPD will be the only product of the reaction of glycerol and benzaldehyde, as it is continuously removed from the reactor and process. Therefore, with the process shown in Figure 8-3, one mole HPD will be expected from each mole glycerol fed into the reactor, or the HPD yield will be 100% (or close to 100% if possible loss in filter is considered), under stable conditions. High HPD yield is a major advantage of the continuous process shown in Figure 8-3. The tosylation process shown in Figure 8-4 is designed strictly based on our lab tosylation procedure described in Chapter 7. It involves three major steps: a) allowing HPD to reactor with TsCl in pyridine at low temperature, so that the hydrochloride formed in the reaction can precipitate from the reaction solution as pyridine-hydrochloride complex. b) mixing the reaction product with large quantity of ice following Step a). The purpose is to further lower the temperature and to force the tosylation product PTD to crystallize from the product solution. and c) separating PTD crystals by filtration High PTD yield has been obtained with this procedure in lab experiments. As an industrial process, however, this procedure has a major disadvantage: mixing reaction product with ice in Step b) creates a difficulty to the recycling of pyridine. At industrial scale, the pyridine solvent has to be recycled, to minimize the process cost and waste discharge. After PTD removal, the pyridine is in a mixture with water in Figure 8-4. To reuse pyridine in tosylation, it has to be first separated from water and then 107 thoroughly dried. Tosyl chloride is attacked by very small quantity of water present in pyridine. To overcome the above difficulty, an alternate process as shown in Figure 8—5 is considered. This design eliminates ice from the process and attempts to crystallize PTD directly from the reaction medium, thus allowing direct recycling of the medium, together with the unreacted starting material in it. The reaction medium for this process may be pure pyridine or a mixture of pyridine with another solvent, ligroin for example, which helps lower the solubility of PTD in the reaction medium. Pyridine, as a base, is needed in the reaction medium to neutralize the hydrochloride released from the reaction. To remove the pyridine-hydrochloride complex precipitated from the reaction medium, the tosylation reactor in Figure 8-5 is incorporated with a circulation loop, which contains a continuous in-line filter. The reaction product sent to the crystallizer is designed to be free of hydrochloride. The crystallizer in Figure 8-5 is of evaporative type, so that the solvent can be partially removed to give a PTD saturated solution in the crystallizer. The solvent removed from the solution may be used to wash the PTD crystal in filtration step. At the time being, the solubility data of PTD in pyridine or other relevant solvents are not available yet. It is also left for speculation wether the tosylation reaction can be carried out in a mix of pyridine with another solvent as smoothly as in pyridine alone. It is clear that further experiments are needed to verify the feasibility of this process. However, if successful, this process has great advantages over the one shown in Figure 8-4, in terms of both processing cost and waste discharge. Like the acetalization process in Figure 8-3, this process can potentially furnish a product yield of 100%, under ideal condition. The detosyloxylation process as shown in Figure 8-6 is designed based on the 108 discovery made in this research: the detosyloxylation of PTD can be accomplished by a hydrolysis (or methnolysis more accurately) reaction carried out in methanol at moderate temperature, followed by a catalytic hydrogenolysis reaction carried out in dioxane. It features a hydrolysis reactor, an extraction column to remove benzaldehyde from the hydrolysis product solution, a distillation column to separate the extracted benzaldehyde from the hexane extractant, a thin-film evaporator/dryer to remove the methanol solvent from the hydrolysis product TPD, and a hydrogenolysis reactor to finally convert TPD to 1,3-PD. Separation of 1,3-PD from the solvent dioxane and other hydrogenolysis products, including TsOH and propanol, is yet to be added to this process. According to the design, both the methanol and the hexane solvents used in this process are recycled. The benzaldehyde reclaimed in this process is also recycled, but back to the Step 1 acetalization process. Figure 8-7, obtained from combining and simplifying the processes in Figures 8-3, 8-5 and 8-6, illustrates the material flow in glycerol dehydroxylation. It is seen from this diagram that the acetalization process, if designed as in Figure 8-3, produces no other product but HPD, at 100% yield. Similarly, the tosylation process, if designed as in Figure 8-5, gives only PTD, at 100% yield. Therefore, the overall yield of the glycerol dehydroxylation process is solely determined by the yield of the detosyloxylation process. In our experiment, a 1,3-PD yield as high as 72% has been achieved in PTD detosyloxylation, under un-optimized conditions. This yield can be taken as the overall yield of the new glycerol dehydroxylation process. For comparison, the theoretical yield of glycerol fermentation is 67 %. From Figure 8-7, it is also seen that production of 1,3-PD is at the expense of tosyl 109 Glycerol, 1 mole V $ ACETALIZATION .———> H20 Yield=100% HPD, 1 mole V TsOH, 1 mole—+ TOSYLATTON —> HCl Yield=100% PTD, 1 mole ,_ _ _ _ _ _ _ ._ _ l HZO/MeOH———h HYDROLYSIS I | 1r TPD Yield = 72% . I l l |DETOSYLOXYLATION I I I H2 —$ HYDROGENOLYSIS : L__..___ _____ J 1,3-PD, 0.72 mole Propanol V TsOH Figure 8-7. Material flow in glycerol dehydroxylation 110 chloride, gaseous hydrogen, as well as glycerol. Benzaldehyde is not consumed in the process because of recycling. Assuming the overall yield of the process is 72%, production of each pound of 1,3-PD will consume 1.68 lb glycerol, 3.61 lb tosyl chloride and 4.12 mole hydrogen, which cost about $1.68, $5.60 and $0.12, respectively. The total raw material cost will be $7.40 per pound of 1,3-PD, of which 76% results from tosyl chloride. The above costs are calculated based on a unit price of $1.00 per pound for glycerol, $1.55 per pound for tosyl chloride, and $0.03 per mole for hydrogen gas. These prices are from the Chemical Market Report, Aceto Corp and AGA Gas, respectively. It is seen from the above that the majority of the raw material cost of 1,3-PD production results from tosyl chloride. The price of this chemical is fairly high on the market, as a result of the small demand and thus small production of this chemical. As a solution to reduce the raw material cost, on—site synthesis of tosyl chloride may be considered. On-site synthesis of tosyl chloride may require a large scale glycerol conversion process to justify it The glycerol conversion processes shown in Figures 8- 3, 8-5 and 8-6 are mostly continuous, and are therefore suitable for large scale production. Another way to improve the economics of the glycerol dehydroxylation process is to recover and sell the r-toluene sulfuric acid (TsOH) produced in this process. As can be seen from Figure 8-7, the glycerol dehydroxylation process produces two by-products: propanol and TsOH. Propanol competes the raw materials with 1,3-PD and is to be minimized from the process; while TsOH is produced at the same time as TsCl is consumed in the process, on the mole to mole basis. Spectrum Chemical MFG has quoted a price of $5.87 per pound on TsOH. Selling of TsOH could basicly compensate the cost of tosyl chloride. Then the net raw material cost for production of 1,3-propanediol from 11] glycerol would be only $1.80 per pound of 1,3-propanediol, which is extremely favorable. PART III VAPOR-PHASE HF SACCHARIF ICAT ION CHAPTER 9 MODELING THE HF ADSORTION PROCESS IN A PACKED-BED REACTOR HF vapor is a potent and selective regent for saccharification of lignocellulose. The vapor-phase HF saccharification consists of two sequential operations: HF adsorption and solvolysis (or simply HF adsorption as it is more commonly called) followed by HF desorption. The current research focuses on the first operation, that is, HF adsorption. Research on vapor-phase HF saccharification can be traced back to the early experiments by German scientists (Fredenhagen and Cadenbach, 1933; Luers, 1938). More recently, this process has been studied by Bentsen (1982), Ostrovski er al. (1984 and 1985), and Franz er al. (1982 and 1987) on pilot-scales. These research was aimed to demonstrate the technical and economic promise of vapor-phase HF saccharification. Because of its empirical and proprietary nature, however, these research provided very limited data for process design and development. The research at Michigan State University (MSU) was started in the late 1980’s, aiming to provide the necessary fundamental chemical engineering data for process development and design. Specifically, Mohring and Hawley (1989) studied the intrinsic adsorption of HF on wood chips. Rorrer er al. (1988) and Rorrer (1989) studied the intrinsic reaction of HF with lignocellulose. Reath (1989) observed the behavior of HF adsorption on wood chips in a bench-scale, packed-bed reactor. These studies have yielded a great deal of useful results, but are not yet complete; as is especially true to the 112 113 research on HF adsorption in a packed-bed reactor. Questions of primary importance to design and Operation of the HF adsorption process in a packed-bed reactor but not yet explored before the current work include: a) under given operating conditions, how long does the adsorption need to be carried out to achieve a certain overall HF loading level? and b) how is the adsorbed HF distributed through the reactor? The current work is a continuation of the research initiated by Hawley, Mohring, Rorrer and Reath, and has focused on modeling the HF adsorption process on wood chips in a packed-bed reactor. A predictive model relating HF loading to the adsorption time and conditions is considered to be the best solution to the two questions raised above. Several breakthroughs have been made in this work. Experimentally, the HF adsorption behavior in the bench-scale, packed-bed reactor has been studied in a broader temperature and HF supply rate ranges than in the previous work, which leads to the finding that the rate of HF adsorption is controlled by two factors: the intrinsic adsorption rate and the HF supply rate. These two rate-limiting mechanisms take control at different phases of the adsorption process and leads to difierent behavior of HF adsorption. Based on this finding, a conceptual model has been developed, which explains all the observed behavior of HF adsorption in the bench-scale, packed-bed reactor. Based on this conceptual model, a mathematical model has also been developed, which can be used to calculate the overall HF loading as a function of the adsorption time. However, the significance of this model is undermined, by its inability to provide information on the HF loading distribution in the reactor. Further work has led to a more advanced model of the HF adsorption process in the packed-bed reactor. This model is developed based on a detailed analysis of the HF flow, 114 the intrinsic HF adsorption and the heat transfer processes occurring in the reactor during HF adsorption. This advanced model provides a great deal of insight into the HF adsorption process in the packed-bed reactor, and can be used to predict the HF loading distribution, as well as the overall HF loading, as a function of the adsorption time and operating conditions. This model has been validated by the experimental data from the bench-scale, packed-bed reactor under study. Based on this work, two papers have been written and been published in Chemical Engineering Science (Wang et al., 1994; Wang et al., 1995a). These papers basically document all the work that has been done in the current research on HF adsorption. To avoid rewriting, only abstracts of these papers are included in this dissertation. The first paper is entitled “Adsorption model of HF on Wood Chips in a Bench- Scale, Packed-Bed Reactor”, representing the experimental work and the work on the more primitive lumped HF adsorption model. the abstract of this paper is quoted below: Abstract: The adsorption process of HF on wood chips in a packed-bed reactor was studied. 7W0 adsorption rate limiting mechanisms exist for this process, namely HF supply control and intrinsic adsorption control. In the initial period of this process, the adsorption rate is determined by the HF supply rate; afler the initial period, the adsorption rate is determined by the intrinsic adsorption rate. Consequently, the HF loading profile for this adsorption process includes two distinct parts: an initial linear part and following non-linear part, corresponding to the two difi’erent rate limiting mechanism. This conceptual model of HF adsorption in the packed-bed reactor was supported by experimental observations from a bench-scale. packed-bed reactor: Based on this conceptual model, a mathematical model was developed to predict the HF loading for a given adsorption time under given operating conditions. To demonstrate its application, the mathematical model was used to analyze the efect of the heat dissipation rate of the reactor on the adsorption process. The second paper is entitled “Modeling the HF Adsorption Process on Wood Chips in a Packed-Bed Reactor”, representing the work on the more advanced detailed HF adsorption model. The abstract of this paper is also quoted below: 115 Abstract: In this work, a detailed model, along with the numerical procedure to solve this model, was developed for the HF adsorption process on wood chips in a packed- bed reactor: This model is aimed to predict the HF loading distribution, as well as the overall HF loading, with respect to the adsorption time. However; the temperature field in the HF reactor may also be obtained using this model, as a function of time. To validate the developed model, the computed overall HF loading profiles were compared with the experimental data acquired previously from a bench-scale, packed-bed reactor filled with big-tooth aspen wood chips, and a good agreement was obtained. By taking advantage of the newly developed adsorption model, the HF loading distribution and temperature field in the bench-scale, packed-bed reactor was examined It was found that a uniform HF loading distribution and temperature field is always obtained at the end of the adsorption, provided that the adsorption is carried out sufi'iciently long. This finding is significant, since a uniform HF loading distribution is always desired in operation of the HF adsorption process. The above modeling work carried out in the current research provides a theoretical foundation to design and control of the HF adsorption process in the packed-bed reactor. CHAPTER 10 SUMMARY AND CONCLUSIONS In the current research. three biomass conversion processes have been investigated, namely sugar hydrogenolysis, glycerol dehydroxylation and HF saccharification, all aiming to produce value-added, useful chemicals from the biomass resources. Results from these investigations are summarized as follows. 10.1 Sugar Hydrogenolysis The research on the sugar hydrogenolysis process has focused on understanding of the reaction mechanism and improvement of the reaction selectivity. In the mechanism study, a series of 1,3-diol model compounds have been hydrogenolyzed, to establish the mechanism of C-C and CO bond cleavage in sugar hydrogenolysis. The experimental work confirmed our theoretical conception that the C-C bond breaks via the retro-aldol reaction of a B-hydroxyl carbonyl precusor and the CO bond breaks via the dehydration of the same carbonyl precusor. Establishment of the above mechanism makes it possible for one to adopt a more rational approach than the current practice in control of the selectivity of sugar hydrogenolysis. The selectivity study in the current research is also canied out with a 1,3-diol model compound, namely 2,4-pentanediol(2,4-PD). Focus of this selectivity study has been placed on understanding of the factors controlling the CC v.s C-O selectivity. In this 116 117 capacity, the effects of temperature, base concentration, hydrogen pressure, catalyst amount and catalyst type on the selectivity are experimentally investigated. Additionally, a mathematical model is also developed and the 2,4-PD hydrogenolysis process is simulated with the developed model, to rationalize the experimental results and to look into ways to improve the selectivity. It is found out that the temperature has little effect on the selectivity; increasing the base concentration can increase the selectivity to certain extent; increase in the amount of catalysts used has an adverse effect on the selectivity; and the selectivity varies widely with the type of catalysts used. The effect of hydrogen pressure on the selectivity is the most interesting. At low pressure, the selectivity increases with the pressure; at medium pressure, the selectivity is not sensitive to the change in the pressure; while at high pressure, the selectivity decreases as the pressure is increased. 0f the seven different types of catalysts that has been studied in our experiments, the best one is barium—promoted copper chromite, which has furnished a C- C vs. 00 selectivity of 2.5 in hydrogenolysis of 2,4—PD. Our simulation study additionally points out that, to improved the selectivity, a catalyst with a reasonable capability to hydrogenate ketones and aldehydes but only moderate ability to hydrogenate cab-unsaturated ketones should be sought. 10.2 Glycerol Dehydroxylation The purpose of glycerol dehydroxylation is to convert glycerol to 1,3-propanediol, a very-high-valued chemical. In this research, a new concept has been presented to selectively transform the second hydroxyl group of glycerol into a tosyloxy group with the help of a group protection technique, and then to remove it by hydrogenolysis. To verify 118 the technical feasibility of this new concept, experiments are carried out in this work to investigate each of the steps involved in the new conversion concept. The results are promising: an overall yield of 72% or higher is found to be achievable with this new conversion concept. Another major accomplishment of this research is that, for the first time, tosylates are hydrogenolyzed catalytically with molecular hydrogen. The standard method to hydrogenolyze tosylates at the current time is to use base metal hydrides, such as lithium aluminum hydride (LiAlH4) and lithium triethylborohydride (Lil-IBEt3). It is discovered in the current research that the tosyloxy group of a tosylate, when assisted by a hydroxy group next to it, can be removed with molecular hydrogen, in the presence of a nickel or ruthenium catalyst. This discovery is vital to the new concept of glycerol conversion. It makes it possible to detosyloxylate TPD economically with gaseous hydrogen. Otherwise, hydrogenolysis of TPD has to be canied out with lithium aluminum hydride or lithium triethylborohydride. Both are too expensive to use at industrial scale. 10.3 HF Saccharification The work on HF saccharification has focused on modeling of the HF adsorption process in a packed-bed reactor. No models have been developed. The first one is phenomenological in nature, and is developed based on experimental observations of the HF adsorption behavior in a bench-scale, packed-bed reactor. The second model is more advanced than the first one. It is developed based on a detailed analysis of the HF flow, the intrinsic HF adsorption and the heat transfer processes occurring in the reactor during 119 HF adsorption. This model provides a great deal of insight into the HF adsorption process in a packed-bed reactor, and it can be used to predict the HF loading distribution, as well as the overall HF loading, as functions of the adsorption time and operating conditions. The above modeling work accomplished in this research provides a theoretical foundation to design and operation of the HF adsorption process in a packed-bed reactor. CHAPTER 11 RECOMMENDATIONS ON FUTURE WORK 11.1 Future Work On Sugar Hydrogenolysis Future work on sugar hydrogenolysis should continue focusing on control of the C- O bond cleavage. The selectivity study may continue being carried out with the 2,4- pentanediol model (2,4—PD) compound (ideally 2,3,4-pentanetriol though, because of its closer analogy to sugars). Analysis of both the products and data are simple with this model compound, and the study may also be assisted by computer simulation, since a process model is now available for this purpose and has proven to be useful. In our simulation study, the process model has predicted some phenomena that have not been observed in the experiment, possibly because the limited range of reaction conditions within which our experiments have been conducted. Included in these predictions are that the selectivity will become independent of the base concentration when the concentration gets too high, and that the selectivity will decrease with hydrogen pressure when the pressure gets too high. In further work, experiments may be performed in a larger base concentration and hydrogen pressure ranges to verify the above predictions. Changing the catalyst is the most efiective way to change the selectivity. The simulation study suggests that catalysts with strong ability to hydrogenate saturated ketones and aldehydes but only moderate ability to hydrogenate or,0unsaturated carbonyls 120 121 have potential to deliver high selectivity. Some homogeneous rhodium catalysts with the above properties have been reported by Mesuoni er al. (1978). These catalysts may be tested in 2,4-PD hydrogenolysis experiments, to see if they can provide improved selectivity. The most promising catalyst found so far in our experiments is barium-promoted copper chromite. It provides a much better selectivity than other catalysts tested, including the plain copper chromite catalyst. These results clearly indicate that the selectivity of copper chromite is enhanced by the presence of barium. Therefore, future work may also be canied out to study the effect of addition of barium, as well as other elements, in preparation of the catalyst on the selectivity of 2,4-PD hydrogenolysis. Finally and most importantly, the results from studying the model compound 2,4- propanediol (2,4-PD) should be applied to sugars in the future, so as to improve the selectivity of sugar hydrogenolysis. The optimal conditions for sugar hydrogenolysis may be different from those for 2,4-PD hydrogenolysis. However, the effects of operating conditions on the former process should parallel to those on the later. This is yet to be verified in the future research. 11.2 Future Work on Glycerol Dehydroxylation Future work on glycerol dehydroxylation has two immediate goals: a) to modify the procedure of tosylation, so as to facilitate the recycling of solvent; and b) to optimize the detosyloxylation condition, so as to increase the yield of 1,3-PD. The standard lab procedure used in this research furnishes high PTD yield. However, introduction of ice following the reaction makes it difficult to recycle the 122 solvent, as the solvent has to be separated from water and thoroughly dried before it can be retumed back to the reactor. Solvent recycling is extremely important at industrial scale. A better process would allow both the reaction and crystallization to be carried out in the same solvent, so that the solvent can be directly recycled as in Figure 8-5. Therefore, future work should check the feasibility to directly crystallize PTD from pyridine. If direct crystallization of PTD from pyridine proves to be difficult, a co-solvent may be found, to help the precipitation of PTD from pyridine. If a co-solvent is needed to facilitate the crystallization, it ought to be verified that the tosylation can proceed in a mixture Of this co-solvent with pyridine as smoothly as in pyridine alone. In both experiments, it may become necessary to proceed the crystallization with a pre- evaporation step to remove a portion of the solvent from the PTD solution. As discussed in Chapter 8, the overall yield of the glycerol dehydroxylation process is determined by the yield of the detosyloxylation step. Although a 1,3-PD yield as high as 72% has been obtained in the current work, it is not optimized against the reaction conditions. Therefore, future work is needed to study the effect of reaction conditions, including temperature, medium acidity, and catalyst, on the product yield. The 1,3-PD yield of the detosyloxylation step is mainly affected by the formation of by-product propanol in TPD hydrogenolysis. Propanol competes with 1,3-PD for starting TPD. The ratio of 1,3-PD to propanol produced in this reaction is determined by the rate of 3—hydropropionaldehyde hydrogenation relative to the 3-hydropropionaldehyde dehydration (Figure 8-2). Any factors affecting this rate has potential to be optimized to improve the yield of 1,3-PD. As a general rule, hydrogenation has a higher activation energy than dehydration. 123 improve the yield of 1,3—PD. As a general rule, hydrogenation has a higher activation energy than dehydration. Therefore, increase in temperature should increase the rate of hydrogenation relative to dehydration, and thus increase the yield of 1,3-PD in TPD hydrogenolysis. In the current work, TPD hydrogenolysis is canied out at a relatively low temperature (140 °C), due to the concern that the 1,3—PD formed in this reaction might be further hydrogenolyzed. In retrospection, this concern may not be well founded, as our experiments on 1,3-diol hydrogenolysis has shown that hydrogenolysis of 1,3-PD is slow even at 210 °C. Hence, the effect of temperature on TPD hydrogenolysis should be investigated in future research, so that the reaction can be carried out at an optimal temperature to maximize the yield of 1,3-PD. In addtion to 1,3-PD and n-propanol, TPD hydrogenolysis also generates p-toluene sulfonic acid (TsOH). In the current work, the TsOH generated during TPD hydrogenolysis is not neutralized. Since dehydration of HPA is catalyzed by acids, neutralization of the reaction medium during the reaction may be helpful to control the formation of by-product n-propanol. This measure to improve 1,3-PD yield should also be investigated in the future research. Future research should also investigate the effect of catalyst on TPD hydrogenolysis. A good catalyst enhances the hydrogenation of 3-hydropropionaldehyde, and thus may improve the yield of 1,3-PD. In addition to the 1,3-PD yield, the possibility of catalyst poisoning by the sulfur containing compounds in TPD hydrogenolysis also needs to be addressed. Although it has been found in our experiments that the Ni catalyst used in TPD hydrogenolysis is still 124 was stopped only after the catalyst completed a turnover number of about 50, to demonstrate the catalytic nature of the reaction. Further work should furnish, and improve if necessary, the overall turnover number that a catalyst can achieve in its lifetime. Theoretically, mesyl chloride may work just as well as tosyl chloride in the new concept of glycerol conversion. Future work may look into the feasibility of using this alternative compound. Because of its smaller molecular weight, mesyl chloride is about 40% cheaper than tosyl chloride per mole, although the prices of these two compounds are similar on per pound basis. This prices difference provides a material cost advantage. 11.3 Future Work on HF Saccharification Focus of the future work on HF saccharification should shift from the HF adsorption process to the HF desorption process. Due to the low price of sugars and the relatively high price of hydrogen fluoride, recovery and reuse of this regent is crucial to the economical viability of the HF saccharification process. Although not reported here, some work has been done on HF desorption in the current research. It was found that, under some conditions, the desorption process was extremely slow, and it took more than 30 hours for the process to complete. Under these conditions, the HF residue level was normally observed to be high and it ranged from 5 to 10% of the HF loaded onto the wood chip substrate. However, some desorption runs were also completed in a time less than one hour, with an HF residue level less than 0.5%. Further work is needed to investigate the reason causing the above difference. It may possibly be linked to the temperature history experienced by the substrate in the HF adsorption process. APPENDIXES APPENDIX A COMPUTER PROGRAMS F OR SIMULATION STUDY APPENDIX A COMPUTER PROGRAMS FOR SIMULATION STUDY This appendix documents the computer programs used to solve the process model of 2,4-PD hydrogenolysis. These programs are wrinen in Matlab language, and must be run in Matlab environment. The function of each program in solving the problem are explained in the comments included in the program. Main Program: % Program SIMUM (Script M file) % Version 4.0, June 1, 1997 % This program is used to simulate the hydrogenolysis reaction of 2,4-pentanediol. % It solves the differential equation system as defined in function M file CPRIMEM, % and displays and saves the results. % t =-- time (not used); c --= concentrations; y = rates of concentration change % Units: time = min; concentration = mol/l; flux and rate = monin % Table of reactant and products: % 1. A, 2,4»pentanediol % 2. B, 4-hydroxyl-2-pentanone % 3. C, acetone % 4. D, 2—propanol % 5.13, acetaldehyde % 6. F, ethanol % 7. G, 3-pentene-2-one % 8.1, 2-pentanone % 9. J, 2-pentanol % 10. M, metal catalyst % 11. MH, metal hydride % 12. OH, hydroxide ion % Additional species: H, molecular hydrogen; H20, water. 125 126 %%% Main Body of Program % Load Parameter (Including initial conditions) clear; clear global %clean work space para %sub-program containing rate constants and initial conditions % Initial Concentrations: c0 = [A, B, C, D, E, F, G, I, J, M, MH, OH]’; % A, B provided by Program PARAM % Reaction time ti = 0; tf = 210; % tf-ti == duration of reaction % Solve the Differential Equation System [t, c] = ode45(‘cprime’, ti, tf, c0, 1e-6, l); % Save the Solution in file: rsl2\sol and “Reaction Conditions” in rsl\condi, if n=2. l=1ength(t); sol = [301; [c(1:10:l,:),t(1:10:l,:)]]; dir = [‘rsl’ int23tr(n)]; eval([‘ !mkdir ‘ dir]) eval([‘!copy para.m ‘ dir ‘/condi’]) eval([‘save ‘ dir ‘Isol sol -ascii’]) % Display Results % Data preparation A=sol(:,l); B=sol(:,2); C=sol(:,3); D=sol(:,4); B=sol(:,5); F=sol(:,6); G=sol(:,7); I=sol(:,8); I=sol(:,9); M=sol(:,10); MH=sol(:,11); OH=sol(:,12); t=sol(:,l3); l = 1ength(t); % Main products figure(l) plot(t,A*1000, t,C*100 t,D*1000,’-.’, t,l*1000,’:’, t,J*1000) axis([0 250 0 120]) title(‘Major Products’) xlabel(‘T1me (min)’) ylabel(‘[Al. [C]. [D]. [I]. [1] (mM)’) grid % Minor products figure(2) plot(t,B*1000, LE*1000,’--’, LG*1000,’-.’ axis([0 250 0 5]); title(‘Minor Products’) xlabel(‘T1rne (min)’) ylabel(‘[B]. [E]. [(3] (mM)’) grid 127 % Selectivity figure(3) S = (C+D) J (G+I+J); P10t(t. S. ‘-’) axis([0 250 0 5]) title(‘Selectivity’) xlabel(‘Trme (min)’) ylabeK‘ s = ( [C]+[Dl )/ ( [G]+[I]+[Jl )’) grid %Base concentration figure(4) plot(t, OH*1000) title(‘Base Concentration’) xlabel(‘Trme (min)’) ylabel(‘[OH] (mM)’) grid eval([ ‘print -deps ‘ dir ‘/base.eps’]) % save the figure %%% End End End End End Sub-Program #1: % Program PARAM (Scrip M file) % Version 4.0, June 1, 1997 % This program is used to simulate the hydrogenolysis reaction of 2,4-pentanediol. % It defines the rate constants and initial conditions % t == time (not used); c = concentrations; y --= rates of concentration change % Units: time = min; concentration = mol/l; flux and rate = monin % Table of reactant and products: % 1. A, 2,4-pentanediol % 2. B, 4-hydroxyl-2-pentanone % 3. C, acetone % 4. D, 2-propanol % 5. E, acetaldehyde % 6. F, ethanol % 7. G, 3-pentene-2-one % 8.1, 2-pentanone % 9. J, 2-pentanol % 10. M, metal catalyst % 11. MH, metal hydride % 12. OH, hydroxide ion % Additional species: H, molecular hydrogen; H20, water. 128 % Kinetic Parameters klf: 435; klb = 67.4; k2f = 752; k2b = 4480; k3f= 1.89; k3b = 12.2; k4f = 11.9; k4b = 1.24; k5f: 14100; 161) = 6590; k6f = 22.1; k7f: 1.16; k7b = 7.48; k8f = 76.2; k8b = 3.32e-5; % Reaction Conditions: % Water Concentration H20 = 46.5; % M (not used) % Hydrogen pressure (base: 5 MPa) H = 0.1; % M, expressed as the concentration of molecular hydrogen % dissolved in the reaction solution. % Initial Base Concentration (0.4 m1 l N NaOH) 0H0 = 0.010; % M % Initial catalyst concentration (base: 0.1 g Ni/K) M0 = 0.028; % M % Initial 2,4-Pentanediol Concentration (base 0.5 g diol in 40 ml water) A0 = 0.12; % M % Pass data to CPRIMEM global klf klb k2f k2b k3f k3b k4f k4b global k5f k5b k6f k7f k7b k8f k8b global H20 H 0H0 M0 % Conditions to start “STARTM” A=A0; B=0; C=0; D=0; E=0; F=0; G=0; I=0; J=0; OH=OH0; M=M0; MH=0; % Define the number of the directory to save the results n = l; %%% End End End End End 129 Sub-Program: function y = cprime(t, c) % Program CPRIMEM (function M file) % Version 4.0, June 1, 1997 % This program defines the problem of 2.4-pentanediol hydrogenolysis % It is to be used with program SIMU. % t =2 time (not used); c == concentrations; y = rates of concentration change % X’s == reaction fluxes % Units: time = min; concentration = mol/l; flux and rate = monin % Table of reactant and products: % 1. A, 2,4-pentanediol % 2. B, 4-hydroxyl-2-pentanone % 3. C, acetone % 4. D, 2-propanol % 5. E, acetaldehyde % 6. F, ethanol % 7. G, 3-pentene-2-one % 8. I, 2-pentanone % 9. J, 2-pentanol % 10. M, metal catalyst % 11. MH, metal hydride % 12. OH, hydroxide ion % Additional species: H, molecular hydrogen; H20, water. %Get Data from PARA.M global klf klb k2f k2b k3f k3b k4f k4b\ global k5f k5b k6f k7f k7b k8f k8b global H H20 % Flux Expressions X1 = klf*c(l)*c(10)*c(12) - k1b*c(2)*c(ll); X2 = k2f*c(2)*c(12) - k2b*c(3)*c(5)*C(12); X3 = k3f*c(3)*c(11) - k3b*c(4)*c(10)*c(12); X4 = k4f*c(5)*c(1 1) - k4b*c(5)*c(10)*c(12); X5 = k5f*c(2)*c(12) - k5b*c(7)*c(12); X6 = k6f*c(7)*c(l 1); X7 = k7f*c(8)*c(1 l) - k7b*c(9)*c(10)*c(12); X8 = k8f*H*c(10)*c(12) - k8b*c(l 1); 130 % Differential Equations y(1) = -Xl; % A y(2)=X1-X2-X5; %B y(3) = X2 - X3; % C y(4) = X3; % D y(5) = X2 - X4; % E y(6) = X4; % F y(7)=X5-X6; %G y(8) = X6 - X7; % I y(9) = X7; % J y(10)=-Xl +X3+X4+X6+X7-X8; % M you = -y(10); %MH y(12>= y(10); %OH % End End End End End APPENDIX B RE T RO-ALDOL AND DEHYDRA T ION RA TE CONSTANT DERI VA T I ON PROCEDURE APPENDIX B RETRO-ALDOL AND DEHYDRATION RATE CONSTANT DERIVATION PROCEDURE This appendix documents the procedure used to derive the rate constants for the retro-aldol and dehydration reactions of 4-hydroxy-2-pentanone, or the Reactions No.2 and No.5 in Figure 5-5. The basis of this derivation is the kinetic data obtained by Jenson and Hassan (1976) on the hydration of 3-penten-2-one. In Jenson and Carre’s experiments, 3-penten-2-one was allowed to undergo the following reactions at 40 and 50 °C in 0.1 N NaOH aqueous solution: 0 OH 0 II km II km CH3CH=CHCCH3 kt: CH3CHCH2CCH3 —> CH3CHO + CH2COCH3 dehy 3-penten-2-one 4-hydroxyl-2-pentanone acetaldehyde and acetone The hydration rate constant khyd was determined at 40 and 50 °C, and was presented along with an observed rate constant hobs. This observed rate constant is related to khyd, kdehy and kmm as follows: kh ydk retro kdehyd + kretro k (D-l) abs = Thus, if the reaction equilibrium constant for the hydration/dehydration reaction pair, or the ratio of khyd to kdeh), is obtained, the dehydration rate constant kdehy may be calculated 131 Tr) '.. . ‘r 132 from khyd, and the retro-aldol rate constant kmm can then be solved from Eq.(D-l), which is exactly what has been in this work. The relevant data and results are all given in Table D-l. The equilibrium constants in Table D-l are from Jenson and Carre (1974), and are measured in 2.57 M HClO4 solution. Table D-l Rate Constants Directly from Jenson and Hassan (1976) a 4b 4b d 4d 40 °C .50 9.5 3.14 3.03 0.168 50 °C 1.1 21 2.67 7.87 0.435 a. unit for all rate constants: s"; b. from J enson and Hassan (1976); c. from Jenson and Carre (1974); d calculated in this work. The rate constants listed in Table D-l are the products of those in Figure 5-3 for the same reactions with the concentration of hydroxide ion in the solution. As the hydroxide ion concentration at which the above rate constants are obtained are known (0.1 M), the rate constants k5, k_5 and k2 in Figure 5-3 can be determined from the rate constants in Table D2, and then from these rate constants at low temperatures, the rate constants for 220 °C can be extrapolated with the help of the Arrhenius equation. To obtain the rate constant k,2, the equilibrium constant for the retro-aldol and aldol reaction pair is needed. In the current work, it is estimated from the free energy data at 25 “C (again data at other temperature are unavailable) in' Cuthrie (1977) to be 0.168. With this equilibrium constant, k_2 is calculated, and is documented in Table D-2, together with other determined rate constants. 133 Table D-2 Determined Rate Constants k5 (M‘lmin'l) k_5 (M'lmin'l) k2 (M'lmin'l) k_2 (M'zmin'l) 40 °C 0.182 0.570 0.0101 50 °C 0.472 1.260 0.0261 220 °C 14100 6590 752 4480 rd APPENDIX C EXPERIMENTAL DATA OF 2,4-PD H YDR OGEN OL YSI S APPENDIX C EXPERIMENTAL DATA OF 2,4-PD HYDROGENOLYSIS This appendix documents the experimental data of 2,4-pentanediol (2,4—PD) hydrogenolysis, collected in the selectivity of sugar hydrogenolysis. In this study, 2,4-PD is used as a model compound to facilitate the rersearch. 134 135 .28 fl :5. .8 .58 a :5. d .23. mm x8. 3:855 .wcuaunea .3? 829% Ewan»: .e Sam eqmm owdm w}: 26 3d n2 an?“ 8.8 cwmm 33 has 3N ca mvsm amém 3.3 S .w 5.x SN m S R. _ v 302 5.? mad mad 2 .n 3 3.3 n3: 8.? 36 NYE 2 .N mm 3.2 "E: wm.wc 2 .v and— wm.~ ow Exec m~.~ mmdm mm; 3.2 RN mm 3.2 _ cod cod cod cad cod o nomad u_o§=ua-~ uo=o=8=oa-~ Locaaoaofl 8888a 8655... aoEu GE o3 ~89: nag—om x255 20.2 7: “3.3 ”AM We Haw—80E no NZ $3 umbSaU ”um—2 Qm ”2385 ”0. CNN ”damp. so; “Sarah. 136 .28 8 .8: .9 .88 8 .8: d 888 mm 83 35355 $828.38 888 3:83 comp—PA: d 2 .me 5.? ~02 wvdfi min 54 non me. g m mm.wm omfi fl 3.: mm.m mad wS 8.? 3.5m 3.3 on.“ H E .v wad mm _ when 50:“ 2.3 $6 3.1 cod m: no.3 5.8 2.3 omd vnfi and mm 5mm 8.3 3.3 «2. $3 $6 we 3.3 3.5 mmfim cod Rue who mm 3.3 $6 $42 mnfi has Ed mm 3%: m: SH and nmé and mm mm; 2 mnd m3. cod 5% Ed 3 3.3 _ cod Rd cod cad cod m— Nodfi cod cod cod cod cod o 993% u_o:8=og-~ uu=o88=8-m 9.28389: $8388 93.859 8988 CE o3 383 nag—om x285 2082 7: 63m Am 08 Haw—83M 8o :4 $3 sax—«EU :52 ed “8385 ”0. CNN 288. a? 8 “55:28am 137 .28 8 :8. .u .88 8 :8: d .888 mm Moe wcuaosam $5358 .88 8:88 mowed»: d gum on.mm 3.2 «Nam m 2 .w E .2 new amfim Sdm Rd _ mcdm 3.3 3.: m2 8.8 cm. _ m 5.2 3 .3. $4 2 3.: m3 3.? Emma S .n _ Eda 3.3 avd m _ _ S .3” 3.3 Swan dwém mm: 3.2 mm 3.8 om.m~ Eda afimm 8.: 3.: ms 02h 3.3 3 .mm owdm and“ 3.2 mm 8.8 mmd 3.2 no.2 3.3 3.2 9. $2. Eum 3.: m2. 3 .m H 8.2 mm and: Ed a: ddd 2mg Rd 2 Vddfi odd ddd odd odd odd d $5-1m u_o:8=ua-~ uo=o=S=oa-~ 9682888“ 92.98“ 9.2850 cos GE GE .883 353cm 55 v.8 1082 7: Human ”Aw mdv Haw—83v— :o MZ $3 nah—8.5 "am—2 cw ”8:395 uU. cNN ”dz—uh. and. “5883.8 138 .28 8 8:: .9 .88 8 88. .8 J88 mm goo. usuaunem 888288 88.8 888% sowed»: .8 90mm 3.3 mmd 2.5m 3.3 nmdm n: dcsm «v.3 mmd when 3.: 2.5 m3 and... 8.3 end S .Vm 3 .2 03h m: mw.nn 8.5 :6 3.3 3.2 Nvdm mm omdv 3.2 and dodm 3.3 3.3 mm awn 3.: d3: 3:: 5.2 dmdm dc MONA: mad 3 .v 36 mod on.» mm 8.2 8 dmd cvd dad Rd odd 2 dcdfl odd odd odd odd odd d unmim u_o§=3-~ mocoaficoa-~ 93838808 92.888 8885» ao8u :5 ox .22., 85:8 ”cs we moaz z_ 58m ”a «.9 2:89.08 .5 2 $8 ”.886 ”£2 3 ”2:828 ”u. an ”.888. 3?; 80888me 139 .28 8 28: .u .88 8 228. .8 .888 mm “282 mega—28m 382822222 83.8 232238 53.8.2: .3 3.3 3.3 w 2 d mwdw cad 2md¢~ new 3.3 d: m end 233 3% 3.3 m: 02 . 2 m 3dm vmd 3.? 36 3.3 on 2 3.3 dNSN New 2%. 22V and 02 .mm m 2 2 $.22V mmdm mm.n 8.9. 2mm 23% mm 8.3. 3.22 36 no.3 and 23mm dd 3.? 3.3 3.x 3.6m mm.» cmdm mm «0.3 3.3 3.2 2 3.3 3.3 3.3 9. mod» 2m.m and no.3 mad 3.3 mm 3622 3d 3.2 mmd 2N.2 cmd m2 dw.3 2 odd ddd ddd odd odd d mania U228822.~ uu=o=8:3-~ 32.53.38 Q8288 9298223 80822 CE o3 $83 H253cm ”CE v.9 2.2082 22 ”8mm 22w 2.8 823236222 28 22 33 ”2822380 3&2 ed ”833.5 uU. CNN 22280.2. 35-2% 258282282 140 .28 fl EB .u .58 fl is d .mEE mm “82 wage—BE .3322.“ San commas Swan»: .3 3.3 v0; Zu— wmwv 8N No.3 new oodm soda em; afiwm :4 mmém new mode Exam hm; 3.9. 8; 86m c3 Kdn 3.8 _m._ 3.5 mag no.3 5: 32b 3.8 $6 5.3 EA 8.0m n g _ 8.: 3.2 ma; 2 .mm avd 5.3 E. 5.3 3.2 S .N ma: Ed 8.: mm 3.3 26 o; «E: :3 RA: ov 36: em; cm; 2: mm; aofi mm 8.3 _ Ed Rd cod wvd cod 3 862 cod cod cod ocd cod o gaméd u_o:5:oa-~ uu=o=8=o9m Locum—83$ 90:88“ L235". ace: c3 as has: 352% was «.8 322 z_ ”3.8 ”a 2: 23383. .5 _2 $3 “5:38 ”£2 ow ”same; no. as as”... am: “58.23%. 141 .28 fl :5. .9 .58 fi :5. d .22: 3 good $5855 .wcaaonoa .3? 3:9? unwed»: d wmdc 3.3 mi Eda cud 3.5 nmv 2w: wmdfi end 2 .3 3d omdm mum and» 8.2 and vmdm and 3.2 n 3 8.3 ms; Rd 3.3 and 3.2 nmm 2 .3 3.2 2 d 2.2 and 33; m3 whmm No; _ odd 2.4: cod Nwfl o2 mafia dmdfi odd 3.2 cod 3.: mi 8.3 and 3d 3.3 cod mg: no— wodd mod 2; :2: and 3d ow 862 Re o: and 3: 2d 3 34 2 mod :4 :4” 2 4 Sum om mod: and v: dmd mwd cod 2 add: 3 d and ovd de odd 2 3d: odd odd cod odd 8d d ooméd o_o=3=od-m uoco:3:od-m u_o§qu8_ $888 $2350 cue: GE GE 333 ”EQZOm ”CE v.8 IOaZ 7: Hamwm Km 2: Haw—030— cc :4 $3 umbfiau 3&2 o.m ”2:305 ”U. of ”9:8. 3-; ~585qu 142 .28 fl as .u .58 fl is d £58 mm x03 wcuaonpfl $.53:an 5% 3:3“ Swen»: .3 03% 8.3 3.2 Eév «v.3 wfimm wt mafia 3.3 3.3 3.; on: owém mi omém 5.3 8.2 8.3. mw.3 8.5 m: m fiom 3.8 3.9 afiwm an: «Ema ca emdm whmm mg; 2.9. %w— mhAm ms cufim S._~ mm: 3.2 «WE mmdm mm 93m 8.2 2.2 8.8 3.: owda 9 mode 3.x mm: mmé 3.2 cmdm an and: Rd S.— :.d 24 and 2 Sdfi cod cod cod cod 86 c “Vania o_c§=oq-~ cococficua-~ ufigqoaofl 92588 $2350 aoau CE 9: 883 sag—om 5.: 18:92 7: Human Aw 2: Haw—3.2M .8 _z ammo ambaso 3&2 ed ”2835 ”U 3N “95h. $7; EoEtaxm 143 .28 fl ES .u .58 fl is d .32: cm moo. mango—:5 .wcuaonoa “3% 3:3“ Swen»: d 8.3 wwdm EN mm. 3. ~ fin 36m com no.3. mmdm $3 $3? mm.m 3.3 at mngn Séu v3 swam 3d Emma 9: 03% mndu oo.~ Sam cad 3.3 E _ mode cadu Rd Sfim 2 .m mnfim ow mvdc 3.2 cm.~ mnfia Dam whmm me $65 33; cad Eda 3am ondm mm 3.3 Q .5 Se Sufi 34. 3.3 mm cad 2 cod an; RA 3 .N an; em mas: we; end cod we.“ ocd S 8.2 H 86 cod cod cod 86 o nan—{N u_c§:oa-m uo=o§=8-~ UBSEEQSM 02553 $935“. $55 as g has; €28 ”as : :92 z_ 53m x» 2: 2:388. .5 2 $8 "£38 ”£2 ow ”2:805 U. 8N as“; a. _- a Esteem .28 a :5. .9 .58 fl ES d .82: on moo. $552.2“— .wcuaonoa 3% 3:95 comet»: d 144 3.8 5.2 vw.m wad £1 and cum omda 2.2 36 awfi an; and new 3.2: no.2 $.N «he an; Rd can 3M2 c2. mvd Sum mm; 3% 93 $62 2 .o chm Nod EA and o: 8.2: Rd E .v Rd 8; Rd cm 3.2: mud SM 3; E..— 2: on vafi _ _ 93 Ed mad 3; cod on 3.2 ~ cod cod cod cod cod 9 993$ u_o§=oa-~ uu=o=8=u9~ u_o§aoa8_ 02588 98550 no.5 GE o3 833 sag—om x285 :92 7: Human xw fie Haw—8&2 .8 :4 $3 sax—Smu ”am—2 Qm ”8835 ”U. com ”dank. as i .Sauaxm 14S .28 fl :5 .9 .58 a :5. .a .mEE ON x02 35855 $532.2.“ 8% coins coma—i: d 3.2. >03 and nndv nvd sown com 3% Rd _ $6 no.5 34 u hnfim 9: news 8.: win mg: mod «55 a: 3.8 v0.3 $6 czm 3.2 3.8 on canw can and no.3 and 3.2 on Sn: and own 2: Sun bng on 3.2 _ cod cad cod cod cod o can—in Leas—San mocccficoan 9352.883 92588 9855» £25 :5 8. as; 3828 was we :92 z_ ”88 ”a 2: .5383. .5 _z *3 ”£36 an: 3 ”2:32“. ”o. 8N ”deep an _ - a sagas 146 .28 fl :5. .9 .EE 3 on: .o .mEE on 969 managea— .w=uau:2o 8cm common 538»: .99 no. 3 n omfim 3.3 Son 3 .3 oo.o won oofin omnm no.3 _ Eunn no.3 3.3 w3 nofin no.3 vm.§ “Nan and 3 3.3 m3 3 3.3 M: .3 3.3 onfim 99.3 3.3 mm 3.3. no.3 onfi 8.3 8.3 o_ .3 on wmnw nod and 39.: v~.o3 3.: nv 8.83 nm.~ ban on.~ 36 con mm oo.oQ oo.o oo.o oo.o oo.o oo.o o 90min 9_o=3coo-~ 99:o=8=9o-~ 9.059398% 92.289 9355» one: o5 as .32, 95:8 nos «.8 :02 E ”88 x9 8.8 2:33.92 8 _z .98 9:36 am: 2 ”93329 6. SN ago... 9: i acacaxm 147 .28 mm :5. .u .58 fl ES d v.58 om mos 393:8 .3585...“ Eda BEE“ come—i: d 3.3 mad SN and «2. 3.3 ova $63 ~o.m and $6 3.0 3.3 o3 3 .2: and Q3 mfiv and 8H ofi 8.2: $3 23 2d 9.6 $6 ma :3 2 8.3 me; 8.— omé mmfi an 8.33 and K5 mad 23 wad cm 863 ood oo.o ocd oo.o cod o chain u_o:8=uq-~ coco:S=8-~ 985238.383 gaseous 93285“. goat :2 as .2“; 352cm 5... Ne :92 z_ ”88 x39 2:39.30. 8 2 $8 ”236 E2 2 ”game“. G. en ”deep 2 i aganaxm 148 .28 3 33:: .9 .58 m3 33:: .3 .2338 on 383 mango—32m .wcuaoaoa 38% 36333393 soap—gm d nib 3.3 and 3.5.3 mod 3 3 .wm ovm no.3 ad cad mad 3 3.3 3.3 333 5.3 cad 22. m 3 .3 oo.~ $63 83 3.83 mud $6 86 v03 .33 3 cm @353 8N Dav 3.3 .m 3.3 .3 med co 336m 33 8.3 23.3 35.3 333 33.33 cm 8.83 oo.o ocd oo.o ocd oo.o o gamim U32:33:23.~ coccaficoa-m u3o§n3oacm3 $588.“ U32:23:. cop—33 CE 93 38“? ”352cm ”CE adv 3.303373 7: 63m ”3w 33.8 333338903 35 373 $3 ”3.333330 ”#333 Wm ”2:805 HU. SN 2380.3. as - :3 2583.86 149 .23: .33 3335 .9 .3338 m3 33:: d .2338 cm 3302 3333.35 $53355 38.3 35339333 Swear .a com 2.3 $3 mods 3 3 .m we. 3 w 33mm 3.3M .33 3 ow.~3 «3.? mc.m 330.3 3.3 3.3 an 5.3 3w 8.2 3w mad mmd on 3N: 3.03 3.3 8.3 3.3 033.3 cm 3 .om3 oo.o oo.o oo.o oo.o oo.o o %EAxw 9323859~ coco—38:3-N Legacaofl 92888 $9850 Q2:33 cs 9.3 .22., €238 ”cs we :02 23 ”3.8 ”a 3.8 23385. .5 32 $3 “2.3338 3,32 3: ”2:32; 6. 8m ”deep E -333 §§§£ 150 .28 m3 333.: .9 .3338 3 33:: d .238 on 3383 wanna—323 .wfiaonoa .333“ 8333393 5on»: d 3.333 335.3 3933 msm 3.3 $63 o3N 336w wm.m3 we. 3 33.333 mod 3303 33 no.3 «33.333 ms. 3 mad 3 mmd 2.3.3 333 3 033.8 ems om.3 $3.3 3 9.3 3333.3 3 cm 3.83 mod 3.3 3.6 :3 3.0 cm 3.3 3 Rd .3335 mod 3 3.3 mad om m3 .om3 cod oo.o cad oo.o oo.o o mania u3o§3=o9m coco—38:3-N 93.35.38.383 92538“ U323.233». 32:33 as $3 .395 ”382% ”as we :02 23 ”33 ”a 3.9 532833 8 32 $3 ”335 5&2 o.» ”2932.3 no. com ”deep am 3 - :3 geostaxm 151 .2333 93 333333 .9 .3338 .33 333333 .9 £33333 mm 9383 wcuaonam .wcu339339333 393.339 33933333333 339333;: .39 and 3.8 3.3 aodm N23 anon 3.3 chm co. 3 m no.3 39mm 5.3.3 mm. 3 n 33 29.x wvdm 8.393 333.29 2.63 $.39 3 3 mode 353 as. mm.mm 8&3 wmdm mm 99.3 396 min 5.3.9 3..m 39.333 mm 3933.3 3 mod aim N03 Sum mod mm 3 .om3 oo.o oo.o oo.o cod oo.o 33 993-33 93o33833933-m 99=o33333933-~ 933953383383 99383839 9323333339 9933333 332 333.3 .2333 ”35:8 333.: 333 33an 23 393m 339 3.8 33333938293 35 32 #8 ”333338 3332 33.33 ”2339.23 6. 8a H333353. % 3 - :3 33323333233333 152 .2333 93 3333: .9 .3333: 3.33 333333 .33 £33333 mm 3383 33333393723 .waoaonéa 393333 33933333333 339333133 .39 N339 omdm who ohvm go 2.3. ovm v3 .99 omdm 3 .o mvom mm .o 33.39 3.3 m 3N.3m Sow moo 8.33m who owfim $33 goo oodm oo.o wmdu o 3 .o 3.3. on 3 when 8.3 So no.5 3o anon o~3 Ego oo.om moo 33.3 moo oo.3m ca wwso oo.o3 woo when oo.o vwfim co No.23 3.333 32. ~03 cos :63 cm 2.333 3 oo.o oo.o oo.o 3mm oo.o 33 993-33 932383399~ 9232383399~ 9339332383383 993839939 93233359 o3.333333 333: So 3933 33339>3om 33m omdo 3.3092 733 383mm 33333338903 35 32 owoo ”3.3333335 8332 Wm 32358333 «I ”D. Em 32333893333393. 93 -6333 353333.39qu 153 .328 .33 333333 .9 .3333: .33 333333 .o 358 mm 3383 $5833-23 .wcuaonéa 393333 33933333? 383333.333 .39 «Gnu n3én no.3. oodn 3339.3. no.3:9 ova fixnn 3n. 3 n nn.o noon nn.o on: SN 35933 no.2... nn.o on. 3 n nn.o 3 .an 33 no.3. 2.3.3 nn.o nv.3n 3 .n ~3dn on 3 «now 3.3 3o.o Dunn vwo anfin oN3 anno no.3 3n.n oo.om 3o.n monn om «ado 332.3 33.33 333.8 333.333 nodn co cons cod 35 no.3 3 333.333 n33 on 5.3.3 3 3333.33 NN.3 oo.o can oo.o o 993$ 933333833933-~ 9933o3333333933-~ 9333533383383 9933039933 933333333339 33933333 33333 33393 393333 3333938 ”Q 3233 20332 7: 39335. ”33333833333933? 35 373 ”39333330 3333332 on 3933338333 «E ”U. o3~ 38333333233333 9N-~# 33393333393333m 154 .2333 3.33 333333 .9 .3338 «3 333333 .33 £33333 mm 3383 w333333033-o333 .w33333323é33 333333 3333333333“ 383333.333 .33 36m 3333.5 353 30am mus wfiwm ova 3.3.0 23.3.3 9mm Sign mud 339mm o3~ 8.8 3.3.3 ~33 wuém ems aw.~m 33333 3.2. 3.2 wqm :83 333.6 @363 on 3 8.8 3.3.m3 8.3. 3963 23. mmén 332 3333.3 oo.o 33:3 a3.m3 mos 3.333 ca 325 E.» cod 330.333 mmfi 8.3 cc no.3 33.3. 333 mod 36 333.333 3333 3.33 3333.33 333.33 3333.33 mud $3 33 393-3% U33333333333333.N $533353: Q33333333333333933 U33333333833 U333333333333 3323333 333: 33333 303333 33co>3om 33% 33233 330332 7: 39.35 332 33333 ”3933333330 333.332 Wm 6.3358333 N33 5. 333a 32333333233338. amén 333233330qu .2333 .33 333333 .3 .3338 3.3 333333 .3 £33333 nu 3333333 333333-233 .w333333o33é333 3333333 333333333 383333.333 .33 155 R633 3m.w 3 3.33 333.2 333.» 333.333 333% 331.5 2.5 cud 3.33 33.3. 3.333 333m 333.3 03 .m owN ohm 3 3 .33 3.333 33w3 5.3 3.336 aim 3.3: 3 3.5 03.33 cm 3 3m.3333 3333.33 nn.o 3.3 .c nn.o 3.333 33~3 $.83 3~.m «Wm 31.33 3333.0 wed 3333 3.33333 3.3 and ~w.~ 333mm 533.3 330 m 3 .m3 3 3.33.33 333.3 3303 mmé and 33m 33m.33~3 3333.33 3333.33 3333.33 an .3 3333.33 33 mania u3o33a333o33-~ 32333333333323-~ U333333333333333333 33333333833 U333333333333 323333 333: 33333 33333 3333033333..“ 33% 33m.333 3303373 23 ”83333 3330 30:33 “39333330 333332 n.m 383338333 N33 30. 333 3233382333393. 3331? 333033333235 .2333 33 333333 .9 .333333 33 333333 .33 £33333 mm 3383 w33333393_-9333 .w333333933-9333 393.333 33933333333 3393333333 .9 156 3.8 2.333 and 2mm $3.3 m3 .3. ova 335.3 3&3 32. end wad 3 3 .v 333 333.8 333.333 33. Rd w3 .n «Wm 33 no.3 3.33 36 333.~ 3.N n33.m 33 99.83 m 3 6 ~33 no.3 3333.3 nn.~ 333 339.83 $6 86 3.3 «3 .N Sta 33a 3.3.333 2.3. saw 33 3333.3 «5.3 cc ~03 33 mmd 339m 5.33 2.3 3.3 3 333 .3 3 33.33 3333.~ 3333.33 3.3 .3 3333.33 33 993-3% 93333333333933-N 9933o33~333933-~ 9323333383383 933333339933 933333333339 33933333 33333 33393 38333 33339>3om 3393333333 330332 733 3835 3330 39333333 3333333330 3332 33.m 383339.333 N3.3 ”0. 333m ”23333339333383. 9m-~# 33333333339qu .2333 m3 333333 .9 .3333: 33 333333 .9 £33333 mm 9383 w333333933-9333 .w333333933-9333 3933“ 33933333333 comet»: .9 157 333. 3 m 3.33 mad £9.03 3&6 Exam ova awNw 335 3333.39 8% 3 was 3.33m 333 3.3933 wvd wad 5.93 333.9 33363 33333 333.3 mm.» mnd 3&3 933.3. and 3 33m3 03.3333 33:. mad 3.3 3 33:. 33383 33 2.3 mad mn.m 3.x 333.5 3333.33 cm 33.3 mad 3o.m m3; 333; 3.333 3333 3333.93: ~33.m $.m v3 .3. 3333.0 $6 on 333 63 3 33m.3 39.3 33033 3.3 @333 33 993-3% 93333333333933-N 99353333333933-~ 93933933933383 99333339933 93933333339 9933333 333: 33393 393333 ”33393333 33% 3.333 330332 Z3 39.3333 3330 393333.33 339333333330 33332 33% 3933339333 N33 30. 33mm ”93333333233338. 99-933 335833233333 158 .2333 93 333333 .9 .333333 93 333333 .33 .9333333 mm 3333333 333339339333 .w333333933-u333 393333 33933333333 383333.333 .9 $9.339 3333.333 m3 .0 2&3 3 36 3&3 ova 3.2. N383 33.0 93.33 333.3. 333 .m 3 333 93.2. 33.3.3 39.0 3.3 3 336 $3.333 333 3.3.3. 3333.33 3.3. 3903 3 3 3.33 3&3 33m 3 8.33 339.93 3% 906 m2. mm.n3 33N3 3.8 333.3 N333. 3333.33 av.» 83 3 33¢ mc.mw 3&3 3333.3. 3333.3. mod 33.333 330 Eda 3.3. 9% 333.39 2.3. 332‘ cm 3339.2 3 and 3.3 3333.33 NON 3333.33 33 993-3% 93o3333333933-~ 99353333333933-N 9393393383383 9933333833 9333333333333 33933333 33:3 3393 393333 333393338 33% 3.333 33333373 733 383333 3330 393333 ”5.3393330 3333332 33% 323398333 «313 ”U. 33mm ”23333339333338. awg 33323333933383 .2333 .33 333333 .9 .3333: 33 333333 .9 3333333 mm 3333333 333339336333 .w333333933-9333 33.333 33933333333 3393333333 .9 159 wmdw 333d m3...V mud no.» 2.33 gm 3.33m $.33 333.9 33033 no.» m3: 333 339mm 8.33 39.39 3.3. 33:. 93.2 33 2.8 and 3 3w 3.3. and 3.3 3 33m 3 3333.3 nn.o 3:9 23 3.33 3.333 33 8.33333 2mm mmé 333.n 33:. 2.3. 3333 33333 mod RM mod 3% 3N.m co m: 33 mod 3.30.3 $3 35d and 3333 3339.332 3339.33 933.3 3333.33 m3..3 3333.33 33 9033-33 93333333333933-~ 99333333933323-~ 93335333333383 9933333833 9333333333333 9933333 3339 3393 393333 33339>3om 33% 33233 3303373 733 383333 3330 393333.33 ”6.3333333 333.333 33% 3933398333 a: 30. 33mm 39.33333339333339H 9a-? 3339333339qu 160 .2333 93 333333 .9 .333333 333 333333 .9 3.333333 mm 3333333 ”3333339336333 .w333333933-9333 393.333 33933333333 33933333: .9 9N: $.3N 99d mmdm mad amdm 339m 339.9 933.3m 99.m 333. 3 93 .9 N3 bu 333m 33.933 99.93 cod cad 3 $.33 3.3.333 33333 No.3 $3 3 33..~ $333 3.33.m m 3 .N3 33m 3 933.33 933.333 33 m9.m %N 93.3 3 33N3 9.933 mm .333 3333.33 23. 3 3 93.9 99.93 ca 33033 39.3. 3.9 933.3. 99m 3339.3 3 339 93.933 33.3 933.3 903 SN and 3333 3.33 3.3..33 39.3 $.33 and 3.3 33 9033-9.n 93333333333933-N 99333333833933-~ 93933333383383 99333339933 933333333339 9933333 3339 3393 393333 33339>3om 33.9 3N.333 $333373 23 39.9.3333 3330 393333.33 3393333333 3333332 33.n 323338333 3: ”U. can 39333393933th 99.93 333933333me 161 .2333 33 333333 .9 .3333: 93 333333 .9 3.333333 mm 3333333 w3333339339333 .w3333333333-9333 393333 33933333333 339M353»: .9 3.9.99 933.33N madm 3.2 95.93 w3..9m 33w3 $3.93. 39.93 3 .93 m9.~3 99.~3 233 339 3 8.3 933.3 93.93 33.3 3 $3.3 3 mm.w3 33~3 3m.~w 3333.333 $.93 3.3 .333 9m.m 3 99.53 3 339.33 933.33 99.m3 mm.w 39.3 333.9 3 339 339.3 99 .9 330333 99.9 59.3 3 933.9 3333 933.333 99.3 9~.m 3.3 5.9 9nd 33 9033-9.N 9333339333933-~ 99333333833933-~ 93335333333383 99333339933 933333333339 9933333 333: 3393 393333 33339>333m 33% 3333 330332 733 383333 3 33D 9393333 33933333330 3333332 33.m 33333333933333 N3.3 6. 33mm 39.333333333333339h 33333 .~# 33393333393333m3 162 .2333 93 333333 .9 .333333 93 333333 .9 £33333 mm 3333333 w333333933-9333 .w333333933-9.333 393333 33933333333 33933333333 .9 339.99 339.99 93.93 90339 933.93 3.9.3.3“ 3393 339.99 9m.mm 99.333 3 Q99 93.333 339.99 3393 mm .339 99.333 99.9 whom 99.333 m9.wm 3393 99.99 93 .93 3.33.9 99.339 99.33 333.99 339 339.99 39.9 3 33.333 3333.3.3 99.93 93.99 339 99.93. 99.333 99.33 99.333 339.3 3 33.3.3 33m 339.93 3 an .3 m9.~ 93 .3 R9 99.9 33 9939.9 93333393339339 99333333833933-~ 93933339333983 993333333933 933333333339 93333333 3339 3393 393333 33339>333m 33.9 339.333 3333332 733 383333 3330 3393333 ”39339390 3333332 33.9 39333338333 33.3 30.. 3399 ”93333333933893. 3.3 3-93 3533335333333 163 .2333 33.3 333333 .9 .333333 333 333333 .9 3.3333333 99 3333333 93333339339333 93333339339333 3333333 33933333333 999939.333 .9 99.933 39.93 933.3 99.93 99.9 99.9 3 3399 933.933 99.93 99.3 99.93 99.9 933.9 3 3339 99. 39 99.3 3 99.3 339.93 99.9 933.93 3393 333.99 3 3.33 39.3 99.9 99.9 99.333 339 3 93 .99 39.333 99.9 933.9 99.9 93.33 3393 339.3333 99.9 33.9 933.9 3333.9 99.3. 339 99.9333 39.9 39.9 39.9 99.9 99. 339 93.9333 99.9 339.3 99.9 339.9 99.3.. 339 933.93 3 3333.3 933.3 3333.33 99.3 3333.33 33 90339.9 93333393339339 993333333335339 933333339333983 99333339933 933333333339 93333333 3393 3393 39393 333333>333m 33.9 39.333 330332 23 383333 3330 333333333 33933333330 333332 33.9 323389333 933 30. 33339 32333333933893. 3393 -99 33393333393335 164 .2333 93 333333 .9 .333333 93 333333 .9 9333333 99 3333333 93333339339333 93333339339333 393333 33933333333 33993333333: .9 99.99 39.93 99.9 99.93 933.9 9 3 .9 3 3399 99.99 99.9 3.9.9 99.333 99.9 93.93 3339 99.3.9 99.9 99.9 99 .9 339.9 93 .33 3393 93.99 99.9 39.9 99.9 333.9 99.9 339 3 99.9333 39 93 .9 99.9 99.9 333 .9 3393 3.9.9333 99.9 339.9 99.9 99.9 33..3. 339 99.9333 99.9 99.9 99 .9 99.9 93 .9 339 99.9333 99.9 99.9 99 .9 99.9 933.9 339 339.3393 339.33 99. 3 3333.33 93 .9 3333.33 33 90339.9 93333333333999 99333333835339 93933933939993 993333393333 933333333339 93333333 33333 3393 393333 3333233339 339 339.333 3.30332 23 3993333 3330 3393333333 33933333330 333332 33.3. 393333399333 933 ”U. 3399 32333333933893. .933 -93 33323333233333 LIST OF REFERENCES LIST OF REFERENCES Alexander, 1954, Studies on enol titration. II. Enol content of some ketones and esters in the presence of methanol. J. Org. Chem, 41, 3299-3302. Andrews, M. A. and Klaeren, S. A., 1989, Selective Hydrocracking of Monosaccharide Carbon-Carbon Single Bonds under Mild Conditions. Ruthenium Hydride Catalyzed Formation of Glycols. J. Am Chem Soc., 111, 4131-4133. Arena B.J., 1983, Hydrogenolysis of polyhydroxylated compounds. US Patent 4,401,823. Ariono, D., Moraes de Abreu, C., Roesyad, A., Declercq, G. and Zoulalian, A., 1986, Hydrogenolyse sur catalyseurs solides de solutions de glucose. Bull. Soc. Chim Fr., 703-710. Baggett, N., Buck, K. W., Foster, A. B., Randall, M. H. and Webber, J. M., 1965, Aspects of Stereochemistry. Part IV. Absolute configuration of Benzylidene derivatives of some acylic polyhydric alcohols. J. Chem Soc., 3394—3400. Baggett, N., Brimacombe, J. 8., Foster, A. B., Stachey, M. and Weiffen, D. H., 1960, Aspects of Stereochemistry. Part IV. Configuration and some reactions of the 1,3-O- Benzylideneglycerols (5-hydroxy-2-phenyl-l,3-dioxans). J. Chem Soc., 2574-2581. Barbirato, F., 1996, 3-Hydroxypropionaldehyde, an inhibitory metabolite of glycerol fermentation to 1,3-propanediol by enterobacterial species. Appl. Environ. Microbiol, 62, 1448-1451. Barrier J. W. and Bulls M. M., 1992, Feedstock availability of biomass and wastes. In Emerging Technologies for Materials and Chemicals from Biomass (Edited by Rowell R. M. et al.), pp 410-421. Americal Chemical Society, Washington DC. Biebl, H., 1995, Fermentation of glycerol to 1,3-propanediol: use of cosubstrate. Appl. Microbiol Biotechnol, 44, 15-19. Boelhouwer, C., Korf, D. and Waterman, H. I., 1960, Catalytic hydrogenation of sugars. J. Appl. Chem. 10, 292-296. Boenigk, R., Bowien, S. and Gottschalk, G., 1993, Fermentation of glycerol to 1,3- propanediol in continuous cultures of citrobacter freundii. Appl. Microbiol Biotechnol, 38, 453-457. 165 166 Brimacombe, J. 8., Foster, AB. and Haines, A. H., 1960, Aspects of Stereochemistry. Part V. Some properties of 1,2-O-methylideneglycerol and related compounds. J. Chem Soc., 2582-2586. Cameron, D. C., 1998, Metabolic engineering of propanediol pathways. Biotechnology Progress, 14, 116-125. Casey, M., Lenoard, J. and Lygo, B., 1990, Advanced Practical Organic Chemistry. pp133-134. Blackie Academic and Professonal: London. Chaloner P. A., 1986, Handbook of Coordination Catalysis in Organic Chemistry; Butterworths: Boston. Chang, F. W., Kuo, K. T. and Lee, C.N., 1985, A kinetic study on the hydrogenolysis of sorbitol over Raney nickel catalysts. J. Chin. I. Ch.E., 16 , 17-23. Chao, J. C., Huibers, D. T. A. and Derk, T. A., 1982, Catalytic hydrogenolysis of alditols for producing glycerol and polyols. Chem Abtra., 98: 18453. Chapman, P., 1997, Shell advances plans for new polyester. Chemical Market Reporter, 250, 1. Chapman, P., 1996, Dupont in race to commercialize a new polyester. Chemical Market Reporter, 251, 3. Che, T. M., 1987, Catalytic conversion of glycerol and synthesis gas to propanediol. US Patent 5093537. Chong, J. M. and Sokoll, K. K., 1993, Synthesis of 2-O-protected glycerol derivatives. Org. Prep. Pro. int, 25, 639-647. Chum, H. L. and Power, A. J ., 1992, The promise and pitfalls of biomass and waste conversion. In Emerging Technologies for Materials and Chemicals from Biomass (Edited by Rowell R. M. et al.), pp 339-353. America] chemical Society: Washington DC. Clark, 1. T., 1958, Hydrogenolysis of Sorbitol. Ind. Eng. Chem, 50, 1125-1126. Clausen, E. C. and Gaddy, J. L., 1987, The production of chemicals and fuels from municipal solid wastes. In Global Bioconversions, vol II (Ed. by Mse D. L.), pp 61-73. CRC: Boca Raton, FL Collman, J. P., Hegedus, L. S. and Finke R. G., 1987, Principles and Applications of Organotransition Metal Chemistry. Universe Science Books, Mill valley, CA. 167 Connor, R. and Adkins, H., 1932, Hydrogenolysis of Oxygenated Organic Compounds. J. Am Chem Soc, 54, 4678-4690. Cook, J. and Mailis, P. M., 1981, Formaldehyde as a Hydrogen-donor to Aldehydes and Ketones in Metal-catalyzed Reactions in Water: J. Chem Soc., Chem Comm, 924- 925. DeAthos, S. J., 1989, A preliminary study of the conversion of glycerol to 1,3- propanediol. MS. Thesis, Chemical engineering Department, Michigan State University, East Lansing, Michigan. Dobinson, B. and Foster, A. B., 1961, Aspects of Stereochemistry. Part IV. Configuration of the 5-hydroxy-2-p-nitrophenyl-1,3-dioxans and futher observations on trans-5- hydroxyl-2-phenyl-1,3-dioxan. J. Chem Soc., 2338. Dubeck, M. and Knapp, G. G., 1984, Sulfide-modified ruthenium catalyst. US patent 4,430,253. Chem Abstr., 102:167115. Dubois, J. E. and Vrellard, H., 1964, Mecanisme de la decetolisation. Tetra. Lett., 1809- 1815. Edashige, Y., Ishii, T. and Hiroi, T., 1990, Anhydrorous hydrogen fluoride sacchrification of lignocellulosis materials. Chem Abstr. 120:301338. Fieser, L. F. and Fieser, M, 1967, Regents for Organic Synthesis, Vol.1, pp1180. Weiley: New York. Flowers, H. M., 1971, Protection of the hydroxyl group. In The chemstry of the Hydroxyl Group, Part 2. Ed. by Patai, S., pp 1001-1004. Interscience: London. Fumey, T. D., 1995, A mechanism and selectivity study of sugar and sugar alcohol hydrogenolysis using 1,3-diol model compounds. MS. Thesis, Chemical Engineering Department, Michigan State University, East Lansing, Michigan. Franz, R., Erckel, R., Riehm, T. and Worenle, R., 1982, Lignocellulose saccharification by HF, in Energy fiom Biomass: Second E. C. Conference (Edited by A. Strub, P. Chartier and G. Schleser), pp. 873-878. Applied Science, London. Goldstein I. S., 1992, Chemicals and fuels from biomass. In Emerging Technologies for Materials and Chemicals fi'orn Biomass (Edited by Rowell R. M. et al.), pp 332-353. Americal Chemical Society, Washington DC. Goldstein I. S., 1981, Organic Chemicals from Biomass. CRC: Boca Raton, FL 168 Grohman, K., Wyman, C. E. and Himmel M. E., 1993, Potential for fuels from biomass and wastes. In Emerging Technologies for Materials and Chemicals from Biomass (Edited by Rowell R. M. et al.), pp 354-392. Americal Chemical Society, Washington DC. Gunzel, B., Yonsel, S. and Deckwer, W. D., 1991, Ferrnentative production of 1,3- propanediol from glycerol by clostn'dium butyicum up to a scale of 2 m2. Appl. Microbiol Biotechnol, 36, 289-294. Guthrie, J. P., 1977, Equibilibrium constants of a series of simple aldol condensations, and linear free energy relations with other carbonyl addition reactions. Can. J. Chem 56, 962-972. Guthrie, J. P., 1974, The aldol condensation of acetaldehyde: the equibilibrium constant for the reaction and the rate constant for the hydroxide catalyzed retroaldol reaction. Can. J. Chem 52, 2037-2040. Hawley, M. C., Downey, D. W., Selke, S. M. and Lamport, D. T. A., 1986, Hydrogen fluoride saccharification of lignocellulose materials, in Cellulose: Structure, Modification, and Hydrolysis (Edited by R. A. Young and R. M. Rowell), pp. 297-321. Wiley, New York. Hawley, M. C., Selke, S. M. and Lamport, D. T. A., 1983, Comparision of hydrogen fluoride saccharification of lignocellulose materials with other saccharification technologies. Energy Agric. 2, 219-244. Hesse], L. W., van Lohuizen, O. E. and Verkade, P. E., 1954, Glycerol-or— and [5- monoiodohydrin. Rec. Trav. Chim, 73, 842-848. Hibbert, H. and Carter, N. M., 1928, Studies on the reactions relating to carbonydrates and polysaccharides. XVIL. Structure of the isomeric methylidene glycerols. J. Am Chem Soc., 50, 3120-3127. Hill, H. S., Whelen, M. S. and Hibbert, H., 1928, Studies on the reactions relating to carbonydrates and polysaccharides. XV. The isomeric benzylidene glycerols. J. Am Chem Soc., 50, 2235-2242. Hydrocarbon Research Inc., 1981, Two step conversion of aldoses to polyols. Neth. Appl. NL 8.105.092. Irvine, J. C., Macdonald, J. L. A and Soutar, C . W., 1915, XXXIX. Condensation of acetone and benadehyde with glycerol. Preparation of glycerol or-methyl ether. J. Chem Soc., 107, 337-351. James, B. R., 1973, Homogeneous Hydrogenation. John Wiley & sons, New York. 169 Jensen, J. L. and Hashtroudi, H., 1976, Base-catalyzed hydration of a,B-unsaturated ketones. J. Org. Chem 41, 3299-3302. Jensen, J. L. and Carre, D. J., 1974, Kinetics and Mechanism of Reaction of 3-Buten-2- one and Related Compounds in Aqueous Perchloric Acid. J. Org. Chem 39, 2103- 2107. Jochims, J. C. and Kobayashi, Y., 1974, Sterospecific synthesis of 5-substituted 1,3- dioxanes by application of Dimroth’s principle. Tetra. Lett., 575-578. Juaristi, E. and Antunez, S., 1992, Conformational analysis of 5-substituted 1,3-dioxanes. 6. Study of the attractive gauche effect in O-C-C-O Segments. Tetrahedron, 48, 5941- 5950. Kuhad, R. C., 1993, Lignocellulose biotechnology: current and future prospects. Critical Reviews in Biomass, 13, 151-172. ‘ Latosh, M. V., 1984, Gas chromatographic identification of polyols in products of carbonhydrate hydrogenolysis. Chem Abstr. 102:26593. Lipinsky, E. S., 1981, Chemical from biomass: petrochemical substitution options. Science, 212, 1465-1471. Luers, H., 1938, Das holzverzucherungsverfahren mit fluorwasserstoff von Hock und Bohunek. Halz Rah- werkst 1, 342-344. Maksimenko, O. A., Zyukova, L. A., Ignat’eva, E. V. and Fedorovich, R. M., 1975, Quantitative gas chromatographic analysis of the hydrogenolysis products of carbohydrate. Chem Abstr., 84:31333. Masters C., 1981, Homogeneous Transition-metal Catalysis: A gentle Art. Chapman and Hall, New York. Matheson, N. K. and Angyal, S. J ., 1952, The replacement of secondary tosyloxy-groups by iodine in polyhydroxy-compounds. J. Chem Soc., 1133-1138. Mestroni, G., Spogliarich, R., Camus, A. Martinelli, F. and Zassinovich, G., 1978, Selective hydrogenation of unsaturated ketones by rhodium (1) complexes containing 2,2’-bipyridine ligand. J. argnametal. chem, 157, 345-352. Mohring, W. R. and Hawley, M. C., 1989, Gravimetric analysis of hydrogen fluoride vapor adsorption by big-tooth-aspen wood. Ind. Eng. Chem Res. 28, 237-243. Molnar, A., Felfoldi, K. and Bartok, M., 1981, Transformation of 1,3-diols on homogeneous and heterogeneous rhodium catalysts. J. Mol. Catal, 11, 225-232. 170 Montassier, C., Menezo, J. C., Hoang, L. C., Renaud, C. and Barbier, J., 1991, Aqueous polyol conversion on ruthenium and on sulfur-modified ruthenium. J. Mol. Catal, 70, 99-100. Montassier C., Giraud D., Barbier J. and Boitiaux J. P., 1989, Transformation de polyols par catalyse heterogene en phase liquide sur les metaux. Bull. Soc. Chim Fr., 148-155. Montassier C., Giraud D. and Barbier J., 1988, Polyol Conversion by Liquid Phase Heterogeneous Catalysis over Metals. In Heterogeneous Catalysis and Fine Chemicals; Guisnet M. et. al. Ed.; Elsevier: Amsterdam. Neilsen, A. T. and Houlihan, W. J., 1968, The Aldol Condensation. In Organic Reactions, val.6, Adams R. et. al. Ed., Weiley: New York. Palsson, B. 0., Fathi-Afshar, S., Rudd, D. F. and Lightfoot, E. N., 1981, Biomass as a source of chemical feedstocks: an economic evaluation. Science, 213, 513-517. Perry, R. H., Green, D. W. and Maloney J. 0., 1984, Perry’s Chemical Engineer’s Handbook, 6th. pp 3-144 to 3-146. MaGraw-Hill, New York. Piasecki, A. and Burczyk, B., 1980, Acetals and ethers. Part VI. Synthesis of selected cis- and trans-2-alkyl-4-hydroxylmethyl-1,3-dioxolanes and cis- and trans-Z-alkyl-S- hydroxy-1,3-dioxanes. Pol. J. of Chem, 54, 367-372. Pignolet L. H, 1983, Homogeneous Catalysis with Metal Phasphine Complexes. Plenum Press: New York. Pikkov, L. M., Sokovov, N. M. and Tearo, E. N., 197 3, Isolation of glycerol from products of sucrose hydrogenolysis in spray apparatus. Chem Abstrz, 80:97637. Reath, M. L., 1989, Bench scale studies of HF saccharification of wood with 1.5 liter packed-bed reactor. MS. Thesis, Chemical Engineering Department, Michigan State University, East Lansing, Michigan. Reimann, A., 1998, Production of 1,3-propanediol by Clostrium Butrium in continuous culture with cell recycling. Appl. Microbial Biotechnol, 49, 359-363. Roner, G. L., 1989, Solvolysis of a single lignocellulose particles by anhydrous hydrogen fluoride vapor: reaction rate studies, sugar yield studies and development of a detailed kinetic and heat transfer model. Ph.D. dissertation, Chemical Engineering Department, Michigan State University, East Lansing, Michigan. Rorrer, G. L., Mohring, W. R., Hawley, M. C. and Lamport D. T. A., 1988, Adsorption and reaction processes of HF solvolysis of wood and pure cellulose by anhydrous hydrogen fluoride vapor. J. Energy Fuels, 2, 556-566. 171 Rylander, P. N., 1985, Hydrogenation Methods. Academic Press, Orlando, Florida. Selke, S. M., Hawley, M. C. and Lamport, D. T. A., 1983, Reaction rates for liquid phase hydrogen fluoride saccharification of wood, in Wood and Agriculture Residues: Research on Use for Feed, Fuels, and Chemicals (Edited by E. J. Soltes), pp. 329-349. Academic Press, New York. Selke, S. M., Hawley, M. C., Hardt, H., Lamport, D. T. A., Smith, G. and Smith, J., 1982, Chemicals from wood via HF. Ind Eng. Chem Prod. Res. Dev., 21, 11-16. Showler, A. J. and Darley, P. A., 1967, Condensation products of glycerol with aldehydes and ketones: 2-substituted m-dioxan-S-ols and 1,3-dioxolane-4-methanols. Chem Rev., 67, 427-440. Smith, G. G. and Yates, B. L., 1964, Pyrolysis Studies. XV. Thermal retrograde aldol condensation of B-hydroxyl ketones. J. Chem Sci., 7242. Smith, J., Lamport, D. T. A., Hawley, M. C. and Selke, S. M., 1983, Feasibility of using anhydrous hydrogen fluoride to “crack” cellulose. J. Appl. Polym Sci.: Appl. Polym Symp., 37, 641-651. Sohounloue, D. K., Montassier, C. and Barbier, J., 1983, Catalytic Hydrogenolysis of Sorbitol. React. Kinet. Catal. Lett., 391-397. Sokolvo, N. M., Tearo, E. and Siirde, B., 1977, Isolation of glycols and glycerol from sucrose hydrogenolysis products by extraction. Chem Abstr., 88:23287. Sokolvo, N. M., Shtrom, M. 1., Zhavronkov, N. M., 1975, Separartion and purification of sucrose hydrogenolysis products. Chem Abstrz, 83:166124. Stefoglo, E. F., Ermakova, A. Vasyunina, N. A. and Klabunovskii, E. 1., 1973, Kinetic model for the preparation of glycerol by the catalytic hydrogenolysis of glucose. Chem Abstr., 80:61336. Stinson, S. C., 1995, Fine and intermediate chemicals makers emphasize new products and processes. C&E News, 73, 10. Swinnen, J. and Tollens, B., 1991, Bulk chemicals from biomass in the EC: feasibility and potential outlets. Biaresaurce Technology, 36, 277-291. Szeja, S., 1983, Separation of 2-substituted 5-hydroxy-l,3-dioxanes by esterification in catalytic two-phase system. Pol. J. of Chem, 57, 609-612. Szmant, H. H., 1986, Industrial Utilization of Renewable Resources. Technomic Publishing Company: Lancaster, PA. 172 Tanikella, M. S. S. R., 1983a, Hydrogenolysis of polyols to ethylene glycol. Eur. Appl. EP 72,629. Tanikella, M. S. S. R., 1983b, Hydrogenolysis of polyols to ethylene glycol in non- aqueous solvents, Eur. Appl. EP 86,296. Tipson, R. S., 1953, Sufonic esters of carbohydrate. In Advances in Carbohydrate Chemistry, voL 8. Academic: New York. Tronconi, E., Ferlazzo, N., Forzatti, P., Pasquon, 1., Casale, B. and Marini, L., 1992, A Mathematical Model for the Catalytic Hydrogenolysis of Carbohydrates. Chem Eng. Sci., 47, 2451-2456. Trosr, B. M., 1991, Carrrprehensive Organic Synthesis, vol. 6, pp 659. Pergamon, New York. Van Ling, G, Driessen, A.J., Piet, AC. and Vlugter, LC, 1970, Continuous Production of Glycerol by Catalytic High Pressure Hydrogenolysis of Sucrose. Ind. Eng. Chem Prod. Res. Develop., 9, 210-212. Van Ling G. and Vlugter J. C., 1969, Catalytic Hydrogenolysis of Saccharides II. Formation of Glycerol. J. Appl. Chem, 19, 43-45. Van Ling G., Ruijterman G. and Vlugter J. C., 1967, Catalytic Hydrogenolysis of Saccharides I. Qualitative and Quantitaive Methods for the Identification and Determination of the Reaction Products. Carbohyd. Res., 4, 380-386. Verrna, R. and Gehlawat, J. K., 1989, Kinetics of hydrogenation of D-glucose to sorbitol. J. Chem Tech. Biatech., 46, 259-301. Wang, K., Fumey, T. D. and M. C. Hawley, 1995b, A mechanism study of sugar and sugar alcohol hydrogenolysis using 1,3-diol model compounds. Ind & Eng. Chem, 34, 3766- 3770. Wang, K., Fumey, T. D. and Hawley, 1995a, Modeling the HF adsorption process on wood chips in a packed-bed reactor. Chem Eng. Sci, 50, 2883-2897. Wang, K., Hawley, M. C. and Reath, M. L., 1994, Adsorption model of HF on wood chips in a packed-bed reactor. Chem Eng. Sci, 49, 2321-2329. Weclas, L., Sokolowski, A and Burczyk, B, 1982, Acetals and ethers. Part X. Synthesis of selected cis- and trans-2-alkyl-4-[(2-hydroxyethoxy)methyl]-1,3-dioxolanes and cis- and trans-2-alkyl-5-(2-hydroxyethoxy)-1,3-dioxanes.Pal. J. of Chem, 56, 485-490. Wilson, R. A and Martin C. G., 1992, Paper tree, A solution to environmental problems. Paper Technology, 2, 26-31. 173 Wise D. L., 1983, Organic Chemicals from Biomass. The Benjamin/Cummings Pulishing Company: Menlo park, CA. Yates, B. L. and Quijano, J., 1969, The thermal decomposition of B-hydroxyl ketones. J. Org. Chem, 34, 2506-2068. Young, C. L., 1981, Solubility Data Series, Volume 5/6, Hydrogen and Deuterium. Pergamon Press, New Yorek. Zeng, A. P., Biebl, H, Schlieker, H. and Deckwer, W.D., 1993, Pathway analysis of glycerol fermentation by Klebsiella pneumoniae: Regulation of reducing equivalent balance and product formation. Enzyme Microb. TechnoL, 15, 770-773.