Vous êtes sur la page 1sur 271

ABSTRACT

The energy demand of the world has been exponentially increasing over recent years caused by economic development and increasing population. Fossil fuels have been the dominant sources of power. However, distribution of fossil fuels is largely geographically dependent. Combustion of fossil fuels generates a variety of air pollutants, which contribute to global warming and have caused major environmental concerns. Hydrogen has emerged as an alternative energy source to replace fossil fuels because it can be produced from a variety of sources including renewable sources and combustion of hydrogen produces only water, making it environmentally friendly. Catalytic hydrogen production technology is an integral part of the future energy portfolio. Research concentrates on improving process efficiencies and reducing emission levels for hydrogen production from fossil fuel and investigating effective ways to generate hydrogen from renewable sources. This work examines catalytic hydrogen production from water-gas shift reaction via coal-derived synthesis gas and steam reforming of propane and bioethanol. Efforts have been made in catalyst development, performance evaluation and catalyst characterization to elucidate active sites and reaction network during catalytic reactions. ii

Fe-Cr catalysts are used in commercial high temperature water-gas shift reactors. This catalyst requires high temperatures to have sufficient activity. However, high temperatures lead to catalyst sintering and limit the conversion due to thermodynamic equilibrium. This brings about the need for a second shift reactor, which will operate at a lower temperature and push the thermodynamic limitation for the conversion. The two-stage operation makes this process costly and impractical. A small amount of Cr6+ is contained in the commonly used Fe-Cr formulation and causes environmental concerns. In this research, Fe-Al-Cu was developed to replace Fe-Cr catalysts for water-gas shift reaction. Catalyst preparation methods were found to play a critical role in determining catalytic performance. Different from conventional impregnation and coprecipitation methods, sol-gel technique was introduced for Fe-Al-Cu catalyst preparation. Sol-gel catalysts demonstrated high activities over a wide range of reaction temperatures compared with commercial catalysts. Characterization results revealed that Cu promoter was uniformly distributed in the iron oxide structure for sol-gel prepared catalysts and iron oxides with different crystal structures were formed in sol-gel catalysts as well. These properties contribute to high catalyst activity and stability. Cu loading amount was optimized as an important synthesis parameter and catalyst performance was further improved. In addition to these efforts, catalysts were evaluated in the iii

presence of H2S, which is a major impurity from coal derived synthesis gas. A series of Fe-based catalysts were compared for their sulfur resistance and a deactivation mechanism was proposed based on characterization studies for both fresh and poisoned catalysts. The Cu component in Fe-Al-Cu catalysts was found to be more susceptible to H2S and sulfidation of Cu leads to initial catalyst activity loss. Another focus of this research is on hydrogen production from steam reforming. Propane steam reforming was studied over Ni-Al2O3 catalysts that were prepared by a conventional impregnation method and a one-step sol-gel technique. Impregnated Ni-Al2O3 catalysts showed high initial activity but deactivated severely with time-on-stream of propane steam reforming whereas sol-gel Ni-Al2O3 catalysts maintained relatively stable performance under the same testing conditions. It was found that sol-gel preparation yields strong interaction between the active metal and the support, thus forming highly dispersed small Ni crystallites on the surface. This was proved to be beneficial for coke suppression and catalyst stability. Similarly in Co-ZrO2 ethanol steam reforming studies, sol-gel prepared Co-ZrO2 catalysts exhibited higher coking resistance. The phase evolution with temperature for sol-gel Co-ZrO2 catalysts illustrates the crystallite transition from cubic to tetragonal and finally to monoclinic ZrO2. Co was found capable of stabilizing cubic ZrO2 phase, which iv

may possibly explain the stability of sol-gel Co-ZrO2 catalysts. It was demonstrated that support basicity/acidity correlates with product distribution and catalyst stability. CeO2 was combined with ZrO2 to support Co metals. This mixed oxides catalyst showed less coking and higher activity and stability during ethanol steam reforming.

Dedicated to my husband, Guoyong Sun

vi

ACKNOWLEDGMENTS

I acknowledge my adviser, Umit S. Ozkan, for her support throughout my PhD studies. She walked me into the catalysis world. Without her guidance in research and help in life, everything today would not be possible. I want to express my great appreciation to Dr. Rick B. Watson. Thanks to him that I was able to learn all laboratory instruments in a short time and his way of doing things influences me during my PhD life. I am grateful to Dr. Xueqin Wang for his supervision on the project and stimulating discussions. I enjoy the Happy Hours with Paul Matter, Erik Holmgreen, Matt Yung, John Kuhn and Matt Woods. They never left out the only female group member during my first two years. My communication skills and understanding of American culture benefit from those experiences. I also want to thank Dr. Bing Tan, Hua Song, Preshit Gwades collaboration on the same projects. I want to acknowledge all the other group members. We create a great group atmosphere together to get our research going. I enjoy every minute in the group. Finally, I want to thank my husband, Guoyong Sun, and my family members in China, whose love and morale support help me survive difficult days during the past five years.

vii

VITA

April 27, 1981Born-Shijiazhuang, Hebei, P.R.China

July, 2003..B.S Biochemical Engineering Tianjin University Tianjin, P.R.China September, 2003-presentGraduate Research Associate The Ohio State University Columbus, Ohio

PUBLICATIONS

1. Natesakhawat, S., Wang, X., Zhang, L., and Ozkan, U.S., Development of chromium-free iron-based catalysts for high-temperature water-gas shift reaction, Journal of Molecular Catalysis A: Chemical 260 (2006) 82-94 2. Song, H., Zhang, L., Watson, R.B., Braden, D., and Ozkan, U.S., Investigation of bio-ethanol steam reforming over cobalt-based catalysts, Catalysis Today 129 (2007) 346-354 3. Song, H., Zhang, L., and Ozkan, U.S., Effect of synthesis parameters on the catalytic activity of Co-ZrO2 for bio-ethanol steam reforming, Green Chemistry 9 (2007) 686-694

viii

FIELDS OF STUDY

Major Field: Chemical Engineering Area of Interest: Heterogeneous Catalysis

ix

TABLE OF CONTENTS

Abstract...ii Dedicationvi Acknowledgments.....vii Vita..viii Table of contents.x List of tables..xviii List of figures..xx

PART 1: Hydrogen production via the water-gas shift reaction from coal derived synthesis gas........................................................................................................ 1 CHAPTER 1: Introduction to water-gas shift for hydrogen production from coal .. 2 CHAPTER 2: Literature review ............................................................................. 6 2.1 Overview of high temperature water-gas shift catalysts ............................. 7 2.2 Overview of low temperature water-gas shift catalysts............................. 11 2.3 Overview of sour gas shift catalysts ......................................................... 15 2.4 Overview of water-gas shift reaction mechanism ..................................... 15

CHAPTER 3: Experimental methods.................................................................. 19 3.1 Catalyst preparation ................................................................................. 19 3.1.1 Precipitation-impregnation ................................................................ 19 3.1.2 Co-precipitation................................................................................. 20 3.1.3 Sol-gel............................................................................................... 21 3.2 Catalyst characterization .......................................................................... 21 3.2.1 N2 physisorption analysis .................................................................. 22 3.2.2 Temperature programmed reduction ................................................ 22 3.2.3 Temperature programmed reaction .................................................. 22 3.2.4 X-ray photoelectron spectroscopy..................................................... 23 3.2.5 Diffuse reflectance infrared Fourier transform spectroscopy............. 23 3.2.6 X-ray diffraction................................................................................. 24 3.2.7 Mssbauer spectroscopy .................................................................. 24 3.3 Reaction testing........................................................................................ 25 3.3.1 Steady state reaction ........................................................................ 26 3.3.2 Catalyst evaluation in the presence of H2S ....................................... 27 CHAPTER 4: Investigation of highly active Fe-Al-Cu catalysts for water-gas shift reaction............................................................................................................... 28 4.1 Overview of Fe-based catalysts for water-gas shift reaction .................... 28 4.2 Experimental procedures ......................................................................... 30 xi

4.2.1 Catalyst preparation .......................................................................... 30 4.2.2 Reaction studies ............................................................................... 31 4.2.3 Catalyst characterization................................................................... 31 4.3 Results and discussion............................................................................. 34 4.3.1 Reaction data for Fe-Al-Cu catalysts prepared by different methods 34 4.3.2 Physisorption isotherm and pore size distribution measurement ...... 40 4.3.3 Temperature programmed reduction of Fe-Al-Cu catalysts .............. 44 4.3.4 XPS analysis of Fe-Al-Cu samples ................................................... 47 4.3.5 XRD of Fe-Al-Cu catalysts ................................................................ 51 4.3.6 Mssbauer spectra of Fe-Al-Cu catalysts ......................................... 54 4.3.7 Temperature programmed reaction of Fe-Al-Cu catalysts ................ 58 4.4 Conclusions.............................................................................................. 60 CHAPTER 5: Effects of Cu loading on the catalytic performance of Fe-Al-Cu for water-gas shift reaction ...................................................................................... 62 5.1 Overview of catalyst synthesis and Cu promoter in water-gas shift catalysis ....................................................................................................................... 62 5.2 Experimental procedures ......................................................................... 64 5.2.1 Catalyst preparation .......................................................................... 64 5.2.2 Reaction studies ............................................................................... 64 5.2.3 Catalyst characterization................................................................... 65 xii

5.3 Results and discussion............................................................................. 67 5.3.1 Reaction performance for Fe-Al-Cu catalysts with different Cu loadings ..................................................................................................... 67 5.3.2 Surface area and pore volume measurement ................................... 68 5.3.3 In-situ X-ray diffraction during catalyst reduction .............................. 72 5.3.4 Bulk structure analysis by Mssbauer spectroscopy......................... 80 5.4 Conclusions.............................................................................................. 86 CHAPTER 6: Deactivation characteristics of Fe-Al-Cu water-gas shift catalysts in the presence of H2S ........................................................................................... 87 6.1 Overview of water-gas shift catalyst deactivation studies in the presence of H2S................................................................................................................. 87 6.2 Experimental procedures ......................................................................... 91 6.2.1 Catalyst preparation .......................................................................... 91 6.2.2 Reaction studies. .............................................................................. 91 6.2.3 Catalyst characterization................................................................... 92 6.3 Results and discussion............................................................................. 93 6.3.1 Catalytic performance in the presence of H2S .................................. 93 6.3.2 Surface interaction between reactants and catalysts surface probed by DRIFTS ................................................................................................. 97 6.3.3 Surface changes characterized by XPS.......................................... 101 xiii

6.3.4 Phase composition and oxidation states examined by Mssbauer spectroscopy............................................................................................ 109 6.3.5 Crystal phases examined by XRD analysis..................................... 114 6.4 Conclusions............................................................................................ 117 CHAPTER 7: Conclusions and recommendations............................................ 118 7.1 Summary of water-gas shift work ........................................................... 118 7.2 Future work for water-gas shift catalyst development ............................ 120 PART 2: Hydrogen production from propane and ethanol steam reforming ..... 122 CHAPTER 8: Introduction to steam reforming .................................................. 123 CHAPTER 9: Literature review on steam reforming ......................................... 125 9.1 Literature review of propane steam reforming........................................ 125 9.2 Literature review of bio-ethanol steam reforming ................................... 128 9.2.1 Overview of support materials for ethanol reforming catalysts........ 130 9.2.2. Overview of active metals or promoters for ethanol reforming catalysts ................................................................................................... 134 9.2.3 Overview of active sites and reaction mechanism during ethanol steam reforming ....................................................................................... 136

xiv

CHAPTER 10: Effect of preparation method on structural characteristics and propane steam reforming performance of nickel supported on alumina catalysts ......................................................................................................................... 139 10.1 Overview of steam reforming catalyst preparation ............................... 139 10.2 Experimental procedures ..................................................................... 140 10.2.1 Catalyst preparation ...................................................................... 140 10.2.2 Characterization ............................................................................ 141 10.2.3 Reaction studies ........................................................................... 144 10.3 Results and discussion......................................................................... 144 10.3.1 Physical surface area and metallic Ni surface area ...................... 144 10.3.2 XPS of Ni species in calcined and reduced catalysts.................... 146 10.3.3 X-ray diffraction ............................................................................. 149 10.3.4 TPR profiles of Ni-Al2O3 catalysts ................................................. 154 10.3.5 Steam reforming of propane over reduced Ni-Al2O3 catalysts....... 157 10.3.6 Coke deposition characterization on post-reaction catalysts......... 160 10.3.7 TEM study on post-reaction samples ............................................ 165 10.3.8 Coking models of propane steam reforming ................................. 168 10.4 Conclusions.......................................................................................... 171 CHAPTER 11: Investigation of bio-ethanol steam reforming over cobalt-based catalysts ........................................................................................................... 172 xv

11.1 Overview of bio-ethanol steam reforming over cobalt-based catalysts. 172 11.2 Overview of effect of synthesis parameters on the catalytic activity of CoZrO2 for ethanol steam reforming ................................................................. 176 CHAPTER 12: Studies on sol-gel prepared cobalt supported on zironia catalysts for ethanol steam reforming.............................................................................. 179 12.1 Overview of sol-gel preparation studies ............................................... 179 12.2 Experimental procedures ..................................................................... 180 12.2.1 Catalyst preparation ...................................................................... 180 12.2.2 Catalyst characterization............................................................... 181 12.3 Results and discussion......................................................................... 183 12.3.1 Temperature programmed reduction............................................. 183 12.3.2 X-ray diffraction patterns of Co-ZrO2 catalysts .............................. 186 12.3.3 Temperature programmed reaction............................................... 192 12.3.4 Diffuse reflectance infrared Fourier transform spectroscopy......... 198 12.4 Conclusions.......................................................................................... 200 CHAPTER 13: Modification of cobalt supported on zirconia catalysts for ethanol steam reforming................................................................................................ 201 13.1 Overview of Co-ZrO2 catalysts modification for ethanol steam reforming ..................................................................................................................... 201 xvi

13.2 Experimental procedures ..................................................................... 201 13.2.1 Catalyst preparation ...................................................................... 201 13.2.2 Catalyst characterization............................................................... 202 13.2.3 Reaction testing ............................................................................ 204 13.3 Results and discussion......................................................................... 205 13.3.1 Sol-gel Co-ZrO2 catalysts with different Co loadings..................... 205 13.3.2 Examination of CeO2 to modify ZrO2 support properties and improve activity and stability .................................................................................. 211 13.3.3 Evaluation of catalysts during ethanol steam reforming and surface acidity measurement................................................................................ 217 13.4 Conclusions.......................................................................................... 222 CHAPTER 14: Conclusions and recommendations.......................................... 223 14.1 Summary of propane steam reforming work ........................................ 223 14.2 Future work on propane steam reforming ............................................ 225 14.3 Summary of ethanol steam reforming work.......................................... 226 14.4 Future work on ethanol steam reforming.............................................. 227 REFERENCES ................................................................................................. 228 APPENDIX A .................................................................................................... 245

xvii

LIST OF TABLES Table 2.1: Catalysts used in HT-WGS process .................................................. 10 Table 2.2: Catalysts used in LT-WGS process ................................................... 14 Table 3.1: GC oven temperature program.......................................................... 26 Table 3.2: GC valve-switching program.............................................................. 26 Table 4.1: Surface area and pore volume of Fe-Al-Cu catalysts ........................ 41 Table 4.2: pH values at the onset of precipitation for components in Fe-Al-Cu catalysts ............................................................................................................. 47 Table 4.3: Mssbauer parameters calculated from spectra of Fe-Al-Cu catalysts collected at 25 C ............................................................................................... 56 Table 5.1: BET surface area and pore volume measurements for Fe-Al-Cu catalysts with different Cu loadings .................................................................... 71 Table 5.2: Mssbauer parameters calculated from spectra of Fe-based catalysts collected at 25C ................................................................................................ 82 Table 6.1: XPS binding energies of phases present in Fe only and Fe-Al-Cu catalysts ........................................................................................................... 107 xviii

Table 6.2: Mssbauer parameters derived from the sample spectra collected at 25C ................................................................................................................. 112 Table 9.1: Supports used in ethanol steam reforming catalyst preparations .... 133 Table 9.2: Active metals used in ethanol steam reforming catalyst preparations ......................................................................................................................... 136 Table 10.1: Surface area and metallic Ni surface area..................................... 146 Table 10.2: Content of Ni species for catalysts prepared by impregnation and solgel methods: calcined and reduced samples.................................................... 148 Table 10.3: Ni crystallite size after reduction at different temperatures ............ 154

xix

LIST OF FIGURES Figure 4.1: Reaction rate comparison of Fe-Al-Cu catalysts prepared by different methods: the inset is CO conversion rate per physical surface area per minute for Fe-Al-Cu catalysts .............................................................................................. 38 Figure 4.2: CO conversion between sol-gel Fe-Al-Cu catalyst and commercial Fe-Cr-Cu catalyst ............................................................................................... 39 Figure 4.3: (a) Adsorption-desorption isotherms of N2 at 77 K for Fe-Al-Cu catalysts. Solid symbols: adsorption isotherms; Open symbols: desorption isotherms. 1-step isotherms are shifted vertically by 100; P0 is the saturation vapor pressure of N2 at 77 K. (b) Pore size distribution for Fe-Al-Cu catalysts obtained from BJH adsorption ............................................................................ 42 Figure 4.4: TPR profiles for Fe-based catalysts.................................................. 45 Figure 4.5: Fe 2p region of the X-ray photoelectron spectra for Fe-Al-Cu catalysts ........................................................................................................................... 49 Figure 4.6: Cu 2p region of the X-ray photoelectron spectra for Fe-Al-Cu catalysts ............................................................................................................. 50 Figure 4.7: XRD patterns of Fe-Al-Cu catalysts.................................................. 53 xx

Figure 4.8: Mssbauer spectrum for 1-step Fe-Al-Cu......................................... 57 Figure 4.9: H2O-TPRxn for Fe-Al-Cu catalysts (a) m/z=18 H2O consumption (b) m/z=2 H2 generation........................................................................................... 59 Figure 5.1: CO conversion for Fe-Al-Cu catalysts with different Cu loadings at various reaction temperatures ............................................................................ 69 Figure 5.2: Time-on-stream testing for Fe-Al-Cu catalysts with Fe/Cu=5 and Fe/Cu=1 using equal surface area in the reactor................................................ 70 Figure 5.3: In-situ X-ray diffraction patterns during reduction with 5% H2/N2 for Fe-Al-Cu catalyst (Fe/Cu=5). Inset: TPR profile for the same catalyst ............... 77 Figure 5.4: In-situ X-ray diffraction patterns during reduction with 5% H2/N2 for Fe-Al-Cu catalyst (Fe/Cu=1). Inset: TPR profile for the same catalyst .............. 78 Figure 5.5: X-ray diffraction patterns for Fe-based catalysts .............................. 79 Figure 5.6: Mssbauer spectrum for Fe-Al-Cu-SG with Fe/Al=10, Fe/Cu=20..... 83 Figure 5.7: Mssbauer spectrum for Fe-Al-Cu-SG with Fe/Al=10, Fe/Cu=5....... 84 Figure 5.8: Mssbauer spectrum for Fe-Cu-SG with Fe/Cu=5 ........................... 85 Figure 6.1: Reaction activity for both fresh and poisoned Fe-based catalysts.... 95 xxi

Figure 6.2: Comparison between Fe only and Fe-Al-Cu catalysts before and after 50 ppm H2S exposure for 24 hrs ........................................................................ 96 Figure 6.3: DRIFT spectra during CO and H2O temperature programmed reaction: (a) Fe-Al-Cu catalyst; (b) 50 ppm H2S poisoned Fe-Al-Cu catalyst .................. 100 Figure 6.4: X-ray photoelectron spectra for Fe-only catalysts after exposed to different concentrations of H2S: (a) Fe 2p region; (b) S 2p region .................... 106 Figure 6.5: X-ray photoelectron spectra for Fe-Al-Cu catalysts after exposed to different concentrations of H2S: (a) Fe 2p region; (b) Cu 2p region; (c) S 2p region ............................................................................................................... 108 Figure 6.6: Mssbauer spectra of: (a) freshly calcined Fe-Al-Cu catalysts; (b) FeAl-Cu catalysts poisoned with 1000 ppm H2S/N2 at 400oC for 72 hrs............... 111 Figure 6.7: X-ray diffraction pattern of commercial FeS after air exposure....... 113 Figure 6.8: X-ray diffraction patterns of freshly calcined and sulfur poisoned catalysts: (a) Fe-only catalyst; (b) Fe-Al-Cu catalysts....................................... 116 Figure 10.1: X-ray photoelectron spectra (Ni 2p region) of Ni-Al2O3 catalysts.. 147 Figure 10.2: X-ray diffraction patterns of calcined Ni-Al2O3 catalysts ............... 150

xxii

Figure 10.3: X-ray diffraction patterns of Ni-Al2O3 catalysts after reduction...... 151 Figure 10.4: Temperature-programmed reduction profiles of Ni-Al2O3 catalysts ......................................................................................................................... 156 Figure 10.5: Propane conversion as a function of time-on-stream in steam reforming reaction of propane over Ni-Al2O3 catalysts...................................... 159 Figure 10.6: Temperature-programmed oxidation profiles of carbon deposited on post-reaction Ni-Al2O3 catalysts........................................................................ 161 Figure 10.7: X-ray diffraction patterns for post-reaction Ni-Al2O3 catalysts....... 163 Figure 10.8: Raman spectra of calcined and post-reaction Ni-Al2O3 catalysts . 164 Figure 10.9: TEM images of post-reaction Ni-Al2O3 catalysts after propane steam reforming for 28 hours: (a) and (b) post-reaction impregnation catalyst; (c) and (d) post-reaction sol-gel catalyst ............................................................................ 167 Figure 10.10: Proposed coking models of propane steam reforming over reduced Ni-Al2O3 catalysts: (a) impregnation catalyst; (b) sol-gel catalyst ..................... 170 Figure 12.1: Temperature programmed reduction profiles for Co-ZrO2 catalysts: IWI catalyst and sol-gel catalysts calcined at different temperatures................ 185

xxiii

Figure 12.2: X-ray diffraction patterns of Co-ZrO2 catalysts: (a) IWI-Co-ZrO2; (b) SG-Co-ZrO2 calcined at different temperatures................................................ 188 Figure 12.3: In situ X-ray diffraction for sol-gel Co-ZrO2 catalyst calcined at 400oC during 5% H2/N2 reduction................................................................................ 189 Figure 12.4: In situ X-ray diffraction for sol-gel Co-ZrO2 catalyst calcined at 500oC during 5% H2/N2 reduction................................................................................ 190 Figure 12.5: X-ray diffraction patterns of ZrO2 supports calcined at 400oC and 600oC................................................................................................................ 191 Figure 12.6: Temperature programmed oxidation for post-reaction Co-ZrO2 catalysts ........................................................................................................... 193 Figure 12.7: Temperature programmed reaction for IWI Co-ZrO2 catalyst ....... 196 Figure 12.8: Temperature programmed reaction for sol-gel Co-ZrO2 catalyst .. 197 Figure 12.9: DRIFT spectra during temperature programmed desorption: (a) IWI Co-ZrO2 catalysts; (b) sol-gel Co-ZrO2 catalysts .............................................. 199 Figure 13.1: In situ X-ray diffraction for sol-gel 10%Co-ZrO2 catalysts during reduction with 5% H2/N2 ................................................................................... 206

xxiv

Figure 13.2: In situ X-ray diffraction for sol-gel 30%Co-ZrO2 catalysts during reduction with 5% H2/N2 ................................................................................... 207 Figure 13.3: Ethanol steam reforming reaction data at different temperatures for sol-gel 20%Co-ZrO2 catalysts........................................................................... 210 Figure 13.4: (a) m/z=44 signal evolution during CO2 pulse chemisorption on CeO2 support; (b) calculated CO2 adsorption amount comparison................... 212 Figure 13.5: Ethanol+H2O TPD in situ DRIFTS on ZrO2 support...................... 215 Figure 13.6: Ethanol+H2O TPD in situ DRIFTS on CeO2 support..................... 216 Figure 13.7: Ethanol steam reforming reaction data at different temperatures for sol-gel 20%Co-40%CeO2-40%ZrO2 catalysts .................................................. 218 Figure 13.8: NH3 TPD using DRIFTS over 20%Co-ZrO2 catalyst ..................... 220 Figure 13.9: NH3 TPD using DRIFTS over 20%Co-40%CeO2-40%ZrO2 catalyst ......................................................................................................................... 221

xxv

PART 1

HYDROGEN PRODUCTION VIA THE WATER-GAS SHIFT REACTION FROM COAL-DERIVED SYNTHESIS GAS

CHAPTER 1

INTRODUCTION TO WATER-GAS SHIFT FOR HYDROGEN PRODUCTION FROM COAL

Although renewable sources have emerged in recent years as promising alternatives for hydrogen production, decades of research and development work is still needed before technology commercialization. For the near term, fossil fuel continues to be the primary source for hydrogen production. It has been reported that around 95% of the hydrogen in the US and 50% of the worlds hydrogen supply is from steam reforming of natural gas[1]. Natural gas reforming is a wellestablished technology and remains the cheapest way for hydrogen production so far. From an optimistic estimation[2], the minable reserves of oil and natural gas on earth will be depleted in around 45 and 65 years, respectively. However, coal is an abundant fossil fuel in US and its reserves can last about 250 years[3]. Utilization of coal for power generation has been attracting more and more interest. In addition to carbon and hydrogen, there are small quantities of other elements present in coal, such as sulfur and nitrogen. Conventional coal combustion for power generation produces a variety of pollutants, including sulfur oxides, nitrogen oxides, carbon dioxide, volatile organic compounds, etc, which 2

contribute to global warming and have caused major environmental concerns. The combustion process is of low efficiency with a high level of heat waste. How to improve process efficiency and reduce pollutant emissions are two major issues to address in coal utilization. Different from conventional coal combustion, coal gasification can break coal down into smaller molecules under high temperature and pressure in the presence of steam or oxygen[4]. This process leads to production of synthesis gas, including CO, H2, CO2, CH4, H2O, H2S, COS, and nitrogen compounds. Synthesis gas clean-up is easier to achieve and after purification, synthesis gas can then be used to manufacture a great variety of industrially important chemicals. It can also be used to generate pure hydrogen to power fuel cells, which have been proven to be the most efficient devices to convert chemical energy into electricity. Another advantage of coal gasification is significant improvement in process efficiency in contrast to conventional combustion. The integrated gasification combined cycle (IGCC) has been demonstrated by US Department of Energy to be the most efficient and cleanest route to produce power, fuels and chemicals from coal. The water-gas shift (WGS) reaction is an integral step in synthesis gas conversion to reduce CO level and concentrate H2 before further H2 purification: CO + H2O CO2 + H2 (Hrxn,298K = -41.2 kJ/mol). Development of an advanced catalytic WGS system is essential for the success of producing hydrogen from coal.

The main objectives for the PhD studies are catalyst development for the WGS reaction and an understanding of catalyst surface and structural properties on catalytic performance. The existing technology for WGS reaction includes two stages, a high temperature shift (HTS) and a low temperature shift (LTS) stage. The reaction, being exothermic and reversible, is limited by thermodynamic equilibrium at high temperatures. The first stage is needed to accelerate the kinetics by operating at high temperatures, and the second stage is needed to increase the conversion by pushing back the thermodynamic equilibrium limit. The existing technology suffers from drawbacks such as toxicity of Cr in the HTS catalyst formulation, inactivity of the HTS catalysts at lower temperatures, and metal sintering, especially for the LTS catalysts, and low-sulfur tolerance and the need for high steam for both stages. In our previous WGS work[5], we have demonstrated that aluminum can be an excellent replacement for chromium in Fe-based high temperature WGS catalysts and Fe-Al-Cu catalysts exhibit a significant improvement in WGS activity. In this research, effect of catalysts synthesis techniques on catalyst performance was examined in Fe-Al-Cu system. In addition to conventional impregnation and coprecipitation methods, sol-gel technique was introduced to prepare Fe-Al-Cu catalyst. Activity evaluation of this catalytic system has shown superior performance compared with commercial catalysts. Additionally, sol-gel Fe-Al-Cu catalysts exhibited high activities over a wide temperature window, thereby offering the potential to bring the high-temperature shift and lowtemperature shift operations together to reduce operation cost. Extensive 4

characterization studies were conducted to correlate catalyst physical and chemical properties with their performance during the reaction. Another focus of the work is catalyst evaluation in the presence of H2S, a major impurity from coal derived synthesis gas. Efforts were made to examine catalyst surface and structural changes upon sulfur poisoning. Catalyst deactivation mechanism was proposed by combining catalyst performance after sulfur exposure and characterization results. These studies provide guidance for poisoned catalyst regeneration and sulfur tolerant catalyst design.

CHAPTER 2

LITERATURE REVIEW

The WGS reaction was first reported by a British patent in 1888[6]. It is moderately exothermic with thermodynamic limitation at high temperature. However, reaction rate in the gas phase is negligible at practical temperatures. A breakthrough discovery was made by Bosch and Wild in 1912. Mixtures of oxides of iron and chromium were found to catalyze the WGS reaction at 400-500oC and reduce CO concentration to around 2%. Thereafter, WGS has been widely used in industry, from the Haber-Bosch ammonia synthesis at the beginning of the 20th century to a variety of hydrogen production processes from synthesis gas. However, a considerable amount of CO still exists after a high temperature WGS. Another milestone in WGS catalyst development is the discovery of low temperature WGS catalysts[7]. A US patent in 1931 reported the invention of copper-based catalysts. This type of catalyst operates at low temperature (200-300oC) and can reduce CO concentration to less than 0.3%. This invention has made profound influence in many industrial processes where low CO and high H2 levels are desired. In industry, WGS is typically performed via two stages. The first high temperature WGS (HT-WGS) stage at 320-450oC 6

can bring CO concentration to 3-5% in kinetically-controlled region. Fe-Cr is the widely used catalyst for this process. To further reduce CO level to 0.3-1%, a second low temperature WGS (LT-WGS) stage (180oC-240oC) is needed[8]. Cu/ZnO/Al2O3 or precious metal-based catalysts are commonly used in this stage. However, both types are sensitive to sulfur contamination in the feed. About 20 years ago, a new class of sulfur resistant WGS catalysts, with Co-Mo or Ni-Mo sulfides as the formulation, was proposed for WGS reaction starting from sulfide species containing feed, with the process known as sour gas shift. WGS reactors require large volumes due to slow kinetics at low temperatures. It is estimated by Qi and his coworkers[9] that the volume of WGS catalysts, including both LT-WGS and HT-WGS, is around 6 times that of the reforming catalysts, which results in a large capital investment. Research on WGS catalyst has been undergoing for many decades. A large number of studies have been published on WGS catalysts, including HT-WGS, LT-WGS and more recently sour gas shift work. Catalyst formulations, preparation methods, reaction kinetics and mechanisms have been explored extensively.

2.1 Overview of high temperature water-gas shift catalysts

The active site for Fe-based HT-WGS catalysts is magnetite, which can be achieved by partial reduction of hematite with hydrogen or synthesis gas in the temperature range of 315oC to 460oC[10]. Over-reduction of hematite to FeO or metallic Fe during activation will drastically decrease catalytic activity. Sintering 7

of magnetite at high temperature is another issue[11]. Different promoters are introduced to prevent over-reduction of catalysts and minimize sintering of magnetite at high temperatures. Species that have an ionic radius similar to the iron atom are expected to go into the magnetite lattice, rather than to segregate as another phase[12]. Those species are of great interest for their potential ability to modify the magnetite structure. The classical HT-WGS catalyst formulation is Fe-Cr and it has been widely used in industry. A typical composition for Fe-Cr commercial catalyst contains ca. 92% iron oxide and ca. 8% chromium oxide[13]. The presence of Cr2O3 was reported as a structural promoter to prevent the rapid thermal sintering of iron oxides and loss of surface area at high temperature. To further improve Fe-Cr catalytic performance, researchers have examined other promoters including Hg, Ag, Ba, Pb, B, Rh, Ce Zn, Cu and Co[14-17]. It has been found that promoters can facilitate hematite reduction, reduce the apparent activation energy or expedite the oxidation-reduction reaction cycle. However, the commonly used Fe-Cr based catalysts contain a weight percentage of around 2% Cr6+[14]. Cr6+ is a highly toxic component and its health hazards remain as major concerns during catalyst manufacture and disposal. There have been efforts reported in the literature to replace Cr in the formulation while maintaining the activity, selectivity and stability of the catalyst for WGS. Rangel and coworkers have performed extensive studies on Cr-free Fe-based catalysts. Fe-ThCu was shown to be promising for industrial applications compared with Fe-CrCu catalyst[18]. Cu was validated to be a structural promoter whereas Th can stabilize iron oxides from sintering and further reduction. Additionally, 8

promotional effects of Al and Cu were examined[19-22]. Al was found to help prevent sintering, minimize surface area loss and suppress over-reduction of hematite. Fe-Al-Cu exhibited superior performance compared with commercial catalysts. Vanadium was another element incorporated into the iron oxide structure[23]. Vanadium improved catalytic activity both texturally and structurally. Liu et al. reported their work on Fe-Al-Ce catalysts for HT-WGS[14, 24]. Fe-Al-Ce was also shown to be an excellent Cr-free WGS catalyst, evidenced by high activity and thermal stability. More recently, Cu-CeO2 catalyst was evaluated in a membrane reactor with H2 separation[25]. Cu-CeO2 demonstrated excellent performance under CO2 rich conditions, which is unlike commercial Fe-Cr catalysts that deactivate dramatically under this feed condition.

Catalysts Fe-Cr promoted by Hg, Ag, Ba, Cu, Pb, B Fe-Cr-Cu Fe-Cr-Rh Fe-Cr-Ce Fe-Th-Cu Fe-Al Fe-Al-Cu Fe-V Fe-Al-Ce Cu-CeO2

Research highlights Promotional effect order Hg>Ag, Ba>Cu>Pb>unpromoted>B Reduced apparent activation energy by Cu Accelerated H2/H2O cycle by Rh Facilitated redox cycle by CeO2 Superior activity compared with Fe-Cr-Cu Comparable activity with Fe-Cr High activity and stability compared with Fe-Cr-Cu More active than Fe-Cr Acceptable activity and high thermostability Excellent performance in CO2 rich conditions

References [26] [11, 27] [15] [28] [29] [19, 20] [21, 22] [23] [14, 24] [25]

Table 2.1: Catalysts used in HT-WGS process

10

2.2 Overview of low temperature water-gas shift catalysts

The current commercial LT-WGS catalyst is Cu-Al-Zn. Metallic Cu is the active site for LT-WGS reaction, which will be discussed in the reaction mechanism section. Research work on LT-WGS has been largely focused on how to maximize Cu distribution on the support. This was achieved by varying preparation methods and support oxides, optimizing pretreatment conditions or adding other promoters. Commercial Cu-Zn-Al is usually prepared by a coprecipitation method. Takehira et al. introduced a new technique, homogeneous precipitation using urea hydrolysis, for preparation of Cu-Zn and Cu-Zn-Al catalysts[30]. A highly dispersed Cu on the surface was obtained, which was attributed to a strong interaction between Cu and Al2O3 support assisting the formation of small and uniformly distributed Cu particles on the surface. Additionally, ZnO played a role in stabilizing and distributing Cu crystallites on the support[31]. Various oxide supports have been examined for LT-WGS catalyst preparation including Al2O3, ZnO, CeO2 and TiO2. CeO2 supported catalyst is originally used as a three-way catalyst in automobile catalytic converters. CeO2 has been heavily studied in WGS system recent years because of its extremely high oxygen storage capacity. CeO2 supported catalysts have the potential to surpass Cu-ZnO in catalytic performance. Rnning et al. doped Ce into Cu-Al-Zn catalysts and Ce was found to act as both an active component and a promoter to improve the reducibility and modify oxidation-reduction during WGS[32]. 11

Supported noble metal catalysts are also being examined to obtain highly concentrated H2 for fuel cells. Jacobs and coworkers carried out a screening of promoters for metal/ceria systems[33] including Pt, Ni, Co and Fe. Pt/CeO2 exhibited the highest extent of reduction and was an order of magnitude more active than other promoters from kinetic studies. Gorte et al. observed a relationship between extent of reduction and catalytic performance through their kinetic measurements of activation energies on CeO2-supported Pd, Pt and Rh catalysts[34, 35]. Song and coworkers[36] investigated Cu-Pd bimetallic catalysts supported on nano-structured CeO2 for oxygen-assisted WGS to achieve CO levels suitable for PEM fuel cell feedstock. The urea gelation method yields CeO2 with high surface areas and nano structures to support the active metals. Supported Au on CeO2 catalyst was found capable of providing high concentrated H2 for fuel cells through WGS[37]. However, stability during WGS in H2-rich reformate gas is an issue with Au catalytic system. Zhang and coworkers doped ZrO2 and Nb2O5 promoters to Au/Fe2O3 catalysts, which enhances Au dispersion and catalytic stability[38]. TiO2 is another alternative support. Pd/TiO2 and Ir/TiO2 catalysts were evaluated in WGS system and showed promising WGS activities[39]. Kondarides and coworkers examined Pt, Rh, Ru and Pd supported on TiO2 and found that the WGS activity was strongly affected by both the nature of metal oxide carrier and the active metal. Crystallite size of TiO2 can influence interaction between metal and the support and creation of oxygen vacancies, which are believed beneficial for the reaction[40, 41]. 12

Another WGS catalytic system intended to be used for H2 generation for PEM fuel cell is Cu-Mn with spinel structure[42, 43]. When Mn in the structure is partially replaced by Fe or Al, activity level can be pushed higher. Co-Mo carbide system was studied by Nagai et al.[44]. This type of catalyst showed high activity at the beginning of the reaction, but it suffered deactivation after even 5 hours time on stream probably caused by carbide phase transformation by the reactants.

13

Catalysts Cu-Zn-Al Cu-Zn-Al-Ce CeO2 supported Pt, Pd, Ni, Co and Fe Cu-Pd/CeO2 Au/CeO2 ZrO2 or Nb2O5 promoted Au/CeO2 Re-Pd/TiO2 and Re-Ir/TiO2 Rh, Ru, Pd and Pt supported on TiO2 Cu-Mn spinel Co-Mo carbide

Research highlights Homogeneous precipitation leads to highly dispersed Cu on the support Ce acts as both an active component and a promoter Pt/CeO2 and Pd/CeO2, Ni/CeO2 formulations are highly active for WGS; Cu distribution varies depending on CeO2 support reducibility High surface area CeO2 was synthesized and bimetallic formulation provided high activity for oxygen-assisted WGS This catalyst yields high concentrated H2 directly for fuel cell but with low stability Promoters significantly enhanced the activity and stability Re- Ir/TiO2 is highly active for LT-WGS A combination of Pt/TiO2 helped to create oxygen vacancies and facilitate the reaction It can be used for H2 generation for PEM fuel cell It is highly active at the beginning but deactivates after 5 hours

References [30] [32] [33-35]

[36] [37] [38] [39] [40, 41] [42, 43] [44]

Table 2.2: Catalysts used in LT-WGS process

14

2.3 Overview of sour gas shift catalysts

Sour gas shift is used to convert raw gases from coal or crude oil gasification that contain sulfur and traces of hydrocarbons. This process operates around 350oC. Catalysts are typically based on fully sulfided molybdenum, which are reported to be the active sites for WGS. Therefore, a certain amount of sulfur species is needed in the stream to maintain the active sites and this type of catalysts are completely insensitive to sulfur. Edreva-Kardjieva and coworkers studied the WGS reaction in Ni-Mo-Al system[45]. MoS2 is the active form for reaction. Presence of potassium can modify the selectivity and activity. The same group examined the effect of potassium introduction order on Mo-Al catalytic performance[46]. Deposition of potassium results in marked differences in activity by modifying the textural properties including surface area, pore volume and distribution of MoOx species.

2.4 Overview of water-gas shift reaction mechanism

Reaction mechanism is an important part of catalysis research. Important aspects include detailed understanding of surface and bulk structural characteristics and elementary steps involved in reaction network. These studies help elucidating reaction phenomena observed during testing and provide guidance for catalyst design. Rhodes and his coworkers in 1995 made a comprehensive review of the WGS reaction mechanism[13]. For Cu-Zn-Al 15

system, it is believed that metallic Cu is the active phase during reaction. In activated catalysts, islands of metallic Cu exist on zinc oxide alumina. However, there have been difficulties in detecting metallic Cu by characterization techniques because metallic Cu is extremely sensitive to oxidizing environment. Even under atmospheric conditions, metallic Cu can be easily oxidized to CuI and CuII. Reaction mechanism occurring in Cu-Zn-Al system has been a debated area for several decades. Two distinct mechanistic pathways were proposed: an associative mechanism and a regenerative mechanism. Both mechanisms were originally proposed by Armstrong and Hilditch in 1920[47]. For the associative pathway, CO and H2O are first adsorbed onto the surface and an intermediate structure is formed involving CO, H2O and catalyst surface. Decomposition of the intermediate species produces CO2 and H2. The regenerative pathway can be described by the following two steps.
H2 O + Red H2 + Ox CO + Ox CO 2 + Red

This can be explained by a cyclic change in catalyst oxidation states. Catalyst surface can be oxidized by H2O to produce H2 and reduced back to the original state by CO to produce CO2. In Armstrong and Hilditchs mechanism studies, they stated that WGS proceeds on Cu-based catalysts through the associative mechanism. The intermediate species was suggested to be a formic acid structure. Researchers attempted to confirm this assumption. By using chemical trapping experiments, infrared spectroscopy, kinetic calculations and reaction data, they were able 16

testify the presence of formate species on the surface during WGS reaction. Recently researchers proposed the regenerative pathways on Cu-based catalysts. However, evidence is needed to ascertain that oxidation of Cu by H2O during the reaction would be rapid enough to account for the observed overall WGS reaction rate. On the other hand, there is much less debate regarding the regenerative mechanism in Fe-based catalysts. The active phase is Fe3O4. Fe3O4 has both Fe2+ and Fe3+ oxidation states. Half of Fe3+ is in tetrahedral site. The other half of Fe3+, together with Fe2+ in a ratio 1:1, occupy the octahedral site. Mssbauer studies have shown rapid electron hopping between Fe2+ and Fe3+ in the octahedral sites. Electron hopping between Fe2+ and Fe3+ facilitates oxidation by H2O and reduction by CO during WGS reaction. Hu et al. proposed a slightly different mechanism for Fe-based catalysts during their studies on Fe-Ce-Cr catalysts[28]. They attributed the WGS reaction to a two-step process. For the first step, reduced catalytic surface splits the HO-H bond, forming adsorbed OH group on the surface and H2 in the gas phase from two neighboring H atoms. For the second step, CO interacts with OH groups to form a formate species and in the presence of H2O, the formate species can decompose into CO2 and H2. The above mentioned mechanism was experimentally demonstrated by Shido and Iwasawa[48-50] for the first time in 1991 when they studied WGS reaction on ZnO surface. It was named reactant-promoted reaction mechanism. Shido and Iwasawa also examined CeO2 and Rh-CeO2 systems for WGS. Ce-O pair is suggested to be the active site. Terminal OH groups adsorbed on the 17

reduced Ce surface can react with CO and form bidentate formates. Bidentate formate could decompose to H2 and unidentate carbonate. The decomposition rate can be accelerated by H2O molecules in the feed. H2O also facilitates decomposition of carbonates to form CO2. This was validated through transient isotope experiments[51]. Panagiotopoulou et al.[41] and Naito et al.[52] proposed similar pathways in TiO2 supported catalytic system. In contrast to Shido and Iwasawas theory, Gorte et al. proposed a simple redox mechanism as shown below in CeO2 supported Pt, Pd and Rh system, which were consistent with their kinetic measurements[34, 35, 53].
CO + COad H2O + Ce2O3 2CeO2 + H2 COad + 2CeO2 CO2 + Ce2O3 +

Here CeO2 can oxidize CO adsorbed on the metal surface and Ce2O3 can be oxidized by H2O to produce H2 and recover the Ce surface as well.

18

CHAPTER 3

EXPERIMENTAL METHODS

3.1 Catalyst preparation

Based on previous studies on WGS project from this group, Fe-Al-Cu is a highly promising formulation for the WGS reaction. Fe-Al-Cu catalysts were prepared by employing different methods, including precipitation-impregnation, co-precipitation and sol-gel.

3.1.1 Precipitation-impregnation

Metal precursors used were purchased from Aldrich and were in nitrate form. Ammonium hydroxide (29.63 volume%, Fisher Scientific) was used as the precipitating agent. During precipitation-impregnation preparation, desired amount of aluminum and iron nitrates were measured and 0.5 mol/L solution for each nitrate was made. The two solutions were mixed and pH was adjusted using ammonium hydroxide. After the desired pH value was reached, the dark brown solution was stirred vigorously for an additional 30 min. The precipitate 19

was washed using demineralized distilled water with a filtration system. This step is the co-precipitation step to form Fe-Al precipitates. To introduce Cu, the Fe-Al filter cake was impregnated with a calculated amount of 0.5 mol/L copper nitrate aqueous solution under stirring. After this, the sample was dried in over at 110oC overnight. Finally, catalysts were calcined in air at 450oC for 4 hrs at a ramping rate of 5oC/min. This preparation process is named co-precipitationimpregnation, which is 2-step for short.

3.1.2 Co-precipitation

Fe, Al and Cu nitrate precursors from Aldrich were used for this method. All precursors, in 0.5 mol/L aqueous solutions, were mixed and co-precipitated in one pot. Sodium carbonate (2mol/L) was used as the precipitating agent instead of ammonium hydroxide because ammonium could form a complex with Cu at high pH values. The precipitate was centrifuged and washed with demineralized distilled water. The washing step was repeated many times to completely remove sodium ions remaining in the filter cake. The washed filter cake was dried overnight in an oven at 110oC. The dried sample was ground into fine powder and calcined in air at 450oC for 4 hrs at a ramping rate of 5oC/min. This preparation process is named co-precipitation, which is 1-step for short.

20

3.1.3 Sol-gel

In sol-gel preparation, in addition to Al and Cu nitrate precursors (Aldrich), organic iron precursor, iron (III) acetylacetonate (C5H8O2)3Fe (Aldrich), was used. C2H5OH and NaOH were used as the solvent and precipitating agent, respectively. Initially, iron (III) acetylacetonate was dissolved in ethanol at a certain temperature (0.2 mol/L). Aqueous solutions of aluminum (0.02 mol/L) and copper nitrates (0.01 mol/L) were added dropwise into the iron (III) acetylacetonate-ethanol mixture. The pH of the resulting solution was measured and adjusted by adding NaOH (0.5 mol/L) dropwise. The precipitate was vigorously stirred for 30 min, then thoroughly rinsed with demineralized distilled water to remove sodium ions. The remaining solid was then dried overnight in an oven at 110oC. The dry samples were ground to a fine powder and were

calcined under air at 450oC (ramp rate = 5oC/min) for 4 hrs.

3.2 Catalyst characterization

A variety of characterization techniques were employed in this thesis work to examine catalyst surface and structural properties of the catalysts at different stages, which can be used to correlate with catalytic performance and guide catalyst design. The follows are techniques used for WGS work.

21

3.2.1 N2 physisorption analysis

During N2 physisorption analysis, adsorption isotherms are collected for the amount adsorbed by sample material at different relative pressures at 77 K. The Brunauer-Emmett-Teller (BET) method is used to evaluate the surface area from physisorption isotherm data. In this work, surface areas of catalysts were measured with a Micromeritics ASAP 2010 instrument. Samples were first degassed under vacuum at 130oC overnight. Pore volumes and pore size distribution were obtained using the Barret-Joiner-Halenda (BJH) adsorption curve.

3.2.2 Temperature programmed reduction

Temperature programmed reduction (TPR) was performed using a laboratory-made gas flow system. During the reduction, the effluent from the reactor, after a drying step through a silica gel water trap, was sent to a TCD (Thermal conductivity detector) connected to a data-acquisition computer to monitor H2 consumption.

3.2.3 Temperature programmed reaction

Temperature programmed reaction is a useful technique to study catalyst dynamic reaction properties. It is conducted using a laboratory-made gas flow 22

system. The effluent species can be analyzed by a mass spectrometer. An Infrared spectrometer can also be used to probe the interaction between gas phase and catalyst surface during temperature programmed reaction process.

3.2.4 X-ray photoelectron spectroscopy

X-ray photoelectron spectroscopy (XPS) was employed to determine surface composition and oxidation states. An AXIS Ultra XPS spectrometer was used. It is operated at 13 kV and 10 mA with monochromator Al K radiation (1486.6 eV). For samples that are not air sensitive, carbon tape was used to load the powder sample. If samples are sensitive to air, an Ar-purged glove box and sample transfer cart were used. During the analysis, a survey spectrum was first collected from 1400 eV to 0 eV. This was followed by region scans where desired element regions can be scanned in details.

3.2.5 Diffuse reflectance infrared Fourier transform spectroscopy

Infrared spectroscopy is a widely used technique in catalysis to probe surface properties, gas-solid interactions and reaction intermediates. Weakly reflected power samples are typically analyzed by diffuse reflectance infrared Fourier transform spectroscopy A Thermo 6700 FT-IR spectrometer equipped with a diffuse reflectance Fourier transform infrared spectroscopy cell, an MCT detector and a KBr beam splitter was used. DRIFT spectra were collected with a 23

500-scan data acquisition at a resolution of 4 cm-1 in a controlled gas atmosphere and temperature using an environmental chamber with Zn-Se windows.

3.2.6 X-ray diffraction

X-ray diffraction (XRD) was used for the detection and identification of bulk crystalline phases within the catalyst. Bruker D8 advance diffractometer was used for data collection. A 9-sample holder accessory was employed to collect diffraction patterns. The X-ray source was Cu K radiation operated at 45 kV and 20 mA. The International Center for Diffraction Data (ICDD) library was used for crystal phase identification. With controlled atmosphere and temperature capability, structural changes during catalyst calcinations, oxidation or reduction can be examined. In situ XRD patterns were obtained using a Scintag XDS diffractometer equipped with an HTK-1200 oven, temperature control and gas flow capability.

3.2.7 Mssbauer spectroscopy

Mssbauer spectroscopy is a unique technique for detection of iron elements in Fe-based WGS catalysts. Information obtained from this technique includes identification of phases, determination of oxidation states, structure, particle size and kinetics of bulk transformations. Gamma rays are produced from 24

a radioactive

57

Co source. The source is vibrated to create a range of energies


57

via the Doppler effect.

Co/Rh -ray source and a conventional constant

acceleration Mssbauer spectrometer were used during spectra collection. The integrated areas under each deconvoluted peak have been used to obtain the relative populations of the different iron species. An equal free recoil fraction for all species was assumed for all calculations.

3.3 Reaction testing

Reaction testing is a very important part in catalyst development. Catalysts were evaluated under steady state reaction testing with or without the presence of H2S, which is a major impurity present in coal derived synthesis gas Reaction experiments were performed using a fixed-bed flow system with a stainless steel reactor (1/4 in. OD). The feed and products were analyzed online using an automated Shimadzu GC-14A equipped with FID (Flame ionization detector) and TCD (thermal conductivity detector) detectors. Separations were performed under Ar using 2 columns: Porapak Q (12 ft x 1/8 in. SS, 80/100 mesh) and molecular sieve 13X (5 ft x 1/8 in. SS, 60/80 mesh). Porapak Q was used for separation of CO2 and H2O. Molecular sieve column was used for separation of N2, H2 and CO. Table 3.1 lists the GC oven temperature program and Table 3.2 is the GC valve-switching program during an analysis run.

25

Time (min) 0 3.5 7.5 33

Temperature (oC) 80 80 120 120

Program Initial setting Hold at 80oC for 3.5 min Ramping at 10oC/min starting from 3.5 min o Hold at 120 C for 25.5 min with the total program time 33 min

Table 3.1: GC oven temperature program

Time (min.) 0.0 0.5 3.5 14.5

Valve Number 1 1 2 2

Valve Position Inject Load Load Inject

Table 3.2: GC valve-switching program.

3.3.1 Steady state reaction

Catalysts were first reduced in situ with 10% CO, 10% H2O, 7.5% H2, 5% CO2 balanced with 67.5% N2 at 350oC for 2 hrs. Reactions were then performed at desired temperatures under the same feed stream. When catalysts were evaluated at different temperatures, low temperatures were performed first before moving to the next higher temperature. Catalysts were held at each temperature for 10 hrs to ensure steady state had been reached.

26

3.3.2 Catalyst evaluation in the presence of H2S

For catalyst evaluation in the presence of H2S, catalysts were first reduced at 350oC for 2 hrs in a clean simulated coal gas mixture with 10% H2O, 10% CO, 5% CO2, 7.5% H2 and 67.5% N2. Fresh catalyst activity was measured with the same feed at 400oC. To achieve steady state catalytic performance, catalysts were kept on stream for at least 5 hrs. After this, catalysts were treated with 50 ppm H2S/N2 at 400oC for a certain amount of time. Reaction activity was again collected with the same clean coal gas feed as before sulfur exposure.

27

CHAPTER 4

INVESTIGATION OF HIGHLY ACTIVE FE-AL-CU CATALYSTS FOR WATERGAS SHIFT REACTION

4.1 Overview of Fe-based catalysts for water-gas shift reaction

There has been extensive research on HT-WGS catalysts. Cr2O3 in the commercial catalyst formulation was reported as an efficient promoter to prevent rapid thermal sintering of iron oxides and loss of surface area at high temperatures. However, Cr6+ present in the commonly used Fe-Cr based catalysts is a highly toxic component and its health hazards remain as major concerns during catalyst manufacture and disposal. Efforts to replace Cr in the formulation while maintaining the activity, selectivity and stability of the catalyst for WGS has been reviewed in chapter 2. In our earlier work on Fe-based WGS[5], we investigated potential textural promoters (Al, Ga, Mn) for Cr replacement and structural promoters (Co, Zn, Cu) for further improvement of the WGS activity. Continued efforts have been focused on Fe-Al-Cu because of its high activity and stability seen in reaction evaluations. The preparation method was found to play a key role in determining 28

the catalytic performance. In that study, two methods were used for Fe-Al-Cu catalyst preparation. The 2-step synthesis of Fe-Al-Cu involved a consecutive coprecipitation-impregnation process, in which Cu was impregnated onto the FeAl surface formed by coprecipitation. A surface enrichment of Cu was observed from characterization results (Temperature programmed reduction, X-ray diffraction and X-ray photoelectron spectroscopy), which may lead to sintering of metallic Cu at high temperatures and thus loss of the WGS activity. The 1-step synthesis of Fe-Al-Cu involved coprecipitating Fe, Al and Cu together in one batch. In 1-step Fe-Al-Cu catalysts, although there is still metallic Cu on the surface, it is conceivable that a portion of Cu species are incorporated into the iron oxide structure and form a solid solution. Therefore, sintering of Cu at high temperatures is not severe and this contributes to a significant boost in WGS activity under harsh conditions. Similar phenomena were reported in Fe-Cr system by Kundu and coworkers[54]. In their work, the interaction between Fe2O3 and Cr2O3 was found to be insignificant in the wet mixing catalyst preparation whereas significant interaction took place for the co-precipitation method, through which Cr3+ ions were incorporated into the octahedral vacant sites of the -Fe2O3 lattice and dramatically improved the activity. This chapter presents our work on sol-gel Fe-Al-Cu catalyst preparation. During sol-gel preparation, an organic iron precursor was used to form a gel together with Al and Cu, yielding a highly uniform distribution of promoters in the iron oxide matrix. Catalysts were evaluated under steady state reaction conditions with a fixed-bed reactor. Various characterization techniques including 29

temperature programmed reduction (TPR), X-ray photoelectron spectroscopy (XPS), X-ray diffraction (XRD) and temperature programmed reaction (TPRxn) were used to examine surface and structural properties of Fe-Al-Cu catalysts prepared by different methods and correlate these characteristics with their WGS performance.

4.2 Experimental procedures

4.2.1 Catalyst preparation

All chemicals used during catalyst preparation were purchased from Sigma-Aldrich. Fe-Al-Cu catalysts were synthesized by three different techniques: coprecipitation-impregnation (2-step), coprecipitation (1-step) and a sol-gel method. 2-step and 1-step preparations have been described in a previous paper[5]. In sol-gel preparation, catalyst precursors were iron (III)

acetylacetonate (C5H8O2)3Fe, aluminum nitrate (Al(NO3)3.9H2O) and copper nitrate (Cu(NO3)2.3H2O). C2H5OH and NaOH were used as the solvent and precipitating agent, respectively. The following synthesis variables were controlled during the synthesis: Fe/Al = 10, Fe/Cu = 20, pH = 9. The conditions were also maintained for Fe-Al-Cu catalysts prepared by the other two methods. Initially, iron (III) acetylacetonate was dissolved in ethanol at 50oC (0.2 mol/L). Aqueous solutions of aluminum (0.02 mol/L) and copper nitrates (0.01 mol/L) were added dropwise into the iron (III) acetylacetonate-ethanol mixture. The pH 30

of the resulting solution was measured and adjusted by adding NaOH (0.5 mol/L) dropwise. The precipitate was vigorously stirred for 30 min at 50oC, then centrifuged and thoroughly rinsed with demineralized distilled water (1200 rpm and 5 min for each washing, 10 times) to remove sodium ions. The remaining solid was then dried overnight in an oven at 110oC. The dry samples were ground to a fine powder and were calcined under air at 450oC (ramp rate = 5oC/min) for 4 hrs.

4.2.2 Reaction studies

All reactions were performed with a WHSV (Weight Hourly Space Velocity) =0.06 m3(g cata)-1h-1. Catalysts were first reduced in situ with 10% CO, 10% H2O, 7.5% H2, 5% CO2 balanced with 67.5% N2 at 350oC for 2 hrs. Reactions were then performed at desired temperatures under the same feed stream. For reaction testing over the temperature range from 250oC to 400oC, reactions were conducted from low to high temperatures and catalysts were held at each temperature for 10 hrs to reach steady state.

4.2.3 Catalyst characterization

Surface areas of calcined catalysts were measured with a Micromeritics ASAP 2010 instrument. Samples were first degassed under vacuum at 130oC overnight. By using the adsorption-desorption isotherm at 77K, surface areas 31

were calculated by the Brunauer-Emmett-Teller (BET) method. Pore volumes were obtained using the Barret-Joiner-Halenda (BJH) adsorption curve. Temperature programmed reduction (TPR) was performed using a laboratory-made gas flow system described in detail elsewhere[55]. For all experiments, 100 mg of catalyst were first sandwiched between quartz wools in a 1/4 in. O.D. quartz U-tube reactor. The catalyst was then treated under N2 at 450C for 1 hr. After cooling to room temperature under N2, using 10% H2/N2 (30 cm3(STP)/min) as a reducing agent, the temperature of the catalyst was raised at a ramp rate of 10C/min to 800C and held for 10 min. During the TPR process, the effluent from the reactor, after a drying step through a silica gel water trap, was sent to a TCD connected to a data-acquisition computer to monitor H2 consumption. X-ray photoelectron spectroscopy (XPS) was performed using an AXIS Ultra XPS spectrometer, operated at 13 kV and 10 mA with monochromator Al K radiation (1486.6 eV). For reduced samples, the same reducing agent as specified in the reaction section was used to perform the reduction at 350oC for 2 hrs. At the end of 2 hrs, inert gas (N2) was used to purge and cool down the reactor. The sample was sealed in N2 before transferring to an Ar-purged glove box. For all samples, they were transferred to XPS analysis chamber without exposure to air. Samples were loaded onto double-sided carbon tape. Charge neutralization was used to reduce the effect of charge built-up on samples. All binding energies were referenced to C 1s of 284.5 eV.

32

Calcined powder catalysts X-ray diffraction (XRD) patterns were obtained using a Bruker D8 X-ray diffractometer equipped with a 9-sample holder. The Xray source was Cu K radiation. 2 diffraction angle was varied from 20o to 90o during the measurement. The International Center for Diffraction Data (ICDD) library was used for phase identification.
57

Co/Rh -ray source and a conventional constant acceleration Mssbauer

spectrometer were used during Mssbauer spectra collection. Isomer shifts are given with respect to -Fe. All spectra were taken at room temperature and ambient atmosphere. The integrated areas under each deconvoluted peak have been used to obtain the relative populations of the different iron species. An equal recoil free fraction was assumed for all iron species. Temperature programmed reaction (TPRxn) experiments were conducted to study redox features of Fe-Al-Cu catalysts. First, catalysts (50 mg) were

loaded into a inch. O.D. quartz reactor and were pretreated under 30 cm3 (STP)/min of He at 450oC for 30 min with a ramp rate of 10oC/min. Subsequently, catalysts were reduced in situ with 10% H2/He for 2 hrs at 350oC followed by cooling to room temperature under He. H2O-TPRxn experiments were initiated by raising the catalyst temperature to 400oC with a ramp rate of 5oC/min under 2% H2O/He flow (32 cm3 (STP)/min). Product gases eluting from the reactor were analyzed with a Thermo Finnigan TRACE DSQ Mass Spectrometer (MS).

33

4.3 Results and discussion

4.3.1 Reaction data for Fe-Al-Cu catalysts prepared by different methods

In this chapter, a sol-gel technique was employed to introduce Cu into the iron oxide matrix. Sol-gel synthesis offers many advantages over traditional preparation methods, including homogeneity, greater control of surface and microstructural properties, improved thermal stability of the supported metals and ease of incorporating additional elements, as reviewed by Gonzalez et al.[56]. As presented in Figure 4.1, reaction rate (expressed as CO conversion rates on equal catalyst weight basis) comparison for Fe-Al-Cu catalysts prepared by different methods is plotted against time-on-stream. All catalysts were first activated with a reducing agent described in the experimental section. Steady state reactions were conducted at 400oC. Catalyst performance was stabilized after 15 hrs on stream. WGS catalytic activity followed this trend: 2-step < 1-step < sol-gel. Sol-gel prepared Fe-Al-Cu exhibited significantly higher WGS catalytic activity compared with the other two methods. The inset in Figure 4.1 illustrates CO conversion rates normalized by surface area for those three Fe-Al-Cu catalysts at different reaction temperatures. Physical surface area discussions will be presented in the following section. It is evident that sol-gel Fe-Al-Cu demonstrated higher activity compared with the other two catalysts over the whole temperature range with the difference being more pronounced at higher reaction temperatures. 34

As reported in our previous work[5], how Cu was introduced into the iron oxide structure can result in dramatic differences in reaction activity even with the same composition. For 2-step Fe-Al-Cu, there was clearly segregation of metallic Cu on the surface after reduction. The enriched metallic Cu on the surface is believed to provide active sites for the reaction. Especially at low reaction temperature, Cu provided active sites in a similar way as in low temperature shift catalysts. This conclusion was also reported by Andreev et al.[17] and Idakiev et al.[57] where they studied influences of the addition of CuO on Fe-Cr WGS activity and ascribed the promotional effect to Cu providing new active sites and reacting in a similar manner as metallic Cu in Cu/ZnO LT-WGS catalysts. However, enriched Cu species on the surface tend to crystallize and sinter at high temperatures. This diminishes high temperature performance of the catalyst sharply, which can be evidenced by the reaction result comparison between 2step and 1-step Fe-Al-Cu catalysts in our work. Flytzani-Stephanopoulos et al.[25] also observed rapid deactivation caused by Cu sintering at 450oC in Cu-CeO2 catalytic system during the high temperature WGS process. For 1-step Fe-Al-Cu catalyst[5], we speculated that Cu can be more uniformly distributed in the iron oxide structure compared with Cu impregnation 2-step method. Although a major portion of Cu is still in the metallic phase after reduction, it is not segregated as large particles on the surface. Instead, much smaller Cu particles are spaced by iron oxides, which could retard sintering of Cu at high temperatures; whereas Cu in the iron oxide structure can modify electron transfer properties, which can be supported by Hutchings et al.s work[11, 27]. 35

The Fe-Cr-Cu catalysts prepared using co-precipitation method were reported to exhibit a significant enhancement in WGS activity compared with Fe-Cr. CuO and Cr2O3 were identified to form a solid solution within the Fe3O4 phase and no metallic Cu was observed on the surface for freshly activated catalysts. The promotional effect of Cu was attributed to Cu2+ forming a solid solution with iron oxides and modifying the electronic properties of the Fe3O4 structure during WGS sequences. Similarly in our earlier work, a higher reaction activity was observed for 1-step Fe-Al-Cu catalysts, particularly at high reaction temperatures where surface Cu species suffer from sintering and electronic modification by Cu comes into play. However, the functionality we are attributing to Cu is distinct from other researchers because in our work, Cu is believed to exhibit dual contributions by providing both active sites (surface Cu) and electronic enhancement (structural Cu). These two features promote the catalytic performance over a wide temperature range in tandem although a portion of Cu on the surface experiences gradual crystallization and sintering at higher temperatures. More structurally incorporated Cu species is beneficial for a stable WGS catalyst, particularly for high-temperature operation. It facilitates the electron transfer between Fe2+ and Fe3+ in octahedral sites in magnetite[58, 59]. Figure 4.2 shows a comparison of the same sol-gel prepared Fe-Al-Cu catalyst with a commercial Fe-Cr-Cu catalyst. Both catalysts were reduced at 350oC for 2 hrs. Steady state reaction data at different temperatures were collected. As seen in the figure, Fe-Al-Cu-SG catalyst yielded much higher activity than the commercial catalyst at every temperature tested. The fact that 36

this superior activity was maintained throughout the entire temperature range indicates that this preparation technique widens the operating temperature range, showing promise to combine high temperature and low temperature water-gas shift stages together. It should be pointed out that reaction conditions were strictly controlled and were maintained far away from equilibrium, and were therefore in the kinetically-controlled region where we could measure real differences between various catalysts. To examine the long-term stability of these catalysts, the sol-gel catalyst was kept on-stream for three days with no sign of deactivation. To elucidate the catalytic differences observed for these FeAl-Cu catalysts, characterization experiments were conducted using various techniques to reveal the physical and chemical properties created by preparation methods.

37

Figure 4.1: Reaction rate comparison of Fe-Al-Cu catalysts prepared by different methods: the inset is CO conversion rate per physical surface area per minute for Fe-Al-Cu catalysts

38

Figure 4.2: CO conversion between sol-gel Fe-Al-Cu catalyst and commercial Fe-Cr-Cu catalyst

39

4.3.2 Physisorption isotherm and pore size distribution measurement

Surface area and pore volume of calcined Fe-Al-Cu catalysts obtained from N2 physisorption measurement at 77 K are shown in Table 4.1. The 2-step Fe-Al-Cu catalyst showed a slightly higher BET surface area (smaller pore volume) compared with 1-step Fe-Al-Cu. However, sol-gel Fe-Al-Cu displayed a larger pore volume and a much smaller BET surface area. To further illustrate physisorption differences between Fe-Al-Cu catalysts, the adsorption-desorption isotherms and pore size distribution derived from BJH adsorption are plotted in Figure 4.3. Hysteresis loops are observed on all samples, implying a type IV gas adsorption isotherm as classified by IUPAC (international union of pure and applied chemistry)[60]. According to the IUPAC classifications, 2-step and 1-step Fe-Al-Cu exhibited H2 type hysteresis loops whereas the sol-gel Fe-Al-Cu showed an H3 type. The loop type has been correlated with specific pore structures[60]. Type H2 hysteresis loop initially was attributed to the presence of pores with narrow necks and wide bodies. However, this assignment was considered to be oversimplified recently and in a review by Jaroniec and coworkers[61], H2 hysteresis loops were considered to be characteristic of relatively uniform channel-like pores. Isotherms with type H2 loops tend to level off when the pressure is close to the saturation pressure. However, type H3 loops do not display this limit at higher relative pressures and have been associated with aggregates of plate-like particles with formation of slit-shaped pores[60, 61]. This explanation was questioned by experimental results[61]. The 40

H3 type loop has recently been attributed to large mesopores or macropores surrounded by a matrix of much smaller pores[62-65]. The presence of large pores could be caused by collapse of layered structures[61] or pore defects formed during material synthesis[65]. Pore size distributions from BJH adsorption branches are presented in Figure 4.3 (b). The dV/dlogD values in the y axis are calculated from BJH adsorption branch. V is the adsorption volume and D is the pore diameter. Both 2-step and 1-step catalysts demonstrated narrow uniform mesopore distribution. The sol-gel Fe-Al-Cu sample was featured by a broader pore size distribution, ranging from mesopores to macropores.

Catalyst 2-step 1-step Sol-gel

BET surface area (m2/g) 88 80 57

BJH adsorption pore volume (cm3/g) 0.19 0.23 0.37

Table 4.1: Surface area and pore volume of Fe-Al-Cu catalysts

41

(a)

(b) Figure 4.3: (a) Adsorption-desorption isotherms of N2 at 77 K for Fe-Al-Cu catalysts. Solid symbols: adsorption isotherms; Open symbols: desorption isotherms. 1-step isotherms are shifted vertically by 100; P0 is the saturation vapor pressure of N2 at 77 K. (b) Pore size distribution for Fe-Al-Cu catalysts obtained from BJH adsorption 42

The observed isotherms and pore size distributions are closely dependent on catalyst synthesis. The co-precipitation method yields small and uniform particles. Because 2-step catalyst is made from impregnation of Cu onto the surface of Fe-Al, the pore size distribution is largely determined by Fe-Al catalyst structure, which is prepared using the co-precipitation method as well. This explains similarities in isotherm and pore size distribution between 2-step and 1step Fe-Al-Cu catalysts. Sol-gel Fe-Al-Cu synthesis starts with an organic iron precursor. The pore volume and surface area is closely related with the extent of precursor hydrolysis[56]. Sufficient hydrolysis yields smaller pore volume and higher surface area. The large pore volume and low surface area from this solgel catalyst suggests an inadequate hydrolysis. It is postulated that in the hydrolyzed network formed during sol-gelling process, some of the hydrophobic tails from the organic precursor stay intact instead of entering the aqueous phase and being hydrolyzed. When calcined in air, the organic species would be burned off and the occupied space are released, which possibly explains the large pore volume. On the other hand, incomplete gelling process may lead to the defects in iron oxide structures, as suggested from subsequent characterization results. In addition, during catalyst calcination at high temperatures, a portion of small pores could collapse and lead to formation of large disordered pores, which provides another explanation for our experimental results.

43

4.3.3 Temperature programmed reduction of Fe-Al-Cu catalysts

TPR was performed on Fe-Al-Cu catalysts to examine the reducibility. Although with the same formulation as prepared (Fe/Al=10, Fe/Cu=20), variation in synthesis methods led to distinct reducibility behavior (Figure 4.4). The 2-step Fe-Al-Cu catalyst, which was reported in a previous publication[5], showed three main reduction features with a strong and well-resolved reduction peak from Cu species around 220oC. The reduction feature at 300oC was assigned to the partial reduction of Fe2O3 to Fe3O4. With further increase in the reduction temperature, Fe3O4 was progressively reduced to metallic Fe, without exhibiting distinct maxima. There is no temperature maximum that can be assigned to the formation of the intermediate FeO phase during reduction and this observation is consistent with the fact that FeO is thermodynamically unstable with a tendency to disproportionate to Fe3O4 and Fe below 570oC[15]. For 1-step Fe-Al-Cu catalyst, a broad peak was observed in the low temperature region. This peak was a combination of three reduction features. The main peak resulting from CuO reduction was displaced to 260oC with a weak shoulder remaining at 220oC, which may be due to the easily reduced CuO species on catalyst surface. The temperature shift for the main CuO reduction peak could be a result of better dispersion of CuO in the iron oxide matrix in the 1-step preparation method. The reduction of Fe2O3 to Fe3O4 appeared to have shifted from 300oC to 290oC. The peak from further reduction of magnetite changed little compared to unpromoted Fe-Al catalyst. When sol-gel method was 44

Figure 4.4: TPR profiles for Fe-based catalysts

45

used to synthesize Fe-Al-Cu catalysts, only one reduction peak (around 290oC) was observed at lower temperature, which can be ascribed to a uniform distribution of Cu in the iron oxide structure that all species were reduced in the same temperature range. In the TPR study conducted by Rhodes and his coworkers[11], no CuO reduction peak was observed for Fe-Cr-Cu catalysts because Cu2+ was found to exist in Fe3O4 solid solution together with Cr3+. Rangel et al.[29] reported a metallic Cu reduction peak at 220oC for their Fe-ThCu catalysts. TPR reducibility studies help us probe the distribution of Cu in the catalyst structure. It can be concluded that the preparation method plays a key role in determining the Cu dispersion in the iron oxide structure. The uniform distribution of the promoters in the iron oxide matrix is also consistent with the observations we made comparing the pH values at the onset of precipitation (for the pH measurement, same precursor concentrations were employed as during catalysts synthesis) for different components in the sol-gel versus precipitation techniques (Table 4.2). Fe, Al and Cu species precipitate in a wider pH range in an aqueous solution, which is the synthesis medium in both 1- step and 2-step preparation techniques. In the sol-gel synthesis, however, the pH for hydrolysis, precipitation and gellation steps are very close to each other, providing a more uniform distribution of the three species.

46

T (oC) Precipitation (room temperature) Sol-gel (50oC)

Fe(NO3)3 1.2 --

pH of precipitation Al(NO3)3 Cu(NO3)2 3.3 8.8 4.8 8.9

(C5H8O2)3Fe -9.0 (in ethanol)

Table 4.2: pH values at the onset of precipitation for components in Fe-Al-Cu catalysts

4.3.4 XPS analysis of Fe-Al-Cu samples

XPS results of calcined (450C, 4 hrs) and reduced (350C, 2hrs) Fe-AlCu catalysts are shown in Figure 4.5 (Fe 2p region) and Figure 4.6 (Cu 2p region). For all freshly calcined catalysts, the dominant iron phase detected on the surface was Fe2O3 located at 710.9 eV (Fe 2p3/2). After reduction under the same conditions, the Fe 2p peak was shifted to 710.4 eV, which was indicative of the magnetite (Fe3O4) phase. For reduced 2-step and 1-step Fe-Al-Cu, in addition to magnetite on the surface, there were shoulder bands at 706.6 eV, which can be ascribed to over reduction of magnetite to a metallic iron phase. However, with the sol-gel technique, metallic iron was not noticeable after catalyst reduction. For Cu 2p region in Figure 4.6, although the Cu species on all calcined catalysts were Cu2+, the surface concentration of Cu species was significantly lower on sol-gel catalysts than 1-step and 2-step catalysts. For reduced 1-step and 2-step Fe-Al-Cu catalysts, intense metallic Cu bands located 47

at 932.0eV (2p3/2) were observed whereas only weak metallic Cu signals were detected for sol-gel Fe-Al-Cu. This finding is consistent with TPR results which showed that Cu species were incorporated into the iron oxide matrix much more uniformly and less Cu stayed on the surface for sol-gel Fe-Al-Cu.

48

Figure 4.5: Fe 2p region of the X-ray photoelectron spectra for Fe-Al-Cu catalysts

49

Figure 4.6: Cu 2p region of the X-ray photoelectron spectra for Fe-Al-Cu catalysts

50

4.3.5 XRD of Fe-Al-Cu catalysts

XRD was performed to examine the crystalline phases present in Fe-AlCu catalysts. As shown in Figure 4.7, for calcined Fe-Al-Cu catalysts, hematite (-Fe2O3, ICDD # 01-1053) is the only crystalline phase in both 2-step and 1-step Fe-Al-Cu. CuO could be in the form of extremely small crystallites that is out of the detection limit for XRD or it is in an amorphous phase. Sol-gel Fe-Al-Cu catalysts demonstrated a mixture of two crystal phases: hematite (-Fe2O3, ICDD#01-1053) and maghemite (-Fe2O3, ICDD#04-0755), exhibiting an important structural difference from the other two catalysts. Two types of iron oxides are used as catalysts for WGS as reported by Liu and coworkers[14]: -Fe2O3 and -Fe2O3. -Fe2O3 has a hexagonal structure whereas -Fe2O3 possesses an imperfect spinel structure. In -Fe2O3, Fe3+ ions occupy the tetrahedral and octahedral positions with cation vacancies located at the octahedral positions, which can be represented by

(Fe83+)A(Fe40/33+8/3)BO32[12]. A denotes the tetrahedral sites whereas B indicates the octahedral sites and signifies the vacant cation sites. Both Fe2O3 structures need to be reduced to Fe3O4, which is the active phase for WGS. Fe3O4 has an inverse spinel structure with the oxygen ions forming a face-centered cubic close-packed lattice. In this lattice, the cations are located in tetrahedral A sites and octahedral B sites, with a unit cell written as (Fe83+)A(Fe83+Fe82+)BO32[66]. The electron hopping between Fe2+ and Fe3+ in the octahedral sites facilitate the oxidation and reduction cycle during WGS[54]. Because of the structural 51

similarity, -Fe2O3 can be converted into Fe3O4 through topotactic transformation, which significantly reduces sintering during catalyst activation compared with Fe2O3 under the same operating conditions[67]. Additionally, -Fe2O3 has been proven to be more active for WGS due to a smaller activation energy in comparison with -Fe2O3[54]. Another advantage for -Fe2O3 stands out in promoted Fe-based catalysts. In Fe-Cr system[14], it is reported that Cr3+ ions are incorporated into the -Fe2O3 structure by a continuous cationic interdiffusion process, which limits the doping amount into the lattice. However, the cation vacancies in -Fe2O3 make it markedly easier for the promoter ions to enter the iron oxide structure. -Fe2O3 can be formed by adjusting preparation parameters including the ratio of Fe2+/Fe3+ in the iron solution and the aging time[68]. In our work, -Fe2O3 is obtained by using different iron precursors and preparation methods. Formation of -Fe2O3 is also beneficial for successful incorporation of Al and Cu promoters into the iron oxide matrix, which leads to a uniform distribution of promoter species (TPR) and a smaller surface metallic Cu concentration (XPS).

52

Figure 4.7: XRD patterns of Fe-Al-Cu catalysts

53

4.3.6 Mssbauer spectra of Fe-Al-Cu catalysts

Complementary to the iron oxide crystal phase identification with XRD, Mssbauer spectroscopy was used to provide a more accurate analysis of the oxidation state and composition of bulk iron species. The spectra which showed different iron oxide species hyperfine parameters are listed in Table 4.3. Integrated area for each species is used to obtain its relative intensity. A representative spectrum was shown for 1-step Fe-Al-Cu in Figure 4.8. Mssbauer spectra for Fe-Al-Cu with Fe/Al=10 and Fe/Cu=20 catalysts prepared by 2-step, 1-step and sol-gel methods are composed of one sextet and one central doublet. The sextet lines can be assigned to large -Fe2O3 particles (>20 nm)[69]. However, both small superparamagnetic -Fe2O3 particles and promoter substituted -Fe2O3 give a central doublet in Mssbauer spectroscopy with similar parameters[54, 69]. Promoter substituted -Fe2O3 is a result of strong interaction between promoters and -Fe2O3, with foreign ions replacing octahedral vacant sites in -Fe2O3[54]. Therefore, the observed doublets for FeAl-Cu catalysts could come from contributions of both components. 2-step Fe-AlCu manifests a smaller doublet spectral area compared with the other two catalysts (34% doublet in contrast with 88% from 1-step and 85% from sol-gel). It is anticipated that in 2-step prepared catalysts the extent of interaction between promoters and iron oxide structure is less compared with 1-step and sol-gel synthesis, as seen in TPR and XPS studies which showed a separate CuO phase in 2-step catalysts and more Cu species staying on the surface instead of 54

entering and interacting with the bulk iron oxide structure. The relative ratios of small -Fe2O3 particles over promoter substituted -Fe2O3 cannot be determined from Mssbauer analysis for those Fe-Al-Cu catalysts. However, observation of -Fe2O3 crystal phase from XRD for sol-gel Fe-Al-Cu confirms the presence of promoter substituted -Fe2O3 in this catalyst whereas 1-step Fe-Al-Cu does not display this feature in its diffraction pattern. Therefore, it is possible that sol-gel catalyst has a higher portion of Fe2O3 in structure compared with 1-step catalyst. The structural difference between Fe-Al-Cu catalysts can be evidenced from XRD and Mssbauer spectra analysis. -Fe2O3 was formed in sol-gel catalysts whereas -Fe2O3 phase is dominant in the other catalysts we studied. It may be easier for Cu to enter into the -Fe2O3 structure because of vacancies in the octahedral sites, which may improve stability of -Fe2O3 as well[54]. Furthermore, catalyst activity is significantly improved as evidenced from our reaction testing. Effect of Cu loading amount in formation of -Fe2O3 structure and further improvement in catalytic performance will be examined in the next chapter.

55

Sample Fe-Al-Cu-2-step (Fe/Al=10, Fe/Cu=20) Fe-Al-Cu-1-step (Fe/Al=10, Fe/Cu=20) Fe-Al-Cu-SG (Fe/Al=10, Fe/Cu=20)

Splitting doublet sextet doublet sextet doublet sextet

Relative intensity (%) 34 66 88 12 85 15

(mm/s) 0.35 0.37 0.36 0.40 0.35 0.38

(mm/s) 0.78 -0.23 0.85 -0.26 0.73 -0.23

H (T) 0 49.1 0 53.1 0 51.0

-isomer shift (given with respect to -Fe); -quadrupole splitting; H-internal magnetic field

Table 4.3: Mssbauer parameters calculated from spectra of Fe-Al-Cu catalysts collected at 25 C

56

Figure 4.8: Mssbauer spectrum for 1-step Fe-Al-Cu

57

4.3.7 Temperature programmed reaction of Fe-Al-Cu catalysts

Figure 4.9 shows the generation of H2 (m/z = 2) and the consumption of H2O (m/z = 18) observed over Fe-Al-Cu catalysts during H2O-TPRxn. In this experiment, catalysts which are pre-reduced in situ, are subjected to a linear temperature program under a flow of H2O/He, while the H2 and H2O signals are monitored by a mass spectrometer. An increase in H2 intensity was accompanied by a decrease in H2O intensity. Catalysts experienced oxidation by H2O while producing H2 during this process. It is conceivable that the occurrence of the oxidation reaction is highly relevant to oxygen vacancies present in the sample. H2 signal was demonstrated to be the highest for sol-gel Fe-Al-Cu, implying that sol-gel catalyst may contain a larger amount of oxygen vacancies to react with H2O and form H2 than the other two catalysts. Several possibilities can be used to explain this phenomenon. As deduced from physisorption results, defect sites formed during sol-gel synthesis possibly provide sites for H2O adsorption and subsequent oxidation. The structural differences detected from XRD

measurement may also be relevant. Sol-gel catalysts consist of both - and Fe2O3 structures. The cation vacancies in -Fe2O3s imperfect spinel structure could promote the formation of oxygen vacancies. After oxygen vacancies in the catalysts were completely depleted, H2 signal relaxed to the baseline level as indicated from Figure 4.9.

58

(a)

(b)

Figure 4.9: H2O-TPRxn for Fe-Al-Cu catalysts (a) m/z=18 H2O consumption (b) m/z=2 H2 generation

59

It has been revealed from literature studies that oxygen vacancy on catalyst surface acts as an important component in WGS reaction

mechanism[68]. During WGS reaction, oxygen vacancies could promote the dissociative adsorption of H2O on the surface by splitting it into OH and H. OH groups replenish oxygen vacancies and interact with adsorbed CO to generate surface formates. A further decomposition of formate species yield CO2 and H2. Another group of researchers[29, 70, 71] believe that instead of dissociative adsorption of H2O into OH and H groups, oxygen from the H2O molecules can replenish the oxygen vacancies and release H2, which is the H2O reduction step during the WGS reduction-oxidation (red-ox) cycle. Our studies appear to support the red-ox mechanism as discussed in our previous publication[5]. Therefore, creation of oxygen vacancies during catalyst preparation or pretreatment favors the WGS redox cycle. The large amount of oxygen vacancies in sol-gel samples is beneficial for the reaction.

4.4 Conclusions

A novel sol-gel synthesis technique was used to prepare Fe-Al-Cu catalysts and compared with previously reported 2-step and 1-step Fe-Al-Cu catalysts. The sol-gel preparation provided the highest catalytic performance during the WGS reaction. TPR and XPS analysis showed that for sol-gel preparation, Cu was distributed more uniformly in the iron oxide structure. This not only prevented Cu sintering at high temperatures, but also ensured efficient 60

electron transfer properties of Cu. Both -Fe2O3 and -Fe2O3 crystal structures were created during sol-gel process. Formation of -Fe2O3 greatly enhances catalytic performance and incorporation of promoters into the iron oxide structure. Additionally, H2O TPRxn experiments revealed that sol-gel method created more oxygen vacancies during the WGS reaction and that this facilitated the oxidation and reduction cycle and thus enhanced the WGS reaction rate.

61

CHAPTER 5

EFFECTS OF COPPER LOADING ON THE CATALYTIC PERFORMANCE OF FE-AL-CU FOR WATER-GAS SHIFT REACTION

5.1 Overview of catalyst synthesis and Cu promoter in water-gas shift catalysis

Cu is an important element in WGS catalysis. Cu-ZnO-Al2O3 is a classical low temperature WGS catalyst. Reduced Cu (Cu0 or CuI or both) is believed to constitute the active sites[13]. Both metallic Cu and oxygen vacancies in ceria were identified to be the active sites in Cu-CeO2 catalytic system[72]. Cu is also a widely used promoter in high temperature WGS catalysts although its promotional function is still debated. One agreement is that Cu provides active sites in a similar way as in low temperature shift catalysts, especially at a lower reaction temperature. This has been reported in studies by Andreev et al.[17] and Idakiev et al.[57], where they examined the effect of CuO addition on the WGS performance of Fe-Cr and observed increased activity, which was attributed to the formation of active Cu assemblies. Another argument is that Cu is incorporated into the iron oxide structure and boosts the activity by modifying electron transfer properties. In studies by Hutchings et al.[11, 27], doping of only 62

2 wt% CuO in the Fe-Cr composition resulted in an evident enhancement in catalyst activity. They attributed effect of Cu to Cu2+ forming a solid solution with iron oxides and modifying electronic properties of the Fe3O4 structure during WGS reduction-oxidation sequences instead of directly providing active sites. From their experimental studies, no metallic Cu particles were detected using high resolution electron microscopy and the high spatial resolution energy dispersive X-ray analysis showed Cu to be dispersed in a solid solution of iron oxides without formation of a discrete CuO phase. As reported recently[73], our studies lead us to believe that Cu fulfills a dual function by providing both active sites (surface Cu) and electronic enhancement (structural Cu). These two features promote catalytic performance over a wide temperature range in tandem. To ensure catalyst stability for high temperature WGS operation, more structurally incorporated Cu species are desired, which is easier to achieve with sol-gel preparation, as shown in our studies. In addition to the synthesis technique, the Cu loading level is also likely to affect the WGS catalytic performance and structural characteristics of these Fe-Al-Cu catalysts. In this study, sol-gel Fe-based catalysts with different Cu loadings were prepared and evaluated over a wide range of temperatures in WGS reaction, in an effort to achieve a better insight into how the Cu loading influences catalyst physical and chemical properties and catalytic performance. In situ X-ray diffraction (XRD) was employed to study evolution of crystal phases during catalyst reduction for Fe-Al-Cu catalysts with different Cu loadings. Iron oxide

63

structures for catalysts with different Cu loadings were analyzed by XRD and Mssbauer spectroscopy measurement.

5.2 Experimental procedures

5.2.1 Catalyst preparation

Sol-gel catalyst preparation was performed as described in our previous work[73] with slightly modified conditions based on synthesis parameter optimization. The gelling temperature was maintained at 60oC during catalyst synthesis to increase the solubility of the organic iron precursor in ethanol solvent. The pH was kept at 11 to ensure complete precipitation of precursors. Fe/Al molar ratio was controlled at 10 for all catalysts. Fe/Cu ratio was varied from Fe/Cu=20, 10, 5 to 1. Fe-only, Fe-Al (Fe/Al=10) and Fe-Cu (Fe/Cu=20 and 5) catalysts were also prepared using the same sol-gel method as for Fe-Al-Cu catalysts.

5.2.2 Reaction studies

Catalyst testing was conducted using the same setup in previous studies[73]. Reactions for Fe-Al-Cu catalysts with different Cu doping levels at different temperatures were conducted with a WHSV (weight hourly space velocity) of 0.06 m3(g cata)-1h-1. Catalysts were first reduced at 350oC for 2 hrs 64

with a stream of 10% CO, 10% H2O, 7.5% H2, 5% CO2 and 67.5% N2. Catalysts were evaluated at different reaction temperatures, covering a range from 250oC to 400oC. Low temperature reaction data at steady state were first collected before temperature was ramped to the next stage. Catalysts were held at each temperature for at least 5 hrs to achieve steady performance. Selected Fe-Al-Cu catalysts (Fe/Cu=5 and Fe/Cu=1) were evaluated for time-on-stream studies at 400oC. For those experiments, equal surface area (4 m2) in the reactor was used. Flow rates were maintained the same during these runs. Catalyst pretreatment and feed stream compositions were kept the same as other reaction experiments.

5.2.3 Catalyst characterization

Surface areas of calcined catalysts were measured with a Micromeritics ASAP 2010 instrument. Samples were first degassed under vacuum at 130oC overnight. By using the adsorption-desorption isotherm at 77K, surface areas were calculated by the Brunauer-Emmett-Teller (BET) method. Pore volumes were obtained using Barret-Joiner-Halenda (BJH) adsorption curve. In situ XRD patterns were obtained using a Scintag XDS diffractometer equipped with an HTK-1200 oven, temperature control and gas flow capability. The X-ray source was Cu K radiation operated at 45 kV and 20 mA. 2 diffraction angle was varied from 20o to 90o during the measurement. For in situ reduction experiments, XRD patterns were collected first at 30oC. After that, air was purged from the instrument chamber and 5% H2/N2 was introduced to the 65

system. The temperature was ramped linearly to 800oC. During the reduction process, XRD patterns were taken starting from 100oC with 100oC interval and 30 min stabilization at each step. The International Center for Diffraction Data (ICDD) library was used for crystal phase identification. TPR was conducted with a laboratory built flow system. 100 mg catalyst was first sandwiched between quartz wools in a 1/4 in. O.D. quartz U-tube reactor. The catalyst was then treated under N2 at 450C for 1 hr. After cooling to room temperature under N2, using 5% H2/N2 as a reducing agent, the temperature of the catalyst was raised at a ramp rate of 10C/min to 800C and held for 10 min. During the TPR process, the effluent from the reactor, after a drying step through a silica gel water trap, was sent to a TCD (thermal conductivity detector) connected to a data-acquisition computer to monitor H2 consumption. A
57

Co/Rh -ray source and a conventional constant acceleration

Mssbauer spectrometer were used during spectra collection. Isomer shifts are given with respect to -Fe. All spectra were taken at room temperature and ambient atmosphere. The integrated areas under each deconvoluted peak have been used to obtain the relative populations of different iron species. An equal recoil free fraction was assumed for all iron species.

66

5.3 Results and discussion

5.3.1 Reaction performance for Fe-Al-Cu catalysts with different Cu loadings

As discussed earlier, Cu can provide active sites and improve electron transfer properties between Fe2+ and Fe3+. Cu loading amount is an important parameter for catalyst optimization. In this study, with Fe/Al molar ratio fixed at 10, four Fe/Cu molar ratios were used to prepare Fe-Al-Cu catalysts: 20, 10, 5 and 1. Figure 5.1 presents the effect of Cu loadings on reaction activity. For comparison the commercial Fe-Cr-Cu catalyst was also included. As shown in Figure 5.1, over a wide temperature range, sol-gel Fe-Al-Cu exhibited significantly better performance compared to the commercial catalyst. At all the other temperatures except 400oC, as Cu loading increased, CO conversion first increased, reached a maximum at Fe/Cu=5, then declined at Fe/Cu=1. At 400oC, the Fe/Cu=5 Fe-Al-Cu showed highest CO conversion, reaching equilibrium under reaction conditions used in this work. The difference among the other three catalysts appears marginal. These Fe-Al-Cu catalysts have varying physical surface areas (presented in the following section). To ensure that the activity trend we obtained is not merely caused by a surface area difference, Fe-Al-Cu with Fe/Cu=5 and Fe/Cu=1 were selected for long term activity evaluation on equal surface area basis. As can be seen in Figure 5.2, when tested using equal surface area in the reactor, the Fe/Cu=5 catalyst still gave a higher initial conversion than the 67

Fe/Cu=1 catalyst, but the difference was not as pronounced (56% versus 51%, respectively). Both catalysts showed some drop in activity in the first 30 hours on stream, however, the decrease was more pronounced for Fe/Cu=1, over which the conversion leveled off around 40% whereas the Fe/Cu catalyst maintained a conversion level of over 50%. This experiment showed that the better performance observed over the Fe/Cu=5 catalyst at equal WHSV is not simply a surface area effect, but this catalyst has indeed the highest intrinsic activity. These results also suggest that at higher Cu levels, deactivation of the catalyst with time-on-stream may be an issue.

5.3.2 Surface area and pore volume measurement

As shown in Table 5.1, both surface area and pore volume of Fe-Al-Cu catalysts decrease as Cu loading increases. Fe-Al-Cu with Fe/Cu=20 has a BET surface area of 89 m2/g. When Cu loading was increased to Fe/Cu=1, only a 38 m2/g surface area was obtained. This could possibly be caused by partial blockage of the pore volume or coverage of iron oxide surface by Cu. Another explanation is that when Cu loading is high, segregation of Cu particles may occur, which possibly leads to a smaller surface area.

68

Figure 5.1: CO conversion for Fe-Al-Cu catalysts with different Cu loadings at various reaction temperatures

69

Figure 5.2: Time-on-stream testing for Fe-Al-Cu catalysts with Fe/Cu=5 and Fe/Cu=1 using equal surface area in the reactor

70

Catalyst Fe-Al-Cu, Fe/Cu=20 Fe-Al-Cu, Fe/Cu=10 Fe-Al-Cu, Fe/Cu=5 Fe-Al-Cu, Fe/Cu=1

BET surface area (m2/g) 89 70 62 38

BJH adsorption pore volume(cm3/g) 0.54 0.45 0.43 0.09

Table 5.1: BET surface area and pore volume measurements for Fe-Al-Cu catalysts with different Cu loadings

71

5.3.3 In-situ X-ray diffraction during catalyst reduction

In-situ X-ray diffraction analysis was conducted on Fe-Al-Cu catalysts to monitor crystal phase changes during reduction. Representative diffraction patterns are shown in Figure 5.3 and Figure 5.4 for Fe-Al-Cu catalysts with Fe/Cu=5 and Fe/Cu=1, respectively. It should be noted that maghemite (-Fe2O3, ICDD # 04-0755) and magnetite (Fe3O4, ICDD # 11-0614) have very similar diffraction patterns in the 2 range we used (2=20o-90o) with only slight differences in relative intensities[74]. Therefore, temperature programmed reduction with H2 consumption monitored by a TCD detector was used to assist crystal phase identification in XRD patterns and TPR profiles for these two catalysts are presented as inset in the corresponding XRD figures. As shown in Figure 5.3 for Fe-Al-Cu (Fe/Cu=5) catalyst, maghemite is the only visible crystal phase at 30oC, with characteristic lines at (220)-30.26o, (311)35.59o, (400)-43.47o, (511)-57.17o and (440)-62.74o (hkl planes with

corresponding 2 values). Catalyst reduction started around 150oC as indicated by TPR profiles. Therefore, the XRD pattern at 200oC in Figure 5.3 can be ascribed to a mixture of maghemite and magnetite phases from partial reduction of maghemite. Further reduction of the catalyst occurred as temperature continued to increase above 200oC. A mixture of magnetite, metallic Cu (ICDD # 04-0836) (diffraction lines: (111)-43.30o and (200)-50.42o) and metallic Fe (ICDD # 06-0696) (diffraction lines: (110)-44.68o, (200)-65.02o and (211)-82.33o) phases coexisted at 400oC. Metallic Fe and Cu signals continued to grow with increasing 72

temperature. By 600oC, visible phases were metallic Fe and metallic Cu. TPR profile also shows a complete reduction, starting at 500C and reaching completion around 700C. Unlabelled diffraction lines detected at 800oC were from exposed sample holder (Al2O3, ICDD # 01-1296) due to catalyst volumetric shrinkage at high temperatures. With a higher Cu doping level, Fe-Al-Cu with Fe/Cu=1 displays a separate CuO phase (ICDD # 01-1117) (CuO diffraction lines: (-111)-35.75o and (111)38.96o) at 30oC besides maghemite (Figure 5.4). Metallic Cu crystals from CuO reduction appeared as early as 200oC. TPR profile (Figure 5.4, inset) shows two broad reduction features below 300oC, one around 200oC and the other around 240oC. These two features can be assigned to reduction of CuO and partial reduction of maghemite, respectively. By 400oC, CuO was completely converted to metallic Cu as seen in the XRD patterns. In addition to the magnetite phase, metallic Fe from over-reduction of magnetite was also detected. Metallic Fe and Cu were dominant phases from 600oC onwards. It appears that there is a limit as to how much CuO can be incorporated into the iron oxide structure. Although we could not identify the exact Cu loading where this occurs because of the limited number of Cu loading levels we used, it is possibly between Fe/Cu=5 and 1. Above this limit, free CuO phase is detected from catalysts as prepared and Cu species on the surface tend to aggregate and sinter at higher temperatures, which can partially explain the lower activity for FeAl-Cu (Fe/Cu=1) catalyst in comparison with Fe-Al-Cu (Fe/Cu=5) catalyst. Although Cu can provide active sites for WGS, its segregation as a separate 73

phase markedly diminish the catalytic performance. Kundu and coworkers reported similar findings from their studies on Fe-Cr catalysts with 0 to 20% Cr doping range[54]. When Cr concentration was less than 15%, Cr3+ was suggested to enter into the octahedral vacant sites of the -Fe2O3 lattice without segregating as a discrete Cr2O3 phase. Attempt to incorporate more Cr during catalyst preparation no longer led to crystallization of the Fe and Cr species in a single phase and separate Cr2O3 phase was observed in the final product, which decreases WGS performance of the catalyst. Although as we have reported previously[73], both hematite and maghemite crystal phases were observed for Fe-Al-Cu with Fe/Cu=20 catalyst at room temperature, when Fe/Cu is 5, maghemite is the only evident crystal phase. With more Cu in Fe/Cu=1 catalyst, on the other hand, CuO crystal phase was seen in addition to maghemite. These observations seem to suggest that the Cu loading controls the crystalline phases present in these Fe-Al-Cu materials. This assertion was further examined by preparing a series of Fe-based catalysts, including Fe-only, Fe-Al and Fe-Cu catalysts using the sol-gel technique. XRD patterns for these materials are presented in Figure 5.5. Hematite (-Fe2O3) is the major crystal phase for Fe-only catalyst. In the Al-promoted samples, -Fe2O3 structure is still the dominant phase. Two Fe-Cu catalysts are included in Figure 5.5, with Fe/Cu=20 and Fe/Cu=5. For sol-gel FeCu catalyst with Fe/Cu=20, -Fe2O3 is the main phase. However, when Fe/Cu ratio is 5, a diffraction pattern corresponding to -Fe2O3 is observed along with much weaker diffraction lines of -Fe2O3. Although the presence of Al does not 74

seem to affect the crystalline phase of Fe2O3, Cu clearly plays a role in promoting the formation of the -Fe2O3 phase. We could not completely rule out the possibility of substitution of Al3+ in -Fe2O3 vacant sites. DeBoer and Selwood reported Al3+ ions replacing octahedral sites in -Fe2O3[75]. Presence of Al3+ in Fe3O4 lattice was also reported from Wielers et al.s work[76]. XRD results in Figure 5.5 again point to the earlier statement we made that Cu doping level is an important factor that could control the formation -Fe2O3 structure, with or without Al. When Fe/Cu=5, -Fe2O3 was the major crystal phase, for both Fe-Cu and Fe-Al-Cu catalysts. -Fe2O3 is less thermally stable compared with -Fe2O3. Pure -Fe2O3 can completely transform into -Fe2O3 in the temperature range of 450oC to 600oC[54]. However, incorporation of foreign ions into the iron oxide lattice can significantly enhance the stability of -Fe2O3. Topse and Boudart conducted Mssbauer studies on Pb promoted Fe-Cr catalysts and found that Pb4+ enters the tetrahedral sites in -Fe2O3 structure and improves catalyst stability and activity[77]. In Kundu et al.s work[54], Cr3+ was reported to partially occupy the octahedral lattice sites in -Fe2O3. This strong interaction between Cr and Fe2O3 not only suppresses the transition from to phase, but also provides a boost to WGS activity. Lund and Dumesic[78-81] published a series of studies on the strong oxide-oxide interactions in SiO2 supported magnetite catalysts. A substitution of Fe3+ in the tetrahedral sites near the surface of Fe3O4 by Si4+ was confirmed by XRD and Mssbauer results and was reported to suppress the transition of -Fe2O3 into -Fe2O3 structure. 75

Effect of catalyst synthesis on the crystal structure of the iron oxide phase was examined in a previous study[73] where it was found that sol-gel prepared Fe-Al-Cu was composed of both - and -Fe2O3 phases, which was distinct from the dominant -Fe2O3 phase found in catalysts prepared with

precipitation/impregnation techniques (1-step and 2-step). In view of our work so far, it appears that the formation of the -Fe2O3 phase may be due to two factors, namely the preparation method and the presence of the promoters. As seen in Figure 5.5, addition of Cu and the amount of Cu incorporated into the iron oxide structure seems to control the phase composition of the iron oxide, either by promoting the formation of the -Fe2O3 phase or by stabilizing it.

76

Figure 5.3: In-situ X-ray diffraction patterns during reduction with 5% H2/N2 for Fe-Al-Cu catalyst (Fe/Cu=5). Inset: TPR profile for the same catalyst

77

Figure 5.4: In-situ X-ray diffraction patterns during reduction with 5% H2/N2 for Fe-Al-Cu catalyst (Fe/Cu=1). Inset: TPR profile for the same catalyst

78

Figure 5.5: X-ray diffraction patterns for Fe-based catalysts

79

5.3.4 Bulk structure analysis by Mssbauer spectroscopy

Bulk oxidation states and composition of the Fe-based catalysts were determined using Mssbauer spectroscopy. Spectra were deconvoluted to fit different iron oxide species. Parameters deduced from these spectra are listed in Table 5.2. Integrated area for each species is used to obtain its relative intensity. Representative spectra are shown in Figure 5.6, Figure 5.7and Figure 5.8. XRD employed in our work examines crystal phases present in the bulk within instrument detection limit whereas Mssbauer spectroscopy provides a more accurate analysis of the bulk iron species, their oxidation states and composition because this technique is not limited to crystallite detection and is extremely sensitive toward the iron element. Mssbauer spectrum acquired for sol-gel Fe-only catalyst consists of a sextet and a central doublet. Parameters calculated from the sextet (37%) imply large -Fe2O3 particles (>20 nm) whereas the doublet (63%), with an isomer shift of 0.35 mm/s and a quadrupole splitting of 0.81 mm/s, can be assigned to superparamagnetic small -Fe2O3 particles[69, 79, 82]. Fe-Al-Cu with Fe/Cu=20 spectrum (Figure 5.6) can also be deconvoluted into a sextet and a doublet with the sextet lines attributed to large -Fe2O3 particles. As has been discussed in our previous paper[73], the doublet is caused by either small -Fe2O3 particles or promoter substituted -Fe2O3 or both[54, 69]. Promoter substituted -Fe2O3 is a result of interaction between promoters and -Fe2O3, with foreign ions replacing the octahedral vacant sites in -Fe2O3[54]. 80

Different from Fe/Cu=20 catalyst, Fe-Al-Cu with Fe/Cu=5 spectrum (Figure 5.7) demonstrates a fairly weak sextet, with 0.34 mm/s isomer shift and no quadrupole splitting, which is characteristic of -Fe2O3 and accounts for 16% of the total intensity. The central doublet, which is of 84% of the total intensity, was ascribed to foreign ions substituted -Fe2O3 structure[54]. Another sol-gel sample we examined is Fe-Cu with Fe/Cu=5. Analysis of the spectrum shows a very small amount of large -Fe2O3 (7%) and 25% of promoter-substituted Fe2O3 indicated by a central doublet. The major bulk (~68%) is composed of Fe2O3, which was fitted using a distribution of sextets, implying that distribution of Cu in the structure may not be homogeneous. Mssbauer spectroscopy results are consistent with observations from XRD analysis and formation of -Fe2O3 is confirmed in Cu-promoted sol-gel Febased catalysts. Appearance of -Fe2O3 is largely dependent on Cu incorporation amount in the iron oxide structure.

81

Sample Fe-SG Fe-Al-Cu-SG (Fe/Al=10, Fe/Cu=20) Fe-Al-Cu-SG (Fe/Al=10, Fe/Cu=5) Fe-Cu-SG (Fe/Cu=5)

Splitting doublet sextet doublet sextet doublet sextet doublet sextet sextet

Relative intensity (%) 63 37 85 15 84 16 25 7 68

(mm/s) 0.33 0.36 0.35 0.38 0.34 0.34 0.31 0.38 0.31

(mm/s) 0.78 -0.23 0.73 -0.23 0.80 0.00 0.75 -0.23 -0.03

H (T) 0 50.8 0 51.0 0 47.2 0 51.8 48.0

-isomer shift (given with respect to -Fe); -quadrupole splitting; H-internal magnetic field

Table 5.2: Mssbauer parameters calculated from spectra of Fe-based catalysts collected at 25C

82

Figure 5.6: Mssbauer spectrum for Fe-Al-Cu-SG with Fe/Al=10, Fe/Cu=20

83

Figure 5.7: Mssbauer spectrum for Fe-Al-Cu-SG with Fe/Al=10, Fe/Cu=5

84

Figure 5.8: Mssbauer spectrum for Fe-Cu-SG with Fe/Cu=5

85

5.4 Conclusions

This chapter is a continuation of the investigations on surface and bulk structural properties in sol-gel prepared Fe-Al-Cu catalysts. The WGS performance changes significantly with catalyst preparation and Cu loading levels. Catalyst physical surface area showed a decreasing trend with higher Cu concentration, possibly due to pore blocking. Appearance of -Fe2O3 structure was found to depend on both catalyst synthesis and Cu loading amount, which was supported from iron oxide crystal phase, oxidation states and compositions obtained from XRD and Mssbauer studies.

86

CHAPTER 6

DEACTIVATION CHARACTERISTICS OF FE-AL-CU WATER-GAS SHIFT CATALYSTS IN THE PRESENCE OF HYDROGEN SULFIDE

6.1 Overview of water-gas shift catalyst deactivation studies in the presence of H2S

Hydrogen production from coal derived synthesis gas is a promising route to provide hydrogen feedstock for chemical industry or fuel cells. The water-gas shift (WGS) reaction is an integral step bridging a fuel gasifier and downstream H2 purification. A variety of impurities can be found in the synthesis gas from coal gasification, among which sulfur compound being a major one. Under WGS conditions, sulfur is most likely to be in the form of H2S[83]. Although in industrial operations large desulphurization units are usually installed after the gasifier, up to a few hundred ppm levels of H2S can still be present in coal syngas due to incomplete removal[84]. H2S is capable of deactivating WGS catalysts. As reported by Xue et al.[83], Cu-Zn low temperature WGS catalyst is extremely sensitive to H2S and 50 ppm H2S can lead to complete and irreversible activity loss. Fe-Cr catalysts are more resistant to sulfur, but they still exhibit a partial 87

decrease in activity upon sulfur exposure. Studies of catalyst deactivation in the presence of H2S are essential to provide a fundamental understanding of interactions between H2S and active sites and how this interaction affects the catalytic reaction between reactants and catalyst surfaces. Examination of deactivation characteristics is an indispensable step in design of sulfur tolerant catalysts and study of catalyst regeneration[85]. Because of the electronic structure of sulfur, it can bond strongly to transitional metals even at extremely low gas phase concentrations. This affinity to metal could cause marked activity loss in many catalytic reactions[86, 87]. Sulfur could influence catalysts both by physical and chemical adsorption[84-86, 88]. H2S could dissociatively adsorb onto catalyst surfaces by splitting into HSand H+ with the possibility of HS- further breaking down into S2-. Coordinative bonding between H2S molecule and the surface is another type of physical adsorption[89]. These interactions could block or inhibit active sites for catalytic reaction and lead to activity loss. Chemisorption of sulfur occurs through multiple coordination sites, which could induce surface reconstruction of the metal surface. This surface rearrangement is particularly detrimental to structure-sensitive reactions. For example, the carbon formation reaction during hydrocarbon steam reforming requires a relatively large ensemble of Ni atoms[90]. Presence of a small amount of sulfur in the feedstock facilitates surface rearrangement of Ni atoms into smaller assemblies, which significantly suppresses hydrocarbon decomposition and promotes the reforming reaction[88]. Similar phenomena

88

were observed in Pt catalytic systems where rearrangement of Pt particles caused by sulfur facilitates or suppresses certain reaction pathways[91, 92]. There have been extensive fundamental studies on sulfur poisoning of metals (review by Bartholomew[86]). Influence of sulfur on catalytic performance have been widely examined in different catalytic systems, such as

hydrogenolysis[91], hydrogenation[16, 87, 93, 94], steam reforming[88, 95, 96] and Fischer-Tropsch synthesis[97-102]. With increasing energy demands in recent years, coal derived synthesis gas has been used as the feedstock for solid oxide fuel cells (SOFC). Effects of H2S on SOFC anode materials have also been heavily investigated[84, 103]. However, very few studies on H2S effects in WGS catalytic system are available in literature. Although sulfur-resistant WGS catalyst has been developed for conversion of raw gases from coal or crude oil gasification which contain sulfur and traces of hydrocarbons, this molybdenumbased catalyst is only functional when a certain amount of H2S (>1%) is present in the feed to sustain the active sites[45, 46, 104]. This limitation necessitates research on catalyst formulations that are operative both under H2S-free and low concentration H2S (<1000 ppm) containing feed. Progress has been reported by Xue et al.[83] and Hutchings et al. [105-109]. Xue and his coworkers evaluated a number of WGS catalysts using sulfur containing feed. Co-Cr catalysts demonstrated high WGS activity both in the absence and in the presence of H2S (200 ppm) compared with Fe-Cr catalysts. In Hutchings and coworkers studies, Co-MnO and CoCr2O4 catalysts exhibited high resistance to H2S up to 240 ppm. This phenomenon was attributed to Mn and Cr functioning as sacrificial and 89

protecting the active Co component in the formulation. The stable spinel structure in CoCr2O4 catalysts also makes sulfidation reactions difficult. Sulfate instead of sulfide species were detected on the surface, leading only to a partial activity loss. This can be explained by the order of poisoning activity for sulfur species reported by Forzatti et al.[85]: H2S>SO2>SO42-. However, there are major drawbacks for Co-based WGS catalysts including lack of long term stability and methane formation due to undesired side reactions. Therefore, development of other alternative catalysts for coal derived synthesis gas application is still needed. In an effort to acquire an understanding of the effect of H2S on catalyst surface and structural properties during WGS that would assist in sulfur-tolerant catalyst design, this chapter examines a series of Fe-based catalysts and the surface and structural changes that they go through as a result of H2S poisoning. Characterization techniques were used including diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) to probe how reactants interact with catalysts before and after sulfur deactivation, X-ray photoelectron spectroscopy (XPS) to examine catalyst surface composition and oxidation states upon sulfur exposure, X-ray diffraction (XRD) and Mssbauer spectroscopy to identify structural changes.

90

6.2 Experimental procedures

6.2.1 Catalyst preparation

Preparation of catalysts evaluated in this study has been described in detail in our previous work[73]. Fe-only, Fe-Al and Fe-Al-Cu catalysts were prepared using sol-gel technique. Catalyst precursors were iron (III)

acetylacetonate (C5H8O2)3Fe, aluminum nitrate (Al(NO3)3.9H2O) and copper nitrate (Cu(NO3)2.3H2O). C2H5OH and NaOH were used as the solvent and precipitating agent, respectively. The gelling temperature was maintained at 60oC during catalyst synthesis. The pH was kept at 11 and Fe/Al molar ratio was controlled at 10. For Fe-Al-Cu catalyst, Fe/Cu ratio was kept at 5 in this work. Catalysts were calcined under air at 450oC (ramp rate = 5oC/min) for 4 hrs.

6.2.2 Reaction studies.

All reactions were performed with a WHSV (Weight Hourly Space Velocity) = 0.06 m3 (g cata)-1 h-1. For catalysts evaluation in the presence of H2S, catalysts were first reduced at 350oC for 2 hrs in a clean simulated coal gas mixture with 10% H2O, 10% CO, 5% CO2, 7.5% H2 and 67.5% N2. Fresh catalyst activity was measured with the same feed at 400oC. To achieve steady state catalytic performance, catalysts were kept on stream for at least 10 hrs. After this, catalysts were treated with 50 ppm H2S/N2 at 400oC for a certain amount of time 91

(2.5 hrs and 24 hrs were used in this paper). Reaction activity was again collected with the same clean coal gas feed as before sulfur exposure. Other details about reaction studies can be found in our previous publication[5, 73].

6.2.3 Catalyst characterization

A Thermo 6700 FT-IR spectrometer equipped with a DRIFTS cell, an MCT detector and a KBr beam splitter was used. DRIFTS spectra were collected with a 500-scan data acquisition at a resolution of 4 cm-1 in controlled gas atmosphere and temperature, using an environmental chamber with Zn-Se windows. For all experiments, catalysts were first pretreated in He at 450oC for 30 min. This was followed by a reduction procedure with 10% H2/He at 350oC for 2 hrs. After reduction, the chamber was purged in He for 30 min at 400oC and background spectra were collected after the flushing. 2% CO was introduced into the chamber for 15 min and sample spectrum was taken. A temperature programmed reaction was performed with 2% CO and 2% H2O starting from 30oC. Temperature was ramped progressively until 400oC. Sample spectra were collected at 30oC, 100oC, 200oC, 300oC and 400oC after holding at each temperature for 15 min. XPS was performed using an AXIS Ultra XPS spectrometer, operated at 13 kV and 10 mA with monochromated Al K radiation (1486.6 eV). Sample treatment with H2S was conducted at 400oC and at the end of the treatment, inert gas (N2) was used to purge and cool down the reactor. The sample was sealed 92

in N2 before transferring to an Ar-purged glove box to eliminate oxygen exposure. All binding energies were referenced to C 1s of 284.5 eV.
57

Co/Rh -ray source and a conventional constant acceleration Mssbauer

spectrometer were used during spectra collection. Isomer shifts are given with respect to -Fe. All spectra were taken at room temperature and ambient atmosphere. The integrated areas under each deconvoluted peak have been used to obtain the relative populations of different iron species. An equal freerecoil fraction for all species was assumed for all iron species. A 9-sample holder accessory was employed to collect diffraction patterns. The X-ray source was Cu K radiation operated at 45 kV and 20 mA. 2 diffraction angle was varied from 20o to 90o during the measurement. International center for diffraction data (ICDD) library was used for crystal phase identification.

6.3 Results and discussion

6.3.1 Catalytic performance in the presence of H2S

Different catalysts were evaluated with exposure to 50 ppm H2S for 2.5 hrs at 400oC and the results are presented in Figure 6.1. As shown in Figure 6.1, both sol-gel (SG) prepared Fe-Al-Cu and commercial Fe-Cr-Cu catalysts demonstrated a partial activity loss after sulfur exposure. However, under these testing conditions, SG prepared Fe-only and Fe-Al catalysts maintained over 93

90% of their fresh catalyst activities. This interesting observation impelled further investigation on sulfur resistance of catalysts with or without Cu promoter. In another test, with all the other conditions the same as in Figure 6.1, Fe-only and Fe-Al-Cu catalysts were exposed to 50 ppm H2S for 24 hrs at 400oC and reaction results are displayed in Figure 6.2. Although the initial fresh catalysts activities are dramatically different for promoted and non-promoted catalysts, both Fe-only and Fe-Al-Cu catalysts exhibited the same CO conversion levels after long time exposure to H2S. By combining reaction data in Figure 6.1 and Figure 6.2, an initial hypothesis on the role of Cu during catalyst deactivation can be suggested. It appears that the extent of activity loss is dependent on concentration of Cu in the catalyst. For the testing condition in Figure 6.1, Fe-based catalysts without Cu in the formulation shows very little deactivation considering the error range. A 40% decrease in CO conversion was observed for Fe-Al-Cu catalyst whereas commercial Fe-Cr-Cu lost 30% conversion after it was poisoned. This variation might be dependent on Cu doping levels in these two samples. Fe-Al-Cu contains 16 wt% CuO as prepared whereas less than 5 wt% CuO is present in the commercial catalyst. When exposure time was extended to 24 hrs, Fe-only catalyst also deactivated, probably due to interaction between iron species and H2S. To verify this speculation, extensive characterization studies were performed to examine surface and bulk changes of the catalysts after deactivation.

94

Figure 6.1: Reaction activity for both fresh and poisoned Fe-based catalysts

95

Figure 6.2: Comparison between Fe only and Fe-Al-Cu catalysts before and after 50 ppm H2S exposure for 24 hrs

96

6.3.2 Surface interaction between reactants and catalysts surface probed by DRIFTS

DRIFT spectroscopy was employed to examine catalyst surface changes upon sulfur exposure. CO and H2O temperature programmed reaction experiment was conducted on both freshly calcined and H2S poisoned (50 ppm H2S, 400oC for 5 hrs, cooled in N2) Fe-Al-Cu catalysts. Figure 6.3 illustrates species evolution with temperature for Fe-Al-Cu catalysts. As shown in Figure 6.3(a) for fresh Fe-Al-Cu catalyst, strong gas phase and weakly adsorbed CO bands were observed at 2180 and 2130 cm-1, respectively. Adsorbed CO results in the formation of CO2, evidenced by bands at 2360 and 2340 cm-1[5]. For this sample, CO2 was clearly seen even at room temperature, which is possibly caused by interaction between CO and the highly reactive surface oxygen through a redox mechanism, as discussed previously[5]. Weakly adsorbed carbonate species (1800-1000 cm-1[110]) were also present for this sample. For poisoned Fe-Al-Cu catalyst, similar CO and CO2 bands were observed in the 2400-2000 cm-1 region. However, CO2 signal intensity became less intense, accompanied by stronger adsorption bands from unreacted CO. Bands in the 1800-1000 cm-1 region are evident on the poisoned sample. Starting from room temperature CO adsorption, the band at 1650 cm-1 is associated with bidentate bicarbonate (-HCO3-) and the bands at 1540 and 1390 cm-1 can be assigned to bidentate formate (-HCOO-)[111, 112]. The presence of bidentate formate species can be further confirmed by bands located around 2970 and 2873 cm-1 97

caused by C-H stretching, although it was very weak at low temperature and became visible at higher temperatures during our experiment. By 200oC, the bidentate bicarbonate band at 1650 cm-1 grew stronger whereas a slight intensity decrease was observed for the bidentate formate bands at 1540 and 1390 cm-1, implying that conversion from bidentate formate to bidentate bicarbonate began to occur. New band at 1250 cm-1 appeared, which can be assigned to carbonate species[111, 112], which could be another product derived from bidentate formate. With further increase in temperature, all bands in this 1800-1000 cm-1 region showed a decrease in intensity. As reported in the literature[113], adsorbed CO and OH on the surface can react to produce formate species. In the presence of oxygen, formate can convert to bicarbonate and further oxidation leads to the formation of surface carbonate. This is consistent with our experimental observation. Combining the reaction testing and DRIFTS studies on both freshly calcined and H2S poisoned catalysts, it is speculated that a relationship may exist between surface carbonate species and catalyst deactivation. There are literature studies on effects of carbonate species on WGS catalyst activities[35, 113-115]. Carbonate and/or formate species that are strongly bonded to the surface could partially block the active sites and inhibit the adsorption of reactants on these sites, resulting in a reduced WGS activity. Formation of carbonate and/or formate species was also reported to influence catalyst electronic properties[113]. Formate species were proposed to be the reaction intermediates by Shido and Iwasawa[48-50]. They concluded that adsorbed terminal OH groups on the 98

surface can interact with CO and form bidentate formates. This reaction intermediate can decompose to H2 and unidentate carbonate and this process can be facilitated by the presence of H2O. CO2 is generated from carbonate decomposition. However, the formate and carbonate species observed from our studies may not function as precursors for final products formation. At the end of the temperature programmed reaction with CO and H2O reactants, the system was purged in He at 450oC. Weakly adsorbed CO and CO2 bands disappeared. However, no changes in the 1800-1000 cm-1 region can be detected, indicating that these species are rather stable and not likely to be reactive intermediates. Similar observations were reported by Liu et al.[113]. They found that carbonates could not be removed by heating up to 500oC and suggested accumulated carbonates that are strongly bound to the surface as the cause for catalyst deactivation. Based on the experimental results for fresh and poisoned Fe-Al-Cu catalysts, it is possible that the formate/carbonate species detected on poisoned catalysts could be caused by the surface change induced by H2S. Formate/carbonate species partially block the active sites for the WGS reaction and further suppress the activity. To get a better understanding of catalyst changes upon poisoning, including surface and bulk structure and electronic properties, a series of characterization work was performed in the following sections.

99

(a)

(b) Figure 6.3: DRIFT spectra during CO and H2O temperature programmed reaction: (a) Fe-Al-Cu catalyst; (b) 50 ppm H2S poisoned Fe-Al-Cu catalyst

100

6.3.3 Surface changes characterized by XPS

XPS provides surface composition and oxidation state changes upon sulfur exposure. The Fe 2p and S 2p regions for Fe-only catalysts after experiencing different treatments are illustrated in Figure 6.4. Approximate binding energy positions for possible phases are tabulated in Table 6.1. As shown in the Fe 2p region, Fe3O4 was observed on reduced catalyst surface with 2p3/2 binding energy located at 710.4 eV. Two H2S treated catalysts were examined. Fe-only-50ppm is Fe-only catalysts with exposure to 50 ppm H2S at 400oC for 24 hrs. Fe-only-1000ppm represents Fe-only catalysts that are dosed with 1000 ppm H2S at 400oC for 72 hrs. This harsh H2S treatment was employed to obtain noticeable changes on catalyst surface after poisoning. As shown from Table 6.1, peak positions in the 2p region for Fe species are very close to each other, which creates challenges for phase identification. For Fe-only-50ppm sample, a peak centered around 710.5 eV was detected. Possible iron phases include Fe3O4 located at 710.4 eV (2p3/2) due to surface reduction by H2S and FeS located at 711.1 eV (2p3/2) caused by the sulfidation reaction. The peak at 706.8 eV could be attributed to FeS2. Reactions that take place during this process could be FeOy + yH2S FeSy + yH2O[116] The actual reactions that occur are believed to be more complex as discussed later. FeS is reported as a result of Fe3O4 sulfidation by H2S in Twiggs review on WGS catalysis[117]. 101

Fe3O4 + H2 + 3H2S 3FeS + 4H2O Our statement on formation of multiple sulfidation products is not contradictory with the observation of only FeS reported in the literature. The above reaction is taking place in a H2-rich reducing atmosphere whereas the samples during our XPS measurements were pretreated under diluted H2S flow, with no H2. FeS, once formed, could transform into other sulfides or oxysulfides species in the presence of surface oxygen. This is supported by analysis of S 2p region. Examination of the S 2p region provides a better understanding of the sulfidation process. The peaks at 162.3 eV and 161.2 eV correspond to 2p1/2 and 2p3/2 binding energies from S2-, respectively, which confirms existence of FeS on poisoned samples. Different explanations can be found in the literature for the peak at 163.5 eV. In Pal et al.s studies on sulfur poisoned Pt catalysts[92], it was suggested that elementary sulfur with a 2p3/2 binding energy located around 164 eV could be present in the sulfide entities. Bromfield and his coworker[118] examined sulfided Fischer-Tropsch Fe catalysts using XPS and observed similar S 2p spectra as shown in our work. The second doublet around 163.0 eV in their work was ascribed to a different type of S2- or elemental sulfur caused by decomposition of polysulfide species. A deconvolution of the S 2p spectra of the Mo-containing catalysts by Andreev et al.[45] implied presence of subsulfides (S22x :

160.7 eV-2p3/2, 161.7 eV-2p1/2), sulfides (S2-: 161.8 eV-2p3/2, 162.7 eV-2p1/2),

disulfides (S22-: 162.6 eV-2p3/2, 163.5 eV-2p1/2) and elemental sulfur (S0: 163.5 eV-2p3/2, 164.5 eV-2p1/2). Oxysulfide species were also reported with the S 2p region in 162.3-163.2 eV range[104], which were formed by a partial replacement 102

of sulfur in the sulfidation product with oxygen atoms. Therefore, the S 2p region of our Fe-only poisoned catalysts could possibly consist of sulfide, subsulfide, oxysulfide and elemental sulfur species. Presence of subsulfides (FeS2) can be further supported by Fe 2p region analysis, which showed FeS2 band at 706.8 eV. Similar XPS spectra were collected over Fe-Al-Cu catalysts (Fe/Al=10, Fe/Cu=5, SG preparation). The Fe 2p regions in Figure 6.5(a) are composed of the same features as the corresponding Fe-only catalysts but with variation in intensity. More iron oxides and less FeS2 were noticed on the sample treated with 1000 ppm H2S. Examination of the Al 2p regions for Fe-Al-Cu catalysts (results not included) did not show any noticeable intensity variations between fresh and poisoned samples, suggesting that the interaction between H2S and Al2O3 is insignificant. In Ziolek et al.s studies on H2S adsorption on metal oxides[89], a small amount of sulfur adsorption (~0.2 wt% sulfur) was detected on Al2O3. This discrepancy could possibly be caused by H2S adsorption conditions employed in their work, in which ~2.8 volume% H2S was adsorbed onto bulk Al2O3. The Cu 2p region (Figure 6.5(b)) for sulfided samples consists of metallic Cu (932.0 eV-2p3/2) and Cu2S (931.5 eV-2p3/2). For the S 2p region (Figure 6.5(c)), in addition to the sulfide, subsulfide and oxysulfide species observed in poisoned Fe-only catalysts, a low intensity peak around 168 eV appeared, which was due to sulfate species. Sulfate species were formed as a result of sulfide species oxidation. This could occur either during the H2S treatment step using neighboring surface oxygen or due to surface phase rearrangement between sulfur and oxygen atoms during the cooling period under 103

inert gas[103]. The sulfide species formed under H2S flow could interact with surface oxygen under H2S lean conditions and lead to formation of sulfate species. The reduction-oxidation cycle has been discussed in our previous publication on Fe-based catalysts[5]. The active site for WGS is Fe3O4 (magnetite), which has an inverse spinel structure that can be represented as (Fe83+)A(Fe83+Fe82+)BO32. A and B represent tetrahedral and octahedral sites, respectively. During the reaction, the electron transfer between Fe2+ and Fe3+ in the octahedral sites in Fe3O4 promotes CO oxidation and H2O reduction reactions to produce CO2 and H2. Cu can play a dual role, by providing active sites (possibly with higher intrinsic activity) and by improving electron transfer properties between Fe2+ and Fe3+ during WGS. When catalysts are exposed to H2S, sulfidation reactions take place on the iron oxide surface as evidenced from XPS spectra. Furthermore, damage to the Fe3O4 active sites drastically influences electron transfer during the redox cycle. Therefore, an evident decrease in WGS activity was observed. When Cu is incorporated into the iron oxide structure, as a fresh catalyst, Cu significantly enhances the WGS activity. However, Cu species can interact strongly with H2S and form Cu2S. As suggested from our reaction studies, once Cu on the surface is poisoned, it possibly acts as an inert diluent and does not function as a promoter any more. This further diminishes catalytic activity. In addition, Lahtinen et al. studied influence of sulfur poisoning on CO adsorption on Co (0001) surface. It was revealed from their work that adsorption of CO on sulfur-precovered surface 104

decreases significantly. More importantly, a redistribution of CO adsorption sites on the surface is induced by the presence of sulfur[119]. Therefore, damage to CO adsorption sites decelerates progression of the redox cycle, which may be another reason for catalyst deactivation in the presence of H2S. A more severe activity loss in Cu-containing catalyst was also observed in our reaction experiments. As shown in Figure 6.1, under the same H2S treatment conditions (50 ppm, 400oC for 2.5 hrs), Fe-only and Fe-Al catalysts did not demonstrate much activity loss in stark contrast with a 40% decrease in CO conversion for Fe-Al-Cu catalysts. After an extended period of H2S exposure (Figure 6.2), Cu-promoted catalysts completely lost its advantage with respect to un-promoted catalysts.

105

(a)

(b)

Figure 6.4: X-ray photoelectron spectra for Fe-only catalysts after exposed to different concentrations of H2S: (a) Fe 2p region; (b) S 2p region

106

Phases Fe2O3 Fe3O4 FeO Fe FeS FeS2 CuO Cu Cu2S S2S4+ SO42-

Region 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2 2p3/2

Approximate position (eV) 710.9 710.4 709.4 706.6 711.1 706.8 933.2 932.0 931.5 161.2 166.0 168.0

Table 6.1: XPS binding energies of phases present in Fe only and Fe-Al-Cu catalysts

107

(a)

(b)

(c) Figure 6.5: X-ray photoelectron spectra for Fe-Al-Cu catalysts after exposed to different concentrations of H2S: (a) Fe 2p region; (b) Cu 2p region; (c) S 2p region 108

6.3.4 Phase composition and oxidation states examined by Mssbauer spectroscopy

Mssbauer spectroscopy was conducted to examine catalyst composition and oxidation states in the bulk phase. 1000 ppm H2S/N2 was used to poison FeAl-Cu (Fe/Al=10, Fe/Cu=5, SG preparation) catalysts at 400oC for 72 hrs. This is the same procedure used prior to XPS analysis. However, unlike the XPS analysis, samples were exposed to air after H2S treatment and spectra were collected under ambient atmosphere as well. Spectrum for the freshly calcined Fe-Al-Cu has been discussed previously[73] and is included here for comparison (Figure 6.6(a)). All spectra were deconvoluted to fit different iron species. Parameters obtained from these spectra (both freshly calcined and H2S-poisoned Fe-Al-Cu catalysts) are listed in Table 6.2. Integrated area for each species is used to obtain its relative abundance. Poisoned Fe-Al-Cu spectrum with peak deconvolution is shown in Figure 6.6(b). As has been discussed previously for calcined Fe-Al-Cu, the major iron oxide phase is -Fe2O3. Additionally, there could be a small portion of -Fe2O3 particles that are of very small particle size or promoter substituted -Fe2O3. Poisoned Fe-Al-Cu catalysts consist of one doublet and five sextets from spectrum deconvolution. The major doublet phase (60% derived from composition calculation) could be unreacted iron oxide species (promoter substituted -Fe2O3 or small -Fe2O3 particles) as observed in calcined catalysts. Another possibility for the doublet is FeS2 from surface iron species sulfidation[120]. Each sextet accounts for only a small portion. Three of 109

them (indicated in Table 2) could be assigned to Fe1-xS[120]. For the sextet that has an isomer shift of 0.23 cm-1, no reference could be found in the literature. The last sextet, which is an insignificant component in the sample (2% concentration), matches -Fe2O3 phase[69]. Presence of -Fe2O3 is not surprising as evidenced from XRD analysis in the following section. FeS is extremely air sensitive and decomposes readily to FeSx and -Fe2O3 even at room temperature from our XRD analysis on commercial FeS reference (Figure 6.7), which also explains why no stoichiometric FeS was detected in the poisoned sample. Fe1-xS is probably formed as a result of coordination environment changes among Fe, S and O atoms[120].

110

(a)

(b)

Figure 6.6: Mssbauer spectra of: (a) freshly calcined Fe-Al-Cu catalysts; (b) FeAl-Cu catalysts poisoned with 1000 ppm H2S/N2 at 400oC for 72 hrs 111

Sample

Splitting

Relative intensity (%)

(mm/s)

(mm/s)

H (T)

Species Small -Fe2O3 or promoter substituted Fe2O3

Fe-Al-CuSG (Fe/Al=10, Fe/Cu=5)

doublet sextet

84 16

0.34 0.34

0.80 0.00

47.2 -Fe2O3 Small -Fe2O3 or promoter substituted Fe2O3 or FeS2 Fe1-xS Fe1-xS Fe1-xS ? -Fe2O3

doublet Fe-Al-CuSG-H2S (Fe/Al=10, Fe/Cu=5) sextet sextet sextet sextet sextet

60 11 9 9 9 2

0.33 0.69 0.66 0.68 0.23 0.32

0.65 0.07 0.08 0.18 0 -0.20

0 30.4 26.1 23.1 35.3 50.0

-isomer shift (given with respect to -Fe); -quadrupole splitting; H-internal magnetic field

Table 6.2: Mssbauer parameters derived from the sample spectra collected at 25C

112

Figure 6.7: X-ray diffraction pattern of commercial FeS after air exposure

113

6.3.5 Crystal phases examined by XRD analysis

XRD was used to provide information on crystal phases present for calcined and poisoned catalysts. Freshly calcined, 50 ppm H2S (400oC, 24 hrs) and 1000 ppm (400oC, 72 hrs) H2S poisoned catalysts, which are the same as samples for XPS and Mssbauer analysis, were examined. XRD patterns are shown in Figure 6.8(a) for Feonly catalysts and Figure 6.8(b) for Fe-Al-Cu catalysts. The dominant crystal phase for calcined Fe-only catalyst is -Fe2O3. For the Fe-only-50ppm catalyst, Fe3O4 phase was observed, which is caused by Fe2O3 reduction. No sulfur related phases were detected. As reported in literature studies[121], when H2S concentration is low, only surface sulfide species will be formed because bulk sulfide is not thermally stable. However, 1000 ppm poisoned Fe-only catalyst demonstrates the presence of both FeS2-x and -Fe2O3, which are likely to be products from iron sulfide species decomposition in air as indicated from XRD pattern in Figure 6.8. It is also noticed that the diffraction lines for 1000 ppm poisoned catalysts are narrow and intense, implying that large crystals are generated during the sulfidation and oxidation process. Calcined FeAl-Cu catalysts are composed of -Fe2O3 phase whereas Fe-Al-Cu-50ppm consists of Fe3O4 crystallites. After 1000 ppm H2S exposure, Fe-Al-Cu catalysts contain multiple crystal phases: -Fe2O3, -Fe2O3, FeS2-x and Cu2S. This provides complementary information to Mssbauer results and confirms some of the assignments as well. Both iron and copper sulfided crystallites were detected. 114

Another indication from XRD patterns presented in Figure 6.8 is that larger Fe2O3 crystals were formed on poisoned Fe-only catalyst compared with poisoned Fe-Al-Cu catalyst. This could be explained by presence of promoters that can prevent or minimize aggregation of iron species to form large crystallites.

115

(a)

(b) Figure 6.8: X-ray diffraction patterns of freshly calcined and sulfur poisoned catalysts: (a) Fe-only catalyst; (b) Fe-Al-Cu catalysts

116

6.4 Conclusions

A series of Fe-based catalysts were evaluated for WGS reaction and characterized after they were exposed to H2S. All catalysts exhibited activity loss. However, the extent of deactivation varied. Cu-containing catalysts were more sensitive to H2S compared to Cu-free samples. XPS results suggested that both Fe and Cu could be sulfided by H2S. This modification induces changes in active sites on the surface, which was revealed from DRIFTS studies that showed carbonate species formed on poisoned catalyst surface, which are likely to suppress the WGS reaction. Formation of sulfide species was further substantiated by XRD and Mssbauer results. A deactivation mechanism was proposed from reaction and characterization studies. Catalyst activity loss was attributed to surface active sites changes induced by FeS and CuS2 formation, with Cu more susceptible to H2S poisoning and contributing to initial activity loss. It appears that in designing sulfur-tolerant catalysts, promoters which are tolerant to H2S, but capable of facilitating the WGS reaction would be better candidates. Another alternative is promoters that can protect iron oxides from sulfidation by functioning as sacrificial sulfur adsorption sites[96].

117

CHAPTER 7

CONCLUSIONS AND RECOMMENDATIONS

7.1 Summary of water-gas shift work

Fe-Al-Cu was developed to replace Fe-Cr catalysts for water-gas shift reaction. Catalyst preparation methods were found to play a critical role in determining catalytic performance. Sol-gel Fe-Al-Cu catalysts exhibited highest activities over a wide range of reaction temperatures compared with 2-step and 1-step Fe-Al-Cu and commercial Fe-Cr-Cu catalysts. The enhanced catalytic performance was attributed to uniformly distributed Cu promoter in catalyst structure and creation of new iron oxide crystal phase for sol-gel catalysts from detailed catalyst characterization. A much lower surface Cu concentration was detected for sol-gel Fe-Al-Cu catalyst from XPS analysis, implying that more Cu tends to stay in the bulk for sol-gel prepared catalyst. Cu distribution in the iron oxide structure was more uniform, evidenced from TPR experiment. Uniformly dispersed Cu and incorporation of more Cu in the bulk greatly prevent Cu sintering at high temperatures, which is a very serious issue for low temperature shift Cu-based 118

catalysts. Both -Fe2O3 and -Fe2O3 crystal structures were created in sol-gel Fe-Al-Cu catalyst as seen through the X-ray diffraction patterns. Formation of Fe2O3 was found to greatly enhance catalytic performance and promote incorporation of foreign ions into the iron oxide structure. Cu that stays in iron oxide structure efficiently modifies electron transfer properties and accelerates the redox cycle during water-gas shift. H2O TPRxn experiments revealed that solgel method created more oxygen vacancies during the WGS reaction and that this facilitated the redox cycle as well. Cu loading amount was optimized as an important synthesis parameter and catalyst performance was further improved. It was concluded that both synthesis technique and Cu loading level are important in formation of -Fe2O3 crystal phase. When Fe/Cu=5, -Fe2O3 is the dominant crystal phase in the catalysts (both Fe-Cu and Fe-Al-Cu sol-gel catalysts) from XRD measurement. Mssbauer spectroscopy studies further confirmed presence of -Fe2O3 and a better of Cu in the iron oxide structure for sol-gel catalysts. With an attempt to acquire an understanding of effect of H2S, which is a major impurity from coal derived synthesis gas, on catalyst surface and structural properties during WGS to assist sulfur tolerant catalyst design, catalysts were evaluated in the presence of H2S. It was found that all Fe-based catalysts suffered from activity loss after they were exposed to H2S. Cu-containing catalysts were more susceptible to H2S and sulfidation of Cu leads to initial catalyst activity loss. Long exposure to H2S can also sulfide Fe species in the formulation. Formation of sulfides possibly induces changes in active sites on the 119

surface, which can be used to explain appearance of carbonate species on poisoned catalyst from DRIFTS studies. Formate/carbonate species could partially block the active sites for the WGS reaction and further suppress the activity. When catalysts were dosed with higher concentration of H2S for an extended period of time, a small amount of sulfide species were detected by XRD and Mssbauer spectroscopy.

7.2 Future work for water-gas shift catalyst development

One of the advantages of sol-gel preparation for Fe-Al-Cu catalysts is that Fe3+, Cu2+ and Al3+ are precipitated at a very narrow pH window, which ensures a more homogenous mixing of three metal components in the final product compared with the precipitation or impregnation methods. However, because NaOH is used during sol-gel preparation, lengthy filtering, rinsing and drying have to be performed to remove Na ions (Na ions inhibit WGS performance). Therefore, for future work, new gelling agents can be examined so Fe, Al and Cu components can precipitate with the aid of the gelling agents instead of adding NaOH. This will eliminate the time-consuming catalyst washing step. Possible gelling agents include ethylene glycol[122] and propylene oxide, both of which are mild gelling agents and can assist the formation of network between different components. Alternative catalysts can be explored in addition to the Fe-based catalytic system. One highly promising area is CeO2 supported catalysts. CeO2 is

120

a widely used catalyst support for its oxygen storage capacity. Cu and Co can be investigated as the active metal dispersing on the CeO2 support. From catalyst evaluation in the presence of H2S, Cu-containing Fe-based catalysts are susceptible to H2S and serious activity loss was detected. Initial catalyst activity loss was largely caused by interaction between H2S and exposed Cu on the surface. Therefore, to improve sulfur tolerance, ferrite structure (CuFe2O4) can be pursued as a candidate for WGS. CuFe2O4 has a much higher percentage of Cu than the current Fe-Al-Cu catalyst. The current Fe-Al-Cu catalysts suffered from Cu sintering when the Cu doping level is high. However, in ferrite structure, Cu atoms are distributed uniformly as part of the crystal lattice. Formation of CuO will not occur until the temperature reaches 1000C. More importantly, decent activity has been reported for CuFe2O4 catalyst with extremely low surface area (1.0 m2/g) by Tanaka[123]. A much higher activity can be expected if CuFe2O4 with a much higher surface area can be synthesized. Promoters can be incorporated to further boost the stability and activity. For the H2S testing reported in this work, 50 ppm H2S/N2 was used. Catalysts were exposed to H2S for a certain period of time. Reaction activity was collected with clean coal derived synthesis gas. For future, catalyst evaluation with continuous H2S containing synthesis gas can be performed. As revealed from our H2S related work, catalysts would be more sulfur tolerant if it is under a more oxidizing atmosphere. Feed stream compositions can be adjusted, particularly increasing the H2O percentage in the feed, to minimize poisoning of catalysts by H2S. 121

PART 2

HYDROGEN PRODUCTION FROM PROPANE AND ETHANOL STEAM REFORMING

122

CHAPTER 8

INTRODUCTION TO STEAM REFORMING

Steam reforming is the most widely used fuel processing technology for hydrogen production. Other hydrogen production technologies, for example, eletrochemical, photobiological or photoelectrochemical processes, have limited applications due to either cost issue or inadequate development for industrial applications[124]. Among the current world H2 feedstock supply, almost 50% is generated from steam reforming. In the US market, steam reforming accounts for 95% of the H2 formation[125]. Steam reforming was first developed in 1926 and has become a well-established conventional technology in industrial

operation[126, 127]. H2 can be commercially manufactured from a variety of sources through steam reforming. Natural gas has been the dominant fuel supply for this process over many decades. With increasing energy demand and depleting natural gas reserves, renewable sources have become very attractive feedstock alternatives for hydrogen production. Bio-derived liquids (such as bio-ethanol, acetone, acetaldehyde, etc) from biomass fermentation are popularly researched energy sources in recent years. In addition to the sustainability of bio-derived liquids, 123

steam reforming of bio-derived liquids addresses environmental concerns involved during utilization of fossil fuels because this process does not emit pollutants such as VOC, NOx or SO2. CO2 produced can be recycled back to the atomosphere during plant growth.

124

CHAPTER 9

LITERATURE REVIEW ON STEAM REFORMING

9.1 Literature review of propane steam reforming

Steam reforming of hydrocarbons has seen increasing interest due to the necessity of efficient and cost-effective reforming technologies for hydrogen production for fuel cell applications. In practice, steam reforming of

hydrocarbons, especially that of methane, is performed at high temperatures over Ni-based catalysts where the lower cost of nickel is an important advantage. Current Ni-based catalysts, which are used for natural gas steam reforming processes, suffer from catalyst deactivation by coke formation and sintering of metallic Ni active phase[126, 127]. It is generally agreed that coke deposition is the major cause for activity loss. Boudouard reaction (2COC+CO2) and methane/hydrocarbon decomposition[128] are two major side reactions during hydrocarbon steam reforming, which could lead to filamentous carbon accumulation at the surface of Ni catalysts[129]. The coking process expedites active metal degradation and activity loss. Carbon growth is significantly decelerated on noble metals because 125 carbon could not dissolve in

them[127, 130, 131]. Therefore, supported precious metals such as Pd, Pt, and Rh are adopted for hydrocarbon reforming reactions and have exhibited superior activity and stability[132-137]. However, the cost of the precious metals is still a major drawback. The low-cost and long-proven performance of Ni-based catalysts, therefore, still warrant the effort to optimize these catalysts for steam reforming applications, particularly for fuel processing from existing liquid fuels such as gasoline and diesel. Extensive efforts have been made on development of Ni catalysts with improved resistance to coke formation. The support material has been found to exert a strong influence on the reactivity of CHx surface intermediates formed during hydrocarbon reforming[90, 131]. The number of hydrogen atoms in surface CHx species is relevant to the coking resistance of Ni catalysts[138]. Lower x values tend to result in the formation of carbonaceous deposits, a precursor of surface coke deposition. The nature of the support materials has been investigated in recent years in an attempt to seek supports with the desired acid-base or redox properties for the catalysts or modify existing supports by addition of promoters to enhance the reactivity of surface intermediates or suppress coke dissolution[139]. Alkali metal oxides such as K2O and CaO have been shown to improve coking resistance by enhancing carbon gasification, but at the expense of catalytic activity[126, 140]. The introduction of a small amount of molybdenum or tungsten (0.5 wt% MoO3 or WO3) into Ni catalysts has been shown to increase the coking resistance without loss in catalytic activity[141145]. Alternatively, lanthanides emerge to be promising promoters because they 126

can inhibit coke formation without sacrificing the catalytic activity[146-148]. Zhuang et al.[146] investigated the effect of cerium oxide as the promoter in MgAl2O4-supported Ni catalysts for methane steam reforming at 550C. Cerium promotion showed a beneficial effect by not only decreasing the rate of carbon deposition, but also increasing the catalytic activity. Wang and Lu[147] stated that the addition of CeO2 into Ni/Al2O3 catalysts enhanced the nickel dispersion and reactivity of carbon deposits, leading to an improvement in the catalytic activity and stability in CO2 reforming of methane. Earlier work from our group has shown that incorporation of lanthanides into the Ni matrix reduces coking by inhibiting the diffusion of carbon into Ni particles[149, 150]. Metal-support interaction is another phenomenon that affects carbon deposition in steam reforming. The metal-support interaction was found to correlate with Ni particle size, Ni dispersion and morphology during the reduction procedure[151]. The propensity to coking, to a large extent, depends on the size of Ni particles. It is reported by Edwards and Maitra[90] that it requires an ensemble of 6 or 7 atoms of Ni to catalyze carbon formation and growth whereas only 3 or 4 atoms of Ni are needed for the CH4 reforming reaction. Stronger metal-support interaction has been proven to enhance Ni dispersion dramatically, which would prevent the formation of large Ni cluster not only during catalyst preparation and activation, but also in the high temperature reaction process. However, this interaction between Ni and the support leads to the formation of solid solution or other mixed oxide compounds and sharply decreases the amount of available Ni for the targeted steam reforming reaction. This can be 127

compensated by increasing the doping amount of Ni. Magnesia has been found capable of controlling the size of metallic clusters for Ni-Al2O3 system[90]. The presence of mobile Ti-O in Ni/TiO2 could break the large ensembles of Ni atoms, which stimulate coke formation[131]. Other promoters including B, Ge, Ga and Zn were also examined in modifying metal-support interactions[152]. In addition to doping other elements to enhance the interaction, another strategy is development of novel preparation techniques to improve metal dispersion and coking resistance. Co-precipitation method was adopted in Pelletier et al.s work[139] and it is evidenced that this method is effective in achieving high dispersion and improving coking resistance. Creation of solid solutions between NiO and MgO during catalyst synthesis yield crystals of smaller size with high reaction activity and stability[153-155].

9.2 Literature review of bio-ethanol steam reforming

Bio-ethanol is a highly attractive sustainable energy source for H2 generation. Bio-ethanol can be renewably produced from fermentation of biomass, including starch, sugar, and cellulitic and lignocellulitic materials. Hydrogen production from biomass or biomass-derived liquids can theoretically be a carbon-emission-free process because all carbon dioxide released can be recycled back to the plants through photosynthesis. In addition, ethanol does not contain impurities such as sulfur which is a catalyst poison in many processes. Ethanol is non-toxic and easy to transport. It has been accepted by both 128

academia and industry as an economically and environmentally attractive starting material for H2 production. Although hydrogen production from methane steam reforming has been well established for many years, commercial catalysts for this process are not suitable for bio-ethanol reforming because ethanol has C-C bonds in the molecule and the reforming process involves a complex reaction network. Extensive research on catalytic reforming of bio-ethanol has only been performed in the last ten years. Additional fundamental work is needed to truly understand the reaction mechanism and develop efficient catalysts. In theory, ethanol steam reforming occurs via the following reaction. C2H5OH(l) + 3H2O(l) 2CO 2 + 6H2 ( Hrxn = 347.4kJ/mo l at 25o C) In reality, this reaction never proceeds straightforwardly as shown above. Various components lead to diversity in product distribution. Reaction conditions including temperature, pressure, feed composition and reactor design play a determining role. In addition, depending on types of catalysts, alternative reaction pathways may be favored. Below are some of the possible reactions that may take place during the process. C2H5OH CH4 + CO + H2 (ethanol decomposit ion) C2H5OH CH3CHO + H2 (dehydrogenation) C2H5OH C2H4 + H2O (dehydration) C2H5OH + H2O 2CO + 4H2 (incomplete reforming) 2C 2H5OH (C2H5 )2 O + H2O(dehydrative coupling)
CO + 3H2 CH4 + H2O(methanat ion)

129

CH4 + 2H2O CO2 + 4H2 (steam reforming) CO + H2O CO2 + H2 (water - gas shift) 2CO CO2 + C(Boudouard reaction)

These side reactions not only lead to an extremely complicated reaction network, but also lower the H2 yield. Dehydration and Boudouard reactions can induce coking on the catalyst surface and cause pressure build-up in the reactor, together with lifetime shortage of the catalysts. Bio-ethanol steam reforming has been an active research area in recent years. Researchers have investigated different aspects of this process. Several main areas of interest include: the support material, active metals, promoters, preparation methods, thermal treatment, the reaction mechanism, reaction conditions and process design[124, 156]. Researchers are making efforts to achieve catalysts that are selective to desired products, active in converting ethanol, and stable over long term operation.

9.2.1 Overview of support materials for ethanol reforming catalysts

The support material plays a crucial role in catalyst development. Choice of support can affect loading amount and dispersion of active metals, which can be further correlated with catalyst reducibility, selectivity and activity. Auprtre et al.[157] examined Al2O3, CeO2-Al2O3, CeO2, Ce0.63Zr0.37O2 as support for different active metals. They found that one key property for the support was high OH group surface mobility. CeO2 containing supports were investigated to 130

promote activity with the Ce0.63Zr0.37O2 support being the most promising. Burch et al.[158] examined Al2O3 and ZrO2-CeO2 supported catalysts. Al2O3 supported catalysts had high dehydration activity especially at lower temperatures, which led to coking problems. ZrO2-CeO2 supported catalysts significantly improved reaction performance, and there was no ethylene formation. Haga and his coworkers[159] evaluated catalytic performance of Co supported on Al2O3, SiO2, MgO, ZrO2 and C. Al2O3 was shown to be highly selective because of its ability to suppress CO methanation and ethanol decomposition. Ticianelli et al.[160, 161] studied Co supported on Al2O3, SiO2 and MgO. The interaction between Co species and supports varied, and was related with catalyst reducibility. All supported catalysts had high conversion levels but they promoted different side reactions which led to serious deactivation in less than 10 hrs. Acid sites on Al2O3 facilitated ethanol dehydration, whereas Co/SiO2 favored methanation and Co/MgO produced the highest amount of CO. Llorca and his coworkers[162-164] performed an exhaustive study on supported Co catalysts by exploring properties of MgO, Al2O3, SiO2, TiO2, V2O5, ZnO, La2O3, CeO2 and Sm2O3. The redox and acid-base properties of the oxides can change the catalytic performance dramatically. Acid surfaces promote the dehydration reaction whereas basic surfaces favor the dehydrogenation reaction. The Co/ZnO formulation was proved to be both selective and active in bio-ethanol steam reforming. In another paper[165] by Llorca et al., CO-free H2 was generated through reforming, catalyzed by Co/ZnO catalysts.

131

After researchers realized the importance of the support acid-base property, they attempted to modify it by using mixed oxide supports. Duprez and his coworkers[166] developed MgxNi1-xAl2O4/Al2O3 and MgAl2O4 spinel oxide supports by co-impregnation and solid-solid reactions of precursors, respectively. These supports were non-acidic and moderately basic and were proved to aid the function of active metals to the best extent. The porous structure of the support was also a parameter in determining metal surface densities and dispersion[167]. Depending on the support, interactions between the support and the metals were significantly different[168]. Jacobs et al. studied how Co species interacted with Al2O3, TiO2, SiO2, ZrO2-SiO2 and ZrO2-Al2O3. Differences in Co cluster particle size on the surface, catalyst reducibility and active site availability during reaction were observed. Sun and his coworkers[169] observed similar phenomena regarding effects of support-metal interactions on catalyst selectivity and activity by studying Ni supported on La2O3, Y2O3 or Al2O3. It was concluded that Ni/Y2O3 was a promising catalyst for hydrogen production from ethanol steam reforming with its high activity and long-term stability in low and middle temperature ranges. Verykios et al.[170-172] studied Ni catalysts supported on La2O3, Al2O3, YSZ, TiO2 and MgO. From their testing, Ni/La2O3 catalysts exhibited long term stability due to removal of the coke deposition on Ni surfaces by formation of lanthanum oxycarbonate species. Because of its unique oxygen storage property, CeO2 has been heavily studied as a support for the ethanol steam reforming. Laosiripojana and coworkers[173] employed a surfactantassisted technique to synthesize high surface area CeO2 and used it as an 132

internal pre-reforming catalyst under solid oxide fuel cell temperatures to convert all ethanol and other high hydrocarbon compounds. CeO2 showed excellent resistance toward carbon deposition. Song et al.[174, 175] examined nanocrystalline CeO2 supported Ni-Rh bimetallic catalysts for oxidative steam reforming of ethanol. Their work indicated that CeO2 support properties (crystal size and surface area) can modify the catalysts properties in terms of geometry (metal particle size and dispersion) and electron transfer (metal-support interaction). Zhang et al.[176] studied CeO2 supported catalysts and under harsh conditions (stochiometric ratio of ethanol to water), they observed long term stability for 300 hrs.

Supports Al2O3, CeO2-Al2O3, CeO2, Ce0.63Zr0.37O2 Al2O3, ZrO2-CeO2 Al2O3, SiO2, MgO, ZrO2, Carbon Al2O3, SiO2, MgO MgO, Al2O3, SiO2, TiO2, V2O5, ZnO, La2O3, CeO2 and Sm2O3 MgxNi1-xAl2O4/Al2O3 and MgAl2O4 spinel oxide supports Al2O3, TiO2, SiO2, ZrO2-SiO2 and ZrO2-Al2O3 La2O3, Y2O3 and Al2O3 La2O3, Al2O3, YSZ, TiO2 and MgO CeO2

References [157] [158] [159] [160, 161] [162-164] [166] [168] [169] [170-172] [173-176]

Table 9.1: Supports used in ethanol steam reforming catalyst preparations

133

9.2.2. Overview of active metals or promoters for ethanol reforming catalysts

Another important component for ethanol reforming catalysts is the active metal. The nature of the metals can determine how ethanol is adsorbed on the surface and which reaction pathway is followed. There are two major categories of active metals: noble metals and common metals. Auprtre et al.[157] compared a series of metals for ethanol reforming: Rh, Pt, Ni, Cu, Zn and Fe. A combination of Rh and Ni led to both high conversion and selectivity. Rh/Al2O3 was studied by Cavallaro et al.[177] and Rh greatly promoted the conversion of C2 species produced in the reforming process. Frusteri et al.[178] tested MgO supported Pd, Rh, Ni and Co catalysts for molten carbon fuel cell ethanol reforming. Rh/MgO showed the best performance, though H2 selectivity was moderate. Other metals suffered from serious sintering at high operating temperature (650oC). Rh/CeO2-ZrO2 catalyst was evaluated by Idriss et al.[179] and a H2 yield as high as 5.7 mol per mol of ethanol was achieved at temperature ranges of 350-450oC when water was in excess. Rh was also examined by Verykios and his coworkers[172] and found to be significantly better than Ru, Pt and Pd when metal loading was low. Ru supported catalyst activity was dramatically improved with increased loadings. Piscina et al.[180] studied the Pd/ZnO system and activity was attributed to Pd-Zn alloy formation. However, CO and acetaldehyde selectivities were quite high with this catalyst. Pd was impregnated on Al2O3 by Tsiakaras et al.[181] and they conducted lots of reaction condition studies using this system. 134

Ni is the active metal for commercial methane steam reforming catalysts and is still being used in ethanol steam reforming catalytic system[169, 182-184]. Cu-Ni-K/Al2O3 formulation was developed in Labordes group[185, 186]. This system exhibited high ethanol reforming activity at 300oC which could be ascribed to the combined promotional effects of all components. Metallic Cu facilitated fast ethanol dehydrogenation to acetaldehyde. Ni favored C-C bond breakage to convert liquid products and produce methane and carbon monoxide. Potassium acted as a neutralizer to balance the acidic sites on the support so as to minimize dehydration reaction. Llorca et al.[162-164, 187-189] conducted extensive work on a Co supported system and Co/ZnO was proven to be an excellent ethanol reforming catalyst with high activity and selectivity to desired products. Ir was found to interact strongly with the CeO2 support and this resulted in high dispersion, less sintering and resistance to coking over a long period of time[176].

135

Active metals Rh, Pt, Ni, Cu, Zn and Fe Rh Pd, Rh, Ni and Co Pd Ni Cu-Ni-K Co

References [157] [172, 177] [179] [178] [180, 181] [169, 182-184] [185, 186] [162-164, 187-189]

Table 9.2: Active metals used in ethanol steam reforming catalyst preparations

9.2.3 Overview of active sites and reaction mechanism during ethanol steam reforming

Hydrogen production from ethanol reforming involves complicated reaction network and reactions can take place toward different pathways depending on characteristics of the catalytic system. There are many parameters that can determine which routes the reaction will take. It has been generally accepted that ethanol was initially adsorbed on various metal oxide surfaces by dissociation of OH bond in ethanol molecule, with ethoxy species bonded to surface cation and proton from OH interacting with lattice oxygen to form new OH bond on the surface. After that, depending on nature of the support and active metals, ethoxy species will evolve differently. It may go to dehydrogenation or dehydration, which was proposed in Fornasiero et al.[190]s studies. Acidic surface favors ethanol dehydration to form ethylene whereas basic surface promotes ethanol dehydrogenation to produce acetaldehyde. Ethylene formed could be reformed 136

into CO and H2 in the presence of water or result in coke deposition on the surface. Acetaldehyde formed leads to acetone formation via oxidation reaction. Another part of acetaldehyde can be decomposed into CH4 and CO. There are also CH4 reforming and CO water-gas shift reaction occurring. To address the coking issue, support can be modified to neutralize the acidic sites. K was often used to counterbalance the acidic sites. Researchers have also investigated basic support for ethanol reforming. Frusteri et al.[178] conducted mechanistic studies in MgO-supported system. With the basic support, the main route is ethanol dehydrogenation. Acetaldehyde subsequently decomposes into CH4 and CO. The exit stream is governed by CH4 reforming and the water-gas shift reaction. Not only the support, but also the active metals play a key role in reaction network. Cu/Ni/K-Al2O3 catalyst showed acceptable activity, stability and H2 selectivity during ethanol reforming in Mario et al.s work[186]. In this formulation, Cu promotes fast ethanol dehydrogenation to acetaldehyde. Ni catalyzes C-C bond cleavage of C2 species to form CH4, CO2 or CO and improve H2 yield. K neutralizes acidic sites of Al2O3 and minimizes formation of ethylene and diethylether and thus improves catalyst stability. Therefore, to take account of contributions from both the support and active metal, a bi-functional mechanism was proposed by Duprez et al.[191] in alkylated aromatics selective reforming reaction. Later on, Auprtre et al.[157] discussed the role of supports in ethanol steam reforming. Ethanol was activated on metal particles whereas water was dissociatively adsorbed onto the support. 137

Support can promote the migration of OH groups towards the metal particles, catalyze the reforming reaction and stabilize the metal particles at high temperature under steam. Another parameter in mechanism studies is the effect of reaction temperature. Sun et al.[169] suggested different reaction mechanisms depending on operative conditions. It was found that different reactions dominate in different temperature regimes. One very important technique in reaction mechanism studies is infrared spectroscopy. Spectroscopy can provide information as regard to how reactant molecules interact with catalytic surface and what types of intermediate species are involved in the reaction. There are extensive literatures on ethanol reforming mechanism by using infrared spectroscopy[192-198]. A better understanding of how ethoxy species react on the surface to form final products through intermediate species and how H2O is adsorbed and takes part in the reaction can possibly be obtained through infrared studies.

138

CHAPTER 10

EFFECT OF PREPARATION METHOD ON STRUCTURAL CHARACTERISTICS AND PROPANE STEAM REFORMING PERFORMANCE OF NICKEL SUPPORTED ON ALUMINA CATALYSTS

10.1 Overview of steam reforming catalyst preparation

Propane steam reforming was studied over Ni-Al2O3 catalysts that were prepared by a conventional impregnation (IM) method and a one-step sol-gel (SG) technique. The catalyst prepared with the IM method showed high initial activity. However, it deactivated severely with time-on-stream of propane steam reforming. The catalyst prepared using a SG technique demonstrated stable catalytic performance. Catalysts were characterized with temperature-

programmed reduction (TPR), X-ray photoelectron spectroscopy (XPS), temperature programmed oxidation (TPO), transmission electron microscopy (TEM), X-ray diffraction (XRD) and Raman spectroscopy. It was revealed that, with sol-gel preparation, highly dispersed small Ni crystallites are formed with a strong interaction with the support. This was proved to be beneficial for coke suppression and catalyst stability. 139

10.2 Experimental procedures

10.2.1 Catalyst preparation

Ni-Al2O3 catalysts were prepared using SG and conventional IM methods. Nickel precursor (Aldrich) used during catalyst preparation was in nitrate form. Aluminum tri-sec-butoxide (ATB; Aldrich) was used as aluminum precursor. For SG synthesis, ethanol (Alfa Aesar) was used as the solvent. The following variables were controlled during the synthesis: nickel content = 20 wt%, H2O/ATB molar ratio = 3.6, and pH of the resulting gels = 4.8. Initially, ATB was dissolved in ethanol. The aqueous solution with the desired amount of nickel was added drop-wise into the ethanol solution using a syringe pump at a flow rate of 0.5 cm3/min. The pH of the resulting green gels was measured and adjusted by adding HNO3 (Fisher) or NH4OH (Fisher). The gels were stirred for an additional 15 min and were kept at room temperature for 30 min. The samples were then dried overnight in the oven at 110C to remove the remaining water and ethanol. The dry samples were ground to fine powders and were calcined under O2 at 450 C for 4 hrs. For the IM synthesis, the Al2O3 support was prepared first with the SG technique as described above. Ni species were introduced onto the Al2O3 support using impregnation in aqueous solution.

140

10.2.2 Characterization

The surface areas of oxidic samples calcined at 450C in O2 were measured using the BET method with a Micromeritics ASAP 2010 instrument, using N2 as the adsorbent at liquid N2 temperature (77 K). The samples were degassed at 130C for 12 hrs before surface area measurements. The volumetric measurement of H2 chemisorption was conducted using a Micromeritics ASAP 2010 Chemisorption system. Prior to adsorption measurements, calcined samples were reduced in-situ under H2 at the desired reduction temperature (600C) for 2 hrs followed by evacuation to 10-5 mmHg and cooling down to 35C. The adsorption isotherms were measured at equilibrium pressures between 50500 mmHg. The first adsorption isotherm was established by measuring the amount of gas adsorbed as a function of pressure. After completing the first adsorption isotherm, the system was evacuated for 1 hr at 10-5 mmHg. Then a second adsorption isotherm was obtained. The amount of probe molecule chemisorbed was calculated by taking the difference between the two isothermal adsorption amounts. Temperature-programmed reduction of catalysts was conducted using a laboratory-made gas flow system. The catalyst sample (50 mg) was loaded in a 1/4 in. O.D. quartz U-tube reactor. The catalyst was then re-calcined under 10% O2/He at 450C for 1 hr followed by cooling to room temperature under Ar. The reduction was performed with 10% H2/Ar (40 cm3 (STP)/min). The reactor temperature was raised using a ramp rate of 10C/min to 900C and held at that 141

level for 10 min. Effluents from the reactor were passed through a silica gel water trap before it reached the detector to remove moisture formed during reduction. H2 consumption was measured continuously as a function of sample temperature using a thermal conductivity detector (TCD) connected to a data-acquisition system. In situ X-ray diffraction patterns during reduction of catalysts were obtained with a Bruker D8 Advance X-Ray diffractometer equipped with an atmosphere and temperature control stage and using Cu K radiation ( = 1.542 ) operated at 40 kV and 50 mA. The powder diffraction patterns were recorded in the 2 range from 20o to 80o. Reduction was performed in-situ under 5% H2/N2 gas flow using a linear temperature-program between isothermal steps (50C). The catalysts were kept at isothermal steps for 0.5 hr before data collection and the ramp rate between the isothermal steps was 10C/min. A 9-sample holder accessory was employed to collect diffraction patterns when no treatment was needed, including calcined, reduced and post-reaction Ni-Al2O3 catalysts. X-ray photoelectron spectroscopy analysis was performed using an AXIS Ultra XPS spectrometer, operated at 13 kV and 10 mA with monochromator Al K radiation (1486.6 eV). Reduction was performed with 20% H2/N2 at 600C for 2 hrs in a reactor. The samples were then sealed using the valves located at both ends of the reactor. The reactors were then taken to an argon-atmosphere glove box, where the catalysts were removed out of the reactor and mounted on the XPS sample holders using a conductive carbon tape. Sample holders were transferred to the XPS chamber without exposing them to the atmosphere using 142

a controlled-atmosphere transporter. All spectra were corrected using the C 1s signal located at 284.5 eV. Temperature-programmed oxidation of post-reaction catalysts was

performed using a laboratory-made gas flow system. Initially, 10 mg of postreaction catalyst was placed in a 1/4 in. O.D. quartz U-tube reactor and was heated under He (40 cm3 (STP)/min) at 400C for 30 min to remove adsorbed moisture and other species (e.g. CO2). During TPO, the reactor temperature was raised under 5% O2/He (40 cm3 (STP)/min) with a ramp rate of 10C/min to 900C. Effluents from the reactor were continuously monitored as a function of sample temperature using a mass spectrometer (Thermo Finnigan Trace GC ultra-Trace DSQ MS) with He as the carrier gas. Raman spectroscopy was conducted with a Horiba Jobin-Yvon LabRam HR800 spectrometer. The illumination source employed is an internal HeNe laser with 633 nm wave length. The instrument is equipped with a CCD detector and a standard Olympus microscope. The laser power at the sample was controlled below 3 mW to avoid laser heating effects. Transmission electron microscope used in this study was Tecnai TF-20 operated at 100 kV. Samples were ultrasonically dispersed in ethanol, and a couple drops of the suspension were deposited on a standard 3 mm copper grid covered with a holey carbon film.

143

10.2.3 Reaction studies

Propane steam reforming reaction experiments were performed using a stainless steel fixed-bed flow reactor (1/4 in. O.D.). Catalytic performance of different catalysts were compared using an equal surface area loading (9 m2) in the reactor at 500C and C3H8/H2O/N2 = 1/4/95. Prior to a reaction, catalyst was reduced in situ under 20% H2/N2 at 600C for 2 hrs. Effluents from the reactor were analyzed using an automated Shimadzu GC-14A equipped with flame ionization detector (FID) and TCD detectors. Separations were performed under Ar using two columns: Porapak Q (12 ft x 1/8 in. SS, 80/100 mesh) and Molecular Sieve 13X (5 ft x 1/8 in. SS, 60/80 mesh). A GOW-MAC 069-50 ruthenium methanizer operated at 350C was used with the FID for accurate quantification of CO down to 10 ppm concentration level. The post-reaction samples were passivated with 1% O2 in He for 2 hrs at room temperature before being unloaded. Then they were saved in a desiccator for TPO and TEM experiments.

10.3 Results and discussion

10.3.1 Physical surface area and metallic Ni surface area

Table 10.1 lists the physical surface area of calcined catalysts and bare support and the metallic Ni surface area of reduced catalysts. IM-Ni-Al2O3 144

catalyst was prepared by impregnating Ni(NO3)2 onto sol-gel prepared Al2O3 support. SG-Ni-Al2O3 catalyst was prepared by a sol-gel technique with precursors of Ni(NO3)2 and aluminum tri-sec-butoxide in an ethanol solvent. SGNi-Al2O3 catalyst yields a higher surface area compared with IM-Ni-Al2O3 catalyst, which is possibly due to loss of pore size caused by Ni impregnation[139]. This is supported by the pore volume measurement that IM-Ni-Al2O3 has a pore volume of 0.26 cm3/g in comparison with 0.63 cm3/g for SG-Ni-Al2O3. H2 uptakes at mono-layer coverage of the metallic Ni were used to estimate surface Ni atom area, assuming that each surface Ni atom chemisorbs one hydrogen atom (H/Nisurface = 1). It should be noted that H2 chemisorption is a measurement of reducible Ni species on the surface. Reduced SG-Ni-Al2O3 catalysts demonstrate a lower metallic surface area than reduced IM-Ni-Al2O3 catalysts. This can be possibly attributed to two reasons. First, the impregnation method may lead to more Ni species staying on the surface instead of entering the bulk as in sol-gel preparation. Another reason is that for Ni species exposed on the surface in both catalytic systems, variation in extents of reduction may exist. In IM catalysts, large Ni particles or weak interaction with neighboring support could make these Ni species easy to reduce. However, in SG catalysts, as would be evidenced by XPS or XRD results in later sections, there are much smaller Ni crystallites and stronger connection between Ni and the support. Therefore, complete reduction of Ni species on the surface is difficult for the reduction conditions employed during H2 chemisorption.

145

Sample SG-Al2O3 O-450-20% IM-Ni-Al2O3 O-450-20% SG-Ni-Al2O3 R-600-20% IM-Ni-Al2O3 R-600-20% SG-Ni-Al2O3

Surface area (m2/g) 300 204 238

Metallic Ni surface area (m2/g)

9.4 6.8

Table 10.1: Surface area and metallic Ni surface area

10.3.2 XPS of Ni species in calcined and reduced catalysts

Figure 10.1 presents the Ni 2p region of the X-ray photoelectron spectra of Ni-Al2O3 catalysts calcined at 450C and reduced at 600C (O-450 and R-600). For peak identification, the spectra were deconvoluted based on the constraints of equal spin-orbit splitting for Ni 2p peaks, the height ratio of Ni 2p3/2 to Ni 2p1/2 being constant at 2 (theoretical value), and the full width at half maximum (FWHM) of these peaks being equal. For calcined catalysts, the peak located between 855 eV and 857 eV can be interpreted as a combination of two Ni 2p3/2 peaks with binding energies of 854.9 eV and 856.1 eV. There have been previous investigations of the XPS Ni 2p spectral region in the Ni-Mo--Al2O3 system by our group[199] as well as by others[200, 201]. The peak at 854.9 eV suggests the presence of NiO on the surface, which can be supported by strong shake-up lines around 860.7 eV (2p3/2), indicative of Ni2+ species. There has 146

Figure 10.1: X-ray photoelectron spectra (Ni 2p region) of Ni-Al2O3 catalysts. O-450: calcined at 450 C; R-600: reduced at 600 C

147

been no affirmative assignment of the peak at 856.1 eV reported in the literature, with the possibility of NiAl2O4 or Ni2O3. We tend to ascribe this feature to a spinel NiAl2O4 structure on the surface, which can be supported by our preparation methods creating a stronger bonding between Ni and the Al2O3 support. This assignment is also in agreement with findings obtained from other

characterization results. The surface spinel model proposed by Jacono et al.[200] where Ni2+ ions occupy the tetrahedral sites in -Al2O3 again supports this possibility. The two deconvoluted peaks described earlier (NiO-854.9 eV and NiAl2O4-856.1 eV) were integrated and the quantified percentages were shown in Table 10.2. SG-Ni-Al2O3 exhibits a larger portion of NiAl2O4 species, 49% in comparison to 27% for that of IM-Ni-Al2O3.

Sample O-450- IM-Ni-Al2O3 O-450- SG-Ni-Al2O3 R-600- IM-Ni-Al2O3 R-600- SG-Ni-Al2O3

Ni-852.7 70 53

Ni species (%) Ni-854.9 73 51 -

Ni-856.1 27 49 30 47

Table 10.2: Content of Ni species for catalysts prepared by impregnation and solgel methods: calcined and reduced samples

After a reduction step at 600oC, the NiO peak at 854.9 eV disappeared, replaced by a new peak centered at 852.7 eV (2p3/2). This corresponds to 148

metallic Ni species as a result of NiO reduction. It was noted that, over both catalysts, NiO species were completely converted to metallic Ni species under our reduction conditions. However, there were negligible changes for the peak at 856.1 eV. One can conclude that NiO on the support surface can be reduced more easily whereas the Ni species at Ni 2p3/2 of 856.1 eV is hardly reduced at 600C, which is most likely NiAl2O4 species instead of Ni2O3 mentioned earlier. Studies from literature[199, 201] reveal that NiAl2O4 species is much more difficult to be reduced compared to NiO species. From the Ni 2p region of reduced catalysts (Figure 10.1 and Table 10.2), for IM-Ni-Al2O3, a larger portion of metallic Ni (70% compared with 53% for that of SG-Ni-Al2O3) was detected on the surface, which explains the H2 chemisorption measurement that there are more reducible Ni species on the impregnated catalyst surface and the sol-gel catalyst has a substantially lower degree of reduction due to spinel formation.

10.3.3 X-ray diffraction

Complementary to the catalyst surface characterization derived from XPS studies, diffraction patterns were collected for calcined catalysts and catalysts after the same reduction procedures as in the XPS studies, i.e., 20% H2/N2, 600oC, 2 hrs. The reduction was conducted using the reaction system and catalysts were purged and cooled down in inert atmosphere at the end of the reduction period. Catalysts were then sealed before being transferred to the 149

Figure 10.2: X-ray diffraction patterns of calcined Ni-Al2O3 catalysts

150

Figure 10.3: X-ray diffraction patterns of Ni-Al2O3 catalysts after reduction

151

sample holder for XRD analysis to minimize air exposure. Diffraction patterns are presented in Figure 10.2 (calcined catalysts) and Figure 10.3 (reduced catalysts). The International Center for Diffraction Data (ICDD) library was used for phase identification. As discussed earlier, a possible interaction product between Ni and the Al2O3 support is NiAl2O4, which has a similar crystallite structure as Al2O3 with partial occupation of vacancies by Ni in the fcc oxygen packing[202]. Al2O3 is characterized by (311), (400) and (440) diffraction lines at 36.3o, 45.4o and 66.3o (2 values) respectively whereas the peaks at 37.3o, 45.4o and 66.5o represent NiAl2O4 diffraction patterns. Therefore, it is difficult to differentiate between the two crystalline phases. As indicated in Figure 10.2, the diffraction patterns observed at those positions could be due to -Al2O3 or a combination of -Al2O3 and NiAl2O4 phases. However, the difference from sample to sample is dramatic. The broad peaks in SG sample suggest an amorphous nature in the structure whereas IM sample has well-defined crystallites formed. Another distinct difference lies in NiO, characterized by (111), (200), (220) and (311) diffraction lines located at 37.3o, 43.3o, 62.9o and 75.4o (2 values) respectively. During IM catalyst preparation, dispersion of Ni on the support is poor. Ni species tend to aggregate and form large crystallites after the calcination process, which can be evidenced from the narrow NiO diffraction lines in Figure 10.2. The very broad NiO peaks for SG catalysts validate well-dispersed nature of Ni in the sample. After reduction, the SG catalyst maintains its amorphous nature, without large -Al2O3 or NiAl2O4 crystallites separating out from the structure. Again, not much crystallinity from metallic Ni was observed. On IM catalysts, very intense 152

metallic Ni diffraction lines, (111)-44.6o, (200)-52.0o and (220)-76.6o, were shown in Figure 10.3. As discussed earlier, Ni crystallite size formed after reduction is directly related to catalyst stability, with large Ni clusters catalyzing coke formation reaction. In-situ X-ray diffraction patterns were also collected during a step-wise temperature-programmed reduction process where a calcined sample was reduced in 5% H2/N2. 5% H2/N2 was chosen instead of 20% H2/N2 used in reaction testing because of safety regulations imposed by the facility. Diffraction patterns were collected at different temperatures including 500oC, 600oC, 700oC and 800oC. At each temperature, a 30 min holding time was used before data collection. Because of overlapping between Al2O3 (or NiAl2O4) (400) diffraction line and the most intensive metallic Ni (111) diffraction line, Ni (200) diffraction line broadening was used to calculate the metallic Ni crystallite size with Scherrer equation. The FWHM was obtained with GRAM software. The calculated results are reported in Table 10.3. For both Ni-Al2O3 catalysts, they started with the same Ni crystallite size at 500oC (9 nm). With increase in reduction temperature, the crystallite Ni size tends to grow larger, evidenced by a sharper and more intense diffraction peaks. For sol-gel catalysts, metallic Ni size stabilized at 13 nm whereas for impregnated catalysts, it continued to grow with increase in reduction temperature. By 800oC, this size was more than double the size at 500oC. These findings are consistent with other characteristics of the sol-gel preparation technique (high BET surface area, improved thermal stability, small particle size) summarized by Gonzalez et al.[203]. The less metal-support 153

interaction in impregnated catalysts makes the Ni particles easily aggregate into polymeric structures and form bigger Ni particles after reduction.

Catalyst IM-Ni-Al2O3 SG-Ni-Al2O3

d (nm) at different reduction temperatures (C) 500 9 9 600 15 12 700 17 13 800 20 13

Note: Calculation based on broadening of Ni (200) diffraction lines Table 10.3: Ni crystallite size after reduction at different temperatures

10.3.4 TPR profiles of Ni-Al2O3 catalysts

Temperature-programmed reduction experiments on Ni-Al2O3 catalysts were performed using 10% H2/Ar as the reducing agent. As shown in Figure 10.4, TPR profiles exhibit two distinct reduction features accompanied by a shoulder peak at high temperature. A small low-temperature reduction peak around 400C could be assigned to the reduction of trace amounts of NiO clusters or bulk NiO weakly interacting with the support, which was reported by Song and his coworkers[204] as the approximate reduction temperature for unsupported NiO. Also in their publication, Ni species supported on -Al2O3 show a reduction feature at 500-700oC. Correspondingly in our TPR profiles, the broad high-temperature reduction peak around 600C can be ascribed to the reduction of NiOx species that have stronger binding to the support. Sol-gel catalyst 154

displays a slight shift of this reduction peak (at 620oC) compared with impregnated catalyst (at 596oC). Further interaction between NiO and the support would lead to the formation of NiAl2O4, which could not be reduced until 800oC[204]. The tail reduction peak after 800oC for SG-Ni-Al2O3 might have contributions from NiAl2O4 reduction. For IM-Ni-Al2O3 catalyst, it demonstrates a shoulder peak around 765oC. This could possibly be caused by reduction of large NiAl2O4 particles formed in IM catalysts. For SG catalysts, reduction of NiAl2O4 did not occur until 800oC, which is expected because of the well dispersed small NiAl2O4 crystallites in SG-Ni-Al2O3. The H2 consumption can be determined by integrating the peak area under the TPR curve for the two Ni-Al2O3 catalysts. An area ratio of 1.3 was obtained for IM-Ni-Al2O3 over SG-Ni-Al2O3. This suggests that more Ni species can be reduced over impregnated catalysts, which is consistent with our statement that there is stronger metal-support interaction for sol-gel catalysts and this structure difference yields more reducible NiO species for impregnated catalysts.

155

Figure 10.4: Temperature-programmed reduction profiles of Ni-Al2O3 catalysts

156

10.3.5 Steam reforming of propane over reduced Ni-Al2O3 catalysts

Reduced (600C) 20% Ni-Al2O3 catalysts prepared by sol-gel and impregnation methods were evaluated with regard to activity and stability in propane steam reforming at 500C. The feed composition used in the experiments was C3H8/H2O/N2 = 1/4/95. The products are H2 and C-containing species (CO, CO2, and CH4 with CO2 being the main one) from propane steam reforming as reported in our earlier publication[149]. Figure 10.5 presents the propane conversion vs. time-on-stream. The IM-Ni-Al2O3 catalyst shows much higher initial activity than SG-Ni-Al2O3 catalyst. This is expected because there is more reducible Ni species on the surface for impregnated catalysts (evidenced from H2 chemisorption, XPS and XRD results) and formation of mixed oxide phases between Ni and Al2O3 takes away from Ni available for the steam reforming reaction, which was mentioned in Pelletier and his coworkers work as well[139]. However, IM-Ni-Al2O3 catalyst exhibits a much faster deactivation rate in comparison with sol-gel sample. Propane conversion drops to 50% of its initial value after 28 h time on stream. SG-Ni-Al2O3 catalyst demonstrates a relatively stable activity. At the end of 28 hours, the conversion on SG-Ni-Al2O3 is significantly higher than that of IM-Ni-Al2O3. The stability difference presented in Figure 10.5 could be associated with the Ni cluster size. There have been discussions in the literature about the relationship between metallic Ni size and catalyst stability [90, 127, 139, 151, 205, 206]. Larger Ni particle size catalyzes the coke formation reaction, which is 157

detrimental for catalyst stability. As seen in the X-ray diffraction patterns, reduced IM-Ni-Al2O3 catalyst (600oC, 2hrs) has larger metallic Ni crystallites, leading to substantial activity loss. The instability of metallic Ni with increase in reduction temperature expedites the carbon formation reactions. Apparently, catalyst preparation plays an important role in determining Ni dispersion and particle size, metal-support interaction and morphology. In Chen and Rens studies[155], catalyst stability is significantly improved if NiAl2O4 is formed during the pretreatment procedure by varying the synthesis. The solid solution in NiO-MgO catalytic system yields smaller Ni size and less pressure buildup caused by coking[153-155].

158

Figure 10.5: Propane conversion as a function of time-on-stream in steam reforming reaction of propane over Ni-Al2O3 catalysts

159

10.3.6 Coke deposition characterization on post-reaction catalysts

In this section, temperature programmed oxidation, XRD and Raman spectroscopy were used to characterize the propane steam reforming postreaction catalysts. TPO experiments were performed with 5% O2/He (40 cm3 (STP)/min) at a ramp rate of 10C/min from room temperature to 900C. Two peaks were observed from the TPO profile (Figure 10.6). A low temperature peak around 350oC is likely caused by combustion of highly reactive monoatomic carbon species deposited on Ni surfaces (Type I). Another peak of much higher intensity around 630oC can be assigned to filamentous carbon (Type II), which is typically stable and oxidized at higher temperatures[150, 207, 208]. Because of large intensity scale differences for these two carbon oxidation peaks, the low temperature oxidation peak was plotted as an inset in Figure 10.6. It is shown that sample preparation leads to differences in this low temperature feature. For SG catalysts, there is a plateau between 330oC and 380oC resulting from type I carbon oxidation before the major type II carbon combustion takes off. This peak for IM catalyst has lower intensity and it quickly evolves into type II carbon combustion stage. Although the trends of CO2 evolution and O2 consumption are similar for the two post-reaction catalysts, the amount of CO2 produced and O2 consumed during the oxidation of carbon deposits on 20% Ni-Al2O3 catalysts are significantly different. A much larger amount of carbon is deposited on Ni-Al2O3 catalysts prepared by the impregnation method during steam reforming of propane. The ratio of CO2 produced for IM/SG catalysts is about 1.5 as 160

Figure 10.6: Temperature-programmed oxidation profiles of carbon deposited on post-reaction Ni-Al2O3 catalysts

161

calculated by integrating the peak area under the m/z=44 evolution curve. Figure 10.7 presents the diffraction patterns for post-reaction Ni-Al2O3 catalysts. In addition to the crystal phases Ni, NiAl2O4 and -Al2O3 shown in Figure 10.3, a broad peak at 2=26.3o was detected, which is characteristic of hexagonal carbon (002). Another feature at 43.7o is a combination of carbon (100) and (101) diffraction lines. The carbon peak on the sol-gel sample is not noticeable, implying an amorphous nature for the coke deposited. Figure 10.8 shows the Raman spectra of Ni-Al2O3 catalysts. Two distinctive bands are observed for post-reaction IM-Ni-Al2O3 catalyst, at 1320 cm1

and 1597 cm-1. However, on SG-Ni-Al2O3 catalyst, they are very weak and

broad. The 1200-1700 cm-1 Raman spectra region is widely used to study disorder in carbon deposits, with two characteristic lines: D line around 13391345 cm-1 and G line around 1592-1597 cm-1[162]. The peak broadening for solgel sample indicates a very poorly graphitized structure. The more visible peak for impregnated catalysts suggests that a certain ordering of the carbonaceous deposit is developing under steam reforming reaction conditions used in this study. This observation is consistent with XRD patterns reported in Figure 10.7.

162

Figure 10.7: X-ray diffraction patterns for post-reaction Ni-Al2O3 catalysts

163

Figure 10.8: Raman spectra of calcined and post-reaction Ni-Al2O3 catalysts

164

10.3.7 TEM study on post-reaction samples Transmission electron microscope study was conducted to further examine the morphology of carbon deposited on Ni-Al2O3 catalysts after 28 hrs steam reforming of propane at 500C. As shown in Figure 10.9, the carbon deposition occurs on both catalysts. Over the impregnated post-reaction sample, carbon deposition is exclusively in the form of filaments, with Ni particles at the tip of the filaments (Figure 10.9 (b)). Filaments are long and solid. It is also noticed that these filaments are of different diameters, ranging from 100-150 nm to 10-20 nm. Because the size of Ni particles determines the carbon filament diameter[150, 207], the carbon filament distribution is suggestive of a nonuniform distribution of Ni particle sizes. There are two possible reasons that could lead to a non-uniform size distribution. The impregnation method used for the preparation may generate different dimensions of Ni ensembles. Additionally, Ni particles on impregnated catalysts tend to aggregate and form larger particles under reduction and reaction conditions. The presence of large Ni particles accelerates the coke formation reaction, as evidenced by the heavily carboncovered catalysts from the TEM images. For the sol-gel post-reaction sample, fiber growth is markedly suppressed. Although there are filaments, as shown in Figure 10.9 (c), they are significantly shorter. The carbon filament diameter is fairly uniform (10-20 nm). No large Ni ensembles are seen. This is in agreement with the high Ni dispersion in sol-gel preparation. Filaments are hollow in the center, which differs from the solid 165

carbon filament on impregnated catalysts. Another interesting phenomenon is that the Ni particle shapes at the tip of filaments are different. For impregnated catalyst, it is either spherical or diamond-shaped (Figure 10.9 (b)). On postreaction sol-gel samples, pear-shaped Ni particles are observed (Figure 10.9 (c)). For the sol-gel catalyst, part of the sample is free of carbon filaments and the Ni particles appear intact, remaining on the support, as shown in Figure 10.9 (d)). The carbon deposits appear to be mostly amorphous. It is reported that carbon is formed on the surface of metal particle through a successive process of formation, diffusion and dissolution[129]. The rate determining step for whisker carbon formation is the diffusion of carbon through a metal particle[209]. The strong chemical bonding between metal and support in sol-gel samples creates resistance for this procedure. The coking model in the following section further illustrates this.

166

(a)

(b)

(c)

(d)

Figure 10.9: TEM images of post-reaction Ni-Al2O3 catalysts after propane steam reforming for 28 hours: (a) and (b) post-reaction impregnation catalyst; (c) and (d) post-reaction sol-gel catalyst

167

10.3.8 Coking models of propane steam reforming

Combining the surface and structural examination on Ni-Al2O3 catalysts prepared by the sol-gel and impregnation techniques, reaction tests, postreaction sample characterization and earlier studies by our group[149, 150], a simplified coking model is proposed for propane steam reforming over Ni-Al2O3 catalysts (Figure 10.10). During the steam reforming process, propane initially cracks into CHx fragments and dissociatively adsorbs on the surface of metallic Ni particles. Subsequently, the adsorbed CHx fragments react with oxygencontaining surface species to produce CO and H2. The CHx fragments could further decompose to carbon and dissolve into Ni particles. With increased carbon formation on the Ni surface, the concentration gradient induces more carbon to diffuse through the Ni particle towards the Ni-support interface, where the carbon interacts with the Al2O3 surface and accumulates, leading to carbon filament growth. The growth of carbon filament could break the attachment between metal and the support and lift the metal particle up. The whisker carbon results in a significant expansion of the catalyst bed, severe operation problems and substantial activity loss. As stated earlier, carbon growth rate is largely limited by the carbon diffusion step. For catalysts prepared by impregnation method, Ni particles weakly attach to the support, which makes it easier for the particles to detach from Al2O3 surfaces. For catalysts prepared by sol-gel technique, Ni particles bond strongly with the support and part of the particles may be anchored into the 168

Al2O3 matrix. The appearance of pear-shaped Ni particles could be a result of the deformation of Ni particles during carbon filament growth, as proposed by Snoeck[210, 211]. The chemical bonding between Ni particles and the support prevents Ni particles from leaving the support, thus retarding the formation of carbon filament. The hollow filaments originating from the pear shape Ni particles are generated as a result of different diffusion distances across the particle[212]. Part of the coking on the sol-gel sample may also be through an extrusion process where carbon, instead of dissolving and diffusing into the metal particle, extrudes over it, thus leaving the metal particle intact on the surface. This carbon appears to be mostly in amorphous form.

169

(a)

(b)

Figure 10.10: Proposed coking models of propane steam reforming over reduced Ni-Al2O3 catalysts: (a) impregnation catalyst; (b) sol-gel catalyst

170

10.4 Conclusions

Compared with conventional impregnation methods, one pot sol-gel preparation yields Ni-Al2O3 catalysts with highly dispersed Ni particles on the surface and a strong metal-support interaction. Ni crystallite size over the sol-gel sample is relatively small and thermally stable under reaction conditions, which suppresses the carbon filament formation reaction. The chemical binding between metal and support on sol-gel catalysts minimizes Ni sintering at high temperature and protects Ni particles from lifting up from the surface during carbon filament growth. Therefore, a much more stable catalytic performance was achieved with sol-gel catalytic system.

171

CHAPTER 11

INVESTIGATION OF BIO-ETHANOL STEAM REFORMING OVER COBALTBASED CATALYSTS

11.1 Overview of bio-ethanol steam reforming over cobalt-based catalysts

Catalytic ethanol steam reforming has been a very popular academia research topic in recent years and a large amount of literatures can be found in this area. Ni, Cu and supported noble metals, like Rh, Pd and Pt are prevalent active metals for this process[124]. Although noble metals have shown remarkable activity in 500600oC range under high space velocity operation[172, 177, 179, 213], their applications are limited by prohibitively high cost. As a less expensive alternative, cobalt-based catalysts are highly reactive in breaking C-C bond into single C containing components like CO, CO2 or CH4 and have exhibited superior ethanol steam reforming performance[162, 165, 214]. The choice of support materials for Co-based catalysts is an important concern in the steam reforming studies, which has been discussed in the literature section to influence catalytic activity, selectivity and stability during ethanol steam reforming reaction. Cobalt crystallite size, reducibility and 172

chemical nature, which are important variables during reaction, are directly related to the nature of the support and interactions between cobalt and the support. Llorca et al. examined a series of supports for ethanol steam reforming, including MgO, Al2O3, SiO2, TiO2, V2O5, ZnO, La2O3, CeO2, and Sm2O3[163]. Comparable ethanol conversions were observed from cobalt catalysts supported on these oxides. However, the product distribution was distinctly different. Although conventional steam reforming catalysts typically are comprised of active metals (Ni or Co) dispersed on a metal oxide support, the support itself may play a role in some of the intermediate steps during ethanol reforming reactions[157, 159, 162]. The support can provide sites for reactant adsorption. It may catalyze side reactions as well, including ethanol dehydration to ethylene (acidic support)[186] and ethanol hydrogenation to acetaldehyde (basic support)[190]. Another important aspect is influence of the support on the reducibility and ultimate dispersion of active metals and ultimately on steam reforming activity. In many supported cobalt systems, especially those at higher cobalt loadings, all of the cobalt cannot be reduced to the metallic state at reasonable temperatures. It was revealed in Jacobs et al.s studies[168] that strong interactions between Co and the support make reduction of Co difficult during catalyst activation. The type of precursors and support materials also play a role in determining metal dispersion, which is revealed from Kraum and Baerns work[215]. Ho and Su[216] prepared supported Co catalysts using different solvents to dissolve the Co precursor and found that use of ethanol as solvent 173

yields an uniformly distributed Co particles on the surface and metal sintering at high temperatures is effectively inhibited. This was attributed to the presence of surface ethoxy groups from ethanol impregnation, which possibly hinders the aggregation of Co3O4 during cobalt precursor decomposition. Similar solvent effects were observed from Tsubaki et al.s work[217] that solvents can influence interaction between Co particles and the support during preparation and this leads to different particle sizes, reaction performance and product distributions. Both nitrate and carbonyl cobalt precursors were employed in Llorca et al.s studies[165] and Co2(CO)8 was evidenced to yield a catalyst that is highly stable and selective for CO-free hydrogen production at low reaction temperature (623 K). Kaddouri and Mazzocchia[218] employed both impregnation and sol-gel methods for catalyst preparation and concluded that metal-support interaction is closely dependent on catalyst preparation, which yields difference in selectivity and product distributions during ethanol steam reforming. In this study, we examined Co catalysts impregnated on different supports including alumina, titania and zirconia. Our research is concentrated on interactions between metal and supports, nature of active sites, and surface intermediates during reaction. An incipient wetness impregnation method was used for catalyst preparation. Catalysts were tested for their activity in ethanol steam reforming using an in-house-built reactor system. Characterization was performed using TGA, H2 chemisorption, temperature-programmed reaction (TPRxn), laser Raman spectroscopy, and DRIFTS. Initial studies on cobalt-based catalysts supported on -Al2O3, TiO2 and ZrO2 have shown promising results for 174

ethanol steam reforming. As determined by H2 chemisorption, ethanol conversion is found to correlate closely with metal dispersion and hence, the metallic Co sites. The product distribution, on the other hand, is determined by a complex network of competing reactions, including ethanol steam reforming, methanation, WGS, dehydration, and dehydrogenation. Among the supports studied, zirconia is shown to provide the highest metal dispersion and the highest H2 yield. H2 yields as high as 92% (5.5 mol of H2 per mole of ethanol fed) are achieved over a 10% Co/ZrO2 catalyst at 550oC.

Experimental details, results and discussions have been presented in paper titled

Investigation of bio-ethanol steam reforming over cobalt-based catalysts By Hua Song, Lingzhi Zhang, Rick B. Watson, Drew Braden, Umit S. Ozkan The paper is published in Catalysis Today 129 (2007) 346-354

175

11.2 Overview of effect of synthesis parameters on the catalytic activity of CoZrO2 for ethanol steam reforming

As illustrated from our previous work[219] and literature studies, Co catalysts demonstrate high activity for ethanol steam reforming. However, many components may affect catalytic performance of Co catalysts. Among them, parameters during catalyst preparation and conditions during pretreatment step are known to have a great influence on the nature of cobalt and cobalt oxide species formed on the surface. Ruckenstein et al.[220] studied effects of preparation method and heat treatment conditions on Co-MgO reforming catalysts. Calcination temperature and time can lead to difference in solid solution formation and degree of reduction, which are closely relevant with ethanol reforming activity. Goodwin et al.[221] examined calcination and reduction conditions on Ru-promoted Co-Al2O3 catalyst and it was revealed that calcination temperature can influence surface active site density and catalyst reducibility. Enache et al.[222, 223] reported in their work that thermal treatments play an important role in formation of metal-support interaction and crystallinity of Co species on the surface. The nature of active sites in Co-based catalysts is another area of interest. Co is in Co3O4 phase in calcined catalysts and reduction of Co3O4 proceeds via two steps with partial reduction to CoO first and then to metallic Co. Both Ticianelli et al.[160, 161] and Freni et al.[224] identified metallic Co as the active sites for steam reforming of ethanol and reoxidation of metallic Co is validated to 176

be the cause of deactivation. In Llorca et al.s studies[164], it was suggested that both CoO and metallic Co are possibly involved in the ethanol reforming reaction. In this study, effect of synthesis parameters on ethanol steam reforming activity were investigated in incipient wetness impregnation prepared ZrO2supported cobalt catalytic system. Parameters examined include calcination temperature, reduction temperature, and reduction time. Characterization was performed using BET surface area measurement, H2 chemisorption, TPR, temperature programmed oxidation, XRD, XPS, temperature programmed calcination, thermogravimetric analysis, ethanol pulse chemisorption, and temperature programmed reaction techniques. It was found that under different treatments, physical and chemical properties of catalysts vary significantly, affecting the catalytic performance in ethanol steam reforming. Cobalt nitrate precursor was used during catalyst preparation and after calcination, nitrate species transformed into a Co3O4 phase. Co3O4 was first reduced into CoO and then to metallic Co. It was revealed that the maximum in metallic Co area correlates with a series of competing processes, such as transformation from a nitrate to an oxide phase and onset of crystallinity versus reaction with the support at higher calcination temperatures, reduction to metallic state versus sintering at higher reduction temperatures. It was indicated from our experimental results that the maximum in metallic Co area coincides with the maxima in both ethanol adsorption capacity and H2 yield in the ethanol steam reforming reaction, suggesting a strong correlation between metallic Co sites and ethanol reforming activity. Metallic Co could be the active sites for ethanol steam reforming. 177

Experimental details, results and discussions have been presented in paper titled

Effect of synthesis parameters on the catalytic activity of CoZrO2 for bioethanol steam reforming By Hua Song, Lingzhi Zhang and Umit S. Ozkan The paper is published in Green Chemistry, 2007, 9, 686-694

178

CHAPTER 12

STUDIES ON SOL-GEL PREPARED COBALT SUPPORTED ON ZIRCONIA CATALYSTS FOR ETHANOL STEAM REFORMING

12.1 Overview of sol-gel preparation studies

Incipient wetness impregnation (IWI) prepared Co-ZrO2 catalyst have exhibited promising ethanol conversion and hydrogen yield from initial ethanol steam reforming reaction testing reported in our publication[219]. However, long term stability testing conducted over Co-ZrO2 catalysts in our lab revealed serious coking. From literature studies, coking is a primary cause for catalyst deactivation in cobalt based catalysts. It was postulated from Llorca et al.s work[162] that the growth of carbon filament was accompanied by the detachment of metallic cobalt particles from the support. The metallic cobalt particles that catalyze the carbon deposition reaction displayed a much bigger size compared with those in contact with the support. In addition to literature findings, our work in Ni-Al2O3 propane steam reforming catalysts has revealed that sol-gel preparation provides a strong interaction between the active Ni metals and Al2O3 support, which produces small and uniformly distributed Ni 179

particles on the support. These small Ni ensembles on the surface suppress carbon formation and improve catalyst durability during long term time on stream testing. This motives us to employ the sol-gel technique in Co-ZrO2 synthesis for ethanol steam reforming reactions.

12.2 Experimental procedures

12.2.1 Catalyst preparation

Catalysts were prepared using two different techniques: IWI (incipient wetness impregnation) and sol-gel. For IWI preparation, ZrO2 support was purchased from Saint Gobain. The support pellets were ground and then sifted through a 100-150 mesh. The sifted supports were then calcined for 3 hrs under air at 500oC before impregnation. Cobalt precursor was from cobalt (II) nitrate hexahydrate (Aldrich, 99.999%). The surface area and pore volume of the ZrO2 support is 31 m2/g and 0.21 cm3/g respectively. During the IWI synthesis, the amount of dimineralized water used to dissolve cobalt nitrate was determined by support pore volume and impregnation times. For each impregnation, the amount of cobalt nitrate aqueous solution was equivalent to the pore volume of the support. A repeated impregnation and drying in an oven overnight at approximately 95oC procedure was used. The resulting samples were calcined at 400oC for 3 hrs under air and stored in desiccators for use.

180

In sol-gel preparation, the cobalt (II) nitrate (Aldrich, 99.999%) is used as the cobalt precursor and zirconium propoxide (C5H8O2)4Zr, Aldrich, 70wt% solution in 1-propanol) as the ZrO2 precursor. A 2.5 times of water amount that can completely hydrolyze zirconium propoxide was used to dissolve the cobalt precursor. During the synthesis, the cobalt nitrate aqueous solution was added dropwise into the zirconium propoxide under stirring. The hydrolysis and gelling was conducted in a rotary evaporator bath at 65C and rotated for 30 min. The gel produced is then dried in the oven at 80C overnight. Finally the powders are calcined at different temperatures for experimental use.

12.2.2 Catalyst characterization

For TPR experiment, catalyst sample was loaded in a 1/4 in. O.D. quartz U-tube reactor. The catalyst was then re-calcined under N2 at 400C for 1 hr followed by cooling to room temperature under N2. The reduction was performed with 5% H2/N2. The reactor temperature was raised using a ramp rate of 10C/min to 800C. Effluents from the reactor were passed through a silica gel water trap before it reached the detector to remove moisture formed during reduction. H2 consumption was measured continuously as a function of sample temperature using TCD connected to a data-acquisition system. A Bruker D8 Advance X-Ray diffractometer was used to collect in situ Xray diffraction patterns during reduction of catalysts. The diffractometer is equipped with an atmosphere and temperature control stage and using Cu K 181

radiation ( = 1.542 ) operated at 40 kV and 50 mA. The powder diffraction patterns were recorded in the 2 range from 20o to 90o. Reduction was performed in-situ under 5% H2/N2 gas flow using a linear temperature-program between isothermal steps (100C). The catalysts were kept at isothermal steps for 0.5 hr before data collection and the ramp rate between the isothermal steps was 10C/min. For selected catalysts, diffraction patterns were collected during sample cooling stage as well. A 9-sample holder accessory was employed to collect diffraction patterns when no treatment was needed. Ethanol and water temperature programmed reaction (TPRxn)

experiments were performed on IWI and SG prepared Co-ZrO2 catalysts. Samples were first treated at 400oC for 30 min in He, followed by a reduction step with 5% H2/He for 2 hrs at 350oC for IWI samples or 600oC for SG samples. Reactors were cooled down to room temperature and the reactant mixture was introduced by water and ethanol double bubblers at a H2O/C2H5OH molar ratio 10 (ethanol concentration is around 0.63%) and gas hourly space velocity of 10,000 h-1. Temperature was increased at 10oC/min up to 650oC. Effluents were monitored by Cirrus MKS mass spectrometer using scanning mode (m/z=1~60). In another series of experiments, with all the other experimental settings the same, catalysts were kept on stream for steady state reactions at 450oC for 15 hrs. The post-reaction catalysts from the above experiments were saved for TPO measurement. Catalysts were treated at 110oC in He for 1 hr and then cooled to room temperature. 5% O2/He was flowed through the reactor at room 182

temperature. TPO was conducted with a ramping rate of 10oC/min until 800oC. Effluents from the reactor were monitored by Thermo Finnigan Trace GC ultraTrace DSQ MS. A Thermo 6700 FT-IR Spectrometer equipped with a DRIFTS cell, an MCT detector and a KBr beam splitter was used. DRIFT spectra were collected with a 500-scan data acquisition at a resolution of 4 cm-1 in a controlled gas atmosphere and temperature using an environmental chamber with Zn-Se windows. During the experiments, catalysts were first pretreated in He at 400oC for 30 min. This was followed by a reduction procedure with 5% H2/He at 350oC for 2 hrs. After the chamber was flushed with He at 350oC for 30 min, ethanol and water mixture (molar ratio H2O/C2H5OH = 10) adsorption was carried out at room temperature for 1 hr. Another flushing with He at room temperature for 1 hr was followed. Finally, a temperature programmed desorption was performed under He flow. The IR spectra were collected at different temperatures: 30, 100, 200, 300, 400 and 500oC.

12.3 Results and discussion

12.3.1 Temperature programmed reduction

TPR was conducted to examine 10 wt% Co-ZrO2 catalysts reducibilities, which helps understanding of how the Co metal interacts with ZrO2 support. As indicated from the results, IWI-Co-ZrO2 catalyst exhibited clearly two reduction 183

features. From our previous publication[225], the low temperature peak around 300oC was associated with reduction of Co3+ to Co2+ species. Further reduction of Co2+ yielded metallic Co, which can be evidenced by the second reduction peak around 500oC in Figure 12.1. For catalysts synthesized with sol-gel technique, catalysts calcined at 300oC or 400oC displayed two similar reduction features as IWI catalysts before 500oC. However, the lower temperature reduction peak was displaced to around 350oC, implying cobalt species that are more difficult to reduce. The second reduction peak for these two sol-gel catalysts also appeared around 500oC. Additionally, a weak peak around 580oC was detected, which could be correlated with reduction of Co species that has stronger interaction with the support. When sol-gel Co-ZrO2 catalysts were calcined at 500oC or 600oC, reduction features before 500oC became less noticeable, particularly that the peak around 500oC disappeared. For all sol-gel prepared catalysts, by the end of the TPR, H2 consumption was still observed, indicating that reducible species in the sample still exists by 800oC. It can be deduced from the TPR results that sol-gel preparation creates a stronger interaction between the metal and the support. Co species in the sample requires higher temperatures to reduce. It is possible that due to the strong interaction, a portion of the Co species could not be reduced under the reducing conditions employed in our work.

184

Figure 12.1: Temperature programmed reduction profiles for Co-ZrO2 catalysts: IWI catalyst and sol-gel catalysts calcined at different temperatures

185

12.3.2 X-ray diffraction patterns of Co-ZrO2 catalysts

Diffraction patterns of Co-ZrO2 catalysts prepared by IWI and sol-gel method were shown in Figure 12.2. For IWI-Co-ZrO2 catalyst, monoclinic ZrO2 (ICDD #: 02-0464) diffraction lines were observed. Co3O4 phase (ICDD #: 021079) was detected as well. However, no cobalt oxide crystal phases were seen on sol-gel prepared catalysts, which is suggestive of highly dispersed Co species in the catalyst structure for sol-gel catalysts. Sol-gel Co-ZrO2 catalysts calcined at 300oC and 400oC demonstrate an amorphous nature. At higher calcination temperatures, cubic fluorite ZrO2 crystallites were formed in the catalysts. In situ XRD was performed to study crystal phase evolution during catalyst reduction process (Figure 12.3 and Figure 12.4). 5% H2/N2 was used as the reducing agent. Starting from 30oC, XRD patterns were collected every 100oC until 730oC. For Co-ZrO2 calcined at 400oC, crystalline phase appeared when reduction temperature was as high as over 400oC. The diffraction lines became more intense as temperature further increased. For Co-ZrO2 calcined at 500oC, cubic ZrO2 was the dominant phase at room temperature. As reduction proceeded, the crystallites grew larger in size, evidenced from stronger signals at higher temperatures. No new phases were detected by the end of the experiment. Cubic ZrO2 is not thermally stable and converts to tetragonal ZrO2 and finally to the most stable monoclinic form as temperature increases. However, the cubic or tetragonal structures can be stabilized by incorporation of appropriate dopants[226]. The presence of dopants can delay the phase 186

transition so that cubic ZrO2 can maintain its structure at relatively higher temperature. To confirm function of Co in stabilizing ZrO2 structure, ZrO2 support was prepared by sol-gel technique. XRD patterns were collected on ZrO2 support calcined at 400oC and 600oC (Figure 12.5). Stable cubic ZrO2 phase was observed at 400oC calcination temperature. However, for ZrO2 support calcined at 600oC, both monoclinic and cubic phases were detected.

187

(a)

(b) Figure 12.2: X-ray diffraction patterns of Co-ZrO2 catalysts: (a) IWI-Co-ZrO2; (b) SG-Co-ZrO2 calcined at different temperatures

188

Figure 12.3: In situ X-ray diffraction for sol-gel Co-ZrO2 catalyst calcined at 400oC during 5% H2/N2 reduction

189

Figure 12.4: In situ X-ray diffraction for sol-gel Co-ZrO2 catalyst calcined at 500oC during 5% H2/N2 reduction

190

Figure 12.5: X-ray diffraction patterns of ZrO2 supports calcined at 400oC and 600oC

191

12.3.3 Temperature programmed reaction

IWI and sol-gel prepared Co-ZrO2 catalysts were exposed to the same reactant stream described in the experimental section at 450oC for 15 hrs. Postreaction samples were collected for temperature programmed oxidation experiments to measure the amount of carbon deposition. A mass spectrometer was used to monitor the eluting species during oxidation. m/z=44 signal is caused by CO2 formation from carbon combustion. Two CO2 peaks can be observed during the TPO process (Figure 12.6). As discussed in the propane steam reforming chapter, the low temperature peak (220oC for IWI catalyst and 280oC for sol-gel catalyst) can be attributed to combustion of highly reactive monoatomic carbon species deposited on Co surfaces (Type I). Another peak, around 430oC for both catalysts, can be assigned to filamentous carbon (Type II), which is typically stable and oxidized at higher temperatures[150, 207, 208]. The peak intensity between the two catalysts is in stark difference. Markedly more carbon deposition was measured on IWI-Co-ZrO2 post-reaction catalyst whereas much less carbon was formed during ethanol steam reforming reaction for sol-gel prepared Co-ZrO2. It appears that IWI prepared catalysts have the tendency to catalyze more coke formation reaction, which could possibly be correlated with dispersion and Co active metal particle size as inferred from our work on Ni-Al2O3 catalytic system. In another set of experiments, under the same experimental conditions, temperature programmed reaction were performed on IWI and sol-gel prepared 192

Figure 12.6: Temperature programmed oxidation for post-reaction Co-ZrO2 catalysts

193

Co-ZrO2 catalysts to obtain an understanding of how reactants interact with catalysts at different temperatures and lead to variation in product distribution. As shown in Figure 12.7 and Figure 12.8, although with the same composition, IWI Co-ZrO2 catalysts are more active in ethanol steam reforming, evidenced from much higher H2 and CO2 signals. It should also be noted that for IWI catalysts, H2 and CO2 formation started to take off at a lower temperature. In terms of production distribution besides H2 and CO2, above 450oC, IWI catalysts are characterized by gas products, including CO and CH4. CH4 steam reforming began to take place at higher temperatures. Water-gas shift reaction was favored at higher temperature as well. On the other hand, for sol-gel Co-ZrO2 catalysts, liquid products including acetone, ethylene and propylene were still observed above 450oC. However, time on stream testing showed a continuous activity decline for IWI catalysts whereas for sol-gel catalysts, activity was relatively stable in the testing period. This can be further confirmed from the pressure gauge reading at the end of the experiments. For sol-gel catalyst, system pressure was 5 psi. However, IWI catalytic system reached a pressure of 15 psi after 15 hrs time on stream. As indicated from TPR, sol-gel catalysts have less reducible Co species. During ethanol steam reforming reaction, Co metal provides the active sites for C-C cracking. IWI Co-ZrO2 catalysts have more gas products because Co sites help decomposition of ethanol into small molecules. On the other hand, sol-gel prepared catalysts lead to formation of liquid products. Another feature of IWI catalysts is that Co species tend to aggregate and form larger particles on the 194

surface after treatment. Those large particles catalyze the coking reaction, causing carbon deposition on catalyst surfaces. However, in sol-gel catalytic system, highly dispersed Co metals greatly suppress coking. This can be used to explain the TPO results shown in Figure 12.6.

195

Figure 12.7: Temperature programmed reaction for IWI Co-ZrO2 catalyst

196

Figure 12.8: Temperature programmed reaction for sol-gel Co-ZrO2 catalyst

197

12.3.4 Diffuse reflectance infrared Fourier transform spectroscopy

DRIFT spectra were collected during temperature programmed desorption after Co-ZrO2 catalysts were exposed to ethanol and water reactants. Results are presented in Figure 12.9. Strong water and ethanol adsorption bands were observed at 30oC. Water adsorption is characterized by broad O-H stretching band around 3700-3100 cm-1 as well as the band around 1660 cm-1 from molecularly adsorbed water[149]. Adsorption of ethanol yields C-H stretching bands at 2980, 2960 and 2890 cm-1. The band at 1160 cm-1 is attributed to C-C vibration[227]. Formation of monodentate and bidentate ethoxy species from disassociative adsorption of ethanol are evidenced by bands located at 1450, 1380 cm-1 (CH3- bending) and 1070 cm-1 (CCO stretching)[228]. As temperature went up, water adsorption signal became less intense. The adsorbed ethoxy species from ethanol adsorption began to convert into surface acetate species, identified by bands at 1570, 1450 and 1360 cm-1[227]. Linearly adsorbed CO2 was detected at 300oC for IWI Co-ZrO2. Sol-gel catalysts exhibited similar IR bands. However, consistent with temperature programmed reaction results, surface reaction for sol-gel catalysts takes off at a higher temperature. By 500oC, acetate species still remain on the surface whereas for IWI catalysts, no acetate species can be detected, implying complete conversion under this condition.

198

(a)

(b)

Figure 12.9: DRIFT spectra during temperature programmed desorption: (a) IWI Co-ZrO2 catalysts; (b) sol-gel Co-ZrO2 catalysts

199

12.4 Conclusions

Findings from studies on Co-ZrO2 catalysts prepared by both IWI and solgel methods can be summarized as follows. (a). Sol-gel preparation generates a strong bonding between Co metals and the support. Reducible Co species is much less in sol-gel catalysts. (b). Cubic crystal ZrO2 phase was detected from sol-gel Co-ZrO2 catalysts, instead of monoclinic ZrO2 phase used for IWI catalyst preparation. This crystal phase was stabilized by the presence of Co in the structure. (c). Product distribution during ethanol steam reforming is largely dependent on catalyst synthesis technique. More Co sites on IWI catalyst surface facilitate C-C breakage reaction and more gas products are formed. Although more liquid byproducts were produced for sol-gel catalytic system, sol-gel catalysts demonstrate higher stability from time on stream testing. Therefore, efforts could be focused on improvement of ethanol steam reforming activity while maintaining stability for sol-gel catalysts.

200

CHAPTER 13

MODIFICATION OF COBALT SUPPORTED ON ZIRCONIA CATALYSTS FOR ETHANOL STEAM REFORMING

13.1 Overview of Co-ZrO2 catalysts modification for ethanol steam reforming

As concluded from work performed in the previous chapter, sol-gel prepared catalysts showed higher stability during ethanol steam reforming. Co loading amount was examined in order to further improve catalyst activity. To modify surface properties of the catalysts, CeO2 was introduced to Co-ZrO2 formulation. The mixed oxide catalyst exhibited higher activity and stability. Catalyst surface basicity/acidity was measured using probe molecules.

13.2 Experimental procedures

13.2.1 Catalyst preparation

Chemicals used are purchased from Aldrich. Sol-gel method has been described in the previous chapter for synthesis of Co-ZrO2 catalysts. In this 201

chapter, Co-ZrO2 catalysts with different Co loadings (5%, 10%, 20% and 30% by weight) by sol-gel technique were prepared using the same sol-gel preparation in an effort to further improve catalyst performance. Catalysts were calcined at 400oC for 3 hrs in air before use. 20%Co-40%CeO2-40%ZrO2 was prepared using sol-gel technique. Zirconium propoxide was used as Zr precursor. Other precursors used were Ce(NO3)3.6H2O and Co(NO3)2.6H2O. A certain amount of zirconium propoxide (C5H8O2)4Zr, Aldrich, 70wt% solution in 1-propanol) was mixed with aqueous solution of ceria and cobalt nitrates. The mixture was kept at 65oC in a water bath and stirred at a constant speed of 90 rpm for 30 min before transferred to oven for overnight drying at 80oC. Catalysts were calcined in air at 400oC for 3 hrs. Bare supports were prepared for comparison. ZrO2 was prepared by sol-gel method starting from zirconium propoxide using the same procedures as CoZrO2 samples. CeO2 was obtained by calcining Ce(NO3)3.6H2O in air at 400oC for 3 hrs.

13.2.2 Catalyst characterization

X-ray diffraction (XRD) patterns were acquired from 20 to 90 at a step width of 0.0144 using Bruker D8 Advance X-Ray Diffractometer equipped with a CuK source. Detailed data collection procedures are the same as described in the previous chapter.

202

Two sets of experiments were conducted using diffuse reflectance Fourier transform spectroscopy (DRIFTS) technique: ethanol and water temperature programmed desorption (TPD) and ammonia TPD. The instrument is a Thermo NICOLET 6700 FTIR spectrometer equipped with a liquid-nitrogen-cooled MCT detector and a KBr beam splitter. For ethanol and water TPD experiments, sample was first pretreated in situ under helium at 400C for 30 min. Reduction was conducted under 5% H2/He at 600C for 2 hrs. This was followed by a purging step at 600oC with He for 15 min. After background spectra were taken in He at desired temperatures, EtOH and H2O vapors generated by a two-bubbler system with a molar ratio 1:10 were introduced into the chamber at room temperature for 1 hr. After the adsorption, the system was purged in He at room temperature for 1 hr. Sample spectra were collected at certain temperatures during desorption. For NH3 TPD experiments, after experiencing the same pretreatment, reduction and purging step as described above, catalysts was exposed to NH3 at 100oC for 30 min. 100oC was chosen as the adsorption temperature to minimize physical interaction between NH3 and catalyst surface. After that, still at 100oC, inert gas He was used to purge the line for 30 min. A temperature programmed desorption was run and spectra were taken with 50oC interval starting from 100oC until 500oC in He. For CO2 pulse chemisorption, 100 mg catalyst was loaded into a in. Utube reactor. Catalyst was first pretreated under He flow at 400oC for 30 min followed by a reduction step with 5% H2/He at 600oC for 2 hrs. After that, He was 203

used to purge the system at 400oC for 1 hr and then the reactor was cooled down in He to room temperature. CO2 pulse chemisorption was performed at 80oC with a 0.15 mL sample loop and pure CO2 gas. For each pulse, pulse time was kept at 1 min and the following pulse would not be injected until a flat baseline was obtained. The effluent was continuously monitored by a DSQ mass spectrometer. The following fragments were recorded during experiment: m/z=44, 28, 16, 32 and 12. Pulses were stopped when there were no changes in the eluting species peak area, which implied that surface adsorption had been saturated.

13.2.3 Reaction testing

Catalysts were evaluated for their ethanol steam reforming performance using a fixed bed inch stainless steel reactor. Ethanol and water vapor was generated by a heated vaporizer. The liquid mixture was delivered by a HPLC pump. Ethanol and water molar ratio was kept at 1:10. Helium was used as carrier gas for the vaporizer. Nitrogen was used as internal standard. During the reaction testing, catalysts were first pretreated at 400oC for 30 min, which is the same temperature as catalyst calcination. Catalysts were then reduced in 5% H2/He at 600oC for 2 hrs. After the reduction step, the reactor was purged with He at 400oC for 1 hr. Catalytic performances were tested in the temperature range of 300550oC, in 50oC increments. Catalysts were held at each temperature for 2 hrs. Feed and product streams can be analyzed online by a 204

Shimadzu Scientific 2010 gas chromatography (GC). Separation was achieved using three columns. A Carboxen column and a 5 molecular sieve column were connected in tandem to a pulse discharge helium ionization detector (PDHID). A 30 m long Q-Plot column was connected with a methanizer and a flame ionization detector (FID) for measurement carbon containing species. He was used as the carrier gas for all columns.

13.3 Results and discussion

13.3.1 Sol-gel Co-ZrO2 catalysts with different Co loadings

As revealed from our initial studies on 10%Co-ZrO2 catalysts, sol-gel prepared 10%Co-ZrO2 catalysts exhibited higher stability compared with IWI 10%Co-ZrO2 catalysts. However, catalytic activity needs to be further improved. Because Co is the active component for ethanol steam reforming, effect of Co doping amount was examined in this work. Co loading amount varied, including 5%, 10%, 20% and 30% by weight. In situ X-ray diffraction during catalyst reduction was employed to monitor crystal phases present in the sample and phase changes with increasing temperature. Two representative catalysts (with 10% and 30% Co loadings) XRD patterns were shown in Figure 13.1 and Figure 13.2. For 10%Co-ZrO2 (shown in Figure 13.1), very broad diffraction lines were detected before 500oC. At 500oC, Cubic ZrO2 together with metallic Co phases appeared. Metallic Co diffraction 205

Figure 13.1: In situ X-ray diffraction for sol-gel 10%Co-ZrO2 catalysts during reduction with 5% H2/N2

206

Figure 13.2: In situ X-ray diffraction for sol-gel 30%Co-ZrO2 catalysts during reduction with 5% H2/N2

207

lines grew stronger as temperature further increased. At 800oC, cubic ZrO2 started to transform into more stable tetragonal ZrO2 phase. A small amount of monoclinic ZrO2 was visible as well. The diffraction lines from monoclinic ZrO2 became more intense at 900oC. After sample was cooled from 900oC to 100oC, XRD pattern was collected again. It was observed that the dominant crystal phase was monoclinic ZrO2. Tetragonal ZrO2 present was still visible, but with much less intensity compared with 900oC. This again validates that cubic ZrO2 would transform into thermally more stable tetragonal phase first and eventually to monoclinic ZrO2. For XRD patterns of 30%Co-ZrO2 catalyst shown in Figure 13.2, very clear Co3O4 and cubic ZrO2 diffraction lines were observed even at 100oC. At 300oC, Co3O4 began to reduce to CoO. Metallic Co appeared at 400oC as a result of further reduction of CoO. This is another evidence that reduction of Co3O4 proceeds via two steps, first to CoO and then to Co, as discussed in our publication[225]. Similarly as 10%Co-ZrO2 catalyst, cubic ZrO2 went through the transition to tetragonal and monoclinic phase starting from 700oC. By 900oC, it is a combination of tetragonal and monoclinic ZrO2 phases. Co-ZrO2 catalysts with different Co loadings were evaluated for their ethanol steam reforming activity. Definitions of ethanol conversion and product yield are as follows.

208

H2 Yield % =

moles of H2 produced 100 6 (moles of ethanol fed)

Carbon containing product A Yield % : = (moles of A produced) ( # of carbon in A) 100 2 (moles of ethaonl fed) moles of ethanol converted 100 moles of ethanol fed

EtOH Conversion % =

From the reaction testing results, higher ethanol conversion was achieved for catalysts with higher Co loading. For 30% Co-ZrO2, 60% ethanol conversion was obtained at 350oC. However, more Co on the surface catalyzed coke formation and even at 400oC, serious coking was observed. Reaction data for sol-gel 20% Co-ZrO2 catalyst were presented in Figure 13.3. Acetaldehyde was the main liquid byproduct. Acetone was detected when the reaction temperature increased to 450oC. By 450oC, ethanol conversion was 70%. However, due to coking, reaction data after 450oC could not be collected.

209

Figure 13.3: Ethanol steam reforming reaction data at different temperatures for sol-gel 20%Co-ZrO2 catalysts

210

13.3.2 Examination of CeO2 to modify ZrO2 support properties and improve activity and stability

As discussed in literature review, support acidity and basicity are important parameters that are directly related with catalyst stability and product distribution during ethanol steam reforming. Acidic surface favors ethanol dehydration to form ethylene, which is a precursor for coking. On the other hand, basic surface tends to promote ethanol dehydrogenation reaction to produce acetaldehyde. Further decomposition of acetaldehyde yields gas products CO and CH4 and would not contribute to coking deposited on the surface. Therefore, it is critical to tailor the surface properties during catalyst preparation. From literature, CeO2, with its basic surface, is an attractive support for ethanol reforming catalysts because of much less coking. In our work, CeO2 was used to adjust ZrO2 properties in an attempt to enhance catalyst stability. 20%Co-40%CeO2-40%ZrO2 catalyst was prepared. CO2 pulse chemisorption was used to probe catalyst basicity. There is extensive information on interaction of CO2 molecule with basic sites on metal oxide surface[229]. CO2 can adsorb strongly on basic sites and interact with them in many different ways. As shown in Figure 13.4, CO2 chemisorption capacities were measured on four samples. ZrO2 and CeO2 supports were included as comparison references. 20%Co-ZrO2 and 20%Co-40%CeO2-40%ZrO2 were samples used for ethanol steam reforming reaction. Figure 13.4 (a) is m/z=44 evolution during CO2 pulses on CeO2 support. The area under each pulse was integrated. Based on the area 211

(a)

(b) Figure 13.4: (a) m/z=44 signal evolution during CO2 pulse chemisorption on CeO2 support; (b) calculated CO2 adsorption amount comparison

212

of the peak when CO2 adsorption was saturated on the surface (which correlates with sample loop volume), all pulses were normalized with respect to the sample loop. With information of sample loop volume, temperature and pressure, CO2 adsorption capacity was determined as shown in Figure 13.4 (b). ZrO2 and CeO2 supports demonstrated similar CO2 adsorption amount, implying similar overall basic sites on the surface. This is consistent as described in literature that ZrO2 has both basic and acidic sites whereas CeO2 only has basic adsorption sites. For two catalyst samples, 20%Co-ZrO2 has very low CO2 adsorption capacity and 20%Co-40%CeO2-40%ZrO2 has the highest CO2 adsorption amount. Enhanced CO2 adsorption capacity on mixed oxides supported Co catalysts is possibly caused by interaction between mixed supports and between Co and the support. To further examine characteristics of ZrO2 and CeO2 support during ethanol steam reforming reaction, ethanol and water TPD experiments were conducted on these two samples. Results are shown in Figure 13.5 and Figure 13.6. The DRIFT spectra were markedly different between them, particularly in the C-O bending region between 1800 and 1000 cm-1. For both supports, broad O-H stretching was observed in the region between 3800 and 3500 cm-1. Those features originate from either molecularly adsorbed ethanol through the formation of hydrogen-bridge bonding with the OH groups of the support or from OH group vibration in H2O molecules. However, for CeO2, after 450oC, very weak OH stretching bands were detected, indicating that CeO2 may be more capable of catalyzing reactions between ethanol and water. As discussed in previous 213

DRIFTS results, the region 3000-2700 cm-1[187] can be assigned to C-H stretching. Ethanol was adsorbed as two distinct forms on the surface. One is the band around 1150 cm-1 for monodentate ethoxide and the other is the band around 1050 cm-1[195] for bidentate ethoxide. Bidentate ethoxide was the major adsorbed ethanol form on ZrO2 surface before 200oC whereas CeO2 support displayed both monodentate and bidentate ethoxides on the surface. This could be attributed to partial reduction of CeO2 during the reduction step that there was not sufficient lattice oxygen in the support to form bidentate ethoxides. For ZrO2 support, bands at 1580, 1470, and 1430 cm-1 are indicative of surface acetate species[194, 196, 228, 230]. For CeO2 support, chemisorbed water band appeared at 1660 cm-1. Bands at 1590, 1550 and 1450 cm-1 show that there might be both acetate and carbonate intermediates co-existent on the surface[229], which could be explained by differences in surface properties including oxygen availability, basic sites distribution and strength. As temperature further increased, acetate species began to transform to carbonate form and by 450oC, carbonate was the major species.

214

Figure 13.5: Ethanol+H2O TPD in situ DRIFTS on ZrO2 support

215

Figure 13.6: Ethanol+H2O TPD in situ DRIFTS on CeO2 support

216

13.3.3 Evaluation of catalysts during ethanol steam reforming and surface acidity measurement

20%Co-40%CeO2-40%ZrO2 catalyst was tested for its ethanol steam reforming activity and results were shown in Figure 13.7. The catalyst was very stable during the testing period. At 300oC and 350oC, acetaldehyde was the main byproduct, which limits H2 yield. At 400oC, acetaldehyde production is significantly reduced. However, a very high acetone yield (>60%) was seen. 100% ethanol conversion was obtained starting from 400oC. From 450oC, no liquid products were detected. Main products include H2 and CO2 with a small amount of CH4 and CO. The performance has been improved compared with 20% Co-ZrO2 catalyst. In addition to the surface basicity measurement in the previous section, surface acidity of the catalysts was measured as well. There are a large variety of probe molecules that can be used for qualitative assessment of acidic sites and their distribution. Among them, pyridine, amine and ammonia are the most common. Ammonia was chosen in this work as the base probe molecule. It is a smaller molecule compared with pyridine and can detect most of the different types of acid sites[231]. Ammonia interacts with acidic protons, electron acceptor sites and hydrogen from neutral or weakly acidic hydroxyls. Therefore, ammonia and metal oxides surface can coordinate through different ways: hydrogen bonding formed by hydrogen from NH3 molecules and oxygen atom from oxides (weakest interaction), Brnsted acid sites with surface donating protons to 217

Figure 13.7: Ethanol steam reforming reaction data at different temperatures for sol-gel 20%Co-40%CeO2-40%ZrO2 catalysts

218

adsorbed NH3, Lewis acid sites with NH3 coordinating to an electron-deficient metal atom from the surface (strongest interaction), a complete transfer of protons from Brnsted acid sites to NH2 or NH formed through ammonia dissociation. DRIFTS experiments were performed to the interaction between NH3 and catalyst surface. Results are presented in Figure 13.8 and Figure 13.9. Evident differences were noticed on the spectra between 20%Co-ZrO2 and 20%Co-40%CeO2-40%ZrO2 catalysts. For 20%Co-ZrO2 catalyst, the band at 1450 and 1500 cm-1 could be due to asymmetric deformation mode of NH4+ ions as a result of ammonia protonation. With increase in temperature, these bands shifted to lower frequencies (1440 and 1480 cm-1 respectively), suggesting that the bonding strength became weaker when temperature increased. The corresponding symmetric mode was located around 1610 cm-1[231], the intensity of which decreased with rising temperatures. The band at 1260 cm-1 was attributed to co-ordinatively bonded ammonia. The bands at 1110 and 1050 cm-1 were assigned to weakly adsorbed NH3 with oxygen from oxide support or from hydroxyl groups attached to the surface. For 20%Co-40%CeO2-40%ZrO2 catalyst, Brnsted acid sites characterized by bands at 1620, 1450 and 1420 cm-1 were much weaker, indicating less acidic surface. This may help explaining higher stability of the catalyst during ethanol steam reforming.

219

Figure 13.8: NH3 TPD using DRIFTS over 20%Co-ZrO2 catalyst

220

Figure 13.9: NH3 TPD using DRIFTS over 20%Co-40%CeO2-40%ZrO2 catalyst

221

13.4 Conclusions

Studies on Co loading amount showed that introduction of more Co component during catalyst preparation is beneficial for ethanol conversion. However, more Co on the surface leads to formation of large Co crystallites, which catalyzes coking formation reaction. Addition of CeO2 into Co-ZrO2 was found to improve both catalyst activity and stability. CO2 and NH3 were used as probe molecules to measure surface basicity and acidity respectively. 20%Co40%CeO2-40%ZrO2 catalyst showed higher surface basicity and less interaction with basic NH3 molecules.

222

CHAPTER 14

CONCLUSIONS AND RECOMMENDATIONS

14.1 Summary of propane steam reforming work

Catalyst preparation methods, including conventional impregnation and one pot sol-gel technique, were studied in Ni-Al2O3 catalytic system for propane steam reforming. Steady state reaction testing showed a higher initial activity for impregnated Ni-Al2O3 catalyst. However, this catalyst suffered from drastic activity loss with time on stream whereas sol-gel Ni-Al2O3 catalyst was relatively stable. A variety of characterization techniques were employed to examine catalyst physical and chemical properties caused by catalyst preparation. From H2 chemisorption measurement, reduced SG-Ni-Al2O3 catalysts demonstrate a lower metallic Ni surface area than reduced IM-Ni-Al2O3 catalysts, implying that the impregnation method may lead to more Ni species staying on the surface instead of entering the bulk as in sol-gel preparation. Another contribution could be variation in extents of reduction for Ni species exposed on the surface. This was supported by XPS results that more spinel NiAl2O4 and less NiO species were detected on sol-gel Ni-Al2O3 catalyst surface. NiAl2O4 species are possibly 223

formed as a result of strong interaction between Ni species and the Al2O3 support. This form of Ni species is difficult to reduce under our reduction condition, which again was confirmed from TPR results. A larger H2 consumption was detected for impregnated catalysts, suggestive of more reducible species in the catalyst. The strong interaction between Ni and the support in sol-gel catalyst decreased the amount of reducible Ni. XRD patterns showed that for both calcined and reduced catalysts, impregnated Ni-Al2O3 catalyst has well defined crystallites whereas sol-gel catalyst displays an amorphous nature, implying that Ni species in sol-gel catalysts is well dispersed and stable after catalyst activation. As revealed from literature studies, Ni crystallite size formed after reduction is directly related to catalyst stability, with large Ni clusters catalyzing coke formation reaction. This is consistent with our observation from reaction testing. Relatively small and thermally stable Ni species over the sol-gel sample greatly suppresses the carbon filament formation reaction. On the other hand, the chemical binding between metal and support on sol-gel catalysts minimizes Ni sintering at high temperature and protects Ni particles from lifting up from the surface during carbon filament growth. Therefore, a much more stable catalytic performance was achieved with sol-gel catalytic system. Post-reaction catalysts were characterized by TPO, XRD, Raman spectroscopy and TEM. Markedly more carbon deposition was observed over impregnated Ni-Al2O3 catalyst. Additionally, carbon formed on impregnated catalyst is highly crystallized in comparison with sol-gel catalyst. 224

14.2 Future work on propane steam reforming

More work can be pursued over Ni-Al2O3 catalytic system to correlate catalyst physical and chemical properties and propane steam reforming catalytic performance. Although impregnated and sol-gel Ni-Al2O3 catalyst properties have been examined by different characterization techniques as summarized, a better understanding on influence of metal-support interaction on Ni particle diameter, Ni dispersion on the surface and how these properties eventually affect catalyst activity, stability and product distribution is still needed. A detailed investigation on coke formation reaction and Ni crystallite size and thermal stability through steady state reaction testing and kinetic studies would be interesting. In addition, more efforts can be put into evaluation of catalyst pretreatment conditions. Mild reducing agents can be employed to minimize sintering of Ni metals during catalyst activation. Catalyst physical and chemical properties under the new reducing conditions can be performed. Catalysts need to be reevaluated and relate the performance with catalyst properties. In a long term, different support materials can be selected to support the active metal. As has been discussed in the literature review, support plays an important role in dispersing the active metal. It may also affect interaction between reactants and catalyst surface and determine the reaction pathway and eventually the product distribution. This is another alternative route to minimize coke deposition during reforming process.

225

14.3 Summary of ethanol steam reforming work

Co-ZrO2 was chosen as our initial catalytic system for ethanol steam reforming work. Co-ZrO2 catalysts were prepared by both IWI and sol-gel methods. It was found that sol-gel preparation generates a strong bonding between Co metals and the support. Reducible Co species is much less in solgel catalysts. XRD results showed cubic ZrO2 crystal phase from sol-gel Co-ZrO2 catalysts, which is different from monoclinic crystal phase for commercial ZrO2 support used for IWI catalyst. Cubic ZrO2 structure may possibly be stabilized by incorporation of Co in the structure. Evaluation of catalyst during ethanol steam reforming reaction showed a higher stability for sol-gel catalysts. However, more liquid byproducts were produced for this system. Later efforts were focused on improvement of ethanol steam reforming activity while maintaining stability for sol-gel catalysts. Examination of Co loading amount indicated that more Co active sites improve ethanol conversion. For 30% Co-ZrO2, 60% ethanol conversion was obtained at 350oC, which makes it promising for low temperature operation. On the other hand, more Co on the surface catalyzed coke formation and serious coking was observed even at 400oC. CeO2 was added to Co-ZrO2 formulation during catalyst preparation in an attempt to modify catalyst surface basic/acidic properties and improve stability. The mixed oxide catalyst exhibited higher activity and stability with negligible liquid product selectivity when temperature was above 450oC. Catalyst surface 226

basicity was measured using CO2 as the probe molecule. The enhanced CO2 adsorption capacity on mixed oxides supported Co catalysts is attributed to possible surface modification by CeO2 and Co-CeO2-ZrO2 catalyst has a more basic surface. Complementary with basicity measurement, NH3 was used to probe surface acidity. Stronger interaction was observed between NH3 and CoZrO2 catalyst due to an acidic surface.

14.4 Future work on ethanol steam reforming

Preliminary work has been conducted to study effect of surface acidity/basicity on catalyst activity, selectivity and stability. More studies are required in order to establish a relationship between surface property and reaction performance. An extensive acidity/basicity measurement needs to be performed on a variety of catalysts with variation in support and cobalt loading amount. DRIFTS can be used to examine how probe molecules interact with different surfaces. Probe molecule adsorption/desorption properties can be further characterized through temperature programmed desorption experiment in conjunction with a mass spectrometer. ZrO2-CeO2 is the only mixed oxide support investigated in this work. Other basic supports, including La2O3, ZnO, MgO, deserve future efforts. Promoters can be incorporated to enhance the performance of Co. In some initial lab testings, 2% Cu dramatically increased 20%Co-ZrO2-CeO2 catalyst stability.

227

REFERENCES

[1] [2] [3] [4]

J.N. Armor, Appl. Catal. A: Gen 176 (1999) 159. M. Laniecki, and M. Ignacik, Catal. Today 116 (2006) 400. EIA and BP statistical review of world energy (2005). S.C. Stultz, and J.B. Kitto, STEAM: Its Generation and Use, 40th ed., Babcock & Wilcox, chapter 17 (1992). S. Natesakhawat, X. Wang, L. Zhang, and U.S. Ozkan, J. Mol. Catal. A: Chem. 260 (2006) 82. D.E. Ridler, and M.V. Twigg, Catalyst Handbook (2nd. Ed.), ICI, Wolf Publishing Ltd., London (1989). O. Ilinich, W. Ruettinger, X. Liu, and R. Farrauto, J. Catal. 247 (2007) 112. A.F. Ghenciu, Curr. Opin. Solid State Mater. Sci. 6 (2002) 389. A. Qi, B. Peppley, and K. Karan, Fuel Process. Technol. 88 (2007) 3. J.C. Gonzalez, M.G. Gonzalez, M.A. Laborde, and N. Moreno, Appl. Catal. 20 (1986) 3. C. Rhodes, and G.J. Hutchings, Phys. Chem. Chem. Phys. 5 (2003) 2719. H. Topsoe, and M. Boudart, J. Catal. 31 (1973) 346. 228

[5]

[6]

[7] [8] [9] [10]

[11] [12]

[13] [14] [15] [16]

C. Rhodes, G.J. Hutchings, and A.M. Ward, Catal. Today 23 (1995) 43. Q. Liu, W. Ma, R. He, and Z. Mu, Catal. Today 106 (2005) 52. Y. Lei, N.W. Cant, and D.L. Trimm, J. Catal. 239 (2006) 227. L. Hu, G. Xia, L. Qu, M. Li, C. Li, Q. Xin, and D. Li, J. Catal. 202 (2001) 220. A. Andreev, V. Idakiev, D. Mihajlova, and D. Shopov, Appl. Catal. 22 (1986) 385. J.L. Rangle Costa, G.S. Marchetti, and M.C. Rangel, Catal. Today 77 (2002) 205. J. Moutinho, T. Souza, and M.C. Rangel, React. Kinet. Catal. Lett. 77 (2002) 29. J. Moutinho, T. Souza, and M.C. Rangel, React. Kinet. Catal. Lett. 83 (2004) 93. A.O. Souza, and M.C. Rangel, React. Kinet. Catal. Lett. 79 (2003) 175. G.C. Arajo, and M.C. Rangel, Catal. Today 62 (2000) 201. I.L. Jnior, J.-M.M. Millet, M. Aouine, and M.C. Rangel, Appl. Catal. A: Gen 283 (2005) 91. Q. Liu, W. Ma, Z. Mou, and Q. Zhang, Prepr. Pap. - Am. Chem. Soc., Div. Fuel Chem. 50 (2005) 424. M. Qi, and M. Flytzani-Stephanopoulos, Ind. Eng. Chem. Res. 43 (2004) 3055. C. Rhodes, B.P. Williams, F. King, and G.J. Hutchings, Catal. Commun. 3 (2002) 381. 229

[17]

[18]

[19]

[20]

[21] [22] [23]

[24]

[25]

[26]

[27]

M.A. Edwards, D.M. Whittle, C. Rhodes, A.M. Ward, D. Rohan, M.D. Shannon, G.J. Hutchings, and C.J. Kiely, Phys. Chem. Chem. Phys. 4 (2002) 3902. Y. Hu, H. Jin, J. Liu, and D. Hao, Chem. Eng. J. 78 (2000) 147. J.L. Rangel Costa, G.S. Marchetti, and M.C. Rangel, Catal. Today 77 (2002) 205. T. Shishido, M. Yamamoto, D. Li, Y. Tian, H. Morioka, M. Honda, T. Sano, and K. Takehira, Appl. Catal. A: Gen 303 (2006) 62. M.S. Spencer, Top. Catal. 8 (1999) 259. M. Rnning, F. Huber, H. Meland, H. Venvik, D. Chen, and A. Holmen, Catal. Today 100 (2005) 249. G. Jacobs, E. Chenu, P.M. Patterson, L. Williams, D. Sparks, G. Thomas, and B.H. Davis, Appl. Catal. A: Gen 258 (2004) 203. T. Bunluesin, R.J. Gorte, and G.W. Graham, Appl. Catal. B: Environ 15 (1998) 107. S. Hilaire, X. Wang, T. Luo, R.J. Gorte, and J. Wagner, Appl. Catal. A: Gen 215 (2001) 271. E.S. Bickford, S. Velu, and C. Song, Catal. Today 99 (2005) 347. Y. Denkwitz, A. Karpenko, V. Plzak, R. Leppelt, B. Schumacher, and R.J. Behm, J. Catal. 246 (2007) 74. F. Zhang, Q. Zheng, K. Wei, X. Lin, H. Zhang, J. Li, and Y. Cao, Catal. Lett. 108 (2006) 131. Y. Sato, Y. Soma, T. Miyao, and S. Naito, Appl. Catal. A: Gen 304 (2006) 78. 230

[28] [29]

[30]

[31] [32]

[33]

[34]

[35]

[36] [37]

[38]

[39]

[40] [41]

P. Panagiotopoulou, and D.I. Kondarides, Catal. Today 112 (2006) 49. P. Panagiotopoulou, A. Christodoulakis, D.I. Kondarides, and S. Boghosian, J. Catal. 240 (2006) 114. Y. Tanaka, T. Takeguchi, R. Kikuchi, and K. Eguchi, Appl. Catal. A: Gen 279 (2005) 59. Y. Tanaka, T. Utaka, R. Kikuchi, T. Takeguchi, K. Sasaki, and K. Eguchi, J. Catal. 215 (2003) 271. M. Nagai, and K. Matsuda, J. Catal. 238 (2006) 489. A.A. Andreev, V.J. Kafedjiysky, and R.M. Edreva-Kardjieva, Appl. Catal. A: Gen 179 (1999) 223. R.N. Nickolov, R.M. Edreva-Kardjieva, V.J. Kafedjiysky, D.A. Nikolova, N.B. Stankova, and D.R. Mehandjiev, Appl. Catal. A: Gen 190 (2000) 191. E.F. Armstrong, and T.P. Hilditch, Proc. R. Soc. 97 (1920) 265. T. Shido, and Y. Iwasawa, J. Catal. 141 (1993) 71. T. Shido, and Y. Iwasawa, J. Catal. 129 (1991) 343. T. Shido, and Y. Iwasawa, J. Catal. 136 (1992) 493. G. Jacobs, G.M. Patterson, U.M. Graham, D.E. Sparks, and B.H. Davis, Appl. Catal. A: Gen 269 (2004) 63. Y. Sato, K. Terada, S. Hasegawa, T. Miyao, and S. Naito, Appl. Catal. A: Gen 296 (2005) 80. T. Luo, and R.J. Gorte, Catal. Lett. 85 (2003) 139.

[42]

[43]

[44] [45]

[46]

[47] [48] [49] [50] [51]

[52]

[53]

231

[54]

M.L. Kundu, A.C. Sengupta, G.C. Maiti, B. Sen, S.K. Ghosh, V.I. Kuznetsov, G.N. Kustova, and E.N. Yurchenko, J. Catal. 112 (1988) 375. U.S. Ozkan, Y. Cai, M.W. Kumthekar, and L. Zhang, J. Catal. 142 (1993) 182. R.D. Gonzalez, T. Lopez, and R. Gomez, Catal. Today 35 (1997) 293. V. Idakiev, A.D. Mihajlo, B. Kanev, and A. Andreev, React. Kinet. Catal. Lett. 33 (1987) 119. H. Topsoe, J.A. Dumesic, and M. Boudart, J. Catal. 28 (1973) 477. D.G. Rethwisch, and J.A. Dumesic, Appl. Catal. 21 (1986) 97. K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouqurol, and T. Siemieniewska, Pure Appl. Chem. 57 (1985) 603. M. Kruk, and M. Jaroniec, Chem. Mater. 13 (2001) 3169. B. Tan, S.M. Vyas, H.-J. Lehmler, B.L. Knutson, and S.E. Rankin, J. Phys. Chem. B 109 (2005) 23225. B. Tan, S.M. Vyas, H.-J. Lehmler, B.L. Knutson, and S.E. Rankin, Adv. Funct. Mater. 17 (2007) 2500. B. Tan, H.-J. Lehmler, S.M. Vyas, B.L. Knutson, and S.E. Rankin, Adv. Mater. 17 (2005) 2368. H.-P. Lin, S.-T. Wong, C.-Y. Mou, and C.-Y. Tang, J. Phys. Chem. B 104 (2000) 8967. C.R.F. Lund, J.E. Kubsh, and J.A. Dumesic, Water gas shift over magnetite-based catalysts: Nature of active sites for adsorption and catalysis. American Chemical Society, Washington, D.C., 1985, 313. 232

[55]

[56] [57]

[58] [59] [60]

[61] [62]

[63]

[64]

[65]

[66]

[67] [68]

G. Parravano, Ind. Eng. Chem. 49 (1957) 266. J.M. Pigos, C.J. Brooks, G. Jacobs, and B.H. Davis, Appl. Catal. A: Gen 319 (2007) 47. G. Doppler, A.X. Trautwein, and H.M. Ziethen, Appl. Catal. 40 (1988) 119. K.G. Azzam, I.V. Babich, K. Seshan, and L. Lefferts, J. Catal. 251 (2007) 163. K.G. Azzam, I.V. Babich, K. Seshan, and L. Lefferts, J. Catal. 251 (2007) 153. X. Wang, J.A. Rodriguez, J.C. Hanson, D. Gamarra, A. Martnez-Arias, and M. Fernndez-Garca, J. Phys. Chem. B 110 (2006) 428. L. Zhang, X. Wang, and U.S. Ozkan, Appl. Catal. A: Gen submitted. G.A. Ferguson, Jr., and M. Hass, Phys. Rev. 112 (1958) 1130. F.E. DeBore, and P.W. Selwood, J. Am. Chem. Soc. 76 (1954) 3365. A.F.H. Wielers, A.J.H.M. Kock, C.E.C.A. Hop, J.W. Geus, and A.M. van der Kraan, J. Catal. 117 (1989) 1. H. Topse, and M. Boudart, J. Catal. 31 (1973) 346. C.R.F. Lund, and J.A. Dumesic, J. Catal. 72 (1981) 21. C.R.F. Lund, and J.A. Dumesic, J. Phys. Chem. 85 (1981) 3175. C.R.F. Lund, and J.A. Dumesic, J. Phys. Chem. 86 (1982) 130. C.R.F. Lund, and J.A. Dumesic, J. Catal. 76 (1982) 93. 233

[69] [70]

[71]

[72]

[73] [74] [75] [76]

[77] [78] [79] [80] [81]

[82]

S. Herreyre, P. Gadelle, P. Moral, and J.-M.M. Millet, J. Phys. Chem. Solids 58 (1997) 1539. E. Xue, M. O'Keeffe, and J.R.H. Ross, Catal. Today 30 (1996) 107.

[83]

[84] M. Gong, X. Liu, J. Trembly, and C. Johnson, J. Power Sources 168 (2007) 289. [85] [86] P. Forzatti, and L. Lietti, Catal. Today 52 (1999) 165. C.H. Bartholomew, P.K. Agrawal, and J.R. Katzer, Adv. Catal. 31 (1982) 135. J. Barbier, E. Lamy-pitara, P. Marecot, J.P. Boitiaux, J. Cosyns, and F. Verna, Adv. Catal. 37 (1990) 279. M.S. Kim, N.M. Rodriguez, and R.T.K. Baker, J. Catal. 143 (1993) 449. M. Ziolek, J. Kujawa, O. Saur, and J.C. Lavalley, J. Mol. Catal. A: Chem. 97 (1995) 49. J.H. Edwards, and A.M. Maitra, Fuel Process. Technol. 42 (1995) 269. C. Kim, and G.A. Somorjai, J. Catal. 134 (1992) 179. Z. Pal, K. Matusek, and M. Muhler, Appl. Catal. A: Gen 149 (1997) 113. M.V. Twigg, and M.S. Spencer, Appl. Catal. A: Gen 212 (2001) 161. N.S. Fgoli, P.C. L'Argentiere, A. Arcoya, and X.L. Seoane, J. Catal. 155 (1995) 95. P.K. Cheekatamarla, and A.M. Lane, J. Power Sources 153 (2006) 157. J.J. Strohm, J. Zheng, and C. Song, J. Catal. 238 (2006) 309. 234

[87]

[88] [89]

[90] [91] [92] [93] [94]

[95] [96]

[97] [98]

C.H. Bartholomew, and R.M. Bowman, Appl. Catal. 15 (1985) 59. N. Koizumi, K. Murai, T. Ozaki, and M. Yamada, Catal. Today 89 (2004) 465. J. Li, and N.J. Coville, Appl. Catal. A: Gen 208 (2001) 177.

[99]

[100] M. Yamada, N. Koizumi, T. Miyazawa, and T. Furukawa, Catal. Lett. 78 (2002) 195. [101] Z.-T. Liu, J.-L. Zhou, and B.-J. Zhang, J. Mol. Catal. 94 (1994) 255. [102] A.L. Chaffee, I. Campbell, and N. Valentine, Appl. Catal. 47 (1989) 253. [103] J.N. Kuhn, N. Lakshminarayanan, and U.S. Ozkan, J. Mol. Catal. A: Chem. 282 (2008) 9. [104] D. Nikolova, R. Edreva-Kardjieva, G. Gouliev, T. Grozeva, and P. Tzvetkov, Appl. Catal. A: Gen 297 (2006) 135. [105] F.M. Gottschalk, R.G. Copperthwaite, M. Van Der Riet, and G.J. Hutchings, Appl. Catal. 38 (1988) 103. [106] F.M. Gottschalk, and G.J. Hutchings, J. Chem. Soc., Chem. Commun. (1988) 123. [107] R.G. Copperthwaite, F.M. Gottschalk, and T. Sangiorgio, Appl. Catal. 63 (1990) L11. [108] G.J. Hutchings, R.G. Copperthwaite, F.M. Gottschalk, R. Hunter, J.R. Mellor, S.w. Orchard, and T. Sangiorgio, J. Catal. 137 (1992) 408. [109] F.M. Gottschalk, and G.J. Hutchings, Appl. Catal. 51 (1989) 127. [110] J. Ryczkowski, Catal. Today 68 (2001) 263. 235

[111] M.M. Yung, Z. Zhao, M.P. Woods, and U.S. Ozkan, J. Mol. Catal. A: Chem. 279 (2008) 1. [112] I.A. Fisher, and A.T. Bell, J. Catal. 172 (1997) 222. [113] X. Liu, W. Ruettinger, X. Xu, and R. Farrauto, Appl. Catal. B: Environ 56 (2005) 69. [114] Q. Fu, W. Deng, H. Saltsburg, and M. Flytzani-Stephanopoulos, Appl. Catal. B: Environ 56 (2005) 57. [115] C.H. Kim, and L.T. Thompson, J. Catal. 230 (2005) 66. [116] S.S. Tamhankar, M. Bagajewlcz, and G.R. Gavalas, Ind. Eng. Chem. Process Des. Dev. 25 (1986) 429. [117] M.V. Twigg, Catalyst Handbook (2nd. Ed.), ICI, Wolf Publishing Ltd., London (1989). [118] T.C. Bromfield, and N.J. Coville, Appl. Surf. Sci. 119 (1997) 19. [119] K. Habermehl-wirze, and J. Lahtinen, Surf. Sci. 573 (2004) 183. [120] N.N. Greenwood, and T.C. Gibb, Mssbauer Spectroscopy, Chapman and Hall Ltd, London (1971). [121] Z. Cheng, and M. Liu, Solid State Ionics 178 (2007) 925. [122] M. George, S.S. Nair, K.A. Malini, P.A. Joy, and M.R. Anantharaman, J. Phys. D: Appl. Phys. 40 (2007) 1593. [123] Y. Tanaka, T. Utaka, R. Kikuchi, K. Sasaki, and K. Eguchi, Appl. Catal. A: Gen. 242 (2003) 287. [124] A. Haryanto, S. Fernando, N. Murali, and S. Adhikari, Energ. Fuel 19 (2005) 2098. 236

[125] S.S. Bharadwaj, and L.D. Schmidt, Fuel Process. Technol. 42 (1995) 109. [126] D.E. Ridler, and M.V. Twigg, Catalyst Handbook (2nd. Ed.), ICI, Wolf Publishing Ltd., London (1989) 225. [127] F. Pompeo, N.N. Nichio, M.M.V.M. Souza, D.V. Cesar, O.A. Ferretti, and M. Schmal, Appl. Catal. A: Gen 316 (2007) 175. [128] K.M. Hardiman, T.T. Ying, A.A. Adesina, E.M. Kennedy, and B.Z. Dlugogorski, Chem. Eng. J. 102 (2004) 119. [129] Q. Ming, T. Healey, L. Allen, and P. Irving, Catal. Today 77 (2002) 51. [130] J.R. Rostrup-Nielsen, and I. Alstrup, Catal. Today 53 (1999) 311. [131] M.C.J. Bradford, and M.A. Vannice, Appl. Catal. A: Gen 142 (1996) 73. [132] J.T. Richardson, and S.A. Paripatyadar, Appl. Catal. A: Gen 61 (1990) 293. [133] A. Igarashi, T. Ohtaka, and S. Motoki, Catal. Lett. 13 (1991) 189. [134] J.R. Rostrup-Nielsen, and J.H. Hansen, J. Catal. 144 (1993) 38. [135] D.L. Qin, J., Catal. Today 21 (1994) 551. [136] R. Craciun, B. Shereck, and R.J. Gorte, Catal. Lett. 51 (1998) 149. [137] X. Wang, and R.J. Gorte, Catal. Lett. 73 (2001) 15. [138] Y.-G. Chen, K. Tomishige, K. Yokoyama, and K. Fujimoto, J. Catal. 184 (1999) 479. [139] L. Pelletier, and D.D.S. Liu, Appl. Catal. A: Gen 317 (2007) 293.

237

[140] J.R. Rostrup-Nielsen, Catalytic Steam Reforming in: Catalysis: Science and Technology, J.R. Anderson and M. Boudart. New York: Springer Verlag 5 (1984) 3. [141] T. Borowiecki, and A. Golebiowski, Catal. Lett. 25 (1994) 309. [142] T. Borowiecki, A. Golebiowski, and B. Stasinska, Appl. Catal. A: Gen 153 (1997) 141. [143] B. Stasinska, A. Golebiowski, and T. Borowiecki, in: Catalyst Deactivation, B. Delmon, G.F. Froment (Eds.), Elsevier, Amsterdam (1999) 431. [144] T. Borowiecki, and A. Machocki, in: Catalyst Deactivation, B. Delmon, G. F. Froment (Eds.), Elsevier, Amsterdam (1999) 435. [145] T. Borowiecki, G. Giecko, and M. Panczyk, Appl. Catal. A: Gen 230 (2002) 85. [146] Q. Zhuang, Y. Qin, and L. Chang, Appl. Catal. 70 (1991) 1. [147] S. Wang, and G.Q. Lu, Appl. Catal. B: Environ 19 (1998) 267. [148] R.M. Sambrook, and J.R.H. Ross, US Patent 4 469 815 (1984). [149] S. Natesakhawat, O. Oktar, and U.S. Ozkan, J. Mol. Catal. A: Chem. 241 (2005) 133. [150] S. Natesakhawat, R.B. Watson, X. Wang, and U.S. Ozkan, J. Catal. 234 (2005) 496. [151] A. Slagtern, Y. Schuurman, C. Leclercq, X. Verykios, and C. Mirodatosy, J. Catal. 172 (1997) 118. [152] B. Jongsomjit, J. Panpranot, and J.G. Goodwin Jr., J. Catal. 215 (2003) 66. [153] B.S. Liu, and C.T. Au, Appl. Catal. A: Gen 244 (2003) 181. 238

[154] H.Y. Wang, and E. Ruckenstein, Appl. Catal. A: Gen 209 (2001) 207. [155] Y. Chen, and J. Ren, Catal. Lett. 29 (1994) 39. [156] P.D. Vaidya, and A.E. Rodrigues, Chem. Eng. J. 117 (2006) 39. [157] F. Auprtre, C. Descorme, and D. Duprez, Catal. Commun. 3 (2002) 263. [158] J.P. Breen, R. Burch, and H.M. Coleman, Appl. Catal. B: Environ 39 (2002) 65. [159] F. Haga, T. Nakajima, H. Miya, and S. Mishima, Catal. Lett. 48 (1997) 223. [160] M.S. Batista, R.K.S. Santos, E.M. Assaf, J.M. Assaf, and E.A. Ticianelli, J. Power Sources 124 (2003) 99. [161] M.S. Batista, R.K.S. Santos, E.M. Assaf, J.M. Assaf, and E.A. Ticianelli, J. Power Sources 134 (2004) 27. [162] J. Llorca, N. Homs, J. Sales, and P. Ramrez de la Piscina, J. Catal. 209 (2002) 306. [163] J. Llorca, P. Ramrez de la Piscina, J. Sales, and N. Homs, Chem. Commun. 2001 (2001) 641. [164] J. Llorca, J.-A. Dalmon, P. Ramrez de la Piscina, and N. Homs, Appl. Catal. A: Gen 243 (2003) 261. [165] J. Llorca, P. Ramrez de la Piscina, J.-A. Dalmon, J. Sales, and N. Homs, Appl. Catal. B: Environ 43 (2003) 355. [166] F. Aupretre, C. Descorme, D. Duprez, D. Casanave, and D. Uzio, J. Catal. 233 (2005) 464. [167] A.Y. Khodakov, R. Bechara, and A. Griboval-Constant, Appl. Catal. A: Gen 254 (2003) 273. 239

[168] G. Jacobs, T.K. Das, Y. Zhang, J. Li, G. Racoillet, and B.H. Davis, Appl. Catal. A: Gen 233 (2002) 263. [169] J. Sun, X.-P. Qiu, F. Wu, and W.-T. Zhu, Int. J. Hydrogen Energy 30 (2005) 437. [170] A.N. Fatsikostas, D.I. Kondarides, and X.E. Verykios, Catal. Today 75 (2002) 145. [171] A.N. Fatsikostas, D.I. Kondarides, and X.E. Verykios, Chem. Commun. 2001 (2001) 851. [172] D.K. Liguras, D.I. Kondarides, and X.E. Verykios, Appl. Catal. B: Environ 43 (2003) 345. [173] N. Laosiripojana, and S. Assabumrungrat, Appl. Catal. B: Environ 66 (2006) 29. [174] J. Kugai, V. Subramani, C. Song, M.H. Engelhard, and Y.-H. Chin, J. Catal. 238 (2006) 430. [175] J. Kugai, S. Velu, and C. Song, Catal. Lett. 101 (2005) 255. [176] B. Zhang, X. Tang, Y. Li, W. Cai, Y. Xu, and W. Shen, Catal. Commun. 7 (2006) 367. [177] S. Cavallaro, Energ. Fuel 14 (2000) 1195. [178] F. Frusteri, S. Freni, L. Spadaro, V. Chiodo, G. Bonura, S. Donato, and S. Cavallaro, Catal. Commun. 5 (2004) 611. [179] C. Diagne, H. Idriss, and A. Kiennemann, Catal. Commun. 3 (2002) 565. [180] A. Casanovas, J. Llorca, N. Homs, J.L.G. Fierro, and P.R. de la Piscina, J. Mol. Catal. A: Chem. 250 (2006) 44. 240

[181] M.A. Goula, S.K. Kontou, and P.E. Tsiakaras, Appl. Catal. B: Environ 49 (2004) 135. [182] F.J. Mario, E.G. Cerrella, S. Duhalde, M. Jobbagy, and M.A. Laborde, Int. J. Hydrogen Energy 23 (1998) 1095. [183] J. Sun, X. Qiu, F. Wu, W. Zhu, W. Wang, and S. Hao, Int. J. Hydrogen Energy 29 (2004) 1075. [184] M.N. Barroso, M.F. Gomez, L.A. Arra, and M.C. Abello, Appl. Catal. A: Gen 304 (2006) 116. [185] F. Mario, G. Baronetti, M. Jobbagy, and M. Laborde, Appl. Catal. A: Gen 238 (2002) 41. [186] F. Mario, M. Boveri, G. Baronetti, and M. Laborde, Int. J. Hydrogen Energy 26 (2001) 665. [187] J. Llorca, N. Homs, and P. Ramrez de la Piscina, J. Catal. 227 (2004) 556. [188] J. Llorca, N. Homs, J. Sales, J.-L.G. Fierro, and P. Ramrez de la Piscina, J. Catal. 222 (2004) 470. [189] J. Llorca, P. Ramrez de la Piscina, J.-A. Dalmon, and N. Homs, Chem. Mater. 16 (2004) 3573. [190] T. Montini, L. De Rogatis, V. Gombac, P. Fornasiero, and M. Graziani, Appl. Catal. B: Environ 71 (2007) 125. [191] D. Duprez, P. Pereira, A. Miloudi, and R. Maurel, J. Catal. 75 (1982) 151. [192] M. Dmk, M. Tth, J. Rask, and A. Erdhelyi, Appl. Catal. B: Environ 69 (2007) 262. [193] A. Yee, S.J. Morrison, and H. Idriss, J. Catal. 186 (1999) 279. 241

[194] J. Rask, M. Dmk, K. Ban, and A. Erdhelyi, Appl. Catal. A: Gen 299 (2006) 202. [195] P.-Y. Sheng, A. Yee, G.A. Bowmaker, and H. Idriss, J. Catal. 208 (2002) 393. [196] L.V. Mattos, and F.B. Noronha, J. Catal. 233 (2005) 453. [197] A. Erdhelyi, J. Rask, T. Kecsks, M. Tth, M. Dmk, and K. Ban, Catal. Today 116 (2006) 367. [198] C. Resini, S. Cavallaro, F. Frusteri, S. Freni, and G. Busca, React. Kinet. Catal. Lett. 90 (2007) 117. [199] X. Wang, and U.S. Ozkan, J. Phys. Chem. B 109 (2005) 1882. [200] M.L. Jacono, M. Schiavello, and A. Cimino, J. Phys. Chem. 75 (1971) 1044. [201] K.T. Ng, and D.M. Hercules, J. Phys. Chem. 80 (1976) 2094. [202] D.K. Kim, K. Stwe, F. Mller, and W.F. Maier, J. Catal. 247 (2007) 101. [203] R.D. Gonzalez, T. Lopez, and R. Gmez, Catal. Today 35 (1997) 293. [204] M.H. Youn, J.G. Seo, P. Kim, and I.K. Song, J. Mol. Catal. A: Chem. 261 (2007) 276. [205] V.R. Choudhary, and A.S. Mamman, Appl. Energ. 66 (2000) 161. [206] J.S. Lisboa, D.C.R.M. Santos, F.B. Passos, and F.B. Noronha, Catal. Today 101 (2005) 15. [207] P. Wang, E. Tanabe, K. Itob, J. Jia, H. Morioka, T. Shishido, and K. Takehira, Appl. Catal. A: Gen 231 (2002) 35. 242

[208] S. Wang, and G.Q. Lu, Ind. Eng. Chem. Res. 38 (1999) 2615. [209] N.M. Rodriguez, J. Mater. Res. 8 (1993) 3233. [210] J.-W. Snoeck, G.F. Froment, and M. Fowlesz, J. Catal. 169 (1997) 240. [211] J.-W. Snoeck, G.F. Froment, and M. Fowlesy, J. Catal. 169 (1997) 250. [212] R.T.K. Baker, M.A. Barber, P.S. Harris, F.S. Feates, and R.J. Waite, J. Catal. 26 (1972) 51. [213] S. Cavallaro, V. Chiodo, S. Freni, N. Mondello, and F. Frusteri, Appl. Catal. A: Gen 249 (2003) 119. [214] J.R. Mielenz, Curr. Opin. Microbiol. 4 (2001) 324. [215] M. Kraum, and M. Baerns, Appl. Catal. A: Gen 186 (1999) 189. [216] S.-W. Ho, and Y.-S. Su, J. Catal. 168 (1997) 51. [217] S. Sun, K. Fujimoto, Y. Zhang, and N. Tsubaki, Catal. Commun. 4 (2003) 361. [218] A. Kaddouri, and C. Mazzocchia, Catal. Commun. 5 (2004) 339. [219] H. Song, L. Zhang, R.B. Watson, D. Braden, and U.S. Ozkan, Catal. Today 129 (2007) 346. [220] E. Ruckenstein, and H.Y. Wang, Catal. Lett. 70 (2000) 15. [221] A.R. Belambe, R. Oukaci, and J.G. Goodwin, Jr., J. Catal. 166 (1997) 8. [222] D.I. Enache, B. Rebours, M. Roy-Auberger, and R. Revel, J. Catal. 205 (2002) 346. 243

[223] D.I. Enache, M. Roy-Auberger, and R. Revel, Appl. Catal. A: Gen 268 (2004) 51. [224] S. Freni, S. Cavallaro, N. Mondello, L. Spadaro, and F. Frusteri, Catal. Commun. 4 (2003) 259. [225] H. Song, L. Zhang, and U.S. Ozkan, Green Chem. 9 (2007) 686. [226] M.O. Zacate, L. Minervini, D.J. Bradfield, R.W. Grimes, and K.E. Sickafus, Solid State Ionics 128 (2000) 243. [227] J.M. Guil, N. Homs, J. Llorca, and P. Ramrez de laPiscina, J. Phys. Chem. B 109 (2005) 10813. [228] J. Rask, A. Hancz, and A. Erdhelyi, Appl. Catal. A: Gen 269 (2004) 13. [229] A. Auroux, and A. Gervasini, J. Phys. Chem. 94 (1990) 6371. [230] L.V. Mattos, and F.B. Noronha, J. Power Sources 152 (2005) 50. [231] K.T. Cheshkova, and D.G. Stoilova, React. Kinet. Catal. Lett. 73 (2001) 237.

244

APPENDIX A LIST OF ACRONYMS

HT: High temperature LT: Low temperature GC: Gas chromatography TPR: Temperature programmed reduction TPD: Temperature programmed oxidation TPRxn: Temperature programmed reaction XPS: X-ray photoelectron spectroscopy XRD: X-ray diffraction TEM: Transmission electron microscopy DRIFTS: Diffuse reflectance infrared Fourier transform spectroscopy BET: Brunauer-Emmett-Teller BJH: Barret-Joiner-Halenda IWI: Incipient wetness impregnation SG: Sol-gel WGS: Water-gas shift

245

Vous aimerez peut-être aussi