.Xwn . u... W .5me .. mt. 1-. .c , .- u. i .fifififi 5%? u . :- .\l}.>. : 3. Jung. .12 a. {8% 11.; z -. .1 a. in». - , l . _ ; . . . _ _ . ‘43...,_.~.€,.w.¥ A. ., . 4 ‘ 3. ._.. . .. A. ‘ T 12h. at... z :0 if}; LIBRARY Michigan State UniverSIty This is to certify that the dissertation entitled ADSORPTION OF BIO-RENEWABLE SUBSTRATES ON SUPPORTED METAL CATALYST IN WATER presented by LARS PEEREBOOM has been accepted towards fulfillment of the requirements for the PhD. degree in Chemical Waring; Major/Professor’s Signature 9/17/07 Date MSU is an atfinnative-ection, equal-opportunity employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/07 p:lCIRC/DateDue.indd-p.1 ADSORPTION OF BIO-RENEWABLE SUBSTRATES ON SUPPORTED METAL CATALYST IN WATER By Lars Peereboom A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree DOCTOR OF PHILOSOPHY Department of Chemical Engineering 2007 ABSTRACT ADSORPTION OF BIO-RENEWABLE SUBSTRATES ON SUPPORTED METAL CATALYST IN WATER By Lars Peereboom Biomass-based organic substrates are attractive feedstocks for chemicals production because they are available in quantity and can undergo a variety of reactions to form useful products. In our efforts to understand the mechanisms of aqueous transformations of these substrates over supported noble metal catalysts, we seek to probe interactions between the substrate and catalyst metal and between the substrate and catalyst support. This dissertation recounts our work to measure the extent of substrate adsorption on the catalyst surface and on the support based on difference—Le. by measuring (via HPLC analysis) the initial and final concentrations of substrate in solution. We have developed a novel flow microreactor system (recirculating batch reactor or RBR) with in-Iine substrate detection and a high catalyst metal to solution ratio. Polyol and organic acid interactions with both the unsupported Ru metal catalyst and the carbon support have been examined, the former in the RBR and the latter in batch systems. Volumetric gas phase chemisorption of H2 and CO onto unsupported Ru catalyst gives 4.5 umol Hzlg Ru and 8.7 pmol COlg Ru, respectively. These quantities are consistent with dissociative H2 chemisorption (H:M = 1). Quantities of biorenewable substrates adsorbed on a “Hz-free” (evacuated) Ru surface from water under argon at 25 ’0 are ~0.6 umol sorbitollg Ru, ~0.8 umol/g Ru for CS alcohols and polyols, and ~4.0 umol/g Ru for C3 organic acids. Degradation products are seen during adsorption of polyols, especially sorbitol. When the metal surface is pre- saturated with H; and is under H2, no significant adsorbtion of polyols is seen. In contrast, the adsorption of acids is unaffected by the presence of hydrogen. Sequential adsorption of glycerol (GO) followed by propylene glycol (PG) shows that GO inhibits the adsorption of PG, while reversing the order of adsorption shows no inhibition of GO adsorption by PG. Single and multiple component physical adsorption on activated carbon catalyst supports was examined over a broad range of concentrations (0.005 M to 1.5 M) and temperature (298 - 433 K). To obtain adsorption data relevant to glycerol hydrogenolysis, adsorption of a combination of GO and PG (0.5 M total) was examined over the same temperature range. An extended Langmuir isotherm model, using single components’ adsorption parameters, accurately predicted the two- component adsorption behavior. Adsorption of combined lactic acid (LA) and propionic acid (PA) on activated carbon was not predictable with an extended Langmuir model. We therefore proposed a new model with two modes of adsorption: the filling of micro-pores, and adsorption on macro-pore surfaces. When this new model was expanded to multiple components, it accurately predicted the mulficomponent LA + PA adsorption using single component adsorption parameters. Adsorption of biorenewable substrates on catalyst metal and on the catalyst support has been characterized. Adsorption in both cases can significantly influence the rate of reaction of these biomass substrates to value-added products. In memory of Ilse Morgan ACKNOWLEDGMENT I would like to thank Dr. Dennis Miller for his guidance and support as my research advisor through my Ph.D. studies (especially the editing). I also thank Dr. James E Jackson, Carl T. Lira, Lawrence T. Drzal, and Greg M Swain for serving on my graduate committee. Special thanks to Lisa Benton (my wife) for emotional and financial support. She put up with me for many years of graduate school. TABLE OF CONTENTS LIST OF FIGURES .............................................................................................. X LIST OF TABLES ............................................................................................. XVII 1 INTRODUCTION ............................................................................................. 1 1.1 Significance ............................................................................................. 2 1.2 Objectives ............................................................................................... 3 2 LITERATURE REVIEW ................................................................................... 5 2.1 Hydrogenolysis of glycerol ...................................................................... 5 2.1.1 Kovacs’s et al. ................................................................................ 5 2.1.2 Lahr and Shanks ............................................................................ 6 2.1.3 Tomishige et al. .............................................................................. 7 2.1.4 Suppes et al .................................................................................... 8 2.1.5 Rosier et al. .................................................................................... 9 2.1.6 Perosa and Tundo ........................................................................ 10 2.2 Aqueous phase adsorption on activated carbon ................................... 11 2.3 Species adsorbed on metal surfaces from liquid phase ........................ 12 2.4 Indirect differential detection ................................................................. 13 3 NOVEL FLOW MICROREACTOR FOR SPECIES ADSORPTION ON METAL CATALYSTS ................................................................................................. 14 3.1 Design strategy ..................................................................................... 14 3.2 System description ................................................................................ 16 3.3 System evaluation ................................................................................. 18 3.3.1 System mixing .............................................................................. 18 3.3.2 Volume determination of reactor, tank, loop, and system ............. 19 vi 3.3.3 HPLC calibration ........................................................................... 20 3.4 Methods of operation ............................................................................ 23 3.4.1 General description of experiment ................................................ 23 3.4.2 Data processing ............................................................................ 24 CATALYST CHARACTERIZATION .............................................................. 29 4.1 Methods ................................................................................................ 29 4.2 Results .................................................................................................. 30 H2lDzo EXCHANGE ..................................................................................... 34 5.1 Experimental ......................................................................................... 34 5.2 Modeling ............................................................................................... 37 5.2.1 Approach to solutions ................................................................... 39 5.2.2 Parameter optimization ................................................................. 41 5.2.3 Results and conclusions ............................................................... 43 AQUEOUS PHASE ADSORPTION ON RUTHENIUM SPONGE .................. 45 6.1 Sorbitol, glucose and gluconic acid adsorption ..................................... 45 6.2 Glycerol Adsorption ............................................................................... 52 6.2.1 Concentration dependence ........................................................... 52 6.2.2 Hydrogen saturation ..................................................................... 55 6.2.3 Temperature effects on GO adsorption ........................................ 58 6.2.4 Propylene glycol (PG), ethylene glycol (EG), and 1,3-propanediol (1 ,3-PDO) adsorption .................................................................... 61 6.2.5 Binary adsorption (GO, PG) .......................................................... 64 6.3 lsopropanol and acetone adsorption ..................................................... 67 6.4 n-Propanol and ethanol ......................................................................... 72 vii 6.5 Organic acid adsorption ........................................................................ 74 6.6 Comparisons and conclusions .............................................................. 78 7 AQUEOUS PHASE ADSORPTION ON ACTIVATED CARBON ................... 80 7.1 Experimental methods .......................................................................... 80 7.1.1 Carbon characterization ................................................................ 80 7.1.2 Isothermal adsorption measurements ........................................... 81 7.1.3 Elevated temperature adsorption experiments ............................. 82 7.1.4 Analysis ........................................................................................ 83 7.1.5 Adsorption modeling ..................................................................... 83 7.1.6 Summary of experiments performed ............................................. 85 7.2 Glycerol and propylene glycol ............................................................... 87 7.2.1 Isothermal adsorption of GO+PG on ROX activated carbon ........ 87 7.2.2 Isothermal adsorption of GO+PG on 3310 activated carbon ........ 91 7.2.3 Elevated temperature adsorption .................................................. 92 7.2.4 Two component adsorption studies .............................................. 94 7.2.5 Conclusion .................................................................................. 100 7.3 Ethanol, n—propanol, EG, and 1,3-PDO adsorption ............................. 101 7.4 Organic acids ...................................................................................... 102 7.5 Multi component adsorption studies relevant to hydrogenation of LA and PA to PG and n-prOH ......................................................................... 107 7.6 Additional notes on adsorption modeling ............................................ 119 8 CONCLUSIONS AND RECOMMENDATIONS ........................................... 122 8.1 Conclusions ........................................................................................ 122 8.2 Recommendations .............................................................................. 124 viii APPENDICES ................................................................................................... 125 Appendix A: Master experiment list ............................................................. 126 Appendix B: Supplemental data .................................................................. 139 References .................................................................................................. 147 LIST OF FIGURES Figure 3-1: Recirculating reactor system ............................................................ 15 Figure 3-2: Dimensions of 316 stainless steel reactor insert ............................... 16 Figure 3-3: Configuration of in-line sampling system .......................................... 18 Figure 3-4: Pulse attenuation at 6mllmin for recirculating system. ...................... 19 Figure 3-5: Refractive index detector linear response curves for 10 pl injection of HO-mix and acid-mix standard solutions. .................................................... 21 Figure 3-6: Comparison of auto injector vs. reactor system injector with HO-mix (Sorbitol, GO, EG, 1,3-PG, MeOH, EthOH, IPA, n-prOH) ............................ 22 Figure 5-1: Corrected and uncorrected data for H2 over D20 with 0.1 g 5% RuIC (Exp 43B), and H2 over 020 with 0.1 g 5% Pth (Exp 43C) ......................... 35 Figure 5-2: Incorporation of deuterium into the gas phase from (DO)ethanol over 0.1 g 5% Ru/C (10 ml reactor 100 psi H2 Exp 47) ....................................... 36 Figure 5-3: H2 + D2 over D20 with 0.1 g 5% RulC (50 psi each in 70 ml reactor, Exp 50) ........................................................................................................ 37 Figure 5-4: Proposed reaction model. ................................................................. 38 Figure 5-5: Reversible reaction model ................................................................ 40 Figure 5-6: Model fit to experimental data for 5% Ru/C (Exp 43B). Dotted lines are predicted values ..................................................................................... 44 Figure 5-7: Model fit to experimental data for 5% Pth (Exp 43C). Dotted lines are predicted values ..................................................................................... 44 Figure 6-1: Adsorption of sorbitol on 10 g H2-free Ru sponge at 25 °C. Initial loading of 5 mg (27.5 umol) and 10 mg (55 pmol). ...................................... 46 Figure 6-2: Adsorption of sorbitol at 25 °C on 10 g Ru sponge using stepwise dosing (Exp 20). Concentration of other species is in sorbitol equivalents. .48 Figure 6—3: Adsorption of glucose at 25 °C on 10 g Ru sponge using stepwise dosing (Exp 23). Concentration of other species is in glucose equivalents..49 Figure 6-4: Effects of H2 addition on adsorbed glucose at 25 °C on 10 g Ru sponge afler stepwise dosing (Exp 23) ........................................................ 50 Figure 6-5: Adsorption of gluconic acid at 25 °C on 10 g Ru sponge using stepwise dosing (Exp 25) ............................................................................. 51 Figure 6-6: Adsorption of glycerol at 25 °C on 10 g Ru sponge comparing stepwise (Exp 24) and fixed dosing (Exp 26). Left solution concentration; right quantity adsorbed. ........................................................ 52 Figure 6-7: Adsorption of GO at 25 °C using (A) fixed dosing: maximum initial concentration vs. total adsorbed, and (8) final solution concentration vs. amount adsorbed. ........................................................................................ 53 Figure 6-8: Adsorption of GO at 25 °C (Exp 67) on Ru sponge prepared with excess water. For clarity EG and PG concentrations are plotted at 10X ...... 54 Figure 6-9: Adsorption of GO at 25 °C (Exp 68) on Ru sponge prepared by reduction on the Micromeritics ASAP 2010 instrument. For clarity EG and PG concentrations are plotted at 10X. ......................................................... 55 Figure 6-10: Comparison of adsorption of GO at 25 °C on H2-saturated (Exp 28) and H2-free (Exp 39) Ru sponge (10 g). ...................................................... 56 Figure 6-11: Effects of higher GO loading on adsorption at 25 °C on H2-saturated (Exp 100) and H2-free (Exp 93) Ru sponge (20 g). ...................................... 56 Figure 6-12: Adsorption of G0 on H2-saturated Ru sponge (10 9) under argon at 50 °C (Exp 30). ............................................................................................ 57 Figure 6-13: Temperature dependence of GO adsorption. (50 pmol GO loading, 20 g Ru sponge). Inset shows magnification of upper region of graph. ....... 58 Figure 6-14: Glycerol decomposition products vs temperature. Multiple experiments are included at 25 and 60 °C. .................................................. 60 xi Figure 6-15: Adsorption of G0 (50 pmol loading) at 80 °C on 20 g Ru sponge (Exp 96) showing the effect of adding H2 and increasing temperature to 200 °C. Quantity GO adsorbed shown on lower graph, solution concentrations on upper graph. Concentrations of E6 and PG are plotted on right axis. ......... 61 Figure 6-16: Adsorption of PG at 25 °C on 10 g Ru sponge comparing loading and H2 saturation. ........................................................................................ 62 Figure 6-17: Adsorption of EG (80.6 pmol) on 10 g Ru sponge at 25 °C. ........... 63 Figure 6-18: Adsorption of 1,3—PDO (50 umol) on 20 g Ru sponge at 25 °C. ..... 63 Figure 6-19: Adsorption of PG followed by GO on 10 g Ru sponge at 25°C. A Exp 42, initial PG loading of 6.6 pmol; and B: Exp 44, initial PG loading of 66 pmol. GO loading of 4.34 umol follows PG loading ...................................... 65 Figure 6-20: Experiment 45, adsorption of GO (6.5 umol) followed by PG (6.5 pmol) at 25 °C on 10 g Ru sponge ............................................................... 66 Figure 6-21: Experiment 49, adsorption of G0 (2.2 pmol) and PG (3.3 pmol) at 25 °C on 10 g Ru sponge ............................................................................. 66 Figure 6-22: Adsorption of IPA (50 umol) at 25 °C on 20 g Ru sponge (Exp 109) showing effect of addition of H2 .................................................................... 67 Figure 6-23: Adsorption of IPA (50 pmol) at 25 °C on 20 9 H2 saturated Ru sponge (Exp 111) under 20 psi H2. .............................................................. 68 Figure 6-24: Competitive adsorption of GO and IPA (50 pmol each) at 25 °C on 20 g Ru sponge (Exp 112) showing the effect of adding H2 ......................... 69 Figure 6-25: Adsorption of acetone (50 umol) at 25 °C on 20 g Ru sponge (Exp 115) showing effect of addition of H2. .......................................................... 70 Figure 6-26: Adsorption (50 pmol each) of acetone after GO adsorption at 25 °C on 20 g Ru sponge (Exp 116) showing the formation of IPA. ...................... 71 Figure 6-27: Comparison of acetone adsorption (50 pmol) at 25 °C on 20 g Ru sponge (Exp 116, 115) Showing the effects of pre-saturation with GO. Open symbols are from presaturated experiment .................................................. 72 xii Figure 6-28: Adsorption (50 pmol loading) of n-propanol (Exp 121) and ethanol (Exp 114) at 25 °C on 20 g Ru sponge. Acetic acid (AA) and PA are formed. ..................................................................................................................... 73 Figure 6-29: Adsorption of n-prOH (50 umol/system) at 25 °C on 209 Ru sponge (Exp 121) showing effect of addition of H2 and increase in temperature to 50 °C. Gas C in legend refers to total gaseous carbon. .................................... 74 Figure 6-30: Concentration dependence of LA adsorption at 25 °C on H2 free Ru surface for different initial loadings. (Exp 54, 56, 57, 58) ............................. 75 Figure 6-31: H2 saturation effects on adsorption of LA at 25 °C on 10 g Ru sponge metal (Exp 54,55*,56,60*). *denotes experiments on hydrogen saturated surface. ........................................................................................ 76 Figure 6-32: Predicted vs. experimental pH as function of initial loading (umol). 77 Figure 6-33: Adsorption of equimolar mixture of LA and PA (120 umol/system) at 25 °C on 20 g of H2-free Ru sponge. (Exp 70) ............................................. 78 Figure 7-1: Experimental and predicted a) Langmuir and b) Freundlich adsorption isotherms for PG and G0 at 25°C on ROX carbon. ..................................... 88 Figure 7-2: Adsorption of PG and GO on ROX carbon on a mass basis. ........... 90 Figure 7-3: Temperature dependence of Langmuir equilibrium constant KA for GO and PG on ROX (A) and 3310 (B) carbon. ............................................ 93 Figure 7-4: Experimental vs predicted adsorption of PG and G0 on ROX carbon and 3310 carbon for all data. ....................................................................... 96 Figure 7-5: Experimental and predicted Langmuir isotherms for mixtures of GO and PG on ROX carbon at 25 °C and 0.05 M, 0.3 M and 0.5 M total species concentration. Abscissa is fraction of GO in the total species (GO + PG) present. Lines are the predicted isotherms. ................................................. 97 Figure 7-6: Experimental and predicted Langmuir isotherms for mixture of GO and PG on ROX carbon at 0.5 M total species concentration and 80 °C, 120 xiii °C, and 160 °C. Abscissa is fraction of GO in the total species (G0 + PG) present. Lines are the predicted isotherms. ................................................. 97 Figure 7-7: Extended Langmuir model predictions of a) GO and b) PG pore concentrations in ROX carbon at 180 °C. Symbols represent trajectory of 1.0 M GO conversion to PG. ........................................................................ 99 Figure 7-8: Experimental and best fit data for isothermal adsorption of n-propanol (n-prOH), ethanol, 1,3-propanediol (1 ,3-PDO), and ethylene glycol (EG) on 3310 carbon at 25 °C. (A-Langmuir, B-Fruendlich) .................................... 102 Figure 7-9: Experimental and best fit data for lsothennal adsorption of formic, acetic, and propanoic acid on 3310 carbon at 25 °C. (A-Langmuir, B- Fruendlich) ................................................................................................. 103 Figure 7-10: Experimental and best fit data for isothermal adsorption of glycolic and lactic acid on 3310 carbon at 25 °C. (A-Langmuir, B-Fruendlich) ....... 104 Figure 7-11: lsotherrnal adsorption of equimolar mixtures of PA+LA (A) and PA+PG (B) on 3310 carbon at 25 °C. ........................................................ 106 Figure 7-12: Isothermal Adsorption of equimolar mixtures of PA+PG (A) and LA+PG (B) on 3310 carbon at 25 °C .......................................................... 106 Figure 7-13 : Low concentration adsorption deviation from Langmuir isotherm at low concentration for LA, PA, PG, and n-prOH on 3310 carbon at 25°C. .. 108 Figure 7-14: Experimental data and WDL fit for the isothermal adsorption of PA, LA, n-prOH and, PG on 3310 carbon at 25 °C ........................................... 112 Figure 7-15: Variable temperature adsorption of 1000 mM LA on 3310 carbon. ................................................................................................................... 1 1 3 Figure 7-16: Experimental vs predicted single component variable temperature adsorption PA, LA, n-prOH and, PG on 3310 carbon from 25-160°C ........ 113 Figure 7-17: Experimental and predicted DWL isotherms for mixture of LA and PA on 3310 carbon at 0.5 M total species concentration and 25 °C, 80 °C, xiv and 150 °C. Abscissa is fraction of LA in the total species (LA + PA) present. Lines are the predicted isotherms. ............................................................. 114 Figure 7-18: Adsorption isotherm for equimolar mixture of LA and PG on 3310 carbon at 25 °C .......................................................................................... 115 Figure 7-19: Adsorption isotherm for LA and PG on 3310 carbon at 25 °C ...... 116 Figure 7-20: Experimental and predicted DWL isotherms for mixture of LA and n- prOH on 3310 carbon at 0.5 M total species concentration and 30 °C, 80 °C, and 150 °C. Abscissa is fraction of n-prOH in the total species (LA + n—prOH) present. Lines are the predicted isotherms. ............................................... 116 Figure 7-21: Experimental and predicted DWL isotherms for mixture of PA and n- prOH on 3310 carbon at 0.5 M total species concentration and 30 °C, 80 °C, and 150 °C. Abscissa is fraction of n-prOH in the total species (PA + n- prOH) present. Lines are the predicted isotherms. .................................... 117 Figure 7-22: Experimental and predicted DWL isotherms for mixture of PA and PG on 3310 carbon at 0.5 M total species concentration and 30 °C, 80 °C, and 150 °C. Abscissa is fraction of PG in the total species (PA + PG) present. Lines are the predicted isotherms. ............................................... 118 Figure 7-23: DWL predicted adsorptions of PA as percentage of actual adsorbed on 3310 carbon at initial concentration from 10 mM to 2000 mM. ............. 118 Figure 7-24: Experimental vs predicted adsorption of PA, LA, n-prOH and, PG on 3310 carbon in binary mixtures only. ......................................................... 119 Figure 7-25: Apparent species adsorption in control experiments without carbon showing the need to correct for vapor fraction of n-prOH. (note data for 150 °C after 14 hours is plotted at 160°C for clarity.) ........................................ 120 Figure B-1: Experimental and predicted Langmuir isotherms for mixture of GO and PG on 3310 carbon at 0.5 M total species concentration and 80°C, 120°C, and 160°C. Abscissa is fraction of GO in the total species (GO + PG) present. Lines are the predicted isotherms. ............................................... 139 Figure B-2: Formula matrix calculation of number of reactions needed. ........... 139 Figure B-3: Matrix of coefficients for ODE’s ...................................................... 140 LIST OF TABLES Table 3-1: Retention times and relative response factors on a weight basis. ..... 23 Table 3—2: Calculated values for a hypothetical adsorption experiment on Ru sponge. ........................................................................................................ 27 Table 41: Physical characterization of Ru sponge. ............................................ 30 Table 4-2: Gas chemisorption on Ru sponge ...................................................... 32 Table 5-1: Assumptions made for H2-D2O exchange model ............................... 38 Table 5-2: Species included in H2, 020 exchange model ................................... 39 Table 5-3: Optimized parameters for experiment 43B and 430 .......................... 43 Table 6-1: Contrasting adsorption properties on Ru sponge of organic acid and polyol. .......................................................................................................... 79 Table 7-1: Carbon characterization by N2 adsorption at 78 K ............................ 81 Table 7-2: General notation used for modeling adsorption on activated carbon.85 Table 7-3: Summary of experiments preformed during the study of aqueous phase adsorption on activated carbon. ........................................................ 86 Table 7-4: Langmuir and Freundlich isotherm coefficients for GO and PG on ROX carbon. ................................................................................................ 89 Table 7-5: Comparison of Langmuir isotherm coefficients for ROX carbon and 3310 carbon. ................................................................................................ 92 Table 7-6: Langmuir isotherm coefficients for GO and PG on ROX carbon ........ 94 Table 7-7: Optimized isothermal adsorption parameters for the DWL model....110 Table 7-8: Optimized isothermal adsorption parameters for the DWL model....111 Table 7-9: Percentage range of optimized parameters whose objective function is within 0.5% of the “best fit” ......................................................................... 112 Table B-1: Unique rate terms for elementary reactions ..................................... 140 xvii 1 INTRODUCTION The Miller group at Michigan State University focuses on the evolving technology of renewable resource refining. We seek to produce commodity chemicals from readily available bio-based feedstocks. Recent group members studied itaconic acid production,‘ lactic acid hydrogenation to propylene glycolz, stereoretentive amino alcohol synthesis3, the processing of sugar alcohols to value-added polyols" 5, and the hydrogenation of glycerol to propylene glycol.‘3 Our current focus is to understand the mechanism of aqueous hydrogenolysislhydrogenation of bio—based intermediates (polyols, organic acids, amino acids, alkyl amines, amides, ...) over noble metals. For example, in glycerol hydrogenolysis the first step in the proposed mechanism (Scheme 1-1) is the dehydrogenation and adsorption of the substrate as an aldehyde-like intermediate. Although the exact mechanism of adsorption is not known, we seek to probe interactions between substrate molecules and the metal catalyst surface to better understand binding energy, adsorbed structure, and competition for surface sites. Scheme 1-1: Proposed sequence of steps for GO hydrogenolysis to EG, PG and LA OH -H2 Hovk/OHv—T: :—H2 Ho\)\/k 60 GA retro [1 aldol OH ke. t /O 9' 20C,OH Glycolcldehyde Ethane-1 ,2-diol + Formaldehyde -H2 11 +H2 OH bOH . -‘3c k.e.t.= keto-enol EG tautomerization 1.1 Significance =MHO\i',OH=—‘ ket HO\/DI:K/OH Trioce encdiol +H20' —H2O OH k.e.t. 0 / T‘ N0 PV enol ‘5 PV $6 -2H2fl +2H2 Global energy demands keep increasing, fueling development of alternative renewable energy supplies and bio-based chemical intermediates. Aqueous phase processing and/or conversion of these bio-based materials is becoming important because of low environmental impact. Unfortunately, little is known about aqueous phase processing/conversion of bio-based materials like glycerol. Many of these processes use metal catalysts which follow the general mechanism in Scheme 1-2. The first step is adsorption of reactant (A) onto the catalyst metal site (M). Even if this is not the rate determining step it is still involved in the reaction kinetics. Understanding the mechanism of this adsorption in aqueous media is important. The other steps are conversion of reactant to product (B) on the metal and then product desorption from metal Scheme 1-2 A+ M——* A*-M (+C) -——> B*-M ——> B+ M For example, the side product of biodiesel formation is glycerol, which has low fuel value and limited industrial uses. On the other hand, the conversion of glycerol to propylene glycol is a good value adding process because propylene glycol has much greater industrial potential as a solvent, heat-transfer fluid, and as a chemical feed stock. Optimizing the conversion of glycerol to propylene glycol is thus an important goal, one which would greatly benefit from a detailed understanding of the reaction mechanism. 1.2 Objectives Understanding the adsorptive properties on active metals and supports of glycerol, glycols, organic acids, and related substrates derivable from renewable resources will give insight into the reactivities of our catalysts. To develop such an understanding, the following objectives need to be accomplished. 1. Develop a system to measure the adsorption of bio-based substrates on metal surfaces under elevated pressures and temperatures. 2. Study the adsorption of glycerol, propylene glycol, ethylene glycol and others individually and in mixtures on ruthenium sponge metal. 3. Study the adsorption of these same species on the carbon support and develop a model for this adsorption. 4. Compare the reactivity and selectivity between supported and unsupported ruthenium catalysts for conversion of selected biorenewable feed stock species, with emphasis on correlating reactivity with the adsorption results. 2 LITERATURE REVIEW 2.1 Hydrogenolysis of glycerol The hydrogenolysis and degradation of polyols has been studied extensively. The dissertation of Mike Shafer from our group has a large collection of references and patents.5 Selected patents dealing with hydrogenolysis of glycerol include Crabtree’s 2005 patent7 that mainly focuses on hydrogenolysis of sugar alcohols, but also includes glycerol. Their system uses an organic phosphine and ruthenium or osmium. The use of organic phosphine alone makes this impractical. Suppes’ 2005 patent“ has a more practical approach and will be discussed below. Glycerol hydrogenolysis over ruthenium on carbon (Ru/C) has been studied by several groups. Kovacs et al. in our group did detailed qualitative mechanistic studies using deuterium labeling.6 Lahr and Shanks focused on kinetics under neutral to basic conditionsg' 1° while Tomishige et al. looked at addition of Amberlyst 15, a strong cationic resin.11 Suppes et al.12 and Rosier et al.” looked at a range of metal catalysts while Perosa and Tundo used Raney nickel.“ 2.1.1 Kovacs’s et al. Kovacs et al.6 propose a mechanistic pathway for catalytic hydrogenolysis of glycerol to propylene glycol, ethylene glycol, and lactic acid over ruthenium on carbon (Scheme 1-1). The first step is dehydrogenation of glycerol to a glyceraldehyde analog, followed by one of two paths: (1) cleavage to the enediol of hydroxyacetaldehyde, which, with or without tautomerization, is reduced to ethylene glycol, or (2) keto-enol tautomerization to the enediol, which then dehydrates and tautomerizes to pymvaldehyde. Subsequent hydrogenation or water addition and rearrangement leads to propylene glycol or lactic acid, respectively. The presence of base promotes enolization and dehydration of the glyceraldehyde, favoring propylene glycol and lactate formation. Kovacs et al. substantiated this reaction mechanism by conversion, isotopic labeling, reaction of proposed intermediates, and computational molecular modeling studies. 2.1.2 Lahr and Shanks Lahr and Shanks studied the hydrogenolysis of glycerol (GO) over 5% ruthenium on carbon with emphasis on understanding the reaction mechanism. As a control, they looked at the degradation of propylene glycol (PG) and ethylene glycol (EG).1o The degradation of both EG and PG were independent of solution concentration (i.e. zero order), while initial rate data for GO conversion gave a reaction order of 0.45 to 1. They also reported that the presence of EG inhibited the conversion of GO, while PG had no effect. Studies of mixtures of EG and PG suggested that EG forms a stronger bond than PG with the Ru surface, thus explaining why EG inhibits GO conversion. Based on these results, they developed a reaction model based on Langmuir—Hinshelwood kinetics where the polyol adsorbs on Ru, undergoes dehydrogenation to form an aldehyde intermediate, desorbs into the solution, and then undergoes degradation with the base. They proposed that the rate limiting step is the dehydrogenation of the adsorbed polyol. In 2005, Lahr and Shanks expanded on this system by studying the effects of sulfur and temperature.9 The degradation rates of PG and EG are similar when they are reacted individually. The apparent activation energy and the relative reaction rates were unaffected by the addition of sulfur. The degradation rate becomes immeasurable at SIRu ratios >0.5. The addition of sulfur only seems to block the active sites and does not alter the degradation chemistry. The relative degradation rate of EGzPG was 4:1 in an equimolar mixture and was unaffected by temperature. This indicates strong competition between EG and PG on the metal surface. For the reaction of glycerol over Ru exposed to sulfur, the apparent activation energy increased while the absolute rate decreased greatly. The selectivity to PG increased with increase in sulfur concentration on the catalyst. l‘5. Shanks proposes that there This effect was also reported by Montassier et. a are two different pathway to PG; the surface dehydration of glyceraldehyde becomes more favored than the solution phase dehydration reaction upon addition of sulfur. 2.1.3 Tomishige et al. In 2005, Tomishige and coworkers11 tested the hydrogenolysis of GO over Ru, Pt, Pd, and Rh on carbon without any acid or base at 180 °C and 80 bar for 10 hours. Under these conditions, the RuIC gave 6.3% conversion with 18% selectivity to PG while the other metals gave conversions of 1% or less but with higher selectivity. They studied RuIC with a range of solid acids of which Amberlyst 15 gave the best conversion, 40% at 140 °C and 80 bar H2. Selectivity to PG was low at 43%. Decreasing the temperature to 120 °C did improve the selectivity to PG (60%) but conversion was only 33% after 40 hours. Curiously, the use of sulfuric acid resulted in only 3.2% conversion under the conditions that gave 43% with Amberlyst 15. They suggest that H2804 interacts with Ru surface sites, perhaps being reduced to sulfide that adsorbs and blocks GO access. 2.1.4 Suppes et al. Suppes et al.12 focused on the use of copper-chromite as a hydrogenolysis catalyst because their reactions with Ru, Pd, Pt, Ni, and Cu catalysts showed that copper-chromite gave 55% conversion and 85% selectivity after 24 hours at the following conditions: 80% GO in solution; 5% bulk copper- chromite; 200 °C, and 200 psi H2. They propose a two step mechanism where glycerol is dehydrated to acetol, and acetol is reduced to PG. They filed a patent8 on a two step process where glycerol is dehydrated under reduced pressure to form acetol. This acetol is then hydrogenated to PG at the same temperatures with the addition of H2 (200psi). No mechanistic details were suggested for the dehydration. All experiments were done in a batch reactor and only conversion of GO and selectivity to PG after 24 hours was reported. Solution pH and the possible presence of lactic, acetic, formic, and other acids was not mentioned. This is 8 somewhat troubling, because they also report that the catalyst loses activity before the reaction goes to completion. They performed a parametric study to understand the effects of catalyst reduction temperature, catalyst loading, reaction temperature, hydrogen pressure, and initial water content. A reduction temperature of 300 °C gave the best conversion and selectivity. Conversion increased with catalyst loading, but selectivity had a maximum of 85% at 5% loading. They speculate that higher loadings result in increases in the decomposition of the product. The effects of temperature follow the same trend as the catalyst loading with a maximum at 200 °C. For these cases, it would be beneficial to have the selectivity and conversion at different times in the reaction. Increasing hydrogen pressure from 50 to 300 psi resulted in a steady increase in the conversion and selectivity. Conversion was 65% with a selectivity of 90% at 300 psi, the maximum value they examined. Decreasing the initial water content increased the conversion but the selectivity went through a maximum of 85% at 20% water. In all cases, they reported that, except for an H2 pressure of 300 psi, the selectivity decreased when the conversion was >55%. This might be a clear indication that PG degradation is important, as reported by Shanks et al.10 2.1.5 Rosier et al. Rosier et al.13 looked at hydrogenolysis of ~20 wt% GO in water over CuOIZnO, PdIC, RhIC catalysts at 180 °C and 80 bar H2. Their conversions were generally <30% after 168 hours and PG selectivities <70%. Rth with H2WO4 gave best selectivity to PG. In switching from water to sulfolane as solvent, the 1,3-propanediol (1,3 PDO) was favored over PG by 2:1, but the 1,3—PDO 9 selectivity was only 12%. The observed rate and selectivity changed if the stainless steel autoclave was lined with Teflon. The lined reactor showed conversions ~twice that of the unlined reactor, and the 1,3-PDOIPG ratio increased from 0.1 to 0.5. Metal leaching was confirmed by testing the reaction mixtures. The addition of iron salts decreased the conversion and increased the 1,3-PDOIPG ratio. The addition of copper salts drastically reduced the conversion while still increasing the 1,3-PDOIPG ratio. They proposed a reaction scheme that includes several different pathways to PG and 1,3-PDO. The complexation of the 1,2- or the 1,3-OH groups to the metal and subsequent hydrogenolysis of the “free” hydroxide is proposed. 2.1.6 Perosa and Tundo Perosa and Tunclo14 used wet Raney Ni as catalyst and neat glycerol at high loading (29 Ni to 89 GO) at 190 °C and 10 atm H2, and obtained 63% GO conversion and selectivity to PG of 77%. The only byproducts that they reported are CO2 and ethanol, identified by GC-MS and NMR respectively. CO2 was identified by GC-MS, but quantities were based on mass loss. Addition of 2 g tetrabutylphosphonium bromide or (C6H13)3014H29PCI decreased the conversion and selectivity. It should be noted that these ionic liquids created a two-phase system at reaction conditions. The lack of 1,3—propanediol was used as evidence that it was not formed in the reaction scheme, but the absence of 1,3-PDO could also result from its rapid degradation as shown by Tomishige". 10 2.2 Aqueous phase adsorption on activated carbon The literature is filled with adsorption studies, but very few deal with aqueous phase adsorption of small molecules on solid adsorbents. Most deal with adsorption of large bio molecules which are relatively easy to detect and will not be discussed in this paper. An extremely useful source is http://adsorption.org/awm/ by Adam W. Marczewski, Ph.D. This site has a great general overview of adsorption and detailed model descriptions and references. Several good reviews exist“"18 that cover general solution phase adsorption. For modeling single species adsorption, several common isotherm models are used including Langmuir, Freundlich, and Sips (Generalized Freundlich). Equations for these isotherms are defined by Marczewski and are shown in Equations 2-1, 2-2, and 2-3. . q kC Lan m r =-—'"———— 2-1 9 UI q H kc ( ) - 1/ Fruendlich q = kFC " (2-2) % Sips q = i (2.3) 1+kC% The Langmuir isotherm is based on monolayer coverage of surface sites that are energetically identical. The Fruendlich isotherm assumes that adsorption energy increases as the quantity adsorbed increases, but there is no maximum coverage. The SIPS model combines both Langmuir and Fruendlich isotherms to 11 allow the system to have a maximum coverage and a variable energy of adsorption. Several recent studies have examined small molecule adsorption on activated carbon '9‘” and on polymer supports“ 324“. Of particular interest is Moon’s28 work on adsorption of sugars on activated carbon. They found that the adsorption could be modeled using either Fruendlich or SIPS isotherms, including binary adsorption using single component parameters. The adsorption of disaccharides was greater than any monosaccharide, with maltose adsorbing to the largest extent. 2.3 Species adsorbed on metal surfaces from liquid phase The literature is filled with studies of gas adsorption onto metal surfaces, but these are not generally relevant to our work. Several spectroscopic methods exist for the study of adsorption from the liquid phase to active metal surfaces. The most promising is attenuated total internal reflection (ATR-IR), which has been used in our labs but not in this work. A general overview of ATR-IR methods and references applicable to the overall thrust of research on hydrogenolysis is provided but will not be discussed?“2 Surface-enhanced Raman scattering (SERS) is also used to examine metal adsorbate interactions. Several good reviews exist but are again outside of the scope of this work.“47 X-ray photoelectron spectroscopy (XPS) is extensively used to study surface composition and has been applied to catalysis but is not as applicable for 12 in situ measurements, particularly liquids, because it is a high vacuum technique. Several reviews on this method are available.“56 Single crystal surfaces are extensively used in gas phase studies and have also been applied in liquid phase environments. One particularly interesting in-situ study stands out.57 Wang et al. used cyclic voltammetry to oxidize and reduce Ru (001) surface while simultaneously measuring x-ray reflectance. They propose that water is chemisorbed on the metal, possibly as a hydroxide. 2.4 Indirect differential detection Indirectly detecting quantities adsorbed on solid materials by measuring differences in liquid-phase concentrations under controlled conditions is one of the oldest and simplest methods. Batch reactors are widely used, either with in situ monitoring or periodic sampling. Shallow bed adsorption is practical when quantities adsorbed are significant."3 No literature has been found covering the adsorption of small organic molecules on bulk sponge metals. Only problems with adsorption on exposed metal in HPLC system are mentioned, and these do not look at quantifying the actual species adsorption. 13 3 NOVEL FLOW MICROREACTOR FOR SPECIES ADSORPTION ON METAL CATALYSTS 3.1 Design strategy One challenge with measuring species adsorbed on the supported metal catalyst is that the ratio of support area to metal surface area is large. In the case of 5% Ru/C, this ratio is >400:1. This large ratio makes it impossible to distinguish adsorption on the metal from that on the support; for mat reason we have to look at adsorption on pure metals only. Reactivity, particle size, and surface area are some of the criteria used in selecting the metal to use. In the case of thhenium, three forms are commercially available, solid (ingot, bar, foil), sponge metal, and powder (Ru black). Ru black has the largest surface area (~10-15 m2/g), unfortunately the particle size is too small to be handled easily and would result in metal loss or plugged filters if used in a flow system. This leaves ruthenium sponge metal which has a low surface area (~0.2 mzlg Ru). This low surface area means that the quantity of substrate adsorbed is extremely small (<1mglg Ru). Maximizing the metal to solution volume ratio is therefore critical; this and the high cost of Ru metal has led us to develop a reactor system that meets the following design parameters: 1. Low total volume <15 ml 2. Metal capacity up to 20 g 3. Reactor operating temperature range 20 - 200 °C 4. Solid liquid mixing and gas liquid mixing 14 5. Addition of two different gases at pressures up to 400 psi 6. Online sampling without disturbing the system 7. Option to flush system with different liquids The reactor system shown in Figure 3-1 meets these criteria. The system consists of an HPLC pump, tubular reactor, inline sampling system, a “tank" (gas liquid inter-phase), and valving to isolate the reactor. HPIC vent systan I 8 —0[ Cooling coil Oil bath (25 -200 °C) t: 8 5 Ta 0 . .3 Er": :5 Heat 5 exchanger ‘3 (coil) 82 Figure 3-1: Recirculating reactor system A fluidized bed reactor was chosen to allow large solid to liquid ratio with good mixing. Initially, a simple 10 mm ID column was used. Unfortunately, in this reactor configuration little axial mixing was observed at 3-6 mllmin. The relatively large particles and high density of ruthenium sponge metal prevented fiuidization. Calculations showed that at 25 °C the bed should be fluidized at ~2 mllmin while 15 at 100 °C more than 6 mllmin was needed. The general solution to this problem is a tapered reactor that allows a velocity gradient Internally tapering a small 316 stainless steel reactor or insert proved to be a technical barrier. Machining a conical insert, on the other hand, was relatively simple, and thus this route was taken to produce the reactor. 3.2 System description The circulation pump is a Bio-Rad 1350 HPLC pump with the intake modified with a simple t-fitting to decrease the volume and allow operation under pressure. The output from the pump goes through a home-made pressure damper consisting of a coil of flattened 1116 SS tubing. The pressure damper is needed to prevent excess pressure buildup during valve switching. 10 mm ‘216 mm Figure 3-2: Dimensions of 316 stainless steel reactor insert. I 100mm I l 20mm The reactor consists of a 10 mm x 100 mm SS HPLC column with 2.0 pm frits at the ends. The machined conical insert (Figure 3-2) is placed in the column to create a semi-fluidized bed in the reactor. This was needed for mixing purposes. A coil of SS tubing was placed before the reactor to act as a heat exchanger to preheat the liquid entering the reactor. A water cooled heat exchanger is placed after the reactor to protect the pump and other systems. 16 The “tank” (where the reactor gas-liquid interface is located) is made from 12mm thick-walled Pyrex. lt facilitates dissolution of gases into solution, and acts as a sight glass, allowing the fluid level in the system to be monitored. Valves and fittings are added to allow monitoring and manipulation of the head space gases. It was found that no additional mixing was needed in the tank. An additional HPLC “sample injector" (Figure 3-1) valve allows the introduction of substrates to the system. The in-Iine sampling system consists of two Rheodyne 7010 HPLC valves configured as shown in Figure 3-3 so as to allow the sampling loop to be flushed with water and the sample to be injected on a Bio-Rad HPX87-H column followed by two detectors; a Perkin-Elmer LC 90 UV detector (210 nm) and a Waters 410 RI detector. It should be noted that the HPLC system can detect dissolved gases and distinguish between He, H2, and Ar, with the response being directly proportional to the headspace pressure. The ability to detect dissolved gas could be used to measure hydrogen consumption on a flow through system. 17 Sample loop From RBR HPLC . T0 RBR Load Mode Waste Sample loop To column V HPLC mo Water To column Wash Mode Figure 3-3: Configuration of in-line sampling system 3.3 System evaluation 3.3.1 System mixing Pulse experiments to evaluate system mixing were performed using lactic acid with the UV detector placed in—line. A 100 pl pulse of 1.0 M lactic acid was 18 added to the recirculating fluid that was flowing at 6 mllmin. Figure 3-4 shows that there is sufficient mixing in the system, provided the time scale of sampling once every 30 minutes. Even if the flow rate is set at 3 mllmin, there should be sufficient mixing for uniform sampling. 0.05 0.045 0.04 0.035 0.03 0.025 0.02 0.015 0.01 0.005 0 UV Adcorbcnce 0 2 4 6 8 10 Tune (min) Figure 3-4: Pulse attenuation at 6mllmin for recirculating system. 3.3.2 Volume determination of reactor, tank, loop, and system A 10 ml syringe was cut lengthwise and clipped around the tank (112’ Pyrex tube) to help with reading the level in the tank. The reading on this cover was calibrated by adding fixed volumes of water using a 3 ml syringe and recording readings. It was found that a reading of 3 corresponded to 2 ml while a reading of 0 was equivalent to 1 ml in the tank The volume of the reactor and loop were determined by first adding 200 pmol GO to the entire system and letting this recirculate overnight to allow GO to saturate the catalyst surface. At this point the system was sampled. Then the recirculating pump was stopped and the tank drained and refilled with 4 ml of HPLC-grade water. The pump was turned on and recirculated for 2 hours while 19 samples were taken at 30 minute intervals. The reactor was placed out of line and the dilution process was repeated. The system volume was then calculated using V=4ml C2 ; this gave a loop volume of 3.82 ml and a reactor volume I 2 of 5.61 ml. 3.3.3 HPLC calibration HPLC analysis was based on external calibration. Linear response from a Waters 410 differential Refractive Index (RI) detector was verified by running standard solutions containing either alcohols or organic acids at concentrations from 1 to 10 mg/ml. The standard solution of alcohols is referred to as HO—mix and contains: sorbitol (SO), glycerol (GO), ethylene glycol (EG), 1,3-propanediol (1 ,3-PG), methanol (MeOH), ethanol (EthOH), isopropanol (IPA), and n-propanol (n-prOH), while the standard acid solution referred to as acid-mix contains glycolic acid (GA), formic acid (FA), acetic acid (AA), and propanoic acid (PA). The RI detector response curves for each mixture are shown in Figure 3-5; this figure demonstrates the linear relationship between concentration and detector response. Because of this linearity, the response factor (RF) was calculated at each point for each compound. The RF had an uncertainty of :I:2% and did not significantly change with time, but did show a dependence on fluctuations in flow rate. Both mixtures were injected using a Waters 717 Autoinjector (10 til/sample) and also were injected through the reactor system injector to get the correlation between the two. Figure 3-6 shows a chromatogram of HO-mix (1 mg/ml) run through both the reactor system injector and the 717 Autoinjector. The peak 20 response ratio between the injectors is 4.7 and was the same for all compounds. This is slightly higher than the expected ratio of 4.0, because the reactor system has a 40 pl loop vs. 10 pl injections from the 717 Autoinjector. Examining the dilution due to sampling of a standard solution in the recirculating reactor, it was found that the actual injection loop was 43 pl. Therefore the actual amount that the 717 Autoinjector injects is less than 10 pl. ~+« Glycolic acid 18 I —x—formic acid 16 . +acetic acid Millions Millions —c— propanoic acid 14 12 4 2% £10 /’ 8 8 , //i s /“ 4 2 / 0 Concentrationfimglmg 10 0 éonccntrafionfimglms 10 Figure 3-5: Refractive index detector linear response curves for 10 pl injection of HO-mix and acid-rnix standard solutions. 21 reactor system 717 autoirjectorj 0 5 10 15 20 25 30 time (min) Figure 3-6: Comparison of auto injector vs. reactor system injector with HO—mix (Sorbitol, GO, EG, 1,3—PG, MeOH, EthOH, IPA, n-prOH) Standards were made at 1% wlw in water and were run using the 717 Autoinjector (10 pl/rnjection). Relative response factors (RRF), which are constant over all conditions, were calculated aid are reported relative to sorbitol on a weight basis ( Table 3-1); the average organic RRF reported was applied to any unknown peaks. The RRF were used to calculate the RF for each species in the reactor system. With each adsorption reaction, the initial loop concentration of the adsorbing species was used to verify the RF for sorbitol by the RRF for that species. 22 Table 3-1: Retention times and relative response factors on a weight basis. 9291mm! _§_§RTat ° BEE Oxalic acid 6.55 0.58 Sodium silicate 8.90 0.14 Gluconic acid 9.10 0.92 Glucose 9.22 1.08 Sorbitol 10.43 1.00 Glyceric acid 11.02 0.71 Boric acid 11.39 2.12 Glyceraldehyde 1 1 .55 0.78 Glycolic acid 12.42 0.64 Lactic acid 13.00 0.73 Glycerol 13.59 0.83 Formic acid 14.11 0.36 Acetic acid 15.41 0.46 Ethylene glycol 16.51 0.66 1 ,2-Propanediol 17.55 0.74 Propionic acid 17.94 0.58 1 ,3-Propanediol 18.07 0.71 Methanol 19.57 0.14 1 ,4-Butanediol 21 .97 0.83 Acetone 22.00 0.49 Ethanol 22.62 0.43 lsopropanol 24.50 0.59 n-Propanol 28.30 0.61 1-Butanol 35.50 0.56 average organic peak 0.60 3.4 Methods of operation 3.4.1 General description of experiment For most experiments, the reactor was charged with 10 g ruthenium sponge (-100 mesh, Aldrich); this Ru sponge was reused in most experiments. The system was flushed at 3 mllmin with argon purge and degassed HPLC- grade water for at least 10 minutes. This was followed by charging the system with ~12 ml of the same water. The system was set to recirculating mode and placed under 250 psi H2 to clean and reduce the catalyst metal surface. The reactor was heated to 200 °C for 2-4 hours. During this time the headspace was 23 exchanged with fresh hydrogen at least three times to remove any methane that was formed during reduction. The reactor was then either cooled to 25 °C while recirculating with H2 saturated water to generate the “H2 saturated surface,” or was flushed at 200 °C with the argon purge and degassed water for at least one hour, followed by cooling to 25 °C to generate the “H2 free surface”. With the reactor out of line, the loop was charged with water to 6.1 ml (~1/3 up the sight glass tank). At this point a sample was taken, followed by the introduction of the substrate of interest. After 30 min, the loop was sampled and the reactor placed in-Iine, followed by automated sampling every 30 min until the experiment was done. UV and RI detector data were collected for 30 minutes and stored for analysis. The level in the sight glass and the headspace pressure was recorded at several times during the experiment. 3.4.2 Data processing The following variables are using in data processing: Cn Concentration of solute in solution at sample n (pmollml) Vm, Total system liquid volume (ml) Va,“ Volume of reactor and heat exchangers (ml) Vino, Volume of loop part of reactor system (ml) Vim Volume of liquid in tank (ml) No Moles added to system (pmol) N... Moles in solution (pmol) Nlcakloet Total moles lost due to leaks (pmol) Nlnj loct Total moles lost due to injections (pmol) N.“ Total moles adsorbed (pmol) Pcount Corrected peak area counts Pcount..." Uncorrected peak area counts 24 RT Retention time (min) RTavg Average retention time (min) RF Response factor (counts/(mglml)) RRF Relative response factor A standard Excel workbook was created and several Visual Basic macros were written to automate the repetitive tasks involved in processing the large amount of data generated with each experiment. The amount adsorbed for a given substrate was calculated using Equation 3-1 and involves the conversion of raw UV and RI detector data into concentration data, finding the system volume, and keeping track of cumulative sampling and leak losses. Moles of Initial moles Moles still in Moles lost due substrate = loaded into - solution - Ito sampling andl (3.1) adsorbed the system leaks Raw UV and RI detector data from chromatograms of each sample were integrated using the Interactive Graphics part of Van'an Star Workstation v6.0 Software. For each sample, the retention time and area for each identified peak was transferred to a Microsoft Excel worksheet and the areas were corrected for fluctuations in retention times using Equation 3-2. RT,g — RT Acorr = A 1+ (34) 3V8 This RT dependent correction of the area was found empirically and is valid for at least 20% deviations in RT. This corrected area was used to calculate solution concentrations using the response factors given in Section 3.3.3. If the measured 25 concentration deviated by more than 5% from the predicted value in the initial sample with the reactor in loop mode, the RF values were sealed for that experiment to match the measured to the predicted concentration based on the actual amount of species introduced. For every identified compound in each sample, the number of moles in solution was given by C. Vim“ where Via.) = View + Vm + Vim, and C" = (HPLC counts) RF... In loop mode, Vm was omitted. Only a limited number of tank volume data points were recorded for each experiment; linear interpolation was used to find Vim for each sample. For each sample, the number of moles lost due to sampling for each identified compound was determined as the solution conmntration multiplied by the sample loop volume (43 pl). For leak losses, it was assumed that the concentration changed linearly between sampling times; therefore, the average concentration between the current and previous samples multiplied by the change in volume over this same interval gives the number of moles lost due to leakage. These two losses were summed and included in computing the substrate quantities adsorbed from the current sample. 26 Table 3-2: Calculated values for a hypothetical adsorption experiment on Ru sponge. Time Vtotal Cn Neel Ninj lost Nlecli lost chs (h.) (ml) (mM) ML (pmol) (pmol) ( mol 0.37 11.26 3.404 38.32 0.15 0.00 7.64 0.88 11.24 3.203 35.99 0.28 0.06 9.77 1.39 11.22 3.128 35.09 0.42 0.12 10.47 1.90 11.20 3.049 34.14 0.55 0.18 11.22 2.40 11.18 2.899 32.41 0.67 0.24 12.78 2.91 11.16 2.852 31.83 0.80 0.30 13.18 3.42 11.14 2.771 30.87 0.92 0.35 13.97 3.93 11.12 2.750 30.59 1.03 0.40 14.08 4.43 11.10 2.742 30.44 1.15 0.45 14.05 4.94 11.08 2.715 30.09 1.27 0.51 14.24 5.45 11.06 2.674 29.58 1.38 0.56 14.58 5.45 11.06 2.674 29.58 1.50 0.56 14.46 New“) = Vmi(i)*Cn(i); Ninjmfl) = Ninjlost(i'1) +.043*Cn(i); Nleak m0): Nleek lost“) 1' (Vtotai(i-1) - Vtorai(i))*(Cn(i-1)+ Cn(i))/2i Nedc“) = No - Neal“) ' Nlni lost“) ' Nleak lost“) The total amount adsorbed was based on the number of moles of carbon in each identified compound and reported as equivalent number of moles of the substrate of interest. Table 3-2 shows calculated values for the adsorption of one 27 component on Ru sponge to illustrate the calculations described above. It should be noted that for cases where products were formed, a negative adsorption value is reported. This is of course a physical impossibility, but expressing it in this manner simplifies the calculations. 28 4 CATALYST CHARACTERIZATION 4.1 Methods For this work, Ru sponge metal (product #267406, Lot# 14021 EB) from Sigma-Aldrich was used. Catalysts were characterized on a Micromeritics ASAP 2010 adsorption instrument, either in a standard analysis cell or by attaching the reactor vessel used in the recirculating batch reactor to the ASAP instrument. Experimental details for nitrogen BET surface area measurements of activated carbon are described in Section 7.1.1; the same procedure was used for Ru sponge metal. Attempted N2 BET measurements of catalyst in the reactor failed, most likely because of flow restrictions through the frit. General procedures for chemical adsorption experiments consist of pretreatment, initial adsorption analysis, evacuation, and repeat analysis. Pretreatment consisted of reducing the sample at 350 °C under a flow of H2 for 2 hours, evacuation for 1.5 hours at 360 °C, cooling to 35 °C under vacuum, and evacuation at 35 °C for 10 minutes followed by leak testing. All analyses were run at 35 °C and consisted of dosing known volumes of analysis gas to the sample chamber and recording the volume added vs. pressure. Following this initial adsorption, the sample chamber was evacuated at 35 °C for 60 minutes to remove physically adsorbed analysis gas, and then dosing was repeated as before. The volume vs. pressure data for each dosing was then plotted and the linear region (usually P>100 Torr) was extrapolated to P=0 to give the quantity of 29 gas adsorbed. The chemically adsorbed quantity of gas is the difference of volume adsorbed in the initial and repeat analysis. The surface area is calculated using the stoichiometry factor (2 for H2, 1 for CO) and the atomic cross-sectional area of Ru (0.0613 nm2). 4.2 Results Surface areas for both physical and chemical adsorption are listed in Table +1. In the table, “fresh” Ru sponge was taken directly from bottle and “used” Ru sponge is a collection of recovered catalyst used in a range of hydrogenolysis experiments with different substrates. Both fresh and used sponges were from the same lot number. Reactors 1 and 2 are mechanically equivalent duplicates used in these studies. Results show a BET surface area (~0.5 mzlg) that does not change upon using the Ru sponge, this area is consistent with results of prior work by Zhang et al.2 Table 4-1: Physical characterization of Ru sponge. Fresh Ru Used Ru Fresh Ru Used Ru Reactor 2 Reactor 1“” Total (N2 BET) 0561:0006 052710.004 - surface area (mzlg) Metal (H2) 0.23 0.12 0.21 0.15(0.13) surface Area (mzlg) Metal (CO) 0.21 surface area: (mzlg) (" Reactor 1 was charged with 89 used and 29 new Ru sponge. Number in parentheses is the amount corrected (by weight fraction) using the new Ru values from Reactor 2. Unlike the N2 BET, the surface areas found by H2 adsorption (0.23, 0.12 mzlg)» show a ~50% decline in the hydrogen active surface area after using the 30 Ru sponge for reaction. The H2 area measured for new Ru sponge in Reactor 2 (0.21 mzlg) is the same within experimental error to that found for new sponge using a standard analysis cell, confirming that using the reactor does not affect the H2 adsorption measurements. Reactor 1 was charged with some (2 9) new and some (8 9) old Ru sponge; the resulting adsorption area of 0.15 mzlg is expected based on the weight fractions for this sample. The adsorption of CO on new Ru sponge showed active Ru area of 0.21 m2/g, the same as found for H2. CO adsorption was not done on reactors for fear of poisoning the metal surface with CO. A set of gas adsorption experiments were run with different gases to compare different molecular adsorption properties (Table 4-2). Interactions of CO and H2 on Ru surfaces have been studied,“ but our goal was to see if there might be two different sites, one for CO and one for H. If there are two independent sites, than there should be little dependence on the order of adsorption. The presence of CO on the surface does not seem to block the adsorption of additional H2 (Exp A6), while the presence of H2 partially blocks the adsorption of CO (Exp A3). Blocking of the CO by hydrogen was reported by Peebles”. Subsequent addition of H2 (Exp A4) showed no significant adsorption. 31 Table 4-2: Gas chemisorption on Ru sponge. Volume Adsorbed: Surface Exp# Sample reduction Ema/98TH area:(mZLgL umolljg_ A1 N2 BET no 0.1146 0.4980 A2 H2 initial yes 0.1020 0.3350 4.55 A3 C0 after H2 no 0.1 100 0.1820 4.91 A4 H2 after H2 and CO no 0.0190 0.0640 0.85 A5 CO with reduction yes 0.1940 0.3200 8.66 A6 H2 after CO no 0.1040 0.3440 4.64 A7 N2 with reduction yes 0.0090 0.0149 0.40 A8 He with reduction yes 0.0005 0.0008 0.02 A9 Methanol attempt yes (~0.05) -— A10 H2 after methanol no 0.0034 0.01 12 0.15 A1 1 Methane yes 0.0306 0.0504 1.36 A12 Methane after H2 yes 0.0016 0.0026 0.07 A13 H2 after Methane yes 0.0450 0.1483 2.01 Methane (Exp A11) adsorbed in a smaller amount than H2; this could be a result of the larger size of methane, or it could be that methane adsorbs with full dissociation. It should be noted that methane did not adsorb on an H2 saturated surface (Exp A12). The adsorption of H2 (2.0 pmollg) after methane (Exp A13) show that there are still sites available. If methane adsorbed with full dissociation, little H2 adsorption would be expected, while if it were adsorbed as a methyl radical (H30) or Methylene (H2C:), then there should be sites for H2. To summarize the characterization of Ru sponge, fresh sponge has the same BET area as used Ru but the fresh sponge H2 area is twice that of used 32 sponge. CO adsorption does not affect the adsorption of H2, but H2 does affect CO adsorption. Several analyses were run to evaluate the possibility of measuring H2 areas after aqueous phase reduction. It was found that a hydrogen-saturated surface flushed with He (to remove water) and then evacuated at 25 °C gave inconsistent results. The adsorbed hydrogen came off the surface during the He flush and evacuation. Measuring gas phase adsorption after liquid phase reduction is not possible. 33 5 Hle20 EXCHANGE It has been observed that when H2 gas and liquid D20 are placed in a reactor over ruthenium catalyst, D and H atoms are exchanged. This was a problem when we were trying to determine the pathways for aqueous hydrogenolysis of polyols using H-D exchange, as under normal experimental conditions (100-150 °C, 500—1500 psi) the H-D ratios equilibrated on a time scale faster (<30 min.) than our reactions, resulting in a statistical distribution based on the overall H-D pool. At 25 °C and 10-200 psi, this exchange happens over several hours. At this time scale, it is possible to perform experiments to elucidate kinetic parameters for this reaction. 5.1 Experimental Two different sets of experiments were run. In the first set (Exp #43a-c), a 10 ml pressure vessel was charged with 2 ml of D2O, 0.1 g dry weight of catalyst, and hydrogen gas (125 psi). The head space was continually sampled through a capillary tube (0.05 mm ID. x 5 m) into a mass spectrometer. The second set of experiments consisted of H-D exchange using a 70 ml Parr reactor equipped with a magnetic stirring bar, gas connections, and sampling port. This reactor was charged with 10ml D2O, 0.1 g dry catalyst or 1.0 g Ru sponge, H2 (50 psi), and D2 (50 psi). The head space was continually sampled through a capillary tube (0.05 mm ID. x 5 m) into a mass spectrometer. A pressure drop of ~25 psi was observed during each experiment. An Ametek Dycor M100M Quadrupole Mass Spectrometer was used for the analysis of the headspace. A HD gas mixture was made by charging a 40 ml pressure vessel with ~O.5 g 5 wt% Ru/C, H2 (150psi), and D2 (150psi). After equilibrating this mixture for several days, the H2:HD:D2 ratios were assumed to be 1:2:1. Instmment response was examined with H2, D2, HD, methane, and Ar. The response ratios were inconsistent and seemed to have some dependence on the degassing and tuning of the instrument. Uncorrected Corrected 1 .7 1 , ,. a- 0.9 Ru/C 5% 0.9 Ru/C 5% 0.0 0.0 0.7 0.7 0.6 0.6 g 0.5 g 0.5 c: 0.4 l: 0.4 0.3 0.3 02 0.2 0.1 0.1 0 0 500 1 000 0 500 1 000 Tlmc (mln) 11m (mln) 0 500 1000 500 Tine (min) Tine (mln) Figure 5-1: Corrected and uncorrected data for H2 over D20 with 0.1 9 5% RulC (Exp 43B), and H2 over D20 with 0.1 9 5% PVC (Exp 43C) 35 After consultation with a technical expert from the mass spectrometer manufacturer, It was concluded that this inconsistency is expected because this instrument is designed and optimized for gases other then hydrogen. Thus no further attempts were made to calibrate and optimize the system. The H-D test mixture was analyzed either before or after each experiment to obtain the relative response factors (RF). The assumption was made that the relative responses were constant with respect to pressure even with a 50 psi pressure drop. Therefore only the ratios of H2, HD, and D2 are plotted. Comparison of raw and corrected data for Experiments 43B and 430 are shown in Figure 5-1. It can be seen that applying the RF correction has little effect on Experiment 43C, while there were significant changes in Experiment 438. Time (min) Figure 5-2: Incorporation of deuterium into the gas phase from (DO)ethanol over 0.1 g 5% Ru/C (10 ml reactor 100 psi H2 Exp 47) It can be seen in Figure 5-1 that, over a Pt catalyst surface, deuterium is rapidly incorporated into the gas phase, while over a Ru surface the process is slower, consistent with water strongly dissociating on the Pt surface while little water dissociates on a Ru surface. Dissociation of a O-H bond in the liquid phase 36 is further confirmed by deuterium incorporation into the gas phase when (DO)ethanol is placed over 5% RuIC under H2 gas (Exp 47, Figure 5-2). This H-O dissociation is slow, while H2 and D2 adsorption\desorption is rapid as seen in Experiment 50 (Figure 5-3). Looking at the data in Figure 5-3, it appears that deuterium incorporation in the gas phase is a lot slower than seen in Exp 438 (Figure 5-1), but this is deceptive because the headspace volume differs by a factor ~10, while the catalyst loadings were the same. The absolute rates should be the same; therefore the apparent rate should be slower with a larger headspace. 1 0.9 - ----------------------------------------------------------------------------- 0.8 « ----------------------------------------------------------------------------- 0.7 « ----------------------------------------------------------------------------- g 0.6 i E 0.5 - O 200 400 600 800 1000 Time (nin) Figure 5-3: H2 + D2 over D20 with 0.1 g 5% RuIC (50 psi each in 70 ml reactor, Exp 50) 5.2 Modeling The general approach to this system involved a Langmuir-type adsorption with a pseudo steady state approximation (PSSA). In general, the Langmuir 37 model looks at adsorption at equilibrium; we wanted to look at the kinetics of adsorption and desorption individually. The model equations are given as follows. k1 XYIaq) + 2Mls) ‘_kz— XMIS) + YMIS) combinations ofXandY k3 X=H Y=H XYO(|) 'I' 2M(s) :ZiLT XM(s) + YOMM §;3 :;3 4 X=D Y=D k5 H20“) + 020“) ? HDO(|) + DHO(|) Figure 54: Proposed reaction model. The kinetic model proposed is shown in Figure 5-4. Several simplifying assumptions were made and are listed in Table 5-1. Assumption #1 is somewhat unrealistic, because H or D is involved in all the reactions and it is generally accepted that isotopes should have some effect. Assumption #9 is also somewhat suspect. We have some evidence that there might be two different sites in aqueous phase reactions, one for hydrogen and one for oxygenated species such as water and organics. Some published calculations by Neurock'31 also suggest a two site model. Table 5-1: Assumptions made for H2-D2O exchange model There is no isotope effect. Dissolved gas concentrations are in equilibrium with gas phase. Only dissolved gases react. There are no transport limitations. Dissolved gases can only react with free sites. @QF‘SPNT‘ Liquid species can react with free sites and with other liquid species. 7. Liquid species are in equilibrium with other liquid species. 38 8. Migration on the metal surface is not considered 9. There is only one type of site on the metal for H-, D-, and 0- adsorption. 5.2.1 Approach to solutions All species involved in this system are shown in Table 5-2. Note that HD, DH and HDO, DHO, which physically are identical pairs, are treated as separate species; this was done to not have to take into account the probability of H or a D reacting. Table 5-2: Species included in H2, D20 exchange model Dissolved Gases: H2 H0 D2 Species on Metal: HM DM HOM DOM M (free sites) Liquid species: H2O HDO DHO D20 To determine the minimum number of reversible independent reactions needed to describe the isotope exchange, a procedure described by J. M. Haile62 was used. In Haile’s method, the formula matrix is generated and the matrix rank is then determined. Then the number of independent reversible reactions needed (Rn) is the number of compounds minus the rank of the formula matrix. Figure B-2 shows formula matrix and calculation to give Rn=9. The 9 reactions are shown in Figure 5-5. 39 H2(aq) + 2M(s) :2 2HM(s) Hem) + 2M(s)‘_:12_ HM(S) + ems) DH(aq) + 2M(s)‘_:21__ DM(s) + HM(s) 02(aq) + 2M(s) :12 2DM(s) H20(|) + 2M(s) HDO(|) + 2M(s) HM(S) + HOM(s) HM(s) + DOM(s) k4 k4 DHO(|) + 2M(s) DM(s) +HOM(s) DM(5) + DOM(5) k4 020“) + 2W8) k4 H200) + DzO(|):_—_::: HDO(|) + DHO(|) Figure 5-5: Reversible reaction model From these 9 reversible reactions, we get 18 elementary reactions, that are all either 2"6 or 3’d order. These reactions will have a rate (velocity) term of the form r=k[A][B] for 2"d order and r=k[A][B]2 for 3rd order. It should be noted that only availability of free metal sites (M) and not their spatial orientation has been included in the model. Because of this, dissociative adsorptions are third order reactions. These terms, when calculated for all 18 reactions, result in 17 unique rate terms shown in Table 84 To get the corresponding 13 ODE’s, one creates a matrix with the coexistence for each term as need to account for the species balance. This matrix is shown in Figure B-3. The vector of ODE’s % = 1; a; can be used as a function block for matlab’s ODE solver. The reason for using this matrix system is that the rate terms are calculated only once in the cycle. If the terms are added as terms in each individual ODE, they would be recalculated each time they are used. This method significantly speeds up the calculations. 40 The ODE solver used from MATLAB was ODE15S. Initially ODE45, a 4th order Runge-Kutta method, was used, but this problem is too stiff for that solver and tended to diverge. The initial conditions of the system were calculated from experimental data and were kept constant throughout this modeling exercise. All variables are listed at the top of the Matlab code. Parameters that were adjusted to observe the system behavior were: k1, Kr, Kg, Kl. The reference rate constant was R1, the forward rate of the “gas” side reactions. Kr is equal to k1lk3, the ratio between the “gas“ side and “liquid “side forward reactions rates, while Kg = k1lk2 and Kl =k3/k4, the equilibrium constants for the “gas“ and “liquid” side reactions, respectively. These values were adjusted to get the model response to be close to experimental. The k1 value of ~1000 (Lzlmolzls) is within a few orders of magnitude of expected reactions with a surface. Generally, adsorption and desorption reactions are rapid and are not the rate limiting steps in catalytic systems. The value of k5 was set at 1000 (L/molls); this was large enough that the proton exchange in the water was close to instantaneous. The value of k5 was varied by a factor of 100 in both directions with no noticeable effect on the model predictions. 5.2.2 Parameter optimization For the experimental data, the raw data for Experiment 43b (5% RulC) and Exp 43c (5% Pth) were used. These data were not corrected for the differences in response factors. The only measurable quantities were the compositions in the gas phase. Dissolved gasses are assumed to be in 41 equilibrium with the head space gas and related by Henry’s law. For simplicity, gas phase concentrations were assumed to be the same as dissolved gas concentrations. The actual dissolved gas concentrations in the liquid are lower by a factor ~1000, which if used would raise the corresponding k values. Initial fitting was done by manually changing k1, Kr, K9, and KI to approximate the results shown experimentally. The final optimization was done using a random walk approach. For each cycle, one parameter was chose at random and increased or decreased by a random percentage, limited such that no change was greater than 10%. The ODE’s were then integrated over the same total time as the experiment, followed by calculation of the overall error using Equation 5-1. Error=2((C:2‘p —c;:'.>2 +(Cff; 42st ace; -C‘3I.>2 +(C3‘é’ -C3'2 2 (54) 0 If this error was lower than the previous cycle, the change was kept; it not, the old values were kept and the cycle was repeated. This is a “brute force” method but it is much simpler to implement than some of the gradient methods, and also it tends to avoid local minima if the changes are large. The possibility of an evolutionary model, in which all parameters are changed at once followed by some selection rules as to the diredion of the next change, was also examined. This method sounds like it might be a better way of doing the optimization if there are a lot of variables. The error calculation is simply the summing of all the squares of the individual errors. The one challenge is that ODE15S uses variable time steps so it is necessary to either interpolate the model point to match the experimental 42 time or the reverse. It was easier to interpolate the experimental data than the model. The experimental data has >5000 points, while the model generally generates 100-200 points depending on the steepness of the curves. 5.2.3 Results and conclusions Figures 5-7 and 5-8 show the experimental results and the calculated results. The model predicts the shape of the curves nicely. The optimized parameters are shown in Table 53. With ruthenium, the gas side reactions are much faster than liquid side, as evident by the large value of Kr. When a molecule of deuterium (D2) makes it to the surface, is rapidly scrambled with the gas phase. In the case of platinum, the liquid and gas exchange rates are almost the same but the liquid side equilibrium lies a lot more towards the surface bound side. This is consistent with water readily dissociating upon adsorbing on Pt as described in the literature”65 Table 5-3: Optimized parameters for experiment 43B and 43C Exp 43B k1=50586; Kr=129.54; Kg=6.9848; KI=8.6846; k5=10000; Exp 43C k1=31093; Kr=2.0048; Kg=5.3232; KI=94.379; k5=10000 43 5 0.8 0 200 400 600 800 1000 Time (min.) Figure 5-6: Model fit to experimental data for 5% RuIC (Exp 43B). Dotted lines are predicted values. 0 200 400 600 800 1000 Time (min.) Figure 5-7: Model fit to experimental data for 5% Pth (Exp 43C). Dotted lines are predicted values. 6 AQUEOUS PHASE ADSORPTION ON RUTHENIUM SPONGE All experiments were performed in the recirculating batch reactor (RBR) using previously described experimental procedures (Section 3.4.1). A summary of all experimental conditions can be found in the master experiment list (Appendix A). Data processing and calculations are explained in Section 3.4.2. Unless noted differently, in the figures used to display data generated with the RBR, the left graph displays solution concentration while the right displays quantity adsorbed. 6.1 Sorbitol, glucose and gluconic acid adsorption Previous work in our lab has looked at transformations of sorbitol (SO) to either propylene glycol (PG)5 or isosorbide“. It has been proposed that the first step in conversion of SO to PG involves the dehydrogenation of $0 at the C1 position to form an “aldehyde” intermediate. To get some mechanistic insight into this first step, information on the adsorption of $0 on Ru metal would be helpful. Initial adsorption work on ruthenium sponge was done by Schafer. He looked at $0 adsorption on hydrogen-saturated metal in a 300 ml Parr reactor with 20 g Ru sponge. No significant adsorption was seen until the temperature was raised above 80 °C. Above 100 °C, ~5-10 mg SOIg Ru was unaccounted for and was assumed to be adsorbed. This is significantly higher than the 0.2-0.5 mg SO/g Ru predicted for the measured metal surface area of ~0.2 mzlg-sponge and an area of ~30 A2 per sorbitol molecule. The appearance of activation with 45 temperature is possibly due to the desorption of hydrogen from the metal surface at ~80 °C, as mentioned in the characterization of the catalyst. The apparent over-adsorption of $0 is probably due to the degradation of sorbitol into a large range of products that were not analyzable by HPLC or were below the detection limits in Shafer‘s system. The RBR was used to look at sorbitol, glucose, and gluconic acid adsorption at 25 °C on Ru sponge metal. Preliminary experiments with sorbitol (SO) showed that on a hydrogen-saturated Ru surface there was no significant adsorption, while on a hydrogen-free surface sorbitol adsorbed and degraded into a large range of products. + Total Material (10mg) —e— Total Material (5mg) +Somitol (10mg) —Q—Sorbitol (5m) +Side Products (10mg) +SideProducts (5mg) 6 ............................................................. fi 5 ‘ n.9._~ __o~0.5 5...? _. H E _ 54 i a T! '2 L 1 __ -__ ...L7 -Aiifiiiiiiiiéfiivéfii‘igAiiiviiiiiiidA o_;_.‘..A.-a““ I I I 0 5 10 15 20 Time(l'louls) Figure 6-1: Adsorption of sorbitol on 10 g H2-free Ru sponge at 25 °C. Initial loading of 5 mg (27.5 pmol) and 10 mg (55 pmol). Figure 6-1 shows that at a high SO concentration (2-5 mM) the metal is saturated in a relatively short time (1 -2 hours), while at lower concentrations adsorption is slower as seen in stepwise additions (Figure 6-2). The presence of a large range of products makes it hard to give an accurate accounting of the amount adsorbed. Using an average response factor and average molecular weight (100 glmol), each of these experiments show an adsorption of ~0.1 mglg Ru (0.55 pmollg). It should be noted that no glucose is seen in adsorption experiments with sorbitol. Control experiments were run to see if this apparent amount of adsorption was really due to the Ru metal and not a result of calculation error or system behavior. First, an empty column (Exp 11) was used but this had a significantly larger volume than a Ru sponge filled column. To keep the volume constant with minimal increase in surface area, 150 pm glass beads (Exp 21) were used. To give a larger surface, a third experiment (Exp22) used Controlled-Pore Glass ® (product# CPG03000D, lot#11D011) with a total surface area of ~15 m2. All three of these experiments showed no adsorption of SO. 47 o on p on —e— [others]" —a— [$0] —I— adsorbed o N p N O G T! O G O U! I' O U'I Concentration (mM) O .5 I l O .5 . .0 . . . 00 Adsorptlon (pmollg Ru) 0 O N (.0 O N ‘IA’c; 4n A _§'r 4 o v v O —L f I n I 0 I i 0 q! n 0 r b a (. J V ‘\ ID (- -) (. .\ 1 fl 0 A 1 O Figure 6-2: Adsorption of sorbitol at 25 °C on 10 g Ru sponge using stepwise dosing (Exp 20). Concentration of other species is in sorbitol equivalents. When sorbitol was adsorbed in the presence of 5 mM H2804 (Exp 14), only one main product was formed at 10.8 minutes retention time on the HPLC. In the presence of acid, it is well known that sorbitol can cyclize by dehydration to form 1,4-anhydro-D-glucitol (sorbitan) and 143,6 dianhydro-D-glucitol (isosorbide).67 These species were identified by comparing their retention times with those determined by Dozeman‘37 on the same Bio-Rad column. The formation of sorbitan at 25 °C with dilute acid only in the presences of Ru sponge metal gives some evidence of the activation of the C1 position. 48 0.50 _ g 1 +iothetsr /M 0.45- , [mucosa] 0.9 0,40 . +[Gluconic acid] 0.8 c ‘ A A 0.35 — +IS°"’“°“ 0.7 E E +Adsorbed ,R . g ~0w~ oea 5 \ a g 0'25 \ I 0.5 E a 8 0.15 - .‘ _. 0.3 v N ;: it I “ 0.10 9 ' I“ . .55 4/“”"' '- 0.2 ’ L t Y r-K'r'iLiv'LK 0.05 - l -_ ’3‘“ 0.1 0.“) ‘V‘ _ \‘ 7‘", - '- jam-:1: I. m A .. T "' I T T T T 0 0 5 10 15 20 25 30 35 40 4s Time (hours) Figure 6-3: Adsorption of glucose at 25 °C on 10 g Ru sponge using stepwise dosing (Exp 23). Concentration of other species is in glucose equivalents. Adsorption of glucose at 25 °C on Ru sponge using stepwise dosing (Exp 23) is shown in Figure 6-3. The total adsorbed material is 0.88 pmollg Ru, which is similar to glycerol under the same conditions but more than sorbitol (0.55 pmollg Ru). The only products detected to any extent are sorbitol and gluconic acid. Both these products are formed to the same extent, suggesting a Cannizzaro-type disproportionation. It should be noted that in the presence of air, gluconic acid is the main product. After stopping the stepwise addition of glucose, the system was left to equilibrate overnight, sampling was started, and then 500 psi H2 was added to the headspace. Figure 6-4 shows that the concentration of gluconic acid jumped up followed by a slow conversion to sorbitol. This is a real change, because gluconic acid is measured by UV 49 detection and not refractive index (RI) detection. Because of glucose and gluconic acid peak overlap in RI, the glucose concentration is subject to error. This shows that some form of adsorbed acid can be removed by adding H2. 0.35 ——o— [othersl' 0.30 <———Added 500 psi H2 +[G'"°°s°l g 0.25 n —9—[GIUCOII'C acid] E, +[Sorbitol] g 0.20 - E c 0.15 — 8 . C 8 0.10 / O 05 ‘ W jw Dom fi f T I I I 62 64 66 68 70 72 74 76 78 Time (hours) Figure 64: Effects of H2 addition on adsorbed glucose at 25 °C on 10 g Ru sponge after stepwise dosing (Exp 23) Adsorption of gluconic acid on a hydrogen-free surface shows no formation of side products. Figure 6-5 shows the adsorption of gluconic acid at 25 °C on Ru sponge (Exp 25); the leveling off and slight decrease seen between 10-23 hours is mainly due to a small leak in the system that was not accounted for in the calculations; the actual amount adsorbed should be ~0.8 mg (0.42 pmollg Ru). The addition of H2 shows a slight increase the solution concentration of gluconic acid. This increase and the increase seen in Exp 23 (Figure 6-4) are taken as evidence that some gluconic acid adsorption is reversible. It should be noted that the observed increase is with in the uncertainty of the data. 50 0.7 — 0.7 [Sorbitol] Added 500 psi H2 ——p~ 0.6 + film 0.6 —c—[Gluconic acid] I A 0.5 ~— 0.5 ‘5 .. s “g- 0.4 0.4 g ’5' i... ‘5 0.3 — A 4:.» l ‘. 0 3 ° g r T “E 8 ' 3 0.2 ' 0.2 < 0.1 0.1 0,0 - “AC-Z1eccecccecceuccccccceccecccceecccccccc‘ excescttsw 0 0 5 10 15 20 25 30 Time (hours) Figure 6-5: Adsorption of gluconic acid at 25 °C on 10 g Ru sponge using stepwise dosing (Exp 25) The stepwise addition of sorbitol and glucose both show a lag time before products appear. This lag seems to be related to the oxidation state of the substrate: it is 5 hours for sorbitol and 10 hours for glucose. Comparing the total material adsorbed shows that glucose (0.88 pmollg Ru) is significantly higher than sorbitol (0.55 pmollg Ru) and gluconic acid (0.42 pmollg Ru). Unfortunately, because of product formation and the large molecule size, this system is too complex to draw meaningful mechanistic conclusions. Therefore, we changed to a smaller substrate, glycerol, as discussed below. 51 6.2 Glycerol Adsorption The adsorption of glycerol was studied by itself and in combination with propylene glycol. The effects of concentration, hydrogen saturation, and temperature were looked at with the hope of gaining mechanistic insight into the adsorption process. Experiments were done to look for evidence that the adsorption is reversible. 6.2.1 Concentration dependence Initial glycerol (GO) adsorption studies were performed at 25 °C using stepwise addition (Exp 24) and a fixed dose of 4.3 pmol (Exp 26) of GO. Figure 6-6 cleariy shows that in both experiments adsorption plateaus at ~0.3 pmollg Ru even though concentrations are 2.6 and 0.05 mM respectively. 3.5 0.40 3.0 A 0.35 A “=5 0 30 E 2.5 g - 5 2.0 E 0.25 E 5 0.20 g 1.5 5 o E 0.15 = 1.0 8 '8 0.10 0.5 < 0.05 0 exp 24 A exp 26 0.0 0.00 0 10 20 30 40 0 10 20 30 40 Time (hours) Time (hours) Figure 6-6: Adsorption of glycerol at 25 °C on 10 g Ru sponge comparing stepwise (Exp 24) and fixed dosing (Exp 26). Left solution concentration; right quantity adsorbed. 52 This result and the results from sorbitol adsorption mentioned before led us to assume that the total quantity adsorbed was independent of concentration. Most subsequent experiments were run based on this assumption, but it was only after examining all experiments at 25 °C that this assumption was validated. Adsorption (pmollg Ru) A .-*.-‘ more 99.69 N‘O 0.0 e e e e 0 “it 0.0 2.0 4.0 Initial Concentration (mM) 6.0 dd Adsorption (pmollg Ru) o o p o c 'o 'cnbe'oio 'ro Q .5 r C 0.0 1.0 2.0 3.0 Final Concentration (mM) 4.0 Figure 6-7: Adsorption of GO at 25 °C using (A) fixed dosing: maximum initial concentration vs. total adsorbed, and (B) final solution concentration vs. amount adsorbed. Figure 6-7 shows maximum initial concentration vs. total adsorbed (A) and also the final solution concentration vs. amount adsorbed (B) after ~18 hours. Maximum initial concentration is the concentration in the loop before the reactor is put inline. It should also be noted that this data has both 10 and 20 gram loadings of ruthenium sponge while the total volume of the system is about the same (11-13 ml). As long as there is enough GO in the system, the adsorption peaks at ~0.8 :I: 0.2 pmol/g Ru. The uncertainty is based on four nearly identical experiments at 7.5 mM initial concentration. The quantity adsorbed, 0.8 pmollg Ru, is only about 10% of the quantity of CO adsorbed via chemisorption (~9 pmol COlg Ru). Yet the value is close to that predicted (1 -1 .2 53 pmollg Ru) by assuming end-on GO adsorption with a 30 A2 per molecule surface area. Most experiments were run for at least 18 hours and adsorption curves seem to level off after ~10—20 hours. This behavior was also observed in stepwise addition of GO (Exp 24). It does not appear that exposure of H2-free Ru surface to water for an extended time changes the adsorption behavior, but even small amounts of polyol decomposition could deactivate the surface. Some noted exceptions to the above behavior are Experiment 67 and 68 (Figure 6-8 and Figure 6-9) which showed that the initial 5 pmol GO added adsorbed completely in the first 15 hours; the subsequent addition of 5 more pmol GO still adsorbed readily. Reduction conditions were not the same as for most experiments. In Experiment 67, the metal was washed with degassed water for 2 days (over the weekend); this was ~2-3 L of water that might have oxidized some metal. 1.2 0.12 -3 1.0 0.1 a E . z 2 =3? 0.8 008-; g E - 0 a”: 06 0065 n it ,2 6 0.4 0.04 F <9 8 E 8 "i < c 3 0.2 0.02 O -1 9 19 29 39 49 59 69 Time (hours) Figure 6-8: Adsorption of GO at 25 °C (Exp 67) on Ru sponge prepared with excess water. For clarity EG and PG concentrations are plotted at 10X. 54 1.2 0.12 -e—GO -e—adsorbed A EG a PG | ‘2“ 0.1 E 5 ,, .. 980 I owe: 3 h g V E .2 0 3 c 0. 0% E 9' o “a E 5 1 c 0 g g 0. 0.04 8 m < i: 8 0. ‘A A A . ‘Ir A A" I 0.02 A AA fl m®-0%D ;I.. A B ACPndPu o I T I 0 -1 4 9 14 19 24 29 34 39 Time (hours) Figure 6-9: Adsorption of GO at 25 °C (Exp 68) on Ru sponge prepared by reduction on the Micromeritics ASAP 2010 instrument. For clarity EG and PG concentrations are plotted at 10X. For Experiment 68, the Ru sponge was dry reduced on the ASAP instrument and flushed only with 10-20 ml degassed water. It is interesting to note that EG and PG did not form initially in Exp 68, while in Exp 67 and most other experiments these products were formed almost from the start. 6.2.2 Hydrogen saturation Generation of the “H2 saturated” and “H2 free“ surfaces is described in Section 3.4.1. Discussion of the nature of the Ru surface is presented in Chapter 4 and 5. Flgure 640 shows adsorption of GO on both H2-saturated (Exp 28) and H2-free (Exp 39) surfaces at 25 °C. The H2-saturated metal shows limited adsorption of ~0.05-O.1 pmol GOlg Ru, while ~0.3 pmol GOIg Ru is adsorbed on hydrogen free metal surface with the formation of trace amounts of EG and PG. 55 0‘40 oI-I25at. eH2free I S 0.35 c g 0.30 ‘W g 0.25 . °“°°°e E 0.20 4V. § 0.15 ’27 C 0.10 L 8 005 M- 0.00 l l 0 5 10 15 Time (hours) Adsorption (mol/g Ru) 9.0 88 0.25 '8 0.15 9.0.0 888 4 g o 0.0 9° n o 9 .0 u.°o° ° °oo° - WV I1 ° 0 Time (hours) Figure 6-10: Comparison of adsorption of G0 at 25 °C on H2-saturated (Exp 28) and H2-free (Exp 39) Ru sponge (10 9). After some minor repairs and the addition of 109 more Ru sponge into the reactor, the effect of H2 saturation was rechecked (Figure 6-11). Total GO added was 50 pmol (7.5 mM max) for each reaction. Unlike Exp 28 and 29, where the headspace gas was H2, both Experiments 93 and 100 used Ar in the headspace. 5.0 4.5 A 4.0 E 3.5 ‘a' 3.0 _o E 2.5 = 20 E 1.5 8 1.0 0.5 0.0 ol-I25at. eH2ere 0 5 10 Time (hours) 15 Adsorption (pmollg Ru) .099 (nude 99.09 NOD-#01 .09 0—3 cHZIrBe 5 10 Time (hours) 15 Figure 641: Effects of higher GO loading on adsorption at 25 °C on H2- saturated (Exp 100) and H2-free (Exp 93) Ru sponge (20 g). GO adsorption (0.3 umol/g Ru) was higher than expected in Exp 100; this might be due to the higher initial concentration of GO or it might just be that the headspace gas composition is critical. Unfortunately, no direct comparison of the effects of the headspace gas during the adsorption run was done, but Figure 6-12 shows adsorption of GO (4.5 pmol loading) at 50 °C under Ar on a H2- saturated surface (Exp 30). The observed adsorption of GO (~0.3 umol/g Ru) is comparable to GO adsorption seen on a Hz-free surface at the same loading (Exp 39), except that the adsorption rate is a lot slower. The higher adsorption under Ar could be explained by desorption of H2 from the Ru surface, freeing up sites for GO to adsorb. At 25 °C, the rate of H2 desorption is practically zero while at 50 °C it is significant. 300"”) Adsorbed (umollg Ru) Concentrationfor 99999999999 O‘Nu-hUIGNOCOO -2024681012141618202224 Time(hours) Figure 6-12: Adsorption of G0 on Hz-saturated Ru sponge (10 9) under argon at 50 °C (Exp 30). The effects of adding hydrogen to the headspace after GO adsorption on a Hz-free metal surface will be discussed in Section 6.2.3. Heating of the H2- saturated system to 100 °C (end of Exp 28 not shown) resulted in rapid decomposition of glycerol with the formation of methane, PG, EG, and trace 57 amounts of CO. It should be mentioned that at 25 °C the addition of H2 (250 psi) appears to stop any further decomposition reactions. 6.2.3 Temperature effects on GO adsorption Experiments were run at 25, 40, 60, and 80 °C under identical conditions (50 pmol GO, 20 g Ru sponge, and standard metal preparation). Gas analysis was not performed on all experiments; those that had no actual analysis done but were corrected with assumed methane production are labeled 'exp#*". "° W 0e .0 p o .l ’9 e 0.9 “‘6“ ,, 08 _ t. O ‘ ‘ 07 - .. 51.31123?" % 0.6 " AP""I‘~ £0 1‘ A mc( ) N E 0'5“ ”I" M9 8°ch 35 :H' 939 6 c #10,, 'o' memes) O 0.4 4 .,, 0.9 ”' . 3 0.3 ’ ’- 800mm) .,-' ammonia) '0 . “ 02 g ' .. - 3 0.8 - . ~ . g ‘ . .5 0.1 1‘] . “ - ‘ can AAA memo“) ' r 250(9‘l93') 0.0 U 0'72: . 5' ' 1b 115 25 25 ' k '0.1 I I l f I l u 0 5 1o 15 20 25 30 35 Thne(houn9 Figure 643: Temperature dependence of GO adsorption. (50 pmol GO loading, 20 g Ru sponge). Inset shows magnification of upper region of graph. Figure 6-13 shows that adsorption does not significantly change at different temperatures and is ~0.8 - 1.0 pmollg Ru. Multiple experiments at the same temperature were added to the graph to show that experimental variations have 58 a greater effect than temperature. Looking at the products formed as a function of temperature (Figure 6-14), it can clearly be seen that GO decomposition goes through a maximum at 60 °C. Scheme 6—1 OH HO OH + Ru H H i OH HOVKfO 4' H2(adfi) H2 H ds Fragmentation, retroaldol, C-Cclavage i c1-c3 #2:. Methane Scheme 6-1 can be used to explain the observed behavior. As the temperature increases, so does the fragmentation of GO, but hydrogen is needed to react with these fragments to remove them from the surface. Surface bound H2 is produced by the adsorption of GO, but can come off into solution at elevated temperatures. This H2 desorption starts to be significant at 60-80 °C. This release of H2 has the effect of oxidizing the metal surface. 59 20 30 4O 50 60 70 80 90 temperature (degc) Figure 6-14: Glycerol decomposition products vs temperature. Multiple experiments are included at 25 and 60 °C. As shown previously, hydrogen saturation inhibits the adsorption of glycerol at 25 °C. Adding H2 to the headspace at 25 ’0 only stopped the adsorption of GO, but at 80 °C (Exp 96) the result is the conversion of adsorbed GO to methane without adsorption of additional GO (Figure 6-15). It should be noted that adding H2 and increasing the temperature to 200 °C results in quantitatively converting all carbon species to methane. Some additional temperature effect experiments were run (Exp 30, 31, 32). Experiment 32 is run stepwise at 80 °C and includes the addition of E6 and PG. Data from these experiments were unreliable due to fact that changes in temperature were made may days after the start of the experiment. 60 9.0 88 i E . E. 5 Added 250 pSI H2 g ‘E 0.20 .° .o E E 0.15 g E g o1o§ 8 2 0.05 :5 Added 250 psi Adsorption (pmollg Ru) 0 5 10 15 20 25 Time (hours) Figure 6-15: Adsorption of GO (50 pmol loading) at 80 °C on 20 g Ru sponge (Exp 96) showing the effect of adding H2 and increasing temperature to 200 °C. Quantity GO adsorbed shown on lower graph, solution concentrations on upper graph. Concentrations of EG and PG are plotted on right axis. 6.2.4 Propylene glycol (PG), ethylene glycol (EG), and 1,3- propanediol (1 ,3-PDO) adsorption Propylene glycol adsorption experiments were run at 25 °C with 6.5 and 66 pmol PG loading. Figure 6-16 shows PG adsorption of ~0.18 umol/g Ru at low loading (Exp 42) and ~0.75 pmol/g Ru at high loading (Exp 41). The amount 61 of PG adsorbed at the high loading is similar to that found for glycerol. This would indicate that the surface is saturated. Comparing adsorption at lower concentrations suggests that PG binds less strongly than GO. 00>: OO ) 99 NO “99‘. 00 ‘r ’1‘ mollgRu 0.0 (no i d Concentration (mM) 2 o +66 umol exp 41 : +6.5 umol exp 42 g 0.3 f , +66umoll‘l2sat.exp51 3 02 J} 1.0 2 0.1 W 0.0 I A AA: A;#‘ AA‘ ATAAA 0.0 T' T 1 T 0 5 10 15 20 O 5 10 15 20 Time (hours) Time (hours) Figure 6-16: Adsorption of PG at 25 °C on 10 g Ru sponge comparing loading and H2 saturation. As with GO, the pro-saturation of Ru with H2 inhibited the adsorption of PG (Figure 646, Exp 51). This finding is consistent with the concept that an initial dehydrogenation step as proposed for glycerol also applies to propylene glycol. The cyclic nature of the noise in Experiment 51 probably stems from the HPLC system, but no satisfactory cause was ever found. Lactic acid and an unidentified peak were the main products and were only seen in the high concentration run with no hydrogen (Exp 41). The area response (RI) for these two peaks is almost the same, suggesting that they might be related. 62 16.0 14.0 12.0 10.0 6.0 4.0 2.0 0.0 Concentration (mM) on O +exp 52 —.— exp 53 5 10 15 20 Time (hours) Time (hours) Figure 6-17: Adsorption of E6 (80.6 pmol) on 10 g Ru sponge at 25 °C. Ethylene glycol was run only at 81 pmol EG loading (Exp 52, 53) and showed an extent of adsorption of ~0.8 - 0.9 pmollg Ru (Figure 6-17), again similar to GO. H2 saturation effects were not checked. 4.0 3.5 E 3.0 E = 2.5 O '5 2.0 § 1.5 E 3 1.0 0.5 0.0 A1,3 FED O 3—l-PA? XAA CIFA 5 10 15 Time (hours) 20 Adsorption (umol/g Ru) 0.8 0.6 0.4 0.2 0.0 —0.2 0 5 10 15 Time (hours) Figure 6-18: Adsorption of 1,3—PDO (50 pmol) on 20 g Ru sponge at 25 °C. Adsorption of 1,3—Propanediol (1 ,3-PDO) on 20 g Ru sponge at 25 °C (Exp 113) gave an extent of adsorption of ~0.8 umol/g Ru (Figure 6-18). This is similar to other polyols. It should be noted that 1,3-PDO showed a range of 63 decomposition products including ethanol, formic acid (FA), acetic acid (AA), and 3-hydroxypropanoic acid (3-HPA). 6.2.5 Binary adsorption (GO, PG) To examine the difference in binding energies of glycerol (GO) and propylene glycol (PG) on Ru, experiments were run with both species simultaneously. In one run, PG was added first followed by GO, in another run, GO followed by PG, and in third run both were added at the same time. These experiments were all run at 25 °C with 10 g Ru sponge. Initial adsorption of PG after 16-20 hours was ~0.18 pmol PGIg Ru for a loading of 6.6 pmol (Exp 42). At a higher loading of 66 pmol (Exp 44), this increased to ~06 umol PG/g Ru. In each experiment, 4.34 pmol of GO was then added and allowed to adsorb for 20 hours (Figure 6-19). In both cases, GO adsorbed after PG in the amount of ~0.2 (Exp 42) and ~0.1 (Exp 41) pmollg Ru, respectively. The total amount adsorbed is similar to the amount GO adsorbed at similar loadings (Exp 39 and 93). A 5 Adsorption (umol! A 0.8 0.7 . 0.6 0.5 0.4 0.3 0.2 0.1 ~ —e— G0 + total 0.0 #— 10 20 30 40 Time (hours) .0 co +GO ‘ —o—total 9 i .o .o G N 99 our .0 u Adsorption (pmol! Ru 9 N F—h pp O-l Time (hours) Figure 6-19: Adsorption of PG followed by GO on 10 g Ru sponge at 25°C. A: Exp 42, initial PG loading of 6.6 pmol; and B: Exp 44, initial PG loading of 66 pmol. GO loading of 4.34 umol follows PG loading. umol) followed by PG (6.6 pmol), there is no adsorption of PG. The amount of GO adsorbed was ~0.27 pmollg Ru before the addition of PG and ~0.32 pmollg Ru at the end of the experiments. It is clear that adsorbed GO inhibits the adsorption of PG, while adsorption of PG leaves sites available for GO to adsorb on. When the order is reversed (Figure 6-20) with initial addition of GO (6.5 65 E .5. i 8 E 8 0 10 20 Time (hours) 30 40 Adsorption (umol/g Ru) 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 -0.05 0 10 20 30 40 Time (hours) Figure 6-20: Experiment 45, adsorption of G0 (6.5 pmol) followed by PG (6.5 pmol) at 25 °C on 10 g Ru sponge. The simultaneous adsorption of GO and PG (2.2 GO, 3.3 PG umol) is shown in Figure 6-21. These results show behavior that is consistent with a G0 preference over PG for Ru and an irreversible interaction with the Ru. 0.6 0.5 0.4 Concentration (mM) 9 o N 0) 9.0 °—L 0 10 20 Time (hours) Adsorption (pmollg Ru) 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 10 20 30 Time (hours) Figure 6-21: Experiment 49, adsorption of G0 (2.2 pmol) and PG (3.3 pmol) at 25 °C on 10 g Ru sponge. 6.3 lsopropanol and acetone adsorption Glycerol and PG are C3 species with multiple hydroxyl groups; each shows adsorption on Ru sponge metal. It is logical to next look at C3 compounds with one hydroxyl group, namely isopropanol (IPA), n-propanol (n~prOH), and ethanol. Figure 6-22 shows the adsorption of isopropanol (IPA) on 20 g Ru sponge at 25 °C. It should be noted that a negative adsorption value seen in figures means that a compound is formed. It can be seen in Figure 6-22 that acetone is formed. This is expected if IPA undergoes dehydrogenation and if acetone is not strongly adsorbed. The addition of hydrogen at 20 hours converts all the acetone formed to IPA 4, 0.6 0 A2-propanol eacetone omet 05 ‘2-propmol eacetone oAds. IPA 3.5 A A . 3 04 A _ z - E 3.0 N g 0.3 g 2.5 g 02 £20 A” | x 5 0.1 3 1.5 250psil-l2 ‘E. 0-0 c -0.1 3 1.0 8 2 -0.2 0-5 W -0.3 0.0 r . m -0.4 0 10 20 0 1O 20 Time (hours) Time (hours) Figure 6-22: Adsorption of IPA (50 pmol) at 25 °C on 20 g Ru sponge (Exp 109) showing effect of addition of H2. It would be expected that if IPA were adsorbed on a H2-saturated surface, we would see no acetone. Figure 6-23 show that this is the case. Also, no additional products are seen. It is uncertain why a small amount of IPA adsorbs, but clearly no dehydrogenation occurs when the surface is first saturated with hydrogen. 67 .5. 3.0 0 10 Time (hours) 20 sorption (pmollg Ru) Ad 10 Time (hours) 20 Figure 6-23: Adsorption of IPA (50 pmol) at 25 °C on 20 9 H2 saturated Ru sponge (Exp 1 11) under 20 psi H2. Figure 6-22 shows a peak that was tentatively labeled as methanol, but this was an impurity that was there at the start of the experiment. It is curious that it disappeared when H2 was introduced, which suggests that it might have been a dissolved gas or a carbonyl species. If it was a carbonyl, we would expect to see a peak in the UV, but none was seen. Unfortunately, no W peak (210 nm) was seen for acetone either at these low concentrations, so we cannot rule out the possibility that this peak is an unsaturated compound. For now, the “methanol” peak is assumed to be an inert molecule of some sort and is ignored. 68 2.0 A IPA e Acetone .. . eds. GO | . ads. IPA a I or a E. c f 8 '8 < 0 1o 20 30 «1 Time (hours) Figure 6-24: Competitive adsorption of GO and IPA (50 pmol each) at 25 °C on 20 g Ru sponge (Exp 112) showing the effect of adding H2. During the adsorption of IPA, acetone is formed in the first 5 hours and then its concentration stays constant When IPA is adsorbed in the presence of GO (Figure 6-24), the maximum amount of acetone (~0.5 pmollg Ru) is formed in about the same time frame as in the adsorption of IPA alone. In the presence of GO, the acetone that was formed slowly converts back to IPA Presumably the hydrogen liberated from adsorbing GO reduces the acetone back to IPA. The next logical experiment was the adsorption of acetone. Figure 6-25 show adsorption of acetone at 25 °C on 20 g of H2-free Ru sponge. It clearly shows that IPA is formed. 69 Time (hours) A2-propanol e acetone 3.5 E 3.0 .5. 2 5 1‘ Added 5 ' '- 250 psi Hz W 5 1.5 OM : 3 1.0 . I 0.5 . 0.0 It . . and o 10 20 0 10 20 Time (hours) Figure 6-25: Adsorption of acetone (50 pmol) at 25 °C on 20 g Ru sponge (Exp 115) showing effect of addition of H2. The observation of acetone hydrogenation with no H2 available can possibly be explained by the dissociation of water on the Ru surface to supply the H2. Another possibility is that when an acetone molecule adsorbs, it dissociates completely (Scheme 62) to liberate six H-M. This is sufficient hydrogen to convert three molecules of acetone to IPA So if the hydrogen comes only from acetone, then the amount of IPA formed cannot exceed three-fourths that of amount of acetone adsorbed. Scheme 62 CO C H + M 114 + 2 114 + 6 I O / H3C CH3 \ f H + 2 3'4 H3C CH3 70 It might be possible to take the carbon all the way to CO2 by a water shift type reaction (Scheme 62) which could generate 10 more H’s. But there is no evidence of CO2 liberation and the ~% ratio of ~0.5 umol IPA formed to ~0.8 umol acetone adsorbed per gram of Ru sponge might suggests that the hydrogen comes from the acetone. The subsequent addition of H2 results in the rapid conversion of all acetone to IPA. 5.0 4.5 .. 4.0 E 3.5 g 3.0 25 8 20 c 1.5 3 10 0.5 0.0 A IPA e acetone - GO / Added acetone '/ 50 (mol [system I «W 45 A - 50 55 Time (hours) Adsorption (pmollg Ru) 2.0 1.5 . 1.0 - 0.5 0.0 -0.5 -1.0 -1.5 AIPA eacetone eadsGO oads.Acetone O O s 0 4“; e A n P ’ . . . A Addedecetone ———-‘"" ‘ ‘ 50mpllsysbm ‘L‘T—— 45 50 55 60 Time(hours) Figure 6-26: Adsorption (50 pmol each) of acetone after GO adsorption at 25 °C on 20 g Ru sponge (Exp 116) showing the formation of IPA. It is possible to convert more acetone to IPA if GO is adsorbed on the Ru surface first. Doing this should make more H2 available. Figure 6-26 shows adsorption of acetone after the saturation of Ru sponge with G0. The quantity of IPA produced is >1.0 pmol after 7 hours; this is twice as much IPA as is produced from acetone by itself. Figure 6-27 shows the direct comparison of these two cases. Because the initial quantities of acetone are not the same, we can only reach the qualitative conclusion that more IPA is produced from acetone 71 in the presence of some hydrogen donating species such as GO than from acetone alone. 5.0 45 _ AZ-propanol eacetone 4'0 . acetone' alPA‘ €35 f .5 3.0 . ’ . 2.5 9., - . O. 2.0 ’34-: . f 8 :73.‘. ..Oee 1: 1.5 ‘ — 8 1° “' ‘ AAALLA . ‘ALA‘ 05‘1‘;t=“‘ 0.0 O 5 10 Time (hours) Figure 6-27: Comparison of acetone adsorption (50 pmol) at 25 °C on 20 g Ru sponge (Exp 116, 115) Showing the effects of pre-saturation with G0. Open symbols are from presaturated experiment. The total quantity adsorbed in IPA/acetone system is ~0.2 pmollg Ru. This is independent of hydrogen saturation or which substrate is used first. It is interesting that this is about the same amount as seen in GO adsorption on a hydrogen-saturated Ru surface. It is possible that there are other adsorption mechanisms at work. 6.4 n-Propanol and ethanol The interaction of two primary alcohols, n-propanol (n-prOH) and ethanol, with H2-free Ru surface was looked at. Both ethanol and 1-propanol adsorbed to he same extent, as seen in Figure 6-28. Both produced the corresponding acids. There was production of methane in both cases but more appeared to form from ethanol than from n-propanol. 72 0 5 10 Time (hours) 15 ”I dsorption (pmollg Ru) .0 .0 P .0 9 .-‘ O N -A O) on O A 9‘: .b A N 10 Time (hours) 15 2‘) 4.0 3.5 E 3.0 a 2.0 a 0 1.5 3 1.0 0.5 0.0 0 ethanol 0 10 20 Time (hours) 30 40 PP.“ cacao 0.4 .0 0 Adsorption (pmollg Ru) 0 is as .r-io eads. 10 A AA —Methane 20 “me (hours) 30 40 Figure 6-28: Adsorption (50 pmol loading) of n-propanol (Exp 121) and ethanol temperature was increased to 50 °C after 18 hours with the following three results: conversion of n-prOH to PA, increased production of methane, and an increase in total quantity n-prOH adsorbed. It might be possible that at higher (Exp 114) at 25 °C on 20 g Ru sponge. Acetic acid (AA) and PA are formed. During Experiment 121, adsorption of n-prOH (Figure 6-29), the temperatures water acts as an oxidant of the Ru, releasing hydrogen and providing oxygen for oxidation of n-prOH to PA. It is also possible that the 73 adsorbed intermediate “propionaldehyde’ undergoes disproportionation to form n-prOH and PA It has been observed that upon addition of H2 the solution concentration of the acids produced drops sharply followed by a steady conversion to the corresponding alcohol. This sharp drop suggests adsorption of PA on the newly reduced surface. It is unlikely that PA reduction takes place, because reduction of the acids is very slow at 25 °C. This same behavior was also seen in the adsorption of gluconic acid. 5.0 45 e PA A n—prOH Ges .. 4.0 E 3.5 .5 3.0 2.5 2.0 I: 1.5 8 1.0 0.5 0.0 Adsorptlon (pmollg Ru) 0 20 40 60 Time (hours) Time (hours) Figure 6-29: Adsorption of n-prOH (50 umol/system) at 25 °C on 209 Ru sponge (Exp 121) showing effect of addition of H2 and increase in temperature to 50 °C. Gas C in legend refers to total gaseous carbon. 6.5 Organic acid adsorption Adsorption of LA was conducted at 25 °C on both H2-free and H2- saturated Ru sponge metal. Initial loadings ranged from 5—190 umol. Adsorption of equimolar mixtures of LA and PA were also examined. 74 Figure 6—30 shows concentration dependent adsorption of LA at 25 °C on 10 g of H2-free Ru sponge surface. In each case the adsorption levels off after ~4-6 hours. There were no products seen. The amount of LA adsorbed seems directly proportional to the initial loading. At 5 pmol loading, the adsorption is ~0.15 umol/g Ru, while at 50 umol/system the adsorption is 1.3 umol/g Ru. At 190 pmol, the adsorption is 4.3 umol/g Ru, which is several times the adsorption seen with polyols. 16.0 A 5umol s 50umol A 5umol e 50umol 14.0 . 93umol - 1soumor 5-° ‘ - 93umol . 190umol E 12") “AW” . I: 4.0 ..""” E 10.0 g 30 e 5 ' e E 8.0 T. c {M 0 o 8 6-0 4 K...- B, 2.0 «——-t = I 8 4'0 3 W 1.0 ~ 2.0 “ 0 0.0 W, 0.0 M , 0 5 10 15 20l 0 5 10 15 20 Time (hours) Time (hours) Figure 6-30: Concentration dependence of LA adsorption at 25 °C on H2 free Ru surface for different initial loadings. (Exp 54, 56, 57, 58) The gas phase was not analyzed, but if there was production of significant amounts of methane it would have shown up in the HPLC analysis; no methane was detected. If we compare the maximum amount of LA adsorbed (4.3 pmollg Ru) to the gas phase H2 and CO adsorption (~45 and 9 pmollg Ru), LA can occupy about two metal sites at the highest concentration examined. Control runs of LA with no Ru present showed no adsorption, but in the presence of air in 75 a control run lactic acid was oxidized to pymvic acid, presumably by the tube walls. The effects of H2 saturation on adsorption of LA at 25 °C on 10 g of Ru sponge is shown in Figure 6-31. Unlike the adsorption of polyols, the adsorption of LA is not limited by H2 saturation. This is not unexpected, because there is no need to dehydrogenate LA on the surface in order to adsorb a molecule of LA. Clearly the mode of adsorption is different thanis seen for polyols. 8.0 3.0 8e ~°M~ . “6m °‘ F ’ 2 - a e w E = 2.0 e ‘9 E 5.0 g 000 . 50m 0 4 0 5 1 5 .Q 0 5umol‘ E ' e Sumo! g ' o e 93umol g 3 o o 5umol' b - . 93umol ‘5 1,0 o 92umor 5 20 e 92urnol' 3 ° 1 o 3 0.5 a . 0.0 dbfil l r I" 0.0 A“ 1 T T —5 O 5 10 15 20 -5 0 5 10 15 20 Time (hours) Time (hours) Figure 6-31: H2 saturation effects on adsorption of LA at 25 °C on 10 g Ru sponge metal (Exp 54,55*,56,60*). *denotes experiments on hydrogen saturated surface. The pH was checked after LA adsorption for Exp 54, 57, and 58 and appeared to be higher than would be expected based on the LA concentration. Figure 6-32 shows measured and calculated pH vs initial concentration. Calculated pH is based on simple weak acid dissolving in water. The model curve is based on the assumption that when a molecule of LA adsorbs, it also generates one molecule of base. 76 .m 9: O O P :3 pH - Adsorption (pmollg Ru) N 9’ O O ..L O .° 0 o L wt {\2 . . \Q """"""""""""""""" ‘ “ ‘hr 0 HAW”) . . . .pH (model) pHaAonly) _. . pH measured 100 120 140 160 180 200 lnkial Loading (movsystem) Figure 6-32: Predicted vs. experimental pH as function of initial loading (pmol). Only one experiment was run with the adsorption of PA and that was the competitive adsorption of LA and PA (120 pmol loading each, Figure 6—33). PA seems to have a slightly lower adsorption (2.5 pmollg Ru) than LA (3.0 pmollg Ru). These quantities are the same as found for LA by itself at similar loading. if we look at the total loading of 240 pmol, we see an adsorption of 5.5 pmollg Ru. This is what would be expected if adsorption is directly proportional to initial concentration, so it is apparent that LA and PA are not competing for surface sites in adsorption and that the surface is not saturated under these conditions. 77 10.0 35 9.0 A 8.0 A 3 0 .e “uéee 6 “7 O E 7.0 u 25 —_._-L¢P°°o,; :5 Ana 7'." 6.0 g 20.8) o 20 v E 5.0 5 o c 15 g 4.0 g - 5 3-0 3 1.0 2.0 2 1 0 0.5 ' 0 LA 0 PA L ’ LA 0 PA 0.0 0.0 . . 5 -5 0 5 10 15 20 0 5 10 15 20 Time (hours) Time (hours) Figure 6-33: Adsorption of equimolar mixture of LA and PA (120 pmol/system) at 25 °C on 20 g of H2-free Ru sponge. (Exp 70) 6.6 Comparisons and conclusions The differences in adsorption properties of the various organic substrates on Ru sponge metal are contrasted in Table 6—1. The total quantities adsorbed for acids is ~4 pmol/g Ru, while that of all polyols, except sorbitol and isopropanol, is ~0.8 pmollg Ru. These results suggest that there two different methods of adsorption. The molecular size does not vary significantly between LA and C3 polyols, ruling out simple sten'c effects as an explanation for the differences observed. For polyol adsorption, a minimum of one terminal hydroxyl group is needed; this is supported by the lack of adsorption in the case of IPA. The inhibition of adsorption of polyols by hydrogen pre-saturation of the Ru surface is evidence supporting dehydrogenation as the first step hydrogenolysis of polyols. 78 Table 6-1: Contrasting adsorption properties on Ru sponge of organic acid and polyol. Organic Acids C3 polyols Adsorption ~4 pmollg Ru Adsorption ~0.8 pmollg Ru No H2 influence Strong H2 influence Consistent with H2 adsorption Consistent with ~30Azl molecule (2 Ru atoms! acid molecule) (~10 atoms/molecule) No degradation Dissociative with degradation 79 7 AQUEOUS PHASE ADSORPTION ON ACTIVATED CARBON 7.1 Experimental methods 7.1 .1 Carbon characterization The carbons used in this study were a 0.8 mm extrudate activated carbon (ROX 0.8, lot #520020, Norit Americas, Marshall, TX) and a powder activated carbon (designation 3310, lot #28850, Johnson Matthey, Sevierville, TN). Both of these carbons have served as support materials for catalysts we have used in GO hydrogenolysis, and lactic acid hydrogenation studies. The total surface area of each carbon was characterized by BET nitrogen physisorption at 78 K over a relative pressure of PlPo from 0.0 to 0.2 in a Micromeritics ASAP 2010 (Micromeritics Instrument, Norcross, GA). Surface area was calculated from the BET equation; micropore volume was determined using the t-plot method and total pore volume was characterized as volume adsorbed at the maximum relative pressure of 0.99. A summary of the carbon characterization results is given in Table 7-1. 80 Table 7-1: Carbon characterization by N2 adsorption at 78 K Carbon Type 3310 ROX BET Surface Area (m’lg) 715.6 833.6 Micropore Area (m’lg) 374.8 585.8 Total Pore Volume (ems/g) 0.654 0.536 Micropore Volume (cm’lg) 0.173 0.272 7.1.2 lsotherrnal adsorption measurements The quantity of material adsorbed onto activated carbon in this study was determined by the difference in initial and final species concentration in solution, which were measured prior to and following exposure to activated carbon catalyst supports, respectively. CA0 _ CA) . Vsolutr'on m CAS=( (7-1 ) carbon Isothermal adsorption experiments at 25°C were performed using 8.5 mL glass vials with Teflon-lined plastic lids. The vials were initially washed in HPLC- grade water, air-dried, and weighed in preparation for experiments. Carbon was weighed and added to the vials based on the total concentration of the solute(s) to be studied (0.1g carbon 5 0.1 M < 0.59 carbon 5 0.750 M < 1.09 carbon 5 2.0 M) in order to maintain at least a 15% change in solution concentration before and after adsorption. A quantity of solution of known concentration was then added to the vials to give approximately 7mL total solution + carbon in the vial. The final vial weight was then recorded for analysis following reaction. The vials 81 were capped and rotated end-over-end on a rotator overnight to ensure thorough mixing and equilibration. Upon removal from the rotator, the vials were either centrifuged or left standing for approximately 30 minutes to allow the suspended carbon in the sample to settle. One milliliter samples were then taken and analyzed via HPLC using the method described below. 7.1.3 Elevated temperature adsorption experiments Adsorption measurements were performed at elevated temperatures (40°C - 160°C) using a Parr 5000 multireactor system (Parr Instrument Company, Moline, IL). This system has six 75mL stainless steel reactors with internal stirring, independent temperature control, and continuous pressure monitoring. For these experiments, the reactors were cleaned and air-dried, and than carbon was added to the reactors according to the solution concentration to be examined: one gram of carbon was used for 0.05 M and 0.2 M concentrations of GO or PG, and four grams were used for a 0.5 M concentration. A Teflon stir- bar and 60mL of solution were added to the reactor. All starting, intermediate, and final weights were recorded. The desired temperature was set and the reactor contents were held for at least two hours at the desired temperature, after which a liquid-phase sample was taken. (The two-hour equilibration time was verified by room temperature experiments in vials for different time periods from 0.5 to 24 hr.) Sampling consisted of removing 1.5mL of waste to clean the sampling line, followed by removal of 1.0 mL as a sample. Typically, multiple temperatures between 40 and 160°C were examined in each experiment. All 82 waste aliquots and samples were collected and weighed at the conclusion of the experiment to check for mass loss via system leaks. Samples were analyzed using the HPLC method described below. 7.1.4 Analysis All samples from adsorption experiments were analyzed using a high pressure liquid chromatography system consisting of a Waters 717 Plus autosampler (Waters Corporation, Milford, MA), a Perkin Elmer pump (Perkin Elmer, Wellesley, MA), a Waters 410 differential refractometer, and a Perkin Elmer LC 90 UV spectrophotometric detector. The system used a Bio-Rad HPX87H column (Bio-Rad Laboratories, Hercules, CA) with 5mM sulfuric acid solution as the mobile phase. The column was operated isocratically at 50°C and a mobile phase flow rate of 0.6 mllmin. A sample injection size of five microliters was used for samples above 0.1 M concentration; ten microliter injections were used for samples below or at 0.1 M. Samples with a concentration above 0.5 M were diluted by a factor of four to maintain an injected concentration between 0.1 M and 0.5 M. A three-point calibration curve was used to determine response factors - no internal standards were thus used in determining species concentrations. 7.1.5 Adsorption modeling The remainder of this chapter will discuss modeling of single component and multi-component adsorption; therefore the general method for fitting 83 experimental data to Langmuir and Freundlich adsorption models is discussed in this section. Unless noted, the notations used for modeling are give in Table 7-2. All data from single component vial adsorption experiments at 25°C were modeled using both the Freundlich and Langmuir isotherms to obtain adsorption constants at room temperature. Freundlich: CA3 = K). C A" (7-2) Langmuir' CAS = KA CA CTA (7-3) ' l + K A C A 2(Cisi - CZEW Goodness ofFit: X2 = "1 n (74) For the Freundlich isotherm, the coefficients K; and n were found by a least-squares fitting of experimental data as CA8 vs CA with predicted CAs using solver in Microsoft Excel. Goodness of fit (x2) is the average residuals squared. The coefficients for the Langmuir isotherm (KA, Cm) were found using the same procedure. Our published adsorption of GO and PG on activated carbon uses the more traditional method, where the Freundlich isotherm coefficients, K; and n were found by a least-squares linear regression of experimental data as In(CA.) vs. ln(CA) to give slope n and intercept In(Kp). Goodness of fit (R2) is given by Microsoft Excel. For the Langmuir isotherm, a plot of experimental data as (CA/CA.) vs. CA gives a slope 1ICTA and an intercept (1/KACTA). The traditional method was more cumbersome to use and also tended to weigh small quantities adsorbed more heavily. Table 7-2: General notation used for modeling adsorption on activated mrbon CA Observed solution concentration of species A in equilibrium with carbon (M) CA0 Initial solution concentration of species A prior to adsorption (M) CA3 Concentration of solute A in activated carbon adsorbent (mol/kg carbon) Cm Maximum concentration of A in activated carbon adsorbent (mol/kg carbon) AHA Heat of adsorption (kJ/mol) Ka Van’t Hoff constant (M") KA Langmuir adsorption model equilibrium constant for adsorption of solute A (M“) K).- Constantin Freundlich isotherm model (mol/M"lkg) mm" mass of carbon in solution (lg) n Exponent in Freundlich isotherm model equation R Ideal gas constant (8.31 JImol/K) T Temperature (K) VWVolume of solution (liter) Subscript A,B Denotes component, either glycerol or propylene glycol 7.1.6 Summary of experiments performed In our lab, the catalyst used for most hydrogenation/hydrogenolysis work is 5% RuIC (PMC 3310). For this reason, the majority of adsorption experiments are on 3310 carbon, the support used on the RulC catalyst obtained directly from the manufacturer. The substrates were selected to match with Chen’s‘58 hydrogenation experiments and include organic acids, primary alcohols, diols, and glycerol. Examination of an additional carbon, Norit ROX, was included to help with our mechanistic investigations of glycerol hydrogenolysis; on that support only GO and PG were studied. Table 7-3 summarizes adsorption experiments covered in this work. 85 Table 7-3: Summary of experiments preformed during the study of aqueous phase adsorption on activated carbon. Compound 3310 carbon ROX carbon Reduced Catalyst (5%RuIC 3310) F orrnic acid (FA) I Acetic acid (AA) I Glycolic Acid (GA) I Propanoic acid (PA) IN V Lactic acid (LA) IN V Ethanol (EtOH) I n-propanol (n-prOH) IN V Ethylene glycol (EG) I,V Propylene glycol (PG) IN IN 1,3-propane diol (1 ,3—PD) I Glycerol (GO) IN IN LA&PA EI,xV LA&PG El, xV LA&n-prOH El, xV PA&PG XV PA&n-prOH EI,xV GO&PG xl,xV xl,xV EG&PG El =isothermal adsorption; El= equimolar isothermal adsorption; xl=variable mole fraction isothermal adsorption; V=variable temperature adsorption; EV= equimolar variable temperature adsorption; xV=variable mole fraction variable temperature adsorption. Single component adsorptions on 3310 carbon were done with organic acids, primary alcohols, diols, and glycerol. Adsorption of equimolar mixtures of LA&PA, LA&PG, PA&PG, GO&PG, and EG&PG were examined. Variable temperature (40°C - 160°C) adsorption measurements of LA, PA, n-PROP, EG, PG, GO, LA&PA, and GO&PG were done on 3310 carbon. Control experiments were carried out in vials or in the reactor without addition of carbon. No degradation or loss of PG or GO was observed at any temperature. Experiments in which total vial or reactor weight after experiment differed from initial weight by more than 1 2%, indicative of a leaking vessel, were discarded. 7.2 Glycerol and propylene glycol 7.2.1 lsothennal adsorption of GO+PG on ROX activated carbon Adsorption experiments were conducted at room temperature (25°C) in glass vials for substrate solution concentrations ranging from 0.01 M to 2.0 M. Figure 7-1 compares the experimental quantities adsorbed with those predicted by the Freundlich and Langmuir isotherm models for individual adsorption of GO and PG on ROX activated carbon. A summary of the calculated parameters at room temperature for each adsorption isotherm is given in Table 7-4. As can be seen, the Langmuir isotherm gives the best fit of GO adsorption over the entire concentration range studied. 87 2.5 1.5 /- 52.0 ‘5 A / a A‘ ‘ 0 w 'l-". "I 5... § - ”x ‘ PG 2 0.5 _ I, " —— Imgmuir PG 1!, I I GO 0.0 . . ' 1 , ---------- Imgmuir GO A 500 1000 1500 2000 4“ 0 100 zoo—1500 Solution concentration (ml) A Solution concentration (ml) 2.5 1.5 A A 1.3 ‘ p 0 2.0 - /¢ 6‘ g / g 1.1 ‘ / g AL --“. O 1/ I r 5 ' 7 ‘g’ 0.7 / _.'.' .»--""" § Freundich PG § ' y; . PG 3 0'5 o co 3 0-3 ’- Fremdich PG A .......... Fremdich so 0.1 Tr————' » I GO 0.0 I I I I I """""" FMdiCh GO B 500 1000 1500 2000: ‘0-1 0 1 fl Solution concentration (mM) B Solution concentration (mM) Figure 7-1: Experimental and predicted a) Langmuir and b) Freundlich adsorption isotherms for PG and GO at 25°C on ROX carbon. For PG, the Langmuir model gives the best fit below 0.75 M, but the Freundlich model better describes the data at concentrations above 0.75 M. The observed isotherm for PG suggests that the quantity adsorbed reaches a plateau at the Langmuir maximum of Cu ~1.64 mol/kg at about 0.75 M in solution and then adsorbs by another mechanism at higher concentration. In Section 7.5 a new adsorption model is proposed that takes in to account two different modes of adsorption, micropore filling, and surface adsorption. In micropore filling the 88 adsorption energy is significantly higher than that found for surface adsorption. Micropore filling dominates at low concentrations, while at high concentrations both modes contribute to the observed adsorption. Filling of the micropores has a relatively constant energy of adsorption and is best represented by the Langmuir isotherm. In the F ruendlich adsorption isotherm, adsorption energies are dependent on coverage, making it better able to predict adsorption at higher concentrations where both adsorption modes are active. Table 7-4: Langmuir and F reundlich isotherm coefficients for GO and PG on ROX carbon. 60 PG GO PG Langmuir Freudlich CTA(mol/kg) 1.77 1.64 K; 0.105 0.202 KA.293K(M") 4.755-03 1.63E-02 n 0.379 0.323 R2 Fit 0.983 0.997 R2 Fit 0.946 0.916 Glycerol shows no tendency to adsorb beyond its Langmuir maximum of Cu = 1.77 mol/kg. While the reported value of the Langmuir Cm for GO is ~8% larger than that for PG, we do not believe that the difference in the two values is significant. Because GO and PG have nearly identical molar volumes of 73.0 mL/mol, the volume of species adsorbed according to the Langmuir model is likewise similar. It is significant, however, that the molar quantity of PG adsorbed is greater than that of GO at any concentration below the Langmuir maximum; this is reflected by the fact that the calculated adsorption equilibrium constant for PG is approximately three times that for GO. Interestingly, the mass of GO and 89 PG adsorbed at any concentration is nearly identical (Figure 7-2); we assume this is coincidental but worth noting. 180 Adsorption (mglg C) essaseéee APG I GO 0 500 1000 1500 2000 Solution concentration (mM) Figure 7-2: Adsorption of PG and G0 on ROX carbon on a mass basis. Using the measured ROX carbon micropore volume of 0.27 cmalg, the micropore concentrations of PG and GO corresponding to Cu at room temperature can be calculated to be 6.07 M and 6.55 M, respectively. This is 44% and 48%, respectively, of the pure component molar concentration of 13.7 M for both PG (1.04 glcma) and G0 (1.26 glcm3). Assuming that the density of PG/water and GOIwater varies linearly with mass concentration, the remaining micropore volume is filled with water (the only other species present) at a concentration of 29.0 M for GO and 31.0 M for PG. This approximate 5:1 molar ratio of water to GO or PG is similar to the result reported by Parsons et al. 69 in their analysis of glycerol aqueous solutions - they report that all water is associated with glycerol and solution behavior changes significantly at a glycerol mole fraction of ~0.15 (5.6:1 molar ratio). Thus it wpears that glycerol retains water associated by hydrogen bonding when adsorbed into the carbon. On the other hand, the continued adsorption of PG beyond the Langmuir model value of 90 Cu may be because PG is less hydrophilic than PG and thus sheds solvating water to reach a higher concentration in the pore as PG solution concentration is increased above 1.0 M. A second paper by Parsons and Koga 7° on PG-water solutions supports this interpretation of lower PG hydrophilicity in solution. The less hydrophilic nature of PG also accounts for its larger adsorption equilibrium constant relative to GO; PG favors residence in the carbon pores rather than in bulk solution. 7.2.2 lsothennal adsorption of GO+PG on 3310 activated carbon Isothermal adsorption on the Johnson-Matthey 3310 powder activated carbon, as with the ROX carbon, follows the Langmuir isotherm model closely for PG and GO solution concentrations up to 0.75 M. A summary and comparison of the results for the ROX and 3310 carbons fit to the Langmuir isotherm are found in Table 7-5. As can be seen, the values of CT for GO and PG are similar on the 3310 carbon, in agreement with the results found for the ROX carbon. If the 3310 carbon micropore volume of 0.17 cm3lg is used, the concentrations‘of species in the micropores at the maximum Langmuir concentration Cu are 6.35 M and 7.0 M for GO and PG, respectively. These values are very close to the values of 6.5 M and 6.1 M obtained on the ROX carbon, indicating that adsorption is strongly associated with the micropore volume of each carbon. Further, if the additional PG adsorption (up to 2.0 mol/kg) on the ROX carbon is considered, the micropore concentration for PG on the ROX carbon becomes 7.35 M, very close to the PG value of 7.0 M for the 3310 carbon. Thus the absolute and relative quantities of PG and GO adsorbed in the micropores of the 91 two carbons are similar; this is supported by the ratio of adsorption constants (erlKeo) at 298 K (Table 7-5), which are 3.6 and 3.4 for the 3310 carbon and the ROX carbon, respectively. However, the affinity of PG and GO for the 3310 carbon is greater than that for the ROX carbon, as the ratio (3310/ROX) of adsorption constants (KA) at 25°C for GO and PG is 1.32 and 1.39, respectively. The greater adsorption affinity for the 3310 carbon may result from differences in surface functionality or in micropore structure, considering the two carbons come from different sources. Table 7-5: Comparison of Langmuir isotherm coefficients for ROX carbon and 3310 carbon. Carbon Type CTA (MOI/k9) KAZG c ("4) G0 ROX 1.77 4.75E-03 3310 1.10 6.25E-03 PG ROX 1.64 1.53E-02 3310 1 .22 2.28E-02 7.2.3 Elevated temperature adsorption Quantities of GO and PG adsorbed onto ROX and 3310 carbon were determined at concentrations from 0.1 to 0.5 M and temperatures from 313 to 433 K. At each temperature, the equilibrium constant (Kit) for the Langmuir model was calculated using CTA determined from isothermal adsorption experiments at 298 K. The equilibrium constant is assumed to vary with temperature based on 92 the Van’t Hoff equation, as the Langmuir model assumes all adsorption sites have equal binding energies and thus the heat of adsorption (AH) is constant. 1n(1<,,)= —%*%+ln(1€o) (7-5) For each species, the plot of In (KA) vs. 1fT (absolute) from experiments was fitted with a straight line via least squares linear regression to give the heat of adsorption (AH) from the slope (-AH/R) and the preexponential constant (K0) from the intercept |n(Ko). The plots for GO and PG on ROX and 3310 are seen in Figure 7-3 and the calculated heats of adsorption and preexponential factors are given in Table 7-6. 9 9 8 . = -1913.7x + 11.771 8 _ y = -2373.3x + 12,516 7- . . R2=0-804 7- - . _ 52:0.8031 6 _ I I I 6 — ‘ A A I g 5 - I g 5 ‘ I r: 4 « ‘E 4 - A "' 3 _ I GO y = ~1765.9x +1004 -. 3 _ . GO ‘ A f 1 A PG R2 = 03593 2 « 3 PG y = -2079.8x + 10.561 0 ‘ ——Lrnear 1 - —|_jnear R2 = 0.7011 . . o r , 0.002 0.0025 0.003 0.0035 0 002 0_0025 (1003 0.0035 A 1” (K) B 1IT (K) Figure 7-3: Temperature dependence of Langmuir equilibrium constant Kit for GO and PG on ROX (A) and 3310(8) carbon. It is seen that the data follow the Langmuir model well over the measured temperature range of 298 K - 433 K for both carbons. The heats of adsorption of GO and PG are both approximately 15 kJ/mol (3.6 kcal/mol) on ROX carbon, while on 3310 carbon G0 is 14 kJ/mol and PG is some what higher at 22 kJ/mol. This is indicative of similar van der Waal’s type forces and weak hydrogen bonding responsible for adsorption of each species. 93 Table 7-6: Langmuir isotherm coefficients for GO and PG on ROX carbon ROX 3310 GO PG GO PG Langmuir CTA (mol/kg) 1.77 1.64 1.10 1.22 KA295K (M’ 1) 4.75E-03 1.63E-02 6.25E-03 2.28E-02 AH (kJ/mol) -15.91 -14.68 -14.02 -22.32 Ko(M“) 7.73E-06 4.36E-05 1.95E-05 4.36E-06 R2 Fit 7.2.4 Two component adsorption studies Aqueous solutions of PG and GO at total species concentrations ranging from 0.05 to 0.5 M and species fractions of PG and GO ranging from 0 to 1.0 were conducted at room temperature on ROX carbon and at elevated temperatures on both ROX and 3310 carbon. The extended Langmuir model (Eq. (7-6) and (7-7)) has been applied as the model of choice for mixed solute adsorption; the denominator of the extended Langmuir model accounts for competitive adsorption of the two species (A=GO; B=PG) into the activated carbon micropores. = KACACTA (76) AS 1+KACA+KBCB KBCBCTB (7_7) ”3 = 1+ KACA + KBCB All coefficients (Cu, Cm, AHA, AHB, KM, K30) required in the extended Langmuir model were taken from the single component results reported in Table 7-6. Figure 7-5 shows the predicted dual component adsorption isothen'ns on ROX carbon at 25 °C and total initial concentrations of 0.05 M, 0.3 M, and 0.5 M. The abscissa in Figure 7-5 is the fraction of glycerol in the total solute present (GO/(GO+PG)) - not mole fraction in solution. The data are reported in this fashion to show comparison of model with experiment and accentuate the different adsorption behavior of GO and PG. It is seen that the extended Langmuir model accurately predicts the adsorption of propylene glycol and glycerol over the studied concentration range. The preference for PG adsorption is clear; in fact, at equimolar concentrations the quantity of PG adsorbed is approximately three times that of GO, in accordance with the ratio of equilibrium constants determined in single component experiments. The adsorption on 3310 shows similar behavior as ROX only with higher adsorption quantities. Predicted and experimental isotherms for GO and PG on 3310 can be found in Appendix B (Figure B-1). It should be noted that predicted GO values are higher then experimentally found but the experimental data has a lot of scatter. 95 ROX Carbon 1.60 1.40 1.20 1.00 predicted adsorption (mmol/g Cl 9.0.0.09 88888 p 8 0.50 1.00 1.50 2.00 experimental adsorption (mmollg C) 1.40 Carbon 3310 g 1.20 2 g 1.00 C 3 030 go 1, 0.60 a U 0 40 g ' 6 ° x=y E, 0.20 0 GO APG 0.00 0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80 experimenhl adsorption (mmollg C) Figure 7-4: Experimental vs predicted adsorption of PG and GO on ROX carbon and 3310 carbon for all data. Performance of the extended Langmuir model for GO and PG on ROX carbon and 3310 is shown in Figure 7-4, which compares experimental to predicted values. All data are included (single and dual component isothermal and variable temperature data) 1.6 14 APG IGO‘ 250‘; ° 60m" 1.2 ‘6 310 E09 30.6 I\‘ 0.4 Tr i - _____ 0.2 \.\ -. I‘ doc-4 . . \ 0.0 02 0.4 0.6 0.8 1.0 390 1.6 14 APG IGOI 25°C ' 300mm 1.2 4 £10 A \ E06 mil—4 0 I, 0.4 T _ .d' 0.2 9 0.02". . . T 0.0 0.2 0.4 0.6 0.9 1.0 no APG I Go 2593 500mm“ 0.00-"’2 . . . 0.0 02 0.4 0.6 0.8 1.0 yeo Figure 7-5: Experimental and predicted Langmuir isotherms for mixtures of GO and PG on ROX carbon at 25 °C and 0.05 M, 0.3 M and 0.5 M total species concentration. Abscissa is fraction of GO in the total species (G0 + PG) present. Lines are the predicted isotherms. Figure 7-8 shows the predicted isotherms at a total initial concentration of 0.5 M and temperatures of 80 °C, 120 °C, and 160 °C. The predicted values again show good agreement with experimental data, and validate the assumption of constant heat of adsorption over the range of concentrations and temperatures investigated. 1.6 1.6 1-6 14 APG IGOl 80°C 14 APG IGOI 120°C. 14 APG 'GOF180‘64 ' 500nm ' WM . m 1.2 1.2 - 12 . a10 ‘ §10A 310 = . ‘ . = . 9 l 0.8 \ . E08 \ E03 X 30.6 ‘ './' 30.6 \f ‘£06 A I g .r o , o ,. o \ 0.4 .X 0.4 ’ .5 0.4 \{ 0.2 _1 0.2- .. ' 02 i ”I" - —"" .... .' 000'". . . . 0.00-‘2' . . . 0.0-4L. . . 0.0 0.2 0.4 0.6 0.3 1.0 0.0 0.2 0.4 0.6 0.9 1.0 0.0 0.2 0.4 0.6 0.3 1.0 too we yeo Figure 7-8: Experimental and predicted Langmuir isotherms for mixture of GO and PG on ROX carbon at 0.5 M total species concentration and 80 °C, 120 °C, and 160 °C. Abscissa is fraction of GO in the total species (G0 + PG) present. Lines are the predicted isotherms. 97 The adsorption data clearly show that PG and GO concentrations in the carbon pore structure are significame different than those in bulk solution. This obviously affects the interpretation and kinetic modeling of reactions taking place over activated carbon catalysts. Figure 7-7 shows the extended Langmuir model predictions of GO and PG pore concentrations, based on the measured micropore volume, in ROX activated carbon at a temperature of 180°C for different bulk GO and PG concentrations. GO pore concentration is only modestly different than GO solution concentration at this temperature, while PG pore concentration is approximately five-fold larger than its corresponding solution concentration. 98 1.2 —0M 5 1.0- ——0.1M .t : —"02M ”*” O 0.84 . /’ ‘/‘ E ----0.4M / .4" .1 ’/n” eeee g 0.6~ —0.7M ’/"’ """"" 8 ’6' ’.-x‘. g 0.4~ /,’ ‘,."‘ 8. /.”,’."' g 0.24 (4“ 0.0 t A 00 02 0.4 08 08 10 GO solution concentration (M) 2 01M 2 13° N ‘ —--0.2M 5 15‘ *— --- 0.4M 1.42 _______________ ——07M 1.2~ """" X ------ . Series1 g 1 ~~n 00.82 "'“""---—_.. —-—--—_-+___—--. 508- _________ 00.4« _—"'-—"—-—-X—_. F 0.2. )L o I 1 T I 0.0 0.2 0.4 0.8 0.8 1.0 B 60 solution concentration (M) Figure 7-7: Extended Langmuir model predictions of a) GO and b) PG pore concentrations in ROX carbon at 180 °C. Symbols represent trajectory of 1.0 M GO conversion to PG. The presence of PG lowers GO pore concentration significantly, while the presence of GO only affects PG pore concentration slightly. The “x” symbols in Figure 7-7 represent an idealized (100% selective) reaction trajectory of GO hydrogenolysis to PG starting with 1.0 M GO solution. During reaction, GO pore concentration declines more rapidly than does solution concentration - converting 70% of GO to PG, which lowers solution GO concentration from 1.0 M 99 to 0.3 M, reduces the corresponding GO pore concentration from 1.15 M to 0.19 M, a six-fold decrease. At the same time, PG concentration increases from zero to 0.7 M in solution, and to nearly 1.8 M in the pore. These differences deviate ’ from bulk solution concentrations to a significant enough extent that they must be accounted for in accurate kinetic modeling of GO hydrogenolysis over activated carbon catalysts. Moreover, the local concentrations may depend on the particular carbon support in use, so that any comparisons between carbon catalysts must include comparison of the adsorption properties of reactants and products. 7.2.5 Conclusion The adsorption of propylene glycol and glycerol over two activated carbon supports has been examined over a broad range of concentrations (0.005 M to 1.5 M) and temperature (298 - 433 K). Adsorption of individual species is well- represented by the Langmuir isotherm, and simultaneous adsorption of the two species is described by an extended Langmuir model that uses the single— component adsorption constants and accounts for competitive adsorption between the two species. The overall quantities of PG and GO adsorbed are comparable and are closely related to the carbon micropore volume. Propylene glycol exhibits an enhanced affinity for adsorption relative to glycerol, with its equilibrium constant approximately 3.5 times that for glycerol. This enhanced adsorption is attributed to the less hydrophilic nature of PG, which favors its presence in the carbon micropores vs. the bulk solution, and also accounts for 100 the lower extent of solvation of PG via hydrogen bonding. The adsorption model shows that pore concentrations in liquid phase reactions using activated carbon catalysts can be significantly different than the corresponding concentrations in the surrounding solution. 7.3 Ethanol, n-propanol, EG, and 1,3-PDO adsorption Adsorption of ethanol, n—propanol (n—prOH), ethylene glycol (EG), and 1,3- propanediol (1 ,3-PDO) are compared in Figure 7-8 and include the Langmuir and Fruendlich predicted adsorption. It should be noted that both models were fit over the entire concentration range. The Fruendlich model predicts the observed adsorption better for all species at concentration >200 mM. The predicted concentration needed to adsorb 1 mmol of substrate per gram carbon is 734 mM for EG (0.017), 213 mM for 1,3-PDO (0.091 ), 162 mM for ethanol (0.49), and 34 mM for n-prOH (1.78), this is consistent with the order of the K0‘" value given in parenthesis. Kow is the distribution coefficient between octanol and water as defined in the literature.71 On 3310 carbon both the Langmuir and Fruendlich adsorption models perform worse than on ROX carbon, which could be due to the large micropore volume found in ROX. In Section 7.5, a new adsorption model is proposed that gives better performance for organic acids, PG, and n—prOH. It is expected that it would also perform better for ethanol, EG, and 1,3-PDO but it was not tested on these compounds. 101 3-0 2.0 I 2" e e 1.8 I A 7 - 532.5 / ————————————— 81 6 .I 2 " . ° u ' [.3 .520 1’9. [Av If", 31.‘ ’ E ’ ”. ___. E1 2 2' E ........... I --------- E - .' /- :1.5 - " :1.0 7 ,/ A 8 n 80.3 s / f- """ II W / O/ .- 3” ‘ ‘ e0.6 , , a?" 20 5 $0.4 zit-L ./ i - ‘en-prOH cEthanol ‘02 fl 1' cn-prOH cEthanol I1,3-PDO AEG ' r I1,3—PDO AEG 0.0 t t r . 0.0 i . r A 0 500 1000 1500 2000. A 0 50 100 150 200 Solution concentration (mM) Solution concentration (mM) 3.5 2.0 , . .. - ’ 1.8 J . . £25 .7 ’ / £14 0 , 4 ,e ’ O r v ........ : ,' /n --------- r :10 / /4,_» 01.5 --f——O'7H / Po 3 - 4 """""""" Q. a , A g - :v ________________ 31.0 i / en—prOH 30-6 - cn-prOH ‘2 c Ethanol 20.4 c Ethanol 0'5 I1,3'Pm o 2 I 1,3'PDO EG ' EG 0.0 . T . ‘ - 0.0 . ‘ . B 0 500 1000 1500 2000 B 0 50 100 150 200 Solution concentration (mM) Solution concentration (mM) Figure 7-8: Experimental and best fit data for isothermal adsorption of n- propanol (n-prOH), ethanol, 1,3-propanediol (1 ,3-PDO), and ethylene glycol (EG) on 3310 carbon at 25 °C. (A—Langmuir, B-Fruendlich) 7.4 Organic acids The effect of chain length was examined in the adsorption of formic acid (FA), acetic acid (AA), and propanoic acid (PA). Figure 7-9 compares the experimental quantities adsorbed with those predicted by the Freundlich and 102 Langmuir isotherm models for individual adsorption of FA, AA, and PA on 3310 carbon. A 3.50 5 .e 3.00 8 a 2.50 6 E 2.00 5 1.50 8 a 1.00 E- 8 0.50 2 000 eFA IAA APA ' cFA IAA APA A 0 500 1000 1500 2000 A 0 50 100 150 200 Solution concentration (mM) Solution concentration (mM) ‘5 3.50 g 3.00 a 2.50 . 6 E 2.00 — 5 1.50 c i 1.00 3 0.50 3 ' g 000 leFA IAA APA 3 ' eFA IAA APA 0 500 1000 1500 2000 B 0 50 100 150 200 3 Solution concentration (mM) Solution concentration (mM) Figure 7-9: Experimental and best fit data for lsothen'nal adsorption of formic, acetic, and propanoic acid on 3310 carbon at 25 °C. (A—Langmuir, B— Fruendlich) As can be seen, the F reundlich isotherm gives a better fit for all three acids over the entire concentration range studied. Examining the adsorption at high concentration, we see that AA and PA adsorb ~3 mmol/g C while the FA is closer to 2.2 mmol/9 C. The last two data points for the FA do not fit the Freundlich model. Including these points in fitting the Freundlich model results in decreasing adsorption (~2 mmol/g C) while overestimating the adsorption at low 103 concentrations. Looking at the predicted concentration needed to adsorb 1 mmol of acid, PA (16 mM) is adsorbed much more strongly than AA (72 mM) which is stronger then FA (176mM). This behavior could simply be explained by the increased organic character from FA500 mM) the adsorbed amount of PA appears less than that for pure PA, but this is merely the result of defining total solution concentration as the sum of the PA and PG concentrations. 105 fruendlich M1 ---------- freundlich Mid: Adsorption (mmollg c N o .5 0| 'o 1 .0 OI .0 o ." .- - 100 200 300 400 Salmon concentration (mM) 500 A PA I PG 3.5—fnnndIchPA(1) o PA+PG a 3.0 / '5 2.5 / 3 E A‘ 0 g 2.0 ~ & 3 1.5 I f 3 1.0 so... _ 0.0 ' L' ' . i 0 500 1000 1500 2000 SohltlonconcentnfloMmM) Figure 7-1 1: Isothermal adsorption of equimolar mixtures of PA+LA (A) and PA+PG (B) on 3310 carbon at 25 °C. Adsorption of equimolar mixtures of PA+PG and LA+PG is shown in Figure 7-12. PG appears not to inhibit adsorption of LA to any significant extent, while LA inhibits the adsorption of PG. The temperature dependence of organic acid adsorption was only examined for LA and PA and will be discussed in the multi- component modeling section. HA I PG A LA I PG 2.5—truandIich PA(1) o PA+PG 2,5 ———fruend|ch LA(1) o LA+PG . 2.0 f + 2.0 E, 1.5 E 1.5 / O 1.0 - i 1.0 . 8 l P 0.5 0.5 — 2 I . g ' ' I I 0.0 1!! . . o.o . . . . 0 200 400 0 100 200 300 400 500 Solution concentration (mM) Solution communion (mM) Figure 7-12: lsothen'nal Adsorption of equimolar mixtures of PA+PG (A) and LA+PG (B) on 3310 carbon at 25 °C. 106 7.5 Multi component adsorption studies relevant to hydrogenation of LA and PA to PG and n-prOH As previously mentioned, a goal of this work is to have a comprehensive model for predicting pore concentrations of reactive species from single component data to be used in kinetic models for hydrogenation reactions of LA and PA to PG and n-prOH. Single compound isothermal and variable temperature data were collected for each species and variable temperature data were collected for combinations of LA + PA, LA + n-prOH, LA + PG(limited), PA + PG and, PA + n-prOH. Data for these experiments will be presented and discussed together with the modeling of this system. Initial attempts to use the extended Langmuir model failed to predict the competitive nature of the adsorption, even though the Langmuir model gave a reasonable fit for single compound adsorptions. It was thought that large differences in the maximum Langmuir concentration Cm for PA (2.3mmollg) and LA (1 .6mmollg) were overestimating the effect of LA, so we adjusted the extended Langmuir model with “weighting factors” 0A3 and 03A. The adsorption is then defined by Equation 7-8 and 7-9, where a” = E12 and a3, = J— TA aAB K C C AS = A A TA (7-8) 1+KACA +aABKBCB KBCBCTB (7'9) BS =1+KACA+aBAKBCB 107 This model is referred as Weighted Extended Langmuir or WEL and is also applied to PG-GO system, where it showed no significant changes (aAewaaAM) and therefore was not used in that system. This correction did not significantly improve our model’s performance in the LA-PA system. A closer look at single compound data showed that at low concentrations (<100 mM), the adsorption of LA, PA, and n—prOH is much stronger (Figure 7-13). 700 5004 wlw” _- ' ---o---lPGl 200 o r '0 0' ‘-°' O O ' ' " -' +[PA1 U ,. A ' f o '0 1 . —-—rLA1 100‘ . "mu-[won o I l T I .....--[a o 200 400 600 300 1000 o 20 40 so so 100 Ce(mmolIL) CarinmoIII-l Figure 7-13 : Low concentration adsorption deviation from Langmuir isotherm at low concentration for LA, PA, PG, and n—prOH on 3310 carbon at 25°C. We therefore propose that there are two different modes of adsorption: the filling of micro-pores, and adsorption on macro-pore surfaces. In the filling of the micro-pores, interactions of the substrate with the walls are much more pronounced compared with planar surface interactions. Each adsorption mode is assumed to follow the weighted extended Langmuir model, and the adsorption is simply the sum of both. This adsorption model will be referred to as Dual Weighted Langmuir (DWL). This approach of using two part adsorption modeling 108 is not new, some of the literaturezz' 72 73 shows models where part of the adsorption is competitive and some is independent. The generalized form of the DWL adsorption isotherm for n competing species is shown in Eq. 7-10. The labeling of constants and concentrations follows the same approach used earlier in this work, where the first subscript refers to the compound (except numbers are used in place of letters (i.e. A and B are now 1 and 2), and the last subscript designates the adsorption mode. C _ leKnlCn + CntZKnZCn (7_10) n: — nt nt 1+—‘—ZC.~..K..C.- 1+CLZC..2K.2C. ml i=1 m2 i=1 _ leKnlCn + CntZKnZCn "3 1+ KMC" 1+ anCn (741) For single component adsorption, Equation 7-10 simplifies to Equation 7- 11. The adsorption now depends of four parameters: Cm,Km and Cman. Cm is limited by the micropore volume (Vmp), and molar volume (V...) of the species adsorbing cannot exceed Vmme. can is the maximum adsorption on macro-pore surfaces and using a Langmuir model should theoretically be limited to a monolayer coverage. This restriction was not applied because of the difficulty of calculating a monolayer coverage and also the possibility that there is multilayer adsorption. All single component isothermal adsorption data on 3310 carbon and ROX carbon were fitted with the Dual Weighted Langmuir model. Optimization was 109 done using Microsoft Excel Solver to minimize the sum of the square of the residuals. It should be noted that in some cases there was no stable solution and it was necessary to assign a fixed value to C11: (maximum adsorption second adsorption mode). Optimized parameters can be found in Table 7-7 Table 7-7: Optimized isothermal adsorption parameters for the DWL model. PA LA n-prOH PG Km (M1) 1.225+00 1.225-01 1.20501 5.45502 Cm (M) 8.00E-01 8.39E-01 1.18E+00 1.005+00 an (M1) 5.59503 3.44504 2395-03 1025-04 Cnt2 (M) 2.25 4.00 2.08 4.00 Extending the DWL model to variable temperature data requires that the temperature dependence of the equilibrium constants Km and Km be known. _fli) K :Koe[ R T (7'12) Equation 7-12 shows that K(T) depends on two constants, Ko and AHIR (H for short). For single component data, this means that we are trying to optimize six parameters (Cm, Kn1o,Hm,Cm2, Km, an ), which is too many degrees of freedom. Langmuir adsorption isotherms have an implied assumption that C... is independent of temperature; therefore Cm and Can were found using isothermal data. The same least square fitting method describe previously was used to find Km, Hm, Km, an, with the additional constraint that the calculated K‘s at 298 °K must equal those found in the isothennal fitting. (Table 7-7) The optimized values are showing in Table 7-8. Generally, the optimization converged on a single 110 solution, except for n-propanol. With n-propanol, the initial guess for these parameters did have some effect on the final optimized values found using Excel Solver. Table 7-8: Optimized isothermal adsorption parameters for the DWL model. PA LA n-prOl-l PG Kn10(M-1) 2.46E-05 2335-05 1255-04 1.18E-05 1572111 (K) -3.225+03 -2.55E+03 -2.075+03 -2.525+03 ano (MT) 1.91E-05 8.87E-06 8885-05 1345-08 Agnz (K) -1.895+03 -1.095+03 -1.125+03 -1.435+03 A simple program was written that would randomly change one of the initial values for Kn1o,Hn1,Kn20, an by a random amount not exceeding 150%. The system was then solved using Microsoft Solver and the final parameters saved. These final parameters were used as the input for the next cycle. The program was set to run overnight but stopped after 13944 cycles when there was a combination that did not converge. The value of the objective function (sum of the square of residuals) was examined and it was found that 82% of all solutions were within 0.5% of the lowest value found. The range of solutions that were within 0.5% of the optimum value are listed in Table 7-9 and varied by more than t7%. These results emphasize the importance of checking the stability of the solutions found using numerical optimization and making sure a global minimum has been found. 111 Table 7-9: Percentage range of optimized parameters whose objective function is within 0.5% of the “best fit”. [ H.- l muss-l ”g L '"(ngl m l 7.79% 5.90% -3.74% -1 .48% 0.49% -7.51% -5.69% 7.04% 2.79% 0.50% 0.00% 0.00% 0.00% 0.00% 0.00% 6.63% -5.02% 7.09% 2.81% 0.44% 7.28% 5.51% -7.25% -2.88% 0.49% 3.50 3.00 O 3'00 e O 2.50 / 9 .4/—_— 9 '5 2.50 6 .0. 2.00 d E ’ E E 2.00 - E : 2' 1.50 ~ g 1.50 - €- 1.00 18:! 1.00 i 3 A n-prOH < 0.50 0 0 PA 412 0-50 —dual LM n-prOl-l 0.00 0 . _. “aim PA 0.00 . . . 0 500 1000 1500 2&0 0 500 1000 1500 Concentration (mmollL) Concentration (mmollL) 2.50 8 6 a a 2.00 4— E E e g E, 1.50 c r: ° ° 100 - a u s E 0.50 3 I LA 3 . PG __ dual LM PG “‘3' LM LA 0.00 . . 5 0 500 1000 0 500 1000 1500 2000 Concentration (mmollL) Concentration (mmollL) Figure 7-14: Experimental data and WDL fit for the isothermal adsorption of PA, LA, n-prOH and, PG on 3310 carbon at 25 °C An Excel sheet that expands Equation 7-10 and 7-12 to model the competitive adsorption of up to 4 compounds has also been developed. lsotherrnal adsorption data for LA, PA, PG and, n-prOH on 3310 carbon at 25 °C 112 is shown in Figure 7-14 and includes the DWL fit The model accurately predicts the experimental data. 0 20 40 60 80 100 1 20 140 1 8 Tempenhue (°C) Figure 7-15: Variable temperature adsorption of 1000 mM LA on 3310 carbon. A 3.5 .4 3.5 o o :3 3.0 i 3.0 ~$2.5 E 2.5 722.0 g 2.0 3 1.5 3 1.5 $1.0 g 1_0 3 0.5 3 0,5 A n 0.0 I 0.0 —X-y 0.0 1.0 2.0 3.0 4.0 o_o 1_o 2,0 3.0 4.0 experlmental adsorption (mmollg C) experimental adsorption (mmollg C) 8- 3.5 =3 3.0 E g 2.5 £20 3 1'3 g0.5 1goo 0.0 1.0 2.0 3.0 4.0 0.0 0.5 1.0 1.5 2.0 experlmental adsorption (mmovg C) experimental adsorptlon (mmollg C) Figure 7-16: Experimental vs predicted single component variable temperature adsorption PA, LA, n-prOH and, PG on 3310 carbon from 25-160°C 113 The variable temperature data were obtained as described above. Adsorption data for a typical single variable temperature experiment are shown in Figure 7-15. The parity plots are shown in Figure 7-16 for single compound system; the DWL reasonably predicts the adsorption on 3310 carbon over a wide temperature range. Figure 7-17 shows the predicted isotherms for LA+PA at a total initial concentration of 0.5 M and temperatures of 25 °C, 80 °C, and 150 °C. The predicted values again show good agreement with experimental data, and validate the assumption of constant heat of adsorption over the range of concentrations and temperatures investigated. PA is preferentially adsorbed on 3310 carbon. COMMItZS‘C COMMItOO‘C CWIHGO'C 250 250 250 _ > e PA —PA" 6 6 0 g 200 A LA ------- LA 9 200 o 200 5 150 . A S 150 E 1.50 ‘5' \ ,. ‘5' 5 g 1 00 5: i 100 i 1.00 0' 050 .9 "\ g 0.50 3 050 3 . ..... e 3 0.00 m """" . 0.00 ' 000 0.00 0.50 1.00 0.00 0.50 1.00 0.00 0.50 1.00 Fraclon LA Fruclen LA Fraction LA Figure 7-17: Experimental and predicted DWL isotherms for mixture of LA and PA on 3310 carbon at 0.5 M total species concentration and 25 °C, 80 °C, and 150 °C. Abscissa is fraction of LA in the total species (LA + PA) present. Lines are the predicted isotherms. Figure 7-18 shows the predicted adsorption isotherms for an equimolar mixture of LA+PG at total initial concentrations from 10 to 500 mM and temperatures of 25 °C to 150 °C. It can be seen that the DWL model overpredicts the adsorption of PG relative to LA If we compare the low concentration fit of DWL for LA and PG, it can be seen in Figure 7-19 that LA is underpredicted 114 while PG is slightly overpredicted. This imperfect fitting may be due to the difficulties of obtaining reliable adsorption data at very low concentrations. Gathering more data to improve the model is possible, but likely not worth it because experimentally it was found that PG generally does not inhibit the hydrogenation of LA to a significant extent, so it would be simpler to leave out PG and any non-adsorbed species in the final model. Adsorption (mmol/g 0) .° .° .0 .O .0 :-* 8 8 8 8 8 8 1.40 E? 0.0 0 100.0 200.0 300.0 400.0 500.0 600.0 Concentration (mmollL) Figure 7-18: Adsorption isotherm for equimolar mixture of LA and PG on 3310 carbon at 25 °C 115 CI 0 .o 8 8 ..9. 8888 p _I O adsorbed pmollg O O O l I p 8 0.0 200.0 400.0 630.0 Kilo 1000.0 0.0 10.0 20.0 30.0 40.0 500 Solution concentration (mM) Solution concentration (mM) Figure 7-19: Adsorption isotherm for LA and PG on 3310 carbon at 25 °C Figure 7-20 shows the predicted isotherms for LA+n—prOH at a total initial concentration of 0.5 M and temperatures of 30 °C, 80 °C, and 150 °C. The predicted values for LA show good agreement with experimental data while the n-prOH is underpredicted at low temperatures. This deviation is probably due to the difficulties encountered in fitting the n-prOH data as described later in this chapter. comm-1m Combhetlonetco‘c Combhetlonettdo'c 2.5 - O M u. 4) 2.5 o u w 2.5 . LA LA. 8 <| ‘ W ....... J 8 . ....... l 6 ‘ W ....... W €20 . 2.0 ""'°" ""0" 2.0 . . £15 , ‘ ‘_.. ' E15 ‘ E15 .r' "A £10 ¥ .+" 1.00 ‘... " E10 \ fl .. ‘ A . 1» -' ' 80.5 \ 0.5 \ 0.5 . "° 2 2 ‘ 0.0 ' 0.0 ' . 0.0 ' 00 05 10 0.0 0.5 10 00 05 1o Blane-plat Freshen-pron Fauna-pron Figure 7-20: Experimental and predicted DWL isotherms for mixture of LA and n- prOH on 3310 carbon at 0.5 M total species concentration and 30 °C, 80 °C, and 150 °C. Abscissa is fraction of n-prOH in the total species (LA + n-prOH) present. Lines are the predicted isotherms. Figure 7-21 shows the predicted isotherms for PA+n-prOH at a total initial concentration of 0.5 M and temperatures of 30 °C, 80 °C, and 150 °C. The 1 16 predicted values for PA show good agreement with experimental data while the n-prOH is underpredicted at low temperatures. This difficulty of n-prOH adsorption is similar to what was seen in LA + n-prOH and will be discussed later in this chapter. Experimental data show that PA and n-prOH compete for the same sites and adsorb in very similar quantities. Therefore, n-prOH should show inhibition on the reaction rate of hydrogenation of PA. Mdfl'c comm-woe W11“? 1) e T! ——-m' A i e F! -——Fl' — ‘ A n-prOt ------- mum—ll ‘0 I I .- l Adeerptlenonmellgq 3 r 1 / WMWC) . . 3 . 4 ./ e l WMMC) O e I I c I I ..o' _p' A. 1 00 t (A 00 ' . 0.00 0.5) 100 0m 05) 1. 0.00 0.50 1.00 Figure 7-21: Experimental and predicted DWL isotherms for mixture of PA and n—prOH on 3310 carbon at 0.5 M total species concentration and 30 °C, 80 °C, and 150 °C. Abscissa is fraction of n-prOH in the total species (PA + n-prOH) present. Lines are the predicted isotherms. Figure 7-22 shows the predicted isotherms for PA+PG at a total initial concentration of 0.5 M and temperatures of 30 °C, 80 °C, and 150 °C. PA greatly inhibits the adsorption of PG, and therefore PA is adsorbs much more strongly than PG. The predicted values for PA show good agreement with experimental values except at high temperatures. This deviation is not due to competition from PG but due to a bias in the DWL model, which tends to overpredict adsorption at high temperatures while underpredicting the adsorption at low temperatures. Figure 7-23 clearly shows this bias. Attempts were made to eliminate this observed bias by adjusting the parameters but these attempts resulted in a 117 significantly worse fit. Lacking a logical basis to eliminate data points that resulted in this bias, the best fit was used and is reported. W‘flrc mast)? mauve -250 1» e PA —n\ 25° e m —m 1m 0 2m \‘ ‘ PG ------- EU" 6 2m ‘ R3 ....... R3. 6 1.0 0 1m 2150 A €1.50 \ E100 " 080 s \\ .- = \ s 100 \Q' 1.00 v\. 3' gm 050 ‘5 ..' 0.50 g 04) .4.. 5 ,,,,, 4 ..... A ..... V 0.20 000 - 0.00 ------- 4 0.00 000 0.50 1.00 000 0.50 1.00 000 Figure 7-22: Experimental and predicted DWL isotherms for mixture of PA and PG on 3310 carbon at 0.5 M total species concentration and 30 °C, 80 °C, and 150 °C. Abscissa is fraction of PG in the total species (PA + PG) present. Lines are the predicted isotherms. 140% Z 130% 12011. 8 I 110% h ; 1m g 90% 2 8091. e. a! 70% m 0 20 40 60 80 100 120 140 160 Temper‘ure (‘6) Figure 7-23: DWL predicted adsorptions of PA as percentage of actual adsorbed on 3310 mrbon at initial concentration from 10 mM to 2000 mM. The DWL model’s performance in dual compound systems is best summarized by looking at the parity plots in Figure 7-24. LA and PA show linearity to X = Y while n-prOH is generally underpredicted. PG shows a significant amount of scatter. The cluster of points between 0.6 and 1.0 mmol/9 C are from one experiment with a 4:1 PG to PA ratio. 1 18 23.0 —— ------- -- U 5 2.5 5 ~— 20 i a 1.5 ’6 i 1.0 .0 0.5 . PA 8 _FY 3 0.0 0.0 1.0 2.0 3.0 experinental adsorption (mmollg C) 0.0 0.5 1.0 1.5 2. experimental adsorption (m mollg C) I LA 0.0 0.5 1.0 experlmertal adsorption (mmovg C) 0.9 0.8 0.6 predlcted adeorptlon (mmollg C) 4 3 2 ePG 1 0 0.0 0.5 1.11 experimental Idsorptlon (mmollg C) 4—1 Figure 7-24: Experimental vs predicted adsorption of PA, LA, n—prOH and, PG on 3310 wrbon in binary mixtures only. 7.6 Additional notes on adsorption modeling n-Propanol data was inconsistent with model. The question arises whether n-prOH behaves differently or the experimental data were flawed some how. Control experiments were run to validate the variable temperature experimental design. Experiments were performed as described earlier, except that no carbon was added. Even though no carbon was added, a carbon loading of 1 gram was 119 assumed to maintain consistency in the calculations. Figure 7-25 shows single component measured adsorption of 100mM n-prOH, LA and, PA without any carbon over a temperature range 25 - 160 °C. Adsorption (mmol/g C) Figure 7-25: Apparent species adsorption in control experiments without carbon showing the need to correct for vapor fraction of n—prOH. (note data for 150 °C after 14 hours is plotted at 160°C for clarity.) PA and LA show no significant deviation from zero as expected. For n-prOH there is a significant deviation at high temperatures due to high vapor pressure of n-prOH. To correct for the fraction of n-propanol in the vapor phase, partial pressure (yP), vapor volume and temperature were used to calculate the quantity (mmols) of n—propanol in the vapor phase. Partial pressure was calculated using modified Raoult’s law( yP = xyP”t ) and the following assumptions: 1. Gas phase acts as ideal gas. 120 2. PS"t for n-prOH can be described by Antoine’s Equation 3. Solution density is ~1 glml 4. y (gamma) is a linear function of x and T with the form y = DT + Ex + F Antoine coefficients (A=7.77374, B=1518.16, C =213.076) where taken from Yaws' Handbook of Antoine Coefficients“. Aspen properties was used to calculate gamma’s over range 05x5.02, 25$Ts160. Least square fitting was used to find D, E and F (D=0.009439, E=-196.09, F=16.955). This correction was applied to all n-prOH data at temperatures greater then 80 °C. 121 8 CONCLUSIONS AND RECOMMENDATIONS 8.1 Conclusions A study of substrate-catalyst interactions has been completed; both metal and carbon support adsorb reacting species and influence apparent reactivity. Methods and a recirculating microreactor apparatus for measuring adsorption on metal surfaces from the aqueous phase have been developed. The key features are a large metal to solution volume ratio, automated sampling, an operating range of 25-200 °C, and headspace gas control. Adsorption of organic acids and polyols on Ru sponge metal from water was characterized. The extent of lactic acid is ~4 pmol/g Ru and is independent of hydrogen saturation of the metal surface. C3 polyols with terminal hydroxyl groups adsorb irreversibly through a dehydrogenation step to an “aldehyde like” species followed by degradation. This adsorption is inhibited by presaturating the metal surface with hydrogen. Adsorption of isopropanol shows little adsorption on either a Hz-free or Hz-saturated Ru surface. On the Hz-free surface formation of acetone, a dehydrogenation product, is seen. Primary alcohols, when adsorbed on a Hz-free Ru surface, show the formation of acids, an oxidation product. 122 Single component and multicomponent competitive adsorption on activated carbon at a range of concentrations and temperatures was characterized. For glycerol and propylene glycol, adsorption of individual species is well-represented by the Langmuir isotherm, and simultaneous adsorption of the two species is described by an extended Langmuir model that uses the single-component adsorption constants and accounts for competitive adsorption between the two species. Propanoic acid (PA) and lactic acid (LA) adsorption is competitive and PA adsorbs three times more strongly than LA. A two part adsorption model consisting of adsorption into the micro pore and adsorption on the surface of the macropores was developed. The model accurately predicts the individual species adsorption and when the individual adsorption parameters are used, competitive adsorption was accurately predicted with the noted exception of n- propanol. Incorporation of the competitive adsorption model into our kinetic model for the hydrogenation of LA and PA was done by Chen“. The combined model performed reasonable well when using reaction rate constants found on unsupported Ru catalyst and the pore concentrations predicted based on competitive adsorption on carbon. H-D exchange on H20-D20-H2—Dz over metal catalyst was characterized. The dissociation of water is the limiting step on 5% RulC but not on 5% PVC. The experimental data can accurately fit the kinetic model. 123 8.2 Recommendations There are several areas that warrant further investigation to elucidate the mechanisms of adsorption that were observed in this work. 0 Simpler systems with substrates containing no carbon atoms or only one carbon atom should be looked at in more detail. a. A thorough study of the interactions, in liquid phase, between water and the metal surface focusing on the actual oxidation state for the metal. b. Characterization of interaction between Ru surfaces and 01 species (ie. methane, methanol, formaldehyde, formic acid, and carbonate). This should include hydrogen saturation affects and temperature dependence. 0 The reversibility of lactic acid adsorption on Ru sponge metal should be examined. . Adsorption behaviors on different metal surfaces should be investigated. One option is adsorption on Raney nickel, which is a known active hydrogenation catalyst. 124 APPENDICES 125 E888 8.28 cc 5:5 52.988 c: 388.8 8 mv or on - . 2 83 8? $820 8328 a 8:88 8838 888$ 5.888 2 8258 882a 95 8588 use? 2: fa: 8 f be 5m 2 88.. 832 8828 8:83 m 888$ E858 8988 >8 222 mv 2.. mm «2888 ed 3. 2 83 82.3 3820 88:3 m :52 2 858 .2358 2.23 888838 88:3 98 Bow 2:833 .8862 .o 50:88 85. 8:88 060m 28 use 826 Eur 8:58 Boa 2:820 .cozaEEano 8 2 8888 3.8 888$ 8858 29:. .88; 9a 88288 no; 38: of. mv 2.. mm 28:2» ed :1 2 83 8? 8820 83:3 N. 28 08am 8 8:58 .2388 888$ E888 new new 8.58:3 8288 SEES. 7. mv 8: mm 28:28 he :1 2 83 2: 8820 8823 a 888$ .5288 83.88 SEER. T. mv 81 mm 2888 No :1 2 83 8.. .898 8:98 a 888$ 5288 04¢: 53 8.888 =8 28.56 am .2558 88:8 3.331 new 58:33 :8. «I0 3.. .88 43 .u 83 83. 8E 8:833. 83 8xm n. ... «502050 8238 5:22.350 .8: 282.898 .8885. u< 52.8%? 126 888$ E088 mv o... .8 222.8 a 3.. 2 83 2:2 _2.28 8 888$ E088 mv 0: 8 28:2» o 3... 2 83 2:2 5.5.8 Nu. 888$ E088 8.8.800 0E8 .508 2. 800 8.5., mv 0... mu 20088 a :1 2 83 quw 8.88m ad vow”... _2Em 8802. 885E888 0E8 mEofio8 8.. 3...... :05 888$ E088 93 8 8:88 swE 88.. E808 v.8 mv 0: mm 2088 o ..m 2 83 8.3 88.8 8: :8 0:... 828.8888?! F _oEsoF «own... 880 8; 8:88 8E 0E: .05: 22.9.8 .2Em 8.888 8808 08 8 8.882 05 00803 :. $.82 8...; 5w"... s.Ev :. 828.8 m. :05 888$ E088 Om 8.5. 8.... E0Et8x0 :8 EaE 8. 8...... mv 0... mm 2888 a :m 2 83 8.9 8.8.8 8388 «a 8.80. E088 381 5.; .8088 ..0. 8838 ”A goEzn. T 2 8888 08E _oEzoP U080 :05 00288 goEsodz 888$ E088 889. 3 8.. £88 .0. .28 .oanF 0028 mv 0: mu 20808 ho am 2 83 8.3 8.88m 82.88 «H E088 8228 o: 5.; 8.8.888 008098 8 mv 0... mu . . 2 83 2: 8.8.8 88:8 .3. 8.28. a. 538.8 127 22 8:82 E228 a 3 832.2 95.9. ho 5:988 a on 2 2:08 22:. .o:_::2 $9.82 956 9:3" 89. 2 E88 0n. :5 0m :5 8::2 m. 00 on 2.. cm. 22:58 or am. aEE -953 .88an 3822: 5 8:988 8268 .o. a 8268 £29m m2: 0.. on"... :< $382 95.6 825 08:8 8.228 N... a :o 5.5.82 cm N... om .28 NT. or am. .8820 vo\wm\volq|m 828.2. 95.0 .22 Bow 3. 9.22662 .8888 9:06 ow. 0.. mm E2628 or 9:2 uEvz 2:820 33:8 N uEmd .22 REES om. 9.083.? ow a: mm 22.8% o: 2:2 uEv... .2820 voatmo «M 3. 9.2:: 7 2:88 953 .22 2::oEm .38 :2 2.88 :8 28 2:83 cm 2.. mm 22:28 or 9:2 9:? 883.0 3: :8 mm 8:988 o: .8268 9N mason aEmd .22 8:988 o: :8 5:82 o: ow 0... mm 82> m2 850 2:2 asz .258 3398 392.2, 95.0 .22 828:: .22 9:7. on 2.. mm 22258 or 3”: 2:2 aENz 6:80. n< 5.2.8.3., 128 898:2 E288 « on .2. mm 23:8 2 3m 283 95 00 E 895an E296 8988 .0635... on :< mm E2629 or :m 2 goo. 25.6 00 No. 8282 E296 aEvdz 3888 ..< on :< mm 22829 or am. 2 :84 aEvd 00 um $92.2 E23». J 8988 00 9:3: 23% on 2 SN 22:59 2 a: 233 8; om. H 0. on «a 82s 828: on EQE .223 8382 E28? 5230: :5 23m 20 2a 20:23.8 om :< pm. 22:29 or am. 2 80.. as: 00 fl qla mm 00 on": 2.. 8908.. anm 325 82.8 3.233 f a 8 8.5.83 o... 8 .23 f 2 :m 2.98.3862 % .w>o. 5:288 .820 o: 22 5:02.: «a 9.32: m. On. :5 82 ___.m 2 258 0m 23 8.52 8. oo 98".. 2 89:2 8.a 225 on. :8 89.83 aENd aEmN h. .8552 8:228:00 5.; 8383, om 2.. mm. 82% 2 HT :62 -953 .2820 36:8. N» .828. "< 5232, 129 .8886 :2: so. on. m. ow .0 232a 8288 2.. .. 8288 9:3 on. 2. 09.2.0 2. ... 32.3mm. 0n. 3 5:63 $9.03. Sci» 00 on. .25? 8288 23 22.. 8.6a 8; 00 cm 2 mm 2222» 2 2... 2 83 25... =2: 00 um 00 .223... 39.83 2. new 00 $9.82 82m? 953 00 825 52 323-9? 0. 3283 2.. cm 2 mm 28:2... 2 3. 2 83 9. can 22.. on. «2 Nn. new 01.0 5.5.2.2 v.92 m2: 22.. 2:28 .2. 2.? .8 8. mites“... a 25 f Ea 08 .em 5.; 832 .22 €va to 2.. of. .2 2 06 f mam an. .0 5.38.2 2.. .6 832.201... 8.62.2 3; 22.. Eu 2.... .228 .2, am? .8 2: 222282 a 8° N... 2.... own. _EN 5.; .232 .22 2.3 to 2... 0.3. _E 2 Own. f a. o... 3 Nn. .o 22.2.22 2. 225 22.. .8 cop 922282 a 23 N... 25 own. .2... 2.; .282 .22 2va 2.2. .2 2 06 f «um 00 25.... .oEamN... 89.83 2. new $9.82 Ema? on. 00 00 88a 22.. .25? on. 8288 22.. cm 2 mm 222% 9 3. 2 83 new... 22.. 0“. «v. .0839 2 2.25 25 oo 2 23.28.22 89.82 Emu? 8.322.. .2.» 25 8288 an. .853... om .< 8.3 222% 2 am 2 0.8.. 92... on. 3 6.2.8. 2 2.2.8.3 130 6:522... oEm... 22. 2.. ... 8.2. 89.82 E226 2.2.8 a 8; 2.2... .3288 on. .22.. T 2. N... om 28 a... 2 2. 2 88.. 2:3 0.. 3.8.: H 62.30.. ES» 5 comm wfitofiao. some .88. 2... 0.3. 2.3 2.; 06 .26 Na + N... .58 :2. ..o 0.322» .... 2. .88 N0.... 3.8.2 2. 00 25... .2506 ...5 22; 5.5.85 0n. 2...; 35% E06? on. ...a 82.2 .22. o. T .28. 8288 00 cm .< mm .2822» 2 2.. 2 88.. 2:8. 8+0... % c205. c. am 025 0.05 02.005 .2 cc 98 Ea... .552... E08 .2. 88 .252. 3 .2 0. 8m .... 8262 22; E222 005% am 9 05 Nn. 05 N.... .0 09.0% 92.585. 06 58 .8 on 23 06 ....2 2.; .208. .22 .53 am .2 2. Nn. + N... 2. 02.22 225 an. 2. fioEE Sn 0.... 5 55:22 22... 2.2.. .5 2.9.2.28 .25 2.? .8 22 s 222208. 2.2.500 8.. N... 2.... 322.300 ....» 2.2.82 ....2 .58. 2.. 0.2. .... 2 f 5. 5.5.8 E2. 0 .0 5.8528... 0.. on 2 mEoom 22.. gm 2.. 22.. N0 2... o... 2 8.2228 32m a 2. 2 «Emma 0.05 .No 05 ...... 5 22.1205. 06 ..28 .8 on 2.... 03 .22 2.; .282 .22 2.2 2.2. .... 2. Nat... % .228. .< 22.2.2. 131 800000. E2200 0.0... 00.0 00.... I0 00.00.... 000.0000 5 0E3. 3 .< mu 0.00020 0. ..m 2 000... .0E000 0.80... 00.8.00 00800.8 00... .000... 0. 0.... 0.0... 800000. E20»... 0.0.. 00.... 000 .0 I0 0.... 000.0000 «1. _oEzmT. n. .< mu 0000008 or :m 0. 000.. .0503 0.80.. moNoRo 000030 00.. N... 0 0...... 00... .0320 00E000 0.... .2. S .o .0050. T .o 02.282. 00080., 0.0.2.0 0.2. .08. ... 0.2. 22...... .o 8.20.2 20%. 2 m... 00 ..00 N... 2 an. 2 08.. 00.0.0 0.00.. 00.5.3 .005... .020 0020.20 f 0... .05 000.0 2.0. 000 .0... 00.8008 0E00 00 2 0E000 0.0.. .r ...0 2 0000 ._ >2. .2. 0.0 .... .0000 000.. 0>0.. 0.00.... ...0E000x0 0.0.. 00.5.00 .00... 00.000000... 00.0 0.0... .0... 0.00 0.203 2 02.0>000 0.00 0.80.. o ..< mm 200 000.. aEmd 0.80.. 00.3.5 |mmn 000.0000 «1. .0 _0E0N. .... 000 0.00 0.250 .0 00000.>0 2. 00... 0.0... 0.... 0. 00.... 0E=_00 00...... ...... 000. 0... 0. 0.00 800000. E228 0.0.4. 0.20.... 2 52.8 2 020000 0.8 2.8.. 00 ... 0L0 0.02.20 2 2 08.. 00.0.0 2.00.. 8.8.3 «I... E08»... 0... 0. 02.0.0080: 2 0:0 00 >0E 0.... .00 02.0.0000 .000... 0 020 0.0 0m .0 800000. E290 E00 02.00830 02.880 00 .2030. on .< 00 0.02.20 2 =0 2 08.. 2.5.0 00 00.00.... 2 300000. E220 009.0000 0m .2000... on .< mm 0.00008 9. 0m 0. 000.. 900.... 0m QUE—.5 Mm 50.2.00. .< ...—0:022 132 35.25 2 o_aE_m 8: 9.3g 23.5 :ano EEomr 05 :o 28 282 he woman $9.83 204. 5.; Eoetwaxo coach 26: a ma; $2... an? 263 moNoBot via .29: E2. mxmma :08 was of. $5 96m acoE_w 628m >3 m_ 35 5E 9mm 8 32a 8.x xwoa < Ema? m5 2 E 9:58. 22:2 E2228 “6 E285 05 02a Boa Ia m_ 805. .830: m E: 05 E 3 2 a: E0; 8225 + E298 22 In 2: So .58 co:&omum ESE—.9.» oz on .< mm 235% or 2 83 _oEzm 0:03 on 96 E: _oEsom 3258 05 :05 E222“. ma; mEF E296 Eo< .28 .5267. a 5 So 3983 <._ _oEsoN... on .< mm 2355 or 2 33 _oanm 283 22 oE: m>o 3:an E8537. c: on .< mm 2355, or 28 goo. _oEsom 283 E893 Eo< I”: mm; In 2t .8288 <._ BEEN... 9 f mm 43. f 2 2 83 BESS 263 .on. 53% BEEN 0n. + ho BE: 5 EB 8988 .8329: ho _oEam on .2 mm 22.5% or 2 83 _oEzoN 6533.: 38:3 «55598 25 .95 «ton _ S _oEzvve o>ao E293 Eo< uca mEoBoa E8 .8; EoEtwaxo £5. 2. .< mu 23:99 9. 9. goo; _oEsooP 283 3323 .6.E8v u< 5.88.2 133 A8823. 888$ Ewfifi 82. 52.5 ..o 3.8. 2% Base... .2. cm 2 mm 285% S a... 2 83 on 00 8.38 3 2a.. 0E8 o5 Sena 898%. E06? (new... 3 <4 .oano new 8988 (n. _oanNz m? .2 mm 9.353 or am 2 goo. Sour (n+5 3:28 3. £38.. 320 can 8: 23 can 056 c. :m coup—mam... 5.832 ho 69.8w. 2? :23 63 a c. 5.5.83 on 2 mEEoE «2 mm Cu .md... am 33 _oeau... 00 3.3.3 ma @3522 8: on 2:8 5.632 «oz. :05 .6267. >3 .oca 95E o 333 5.8.62 693%. Emu? 0E: ...oao 39.83 m.oE:m $5 265 3 .< mm to or am 2 goo. _oEsoim 00 38:8 ad m>wu .298 :3; 883 $3 .33 20E m 338 .55 895%. E2m>m 0E: some 3983 m_oE:m «a... 26% o. ..< mm 235% or :m 0.. goo. _oE:m+m 00 3.2.3 H 2.3 5.55598 .0 $.38 .620 o: 895%. «00 29... =25 to: .2. Eu 2238 E2 5 mfiEofi am 0.83 33.93 ad Boa .o 5.5.33 326:... E... ..ano noon mama oEom 39.0%. :8 83 So: coco 0: 3 96 o..a._EE_m :m «2.2. 3328 mm .628. 2 528.3 134 g £83 8: Sn 2:2..8 3.88.60 a 2.2: 2988 2:8 9:955 22> 80:83 232?. 2:: 5.522 on so: 8:62... um: 0.5.... on .< oo 23:2"... on :m 2 goo; on 00 83:3 mm on 2 .23: no; >3 as: $5 $9.83 2226 3% =5... .26 was» 253 2:28.32 8 2 nu 2.856 cm 3“. 2 33 8 an. 8523 a 80:82 2826 .023...” >8. 5:283 on 2 on 235% cu :m 2 33 on on. 8823 Nm f 5.3 95am; :8: 528 __a .6 2382 «26%. 69.82 222m 83 . 23 man 9.053... 8:283 00 om .< on 2356 on 3. 2 30.. on 00 025:3 a o. on E 3233 320 __a 02982 A«I 9.62. 0320:. 32 o: o9 89.82 229m 2 9.22.5 :23 9.0636 3283 00 om 2 cc 2356 on :m 2 goo; on 00 82.23 a .323 __o 0. com 328$ 222$ “a 2382 2:2.» on? 9823.0 8283 cu .< on. 2856 om am 2 33 on 00 83:3 wlw .0633... 2 E25 3:983 0. oo 2 33205 ... :2? .0237. 0. mm 5 828$ Eovém 8988 00 .22 $3... 8m 5.; 5: Em. ca 2 8.3 2855 om am 2 83 o... 00 SEE Ma 3283 00 _oEanN 2.5 :83 €9.82 22%.» :2 «ma. 20cm .23, t 5.3 3:83 2:0 on .< mu .2.; or am 2 33 on 00 8628 um 8:272: 820 on new 69.82 222$ 3883 Sun 58 3933 00 3:59.. on 2.4: mm 2353 or am 2 goo; on 00 09330 a .6203 2 52.8.2 135 0. oo 3 8 595 $982.83 En 825.228 898%. 82m? 28 8o... moon £9: .o>o 8:88 359:8 on .< SN «2853 om :m o. 83 on 00 8: :3 add .88: 9.8. 88. 8.3. 895%. Emu...» Eofixm «Soot .2: 2: .8. Eco 022. zoo om 2 8 28:2.» om 3. o. ooo.. cm. 00 8.8.3 M4 . 39.5%. E296 =23: .8 v.59. u .23 8.3.... on .... mm 285% on :m 2 ooo... on 00 8.8.8 8H o.o 8 2:23 895%. 8.883 8:88.. =2 8.. 5.6.3.8 o: ow ..< .8 «285% cu :m on 00 8.8.8 add 895%. 8293 ...:o 5.88.. on om. o. ooo.. on 8.8.8 3|. 8:8 398%. 8296 8.5.88 00 .88: 5.3 5296 95.85 om .... .8 2853 on :m o. ooo.. on 00 8.3.8 mg 5.58 :. no.8. mo; 8:on 35 5896 8:88: .2.; 23 .o: cow E2. 688...... Eofixm ooEomoo 6..oE:n.o-~.ovm_oE= We 8:65 on E .8 285% on :m 2 ooo.. on 00 8NF.8 mad E285 $9.8m. 628m 9353...... .25.... 28 8:938 on 2 .8 235% om am 2 83 on 00 8.3.8 3|. 3m 8a :0 ... ”.5 205 00 96539653 8989. cops 5.3 :83 .88, .3 3.52.8 A«... 8.238 898%. E06? co 8.88 B 832.2 8.268 .3 8.88.5 om .< .8 f on :m o. 83 on 00 8.8.8 add .628. 3 5.2.3.3. 136 8.5.88 965358.? 83 82.. new (a. o. 852.8 8208 2: A858. 5288 5.5.8 5 5....> :5 ooo. 5 .28 Co. 5.. g 8 .< .8 2858 8 am 2 goo. on 888,. 3.8.8 a. 28 c. 8858 58 a 5.3 N.... .o 53.8 on. 5 858 5.5.88 21.98.. .2.. .38 85.82% .3 8E 83 055 28 ozooo 2:8 .o 5:52 .858. E298 2.. 5.3 8988 2o; a..oE:...o... 8 .< .8 28:8 8 :m o. ooo. on .893 8.2.8 «Hm 5: o... .o 25.8 2: c. .8. o 83 22:. .9555 8.88 .o $98... 88:85 8.8 .288 82.. A858. .588 82.. 2o: 8988 Coma; .o 9553. F... 8 ..< .8 2858 8 am. o. 53 on can; 8.3.8 3|. 8288 00 2... o: N... 8.28 59. 8:5. 83 8208 A858. .58...." 5... 0&8 . 0 ) 39 getdata %get experiment 1 data 40 Va: zeros(5,1);temp=1 ; temp2=1 ;sheet=0; 41 k=[k1,Kr,Kg,Kl,k5]; 42 R=objectivef(k,tdata1 ,ydata1,Yo);Ro=R+10; 43 while R>.01; 44 dp=1; 45 a = randint(1,1,[1,4]); %randomly pick one to change 46 KTemp=k(a); 47 k(a)=k(a)+(0.5-rand)*k(a)*dp"2; 48 %solve ODE's from 0 to 1000 49 R=objectivef(k,tdata1,ydata1 ,Yo); 50 if (R < R0) ; %if improvement then keep it and save the conditions 51 KR(temp,:)=[R,k]; 52 temp=temp+1; 53 R0=R 54 else %restore original value and continue 55 KF(temp2,:)=[R,k]; 56 k(a)=KTemp; 57 temp2=temp2+1; 58 end 59 if temp2>1000 60 sheet=sheet+1 142 61 xlswrite( 'H20modeltempdata.xls', KF ,sheet) 62 temp2=1 63 end 64 Va(a)=Va(a)+1; [R0,a,temp2] 65 end 66 end 67 13M 1 function dy = larsf(t,Y,k) 2 global Mt Dt Nt NI Ng Ns Ht XX reactor_V 3 %calculate k values 4 y=Y; 5 y(1 :4)=exp(-U2000)*Y(1 :4); 6 k1 =k(1 );k2=k(1 )Ik(3);k3=k(1 )Ik(2);k4=k3lk(4);k5=k(5); 7 %k1=0;k2=0;k3=0;k4=0;k5=k(5); 8 % clear 9 dydy = zeros(13,1); 10 % calculate 11 V=reactor_V/1000;% vol in liters 12 %calculate each reaction rate pates 13 k1y7y7=k1‘y(7)'Y(7); k4y7y7=k4‘Y(7)‘Y(7) ; 14 >0<=lk2*y(5)*Y(5)N; k2'v(5)*Y(6)N; k2‘y(6)"Y(6)N; 1 5 k1 y7y7*y(1 )NN; k1 y7y7'y(2)NN; k1 y7y7‘y(3)NN; 16 k1 y7y7‘y(4)NN; k4y7y7‘y(10)NN; k4y7y7‘y(1 1 )NN; 1 7 k4y7y7*y(12)NN; k4y7y7‘y(1 3)NN; k3‘y(5)*y(8)N; 18 k3‘y(5)*Y(9)N; k3‘y(6)‘Y(8)N; k3*Y(6)'3/(9)N; 19 k5“y(10)‘y(13)N; k5l4*(y(11)+y(12))"2N];%try using average of the HDO and HOD 20 % difine the coat for all the ODE 21A=[100-10000000000000;... 143 220100-1000000000000;... 2301000-100000000000;... 24001000-10000000000;... 25-2-2021101100-1-10000;... 260-2-20112001100-1-100;... 27242-2-2-2-2-2-2-2-2222200;... 2800000001010-10-1000;... 29000000001010-10-100;... 300000000-10001000-11;... 31 00000000-10001 001 -1 32000000000-1 0001 01 -1 330000000000-10001-11]; 34 %calculate all the ODE's 35dy=A’)0(; 36 37 calfra i ns.m 1 % calculate fractions for each spieces 2 g(:,1:3)= [Y(:,1)/Ng, (Y(:,2)+Y(:,3))lNg, Y(:,4)lNg]; 3 l(:,1:3)= [Y(:,10)/Nl, (Y(:,11)+Y(:,12))INI, Y(:,13)INI]; 4 s(:,1:5)= [Y(:,5:6)le, Y(:,7)/Ns, (Y(:,8:9))INs]; 5 % calculate free site fraction and M-X M-OX fractions 6 ss(:,1:3)= [Y(:,7),(Y(:,5)+Y(:,6))/Ns, (Y(:,8)+Y(:,9))le ]; 7 % calculate HID ration in G,S,L 8 HD(:,1:3)= [(Y(:,1)+Y(:,2)I2+Y(:,3)I2)/Ng, (Y(:,5)+Y(:,8)).l(Ns-Y(:,7)), (Y(: , 1 0)+Y(: ,1 1 )I2+Y(:,12)/2)INI]; 144 H202modelfunctions.m 1 %this is a script -— getdate.m 2 %get experimental data 3 tdata1=xlsread('f:\che 801 \data.xls', 'data2', 'time2'); 4 ydata1=xlsread('f:\che 801\data.xls', 'data2','HDF2'); 5 % 6 function dy = larsf(t,y, k) 7 global Mt Dt Nt Nl Ng Ns Ht XX reactor_V 8 %calculate k values 9 k1 =k(1 );k2=k(1 )Ik(3);k3=k(1)Ik(2);k4=k3lk(4);k5=k(5); 10 % clear 11 dydy = zeros(13,1); 12 % calculate 13 V=reactor_V/1000;% vol in liters 14 %calculate each reaction rate pates 15 k1y7y7=k1‘y(7)*>'(7); k4y7y7=k4*y(7)"Y(7) ; 16 xx=lk2‘y(5)*Y(5)N; k2*y(5)"‘y(6)N; k2‘v(6)‘y(6)N; 17 k1 y7y7*y(1)NN; k1 y7y7*y(2)NN; k1 y7y7‘y(3)NN; 18 k1 y7y7‘y(4)NN; k4y7y7*y(10); k4y7y7*y(1 1 )NN; 19 k4y7y7*y(12)NN; k4y7y7‘y(13)NN; k3‘y(5)*y(8)N; 20 k3*Y(5)*y(9)N; k3‘Y(6)‘Y(8)N; k3*y(6)‘Y(9)N; 21 k5*y(10)*y(13)N; k5‘y(11)*y(12)/V]; 22 % difine the coeffor all the ODE 23A=[100-10000000000000;... 240100-1000000000000;... 2501000-100000000000;... 26001000-10000000000;... 27-2-2021101100-1-10000;... 280-2-20112001100-1-100;... 29242-2-2-2-2-2-2-2-2222200;... 3000000001010-10-1000;... 145 31000000001010-10-100;... 320000000-10001000-11;... 3300000000-10001001-1;... 34000000000-1000101-1;... 350000000000-10001-11]; 36 %calculate all the ODE's 37dy=A*)0(; 38% 39 function RR = objectivef(k,tdata,ydata,Yo); 40 global Mt Dt Nt NI Ng Ns Ht XX reactor_V 41 %solve ODE's from 0 to 1000 42 options = odeset( 'RelTol',.00000001,'MaxStep' .10); 43 [t,Y] = ode15$(@(t,Y) larsf(t,Y,k),[0 995],Yo,options); 44 %calculate fractional info 45 calfractions; 46 % extrapolate Experimental data to calculated data 47 Yexp = interp1(tdata,ydata,t, 'pchip'); 48 %calculate least squares 49 R=(g-Yexp)."2; 50 %add all least square values 51 RR=sum(sum(R, 'double'), 'double'); 146 10. References B. Hogle, D. S., K Nagarajan, J.E. Jackson, D.J. Miller Formation and recovery of itaconic acid from aqueous solutions of citraconic acid and succinic acid. Industrial and Engineering Chemistry Research 2002, 41, (9), 2069-2073. Z. Zhang, J. E. J., D.J. Miller, Aqueous-phase hydrogenation of lactic acid to propylene glycol. Applied Catalysis A: General 2001, 219, 89-98. Jere, F. T.; Miller, D. J.; Jackson, J. E., Stereoretentive C-H bond activation in the aqueous phase catalytic hydrogenation of amino acids to amino alcohols. Organic Letters 2003, 5 (4), 527-530. Chopade, S. P.; Miller, D. J.; Jackson, J. E.; Werpy, T. A.; Frye, J. G., Jr.; Zacher, A. H. Catalysts and process for hydrogenolysis of sugar alcohols to polyols. 2001-US536,2001066499, 2001, 2001. Shafer, M. W. Catalytic processing of monosaccharides and sugar alcohols to value-added products. Michigan State University, East Lansing, 2004. Kovacs, D. 6.; Jackson, J. E.; Miller, D. J., Mechanistic investigation of glycerol hydrogenolysis. Abstracts of Papers of the American Chemical Society 2001 , 221, U177-U178. Crabtree, S. P.; Tyers, D. V. Process for the catalytic hydrogenolysis of sugar feedstocks into glycols and glycerol. 2004-GB4391,2005051874, 2004, 2005. Suppes, G. J.; Sutterlin, W. R.; Dasari, M. A. Method of producing lower alcohols from glycerol. 2005-U89901,2005095536, 20050324., 2005. Lahr, D. G.; Shanks, B. H., Effect of sulfur and temperature on ruthenium- catalyzed glycerol hydrogenolysis to glycols. Journal of Catalysis 2005, 232, (2), 386-394. Lahr, D. G.; Shanks, B. H., lGnetic analysis of the hydrogenolysis of lower polyhydric alcohols: Glycerol to glycols. Industrial & Engineering Chemistry Research 2003, 42, (22), 5467-5472. 147 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. Kusunoki, Y.; Miyazawa, T.; Kunimori, K; Tomishige, K, Highly active metal- acid bifunctional catalyst system for hydrogenolysis of glycerol under mild reaction conditions. Catalysis Communications 2005, 6, (10), 645-649. Dasari, M. A.; Kiatsimkul, P. P.; Sutterlin, W. R.; Suppes, G. J., Low-pressure hydrogenolysis of glycerol to propylene glycol. Applied Catalysis a-General 2005, 281, (1-2), 225-231. Chaminand, J.; Djakovitch, L.; Gallezot, P.; Marion, P.; Pinel, C.; Rosier, C., Glycerol hydrogenolysis on heterogeneous mtalysts. Green Chemistry 2004, 6, (8), 359-361. Perosa, A.; Tundo, P., Selective Hydrogenolysis of Glycerol with Raney Nickel. Industrial & Engineering Chemistry Research 2005, 44, (23), 8535-8537. Montassier, C.; Menezo, J. C.; Hoang, L. C.; Renaud, C.; Barbier, J., Aqueous polyol conversions on ruthenium and on sulfur-modified ruthenium. Journal of Molecular Catalysis 1991, 70, (1 ), 99-110. Hahn, R.; Tscheliessnig, A; Zochling, A.; Jungbauer, A., Shallow bed adsorption: Theoretical background and applications. Chemical Engineering & Technology 2005, 28, (1 1), 1241 -1 251. Alduri, B., A Review in Equilibrium in Single and Multicomponent Liquid Adsorption Systems. Reviews in Chemical Engineering 1995, 11, (2), 101-143. Comelissen, G.; Gustafsson, 0.; Bucheli, T. D.; Jonker, M. T. 0.; Koelmans, A. A.; Van Noort, P. C. M., Extensive sorption of organic compounds to black carbon, coal, and kerogen in sediments and soils: Mechanisms and consequences for distribution, bioaccumulation, and biodegradation. Environmental Science & Technology 2005, 39, (18), 6881 -6895. Smith, E. H., Evaluation of Multicomponent Adsorption Equilibria for Organic Mixtures onto Activated Carbon. Water Research 1991, 25, (2), 125-134. Xing, B. S.; Pignatello, J. J.; Gigliotti, 8., Competitive sorption between atrazine and other organic compounds in soils and model sorbents. Environmental Science & Technology 1996, 30, (8), 2432-2440. 148 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. Li, F. S.; Yuasa, A.; Ebie, K; Azuma, Y.; Hagishita, T.; Matsui, Y., Factors affecting the adsorption capacity of dissolved organic matter onto activated carbon: modified isotherm analysis. Water Research 2002, 36, (18), 4592-4604. Antonjuk, A. A; Redkovsky, N. N.; Cari, P. S.; Seidel, A.; Marutovsky, R. M., Determination of lntraparticle lfinetic-Parameters for Adsorption of Binary- Solutions on Activated Carbon. Chemical Engineering Science 1991, 46, (4), 1 035-1039. Leitao, A.; Serrao, R., Adsorption of phenolic compounds from water on activated carbon: Prediction of multicomponent equilibrium isotherms using single-component data. Adsorption-Joumal of the lntemational Adsorption Society 2005, 1 1, (2), 167-179. Abburi, K, Adsorption of phenol and p—chlorophenol from their single and bisolute aqueous solutions on Amberlite XAD-16 resin. Journal of Hazardous Materials 2003, 1 05, (1 -3), 143-1 56. Yaacoubi, A.; Mazet, M.; Dusart, 0., Competitive Effect in Bi-Solute Adsorption onto Activated Carbon - Dss, Alcohols and Phenols as Solutes. Water Research 1991, 25, (8), 929-937. Lu, Q. L.; Sorial, G. A, Adsorption of phenolics on activated carbon - impact of pore size and molecular oxygen. Chemosphere 2004, 55, (5), 671 -679. Dekock, F. P.; Vandeventer, J. S. J., Stochastic—Model for Equilibrium Adsorption onto Activated Carbon. Chemical Engineering Journal and the Biochemical Engineering Journal 1 995, 59, (3), 205-220. Lee, J. W.; Kwon, T. 0.; Moon, I. 8., Adsorption of mono saccharides, disaccharides, and maltooligosaccharides on activated carbon for separation of maltopentaose. Carbon 2004, 42, (2), 371 -380. Liu, X.; Pinto, N. 6., Ideal adsorbed phase model for adsorption of phenolic compounds on activated carbon. Carbon 1997, 35, (9), 1387-1397. Kumar, A.; Kumar, S.; Kumar, S., Adsorption of resorcinol and catechol on granular activated carbon: Equilibrium and kinetics. Carbon 2003, 41, (15), 3015- 3025. 149 31. 32. 33. 35. 37. 38. 39. Lu, Q. L.; Sorial, G. A, The role of adsorbent pore size distribution in multicomponent adsorption on activated carbon. Carbon 2004. 42, (15), 3133- 3142. Braeken, L.; Boussu, K; Van der Bruggen, 8.; Vandecasteele, C., Modeling of the adsorption of organic compounds on polymeric nanofiltration membranes in solutions containing two compounds. Chemphyschem 2005, 6, (8), 1606-1612. Frimmel, F. H.; Assenmacher, M.; Kumke, M. U.; Specht, C.; Abbt-Braun, G.; Grabe, 6., Removal of hydrophilic compounds from water with organic polymers - Part II: Adsorption behavior of industrial wastewater. Chemical Engineering and Processing 2002, 41, (9), 731 -736. Matsui, Y.; Yuasa, A; Ariga, K, Removal of a synthetic organic chemical by PAC-UF systems - l: Theory and modeling. Water Research 2001, 35, (2), 455- 463. Bonalumi, N.; Vargas, A; Ferri, D.; Burgi, T.; Mallat, T.; Baiker, A, Competition at chiral metal surfaces: Fundamental aspects of the inversion of enantioselectivity in hydrogenations on platinum. Journal of the American Chemical Society 2005, 127, (23), 8467-8477. Vargas, A; Ferri, D.; Baiker, A, DFT and ATR-IR insight into the conformational flexibility of cinchonidine adsorbed on platinum: Proton exchange with metal. Journal of Catalysis 2005, 236, (1 ), 1-8. Keresszegi, C.; Ferri, D.; Mallat, T.; Baiker, A, Unraveling the surface reactions during liquid-phase oxidation of benzyl alcohol on Pd/Al203: an in situ ATR-IR study. Journal of Physical Chemistry B 2005, 109, (2), 958-967. Hamminga, G. M.; Mul, G.; Moulijn, J. A, Real-time in situ ATR-FI’IR analysis of the liquid phase hydrogenation of gamma-butyrolactone over Cu-ZnO catalysts: A mechanistic study by varying lactone ring size. Chemical Engineering Science 2004, 59, (22-23), 5479-5485. Lefevre, G., In situ F ourier-transform infrared spectroscopy studies of inorganic ions adsorption on metal oxides and hydroxides. Advances in Colloid and Interface Science 2004, 107, (2-3), 109-123. 150 40. 41. 42. 43. 45. 47. 48. 49. Mul, G.; Hamminga, G. M.; Moulijn, J. A, Operando ATR-FI' IR analysis of liquid- phase catalytic reactions: can heterogeneous catalysts be observed? Vibrah’onal Spectroscopy 2004, 34, (1 ), 109-121. Hind, A. R.; Bhargava, S. K; McKinnon, A, At the solid/liquid interface: FTIRIATR - the tool of choice. Advances in Colloid and Interface Science 2001, 93, (1-3), 91-114. Ninness, B. J.; Bousfield, D. W.; Tripp, C. P., In situ infrared technique for studying adsorption onto particulate silica surfaces from aqueous solutions. Applied Spectroscopy 2001, 55, (6), 655-662. Lazorenko-Manevich, R. M., Adatom hypothesis as a predominant mechanism of surface enhanced Raman scattering: A review of experimental argumentation. Russian Journal of Electrochemistry 2005, 41 , (8), 799-831. Tian, Z 0.; Ron, 8, Adsorption and reaction at electrochemical interfaces as probed by surface-enhanced Raman spectroscopy. Annual Review of Physical Chemistry 2004, 55, 197-229. Brolo, A. G.; Irish, D. E.; Smith, 8. D., Applications of surface enhanced Raman scattering to the study of metal-adsorbate interactions. Journal of Molecular Structure 1997, 405, (1 ), 29-44. Urban, M. W., Surface and Interface Vibrational Spectroscopy - Relevance to Adhesion. Journal of Adhesion Science and Technology 1993, 7, (1 ), 1-47. Otto, A, Surface-Enhanced Raman-Scattering of Adsorbates. Journal of Raman Spectroscopy 1991, 22, (12), 743-752. Guo, Q. L.; Wang, E. G., Geometric and electronic structure of Cu on corundum (0001) surfaces. Science and Technology of Advanced Materials 2005, 6, (7), 795-798. Reniers, F.; Tewell, C., New improvements in energy and spatial (x, y, 2) resolution in AES and XPS applications. Journal of Electron Spectroscopy and Related Phenomena 2005, 142, (1 ), 1-25. 151 50. 51. 52. 53. 55. 57. 58. 59. Duwez, A S., Exploiting electron spectroscopies to probe the structure and organization of self-assembled monolayers: a review. Journal of Electron Spectroscopy and Related Phenomena 2004, 134, (2-3), 97-138. Marcus, P.; Galtayries, A, ESCA and metal surfaces - Alloys, corrosion and surface treatments. Vide-Science Technique Et Applications 2003, 58, (308), 230-+. Sun, C. H.; Berg, J. C., A review of the different techniques for solid surface acid-base characterization. Advances in Colloid and Interface Science 2003, 105, 151-175. Pollini, l.; Mosser, A; Pariebas, J. C., Electronic, spectroscopic and elastic properties of eariy transition metal compounds. Physics Reports-Review Section of Physics Letters 2001, 355, (1 ), 1-72. Zemek, J., Electron spectroscopy of solid surfaces. Acta Physica Slovaca 2000, 50, (5), 577-589. Ronning, C.; Felderrnann, H.; Merk, R.; Hofsass, H.; Reinke, P.; Thiele, J. U., Carbon nitride deposited using energetic species: A review on XPS studies. Physical Review B 1998, 58, (4), 2207-2215. Weitzsacker, C. L.; Xie, M.; Drzal, L. T., Using XPS to investigate fiber matrix chemical interactions in carbon-fiber-reinforced composites. Surface and Interface Analysis 1997, 25, (2), 53-63. Wang, J. X.; Marinkovic, N. S.; Zajonz, H.; Ocko, B. M.; Adzic, R. R., In situ X- ray reflectivity and voltammetry study of Ru(0001) surface oxidation in electrolyte solutions. Journal of Physical Chemistry B 2001 , 105, (14), 2809-2814. Ciobica, l. M.; Kleyn, A. W.; Van Santen, R. A, Adsorption and coadsorption of CO and H on ruthenium surfaces. Journal of Physical Chemistry B 2003, 107, (1 ), 164-172. Riedmuller, B.; Ciobica, I. M.; Papageorgopoulos, D. C.; Frechard, F.; Berenbak, B.; Kleyn, A. W.; van Santen, R. A, CO adsorption on hydrogen saturated Ru(0001). Journal of Chemical Physics 2001, 115, (11), 5244-5251. 152 60. 61. 62. 63. 65. 67. 68. 69. 70. Peebles, D. E.; Schreifels, J. A; White, J. M., The Interaction of Coadsorbed Hydrogen and Carbon-Monoxide on Ru(001). Surface Science 1982, 116, (1 ), 1 1 7-1 34. Pallassana, V.; Neurock, M., Reaction paths in the hydrogenolysis of acetic acid to ethanol over Pd(111), Re(0001), and PdRe alloys. Journal of Catalysis 2002, 209, (2), 289-305. Haile, J. M., Molecular dynamics simulation : elementary methods. Wiley: New York, 1992; p xvii, 489 p. Jacob, T.; Goddard, W. A, Agostic interactions and dissociation in the first layer of water on Pt(111). Journal of the American Chemical Society 2004, 126, (30), 9360-9368. Ogasawara, H.; Brena, B.; Nordlund, D.; Nyberg, M.; Pelmenschikov, A; Pettersson, L. G. M.; Nilsson, A, Structure and bonding of water on Pt(111). Physical Review Letters 2002, 89, (27). Ito, M.; Nakamura, M., Hydration processes of electrolyte anions and cations on Pt(111), Ir(111), Ru(001) and Au(111) surfaces: Coadsorption of water molecules with electrolyte ions. Faraday Discussions 2002, 121, 71 -84. Thor Dozeman, G. J., TLE Catalytic conversion of D-glucitol and isosorbide using zeolite. Dozeman, G. J., Catalytic conversion of D-glucitol and isosorbide using zeolite catalysts. In 1989. Chen, Y. Aqueous-Phase Catalytic Hydrogenation of Organic Acids and Theit Mixtures. Michigan State University, East Lansing, 2007. Parsons, M. T.; Westh, P.; Davies, J. V.; Trandum, C.; To, E. C. H.; Chiang, W. M.; Yee, E. G. M.; Koga, Y., A thermodynamic study of 1-propanol-glyceroI-H20 at 25 degrees C: Effect of glycerol on molecular organization of H20. Journal of Solution Chemistry 2001 , 30, (1 1 ), 1007-1028. Parsons, M. T.; Koga, Y., Interactions in 1-propanoI-(1,2-and 1,3-)propanediol- H20: The effect of hydrophobic vs hydrophilic moiety on the molecular 153 71. 72. 73. 74. organization of H20. Journal of Physical Chemistry B 2002, 106, (28), 7090- 7095. Hansch, C.; Hoekman, 0; Leo, A; Zhang, L. T.; Li, P., The Expanding Role of Quantitative Stmcture-Activity-Relationships (Qsar) in Toxicology. Toxicology Letters 1995, 79, (1-3), 45-53. Sorial, G. A; Suidan, M. T.; Vidic, R. D.; Maloney, S. W., Competitive Adsorption of Phenols on Gac .1. Adsorption Equilibrium. Journal of Environmental Engineering-Asce 1 993, 1 19, (6), 1026-1043. McKay, G.; Alduri, B., Multicomponent Dye Adsorption onto Carbon Using a Solid Diffusion Mass-Transfer Model. Industrial & Engineering Chemistry Research 1991, 30, (2), 385-395. Yaws, C. L. N., Prasad, K; Gabbula, Chaitanya . Yaws' Handbook of Antoine Coefficients for Vapor Pressure (Electronic Edition). Knovel. 154 Illllllllllll