REDOX FOR ENVIRONMENTAL APPLICATIONS

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree

Doctor of Philosophy

in the Graduate School of The Ohio State University

By

Preshit Vilas Gawade, M.S.

Chemical Engineering Graduate Program

*****

The Ohio State University

2012

Dissertation Committee:

Umit S. Ozkan, Advisor

Jeffrey J. Chalmers

Kurt W. Koelling

James Coe

Copyright by

Preshit Vilas Gawade

2012

ABSTRACT

The presented work comprehends a broad spectrum of catalysis for various environmental applications, such as i) hydrogen production via water-gas shift reaction, ii) hydrogen purification for fuel cell applications and iii) catalytic after-treatment of lean-burn engines. This dissertation involves, but is not limited to catalyst development, reaction studies and catalyst characterization for the above-mentioned environmental applications, which can be summarized as follows.

(i) Water-gas shift (WGS) remains an essential step in integrated gasification combined cycle (IGCC) for hydrogen production, as it forms a link between the gasification process and fuel cell operations. The current catalysts for WGS application are based on Fe-Cr and Cu/ZnO/Al2O3, as a high temperature (HT-WGS) and low temperature (LT-WGS) catalysts, respectively. This two-stage WGS process is a consequence of several operational drawbacks of the current catalyst formulations including Cr+6 being carcinogenic. Hence the presented WGS project has a two-fold purpose. First, Cr-free Fe-based catalyst development and second, Cu supported catalyst development for WGS that can be operated over a wide temperature range.

In this dissertation, Cr- free Fe-Al-Cu catalyst prepared through “one-step” sol- gel method using propylene oxide as a gelation agent has been reported. Steady state reactions demonstrated that WGS performance of Fe-Al-Cu was superior as compared to

ii commercial Fe-Cr catalyst. The reaction studies along with complementary catalyst characterization indicated that the amount of copper in iron oxide matrix played a crucial role. The optimized ratio of Fe to Cu was found to be five and any further increase in copper loading resulted in copper segregation from the iron oxide matrix. Thus various catalyst characterization techniques were exploited to understand this phenomenon.

Furthermore, a detailed study was performed to comprehend the formation of surface species during WGS reaction and to evaluate the reaction mechanism over Fe-Al-Cu.

In the quest of exploring Cu-based catalyst for WGS system, Cu supported over various CeO2 nano-morphologies were investigated. Here, nanoparticles (NP) and nanorods (NR) of CeO2 were prepared through hydrothermal precipitation method and copper was supported on these morphologies using a wet impregnation method. In the current findings, copper was more finely and uniformly dispersed over CeO2 nano- particles compared to nanorods, resulting in better WGS activity compared to particle- based samples. Catalyst characterization indicated finely dispersed copper particles in close interaction with ceria nanoparticles, whereas isolated bulk-like copper species were formed over the ceria nanorods. Finally, the formation of surface species during WGS reaction delineated the redox reaction mechanism over Cu/CeO2.

(ii) Hydrogen produced via WGS reaction may contain up to 1-2% CO in stream which can be poisonous to proton exchange membrane (PEM) fuel cell. Preferential oxidation of carbon monoxide (PROX) is considered as an effective and economical way

iii to purify the hydrogen stream for PEM fuel cell applications. The major challenge in this process is to selectively oxidize CO with minimum loss of hydrogen. Hence a non- precious metal catalyst such as, cobalt supported over ceria with a special focus on cobalt loading has been utilized. Both, activity and selectivity were found to be a strong function of cobalt content. In addition, CO and hydrogen oxidation kinetics was studied as a function of cobalt loading. The higher activation energy barrier for hydrogen oxidation compared to CO oxidation indicated higher temperature sensitivity for hydrogen oxidation. The cobalt phase was identified, as Co3O4 and it remained stable under PROX atmosphere. Time-on-stream experiment along with various catalyst characterization techniques indicated no significant contribution from lower valency cobalt species.

Finally, the formation of surface species during PROX reaction demonstrated conversion of carbonate species to more stable polydentate carbonates and formate type species with increase in reaction temperature.

(iii) Lean-burn natural gas fired engines remains a popular choice in the energy market. Despite emission being greatly reduced, exhaust still contains considerable amount of NOx, CO and hydrocarbons. Hence after-treatment to clean up the exhaust is essential. The selective catalytic reduction using hydrocarbons is considered as a promising alternative for conventional after-treatment technology and is well suited, especially for natural gas lean-burn engines.

iv

For this purpose, a “single-stage” de-NOX system composed of a physical mixture of dual-catalyst bed has been developed. This dual catalyst bed was a physical mixture of reduction (Pd/SZ) and oxidation (CoOx/CeO2) catalysts capable of performing three distinct functions, NO oxidation, NOx reduction, and CO and hydrocarbon oxidation.

Here, oxidation catalyst was assumed to play multi-functional role in dual-catalyst bed.

These include oxidizing NO or re-oxidizing partially reduced NOx species, CO oxidation and catalyzing the combustion of un-burned hydrocarbons, which have not participated during SCR reaction.

In this dissertation, the role of oxidation catalyst in dual-catalyst bed was addressed. NO oxidation was studied as a function of cobalt loading in CoOx/CeO2 formulation. The dual-catalyst bed was optimized by varying the reduction to oxidation catalyst ratio in order to achieve significantly high NOx conversion during hydrocarbon-

SCR. The lower cobalt loading in oxidation catalyst in a dual-bed resulted in higher NOx conversion. This observation was associated with lower hydrocarbon oxidation and hence increased hydrocarbon availability for NOx reduction. Kinetic study along with catalyst characterization confirmed the activation of methane molecule via hydrogen abstraction, consequently participating in either in NOx reduction or directly oxidizing over the oxidation catalyst. Moreover, the effect of water vapor was thoroughly investigated over the optimized dual-catalyst bed. The primary focus of this work was to improve the hydrothermal stability of the dual-catalyst bed by changing the various engine exhaust parameters.

v

Dedicated to my parents and my wife

vi

ACKNOWLEDGMENTS

I would like to thank Prof. Umit S. Ozkan for her support and guidance throughout my Ph.D. Her dedication to work, endless energy, support and respect to fellow co-workers has always inspired me.

I have always believed that research is a team-effort. I thank all the current and former group members for contributing directly or indirectly to my research. Especially, I would like to thank Burcu Bayram and Anne-Marie Alexander for their contributions towards my projects. Some of the work presented here was completed with their help.

I thank Ohio State University for providing excellent research facility. I thank department staff and faculty members for their support and help. Especially, I would like to thank Prof. Jeffrey Chalmers and Prof. Kurt Koelling for serving on my committee and providing valuable recommendations.

Financial support from Ohio Coal Development Office and Department of Energy is greatly appreciated. I also thank Caterpillar for funding the emission clean up project.

I would like to acknowledge Ronald Silver (Technical Centre, Caterpillar) for the valuable discussion on the NOx reduction project.

I would like to thank my friends, especially “Columbus-Gang” for their friendship and making my stay at Columbus memorable. I think I will really miss you guys. Finally,

I thank my parents and beautiful wife for their endless support and love.

vii

VITA

December 1, 1982..………………………………………Born, Mumbai, India

October 2004..……………………………………………B.S. Chemical Engineering U.D.C.T., Mumbai, India

December 2007..…………………………………………M.S. Chemical Engineering University of Toledo, Ohio

December 2007-Present..………………………………..Graduate Research Associate. Ohio State University, Ohio

INVENTION DISCLOSURES

I. U.S. Ozkan, P. Gawade and A.M. Alexander, “Addition of alumina binder during palladium over sulfated zirconia preparation for monolith coating: Implications NOx selective catalytic reduction from lean-burn natural gas engines”, Invention Disclosure, Technology Commercialization and Knowledge Transfer, Ohio State University, (2012), (ID: 2012-294) Pending

II. U.S. Ozkan, P. Gawade and A.M. Alexander, “Novel methods to coat supported metal catalysts on monolith cores”, Invention Disclosure, Technology Commercialization and Knowledge Transfer, Ohio State University, (2012), (ID: 2012-293) Pending

PUBLICATIONS

1. I. Soykal, B. Bayram, H. Sohn, P. Gawade, J. Miller and U.S. Ozkan, “Ethanol over Co/CeO2 catalyst: Investigation of effect of ceria morphology”, Applied Catalysis A, Submitted

viii

2. P. Gawade, A.M. Alexander, R. Silver and U. S. Ozkan, “Effect of various engine exhaust parameters on the hydrothermal stability of hydrocarbon-SCR catalysts for lean-burn systems”, In Preparation

3. P. Gawade, A.M. Alexander, R. Clark, and U. S. Ozkan, “The Role of oxidation catalyst in dual-catalyst bed for after-treatment of lean burn natural gas exhaust”, Catalysis Today, Submitted

4. P. Gawade, B. Bayram, A.M. Alexander and U.S. Ozkan, “Preferential oxidation of CO (PROX) over CoOx/CeO2 nanoparticles in hydrogen-rich stream: Effect of cobalt loading”, Applied Catalysis B: Environmental, Submitted

5. P. Gawade, B. Mirkelamoglu, and U. S. Ozkan, “The role of support morphology and impregnation medium on the water gas-shift activity of ceria-supported copper catalysts”, Journal of Physical Chemistry: C, (2010), 114, 18173–18181

6. P. Gawade, B. Mirkelamoglu, B. Tan, and U. S. Ozkan, “Cr-free Fe-based water-gas-shift catalysts prepared through propylene-oxide sol-gel assisted technique”, Journal of Molecular Catalysis A: Chemical 321, (2010), 61-70

7. M. P. Woods, P. Gawade, B. Tan, and U. S. Ozkan, “Preferential oxidation of carbon monoxide on Co/CeO2 nanoparticles”, Applied Catalysis B: Environmental, 97, (2010), 28- 35

FIELDS OF STUDY

Major Field: Chemical Engineering

Area of Interest:

ix

TABLE OF CONTENTS

ABSTRACT …………………………………………………………………………….. ii

DEDICATION…………………………………………………………………………... vi

ACKNOWLEDGMENTS……………………………………………………………… vii

VITA …………………………………………………………………………………... viii

TABLE OF CONTENTS………………………………………………………………….x

LIST OF TABLES.……………………………………………………………………. xvii

LIST OF FIGURES……………………………………………………………………xviii

PART 1: HYDROGEN PRODUCTION via WATER-GAS SHIFT (WGS) REACTION

OVER Fe and Cu-BASED CATALYSTS………………………………………………...1

CHAPTER 1: Introduction to Hydrogen Production via Water-Gas Shift Reaction ……..2

CHAPTER 2: Literature Review.…………………………………………………………4

2.1 Fe-based Catalysts for Water-Gas Shift……………………………………….4

2.2 Cu-based Catalysts for Water-Gas Shift………………………………………6

CHAPTER 3: Experimental Methods……………………………………………………..9

3.1 Steady State Reaction System…………………………………………………9

3.2 Catalyst Characterization…………………………………………………….12

3.2.1 N2- Physisoprtion for BET Surface Area Analysis………………...12 x

3.2.2 Temperature Programmed Reduction (TPR)………………….…...12

3.2.3 Temperature Programmed Oxidation (TPO)………………………13

3.2.4 X-ray Photoelectron Spectroscopy (XPS)…………………………13

3.2.5 X-ray Diffraction (XRD)…………………………………………..14

3.2.6 Diffuse Reflectance Infra-red Fourier Transform Spectroscopy

……………………………………………………….…………...16

CHAPTER 4: Chromium-free Fe-based Water-Gas-Shift Catalysts Prepared Through

Propylene Oxide-Assisted Sol–Gel Technique…………………….……18

4.1 Overview of Cr-free Catalyst Prepared through Propylene Oxide Sol-gel

Technique…………………………………………………………………….18

4.2 Experimental Procedure…………………………………………………...…20

4.2.1 Catalyst Preparation……………………………………………..…20

4.2.2 Catalyst Activity Testing………………………………………..…21

4.2.3 Catalyst Characterization…………………………………………..22

4.3 Results and Discussion……………………………………………………....27

4.3.1 Effect of Gelation Agent on WGS Activity………………………..27

4.3.2 Effect of Cu-loading on WGS Activity…………………………....29

4.3.2.1 Catalyst Activity………………………………………....31

4.3.2.2 X-ray Diffraction………………………………………...34

4.3.2.3 X-ray Photoelectron Spectroscopy………………………41

xi

4.3.2.4 Temperature-Programmed Reduction……………...……48

4.3.2.5 In situ DRIFTS Study Using CO………………………..50

4.3.2.6 Temperature Programmed Re-oxidation………………...53

4.4 Conclusions …………………………………………………………………57

CHAPTER 5: The role of Support Morphology and Impregnation Medium On Water

Gas-Shift Activity of Ceria-Supported Copper Catalysts…………………………….….58

5.1 Overview of Cu/CeO2 Catalyst: Effect of Ceria Morphology and

Impregnation Medium………………………………………………………58

5.2 Experimental Procedure…………………………………………………..…59

5.2.1 Catalyst Preparation……………………………………………….59

5.2.2 Catalyst Activity Testing………………………………………….62

5.2.3 Catalyst Characterization………………………………………….63

5.3 Results and Discussion…………………………………………………...…66

5.3.1 Catalyst Activity…………………………………………………..66

5.3.2 Transmission Electron Microscopy ………………………………69

5.3.3 X-ray Photoelectron Spectroscopy………………………………..71

5.3.4 X-ray Diffraction………………………………………………….77

5.3.5 Temperature-Programmed Reduction…………………………….82

5.3.6 N2O Chemisorption……………………………………………….86

5.3.7 In situ DRIFT During WGS………….……………………………89

xii

5.4 Conclusions …………………………………………………………………94

PART 2: PREFERENTIAL OXIDATION OF CARBON MONOXIDE (PROX) IN

HYDROGEN-RICH STREAMS ………………………………………………………..95

CHAPTER 6: Introduction to Preferential Oxidation of CO (PROX) in Hydrogen…….96

CHAPTER 7: Literature Review……………………………………………………...…98

7.1 Precious Metal Catalysts……………………………………………………..98

7.2 Non-Precious Metal Catalyst……………………………………………...…99

CHAPTER 8: PROX Over CoOx/CeO2 In H2-Rich Streams: Effect of Cobalt Loading

………………………………………………………………....………..101

8.1 Overview of PROX over CoOx/CeO2…….………………………………...101

8.2 Experimental Procedure…………………………………………………….102

8.2.1 Catalyst Preparation………………………………………………102

8.2.2 Catalyst Activity Testing…………………………………………103

8.2.3 Catalyst Characterization..………………………………………..106

8.3 Results and Discussion …………………………………………………….109

8.3.1 Surface Area and Dispersion Measurements……………………..109

8.3.2 Steady State Activity of Performance of CoOx/CeO2…………….112

8.3.3 Oxidation of CO and H2 over CoOx/CeO2………………………..115

8.3.4 Effect of CO2 and water vapor over CoOx/CeO2…………………120

xiii

8.3.5 Catalyst Characterization…………………………………………124

8.3.5.1 XRD…………………………………………………….124

8.3.5.2 TPR……………………………………………………..128

8.3.5.3 XPS……………………………………………………..130

8.3.5.4 In situ DRIFTS during PROX………………………….133

8.3.5.5 In situ X-ray absorption near edge spectra (XANES)

………………………………………………………….139

8.4 Conclusions…………………………………………………………………141

PART 3: CATALYTIC AFTER-TREATMENT OF NATURAL GAS ENGINE

EXHAUST-A DUAL STAGE APPROACH………………………………………..…143

CHAPTER 9: Introduction to Catalytic NOx Reduction……………………………….144

CHAPTER 10: Literature Review: NOx Reduction with Hydrocarbons………………146

CHAPTER 11: The Role of Oxidation Catalyst in Dual Catalyst Bed………………...150

11.1 Overview of Role of Oxidation Catalyst in Dual-Catalyst Bed…………..150

11.2 Experimental………………………………………………………………151

11.2.1 Catalyst Preparation……………………………………………..151

11.2.2 Catalyst Characterization………………………………………..151

11.2.3 Catalyst Activity Testing………………………………………..152

11.2.3.1 NO oxidation over CoOx/CeO2……………………….153

11.2.3.2 Effect of Reduction-to-Oxidation Catalyst Ratio …….154

xiv

11.2.3.3 Hydrothermal Stability: Effect of Co Loading……….154

11.2.3.4 Kinetics of Hydrocarbon Oxidation…………………...155

11.3 Results and Discussion……………………………………………………157

11.3.1 NO Oxidation over CoOx/CeO2…………………………………157

11.3.2 Effect of Reduction- to- Oxidation Catalyst Ratio……………...160

11.3.3 Hydrothermal Stability: Effect of Co Loading………………….164

11.3.4 Kinetics of Hydrocarbon Oxidation……………………………..167

11.3.5 Investigation of Surface Species during CH4-TPD……………...175

11.4 Conclusions………………………………………………………………..183

CHAPTER 12: Effect of Various Engine Exhaust Parameters to Improve the

Hydrothermal Stability during Hydrocarbon-SCR……………………………………..184

12.1 Overview of Improving Hydrothermal Stability…………………………..184

12.2 Experimental and Reaction Studies……………………………………….185

12.2.1 Effect of Hydrocarbon Concentration…………………………..186

12.2.2 Effect of Reaction Temperature………………………………....187

12.3 Results and Discussion……………………………………………...….…188

12.3.1 Effect of Hydrocarbon Concentration……………………….….188

12.3.2 Effect of Reaction Temperature…………………………………200

12.4 Conclusions………………………………………………………………..205

xv

CHAPTER 13: Summary and Future Work …………………………………………...206

13.1 Water Gas-Shift…………………………………………………………...206

13.1 Fe-based Catalyst………………………………………………….206

13.1.2 Cu-based Catalyst……………………………………………….208

13.2 Preferential Oxidation of Carbon Monoxide……………………………...210

13.3 Selective Catalytic Reduction……………………………………………..212

REFERENCES…………………………………………………………...…………….216

APPENDIX A: List of Acronyms………………………………………………………232

APPENDIX B: Sample Calculations………………………………………………...…234

APPENDIX C: Sample Calculations for Methane/Ethane Oxidation Kinetics. ……….236

xvi

LIST OF TABLES

Table 4.1: BET surface area, and pore volume of Fe-Al-Cu catalysts at various Cu- loadings…………………………………………………………………………………..30

Table 4.2: Crystallite size of Fe-Al-Cu catalysts at various Cu-loadings calculated from the XRD patterns through Scherrer equation…………………………………………….37

Table 5.1: Specific surface areas of CeO2-based samples calculated using BET method

……………………………………………………………………………………...…….61

Table 5.2: XPS analysis of the catalysts……………………………..…………………..76

Table 5.3 Cu-dispersion values estimated from N2O chemisorption…………………….88

Table 8.1: Specific surface areas and Co dispersions over CoOx/CeO2 catalysts….…..111

Table 8.2: Activation energies of CO and H2 oxidation over CoOx/CeO2 catalysts……117

Table 8.3: CO conversion (%) over CoOx/CeO2 in the presence of H2O and CO2. Values in parentheses show O2 selectivity to CO2 (%) over CoOx/CeO2 in the presence of H2O and CO2…………………………………………………………………………………122

Table 11.1: Summary of kinetic parameters during CH4 and C2H6 oxidation over oxidation, reduction and dual catalyst bed catalyst…………………………………….171

Table 11.2: Summary of DRIFTS study during CH4-TPD over oxidation, reduction and dual catalyst bed catalyst ………………………………………………...…………….182

xvii

LIST OF FIGURES

Figure 3.1: Schematic of Steady State Reaction System for Water-Gas Shift…………..11

Figure 4.1: WGS activity (CO% conversion) of Fe-Al-Cu catalysts prepared by propylene oxide-assisted sol-gel method with different Cu loadings at (a) S/C = 1 and (b) S/C = 2.

Inset: time-on-stream performance of Fe-Al-Cu (Fe/Cu=5) at 400o C at S/C = 1. Reaction

Conditions: [CO] = 10%, [H2O] = 10% or 20%, [CO2] = 5%, [H2] = 7.5% and balance

3 -1 -1 N2, WHSV = 0.06 m (g cat) h , P = 1 atm……………………………………………33

Figure 4.2: XRD patterns of Fe-based samples. (a) Fe-only (b) Fe-Cu (Fe/Cu=10) (c) Fe-

Al-Cu (Fe/Cu = 10), (d) Fe-Al-Cu (Fe/Cu = 5) and (e) Fe-Al-Cu (Fe/Cu = 2)………….38

Figure 4.3: Transmission electron micrograph (TEM) of Fe-Al-Cu (Fe/Cu = 5)………..39

Figure 4.4: XRD patterns of Fe-Al-Cu samples reduced in syngas mixture

(a) Fe/Cu=10, (b) Fe/Cu=5 and (c) Fe/Cu=2……………………...……………………..40

Figure 4.5: X-ray photoelectron spectra of Cu 2p region of pristine and reduced Fe-Al-Cu samples: (a) Fe/Cu = 10, (b) Fe/Cu = 5 and (c) Fe/Cu = 2……………………………....44

Figure 4.6: X-ray photoelectron spectra of Fe 2p region of pristine and reduced Fe-Al-Cu samples: (a) Fe/Cu = 10, (b) Fe/Cu = 5 and (c) Fe/Cu = 2……………………….……...47

Figure 4.7: Temperature-programmed reduction profiles of Fe-Al-Cu catalysts………..49

Figure 4.8: DRIFTS study of CO interaction with Fe-Al-Cu (Fe/Cu=5)………………..52

xviii

Figure 4.9: Temperature-programmed re-oxidation of reduced Fe-Al-Cu samples at different ramp rates: (a) Fe/Cu = 10, (b) Fe/Cu = 5 and (c) Fe/Cu = 2. Insets show determination of activation energies of re-oxidation following Kissinger method……...56

Figure 5.1: Activity of (a) CeO2-NP, (b) CeO2-NR based WGS catalysts: (●) CeO2, (■)

Cu/CeO2(E) and (▲) Cu/CeO2(T) and (c) time-on-stream activity of Cu/CeO2-NP(E) at

400°C. Reaction conditions: [CO] = 10%, [H2O] = 10%, [CO2] = 5%, [H2] = 7.5% and

3 -1 -1 balance N2, WHSV = 0.06 m (g cat) h , P = 1 atm…………………………………..68

Figure 5.2: TEM images of (a) CeO2-NR and (b) CeO2-NR at higher magnification, (c)

CeO2-NP and (d) CeO2-NP at higher magnification…………………………………….70

Figure 5.3: Ce 3d region of the X-ray photoelectron spectra of (a) CeO2-NP, (b) CeO2-

NR, (c) Cu/CeO2-NP(E), (d) Cu/CeO2-NP(T), (e) Cu/CeO2-NR(E) and (f) Cu/CeO2-

NR(T)……………………………………………………………………………………74

Figure 5.4: Cu 2p region of the X-ray photoelectron spectra in (a) Cu/CeO2-NP(E), (b)

Cu/CeO2-NP(T), (c) Cu/CeO2-NR(E) and (d) Cu/CeO2-NR(T)…………………………75

Figure 5.5: X-ray diffraction patterns of (a) CeO2-NR, (b) CeO2-NP, (c) Cu/CeO2-NR(E)

(d) Cu/CeO2-NR(T) (e) Cu/CeO2-NP(E) and (d) Cu/CeO2-NP(T)………………………80

Figure 5.6: X-ray diffraction patterns collected during in-situ reduction with 5%H2/N2 of

(a) Cu/CeO2-NP(E), (b) Cu/CeO2-NR(E) and (c) Cu/CeO2-NP(T)

……………………………………………………..…………………………………….81

xix

Figure 5.7: Temperature-programmed reduction profiles of (a) CeO2-NR, (b) CeO2-NP,

(c) Cu/CeO2-NR(T) (d) Cu/CeO2-NR(E) (e) Cu/CeO2-NP(T) and (f) CeO2-NP(E) in 5%

H2/He. Inset: Temperature-programmed reduction profile of bulk CuO………………...85

Figure 5.8: In-situ DRIFT spectra collected during WGS reaction over Cu/CeO2-NP(E)

(a)high wavenumber region and (b)low wavenumber region……………………………92

Figure 5.9: In-situ DRIFT spectra collected during WGS reaction over Cu/CeO2-NR(E)

(a)high wavenumber region and (b)low wavenumber region……………………………93

Figure 8.1: Schematic of Steady State Reaction System for PROX……………………105

Figure 8.2: PROX activity and selectivity of () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10% CoOx/CeO2. (Reaction conditions: 1% CO, 1% O2, 60% H2 in He.

3 -1 -1 WHSV=15,000 cm (g cat.) h ). Inset: Time-on-stream CO conversion and O2

o selectivity to CO2 over 10% CoOx/CeO2 at 175 C under above mentioned reaction conditions……………………………………………………………………………….114

Figure 8.3: CO conversion over () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10%

CoOx/CeO2 during CO oxidation. (Reaction conditions: 2% CO and 2% O2 in helium).

Inset shows Arrhenius plot for determination of the activation energy for CO oxidation

…………………………………………………………………………………….…….118

Figure 8.4: H2 conversion over () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10%

CoOx/CeO2 during H2 oxidation. (Reaction conditions: 2% H2 and 2% O2 in argon) Inset shows Arrhenius plot for determination of the activation energy for H2 oxidation

…………………………………………………………………………………………..119

xx

Figure 8.5: CO conversion over () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10%

CoOx/CeO2 during water-gas-shift. (Reaction conditions: 1% CO and 1% H2O in helium)

…………………………………………………………………………………………123

Figure 8.6: X-ray diffraction patterns of (a) CeO2, (b) 1% CoOx/CeO2, (c) 2%

CoOx/CeO2, and (d) 10% CoOx/CeO2…………………………………………………126

Figure 8.7: In situ X-ray diffraction patterns collected during reduction of 10%

CoOx/CeO2 with 5% H2/He…………………………………………………………….127

Figure 8.8: Temperature-programmed reduction profiles of (a) CeO2, (b) 1% CoOx/CeO2,

(c) 2% CoOx/CeO2 and (d) 10% CoOx/CeO2 in 5% H2/He……………...……………..129

Figure 8.9: X-ray photoelectron spectra in the Ce 3d and Co 2p regions of (a) 1%

CoOx/CeO2 (b) 2% CoOx/CeO2 and (c) 10% CoOx/CeO2…………………………...…132

Figure 8.10: In situ DRIFT spectra collected during preferential oxidation of CO over 2%

CoOx/CeO2 (a) high wavenumber region and (b) low wavenumber region……………137

Figure 8.11: In situ DRIFT spectra collected during preferential oxidation of CO over

10% CoOx/CeO2. (a) high wavenumber region and (b) low wavenumber region. Inset: spectra taken at 125°C after the reaction gases are flushed……………………….….138

Figure 8.12: Normalized Co K-edge XANES spectra of 10% CoOx/CeO2. (a) Pristine and

o (b) In situ under PROX at 175 C (Spectra for Co3O4 is included for comparison)…...140

Figure 11.1: Schematic of Steady State Reaction System for HC-SCR………………..156

xxi

Figure 11.2: NO2 yield during NO oxidation over 2%CoOx/CeO2 () and

10%CoOx/CeO2 (). Reaction conditions: 1000ppm NO, 10% O2; 1 atm; GHSV :

-1 100,000 h , (---) thermodynamic equilibrium conversion of NO to NO2…….………..159

Figure 11.3: (a) NOx conversion (b) CH4 conversion over dual catalyst bed as a function of Pd/SZ-10%CoOx/CeO2 ratios, 1:0 ( ), 2:1(), 4:1() and 8:1() under simulated lean exhaust conditions. Reaction conditions: [NO2]=180 ppm, [CH4]=1737 ppm,

[C2H6]= 208 ppm, [C3H8]= 104 ppm, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%; 1 atm;

GHSV : 32,000 h-1……………………………………………………………………..162

Figure 11.4: NOx (bold) and CH4 conversion (void) over Pd/SZ: CoOx/CeO2=8:1 at different cobalt loadings; 2%CoOx/CeO2 (,) and 10%CoOx/CeO2 (,) under simulated lean exhaust. Reaction conditions: 180 ppm [NO2]=180 ppm, [CH4]=1737 ppm, [C2H6]= 208 ppm, [C3H8]= 104 ppm, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%; 1 atm; GHSV : 32,000 h-1…………………………………………………………….….163

Figure 11.5: Reversibility of the effect of water vapor on (a) NOx (b) CH4 conversion over Pd/SZ: CoOx/CeO2=8:1 at different cobalt loadings 2%CoOx/CeO2 (,) and

10%CoOx/CeO2 (,) under simulated lean exhaust. Reaction conditions: [NO2]=180 ppm, [CH4]=1737 ppm, [C2H6]= 208 ppm, [C3H8]= 104 ppm, [CO]= 650 ppm, [CO2]=

o -1 6.5%, [O2]=10%, [H2O]= 0% or 7%; 450 C; 1 atm; GHSV : 32,000 h …………..…166

xxii

Figure 11.6: (a) Rate of methane oxidation Vs. methane concentration and (b) fractional ethane conversion Vs. W/FA,0 over 2%CoOx/CeO2. Inset: comparison of power-law model predictions for hydrocarbon oxidation rate against experimentally determined rates over 2%CoOx/CeO2…………………………………………………………………….172

Figure 11.7: (a) Fractional methane conversion Vs. W/FA0 and (b) fractional ethane conversion Vs. W/FA0 over Pd/SZ. Inset: comparison of power-law model predictions for hydrocarbon oxidation rate against experimentally determined rates over Pd/SZ

…………………………………………………………………………………………..173

Figure 11.8: (a) Fractional methane conversion Vs. W/FA0and (b) fractional ethane conversion Vs. W/FA0 over Pd/SZ: 2%CoOx/CeO2= 8:1. Inset: comparison of power-law model predictions for hydrocarbon oxidation rate against experimentally determined rates over Pd/SZ: 2%CoOx/CeO2= 8:1……………………………………………………….174

Figure 11.9: In situ DRIFT spectra collected during CH4-TPD over i) 2% CoOx/CeO2, ii)

Pd/SZ and iii) 1:8 Dual-catalyst bed 2%CoOx/CeO2-PdSZ a) high wavenumber region b) low wavenumber region………………………………………………………………..181

Figure 12.1: Effect of hydrocarbon concentration () CHx=2050 ppm, () CHx=2500 ppm and () CHx=3050 ppm on (a) NOX and (b) CH4 conversion during NO2 reduction in the presence of H2O. Reaction conditions: [NO2]= 180 ppm, [CHx]= varying, [CO]=

o -1 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=7% ; 450 C ; 1 atm ; GHSV ̴ 32,000h

…………………………………………………………………………………………..191

xxiii

Figure 12.2: Effect of hydrocarbon concentration on hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b)

CH4 conversion. ()[CHx]=2050 ppm, ()[CHx]= 2500 ppm. Reaction conditions:

[NO2]= 180 ppm, [CHx]= varying, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%,

o -1 [H2O]=10% ; 450 C ; 1 atm ; GHSV ̴ 32,000h ………………………………………193

Figure 12.3: The effect of either a methane only hydrocarbon feed or a mixed higher hydrocarbon feed on the hydrothermal stability of a dual catalyst bed Pd/SZ and

CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b) CH4 conversion. ()

[CH4]= 2050 ppm, ()[CHx]= 2050 ppm. Reaction conditions: [NO2]=180 ppm, [CHx]=

o varying, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ; GHSV

-1 ̴ 32,000h …………………………………………………………………………...…..195

Figure 12.4: The effect of different hydrocarbons, as a function of equal carbon basis, on the hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b) CH4 conversion. () [CH4]= 2050 ppm, ()

[C2H6]= 1025 ppm, ()[C3H8]= 683 ppm. Reaction conditions: [NO2]=180 ppm, [CHx]=

o varying, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ; GHSV

-1 ̴ 32,000h ……………………………………………………………………………….197

xxiv

Figure 12.5: The effect of supplementary ethane addition on the hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b) CH4 conversion. () [CHx]= 2050 ppm, () [CHx]= 2050 ppm +

100ppm C2H6. Reaction conditions: [NO2]=180 ppm, [CHx]= varying, [CO]= 650 ppm,

o -1 [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ; GHSV ̴ 32,000h

………………………………………………………………………………………….199

Figure 12.6: Effect of reaction temperature () 425oC () 450oC, and () 475oC on (a)

NOX and (b) CH4 conversion during NO2 reduction in the presence of H2O. Reaction conditions: [NO2]=180 ppm, [CHx]= 2500 ppm, [CO]= 650 ppm, [CO2]= 6.5%,

-1 [O2]=10%, [H2O]=7% ; 1 atm ; GHSV ̴ 32,000h ………………...…………………..201

Figure 12.7: The effect of temperature on the hydrothermal stability of a dual catalyst bed

Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b) CH4

o o o conversion. () 450 C, () 475 C, () 500 C. Reaction conditions: [NO2]=180 ppm, [CHx]= 2050 ppm, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 1 atm ;

-1 GHSV ̴ 32,000h ……………………………………………...……………………….204

Figure 13.1: Various nano-morphologies of ceria prepared through hydrothermal- precipitation (Image Courtesy: Bing Tan)……………………………………...………210

xxv

PART 1:

HYDROGEN PRODUCTION via WATER-GAS SHIFT (WGS) REACTION

OVER Fe and Cu-BASED CATALYSTS

1

CHAPTER 1

INTRODUCTION TO HYDROGEN PRODUCTION via WATER-GAS-SHIFT

REACTION

Hydrogen production via gasification of coal and biomass is a promising and clean technology to produce hydrogen, electricity and other valuable products. Coal is the nation’s most abundant fossil fuel resource, which can last up to several hundred years. It accounts for around 50% of electricity production in United states [1]. On the other hand, power generation via biomass gasification has received a considerable attention in recent years to replace fossil fuels. The current estimated biomass worldwide energy potential is around 100 EJ/year (1EJ= 1018 J) [2], which accounts for 25% of the total energy requirement. However, Fischer, et al. [3] and Hoogwijk, et al. [4] estimated that it will increase to 675-1135 EJ/year in next 40-50 years.

WGS unit remains as an essential step in the overall integrated gasification process for hydrogen generations from syngas regardless of the gasification feedstock and operating conditions. Typically, the syngas mixture derived from gasification processes consists of CO, CO2, H2O, H2, CH4 and traces of other hydrocarbons. It may also content traces of H2S and COS depending upon feedstock, operating conditions and clean-up processes.

2

Water-gas-shift reaction (CO+H2O = CO2 + H2, ΔH = −40.6 KJ/mol) is a slightly exothermic reaction which is favored at low reaction temperature and high steam to carbon ratio (S/C) [5]. The current WGS unit is operated in a “two-stage” mode, namely, high-temperature WGS (HT-WGS) and low-temperature WGS (LT-WGS).

Commercially, HT-WGS operated in the temperature range of 320-450oC over Fe-Cr catalyst, which brings down CO concentration from 10% to 3-5%. HT-WGS is followed

o by LT-WGS, operated around 200-250 C over Cu/ZnO/Al2O3 catalyst which can further decrease CO concentration to 0.5-1% [6-9]. A “two-stage” operation mode is a consequence of several drawbacks of the current catalytic system. For example, Fe-Cr catalyst is practically inactive at low reaction temperature whereas, Cu/ZnO/Al2O3 has sintering issues during HT-WGS.

* Information in this chapter is adapted/taken from P. Gawade, B. Mirkelamoglu, B. Tan and U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 321, (2010), 61-70

3

CHAPTER 2

LITERATURE REVIEW

2.1 Fe-based Catalysts for Water-Gas Shift:

Fe-Cr catalyst is used at commercial level during HT-WGS due to its high stability and reasonable activity. Cr acts as a structural promoter and stabilizer to prevent the particle sintering during high temperature operations. Commercially, 8-14% (wt%) of Cr is incorporated in Fe catalyst as a stabilizer [5, 10, 11]. In Fe-Cr catalyst, iron is mostly in the form of α-Fe2O3 (hematite), which gets converted to WGS active Fe3O4

(magnetite) phase under the reducing environment. However, over-reduction of active phase (Fe3O4) may occur due to exothermicity of the reaction that may lead to the formation of less active metallic Fe [5, 11]. In addition, metallic Fe is believed to promote the side reaction such as, methanation reaction and carbon formation [12].

Hence higher amount of steam is usually used during WGS over Fe-Cr to avoid the over- reduction. Cu-promoted Fe-Cr catalysts have shown promising WGS performance in the past even in the presence of low steam content without forming metallic Fe [10, 13].

Andreev and co-workers reported that copper promoted Fe-Cr had much better WGS activity than cobalt and zinc promoted catalysts [14, 15]. However, as mentioned earlier,

4

Cr+6 in Fe-Cr is highly carcinogenic and posses environmental and health hazards.

Therefore, the substitution of Cr from Fe-based catalysts is essential.

In the past, several attempts have been reported to substitute Cr from Fe-Cr. The oxides of Pb, La, Ca, Th, Al, Mg, Mn and Zr have shown promising potential to replace

Cr [11, 16]. 10% (wt%) PbO doped in Fe-catalyst showed comparable WGS activity to that of commercial Fe-Cr. However, this formulation failed in the presence of sulfur [11].

In the quest of replacing Cr, de Araújo, G.C. and do Carmo Rangel, M. [17] showed that aluminum could be a promising option for Fe-based catalyst. In their findings, a small addition of copper (3wt%) and aluminum significantly improved Fe-catalyst WGS activity even at low steam to gas ratio (S/G=0.4). In addition the catalyst properties and activity was similar to that of commercial Fe-Cr-Cu catalyst. The same research group [5] extended this work by replacing Al with thorium (Th) in Fe-Cu catalyst. In theses findings, thorium acted as a stabilizer and prevented the particle sintering at high reaction temperatures whereas; Cu played the role of structural promoter. The increase in surface area due to thorium addition in Fe-Cu-Th resulted in higher WGS activity than commercial Fe-Cu-Cr catalyst at 370oC.

Ozkan research group extensively studied the effect of incorporation of Al, Ga and Mn as replacement for Cr in Fe-Cr catalyst as well as catalyst preparation variables and methods [18]. The catalyst was prepared using co-precipitation and impregnation method. In these finding, Al emerged as a potential substitute for Cr and the further catalyst improvements were achieved by addition of Cu as a promoter. In the following years, Ozkan, et al. [19, 20] extended this Cr-free work and found that Fe-Al-Cu catalyst prepared using sol-gel method had much improved activity due to uniform distribution of

5 copper in iron oxide matrix than its previous counterpart prepared using precipitation method. Fe-Al-Cu catalyst was prepared using NaOH as a gelation agent and its performance was significantly higher than commercial Fe-Cr-Cu catalyst over a temperature range of 250-400oC. However, as mentioned earlier in spite of excellent

WGS performance, catalyst preparation method was tedious, required critical pH, several washing steps. Therefore, in the present work we report “single-step” sol-gel preparation method using propylene oxide as a gelation agent to obtain highly active and stable Fe-

Al-Cu catalyst.

2.2 Cu-based Catalysts for Water-Gas Shift:

Cu-based catalysts are as extensively investigated as Fe-based catalysts for WGS reactions. Cu supported over various oxides such as, Al2O3, ZrO2, SiO2-Al2O3, CeO2,

MgO, ZnO-Al2O3, Cr2O3, CeO2-ZrO2, CeO2-Al2O3 and Ce(La)Ox have been reported as

WGS catalysts [21-33]. Among these systems, Cu/CeO2 is probably the most widely investigated formulation due to unique nature of CeO2.

CeO2 is widely reported as a support and catalyst for various applications such as

WGS [22, 27, 28, 32, 34-38], preferential oxidation [39-43], NO reduction [44] and methane oxidation [45] due to its excellent oxygen storage capacity and redox property

[35]. It is also known to improve the catalyst reducibility and prevent particle sintering

[46]. Both, precious/noble [38, 47-49] as well as non-precious metals [21, 27, 34, 50,

51] supported over ceria are widely reported as WGS catalyst.

6

Excellent WGS activity and stability has been reported over Cu/CeO2 catalyst due to close interaction between Cu and CeO2 that resulted into improved catalyst reducibility. The formation of Cu0/Cu+1 over ceria support has been considered as an active site for WGS whereas; ceria plays a direct or indirect role in overall reaction process [27, 32, 47, 48, 51]. The exceptional oxygen storage capacity (OSC) of ceria is known to assist the redox cycle of WGS reaction which can be further improved by the incorporation of promoters such as, ZrO2 and La2O3 [46, 52]. The studies [27, 46] have shown the addition of 8% (wt%) La2O3 in Cu-Ce(La)Ox can significantly improved WGS activity and thermal stability up to 600oC.

Various catalyst preparation techniques have been reported for ceria and Cu/CeO2 catalysts preparation such as, co-precipitation, hard-template method, urea gelation co- precipitation, hydrothermal and sol-gel methods ([34, 46, 50, 53] and references therein).

Da Silva and Assaf [21] reported that highly active 5%Cu-CeO2 catalyst could be prepared through hydrothermal and urea co-precipitation method. In their findings, high surface area and better copper dispersion was believed to provide superior WGS performance which can be further enhanced by the preparation of nano-structures of ceria

(cited in [34]).

In recent years, numerous nano-sized morphologies of ceria are studied and debated for structure sensitive reactions such as CO oxidation and preferential oxidation of CO (PROX). It is believed that ceria morphologies affect the redox property and OSC

[54-59]. Mai et al. [54] showed that ceria nanorods and nanocubes were more active towards CO oxidation. According to their claim, the more exposed planes [1 1 0]/[1 0 0] of ceria were responsible for higher oxygen storage. Tana, et al. [60] further supported

7 these finding in similar nano-shaped ceria study for CO oxidation. Pan et al. [57] showed that CeO2 nano-wires prepared through hydrothermal method had better CO oxidation performance than nano-particles and nano-cubes. Yi et al. [61] extended ceria morphology work over Au/CeO2 catalyst for PROX. In their observations, 1%Au/CeO2

(particles and rods) had better PROX activity and O2 selectivity than 1%Au-CeO2-cubes in the temperature range of 30-100oC.

Although broad literature is available on ceria morphology for CO oxidation processes, few studies have been reported on the effect ceria morphology for WGS system. Si and Flytzani-Stephanopoulos [62] have reported the effect of ceria

o morphology over 1%Au/CeO2 catalyst for WGS system over 150-350 C. In these findings, 1%Au-CeO2-nanorods had better WGS activity due to more exposed [1 0 0]/[1

1 0] crystal plane which assisted in stabilizing gold. To the best of our knowledge, Ozkan research group will be one of the first to report and investigate the effect of various ceria morphologies over Cu/CeO2, which will be discussed in details in the subsequent section.

* Information in this chapter is adapted/taken from 1. P. Gawade, B. Mirkelamoglu, B.

Tan and U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 321, (2010), 61-70 2.

P. Gawade, B. Mirkelamoglu, U.S. Ozkan, Journal of physical Chemistry C 114 (2010)

18173-18181.

8

CHAPTER 3

EXPERIMENTAL METHODS

3.1 Steady State Reaction System

The schematic of steady state reaction system for WGS is shown in Figure 3.1.

The system was equipped with four independent mass controllers (Brooks Instruments,

5850E) to regulate the gas flow rates. Two micron particle filters were installed before every mass flow controllers (MFC) to avoid any particulate contamination reaching to

MFC. In addition, either Oxy-trap© or Drierite© was installed in some of the lines (Ar,

H2 and N2) to remove moisture or oxygen contamination. Water vapor was introduced to the system by bubbling nitrogen through heated de-ionized water bubbler. The temperature of water bubbler was monitored and controlled by temperature controller.

The gas lines were heated (c.a. 120o C) to avoid any condensation of water vapor in the lines.

All the steady-state reactions were conducted in a fixed-bed flow reactor using ¼“

O.D. stainless steel rector tube. The catalyst was packed and held inside the reactor using stainless steel frit and quartz wool plug. The reactor was placed in homemade resistively

9 heated furnace using Omega (model CS232) PID temperature controller and the reaction temperature was monitored using Omega K-type thermocouple.

For most of the WGS reactions, 100 mg of catalyst sample was used for each run and the experiments were carried out at a WHSV (weight hourly space velocity) of 0.06 m3 (g cat)-1 h-1. In addition, some steady-state experiments were conducted on equal surface area to ensure that the observed differences were not simply due to differences in surface areas. In order to assure that the activity testing experiments were run in the kinetically controlled regime, reaction conditions were maintained away from equilibrium in all of the activity testing experiments.

Prior to catalytic activity testing, catalyst samples were pretreated in-situ

o o in N2 (67.5 sccm) for 30 minutes at 350 C and then reduced at 350 C for 2 hours in syngas mixture (10% CO, 10% (or 20%) H2O, 7.5% H2, 5% CO2 balanced with 67.5%

o N2). Finally, reaction studies were performed over the wide temperature range (250 C-

400o C) in the above mentioned syngas mixture.

The feed as well as the reactor effluents were analyzed by an on-line gas chromatograph (Shimadzu, GC-14A series) equipped with a thermal conductivity detector (TCD). Argon was used as the carrier gas and the gas separation was performed using a molecular sieve column 13X (5 ft x 1/8 in. SS, 60/80 mesh) and porapak Q column (12 ft x 1/8 in. SS, 80/100 mesh).

10

Figure 3.1: Schematic of Steady State Reaction System for Water-Gas Shift

11

3.2 Catalyst Characterization

3.2.1 N2- Physisorption for BET Surface Area Analysis

The specific surface areas of the samples were measured on either Micromeritics

ASAP 2010 or 2020 accelerated surface area and porosimetry instruments using nitrogen physisorption technique. Prior to analysis, samples were degassed overnight at 130oC under a vacuum of 3µmHg. The specific surface areas were determined by the Brunauer-

Emmett-Teller (BET) method using the adsorption branch of the isotherm. Pore volumes of the samples were determined by the Barret-Joiner-Halenda (BJH) method using the adsorption branches of the N2 physisorption isotherms.

3.2.2 Temperature Programmed Reduction (TPR)

Temperature programmed reduction (TPR) is a bulk characterization technique, which assists to identify the reduction temperature of the sample, metal-support interaction, presence of various components and oxidation states of the sample.

For H2-TPR experiments, usually 50-100 mg of calcined catalyst sample was packed in a 0.25” O.D. U-tube quartz reactor using quartz wool. The catalyst was pretreated in He at 350o C for 30 minutes followed by cooling at room temperature in He.

Then either 5% H2/He or 5% H2/N2 was introduced at room temperature before ramping the temperature at 10° C/min. Hydrogen consumption during the temperature ramp was monitored either using a thermal conductivity detector (TCD) or an online mass

12 spectrometer (Cirrus II, MKS Instruments, 1–300 amu) operated in selective ion detection mode.

3.2.3 Temperature Programmed Oxidation (TPO)

TPO experiments were conducted to derive re-oxidation kinetic parameters over

Fe-Al-Cu catalysts during WGS using Kissinger’s method [63]. This method correlates the temperature at which the maximum rate of oxidant consumption occurs with the activation energy for re-oxidation through a series of temperature-programmed experiments ran at different heating rates. The following equation is used to calculate the

2 activation energy for re-oxidation, ln(φ /Tm ) = (−E /RTm ) + Constant .

Where, Φ is the heating rate and Tm is the temperature at which the oxidant consumption rate is a maximum€ . The details of this study are presented in Section 4.3.2.6

3.2.4 X-ray Photoelectron Spectroscopy (XPS)

XPS is a surface characterization technique, which can provide elemental compositions, oxidation state and dispersion of one component over other [64]. This technique works on photoelectric effect in which desired atom absorbs a photon (hν) and the photoelectrons (core/valance) with binding energy Eb are ejected with kinetic energy

Ek. The following equation is used to calculate binding energy of photoelectrons:

Ek = hν- Eb- φ where; φ is a work function

13

XPS analysis was carried using Kratos AXIS Ultra X-ray photoelectron spectrometer operated under high ultra-high vacuum. The monochromatized Al Kα

(1486.7eV) X-ray source operated at 13 kV and 10 mA. When necessary, the charge neutralizer was operated at a current of 2.1A and a filament bias of 1.3V. For each sample a survey scan was collected from 1400 to 0 eV. Concurrent sweeps of Cu 2p, Fe

2p, Ce 3d, O 1s and C 1s envelopes were collected at an electron pass energy of 20 eV. In addition, the collected spectra were corrected using C 1s envelope located at 284.5 eV.

Analysis of the collected data was performed on XPS Peak 4.1.

3.2.5 X-ray Diffraction (XRD)

XRD is a bulk characterization technique which is used to identify the crystal structures and average crystal sizes. It can be also operated in situ to monitor the changes in crystal structures as a function of temperature and reaction environment. This technique is based on constructive interference of scattered X-ray radiations by the sample [64]. The following Bragg relation is used to derive the lattice spacing:

n λ= 2 d sinθ

Where; n= Order of reflection,

λ= Wavelength of X-rays,

θ= Angle between incoming X-rays and normal to lattice plane,

d= Distance between lattice planes

14

Furthermore, Scherrer formula can be used to calculate the average crystal size using the

peak width

K" < L >= #cos$

Where; = Average crystal size

! K= Constant (usually 1)

λ= Wavelength of X-rays

θ= Angle between incoming X-rays and normal to lattice plane,

β= Peak Width

XRD patterns of the pristine samples were collected on a Rigaku X-ray

diffractometer (X-ray source: Cu Kα radiation, λ=1.5418 Å) operated at 40 kV and 25

mA. The diffraction patterns were collected in the range of 20o to 60o Bragg angle values.

Capillary XRD was used to understand the transformation of crystal phases

during reduction. For capillary XRD analysis of Fe-Al-Cu samples, Bruker D8 Advance

X-ray diffractometer equipped with a capillary sample holder was used. Following in situ

reduction treatment inside the reactor (10% CO, 10% H2O, 7.5% H2, 5% CO2 balanced

o with 67.5% N2 for 2 h at 350 C), the samples were transferred to an Ar-purged glove box

that contains minimal residual air (< 5ppm). The samples were transferred to capillary

sample tubes (0.2 mm O.D.) and the ends of the capillary tube were sealed inside the

glove box with glass sealant. The samples were then transferred air-free to capillary

sample holder accessory on the diffractometer. The diffraction patterns were collected in

15 the 20°-70° Bragg angle range using a Cu Kα X-ray source (λ=1.5418 Å) operated at 40 kV and 50 mA.

For in situ analysis of Cu/CeO2 samples under a flow of 5% H2/N2 (15 ccm), the same

Bruker D8 Advance X-ray diffractometer equipped with an Anton Paar HTK1200 oven was used. The samples were heated from room temperature to 450°C under a flow of 5%

H2/N2 (15 ccm) and diffraction patterns were collected at regular temperature intervals.

3.2.6 Diffuse Reflectance Infra-red Fourier Transform Spectroscopy (DRIFTS)

DRIFTS is a vibrational spectroscopy technique used to identify the adsorbed species on catalyst surface and their interactions. It is an absorption technique, which can be operated in diffuse reflectance mode for the powder samples. For vibrations to be IR active dipole moment of the molecule must change during the vibrations. The FT-IR spectrometer (Thermo Nicolet 6700) equipped with an MCT detector and a DRIFTS chamber with ZnSe windows was used for DRIFTS study. Liquid nitrogen was used throughout the experiments to cool the MCT detector. The DRIFTS spectra were collected in the mid-IR range at a nominal resolution of 4cm-1 and the presented spectra were averaged over 500 scans.

For Fe-Al-Cu catalysts, the samples were reduced in reactor system under the conditions mentioned in Section 3.1. The reduced samples then were recovered under nitrogen atmosphere and transferred to DRIFTS cell. The details of reaction studies over

Fe-Al-Cu catalyst are given in Section 4.3.2.5. For Cu/CeO2 samples, spectra were

16 collected during WGS reaction in 1%CO and 1%H2O. The detailed description of which can be found in Section 5.3.7

* Information in this chapter is adapted/taken from 1. P. Gawade, B. Mirkelamoglu, B.

Tan and U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 321, (2010), 61-70. 2.

P. Gawade, B. Mirkelamoglu, U.S. Ozkan, Journal of physical Chemistry C 114 (2010)

18173-18181.

17

CHAPTER 4

Cr-FREE Fe-based WATER-GAS SHIFT CATALYST PREPARED THROUGH

PROPYLENE OXIDE-ASSISTED SOL-GEL TECHNIQUE

4.1 Overview of Cr-free Catalyst Prepared through Propylene Oxide Sol-gel Technique

The sol-gel method for catalyst preparation is considered to be more effective in terms of offering improved homogeneity, high purity, low calcination temperature and ability to tailor the desired metal oxide phase over traditional preparation methods [65,

66]. The sol-gel method is basically a two-step process in which metal salt undergoes hydrolysis reaction followed by the condensation to form nano-scale structures.

Formation of iron oxide matrix using sol-gel techniques in which the gel formation was induced by heating, adding a base or a gelation agent was reported (cited in [67]). In the past, several researchers synthesized nano-scale Fe as well as spinel oxides using epoxides such as ethylene oxide and propylene oxide [66-68]. These gelation agents were reported to provide high homogeneity, high purity and high surface area [66-68]. The ratio of epoxide to metal salt [67, 68] as well as the choice of solvent play a critical role in gel formation rate and determining the particle size. The epoxides are believed to act as proton scavengers that cause iron to undergo hydrolysis and condensation processes to

18 form an iron oxide matrix [67]. However, it is worthwhile to note that ethylene oxide is highly carcinogenic and toxic which defeats the purpose of environmentally benign catalyst preparation methods. Citric acid is also another gelation agent, which is widely studied for the preparation of highly dispersed mixed-oxides such as bismuth oxide, cerium-zirconium oxide and spinel oxides of cobalt [65, 69-72].

In this chapter, we report development of Cr-free Fe-based formulations using two gelation agents, namely citric acid and propylene oxide. The developed sol-gel preparation methods using these gelation agents are simpler, less time consuming and above all, environmentally benign. Propylene oxide based Fe-Al-Cu system demonstrated significant WGS performance which was comparable to the previously reported Fe-Al-

Cu sol-gel combination [19, 20]. This propylene oxide based gelation technique is a simple “one-pot” technique and requires neither catalyst washing nor critical pH adjustments. Copper is an essential promoter in the Fe-based catalysts [5, 10, 13, 19], therefore, the effect of Cu-loading for catalysts prepared by this technique has also been investigated. In addition to steady-state experiments for WGS activity evaluation, numerous catalyst characterization techniques such as X-ray diffraction (XRD), temperature-programmed reduction (TPR) and oxidation (TPO), X-ray photoelectron spectroscopy (XPS), diffuse reflectance infrared Fourier transform spectroscopy

(DRIFTS) and Brunauer-Emmett-Teller method (BET) were implemented in order to evaluate the surface as well as bulk properties of the catalysts which could be correlated to their WGS performance.

19

4.2 Experimental Procedure

4.2.1 Catalyst preparation

Fe-based catalysts with Cu and Al employed as promoters were prepared through a sol-gel route using propylene oxide and citric acid as gelation agents. Aluminum nitrate

[Al(NO3)3.9H2O] (crystals, Fisher), iron (III) nitrate nanohydrate [Fe(NO3)3.9H2O]

((98+%) ACS reagent, Sigma-Aldrich) and copper (III) nitrate trihydrate [CuN2O6.3H2O]

(Fluka) were used as metal precursors in catalyst preparation.

For preparation of catalysts with the propylene oxide route, metal salt(s) in quantities to yield the desired atomic ratios of Fe, Al and Cu in the final catalyst were dissolved in ethanol at ambient temperature. Propylene oxide was added at once to this solution in excess quantities while continuously stirring the solution. The addition of propylene oxide resulted in rapid heat evolution accompanied by gel formation. The gel was kept at 50°C for 4 hours and then dried overnight at 110°C in a convection oven.

Dried samples were ground to a fine powder and calcined in air for 4 hours at 450°C

(heating rate: 10°C/min). The samples were crushed to a fine powder following calcination. The propylene oxide method was utilized for preparation of Fe-only, Fe-Cu bi-metallic and Fe-Al-Cu ternary samples. The atomic ratio of Fe to Cu in Fe-Cu bi- metallic catalyst was 5, whereas Fe to Cu ratio was varied in Fe-Al-Cu catalysts

(Fe/Cu=10, 5 and 2) while keeping Fe/Al=10 constant.

Another gelation agent that was utilized for the preparation of Fe-based WGS catalysts was citric acid. For the preparation of catalyst sample through the citric acid

20 route, metal salts to yield the desired atomic ratios of the metals and citric acid were dissolved in a 1:1 (wt/wt) solution of ethanol and water at ambient temperature while stirring continuously. The ratio of moles of citric acid to the moles of total nitrate (NO3) contained in the metal salts was kept constant at 1. The solution was kept at 80°C for 3 hours to slowly drive off ethanol and then dried in a convection oven overnight at 110°C.

The sample was then calcined in air at 450°C (heating rate: 10°C/min) for 4 hours. The sample was ground to a fine powder following calcination.

4.2.2 Catalytic activity testing

All the steady-state experiments were conducted in a fixed-bed flow reactor system using 0.25” OD stainless steel reactor. Catalyst samples were held in place by the stainless steel frit and a quartz wool plug and the temperature of the reactor bed was measured by an Omega K-type thermocouple. The reactor was placed inside a home- made, resistively heated furnace, the temperature of which was controlled by an Omega

(model CS232) PID temperature controller. Reagent gases were supplied from Praxair and were used without any further treatment. Independent gas mass flow controllers

(Brooks, 5850E) connected to an electronics control box (Brooks, 0154) were used to control gas flows to the system. Water vapor was supplied to the system by bubbling nitrogen through a heated bubbler containing de-ionized water. The temperature of the bubbler was monitored and controlled precisely by a temperature controller. Gas lines in contact with water vapor containing stream were heated using heating tapes in order to avoid water condensation in the lines.

21

Unless otherwise stated, 100 mg of catalyst sample was used for each run and the experiments were carried out at a WHSV (weight hourly space velocity) of 0.06 m3 (g cat)-1 h-1. In addition, steady-state experiments were conducted on equal surface area to ensure that the observed differences were not simply due to differences in surface areas.

In order to assure that the activity testing experiments were run in the kinetically controlled regime, reaction conditions were maintained away from equilibrium in all of the activity testing experiments.

Prior to catalytic activity testing, catalyst samples were pretreated in-situ in N2

(67.5 sccm) for 30 minutes at 350oC and then reduced at 350o C for 2 hours in syngas mixture (10% CO, 10% (or 20%) H2O, 7.5% H2, 5% CO2 balanced with 67.5% N2).

Finally, reaction studies were performed over the wide temperature range (250oC-400oC) in the above mentioned syngas mixture.

The feed as well as the reactor effluents were analyzed by an on-line gas chromatograph (Shimadzu, GC-14A series) equipped with a thermal conductivity detector (TCD). Argon was used as the carrier gas and the gas separation was performed using a molecular sieve column 13X (5 ft x 1/8 in. SS, 60/80 mesh) and porapak Q column (12 ft x 1/8 in. SS, 80/100 mesh).

4.2.3 Catalyst characterization

The specific surface areas of the samples were measured on Micromeritics ASAP

2010 surface area analyzer using N2 physisorption technique. Prior to analysis, samples were degassed overnight at 130oC under a vacuum of 3µmHg. The specific surface areas

22 were determined by the Brunauer-Emmett-Teller (BET) method using the adsorption branch of the isotherm. Pore volumes of the samples were determined by the Barret-

Joiner-Halenda (BJH) method using the adsorption branches of the N2 physisorption isotherms.

X-ray diffraction (XRD) patterns of the pristine samples were collected on a

Rigaku X-ray diffractometer (X-ray source: Cu Kα radiation, λ=1.5418 Å) operated at 40 kV and 25 mA. The diffraction patterns were collected in the range of 20o to 60o Bragg angle values. The X-ray diffraction patterns of the reduced catalyst samples were collected on a Bruker D8 Advance X-ray diffractometer equipped with a capillary sample holder. Following reduction treatment, the samples were transferred to an Ar-purged glove box that contains minimal residual air (< 5ppm). The samples were transferred to capillary sample tubes (0.2 mm OD) and the ends of the capillary tube were sealed inside the glove box with glass sealant. The samples were then transferred air-free to capillary sample holder accessory on the diffractometer. The diffraction patterns were collected in

the 20°-70° Bragg angle range using a Cu Kα X-ray source (λ=1.5418 Å) operated at 40 kV and 50 mA.

X-ray photoelectron spectroscopy was used for the investigation of the chemical states of copper and iron in the calcined (pristine) and reduced samples. The samples referred to as ‘pristine’ were ground into double sided carbon tape directly following calcination and transferred to the vacuum chamber of the XPS instrument. The reaction system described in Section 4.2.2 was used for preparing the pretreated samples. The

‘reduced’ samples were prepared through treatment in syngas mixture containing 10%

CO, 10% H2O, 7.5% H2 and 5% CO2 in N2 (100 sccm) at 350°C for 2 hours. Following

23 pretreatment, the reactor was flushed with N2 for 30 minutes at 350°C and the reactors were allowed to cool down to room temperature under N2 flow before sealing under N2.

The sealed reactors were transferred to a glove box kept under Ar atmosphere with moisture content lower than 5 ppm. The samples were ground into a carbon tape inside the glove box and then transferred air-free to the vacuum chamber of the XPS instrument.

The X-ray photoelectron spectra were collected on a Kratos Axis Ultra spectrometer

using Kα radiation through Al anode operated at 13kV and 10mA. When necessary, the charge neutralizer was operated at a current of 2.1A and a filament bias of 1.3V.

Concurrent sweeps of Cu 2p, Fe 2p, O 1s and C 1s envelopes were collected. The collected spectra were corrected using C 1s envelope located at 284.5 eV. For comparability of the signal intensities of the envelopes, counts per second, were divided by the number of sweeps and the transmission value at the specific binding energy.

Background correction and peak fitting were done using the XPSPeak 4.1 software.

The adsorption and thermal transformations of CO on Fe-Al-Cu based catalysts was investigated using an FT-IR spectrometer (Thermo Nicolet 6700) equipped with an

MCT detector and a DRIFTS chamber with ZnSe windows. The DRIFT spectra were collected in the mid-IR range at a nominal resolution of 4 cm-1 and the final spectra were averaged over 500 scans. For each experiment, catalyst sample was reduced in syngas mixture in a reactor system at 350o C for 2 hours as mentioned in Section 4.2.1. The reduced catalysts were recovered under N2 atmosphere and transferred to DRIFTS cell to acquire spectra. The sample was pretreated in He (30 sccm) at 400o C for 30 minutes, followed by background spectra collection in He (30 sccm) from 400oC to room temperature while cooling down the sample. Helium was switched off after background

24 collection and system was flushed with CO (30 sccm) for 30 minutes at room temperature. Catalyst temperature was raised under the flow of CO (30 sccm) from room temperature to 400o C stepwise and spectra were collected at regular temperature intervals.

Transmission electron microscopy (TEM) characterization was performed using a

Phillips Tecnai F20 instrument with FEG operated at a voltage of 200 kV. The catalyst images were collected in brightfield mode. Catalyst sample was dispersed in ethanol and the mixture was sonicated for 20 minutes. The sample was then deposited onto 200 mesh copper grid coated with lacey carbon.

Temperature-programmed reduction (TPR) and temperature-programmed re- oxidation experiments were conducted in order to provide insights to the oxidation/reduction characteristics of the samples. For the temperature-programmed reduction experiments, 50 mg of calcined catalyst sample was packed in a 0.25” O.D. U-

o tube quartz reactor. Catalyst was pretreated in He (35 sccm) at 350 C for 30 min and then cooled down to room temperature under the same atmosphere. Finally, temperature of the catalyst bed was raised to 880oC at 10oC/min under the flow of the reducing agent,

5%H2/N2 (20 sccm). Hydrogen consumption during the temperature ramp was monitored by a thermal conductivity detector (TCD).

For the temperature-programmed re-oxidation experiments, 100 mg of sample was packed in a fixed bed quartz reactor with a quartz frit, which is placed inside a fast response furnace (Carbolite, MTF 10/15/130). The dead volume of the reactor was filled with quartz wool. The reactor effluents during pretreatment and temperature-programmed re-oxidation stages were monitored via a residual gas analyzer (MKS – Cirrus II)

25 operated in scanning ion mode. An electron multiplier detector was used to trace mass signals (m/z) 2, 16, 18, 32 and 44. The gas lines from the reactor to the mass spectrometer capillary inlet were heated via a heating cord to prevent condensation during pretreatment and the experiment. Each experiment consisted of a reduction stage where the sample was heated at 10°C/min to 300°C under He atmosphere and then, reduced with 5%H2/He (40 sccm) at the same temperature for an hour. Reduction was followed by He flush (40 sccm) at 300°C prior to cooling to room temperature under helium. Once the sample was at room temperature, 5% O2/He (40 sccm) introduced to the reactor. The mass traces on the mass spectrometer were allowed to stabilize at room temperature and then, a linear temperature program from 30°C to 450°C was enacted with isothermal stages lasting 20 minutes at 30°C and 450°C. The same temperature- programmed re-oxidation experiment was repeated with different heating rates for the determination of activation energies of re-oxidation of Fe-Al-Cu samples. The samples were treated following the same pretreatment procedure and heated at 10, 14 or 18°C/min during the temperature-programmed re-oxidation stage. Kissinger’s method [63] was used for estimating the activation energies of re-oxidation. Kissinger’s method correlates the temperature at which the maximum rate of oxidant consumption occurs with the activation energy for re-oxidation through a series of temperature-programmed experiments ran at different heating rates. The temperature at which the maximum rate of oxidant consumption occurs corresponds to the sample temperature at which the deflection of the oxidant consumption peak takes place (Tm). The activation energy can

2 be estimated through a plot of 1/Tm versus ln (Φ/Tm ) which yields a straight line with the

26 slope being equal to –Ea/R where, Φ is the heating rate in K/min; -Ea is the activation energy in J/mol and; R is the universal gas constant in J/K.mol.

4.3 Results and Discussion

4.3.1 Effect of gelation agent on WGS activity

In our previous publications, we have reported the development of Cr-free Fe-Al-

Cu catalyst formulations which exhibited superior WGS performance over the commercial Fe-Cr catalyst in a wide temperature range [19, 20]. We also showed that the preparation method played a key role on the WGS performance of these catalysts. Our initial work involved catalyst preparation through co-precipitation of Fe and Al from nitrate precursors followed by impregnation of Cu onto the calcined precipitate (2-step) and co-precipitation of Fe-Al-Cu using nitrate precursors and sodium carbonate as the precipitating agent (1-step) [18]. In a later publication, we reported catalyst development through a ‘one-pot’ sol-gel-precipitation route using organic precursors [19]. The Fe-Al-

Cu samples prepared through this sol-gel/precipitation route showed significant enhancement in the WGS activity compared to the 2-step and 1-step routes. However, this sol-gel/precipitation method required critical pH adjustment, which was achieved through addition of NaOH and involved extensive catalyst washing to remove the sodium ions afterwards.

To address the issues associated with catalyst preparation, we have investigated the applicability of a gelation agent-assisted sol-gel ‘one-pot’ route using propylene oxide

27 as the gelation agent. Fe-only, Fe-Cu and Fe-Al-Cu samples were prepared by the propylene oxide assisted sol-gel route. The monometallic Fe catalyst prepared by propylene oxide-assisted sol-gel route exhibited a reasonable CO conversion of 30% at

400oC in a syngas mixture. The textural promoter (Al) and structural promoter (Cu) were incorporated in the catalysts in order to improve their WGS activity. Al and Cu have been reported to be effective promoters [5, 10, 11, 13, 17, 19, 20]. Among the promotional effects that are reported are structural and textural promotion, which include improved surface area, resistance to sintering and change in the electronic properties of the magnetite phase. As expected, further improvement was achieved in the WGS activity by incorporation of Cu and Al in the catalysts prepared by propylene oxides-assisted sol gel technique, achieving CO conversions of 40% and 57% over Fe-Cu (Fe/Cu=5) and Fe-Al-

Cu (Fe/Al=10 and Fe/Cu=5), respectively. The equilibrium CO conversion under these conditions is 64%.

Similar Fe-based water gas shift catalysts were also prepared by using another commonly used gelation agent, citric acid, and tested with regards to their activity in the syngas mixture. The Fe-only catalyst prepared through the citric acid route showed slightly higher WGS activity than the Fe-only catalyst prepared through the propylene oxide route however, the activity was not stable with time-on-stream. Further investigation of the WGS activity of Fe-Al-Cu (Fe/Al=10 and Fe/Cu=5) catalyst prepared through the citric acid route showed that, the activity of this catalyst was not stable either and the initial CO conversion of 40% over this catalyst decreased to less than 20% with time on stream over a period of 10 hours. Although the citric acid-assisted route was shown to be unsuitable for preparation of active and stable Fe-Al-Cu catalysts, the initial

28 catalyst testing results showed that it was possible to prepare active and stable Fe-Al-Cu catalysts through the propylene oxide-assisted sol-gel route. Therefore, the structural properties of Fe-based catalysts prepared through the propylene oxide route and the effect of copper loading on their performance was further investigated

4.3.2 Effect of Cu loading on the WGS activity

Sol-gel Fe-based catalysts with different Cu loadings (Fe/Cu = 10, 5 and 2) and constant aluminum loadings (Fe/Al = 10) were prepared using the propylene oxide- assisted route in order to investigate the effect of copper loading on the WGS activity of these catalysts and to elucidate the structural changes induced by copper incorporation.

Table 4.1 presents the specific surface areas and pore volumes of these Fe-Al-Cu catalysts. The surface area was found to be strongly dependent on copper loading and it decreased with increasing Cu content of the catalysts. A similar trend was observed in pore volume. As it will be discussed in the following sections, incorporation of copper in

Fe-based catalysts resulted in significant changes in the catalyst structure, especially in the major crystalline phases and particle size, as well as catalytic performance.

29

BJH adsorption BET surface area Catalyst pore volume (m2/g) (cm3/g)

Fe/Cu = 10 99 0.1830

Fe/Cu = 5 46 0.1669

Fe/Cu = 2 25 0.0421

Table 4.1: BET surface area, and pore volume of Fe-Al-Cu catalysts at various Cu- loadings

30

4.3.2.1 Catalyst Activity

Figure 4.1a shows the WGS activity of the three Fe-Al-Cu samples a wide temperature range (250-400°C) in syngas mixture at WHSV of 0.06 m3 (g cat)-1 h-1 at a steam-to-carbon (S/C) feed ratio of 1. WGS activity was significantly improved when Cu content was increased from Fe/Cu=10 to 5. The Fe-Al-Cu (Fe/Cu=5) sample was able to achieve CO conversion in excess of 55% in a broad temperature window of 300-400oC.

The equilibrium CO conversion at 400oC under these reaction conditions was 64%.

However, Fe-Al-Cu (Fe/Cu=10) sample was able to achieve a maximum CO conversion of slightly lower than 40%. Further increase in Cu content affected WGS activity adversely, resulting in considerable drop in CO conversion. This catalyst was able to achieve only 27% CO conversion in 300-400°C. When a commercial Fe-Cr catalyst was tested under identical conditions, the CO conversions achieved were much lower, even when compared to the least active of the formulations tested, i.e., Fe-Al-Cu (Fe/Cu=2).

Since the catalysts with different Cu loadings were compared at equal WHSV, a valid question would be if the differences in activities could be due to the differences in surface areas. To address this question, WGS activity of Fe-Al-Cu (Fe/Cu=2) and Fe-Al-

Cu (Fe/Cu=5) samples were also tested at 400°C on an equal surface area basis and the

Fe-Al-Cu (Fe/Cu=2) sample was found to exhibit significantly lower CO conversion than

Fe-Al-Cu (Fe/Cu=5). The CO conversion over Fe-Al-Cu (Fe/Cu=2) sample was 30%, whereas it was 41% over Fe-Al-Cu (Fe/Cu=5) at 400oC on equal surface area basis.

Therefore, decrease in CO conversion over Fe-Al-Cu (Fe/Cu=2) cannot be explained with the decrease in surface area.

31

The Fe-Al-Cu (Fe/Cu=5) catalyst was further tested for stability. The inset to

Figure 4.1a shows CO conversion over this catalyst during WGS in syngas mixture as a function of time-on-stream at 400oC. The catalyst was observed to exhibit stable WGS activity with time-on-stream over a period of 40 hours.

The WGS activity of the Fe-Al-Cu catalysts was further investigated at a higher steam-to-carbon ratio of the reaction feed. Figure 4.1b shows the CO conversions achieved as a function of temperature over Fe-Al-Cu catalysts during WGS reaction at a steam-to-carbon ratio of 2. At low temperatures no significant effect of increasing steam- to-carbon ratio on the WGS activity was observed; however, at higher temperatures, increased steam content of the feed resulted in a significant increase in CO conversion.

Although an increase in the stream content resulted in a shift in CO conversion curves to higher values over all of the samples by accelerating the forward reaction, it had a more pronounced effect over Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/Cu=2) than Fe-Al-Cu

(Fe/Cu=5). At 400°C, the CO conversion over Fe-Al-Cu (Fe/Cu=5) increased from 57% to 72% with increasing steam-to-carbon ratio, whereas the CO conversions over Fe-Al-

Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/Cu=2) almost doubled. As it will be discussed in

Section 4.3.2.6, the re-oxidation of Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/Cu=2) is expected to proceed at a much slower rate than the re-oxidation of Fe-Al-Cu (Fe/Cu=5) during reaction. The fact that Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/Cu=2) catalysts benefit more from the increased steam content of the feed as compared to Fe-Al-Cu

(Fe/Cu=5) implies that the re-oxidation of surface through water splitting during WGS reaction is the rate limiting step over these catalysts under these conditions.

32

Figure 4.1: WGS activity (CO% conversion) of Fe-Al-Cu catalysts prepared by propylene oxide-assisted sol-gel method with different Cu loadings at (a) S/C = 1 and (b) S/C = 2.

Inset: time-on-stream performance of Fe-Al-Cu (Fe/Cu=5) at 400o C at S/C = 1. Reaction

Conditions: [CO] = 10%, [H2O] = 10% or 20%, [CO2] = 5%, [H2] = 7.5% and balance

3 -1 -1 N2, WHSV = 0.06 m (g cat) h , P = 1 atm.

33

4.3.2.2 X-ray diffraction

X-ray diffraction was used to examine the effect of copper incorporation and copper loading on the crystal phases formed during catalyst preparation. The X-ray diffraction pattern of the Fe-only sample showed the presence of α-Fe2O3 (hematite) phase, as evidenced by diffraction features located at Bragg angle (2θ) values of at 24.2o,

33.2o, 35.7o, 40.9o and 49.5o (Figure 4.2a) which can be associated with (012), (104),

(110), (113) and (024) crystal planes, respectively. Hematite phase disappeared with the incorporation of Cu into the iron oxide matrix. The X-ray diffraction pattern of Fe-Cu

(Fe/Cu=5) reflected the formation of γ-Fe2O3 (maghemite) phase instead of hematite phase. The diffraction lines at Bragg angle (2θ) values of 30.1o, 35.6o and 43.2o (Figure

4.2b) were associated with (220), (311) and (400) planes of maghemite, respectively.

This maghemite phase was retained even after the addition of Al (Figure 4.2c-e).

Maghemite is considered to be more active than hematite phase for WGS [11]. In addition, maghemite phase is in an imperfect cubic spinel structure, which is similar to the WGS active phase, Fe3O4 (magnetite) [6, 11, 73, 74]. This imperfect structure of maghemite allows better incorporation of active promoters than hexagonal hematite phase. Consequently, active promoter incorporated in maghemite is less prone to sintering effects [11, 75]. Also Figure 4.2a-c showed that the addition of promoters lead to broadening of diffraction lines which could be correlated to a decrease in the crystallite size.

The X-ray diffraction patterns of Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/Cu=5) samples (Figure 4.2c and 4.2d, respectively) showed diffraction lines located at 2θ values

34 of 30.1o, 35.6o and 43.2o which can be associated with (220), (311) and (400) planes of maghemite. No other phases that could be related to the presence of copper in the structure were detected in either of the samples. Unlike these two catalysts, the sample with the highest amount of copper loading (Fe/Cu=2) consisted of peaks that could be associated with maghemite as well as a separate CuO phase, which was identified through peaks located at 38.6o and 48.6o. The development of a separate CuO phase in the Fe-Al-Cu (Fe/Cu=2) sample but not in Fe-Al-Cu (Fe/Cu=5) suggests that there is a certain amount of copper atoms that could be incorporated into the iron oxide matrix. We have previously observed the separation of copper in the form of CuO domains on Fe-Al-

Cu samples prepared through the sol-gel route and suggested that there was a limit on the amount of copper atoms that could be incorporated into the iron oxide matrix [20]. We have further suggested that the amount of copper that could be incorporated into the iron oxide matrix lied in between Fe/Cu=5 and Fe/Cu=2, with the latter being an upper limit for Cu incorporation. Fe-Al-Cu samples prepared through the propylene oxide-assisted sol-gel route described in this study showed similar behavior, also suggesting that the maximum amount of copper that could be incorporated in the iron oxide matrix is a function of the charge structure of the system, rather than being dependent on the catalyst preparation route. Cu is an active promoter and modifies electronic properties of iron catalyst [8, 19, 20, 76] as long as it is incorporated in the iron oxide matrix. The optimum amount of Cu in iron oxide matrix facilitates the formation of WGS active maghemite phase and improves surface area of Fe-based catalysts, as shown earlier. However, metallic Cu on the surface is more prone to sintering at high operating temperatures [19,

35

20] which can result in poor WGS activity. The catalytic activity results presented in

Section 4.3.2.1 are in agreement with these findings.

The crystallite size in Fe-Al-Cu samples was calculated through the Scherrer equation. The crystallite size of Fe-Al-Cu samples increased with increasing copper loading (Table 4.2) indicating the formation of larger, ordered crystals with the incorporation of increasing amounts of copper into the structure. The formation of crystals with longer range order was also evidenced with the decrease in the specific surface area of the samples with increasing copper loading. The crystallite size calculations were in line with transmission electron microscopy (TEM) images, which show the presence of uniform, nano-sized (~12 nm) particles of Fe-Al-Cu sample, as shown in Figure 4.3. The lattice fringes for (400) plane of maghemite with a d-spacing of

1.70Å is visible in the TEM image of Fe-Al-Cu (Fe/Al=10, Fe/Cu=5) catalyst.

The capillary XRD technique was used to study the reduced Fe-Al-Cu samples in order to avoid oxidation of the catalyst surface during the sample transfer. The reduced samples were pretreated in nitrogen and then reduced in syngas mixture (refer to Section

4.2.2). Figure 4.4 shows XRD patterns of the reduced samples of Fe-Al-Cu at different

Cu-loadings. Magnetite phase (resulting from reduction of maghemite) was observed through peaks associated with (220), (311), (400), (511) and (440) planes at Bragg angle

(2θ) values of 30.1o, 35.4o, 43.3o, 57.4o, and 62.7o, respectively, in Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/Cu=5) samples. The XRD pattern of Fe-Al-Cu (Fe/Cu=2) sample, along with maghemite phase, also contained intense peaks of metallic copper (Cuo) at

43.2o and 50.3° which correspond to diffraction from (111) and (200) planes.

36

Crystallite Size Catalyst (nm)*

Fe/Cu = 10 11

Fe/Cu = 5 12

Fe/Cu = 2 15

Table 4.2: Crystallite size of Fe-Al-Cu catalysts at various Cu-loadings calculated from

the XRD patterns through Scherrer equation

37

Figure 4.2: XRD patterns of Fe-based samples. (a) Fe-only (b) Fe-Cu (Fe/Cu=10) (c) Fe-

Al-Cu (Fe/Cu = 10), (d) Fe-Al-Cu (Fe/Cu = 5) and (e) Fe-Al-Cu (Fe/Cu = 2).

38

Figure 4.3: Transmission electron micrograph (TEM) of Fe-Al-Cu (Fe/Cu = 5)

39

Figure 4.4: XRD patterns of Fe-Al-Cu samples reduced in syngas mixture (a) Fe/Cu=10,

(b)Fe/Cu=5 and (c)Fe/Cu=2

40

4.3.2.3 X-ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy was used to investigate the chemical states of iron and copper over the pristine samples and the changes that take place during catalyst activation via reduction in syngas. Figure 4.5 shows the Cu 2p envelopes of pristine and reduced samples of Fe-Al-Cu catalysts with Fe/Cu ratios of 10, 5 and 2. In order to prevent interaction of the reduced samples with ambient air, the reduced samples were transferred air-free from the reactor to the vacuum chamber of the XPS instrument under an inert atmosphere.

The Cu 2p envelopes of the pristine samples resembled each other regardless of the copper content of the catalyst. Figure 4.5 also shows the curve fitting for identification of photopeaks. In all of the samples, the Cu 2p3/2 component of Cu 2p envelope was located at 933.3eV with shake-up satellite lines at 941.0eV (ΔEB = 7.7 eV) and 943.1eV (ΔEB = 9.8eV). The spin orbit splitting (SOS) of Cu 2p3/2 and Cu 2p1/2 was

19.8eV, and the satellite peak for 2p1/2 was located at 961.7eV (ΔEB = 8.6 eV). The appearance of shake-up satellite peaks is common for paramagnetic species and the observed displacement in these together with the position of the photoelectron lines in the

Cu 2p region suggests the presence of Cu2+ ions in CuO environment as the major component [77, 78].

Following reduction treatment, lower oxidation state copper species were observed together with cupric oxide over all of the samples. It is well known that reduction of copper oxide occurs during irradiation with the X-ray beam in the XPS

41 instrument [79, 80] and it is worthwhile to note that, we have observed the presence of

Cu2+ over pristine samples under similar experimental conditions and irradiation periods in the XPS instrument as the sole copper species. It is plausible that interaction of copper species with iron oxide network leads to increased stability of the CuO species over Fe-

Al-Cu samples compared to pure cupric oxide. In line with the above discussion, the presence of copper species with lower oxidation states over the reduced samples is likely to be associated with the preceding treatment steps rather than reduction during X-ray irradiation.

Curve fitting of Cu 2p3/2 region of the reduced samples suggested the presence of two components –located at 933.3 eV and 932 eV- contributing to this photopeak in all of the reduced samples. Over Fe-Al-Cu (Fe/Cu = 5) and Fe-Al-Cu (Fe/Cu = 10) catalysts, the shake-up lines that were displaced by 7.8eV and 9.8eV with respect to 2p3/2 were also observed. The higher binding energy component together with the shake-up lines was associated with Cu2+ ions whose appearance have been attributed to stabilization through interaction with iron oxide matrix. The location of the lower binding energy component is characteristic of copper species at lower oxidation state, however differentiation of Cu1+ and Cu0 by the photoelectron lines is not possible as the difference in the core level binding energies of copper atoms in these environments is well within experimental uncertainty. Over Fe-Al-Cu (Fe/Cu = 2) the Cu 2p envelope was dominated by lower oxidation state copper species (i.e. Cu1+ or Cu0) with little contribution from a higher binding energy component. Although the binding energy difference between

+1 metallic copper and Cu ions in Cu2O environment is too low to allow spectral differentiation of these species, complementary evidence from the X-ray diffraction

42 patterns presented in Section 4.3.2.2 allows to associate this photopeak with the presence of metallic copper on reduced Fe-Al-Cu (Fe/Cu = 2). The relative contributions of CuO and the lower oxidation state component of the Cu 2p envelope were calculated using the areas under the fitted components. The relative contribution from Cu2+ to the Cu 2p envelopes of reduced Fe-Al-Cu (Fe/Cu=10), Fe-Al-Cu (Fe/Cu=5) and Fe-Al-Cu

(Fe/Cu=2) was 72%, 74% and 18%, respectively. The presence of higher amount of lower oxidation state components over Fe-Al-Cu (Fe/Cu=2) in comparison to Fe-Al-Cu

(Fe/Cu=5) and Fe-Al-Cu (Fe/Cu=10), can be attributed to the copper species which have not been incorporated into the iron oxide matrix over this catalyst. Deprived of the stabilizing effect of iron oxide matrix, this free CuO phase is more prone to reduction and sintering at elevated temperatures and harsher reducing conditions.

43

Figure 4.5: X-ray photoelectron spectra of Cu 2p region of pristine and reduced Fe-Al-Cu samples: (a) Fe/Cu = 10, (b) Fe/Cu = 5 and (c) Fe/Cu = 2

44

Figure 4.6 shows the X-ray photoelectron spectra in the Fe 2p region of pristine and reduced Fe-Al-Cu samples with the three different Fe/Cu ratios of 2, 5 and 10. The

X-ray photoelectron spectra of the pristine samples were similar and contributions of four different photopeaks were resolved. The photopeak located at 711.1 eV is characteristic of core level Fe 2p3/2 photoelectrons of ferric oxide. The 2p3/2 – 2p1/2 spin orbit splitting was 13.5 eV. The high degree of multiplet splitting of the Fe 2p core line resulted in the broad photopeaks observed in this region. Also observed were the satellite lines which were displaced by 8 eV. These satellite peaks were characteristic of Fe2O3. These satellite peaks are commonly used as fingerprints for different oxides of iron as the intensity and the position of the satellite peak is dependent on the environment of the Fe3+ ions.

However, Fe3+ satellites were not the only photoelectrons that have contributions in this

region in the case of Fe-Al-Cu samples. Since monochromatic Al Kα radiation which resulted in higher spectral resolution was used for obtaining well defined peak shapes, this region had contribution from Cu LMM Auger electrons at c.a. 719 eV. The core

3+ level Fe 2p spectra of Fe species in α-Fe2O3 and γ-Fe2O3 environments are almost identical to each other. McIntyre and Zetaruk [81] used multiplet splitting patterns of Fe

2p core levels for identification of different oxides of iron. Through curve fitting of the

Fe 2p3/2 peak these authors showed that the splitting of the two most intense components of the envelope was 1eV for γ-Fe2O3 which was 0.2 eV lower than α-Fe2O3. In line with their findings, the curve fitting of the Fe 2p core level spectra of the pristine Fe-Al-Cu samples using the same parameters that these authors have used yielded a displacement of 1 eV between the two most intense components (not shown). Although more indirect than XRD, X-ray photoelectron spectroscopy provides complementary evidence for the

45 presence of iron in the form of γ-Fe2O3 in the pristine samples. Following reduction treatment, the Fe 2p envelopes of all of the samples were similar to each other; the Fe

2p3/2 photopeak was located at 710.8 eV (SOS = 13.5 eV) with a weakly resolved shoulder that was displaced by 2.5 eV towards the lower binding energy side (Figure

4.6). Another important feature of the Fe 2p envelope of the reduced samples was the disappearance of the shake-up lines, further indicated the transformation of Fe2O3 to

Fe3O4 during reduction treatment. The X-ray photoelectron spectra suggested the presence of magnetite as the major phase and that the adopted reduction procedure resulted in partial reduction of iron as the presence of metallic iron (c.a. 707 eV) was not detected in any of the samples.

46

Figure 4.6: X-ray photoelectron spectra of Fe 2p region of pristine and reduced Fe-Al-Cu samples: (a) Fe/Cu = 10, (b) Fe/Cu = 5 and (c) Fe/Cu = 2

47

4.3.2.4 Temperature-programmed reduction

The effect of copper loading on the reduction characteristics of Fe-Al-Cu samples was investigated with temperature programmed reduction (TPR) in 80°C-880°C region.

Figure 4.7 shows the H2 consumption traces from these experiments. All of the samples exhibited a well-resolved consumption feature in 170-400°C region and another broad consumption band that corresponded to higher temperatures. The low temperature H2 consumption band, which can be associated with reduction of maghemite to magnetite, corresponded to the same temperature range for Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu

(Fe/Cu=5) (Tmax =270°C and 261°C, respectively), whereas the peak shifted to higher temperatures (Tmax =296°C) over Fe-Al-Cu (Fe/Cu=2) sample. The shoulder peak at

o 230 C in H2 consumption trace of Fe-Al-Cu (Fe/Cu=2) was assigned to reduction of CuO to metallic Cu, presence of which has been confirmed through XRD pattern of the reduced sample. The reduction peaks above 600o C (in all samples) were assigned to over-reduction of active magnetite to WGS-inactive sub-oxides or to metallic Fe. The

TPR studies indicated that optimum Cu content (Fe/Cu=5) eased the reduction characteristic of iron oxide matrix (here, maghemite) to form active magnetite phase, whereas excess Cu content retarded the formation of the active phase during reduction.

The segregation of a separate CuO phase was also demonstrated in these experiments.

48

Figure 4.7: Temperature-programmed reduction profiles of Fe-Al-Cu catalysts.

49

4.3.2.5 In situ DRIFTS studies using CO

The interaction of CO with Fe-Al-Cu catalysts was investigated using in-situ

DRIFTS. Figure 4.8 shows the DRIFTS spectra of CO adsorbed on the reduced Fe-Al-Cu sample in the temperature range of 30o C- 400oC. Bands in the range of 1000-1700 cm-1 correspond to OCO symmetric as well as asymmetric stretching vibrations. However, carbonate and formate both contain OCO stretching and therefore it is difficult to distinguish between them. One possible way to is look for bands in the region 2800-3000 cm-1 for C-H stretching [82] as formate species contain both OCO and C-H stretching.

Fe-Al-Cu sample showed C-H stretching vibrations at 2845 and 2950 cm-1, both could be due to bidentate formate or bidentate and bridging type formate species [36]. These C-H stretching features increased in intensity with increasing adsorption temperature. The bands at 2110 and 2173 cm-1 assigned to weakly adsorbed CO and gas phase CO,

-1 respectively. The doublet of weakly adsorbed CO2 was observed at 2327 and 2360 cm

[18].

The bands in between 3500-3800 cm-1 were characteristic of OH groups associated with the catalyst. The bands at 3643 and 3685 cm-1 were assigned to Type II bridging OH group (geminal) [36, 83] on the “reduced” Fe-Al-Cu catalyst. Jacobs et al.

[36] reported that these Type II OH groups reacted with CO to form formate species for

Pt/CeO2 catalytic system. For Fe-Al-Cu, we observed that Type II OH groups diminished as temperature increased under the flow of CO while formate bands increased in intensity. Peaks at 3611 and 674 cm-1 were the characteristic of hydroxyl vibrations

50 associated with Cu [84]. The band at 3713 cm-1 was assigned to Type I OH group

(terminal) [36]. The OH stretching of adsorbed water was observed at 3400 cm-1 [85] at lower temperatures, which disappeared above 100oC.

While the presence of formate species may suggest an involvement as an intermediate in the reaction mechanism, it is also possible that they may be acting merely as spectators [86, 87], and that the redox pathway is more likely to be the mechanism in operation in WGS reactions over these catalysts [18, 19]. Temperature-programmed re- oxidation studies discussed in the next section also support this hypothesis. The catalyst re-oxidation studies suggest that a redox pathway may be the operative mechanism over these catalysts

51

Figure 4.8: DRIFTS study of CO interaction with Fe-Al-Cu (Fe/Cu=5)

52

4.3.2.6 Temperature-Programmed Re-oxidation

Re-oxidation of the catalyst surface constitutes an important step in water-gas shift reaction. Usually, surface oxygen vacancies are created through the oxidation of adsorbed CO and these vacancies act as sites for water adsorption to replenish surface oxygen [18, 88, 89]. Temperature programmed re-oxidation experiments were performed following reduction of Fe-Al-Cu samples to study the re-oxidation kinetics of the catalysts and to determine the apparent activation energies of re-oxidation of the Fe-Al-

Cu samples with Fe/Cu ratios of 2, 5 and 10.

The samples were reduced in 5%H2/He at 300°C and then, re-oxidized in

5%O2/He while being heated at 10, 14 or 18°C/min to 450°C. The reduction temperature for these experiments was chosen in light of the TPR results discussed in Section 4.3.2.4, to enable reduction of γ-Fe2O3 to Fe3O4, but to prevent magnetite from being over- reduced. Figure 4.9 shows oxidant consumption profiles during the temperature programmed re-oxidation of reduced Fe-Al-Cu catalysts. At a heating rate of 10°C/min, maximum oxidant consumption temperature corresponded to 201°C, 169°C and 180°C for Fe-Al-Cu (Fe/Cu = 2) Fe-Al-Cu (Fe/Cu=5) and Fe-Al-Cu (Fe/Cu=10), respectively.

For Fe-Al-Cu (Fe/Cu = 2), the maximum oxidant consumption temperature shifted from

201°C to 211°C as the heating rate was increased from 10°C/min to 18°C/min (Figure

4.9c). A shoulder in oxygen consumption profile was also resolved at 135°C. In line with the H2-TPR results, this O2 consumption feature could be associated with oxidation of

53 metallic copper species on the surface which have not been incorporated to the iron oxide matrix. Over Fe-Al-Cu (Fe/Cu = 5) and Fe-Al-Cu (Fe/Cu = 10), the temperature for maximum oxidant consumption shifted from 168°C to 200°C and from 180°C to 188°C, respectively. The oxidant consumption traces were composed of a single component, therefore no evidence for the formation of metallic copper domains through diffusion of copper species from the iron oxide matrix to the surface during reduction was present over these samples.

In a plot of oxidant consumption as a function of temperature, the deflection point gives the temperature (Tm) at which the oxidant consumption rate is a maximum. The temperature of maximum oxidant consumption rate shifted to higher temperatures as the

2 heating rate (Φ) was increased and a plot of 1/Tm versus ln(Φ/Tm ) was used to determine the activation energy for the re-oxidation process. In line with the above discussion on the selection of the reduction temperature, the apparent activation energies for re- oxidation calculated through Arrhenius plots were linked to re-oxidation of Fe3O4. The

2 inset to Figure 4.9c shows the plot of 1/Tm versus ln (Φ/Tm ) which gives a straight line with a slope of -12.48 that corresponded to activation energy of 103.76kJ/mol for re- oxidation of Fe-Al-Cu (Fe/Cu=2). Similarly, the activation energies for re-oxidation were calculated as 25kJ/mol and 115kJ/mol for Fe-Al-Cu (Fe/Cu=5) and Fe-Al-Cu

(Fe/Cu=10), respectively. Based on the activation energies, the re-oxidation of Fe-Al-Cu

(Fe/Cu = 2) and Fe-Al-Cu (Fe/Cu = 10) is expected to take place at a much slower rate than Fe-Al-Cu (Fe/Cu = 5).

Iron oxide-chromium catalysts have been shown to have apparent WGS activation energies in the order of 110-130 kJ/mol [8, 90, 91]. Hutchings et al. [8, 13] further

54 showed that incorporation of copper into the Fe-Cr system reduced the apparent activation energy for WGS reaction and reported activation energies in the order of 80 kJ/mol for this system. The apparent activation energies of re-oxidation of Fe-Al-Cu samples showed that while the re-oxidation of Fe-Al-Cu (Fe/Cu=2) and Fe-Al-Cu

(Fe/Cu=10) samples proceed at a rate comparable to (if not slower than) the reduction of the surface during oxidation of adsorbed CO with surface oxygen, the re-oxidation of Fe-

Al-Cu (Fe/Cu=5) proceeded at a much faster rate than the reduction. Therefore, WGS reaction over the Fe-Al-Cu (Fe/Cu=5) sample would solely be controlled by reduction kinetics of the surface whereas for the Fe-Al-Cu (Fe/Cu=2) and Fe-Al-Cu (Fe/Cu=10) availability of surface oxygen for CO oxidation would be a rate determining factor.

These results are consistent with the reaction experiments which showed a much more pronounced improvement over the Fe-Al-Cu (Fe/Cu=2) and Fe-Al-Cu (Fe/Cu=10) when steam-to-CO ratio was doubled. The catalysts which had the higher activation energy barriers for re-oxidation benefited more from the increased water concentration than the catalyst which had a lower activation energy for re-oxidation with water. Although the presence of formate species have been observed in the DRIFT spectra which does not allow us to rule out the formate mechanism during WGS reaction, the catalyst re- oxidation studies suggest that a red-ox pathway may be the operative mechanism over these catalysts.

55

Figure 4.9: Temperature-programmed re-oxidation of reduced Fe-Al-Cu samples at different ramp rates: (a) Fe/Cu = 10, (b) Fe/Cu = 5 and (c) Fe/Cu = 2. Insets show determination of activation energies of re-oxidation following Kissinger method

56

4.4 Conclusions

Gelation agent-assisted one-pot sol-gel technique was used for preparation of Fe- based WGS catalysts. Propylene oxide and citric acid were investigated as the gelation agents. Propylene oxide emerged as a more suitable gelation agent than citric acid. The effect of Cu-loading study showed that Fe-Al-Cu (Fe/Cu=5) was the optimum Cu-loading and the excess amount of copper in Fe-Al-Cu (Fe/Cu=2) caused significant decrease in the WGS activity. With increasing copper content, segregation of a copper-containing phase, which is more prone to reduction under reaction conditions, was observed through

XRD and TPR.

The apparent activation energies of re-oxidation of Fe-Al-Cu (Fe/Cu=10) and Fe-

Al-Cu (Fe/Cu=2) were found to be much higher than Fe-Al-Cu (Fe/Cu=5). The apparent activation energies of re-oxidation of the Fe-Al-Cu catalysts suggested that WGS reaction over Fe-Al-Cu (Fe/Cu=5) was controlled by the rate of reduction of the surface through oxygen removal by CO, whereas for Fe-Al-Cu (Fe/Cu=10) and Fe-Al-Cu (Fe/2) reduction and re-oxidation of the surface would proceed at comparable rates. Increasing water concentration in the feed showed a more pronounced improvement in CO conversion over the catalysts with high activation energy barriers for re-oxidation.

* Information in this chapter is adapted/taken from P. Gawade, B. Mirkelamoglu, B. Tan and U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 321, (2010), 61-70

57

CHAPTER 5

THE ROLE of SUPPORT MORPHOLOGY and IMPREGNATION MEDIUM on

WATER-GAS SHIFT ACTIVITY of CERIA-SUPPORTED COPPER CATALYSTS

5.1 Overview of Cu/CeO2 Catalyst: Effect of Ceria Morphology and Impregnation

Medium

In the present chapter, we report the effect of ceria nanostructure on the WGS activity of Cu catalysts. Two different nanostructures of ceria, nano-particles (NP) and nanorods (NR) were prepared using precipitation-hydrothermal method. Wet impregnation method was used to incorporate copper over ceria using, ethanol (E) or tetrahydrofuran (T) as the impregnation medium. To the best of our knowledge, the effect of CeO2 morphology on WGS reaction over Cu/CeO2 catalytic system has not been previously reported.

Steady-state reaction experiments showed that the catalysts supported over CeO2 nano-particles achieved significantly higher CO conversions than catalysts supported over nano-rods, which achieved only marginally higher conversions than the bare support. The catalysts were characterized by transmission electron microscopy, X-ray diffraction, X-ray photoelectron spectroscopy, N2O chemisorption and H2-temperature-

58 programmed reduction. Characterization of the catalysts showed that copper dispersion and catalytic activity were strongly correlated. Copper was observed to be better dispersed over CeO2 nano-particles, while larger, crystalline CuO domains with properties closer to that of bulk CuO were identified over CeO2 nano-rods. The finely dispersed copper species constitute the active sites for WGS reaction over Cu/CeO2 catalysts.

5.2 Experimental Procedure

5.2.1 Catalyst Preparation

Both of the CeO2 supports (nano-particles and nano-rods) were prepared via precipitation from an aqueous solution of 0.4 M cerium (III) nitrate hexahydrate

(99.999% trace metals basis, Sigma-Aldrich), by an aqueous solution of sodium hydroxide (2.7M) at room temperature under constant stirring. For the preparation of ceria nano-particles (NP), the precipitates were collected by filtration and washed with deionized water until the pH of the filtrate was 7. The particles were, then, dried at 110°C overnight in a convection oven followed by calcination at 400°C for 3 hours in air at

10oC/min. For the preparation of nano-rods (NR), the precursor solution and precipitates were transferred to a Pyrex© bottle, sealed and kept at 110°C for 24 hours. The precipitates were then filtered, washed, dried and calcined following the above- mentioned procedure for the preparation of the nano-particles.

59

The Cu/CeO2 catalysts were prepared through impregnation of the nano-structured

CeO2 supports with copper via a conventional wet-impregnation method. Two different organic solvents, namely ethanol (E) and tetrahydrofuran (T) were used as the impregnation media. For impregnation, 1.9 g of ceria support was added to a 40 ml solution of 0.04 M copper (II) nitrate tri-hydrate (98%, Fluka) in ethanol or tetrahydrofuran and kept overnight at room temperature while stirring continuously. The catalysts were dried at 100°C overnight in a convection oven and calcined at 400°C for 3 hours at 10oC/min under a flow of air.

The specific surface areas of the catalysts were measured using N2 physisorption technique on a Micromeritics ASAP 2020 accelerated surface area and porosimetry instrument. Before measurement, samples were degassed for 12 hours at 130°C under a vacuum better than 3 µmHg. The surface areas of the samples were determined using the

Brunauer-Emmett-Teller (BET) method and presented in Table 5.1.

60

Sample Surface Area (m2/g)

CeO2-NP 123

CeO2-NR 115

Cu/CeO2-NP(E) 112

Cu/CeO2-NP(T) 94

Cu/CeO2-NR(E) 106

Cu/CeO2-NR(T) 103

Table 5.1: Specific surface areas of CeO2-based samples calculated using BET method

61

5.2.2 Catalyst Activity Testing

The steady-state reaction experiments were conducted in a fixed-bed flow reactor system using ¼” OD stainless steel reactor with a stainless steel frit. The reactor was placed inside a resistively heated furnace and the reactor temperature was controlled by an Omega CSC232 PID temperature controller. Water vapor was supplied to the reactor system by bubbling nitrogen through a heated bubbler containing de-ionized water. Gas lines in contact with water-vapor-containing stream were heated in order to avoid water condensation in the lines. 100 mg of catalyst sample was packed inside the reactor for each run and the samples were pretreated in-situ in N2 (67.5 sccm) for 30 minutes at

o 350 C and then reduced in a syn-gas mixture with a composition of 10% CO, 10% H2O,

7.5% H2, 5% CO2 in N2 for 2 hours at 350°C.

Reaction studies were performed in 250-400°C range at a WHSV (weight hourly space velocity) of 0.06 m3 (g cat)-1 h-1. The feed syn-gas mixture constituted of 10% CO,

10% H2O, 7.5% H2, 5% CO2 and balance N2. The experiments were run in the kinetically controlled regime by maintaining the reaction conditions away from equilibrium during all of the activity testing experiments. The effluents from the reactor were monitored by an on-line gas chromatograph (Shimadzu, GC-14A series) equipped with molecular sieve

13X (5 ft x 1/8 in. SS, 60/80 mesh) and porapak Q (12 ft x 1/8 in. SS, 80/100 mesh) columns, and a thermal conductivity detector (TCD).

62

5.2.3 Catalyst Characterization

The X-ray diffraction patterns of the Cu/CeO2 catalysts as well as the CeO2 supports were collected in the 20-60° diffraction angle range on a Rigaku X-ray diffractometer using Cu Kα radiation (λ=1.5418 Å) through a tube operated at 40 kV and 25 mA. The transformation of crystal phases during reduction of the catalysts was monitored by in- situ XRD. A Bruker D8 Advance X-ray diffractometer using monochromatic Cu Kα radiation (λ=1.5418 Å) through a tube operated at 40 kV and 50mA and equipped with an

Anton Paar HTK1200 oven was used for collection of the X-ray diffraction patterns during in-situ reduction of Cu/CeO2 samples. The samples were heated from room temperature to 450°C under a flow of 5% H2/N2 (15 ccm) and diffraction patterns were collected at 30°C, 100°C, 200°C, 300°C, 400°C and 450°C. Following the collection of the XRD pattern at one temperature, the samples were heated under 5%H2/N2 flow at a rate of 12°C/min to the next set point where they were kept isothermally under the same flow for 30 minutes prior to collection of the XRD pattern. The identification of the crystalline phases through the collected diffraction patterns was done by using

International Center for Diffraction Data (ICDD) database.

Transmission electron microscopy (TEM) images were collected in brightfield mode on a Phillips Tecnai F20 instrument with FEG operated at a voltage of 200 kV. The samples were dispersed in ethanol and sonicated for 20 minutes before being deposited onto 200 mesh copper grid coated with lacey carbon/formvar.

63

The chemical states of cerium and copper in the Cu/CeO2 catalysts were investigated on a Kratos AXIS Ultra X-ray photoelectron spectrometer, using monochromatized Al Kα (1486.7eV) X-ray source operated at 13 kV and 10 mA. For each sample a survey scan was collected from 1400 to 0 eV. Next, concurrent sweeps of

Cu 2p, Ce 3d and O1s regions were collected at an electron pass energy of 20 eV.

Analysis of the collected data was performed on XPS Peak 4.1.

The temperature-programmed reduction profiles (H2-TPR) of CeO2 supports and

Cu/CeO2 catalysts were collected on an online mass spectrometer (Cirrus II, MKS

Instruments, 1–300 amu) operated in selective ion detection mode. 0.1 g of catalyst sample was packed inside a ¼” OD U-tube reactor made of quartz using quartz wool plugs. The catalyst was pretreated in He at 350oC for 30 minutes followed by cooling at room temperature in He. Then 5% H2/He (35 ccm) was introduced to the reactor at room temperature and the mass signals were allowed to stabilize for at least 30 minutes before ramping the temperature at 10°C/min to 550°C.

N2O chemisorption technique outlined earlier by Jensen et al. [92] was used to determine the dispersion of Cu over CeO2 supports by oxidation of surface Cu atoms to

Cu2O. For each run, 150 mg of sample was packed in a fixed bed quartz reactor with a quartz frit bed which is placed inside a fast response furnace (Carbolite, MTF 10/15/130) and pretreated with 5%H2/He (30 ccm) at 250°C for 2 hours. The reactor was then flushed with He at the same temperature and cooled under the same atmosphere. N2O chemisorption was carried out by introduction of a 4%N2O/He containing stream to the reactor at 40°C. The effluents in the m/z=12 to m/z=46 range were monitored via an on- line mass spectrometer (MKS – Cirrus II) and throughout the experiments N2O and N2

64 were the only species detected in the reactor effluent. The mass spectrometer was calibrated for instrumental sensitivity factors and the contribution of m/z=28 fragment of

N2O to the m/z=28 trace. For the quantitative determination of N2 evolving from the surface, a known amount of N2, which approximately matches the amount of N2 obtained during N2O chemisorption runs, was injected under the same flow conditions using an external sampling loop and a six-port valve using the same experimental set-up with a blank reactor. The number of moles of oxygen consumed is calculated through N2 evolution and gives the total amount of oxygen consumed via the surface reaction and oxygen uptake through Fickian diffusion of oxygen atoms from the surface to the bulk. A plot of Cu conversion versus square root of time gives a straight line once the surface is covered, and can be used for diffusion correction [92].

In-situ diffuse reflectance Fourier transform spectroscopy (DRIFTS) experiments were performed during WGS reaction on Cu/CeO2-NP(E) and Cu/CeO2-NR(E) catalysts to study the effect of support morphology on the formation of adsorption products and reaction intermediates. The FT-IR spectrometer (Thermo Nicolet 6700) was equipped with an MCT detector and a DRIFTS chamber with ZnSe windows. The DRIFTS spectra were collected in the mid-IR range at a nominal resolution of 4cm-1 and the presented spectra were averaged over 500 scans. The catalyst was pretreated in He (30 sccm) for 30

o min at 350 C, then reduced in 5%H2/He (30 sccm) at the same temperature for 1h.

Afterwards, the reaction chamber was purged with He (30 sccm) at 400oC for 30 min.

The background spectra were collected in He at regular temperature intervals while cooling down. The sample spectra were acquired under a stream of 1% CO and 1% H2O at temperature intervals of 50oC up to 400 oC.

65

5.3. Results and Discussion

5.3.1 Catalyst Activity

Figure 5.1a shows CO conversions as a function of temperature during water gas- shift reaction over Cu-based catalysts supported on CeO2 nano-particles. Copper was impregnated over CeO2 support via a wet-impregnation route using either ethanol (E) or

THF (T) as the impregnation medium. The bare CeO2-NP support was also tested under the same conditions for WGS activity and was observed to show no activity for WGS.

The best performing Cu/CeO2 catalyst achieved CO conversions that are comparable to our previously reported Fe-Al-Cu catalyst [93], reaching 36% and 49% at 400°C for THF and ethanol impregnated samples, respectively. Furthermore, the CO conversions observed over Cu/CeO2-NP(E) were, on the average, 40% higher than those achieved over the commercial Fe-Cr-Cu catalyst tested under the same conditions in 300-400°C range [19, 20].

The WGS activity (% CO conversion) achieved over CeO2-NR based samples was observed to be significantly lower than that of the CeO2-NP supported catalysts

(Figure 5.1b). The bare CeO2 nano-rod support showed no activity. Below 350°C, copper impregnation onto CeO2 nano-rods had no practical effect on the WGS activity and above that temperature the catalysts were able to achieve only marginally higher activity than the bare support. A maximum CO conversion of 15% was achieved over the nano-rod supported samples at 400°C, regardless of the impregnation medium. The rates of reaction were calculated at 400°C. In line with the above results, the NP-based samples

66

-2 -1 showed the highest rates of reaction (0.29 µmol m s for Cu/CeO2-NP(T) and 0.33

-2 -1 µmol m s for Cu/CeO2-NP(E)), while the rates of reaction over Cu/CeO2-NR(T) and

-2 -1 -2 -1 Cu/CeO2-NR(E) were calculated to be 0.10 µmol m s and 0.11 µmol m s , respectively.

The best performing catalyst, Cu/CeO2-NP(E), was further tested for stability with time-on-stream. Figure 5.1c shows the CO conversion attained over this catalyst as a function of time-on-stream at 400°C under the same feed conditions used in catalyst screening studies. The CO conversion over this catalyst was stable at 46% over a period of 55 hours.

As it will be discussed in the following sections, the support morphology has a significant effect on both the size of copper clusters on the surface, their redox behavior and resistance to sintering under reducing conditions that accounts for the significant difference in the WGS activity of Cu/CeO2 catalysts. The impregnation medium was also observed to have an effect on the copper segregation over these catalysts that could be correlated with the activity differences.

67

Figure 5.1: Activity of (a) CeO2-NP, (b) CeO2-NR based WGS catalysts: (●) CeO2, (■)

Cu/CeO2(E) and (▲) Cu/CeO2(T) and (c) time-on-stream activity of Cu/CeO2-NP(E) at

400°C. Reaction conditions: [CO] = 10%, [H2O] = 10%, [CO2] = 5%, [H2] = 7.5% and

3 -1 -1 balance N2, WHSV = 0.06 m (g cat) h , P = 1 atm.

68

5.3.2 Transmission Electron Microscopy

Figure 5.2 shows the TEM images of the cerium oxide nano-particles and nano- rods used as the supports. The CeO2-NR is comprised of nano-rods with diameters of

5nm. Figure 5.2b shows a high resolution TEM image of a representative CeO2 nano-rod with the lattice fringes for the [2 0 0] plane of CeO2 with an interplanar spacing of 2.7Å.

The CeO2-NP is comprised of nano-particles with diameters approximately 8nm (Figure

5.2c). Figure 5.2d shows a representative CeO2 nanoparticle with [1 1 1] plane orientation with an interplanar spacing of 3.1Å. Copper impregnation and subsequent heat treatments did not result in any change in the morphology of any of the supports that was observable through TEM (not shown).

69

Figure 5.2: TEM images of (a) CeO2-NR and (b) CeO2-NR at higher magnification, (c)

CeO2-NP and (d) CeO2-NP at higher magnification.

70

5.3.3 X-ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy was utilized to investigate the chemical states of cerium and copper. Figures 5.3a and 5.3b show the Ce 3d core level spectra of the

CeO2 nano-particles and nano-rods, respectively. The fairly complex structure of the core level Ce 3d spectra is due to the presence of both Ce4+ and Ce3+ derived modes, showing deviations from the ideal stoichiometry. The structures labeled as v (882.6 eV), v´´(889.1 eV) and v´´´ (897.1 eV) and the corresponding u modes – u (901.2 eV), u´´ (907.7 eV) and u´´´ (915.1 eV) – are characteristic of CeO2 while v0 (880.8 eV)and v´ (886.9 eV) and the corresponding u modes – u0 (899.3 eV) and u´ (905.8 eV) are associated with

Ce2O3 [94-96]. No changes were observed in the core level Ce 3d spectra following the impregnation of copper on to the CeO2 supports (Figures 3c through 3f). Table 2 shows the relative contributions of Ce3+ and Ce4+ that were calculated through peak fitting (not shown) and using the areas under the fitted components. The Ce3+-to-Ce4+ ratio was observed to be independent of the particle morphology and was fairly constant over the copper impregnated samples.

The core level Cu 2p spectra of the Cu/CeO2 catalysts are shown in Figure 5.4.

For the CeO2 nano-particle supported catalysts, curve fitting in the Cu 2p3/2 region showed the presence of four components located at, 932.3 eV, 933.2 eV, 941.0 eV and

943.1 eV. The lowest binding energy component can be assigned to Cu1+ or Cu0 species, however exact spectral identification of this species through the photoelectron lines is not possible as the difference in the core level binding energies of copper atoms in Cu2O and metallic copper environments is well within experimental uncertainty. The 2p3/2 -2p1/2

71 spin-orbit-splitting (SOS) for this mode was 19.8 eV. The higher binding energy components with 933.2 eV being the main photopeak and 941.0 eV and 943.1 eV being the shake-up satellite lines, are characteristic of Cu2+ ions in CuO environment [77, 78].

The Cu 2p1/2 component for this species was located at 953.0 eV (SOS=19.8 eV) with a satellite peak resolved at 961.7eV (ΔEB = 8.7 eV). Over the CeO2 nano-rod-supported catalysts, the Cu 2p3/2 region was composed of three components located at 933.2 eV,

941.0 eV and 943.1 eV, which are all, in line with the above discussion, assigned to Cu2+ species with no contribution from lower oxidation state copper species.

The relative contributions of the two copper species to the Cu envelope was calculated from the areas under the fitted components and presented in Table 5.2.

Reduction of copper oxide under the X-ray beam in the high vacuum environment of the

XPS instrument is a well known phenomenon [79, 80]. It is plausible that reduction of

CuO taking place under the X-ray beam resulted in the appearance of the lower oxidation state copper species over the CeO2 nano-particle-supported samples. As will be discussed below, the relative contribution of the reduced copper phase over these samples is well correlated with the dispersion of copper over the samples and the WGS activity. The nano-particle supported catalysts were comprised of well-dispersed copper domains, which tend to get reduced more easily than larger particles, while larger copper crystallites were present over the nano-rod supported sample. A control experiment was carried out by collecting the X-ray photoelectron spectra over bulk CuO using the same experimental parameters as the Cu/CeO2 catalysts and formation of the reduced copper phase was not observed over this sample (not shown). The absence of a reduced copper component over bulk CuO further supports the conclusion that the presence of a reduced

72 copper phase over the nano-particle supported catalysts was due to the reduction of small clusters of copper in the high vacuum environment of the XPS instrument.

73

Figure 5.3: Ce 3d region of the X-ray photoelectron spectra of (a) CeO2-NP, (b) CeO2-

NR, (c) Cu/CeO2-NP(E), (d) Cu/CeO2-NP(T), (e) Cu/CeO2-NR(E) and (f) Cu/CeO2-

NR(T)

74

Figure 5.4: Cu 2p region of the X-ray photoelectron spectra in (a) Cu/CeO2-NP(E), (b)

Cu/CeO2-NP(T), (c) Cu/CeO2-NR(E) and (d) Cu/CeO2-NR(T)

75

Ce3+ Ce4+ Cu2+ Cu1+(or Cu0)

CeO2-NP 49 % 51 % - -

CeO2-NR 48 % 52 % - -

Cu/CeO2-NP(E) 47 % 53 % 42 % 58 %

Cu/CeO2-NP(T) 49 % 51 % 55 % 45 %

Cu/CeO2-NR(E) 46 % 54 % 100 % -

Cu/CeO2-NR(T) 44 % 56 % 100 % -

Table 5.2: XPS analysis of the catalysts

76

5.3.4 X-ray diffraction

Figures 5.5a and 5.5b shows the X-ray diffraction patterns of the CeO2-NP and

CeO2-NR supports collected under ambient conditions. Both of the patterns show diffraction lines at 2Θ values of 28.8°, 32.9°, 47.4°, 55.9° and 58.6° that could be associated with [1 1 1], [2 0 0], [2 2 0], [3 1 1] and [2 2 2] planes of CeO2 in face- centered cubic cerianite structure (ICDD #34-394), respectively. A comparison of the relative intensities of the diffraction peaks show that the relative intensity of the [1 1 1]

(at 2Θ = 28.8º) peak is larger in the XRD pattern of nano-rods as compared to nano- particles. The ratio of the intensities of peaks associated with [1 1 1] (at 2Θ = 28.8º) and

[2 2 0] (at 2Θ = 47.4º) calculated from the XRD patterns of CeO2-NP and CeO2-NR showed that the [1 1 1] plane exposure in the nano-rods is 1.5 times higher than in nano- particles. This observation is in agreement with the literature [54, 62], where different ceria nano-structures have been reported to have different crystal plane exposures: [1 1 1] and [1 0 0] are the exposed crystal planes for nano-polyhedra, while nano-rods have [1 1

0] and [1 0 0] plane exposures. The crystallite sizes of the CeO2 samples were calculated from the broadening of the [1 1 1] diffraction peak of CeO2 (at 28.8°) using Scherrer’s method. The crystallite sizes of CeO2-NP and CeO2-NR were calculated as 8 and 6 nm, respectively, which are in line with TEM results.

Figures 5.5c and 5.5d show the X-ray diffraction patterns of Cu/CeO2-NR catalysts impregnated in ethanol or THF media, respectively. The cerianite phase was preserved after impregnation of copper on to the support material and following heat

77 treatment. In addition to the cerianite phase, both of the diffraction patterns showed the formation of copper oxide domains which are observed through diffraction lines at 2Θ values of 35.9° and 38.7°. These diffraction peaks can be associated with [1 1 ] and [1

1 1] planes of CuO in monoclinic tenorite structure (ICDD # 48-1548). Similarly,

Figures 5.5e and 5.5f show the diffraction patterns of Cu/CeO2-NP catalysts impregnated in ethanol and THF media, respectively. Over these catalysts cerianite phase was observed to be the major crystalline phase after copper impregnation. Unlike the nano- rod supported catalysts, diffraction peaks associated with a separate CuO phase were either non-existent or very weakly resolved over these catalysts suggesting the presence of smaller CuO particles.

Neither crystallite size calculations using Scherrer equation nor TEM images of the Cu/CeO2 catalysts (not shown) showed changes in the ceria particle size upon impregnation of the CeO2 supports with copper. Calculation of CuO crystallite size in

Cu/CeO2-NR sample was not attempted because of the low intensity of the CuO diffraction peaks.

The evolution of crystal phases during reduction of Cu/CeO2-NP(E), Cu/CeO2-

NP(T) and Cu/CeO2-NR(E) catalysts was studied by in-situ XRD as a function reduction temperature and the diffraction patterns collected during these experiments are presented in Figure 5.6. In agreement with the foregoing discussion, the peaks at 28.8º, 32.9º, 47.4º,

55.9º and 58.6º were associated with diffractions from [1 1 1], [2 0 0], [2 2 0], [3 1 1] and

[2 2 2] crystal planes of ceria (cerianite, ICDD #34-394), respectively. The CeO2 phase was observed to be the major crystalline phase present over all the catalysts throughout the temperature range studied and no contribution from sub-oxides or metallic cerium

78 was identified. In line with the XRD patterns collected under ambient conditions, an additional peak that has been associated with the presence of crystalline CuO domains

(tenorite, ICDD # 48-1548), was resolved at 35.9º [1 1 ] in the XRD patterns collected at 30ºC. As the reduction proceeds, a new diffraction peak was resolved at 43.5º above

200°C. This new peak can be associated with diffraction from [1 1 1] plane of Cuº (ICDD

# 4-836). This peak is well-resolved over Cu/CeO2-NR(E) and Cu/CeO2-NP(T) catalysts and grows in intensity up to 300ºC, suggesting the formation of larger, more ordered crystallites of metallic copper with increasing reduction temperature. The metallic copper phase is likely to be formed through the reduction of the crystalline CuO particles that were already present over these catalysts rather than agglomeration of dispersed CuO species as this phase is only very weakly resolved over Cu/CeO2-NP(E) catalyst where

CuO is more finely dispersed relative to the other two catalysts.

79

Figure 5.5: X-ray diffraction patterns of (a) CeO2-NR, (b) CeO2-NP, (c) Cu/CeO2-NR(E)

(d) Cu/CeO2-NR(T) (e) Cu/CeO2-NP(E) and (d) Cu/CeO2-NP(T).

80

Figure 5.6: X-ray diffraction patterns collected during in-situ reduction with 5%H2/N2 of

(a) Cu/CeO2-NP(E), (b) Cu/CeO2-NR(E) and (c) Cu/CeO2-NP(T)

81

5.3.5 Temperature-Programmed Reduction

The reduction behaviors of CeO2 supports and Cu/CeO2 catalysts impregnated in ethanol and THF media are studied using temperature-reduction technique and the results are presented in Figure 5.7. In the 50ºC-550ºC range, no H2 consumption features are observed in the H2 consumption traces of the CeO2 supports showing that, regardless of the particle morphology, CeO2 is not reduced under these conditions (Figures 7a and 7b).

This finding is in agreement with the in-situ XRD-TPR results discussed in the previous section where cerianite (CeO2) was observed to be the only crystalline phase in the 30-

450°C range. Regardless of the impregnation media, Cu/CeO2-NR supported copper catalysts exhibit a broad H2 consumption band in the 180ºC-450ºC range (Tmax = 285ºC) with a well-resolved shoulder in the high temperature range (Figures 7c and 7d). Figures

7e and 7f show the H2 consumption traces of Cu/CeO2-NP catalysts impregnated in ethanol and THF media. Both of the traces show a single H2 consumption feature in the

150ºC-250ºC region. The maximum corresponds to 205°C and 225°C for Cu/CeO2-

NP(E) and Cu/CeO2-NP(T), respectively and both traces show a shoulder in the low temperature range. The complex structure of the H2 consumption profiles of both of the samples suggests the presence of multiple species with different reduction potentials.

H2-TPR is extensively utilized for characterization of copper catalysts over ceria and the nature of copper species giving rise to the H2 consumption maxima observed in the TPR profiles is widely debated [26, 28, 35, 39, 51, 97-99]. The H2-consumption features located at lower temperatures (150ºC - 200ºC) are usually associated with well- dispersed copper oxide particles that are in close interaction with CeO2 and are

82 amorphous in nature [97, 98]. The H2 consumption peaks observed above 200ºC are attributed to the presence of larger and isolated crystalline CuO particles [26, 35, 39, 51,

97, 98]. Based on these studies, the low temperature H2 consumption features observed over both Cu/CeO2-NP(E) and Cu/CeO2-NP(T) were assigned to the reduction of finely- dispersed CuO species in close interaction with the CeO2 support. Avgouropoulos and

Ioannides [51] showed that the reduction temperature and the crystallite size of the CuO domains are correlated and that reduction temperature of the copper oxide domains increase with increasing cluster size. The lower temperatures at which the H2 consumption features were observed over CeO2-NP supported catalysts suggest the formation of smaller CuO particles compared to the catalysts supported over CeO2-NR.

The inset of Figure 5.7 shows the H2 consumption profile of bulk CuO during H2-

TPR carried out under the same experimental conditions. Under these experimental conditions, reduction of CuO takes place in 270ºC-600ºC region and goes through a broad maximum around 450ºC. The high temperature shoulder (300ºC-450ºC) in the H2 consumption profile of Cu/CeO2-NR can be associated with the reduction of crystalline copper oxide domains to form metallic copper. The transformation of these copper oxide particles to metallic copper was also observed through the in-situ XRD patterns collected during the reduction of Cu/CeO2-NR(E) (see Figure 5.6b). The slightly lower temperature for the appearance of metallic copper phase in the in-situ XRD patterns when compared to the temperature for the H2 consumption feature associated with the reduction of bulk- like CuO phase is due to the truly transient nature of the TPR experiment versus the stepped temperature profile utilized for the collection of in-situ XRD patterns. These

83 findings are in agreement with various studies in the literature that document the effect of ceria in enhancing the reducibility of copper [28, 35, 99, 100].

84

Figure 5.7: Temperature-programmed reduction profiles of (a) CeO2-NR, (b) CeO2-NP,

(c) Cu/CeO2-NR(T) (d) Cu/CeO2-NR(E) (e) Cu/CeO2-NP(T) and (f) CeO2-NP(E) in 5%

H2/He. Inset: Temperature-programmed reduction profile of bulk CuO.

85

5.3.6 N2O Chemisorption

The dispersion of Cu over the Cu/CeO2 catalysts was studied using N2O as a probe. The results of the N2O chemisorption measurements are presented in Table 5.3.

The results of these measurements further support the results from H2-TPR and XRD studies showing the presence of well-dispersed copper species over Cu/CeO2-NP(E) and

Cu/CeO2-NP(T). The Cu dispersion values are significantly lower for the CeO2 nano-rod- supported catalysts suggesting the formation of larger copper domains over this support.

The catalyst characterization studies showed the presence of copper species with different degrees of dispersion over CeO2 surface, ranging from large copper oxide crystallites to finely dispersed copper oxide particles. They also pointed to a correlation between copper dispersion and catalytic activity. It is also important to note that the possibility of copper occupying the interstitial sites within the ceria matrix cannot be ruled out; however, the preparation technique, i.e. wet impregnation of copper over the ceria support and the processing temperatures make copper migration into the ceria matrix is rather unlikely.

The presented results suggest that CeO2 nano-particles are capable of stabilizing copper oxide as finely-dispersed copper oxide particles which can be transformed into finely- dispersed metallic copper particles under reducing conditions. These finely dispersed clusters harbor the active sites for WGS reaction over Cu/CeO2 catalysts. Over the CeO2 nano-rods, copper oxide forms larger domains and is also present as bulk-like CuO species that get converted to large Cuº particles under reducing conditions. These particles exhibit poor WGS activity. Similar results on the dependence of WGS activity

86 of supported copper catalysts on the size of CuO domains, hence on the reducibility of

CuO has been reported previously [26, 33].

87

Cu dispersion Cu surface atoms

(%) (µmol/g cat)

Cu/CeO2 –NP(E) 18.0 129.6

Cu/CeO2 –NP(T) 13.8 99.4

Cu/CeO2 –NR(E) 1.2 8.6

Cu/CeO2 –NR(T) 1.0 7.2

Table 5.3 Cu-dispersion values estimated from N2O chemisorption

88

5.3.7 In-situ DRIFTS during WGS

The effect of ceria morphology on WGS reaction over Cu/CeO2 catalysts was further investigated by in-situ DRIFT spectroscopy during WGS reaction over Cu/CeO2-

NP(E) and Cu/CeO2-NR(E). Introduction of WGS reaction feed mixture over a pre- reduced Cu/CeO2-NP(E) sample at 50°C resulted in the formation of a sharp absorption band at 2105 cm-1 which can be associated with ν(C=O) vibrations of Cu-carbonyl species [101, 102]. The 2105 cm-1 band disappeared above 200°C after going through a maximum between 100°C-150°C. Due to their high thermal stability, the species were assigned as Cu+ carbonyls.

The 1700-1200 cm-1 region is the fingerprint region for asymmetric and symmetric stretching modes of O-C-O moieties which are present in formate, carbonate and carboxylate species. The complex structure of the spectrum in this region suggests the formation of several different surface species. Following introduction of WGS reaction feed at 50°C, bands located at 1638 cm-1 (shoulder), 1550 cm-1, 1390 cm-1

(shoulder), 1313 cm-1, 1050 cm-1 and 865 cm-1 were resolved. The 1638 cm-1 band decreased significantly in intensity with increasing temperature and disappeared above

100°C. This band was accompanied with a broad band centered around c.a. 3400 cm-1 in the hydroxy vibration region (νOH) which followed the same trend with temperature.

-1 Therefore, the shoulder band at 1638 cm was associated with deformation modes (δHOH) of undissociated water molecules on the surface, rather than νas(C=O) mode of bidentate carbonates. In the absence of bands in 2950-2850 cm-1 region, the absorption bands in

1600-1200 cm-1 region were associated with carbonate species rather than formates. The

89 bands located at 1055 cm-1, 1313 cm-1 and 1550 cm-1 were assigned to bidentate carbonate species [22, 103, 104] while the 1390 cm-1 band was associated with carbonates that are likely to be uni-dentate in nature [22]. The 1390 cm-1 band increased in intensity with increasing temperature and above 100°C the νs(OCO) mode of these

-1 -1 species were resolved at 1084 cm [105]. The 865cm band was assigned to π(CO3) mode of carbonates. 1215 cm-1 band observed in 150°C-250°C region may be attributed to formation of hydrogen carbonate species on the surface [103]. These results suggest that WGS reaction over Cu catalysts supported on nano-structured CeO2 particles proceeds through a red-ox mechanism rather than a formate mechanism and are in agreement with the findings of Wang et al. [32].

Figure 5.9 shows in-situ DRIFT spectra collected during a similar experiment over Cu/CeO2-NR(E). The significantly lower amount of surface species over this catalyst relative Cu/CeO2-NP(E) catalyst was apparent from the fewer absorption bands resolved in the DRIFT spectra. Over this catalyst the temperature window for the formation of copper carbonyl species was shifted by 100°C to higher temperatures and the intensity of the 2105 cm-1 band was significantly smaller, suggesting formation of a lower amount of copper carbonyl species. It is important to note that, the intensities and the temperature windows of appearance of the Cu+-carbonyl bands observed during WGS reaction over Cu/CeO2 catalysts is well correlated with the catalytic activities of these catalysts. Over Cu/CeO2, Gamarra et al. [102] observed a similar correlation between the intensity of the Cu+-carbonyl band and CO oxidation rate during preferential oxidation of

CO. These authors further suggested that the intensity of Cu+-carbonyl band was an indication of the reduction potential of the interfacial sites over dispersed copper oxide,

90 which serve as the active sites for CO oxidation. In line with the findings of these authors, the well dispersed copper entities, the presence of which have been shown by various characterization techniques over Cu/CeO2-NP(E) catalyst, were associated with high WGS activity. The high WGS activity of the catalyst with high copper dispersion together with the results of the DRIFTS experiments imply the presence of a red-ox mechanism which is facilitated by the higher reduction-oxidation potential of small, well- dispersed copper species.

91

Figure 5.8: In-situ DRIFT spectra collected during WGS reaction over Cu/CeO2-NP(E)

(a)high wavenumber region and (b)low wavenumber region.

92

Figure 5.9: In-situ DRIFT spectra collected during WGS reaction over Cu/CeO2-NR(E)

(a)high wavenumber region and (b)low wavenumber region.

93

5.4. Conclusions

Nano-structured CeO2 supports with two different particle morphologies, i.e. nano- particles and nano-rods, were prepared and wet impregnated with copper in two different impregnation media, ethanol and tetrahydrofuran. The morphology of the support was observed to have a strong effect on the WGS activity of the catalysts while the impregnation medium did not have a significant effect on the activity. Copper-catalyst supported over CeO2 nano-particles showed significantly higher activity than its CeO2 nano-rod-supported counterpart while neither bare CeO2 nano-particles, nor CeO2 nano- rods showed activity towards the WGS reaction. The effect was associated with the capability of CeO2 nano-particles in stabilizing CuO in a highly-dispersed state as the

Cu/CeO2-NP catalyst was found to have finely-dispersed CuO particles whereas large, crystalline CuO domains were present over ceria nano-rod-supported samples. Under reducing conditions, the bulk-like CuO domains were transformed into large metallic copper crystallites, which were associated with poor WGS activity of this catalyst.

* Information in this chapter is adapted/taken from P. Gawade, B. Mirkelamoglu and

U.S. Ozkan, Journal of physical Chemistry C 114 (2010) 18173-18181.

94

PART 2:

PREFERENTIAL OXIDATION OF CARBON MONOXIDE (PROX) IN

HYDROGEN-RICH STREAMS

95

CHAPTER 6

INTRODUCTION TO PREFERRENTIAL OXIDATION OF CO (PROX) IN

HYDROGEN

Proton exchange membrane (PEM) fuel cells are popular for their energy efficiency, low operating temperature, compactness and zero-emission. However, the Pt anode used in PEM fuel cells is highly sensitive to CO poisoning and even ppm level CO can significantly compromise PEM performance. Preferential oxidation of CO (PROX) is an important reaction to serve as a connecting link between low-temperature water-gas-shift

(LT-WGS) and the proton exchange membrane (PEM) fuel cell. It offers an economical and efficient technique for bringing down the carbon monoxide (CO) concentration from

0.5-1% to parts per million (ppm) level and purifies the hydrogen stream. Apart form this technique, there are other methods such as methanation, selective membrane separation and pressure swing adsorption, which are proposed for hydrogen purification application.

However, the ease of operation and implementation associated with PROX method makes it more promising than its other counterparts.

PEM fuel cell is operated around 80 oC [106] whereas, LT-WGS operating range is

200-250 oC. Therefore, 80-200 oC is an ideal temperature window for intermediate

PROX step. However, one of the challenges in this temperature range is to selectively

96

oxidize CO with minimal loss of H2 during side reactions. The main as well as the

possible side reactions during PROX can be summarized as follow [107]:

1 KJ CO oxidation: CO + O →CO ΔH 0 = −283 (1) 2 2 2 mol

1 0 KJ H2 oxidation: H + O →H O ΔH = −242 (2) 2 2 2 2 mol € € KJ WGS: CO + H O ↔ CO + H O ΔH 0 = −41.3 (3) 2 2 2 mol € € KJ Methanation: CO + 3H ↔ CH + H O ΔH 0 = −206 (4) 2 4 2 mol € € KJ Methanation: CO + 4H ↔ CH + 2H O ΔH 0 = −165 (5) 2 2 4 2 mol € €

€ Amongst the above-mentioned€ reactions, reaction (1) and (3) are desirable

reactions, whereas, reactions (2) and (4-5) are considered unfavorable for PROX

operations. It should be noted that although methanation reaction assists further removal

of CO from the reformate feed, it also causes the loss of hydrogen and hence brings down

the energy value from the reformate feed [108].

* Information in this chapter is adapted/taken from, P. Gawade, B. Bayram, A.M.

Alexander and U.S. Ozkan, Applied Catalysis B, Environmental, In Review.

97

CHAPTER 7

LITERATURE REVIEW

7.1 Precious Metal Catalysts

Noble metals such as Pt [107, 109-111], Rh/Ru [112], and Ir/Pd [113] are widely studied as PROX catalysts. Pt and Pt-X (X= Sn, Ce, Co and Ni) supported on various supports, such as Al2O3, CeO2, ZrO2, CeO2-ZrO2, MgO, activated carbon, La2O3, SiO2 and SiO2-Al2O3 [107, 111, 113-122] are widely investigated for their PROX activity.

Pt/CeO2 is a promising formulation in terms of PROX performance over Pt/Al2O3 due to the excellent oxygen storage capacity (OSC) of CeO2 [107, 115], which contributes to better CO oxidation. Apart from noble metals, gold is probably as widely studied as Pt- based catalysts for PROX [123-127] due to its higher availability and lower price than Pt

[128]. Au supported on α-Fe2O3, CeO2, γ-Al2O3, ZrO2, TiO2, Co3O4, NiO, MgO and

MnO have been investigated for their PROX activity [125, 126, 128-130]. Among these supports, Au/TiO2, Au/CeO2 and Au/α-Fe2O3 have promising PROX activity [125, 126,

131], especially Au/α-Fe2O3 prepared through co-precipitation method. This catalyst can achieve comparable PROX activity to that of commercially available Pt/γ-Al2O3 even at

98 low temperatures (80 oC) [126, 128]. In spite of the promising performance of both Pt- and Au-based catalysts for PROX, they need to be replaced due to cost factor.

7.2 Non-Precious Metal Catalysts

Copper and cobalt are probably the most commonly used non-precious transition metals as PROX catalyst. CuO/CeO2 and mixed oxides of CuO/ CeO2 are the most widely reported formulations for PROX [132-138]. The synergetic effect between CuO and CeO2 is known to provide high PROX activity [136, 138]. The excellent redox property of ceria support assists the CO oxidation, thereby provides higher activity. The

+4 addition of ZrO2 and Sn in CeO2 further improves its redox property and thermal resistance of the support [136, 139-141]. Chen, et al. [140, 141] found that

CuO/Ce0.9Zr0.1O2 and CuO/Ce0.9Sn0.1O2 had better PROX activity than CuO/ CeO2 due to improved redox property of CeO2.

Cobalt oxide (Co3O4) based catalysts are widely studied for CO oxidation in the absence of hydrogen [142-147]. Both, pure Co3O4 and Co3O4/Al2O3 have been reported for high CO oxidation in the absence of hydrogen [145, 146, 148]. When compared with other transition metals, cobalt oxide has better CO oxidation activity as well as selectivity over a wide temperature range in the presence of hydrogen [149]. Several reports in the past [127, 150-154] have suggested that cobalt oxide can be a promising candidate for

PROX application. It can either provide active sites or can be used as co-dopant in other catalyst formulations. As a co-dopant, it is known to enhance the PROX activity of

Au/CeO2 and Pt/ CeO2. In addition, Omata et al. [155] showed that cobalt supported on

99 alkali earth metal carbonate, such as Co/SrCO3 has excellent PROX activity in the presence of excess hydrogen.

Both, experimental and DFT studies have shown that Co+3 is an active site for CO

+3 oxidation [146, 156, 157]. Co occupies the octahedral positions in Co3O4, which can provide oxygen for CO oxidation resulting in the formation of Co+2. This lower oxidation state cobalt is known to occupy tetrahedral position of Co3O4, which may be re-oxidized

+2 in the presence of oxygen. However, the de-activation of Co3O4 may be observed, if Co fails to re-oxidize [146].

* Information in this chapter is adapted/taken from, P. Gawade, B. Bayram, A.M.

Alexander and U.S. Ozkan, Applied Catalysis B, Environmental, In Review.

100

CHAPTER 8

PROX OVER CoOx/CeO2 IN HYDROGEN-RICH STREAMS: EFFECT OF COBALT

LOADING

8.1 Overview of PROX over CoOx/CeO2

Our group [151, 152] in the past had studied the various potential supports for Co, such as ZrO2, CeO2, Al2O3, SiO2 and TiO2 for their PROX performance. 10% (%wt)

CoOx catalysts were prepared over commercially available supports using wet- impregnation method. Interestingly, in spite of having extremely low surface area (7

2 m /g) and low Co-dispersion (<0.1%), Co/CeO2 had shown promising PROX activity.

Woods et al. [158] further extended this effort to prepare a high surface area 10%

CoOx/CeO2 using precipitation and incipient wetness impregnation (IWI) method. The catalyst had high specific surface area (78 m2/g) and significantly higher Co-dispersion

(3.4%) over nanoparticles of CeO2, thereby providing a much-improved PROX activity than its previous counterpart.

This work is an extension of our previous studies on PROX catalysis over Co/CeO2 catalysts. The catalyst preparation method used offers a simpler and cleaner way to prepare high surface area nanoparticles of CoOx/CeO2. The effect of Co-loading is

101 examined under hydrogen rich conditions as well as in the presence of CO2 and water vapor.

8.2 Experimental Procedure

8.2.1 Catalyst Preparation

High surface area ceria nano-particles were prepared through a precipitation technique using cerium (III) nitrate hexahydrate (99.999%, Sigma-Aldrich) and sodium hydroxide. An aqueous solution of cerium nitrate hexahydrate (0.4 M) was mixed with an aqueous solution of sodium hydroxide (2.7 M) at room temperature to form precipitates of Ce(OH)2. The resulting precipitate was aged at room temperature for 24 h followed by washing in de-ionized water until the pH of the filtrate was neutral. The precipitate was, then, dried at 110 oC overnight, followed by calcination at 400 oC for 3 h. Cerium oxide support prepared through the precipitation method outlined above yields nanoparticles approximately 8 nm in diameter [159].

CoOx/CeO2 catalysts with different cobalt loadings were prepared through a wet- impregnation method using aqueous medium. Aqueous solutions of cobalt nitrate hexahydrate of desired concentrations corresponding to 1%, 2% and 10% CoOx (%wt) loading on the resultant catalyst were prepared at room temperature followed by addition of ceria support into it under constant stirring. The solution was stirred rigorously for 24 h followed by drying at 110 oC. Finally, the catalysts were calcined at 400 oC for 3 h, 10 oC/min ramp.

102

8.2.2 Catalyst Activity Testing

The schematic of steady state reaction system for PROX is shown in Figure 8.1.

Steady-state reactions were performed in a fixed bed reactor flow system using ¼’ OD stainless steel reactor. For each run, the catalyst was packed inside the reactor using quartz wool and was heated using a homemade, resistively-heated furnace. The reaction temperature was measured and controlled by using a K-type thermocouple and an Omega

CSC232 PID temperature controller. The reaction feed as well as effluent from the reactor was monitored using on-line micro-GC (Agilent 3000A) equipped with 0.32 mm

PLOT mol-sieve and PLOT Q columns with thermal conductivity detectors (TCD) using helium (or argon, when appropriate) as the carrier.

Prior to catalytic activity testing, catalysts were pretreated in situ in 10% O2/He

(30ccm) at 300 oC for 30 min, followed by cooling to the desired reaction temperature under He flow. Steady-state PROX reaction studies were performed in the temperature range of 100-175 oC at a WHSV (weight hourly space velocity) of 15,000 cm3 (g cat-1) h-

1 . The feed composition was 1% CO, 1% O2, 60% H2 and balance helium.

CO and H2 oxidation experiments were performed over CoOx/CeO2 catalysts using 2% CO (or H2), 2% O2 in balance helium (or argon). For these experiments, 50 mg sample was packed inside the reactor and the total flow rate was maintained at 130 ccm to ensure the differential reactor operating conditions. For H2 oxidation experiments, carrier gas to the micro-GC was replaced with argon to enable detection of hydrogen by the TCD.

103

The effect of water vapor and CO2 on the PROX activity of the Co/CeO2 catalysts was investigated in 125-175 oC range at a WHSV of 15,000 cm3 (g cat-1) h-1. The feed composition for these studies was 1% CO, 1% O2, 60% H2, 0.5% H2O (or 0.5% H2O and

5% CO2) in balance helium. Water-gas-shift (WGS) and reverse water-gas shift (RWGS) activity of the CoOx/CeO2 formulations were studied in the temperature range of 125-200 oC at a WHSV of 15,000 cm3 (g cat-1) h-1. The feed stream for these studies constituted of

1% CO, 1% H2O or 5% CO2, 60% H2 in helium.

104

Figure 8.1: Schematic of Steady State Reaction System for PROX

105

8.2.3 Catalyst Characterization

Micromeritics ASAP 2020 accelerated surface area and porosimetry instrument was used to measure the surface area of support as well as CoOx/CeO2 catalysts through

N2 physisorption method at liquid nitrogen temperature (77 K). Before N2-physisorption,

o the samples were de-gassed at 130 C for 12 h at a vacuum better than 3 µm Hg.

Brunauer-Emmett-Teller (BET) method was used to determine the specific surface areas of the samples.

The N2O chemisorption technique outlined earlier in detail [160] was used to obtain the dispersion of cobalt in the CoOx/CeO2 catalyst (shown in Table 1). This technique is based on the method developed by Jensen et al. [92] to study the dispersion of Cu over CeO2. For each run, 100 mg of sample was packed in a fixed bed quartz reactor with a quartz frit bed that is placed inside a fast response furnace (Carbolite,

MTF 10/15/130) and pretreated with 5% H2/He (30 ccm) at 400 °C for 2 h. The reactor was then flushed with He for another 30 minutes at the same temperature and cooled under helium to 40 °C. Chemisorption of N2O was carried out by introduction of 4%

N2O/He to the reactor at 40 °C. An on-line mass spectrometer (MKS-Cirrus II) was utilized to monitor the reactor effluents in the m/z = 12 to m/z = 46 range. Prior to the experiment, the mass spectrometer was calibrated for instrumental sensitivity factors and the contribution of the m/z = 28 fragment of N2O to the m/z = 28 trace. Throughout the experiments, N2O and N2 were the only species detected in the reactor effluent. The number of moles of N2 evolved was related to the moles of oxygen consumed and moles of cobalt oxidized by assuming a Co:O stoichiometry of 1:1. As described by Jensen et

106 al. [92], the oxygen uptake is not only due to surface oxidation of Co, but also due to subsurface diffusion. The technique includes a correction for the Fickian diffusion.

Rigaku X-ray diffractometer equipped with Cu Kα radiation source (λ=1.5418 Å) operated at 40 kV and 25 mA was used to obtain X-ray diffraction patterns of Co/CeO2

o catalysts and CeO2 support in the range of 20-70 . In situ X-ray diffraction was performed during temperature-programmed reduction of 10% CoOx/CeO2 on a Bruker-

D8 advanced X-ray diffractometer equipped with an Anton Paar HTK1200 oven using

Cu Kα radiation source (λ=1.5418 Å) operated at 40 kV and 50 mA. The diffraction

o patterns were collected in the temperature range of 30-700 C under a flow of 5% H2/He.

International Center for Diffraction Data (ICDD) database was used for the identification of the crystalline phases from the collected diffraction patterns.

Temperature-programmed reduction (TPR) of CoOx/CeO2 catalysts, as well as the support, was carried out under a flow of 5% H2/N2 (40ccm). For each run, 100 mg of sample was packed in a quartz reactor with a porous quartz frit. The sample was

o pretreated in 10% O2/He (40 ccm) at 400 C for 1 h, followed by purging with helium at the same temperature for 30 min. The sample was then cooled down to room temperature under the same He flow. Finally, H2-TPR profile was collected using online mass spectrometer (Cirrus II, MKS Instruments) in 5% H2/N2 while increasing the temperature at a rate of 10 oC/min.

Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was performed over CoOx/CeO2 samples during PROX. The surface species over CoOx/CeO2 as well as the gaseous components were detected using Thermo Nicolet 6700, FT-IR spectrometer equipped with an MCT detector. 28 mg of the sample was loaded to the

107 controlled atmosphere chamber of the DRIFTS instrument for each run. DRIFTS spectra were collected in mid-IR range with resolution of 4 cm-1 and averaged over 500 scans.

o The catalyst was pretreated in 10% O2/He (40 ccm) for 30 min at 300 C, followed by cooling of the sample under helium to the desired temperature. Then the background spectrum was collected under helium. Finally, PROX gases (1% CO, 1% O2, 60% H2 and helium) were introduced at the same temperature and sample spectra were collected after

10 min under the gas flow.

Kratos AXIS Ultra X-ray photoelectron spectrometer (XPS) equipped with a monochromatic Al-Kα radiation source (1486.7 eV) operated at 13 kV and 10 mA was used to study the chemical states of cobalt and/or cerium in the ceria support as well as the CoOx/CeO2 catalysts. A survey scan was collected in the range of 1400-0 eV for every sample at an electron pass energy of 80 eV before sweeping in the Co 2p, Ce 3d and O 1s regions at an electron pass energy of 20 eV. Analysis of the collected X-ray photoelectron spectra was performed using XPS Peak 4.1.

In situ X-ray absorption near edge (XANES) data over 10% CoOx/CeO2 was collected at the Co K-edge (7709 eV) at the insertion device beamline (10ID-B) of the

Materials Research Collaborative Access Team (MR-CAT) of the Advanced Photon

Source, Argonne National Laboratories. The measurements were made in transmission mode and the Si(111) monochromator was tuned down to eliminate the higher order harmonics in the beam. The catalysts were mixed with SiO2 at a ratio of 1:2 and finely ground to obtain a homogeneous mixture. A ~30 mg of the mixture was then pelletized in a 6mm polished steel die and placed inside a 5 cm long quartz tube (6.5 mm. ID) and supported with quartz wool plugs. The sample was then centered in an in situ XAFS

108 chamber that was fitted with Kapton® windows and allowed continuous flow of the reactants as well as the isolation of the catalyst sample.

XANES data were collected over the pristine 10% CoOx/CeO2 sample in the range of 7500-7750 eV at room temperature. Then the catalyst was pretreated in the

o presence of 5% O2/He (20 ccm) at 300 C for 30 min followed by purging the catalyst with helium at the same temperature for 10 min. The sample was then cooled down to

o 175 C under the same He flow. Finally, PROX feed, 0.5% CO, 0.5% O2, 40% H2, balance helium (200 ml/min) was introduced at 175 oC for 30 min, followed by in situ

XANES data collection at the same temperature. It should be noted that a similar experiment was also attempted over 2% CoOx/CeO2 however, lower cobalt concentration coupled with the highly absorbing CeO2 matrix did not allow collection of XAFS data with sufficient resolution for analysis.

8.3 Results and Discussion

8.3.1 Surface Area and Dispersion Measurements

Table 8.1 presents the specific surface area of CoOx/CeO2 catalysts with different cobalt loadings. The surface area was found be dependent on the amount of cobalt content and it decreased with increase in cobalt loading. Cobalt dispersion was ∼20 % regardless of cobalt loadings, suggesting a similar particle size of the Co phase over the support. It should be noted that these dispersion measurements were within ± 2% error.

Although the particles size seemed to be similar regardless of cobalt loadings, the extent

109 of the support surface populated with these particles or the size of the clusters made of these particles increased with loading.

110

Surface Area Co Dispersion Co Surface Atoms

(m2/g) (%) (µmol/g cat)

CeO2 123 - -

1%CoOx/CeO2 111 17 ± 2 29

2%CoOx/CeO2 106 18 ± 2 62

10%CoOx/CeO2 84 21 ± 2 362

Table 8.1: Specific surface areas and Co dispersions over CoOx/CeO2 catalysts

111

8.3.1 Steady State Activity of Performance of CoOx/CeO2

Figure 8.2a shows the CO conversion as a function of temperature over

CoOx/CeO2 catalysts with different cobalt-loadings during preferential oxidation of carbon monoxide. CO conversion was found to increase monotonically with increasing temperature regardless of the cobalt-content. 10% CoOx/CeO2 showed slightly better CO oxidation activity than the 2% CoOx/CeO2 catalyst especially at temperatures below 150 o C while, the PROX activity of 1% CoOx/CeO2 was significantly lower than its counterparts. The difference in the PROX activity diminished at higher temperatures, especially in case of 10% CoOx/CeO2 and 2% CoOx/CeO2, reaching to approximately

94% of CO conversion at 175 oC over both of these catalysts. In addition, long-term

o reaction study was performed over 10% CoOx/CeO2 for over 40 h at 175 C as shown in

Figure 8.2 (inset). The data indicated no significant drop in activity or selectivity after the steady state was established in the first 10 h. Co3O4 has been considered an excellent candidate for CO oxidation in the presence of hydrogen. However, excess amount of hydrogen under the PROX condition may result in lower oxidation state cobalt phases, including metallic cobalt, which may promote side reactions [149, 151, 152, 155, 157,

161]. Both these side reactions could result in a drop in the CO conversion as well as O2 selectivity to CO2. However, no such phenomenon was observed during the time-on- stream experiment indicating that Co3O4 phase was preserved even in the presence of excess hydrogen. This claim was further supported by the X-ray absorption study

(Section 8.3.5.5) where no lower oxidation states of cobalt phases were observed during

PROX reaction.

112

The selectivity of O2 towards CO2 is illustrated in Figure 8.2b. O2 selectivity was around 100% in the temperature range of 100-125 oC regardless of cobalt content, indicating negligible hydrogen combustion over these catalysts. However, above 125 °C,

O2 selectivity was observed to decrease for all of the three catalysts, indicating the onset

o of hydrogen combustion. At 175 C, O2 selectivity decreased with increased cobalt loading, indicating that hydrogen combustion was more favored at higher cobalt loadings.

o For PROX reaction over CoOx/CeO2 catalyst, 175 C was a more relevant reaction temperature as CO conversion reached nearly 100%. Both 10% CoOx/CeO2 and 2%

o CoOx/CeO2 showed similar CO conversion at 175 C, however, O2 selectivity over 2%

CoOx/CeO2 was much better (71%) than 10% CoOx/CeO2 (58%). Therefore, under the dry PROX conditions, 2% CoOx/CeO2 could be a better choice than 10% CoOx/CeO2 at

175 oC.

It should be noted that the current CoOx/CeO2 catalyst showed superior PROX performance compared to our previously developed [151] cobalt catalysts supported on various commercial supports such as Al2O3, SiO2, ZrO2, TiO2 and CeO2 (not shown).

Furthermore, the PROX performance over the current CoOx/CeO2 catalyst was comparable to our previously developed [158] CoOx/CeO2 formulation (not shown).

However, the current catalyst formulation holds advantages over the previous generation catalysts in terms of ease of synthesis and environmentally benign catalyst preparation.

113

Figure 8.2: PROX activity and selectivity of () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10% CoOx/CeO2. (Reaction conditions: 1% CO, 1% O2, 60% H2 in He.

3 -1 -1 WHSV=15,000 cm (g cat.) h ). Inset: Time-on-stream CO conversion and O2

o selectivity to CO2 over 10% CoOx/CeO2 at 175 C under above-mentioned reaction conditions 114

8.3.2 Oxidation of CO and H2 over CoOx/CeO2

The kinetics of CO and H2 oxidation over the CoOx/CeO2 catalysts were studied to investigate the intrinsic differences between these catalysts in the presence of excess oxygen. The kinetic data for CO and hydrogen oxidation was collected separately in the presence of 2% CO (or H2) and 2% O2. The results obtained during separate CO and H2 oxidation complement the activity and selectivity data shown in Figure 8.2.

Figure 8.3 shows the CO conversion over CoOx/CeO2 catalysts with different cobalt loadings as a function of temperature during CO oxidation. As expected, CO conversion was found to increase with temperature and cobalt loading. The inset shows the Arrhenius plot for CO oxidation, which was used to obtain activation energies shown in Table 8.2. The activation energy of CO oxidation was found to increase with decreasing cobalt content. Activation energy numbers indicated that CO oxidation became more difficult in lower cobalt content samples. Therefore, the difference in CO conversion among different CoOx/CeO2 catalysts was more prominent at lower PROX temperatures. However, with increase in reaction temperature, sufficient energy appears to have provided in order to overcome the activation energy barrier, consequently difference in activity was diminished at 175 oC.

Figure 8.4 shows the H2 oxidation over CoOx/CeO2 catalysts. H2 oxidation increased with temperature and cobalt content. Arrhenius plot for H2 oxidation is shown in inset, which was used to obtain activation energies shown in Table 2. The activation energy of H2 oxidation was found to increase with decreasing cobalt content indicating

H2 oxidation was not favored in samples with low cobalt loading, and hence better O2

115 selectivity was achieved over 2% and 1% CoOx/CeO2 even at higher reaction temperatures. Furthermore, higher activation energies for H2 oxidation compared to CO oxidation showed that H2 oxidation was more temperature sensitive and became more significant at higher PROX temperatures.

116

CO Oxidation H2 Oxidation

(KJ/mol) (KJ/mol)

1%CoOx/CeO2 58.4 93.2

2%CoOx/CeO2 52.7 80.2

10%CoOx/CeO2 40.8 65.5

Table 8.2: Activation energies of CO and H2 oxidation over CoOx/CeO2 catalysts.

117

Figure 8.3: CO conversion over () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10%

CoOx/CeO2 during CO oxidation. (Reaction conditions: 2% CO and 2% O2 in helium).

Inset shows Arrhenius plot for determination of the activation energy for CO oxidation

118

Figure 8.4: H2 conversion over () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10%

CoOx/CeO2 during H2 oxidation. (Reaction conditions: 2% H2 and 2% O2 in argon) Inset shows Arrhenius plot for determination of the activation energy for H2 oxidation

119

8.3.3 Effect of CO2 and water vapor over CoOx/CeO2

Table 8.3 compares the CO conversion among CoOx/CeO2 catalysts in the

3 - presence of water vapor as well as CO2. The WHSV was maintained at 15,000 cm (g cat

1 -1 ) h for all the runs. The introduction of 0.5% H2O in the PROX feed showed the negative effect on CO conversion over CoOx/CeO2 catalysts.

Several studies [107, 135, 153, 162] have shown that water vapor can significantly affect the CO conversion due to various reasons, such as blockage of active sites, reverse water-gas-shift (RWGS) and formation of a less active CO-H2O complex.

In our case, this effect was more prominent over catalysts with lower cobalt content and at low reaction temperatures. 10% CoOx/CeO2 was the least affected catalyst in the presence of water vapor at any given temperature between 125-175 oC, followed by 2%

CoOx/CeO2. Performing a water-gas-shift (WGS) reaction over these catalysts in the presence of 1% CO and 1% H2O supported this observation. As shown in Figure 8.5, CO conversion increased with increasing cobalt loading and reaction temperature. 10%

CoOx/CeO2 had a much better WGS activity compared to other catalysts due to which

95% CO conversion was observed during PROX even in the presence of water.

The combination of 0.5% H2O and 5% CO2 in the feed had more significant impact on CO conversion than water alone. Again, 1% CoOx/CeO2 was the most affected catalyst followed by 2% CoOx/CeO2. CO conversion of 2% CoOx/CeO2 dropped to 76%

o in the presence of CO2 and H2O at 175 C whereas, 10% CoOx/CeO2 was still able to achieve around 90% CO conversion. It should be noted that reverse-water-gas-shift

(RWGS) was carried out in the presence of 60% H2 and 5% CO2 over CoOx/CeO2

120 catalysts (not shown). However, CO2 conversion was below 3-4% in the temperature

o range of 100-200 C indicating that WGS was still a dominant reaction over CoOx/CeO2 even in the presence of CO2, especially over 10% CoOx/CeO2.

O2 selectivites are also presented in the Table 8.3 and were found to decrease in

o the presence of H2O and CO2 at 125 C indicating that hydrogen oxidation was not as affected as CO oxidation at lower reaction temperatures. However, as the reaction temperature was increased, O2 selectivites were seen to converge and in the case of 10%

o CoOx/CeO2 and 2% CoOx/CeO2 at 175 C, they showed an increase in the presence of

H2O and CO2. This observation could be due to an increase in CO oxidation and higher

WGS reaction rates at higher temperatures, as discussed earlier.

121

Temperature 1%CoOx/CeO2 2%CoOx/CeO2 10%CoOx/CeO2

No H2O-CO2 32 (100) 43 (93) 48 (99)

o 125 C 0.5% H2O 19 (82) 31 (91) 34 (93)

0.5%H2O-5%CO2 0 (N/A) 9 (90) 17 (73)

No H2O-CO2 61 (100) 77 (92) 79 (83)

150oC 0.5% H2O 50 (94) 70 (92) 78 (83)

0.5%H2O-5%CO2 16 (81) 36 (84) 53 (85)

No H2O-CO2 89 (86) 95 (70) 93 (58)

o 175 C 0.5% H2O 80 (87) 93 (70) 96 (59)

0.5%H2O-5%CO2 53 (86) 77 (83) 88 (65)

Table 8.3: CO conversion (%) over CoOx/CeO2 in the presence of H2O and CO2. Values in parentheses show O2 selectivity to CO2 (%) over CoOx/CeO2 in the presence of H2O and CO2

122

Figure 8.5: CO conversion over () 1% CoOx/CeO2, () 2% CoOx/CeO2, and () 10%

CoOx/CeO2 during water-gas-shift. (Reaction conditions: 1% CO and 1% H2O in helium)

123

8.3.5 Catalyst Characterization

8.3.5.1 XRD

Figure 8.6 shows the X-ray diffraction patterns over CoOx/CeO2 catalysts and the bare ceria support collected under ambient conditions. The diffraction lines observed at

o o o o o o 2Θ values of 28.44 , 32.96 , 47.56 , 56.2 , 58.6 and 68.9 were assigned to cerianite structure and associated with [1 1 1], [2 0 0], [2 2 0], [3 1 1], [2 2 2] and [4 0 0] planes, respectively. In addition to the cerianite phase, 10% CoOx/CeO2 showed distinct diffraction peaks corresponding to cobalt oxide phase (Co3O4) located at 2Θ values of

36.7o and 65.15o associated with [3 1 1] and [4 4 0] planes, respectively. Unlike 10%

CoOx/CeO2, weakly resolved peaks for Co3O4 were observed over 2% and 1%

CoOx/CeO2 along with the cerianite phase. This observation along with dispersion measurements indicated that increased cobalt loading did not change the particle size, but the extent of the support surface populated with cobalt particles or the sizes of the clusters made up of these particles. Hence the surface more populated with Co3O4 particles and larger clusters demonstrated diffraction lines with measureable intensities.

The evolution of the crystal phases under a reducing environment over 10%

CoOx/CeO2 is shown in Figure 8.7. XRD patterns were collected in situ in the presence of

o 5% H2/N2 over calcined 10% CoOx/CeO2 in the temperature range of 30-700 C. The

o o o o o peaks located at 2Θ values of 28.44 , 32.96 , 47.56 , 56.2 and 58.6 are associated with

124

[1 1 1], [2 0 0], [2 2 0], [3 1 1] and [2 2 2] planes, respectively and are characteristic of the cerianite phase. The intensity of cerianite peaks increased with temperature, indicating an increase in average particle size under the reducing environment, especially

o o above 500 C. The peak at 2Θ value of 36.7 associated with Co3O4 [3 1 1] plane started

o o to disappear at 250 C and a new crystal phase, CoO was observed at 2Θ value of 42.1 .

With further temperature increase, CoO started to disappear at 300 oC and metallic cobalt

o (Co) was observed at 2Θ value of 44.07 .

125

Figure 8.6: X-ray diffraction patterns of (a) CeO2, (b) 1% CoOx/CeO2, (c) 2%

CoOx/CeO2, and (d) 10% CoOx/CeO2

126

Figure 8.7: In situ X-ray diffraction patterns collected during reduction of 10%

CoOx/CeO2 with 5% H2/He

127

8.3.5.2 TPR

The reduction characteristics of CoOx/CeO2 catalysts and CeO2 support under 5%

H2/He is shown in Figure 8.8. A bare CeO2 support did not exhibit H2 consumption showing that CeO2 could not be reduced under the conditions that the TPR was carried

o o out. 10% CoOx/CeO2 showed three distinct reduction features at 290 C, 325 C and

o o around 500 C. The reduction peak at 290 C was assigned to the reduction of Co3O4 to

CoO whereas the peak at 325 oC was assigned to reduction of CoO to Co [163]. These observations were in good agreement with in situ XRD over 10% CoOx/CeO2.

Unlike 10% CoOx/CeO2, lower cobalt content catalysts, 2% CoOx/CeO2 and 1%

o CoOx/CeO2 showed a broad feature in the temperature of 230-390 C, which could be correlated to reduction of Co3O4 to CoO and CoO to Co. The weakly resolved reduction features observed over these catalysts could be due to low cobalt concentration over

CeO2 support.

o The broad peak observed over 10% CoOx/CeO2 around 500 C was assigned to reduction of surface ceria. The transition metals are known to improve the ceria reducibility. In fact, the absence of this reduction feature in 2% CoOx/CeO2 and 1%

CoOx/CeO2 supported the assertion that cobalt was able to improve the reducibility of ceria.

128

Figure 8.8: Temperature-programmed reduction profiles of (a) CeO2, (b) 1% CoOx/CeO2,

(c) 2% CoOx/CeO2 and (d) 10% CoOx/CeO2 in 5% H2/He

129

8.3.5.3 XPS

X-ray photoelectron spectroscopy was used to examine the chemical nature of cerium and cobalt species in CoOx/CeO2 catalysts. Figure 8.9a-c shows Ce 3d envelope of 1% CoOx/CeO2, 2% CoOx/CeO2 and 10% CoOx/CeO2, respectively. All three spectra showed the contribution from both Ce+3 and Ce+4. The peaks labeled as v (884.1 eV), vʹʹ

(889.2 eV) and vʹʹʹ (898 eV) and their corresponding u modes, u (902.7 eV), uʹʹ (907.6

+4 eV) and uʹʹʹ(916.4 eV) were the characteristic features of Ce whereas, v0 (882 eV), vʹ

(887.2 eV) and their corresponding u modes, u0 (900.5 eV) and uʹ (906.6 eV) were the characteristic features of Ce+3 species [94-96]. Ce 3d envelope showed no significant change upon changing the cobalt loading of the CoOx/CeO2 catalysts.

The core-level Co 2p3/2 spectra of CoOx/CeO2 catalysts are shown in Figure 8.9.

The curve fitting in Co 2p3/2 region for 10% CoOx/CeO2 (Figure 8.9c) showed two components located at 779.1 eV and 780.4 eV. The lower binding energy (B.E.) component was assigned to the main photopeak of cobalt oxide (Co3O4) whereas, the higher B.E. shoulder at 780.4 eV was associated with Auger LMM transition [164]. For

2% CoOx/CeO2 (Figure 8.9b), both these peaks were shifted approximately by 1 eV towards higher B.E. values. It is a known fact that binding energy of core-level electrons of Co+3 is smaller than that of Co+2. Bonelle et al. [165] explained this phenomenon by

+3 +2 postulating that Co ions in Co3O4 have smaller effective charge than Co ions in CoO.

Therefore, this final state effect results in a higher binding energy value (usually 0.9 eV apart) of Co+2 ions than Co+3 ions [166]. In addition, the final state effects in Co 2p region accompany a shift in B.E. of O 1s core-level electrons. The binding energy value

130 of O 1s core electrons in Co3O4 is at higher B.E. than CoO [164]. The peaks fitting (not shown) in O 1s region for 10% CoOx/CeO2 showed the higher B.E. values for O 1s core electron than 2% CoOx/CeO2 supporting this assertion. In addition, the satellite feature was observed at 785.6 eV in 2% CoOx/CeO2 along with main photopeak and Auger

LMM transition. The presence of the satellite feature in 2% CoOx/CeO2 could be due the presence of the paramagnetic Co+2 ions [164]. The reduction of cobalt oxide species under X-ray beam in the high-vacuum environment may occur. Therefore it is plausible

+3 that some of Co ions in Co3O4 were reduced to lower oxidation state cobalt oxide.

Hence 2% CoOx/CeO2 may have some contribution from CoO along with Co3O4. In contrast, 10% CoOx/CeO2 showed no such evidence of the presence of paramagnetic

Co+2, indicating a strong presence of Co+3.

131

Figure 8.9: X-ray photoelectron spectra in the Ce 3d and Co 2p regions of (a) 1%

CoOx/CeO2 (b) 2% CoOx/CeO2 and (c) 10% CoOx/CeO2

132

8.3.5.4 In situ DRIFTS during PROX

The investigation of surface species formed during PROX over CoOx/CeO2 was studied using in situ DRIFTS. Figures 8.10 a-b show the in situ DRIFTS spectra collected over 2% CoOx/CeO2, at various temperatures, in the presence of 1% CO, 1% O2, 60% H2 and helium. The bands associated with gaseous and weakly adsorbed CO were observed at 2152 and 2095 cm-1, respectively, and were seen to decrease with increasing

-1 temperature. The doublet of CO2 at 2358 and 2330 cm was found to increase indicating significant CO oxidation at higher PROX temperatures. The strong positive peaks at

1022, 1290 and 1581 cm-1 indicate the formation of bidentate carbonate species [167-

170] at 30 oC. This observation is consistent with the study reported by Luo. et al. [167] over Co3O4-CeO2. In their findings, the formation of bidentate carbonates was observed at 1590-1582, 1291-1272 and 1026 cm-1 during CO oxidation at room temperature, it was also reported that the intensities of these peaks decreased with increasing temperature. A weak positive feature at 1520 cm-1 accompanied by a broad shoulder at 1260-1220 cm-1 was also observed at 30 oC, which may be attributed to monodentate carbonate species.

Takeguchi et al. [171] also reported monodentate carbonate formation on CeO2 supports

-1 at 1520 and 1270 cm after adsorption of both CO and CO2. Although the presence of monodentate species is to be expected, hydrogen carbonate formation at room temperature cannot be ruled out. Hydrogen carbonate bands have also been reported to occur within the 1270-1220 cm-1 region. However these bands are generally accompanied by the additional bands at 1600, 1390-1413 and 1025-1045 cm-1 [168, 170]. It is possible that these bands are masked due to other carbonates and gas phase species. As the

133 reaction temperature was increased under the PROX feed, the intensities of carbonate bands decreased, possibly due to transformation of carbonate species. The presence of negative bands at 1520, 1220 and 1028 cm-1 could suggest that some of the carbonate species were present on the surface when background spectra were collected. The diminished carbonate bands were accompanied by the formation of strong bands at 1463,

1060 and 854 cm-1, which could be assigned to polydentate carbonates. This observation was in agreement with results obtained in a study by Binet et al. [172]. It was reported that adsorption of CO onto ceria at room temperature predominantly results in the formation of bidentate carbonate species, which upon heating convert to polydentate carbonates [172]. In the same study the thermal stability of various carbonate species was discussed and it was found that, the thermal stability increased from bidentate carbonates

< monodentate carbonates

-1 band at 1380 cm , which is attributed to the symmetric stretch of CO2 [168, 169]. This was confirmed in a separate experiment (inset in Figure 8.11a) where the sample chamber was flushed with He after the catalyst (10% CoOx/CeO2) was kept in a PROX reaction medium at 125°C for 30 minutes. After the gas phase species were flushed, a sharp band

-1 - that was observed at 1380 cm along with the CO2 doublet at 2358 and 2330 cm

1diminished with time. Therefore it is plausible that 1380 cm-1 feature was due to

-1 symmetric CO2 stretch. A well-defined positive band at 1211 cm was observed at higher

134 temperatures, which could be assigned to the formation of intermediate carbonate species

-1 due to adsorption of gas phase CO2 [167]. The peak at 1403 cm usually accompanies this feature, however, the presence of the strong band evidenced at 1380 cm-1 due to gaseous CO2, probably masks this feature. Apart from the carbonates, the formate species were observed with increasing temperature at 2929, 2840, 1560-1540 and 1361 cm-1

[152, 170, 173] which could be due to the interaction OH species at 3712 cm-1 (OH-I),

3665 cm-1 (OH-II) and 3610 cm-1 (OH-III) [170, 174] with CO. In addition to the above mentioned formate bands, some negative features were observed at low temperatures around 1548 and 1357 cm-1, which could be assigned to formate species on ceria support

[175]. One possibility is that these formate species were present originally on the catalyst when background spectra were collected and were transformed to other species upon introduction of the PROX reaction medium. Nevertheless, there is still some ambiguity in assigning these two negative bands and the possibility of a mathematical artifact due to positive/negative background subtraction cannot be ruled out.

A similar in situ DRIFTS experiment was performed during PROX over 10%

CoOx/CeO2 as shown in Figures 8.11 a-b. No significant difference was observed between DRIFT spectra collected over 10% and 2% CoOx/CeO2. However it was noted

-1 that the CO2 doublet occurring at 2356 and 2329 cm was more prominent over the 10%

CoOx/CeO2 sample, suggesting significantly higher CO oxidation. When compared with

2% CoOx/CeO2 on common Y-axis scale, the intensities of CO2 doublet were higher in

o the temperature range of 100-125 C over 10% CoOx/CeO2. However, the difference

o between CO2 intensities diminished from 175 C onwards, in agreement with data presented in Figure 8.2. It should be noted that background spectra were collected under

135 helium at each temperature before collecting the sample spectra under PROX feed in order to correct for background CO2. Therefore, it is reasonable to correlate CO2 intensities observed during the DRIFTS with CO oxidation activity presented in Figure

8.2.

136

Figure 8.10: In situ DRIFT spectra collected during preferential oxidation of CO over 2%

CoOx/CeO2 (a) high wavenumber region and (b) low wavenumber region.

137

Figure 8.11: In situ DRIFT spectra collected during preferential oxidation of CO over

10% CoOx/CeO2. (a) high wavenumber region and (b) low wavenumber region. Inset: spectra taken at 125°C after the reaction gases are flushed

138

8.3.5.5 In situ X-ray absorption near edge spectra (XANES)

The transformation of cobalt coordination environment during PROX was monitored using X-ray absorption near edge spectra (XANES). Figure 8.12a presents the

XANES spectra collected over the pristine 10% CoOx/CeO2 in the range of 7500-7750 eV together with Co3O4. The XANES spectrum of pristine 10% CoOx/CeO2 was well aligned with Co3O4. Figure 8.12b presents the XANES spectrum collected during PROX

o at 175 C over 10% CoOx/CeO2. The XANES spectrum remains similar to that of Co3O4 species and reduction of cobalt was not observed.

Several studies in the past have indicated that Co3O4 can be an excellent candidate for CO oxidation in the presence of hydrogen. However, the excess hydrogen content during PROX may reduce Co3O4 to cobalt oxide phases with lower oxidation states, including metallic cobalt, which may promote the side reactions [149, 151, 152, 155, 157,

161]. The present study suggests that Co3O4 phase remains intact in 10% CoOx/CeO2 and no lower valency cobalt was observed during PROX. Having said all this, the possibility of a partial reduction of cobalt species if the catalyst was kept on-line for much longer period or under higher reaction temperatures can not be ruled out.

139

Figure 8.12: Normalized Co K-edge XANES spectra of 10% CoOx/CeO2. (a) Pristine and

o (b) In situ under PROX at 175 C (Spectra for Co3O4 is included for comparison)

140

8.4 Conclusions

CoOx/CeO2 catalyst was investigated for PROX with a special focus on the effect of cobalt loadings. While higher cobalt loadings favored CO oxidation at low PROX temperatures (T ≤ 150 oC) under the dry conditions, oxygen selectivity was seen to suffer

o at higher cobalt loadings. At 175 C, both 10%CoOx/CeO2 and 2%CoOx/CeO2 showed nearly 100% CO conversion however, O2 selectivity over 2% CoOx/CeO2 was much better than 10% CoOx/CeO2. Furthermore, higher activation energies for H2 oxidation compared to CO oxidation were observed, regardless of the Co loading, implying a higher temperature sensitivity for the H2 oxidation reaction. The presence of water and

CO2 in the feed resulted in a drop in PROX performance, possibly due to the inhibition effect over CoOx/CeO2 catalyst, regardless of the cobalt loadings. However, higher cobalt loading was seen to demonstrate better water /CO2 tolerance and assisted water gas-shift reaction.

The cobalt phase was identified, as Co3O4 and was stable in the reducing environments present in PROX reactions. A stable time-on-stream performance over

CoOx/CeO2 along with catalyst characterization using several analytical techniques indicated no significant contributions from lower valency cobalt species, including metallic cobalt under the reducing environment. From the in situ DRIFTS studies conducted on CoOx/CeO2 upon the introduction of PROX reaction gases at room temperature several species were formed on the catalyst surface such as linearly adsorbed

CO, monodentate carbonate and bidentate carbonate species. As the PROX reaction

141 temperature increased, these carbonate species were converted into more stable polydentate carbonates and formate type species. In addition, the intermediate carbonate species were also observed between 100-150 oC, probably due to the interaction of adsorbed gas phase CO with the hydroxylated CoOx/CeO2 surface.

* Information in this chapter is adapted/taken from, P. Gawade, B. Bayram, A.M.

Alexander and U.S. Ozkan, Applied Catalysis B, Environmental, In Review.

142

PART 3

CATALYTIC AFTER-TREATMENT OF NATURAL GAS ENGINE EXHAUST-A

DUAL STAGE APPROACH

143

CHAPTER 9

INTRODUCTION TO CATALYTIC NOX REDUCTION

In the context of evermore stringent requirements for fuel efficiency and CO2 emissions, natural gas-fired reciprocating engines represent an important and increasingly popular choice within the distributed energy market place. Lean-burn natural gas reciprocating engines offer a simple, well-understood technology and have several advantages over stoichiometric gasoline engines [176]. Lean burn conditions are associated with higher engine efficiencies and significantly cleaner engine out emissions.

However, despite emissions being greatly reduced, engine exhaust still contains considerable amounts of NOx species, carbon monoxide and un-burned hydrocarbons.

Some NOx emissions can be reduced through the modification of fuels or by altering the combustion parameters¸ however in order to reach acceptable emission levels the need for effective after-treatment is evident.

Among current catalytic NOx reduction control technologies, three way catalysts and ammonia- or urea–based selective catalytic reduction (SCR) are the most prevalent.

Both methods are highly effective for the current combustion technologies, which operate over a wide temperature window (200-400 oC), yet are unsuitable for the next generation of lean-burn engines. For example, three way catalysts, which are mostly used in mobile

144 applications, quickly deactivate under lean conditions, while the use of ammonia in emission abatement poses hazards in itself, mainly related to ammonia slip, direct ammonia oxidation, corrosion of equipment due to the formation of ammonium salts, and the additional issues related to ammonia storage and handling [177].

The selective reduction of NOx species by hydrocarbons has attracted significant attention as a promising alternative to conventional after-treatment technologies [178-

180]. This type of application is particularly suited to natural gas engines, in which unburned hydrocarbons, particularly methane, are already present in the exhaust stream and capable of acting as the reducing agent. Although methane, which is a major component of natural gas, is readily available, its effective use as a reducing agent in hydrocarbon-SCR systems is somewhat limited. This is mainly due to the competition from the combustion of the hydrocarbon in the presence of excess oxygen and the inherent difficulty of methane activation [181, 182].

* Information in this chapter is adapted/taken from P. Gawade, A.M.C Alexander, R.

Clark and U.S. Ozkan, Catalysis Today, Submitted (2012).

145

CHAPTER 10

LITERATURE REVIEW: NOX REDUCTION WITH HYDROCARBONS

Early work by Iwamoto et al. [183] and Held et al. [184] demonstrated the possibility of selective catalytic reduction of NOx using hydrocarbons (HC-SCR) over zeolite-based catalysts. Since then, numerous studies and catalyst formulations have been reported on HC-SCR with variety of reaction conditions. Typically, the catalysts for HC-

SCR can be classified into two categories based on the support materials, namely zeolites and metal oxides.

Both noble (Pt, Pd, Rh, Ag) and non-noble (Cu, Fe, Co, In Ni) metals supported on zeolites are widely reported for this application [185-195] ([196, 197] and reference therein). Among these formulations, Cu-ZSM-5 is the probably the oldest and most studied formulation until now [198-200] ([196, 197] and reference therein). Despite being the most popular catalyst for HC-SCR for academic purpose, the performance of

Cu-ZSM-5 suffers in the presence of water and SO2 ([197] and references therein). Apart from Cu-ZSM-5, Pt and Pd-based zeolite catalysts are the most commonly used catalysts under the lean conditions. However, the major drawback of using Pt-ZSM-5 is the formation of N2O during HC-SCR, which is a strong greenhouse gas ([197] and references therein). Among the catalysts which have been reported to date, palladium- based catalysts supported on zeolites, such as ZSM-5 and mordenite, have shown the

146 highest activity and selectivity for catalytic NOx reduction under lean conditions [201].

Despite their high activity and selectivity in HC-SCR, zeolitic catalysts suffer from poor hydrothermal stability [193, 194, 202]. Consequently dealumination of the material may occur, resulting in the formation of PdO aggregates, which are thought to be active for methane combustion, and loss of metal dispersion [203].

Apart from zeolite-based catalysts, metal oxide supported catalysts such as, alumina-based and zirconia-based formulations are commonly used for HC-SCR. Ag and

Pt supported on alumina is widely reported for this application [204-210] ([197, 211] and references therein). Despite being widely studied, alumina-based catalysts are known to lose SCR performance in the presence of water and SO2 ([197] and reference therein). In recent years, sulfated-zirconia supported catalysts have attracted attention for HC-SCR due their promising performance under lean and wet/SO2 conditions [212-216]. The development of these acidic non-zeolitic catalysts, for the HC-SCR of NOx species, has therefore attracted much interest, mainly due to an increase in hydrothermal stability compared to zeolitic counterparts.

Early studies by Resasco and co-workers compared the activity of various zeolithic and non-zeolithic supported Pd catalysts [217]. It was reported that Pd- loaded sulfated zirconia has similar activity for NOx reduction by methane to that of

Pd-supported zeolites. The advantage of using zirconia over other well-established supports is its enhanced thermal and chemical stability. The increased acidity of sulfated zirconia is significant since the catalytic activity and selectivity of De-NOx are thought to be a consequence of the metallic phase and acidity of the support.

Studies by Ohtsuka and co-workers have shown that the addition of platinum into a

147

Pd/SZ catalyst [218] improves both long-term hydrothermal stability and NOx reduction activity. Later studies by the same group have shown that NOx reduction activity can be further improved through the use of a Fe dopant [219]. In these studies NOx conversion was reported to be 60% after 700 hours on stream in the presence of 9% water vapor. Similar co-operative effects have also been observed on Co promoted Pd supported catalysts [220-222]. Quincoces et al. have reported that a Co-Pd/SZ catalyst undergoes a reversible decrease of both NOx reduction and

CH4 oxidation activity in the presence of water vapor; it was also noted that in the presence of water vapor the optimum operating temperature shifts to higher temperatures.

Previously we have described an integrated oxidation and selective reduction catalytic system [223-227] where NO was first oxidized to NO2, then reduced selectively with hydrocarbons. This method combines separate oxidation and reduction catalyst components in order to perform three distinct catalytic functions, specifically NOx reduction, CO oxidation and hydrocarbon combustion. The dual-catalyst approach for lean-burn exhaust after-treatment takes advantage of the stronger oxidizing potential of

NO2 compared to NO, which in turn helps to utilize the reducing capability of unburned hydrocarbons in the exhaust. Despite the exothermic oxidation of NO to NO2 being thermodynamically limited at high temperatures, the dual-catalyst mechanism helps to drive the NO oxidation reaction in the forward direction, as a result of NO2 being continuously removed via NO2-SCR. Any gaseous NO that is present in the system, a consequence of partially reduced NO2, can be re-oxidized by the oxidation catalyst. It is clear that the oxidation catalyst assumes a multi-functional role in the dual-catalyst

148 scheme. In addition to oxidizing NO or re-oxidizing partially reduced NO2 species, the oxidation catalyst also plays a role in catalyzing the combustion of un-burned hydrocarbons and the oxidation of carbon monoxide, which have not been consumed during the SCR reaction [228].

* Information in this chapter is adapted/taken from 1. P. Gawade, A.M.C Alexander, R.

Clark and U.S. Ozkan, Catalysis Today, Submitted (2012). 2. P. Gawade, A.M.C

Alexander, R. Silver and U.S. Ozkan, In Preparation (2012).

149

CHAPTER 11

THE ROLE OF OXIDATION CATALYST IN DUAL-CATALYST BED

11.1 Overview of the Role of Oxidation Catalyst in Dual-Catalyst Bed

In this contribution, the multifunctional role of a metal oxide-supported cobalt oxidation catalyst, namely CoOx/CeO2, in a dual-catalyst system is addressed. The effect of Co-loading in NO to NO2 oxidation has been investigated through steady state experiments. In order to optimize the dual bed composition, the ratio of reduction to oxidation catalyst was varied and investigated via steady state experiments under simulated engine exhaust conditions. Cyclic experiments were conducted in the presence of water vapor to understand the effect of different cobalt loadings on the overall hydrothermal stability of the dual-bed system. Both kinetic and in-situ DRIFTS studies are reported with the aim of understanding the behavior of the oxidation catalyst with respect to hydrocarbon oxidation, which consequently affects the overall NOx reduction performance.

150

11.2 Experimental

11.2.1 Catalyst Preparation

Palladium catalyst supported on sulfated zirconia with 0.3 wt% loading, which was employed as the NOx reduction catalyst for the dual-catalyst system, was prepared using a one-pot sol-gel method, details of which can be found elsewhere [223, 224]

Cobalt catalysts supported on ceria were prepared through an incipient wetness impregnation technique and were used as the NO oxidation catalyst component in the dual-catalyst system. The nanoparticle ceria support was prepared using a precipitation method described previously [159]. The calcined ceria support was impregnated with aqueous cobalt nitrate hexahydrate so as to yield catalysts with Co loadings of either 2 or

10wt% [229].

11.2.2 Catalyst Characterization

Surface area measurements were obtained using a Micromeritics ASAP

2010/2020 accelerated surface area and porosimetry instrument using a N2-physisorption method. Prior to analysis, the samples (0.1 g) were degassed overnight at 130 oC under a vacuum of 3µmHg to remove any adsorbed moisture. The specific surface areas of the catalysts, 0.3% Pd/SZ, 10% CoOx/CeO2, and 2% CoOx/CeO2, were determined by applying the Brunauer, Emmett and Teller (BET) method to the nitrogen physisorption

151 isotherms which were determined at liquid nitrogen temperatures (77K) and found to be

38 m2/g, 84 m2/g and 106 m2/g respectively.

Diffuse Reflectance Fourier Transform Infrared Spectroscopy (DRIFTS) was performed over individual reduction (0.3% Pd/SZ) and oxidation (2% CoOx/CeO2) catalysts and a dual-catalyst bed (0.3% Pd/SZ: 2% CoOx/CeO2 =8:1) during methane temperature programmed desorption (CH4-TPD). The DRIFTS instrument (Thermo

Nicolet 6700) was equipped with a liquid nitrogen-cooled mercury-cadmium-telluride

(MCT) detector and a DRIFTS chamber with ZnSe window. The DRIFTS spectra were collected in the mid-IR range with a resolution of 4 cm-1 averaged over 500 scans.

o Samples were pre-treated in situ at 500 C in 10% O2/He for 30 min at a flow rate of

30ccm, followed by helium flush at the same temperature for an additional 30 min. The background spectra were collected while cooling, under a He flow, at regular temperature

o intervals. Adsorption of 10% CH4/He was performed for 30 min at 50 C followed by

He flush for 30 min at the same temperature, to remove physisorbed CH4. Finally, the spectra were acquired, in He, at 50 oC increments up to a maximum temperature of 500 oC.

11.2.3 Catalyst Activity Testing

Steady-state reaction experiments were conducted at ambient pressure in a ¼”

O.D. stainless steel fixed bed reactor system. The schematic of NOx system is shown is

Figure 11.1. A physical mixture of reduction and oxidation catalysts were packed in desired quantities inside the reactor and held in place using quartz wool plugs. The reactor bed temperature was monitored and controlled using an Omega K-type

152 thermocouple, which was upstream of the catalyst bed and an Omega (model CS232) PID controller. The reactor was kept in a resistively heated homemade furnace. Brooks mass flow controllers (5850E) were used to regulate the gas flows to the reactor system. Water vapor was introduced to the reactor by saturating a stream of helium gas through a heated water bubbler containing de-ionized water. A chemiluminescence NOx analyzer

(Thermo-Scientific 42i-HL) was coupled with a micro-GC (Agilent 3000A) equipped with 0.32mm mol-sieve and PLOT Q columns with thermal conductivity detector to analyze the product gases. NO2 yield and conversions of hydrocarbons (XCHx) and NOx

(XNOx) were calculated using the following relationships:

NO2 yield = (moles of NO2 produced) /(moles of NO fed)

(XCHx) or (XNOx)= ([moles in]-[moles out])/(moles in)

11.2.3.1 NO Oxidation Over CoOx/CeO2

o The catalyst samples were pre-treated in 10%O2/He (40ccm) for 30 min at 450 C prior to catalytic activity testing. Steady-state NO oxidation experiments were performed on CoOx/CeO2 with two different cobalt loadings, namely 2 and 10% Co. These experiments were conducted in the temperature range of 200-400oC and at a gas hourly space velocity (GHSV) of 100,000 h-1. The feed composition was made up of 1000 ppm

NO, 10% O2 and balance He.

153

11.2.3.2 Effect of Reduction-to-Oxidation Catalyst Ratio

All the NOx reduction studies using hydrocarbon mixtures were carried out at a

GHSV 32,000 h-1 and under the following simulated engine exhaust composition 180 ppm NO2, 1737 ppm CH4, 208 ppm C2H6, 104 ppm C3H8, 650 ppm CO, 6.5% CO2, 10%

O2, (0-10%) water vapor and balance He, unless otherwise stated. The effect of the ratio of reduction to oxidation catalyst in dual-catalyst bed was evaluated in the temperature

o range of 300-500 C for NOx reduction performance. The ratio of reduction (0.3% Pd/SZ) to oxidation (10% CoOx/CeO2) catalyst was varied from 2:1, 4:1 and 8:1 (by wt) in the dual-catalyst bed. In order to keep the bed volume constant during reaction studies, quartz powder was mixed with the catalyst bed. The findings of these experiments

(section 3.2) revealed that lower amounts of oxidation catalyst in the dual-bed assisted the overall NOx conversion. Therefore, the amount of oxidation catalyst in dual-catalyst bed was further decreased using a lower cobalt-loading sample (2% CoOx/CeO2) and a similar experiment was performed under simulated engine exhaust conditions to evaluate its NOx reduction performance.

11.2.3.3 Hydrothermal Stability: Effect of Co Loading

The hydrothermal stability of the dual-catalyst bed was evaluated in the presence of 7% water vapor at 450 oC under simulated engine exhaust conditions as outlined above. Oxidation catalysts with different cobalt loadings (10% CoOx/CeO2 or 2%

CoOx/CeO2) were mixed with 0.3% Pd/SZ giving a reduction to oxidation catalyst ratio

154 of 8:1 (by wt). The cyclic experiments were conducted in order to understand the reversibility of the water effect on the dual catalyst bed.

11.2.3.4 Kinetics of Hydrocarbon Oxidation

The oxidation kinetics of the hydrocarbon mixture (methane, ethane and propane) over 0.3% Pd/SZ, 2% CoOx/CeO2 and the dual catalyst bed (Pd/SZ: 2% CoOx/CeO2=

o 8:1) was studied in the typical temperature range of 300-500 C with10% O2. Methane, ethane and propane concentrations were varied in the ranges of 889-2583 ppm, 107-311 ppm and 53-155 ppm, respectively. The reactor was treated either as a differential or an integral reactor depending upon the fractional conversion of hydrocarbons.

155

+,-."/# 0#+#(# +,-."/# +,-."/# 0#+#(# 0#+#(# +,-."/# 0#+#(#

+,-."/# 0#+#(#

+,-."/# 0#+#(#

34# J"7K.;/#

?7."/# 2(# @>AA-"/#

1#1#1#1# 1#1#1#1# 1#1#1#1# 2(# +,-."/# 1#1#1#1#

(;<=>."/# 5,/#$>.# @7--#:7-B"# 0"/"#I7>I"# !"# $%# ($)# *$)# (!)# (!)# :"6.# 0+(# 07EE#+-;H#(;6./;--"/# Figure 11.1: Schematic of Steady State Reaction System for HC-SCR

156

11.3 Results and Discussion

11.3.1 NO Oxidation over CoOx/CeO2

It is known that NO2 participates in the activation of hydrocarbons [230, 231], therefore the efficiency of the oxidation catalyst to perform oxidation of NO to NO2 is fundamental to the selective reduction of NO to N2. NO may exist in the engine exhaust, but it can also form through the partial reduction of NO2 during NO2-SCR, and it is necessary to convert it back into NO2. Therefore NO-to-NO2 oxidation over CoOx/CeO2 with different Co loadings was evaluated under steady state conditions, as illustrated in

Figure 11.2. As a reference, the thermodynamic equilibrium limitation (dashed line) for the conversion of NO to NO2 is also shown for the chosen feed conditions. An increase in

NO2 yield was observed with an increase in the reaction temperature upon going from

200 oC to 300 oC. However a further increase in temperature results in a decrease in the

NO2 yield, which may be due to thermodynamic limitation imposed by the system at higher temperatures. As can be observed, both samples reach equilibrium at approximately 300 oC, with a maximum conversion of c.a. 83%. At higher temperatures, conversions closely follow the equilibrium curve. These results suggest that the NO2 yield is largely independent of cobalt loading on the ceria support. These results differ from those previously reported for cobalt loadings on other supports, namely TiO2 and

ZrO2, where it was found that increasing the cobalt concentration had a significant effect on the NO2 yield [228]. The use of CeO2 as a support appears to enhance catalytic activity for NO conversion with reduced cobalt loadings. This may be a direct result of the high

157 oxygen mobility in CeO2 at elevated temperatures [232] and [references within]. It should also be noted that the mixture of oxidation catalyst with reduction catalyst in a dual-catalyst bed is capable of pushing the thermodynamic equilibrium in forward direction due to the continuous removal of NO2 during SCR with hydrocarbons.

158

Figure 11.2: NO2 yield during NO oxidation over 2%CoOx/CeO2 () and

10%CoOx/CeO2 (). Reaction conditions: 1000ppm NO, 10% O2; 1 atm; GHSV :

-1 100,000 h , (---) thermodynamic equilibrium conversion of NO to NO2

159

11.3.2 Effect of Reduction- to- Oxidation Catalyst Ratio

In the previous section, it was found that NO oxidation appears to be independent of the oxidation catalyst employed. However, the oxidation catalyst not only plays a role in NO oxidation, but also in hydrocarbon and carbon monoxide oxidation. The extent to which these reactions occur over the oxidation component may eventually affect the overall NOx reduction performance in the dual bed scheme. In the dual-catalyst bed, the oxidation catalyst plays multi-functional roles by removing carbon monoxide and un- burned hydrocarbons, which have not participated in the NOx reduction. Also NO formed due to the partial reduction of NO2 can be re-oxidized using the oxidation catalyst [225].

Therefore the effect of the reduction-to-oxidation catalyst ratio in the dual-catalyst bed was investigated under the simulated lean exhaust conditions. Figures 11.3a and 11.3b show NOx and CH4 conversion as a function of reaction temperature, with various dual- catalyst bed reduction-to-oxidation catalyst ratios. It is evident from these figures that the lowest NOx and CH4 conversions were obtained when the reduction catalyst (Pd/SZ) was tested in the absence of the oxidation catalyst. This result clearly indicates the importance the presence of CoOx/CeO2 has in the envisioned dual catalyst bed system.

NOx conversion (Figure 11.3a) was found to increase with increasing the reduction-to-oxidation catalyst ratio (Pd/SZ: 10% CoOx/CeO2), from 2:1 to 8:1 whereas, the opposite trend was observed for CH4 conversion (Figure 11.3b). These results would suggest that having the oxidation catalyst in smaller quantities in the dual catalyst bed could favor the NOx conversion at the expense of hydrocarbons conversion. It may be

160 expected that there is a lower hydrocarbon combustion rate over the 2% CoOx/CeO2 oxidation catalyst compared to the 10% CoOx/CeO2 sample, which ultimately results in the presence of a higher hydrocarbon concentration in the reaction medium to allow the

NOx reduction to take place, consequently increasing the NOx conversion. C2H6 and

C3H8 conversion trends were similar to that observed for CH4 conversions (not shown).

Despite this however, the combustion rates for C2H6 and C3H8 were much higher than

o CH4, reaching 100% conversion at temperatures above 350 C, indicating that the complete oxidation of C2H6 and C3H8 was much easier than that of CH4. Furthermore, complete carbon monoxide conversion was observed regardless of the dual-catalyst bed ratio employed (not shown). These results indicate that the oxidation of CO to CO2 occurs readily over the oxidation catalyst and the percentage of metal loading on the support does not appear to have a significant effect on this process. A separate study was conducted, (not shown), to evaluate the effect of the presence of CO, CO2 and higher hydrocarbons on NO2 reduction, as well as hydrocarbon oxidation. However, the findings of this study are beyond the scope of the current publication.

It was observed, from Figure 11.3, that by lowering the amount of oxidation catalyst (10% CoOx/CeO2) in the dual-catalyst bed, the NOx conversion was improved.

The oxidation catalyst amount in the dual-catalyst bed was further decreased by using the lower cobalt-loading sample (2% CoOx/CeO2). Figure 11.4 shows NO2 SCR performance over the dual-catalyst bed (Pd/SZ: 2% CoOx/CeO2 = 8:1). In this case, NOx conversion was further increased due to the lower hydrocarbon oxidation and hence having a higher concentration of hydrocarbons available to participate in the NO2 reduction.

161

Figure 11.3: (a) NOx conversion (b) CH4 conversion over dual catalyst bed as a function of Pd/SZ-10%CoOx/CeO2 ratios, 1:0 ( ), 2:1(), 4:1() and 8:1() under simulated lean exhaust conditions. Reaction conditions: [NO2]=180 ppm, [CH4]=1737 ppm,

[C2H6]= 208 ppm, [C3H8]= 104 ppm, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%; 1 atm;

GHSV : 32,000 h-1

162

Figure 11.4: NOx (bold) and CH4 conversion (void) over Pd/SZ: CoOx/CeO2=8:1 at different cobalt loadings; 2%CoOx/CeO2 (,) and 10%CoOx/CeO2 (,) under simulated lean exhaust. Reaction conditions: 180 ppm [NO2]=180 ppm, [CH4]=1737 ppm, [C2H6]= 208 ppm, [C3H8]= 104 ppm, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%; 1 atm; GHSV : 32,000 h-1

163

11.3.3 Hydrothermal Stability: Effect of Co Loading

In a previous study, we demonstrated that the presence of water in the engine exhaust significantly affects NOx reduction performance [223, 224]. It was reported that the reduction catalyst, Pd/SZ, was considerably affected by the presence of water whereas the inhibition effect on the oxidation catalyst (Co/ZrO2) was not as apparent.

Furthermore, temperature programmed desorption (TPD) studies, in the presence of water vapor, revealed that water vapor negatively affects the NO adsorption over Pd/SZ whereas NO2 adsorption over Pd/SZ was unaffected. This suggests that the choice of oxidation catalyst in a dual bed system could potentially reduce the negative effect of water vapor on the dual catalyst bed, to improve the overall NOx reduction performance

[223, 224].

In line with above discussion and the findings illustrated in Figures 11.3 and 11.4, the oxidation catalyst is likely to be associated with improving the hydrothermal stability.

Therefore, steady state experiments were conducted with different cobalt loadings in the dual-catalyst bed at 8:1 (by wt) reduction-to-oxidation catalyst ratio. Simulated engine exhaust composition was used as a feed, as described in the experimental section, and the reaction experiments were performed at 450 oC in the presence of 7% water vapor.

Figures 11.5a and 11.5b show NOx and CH4 conversion during the cyclic experiment in the presence and absence of water vapor. It was observed that the effect of water vapor was reversible in both dual-catalyst bed samples, as the complete recovery of NOx and

CH4 conversions was observed upon the removal of water vapor from the feed.

164

For the dual-catalyst bed (Pd/SZ: 10% CoOx/CeO2), upon going from dry to wet feeds streams the NOx and CH4 conversions dropped from 80% to 65% and 76% to 30-

32%, respectively. The complete recovery of NOx and CH4 conversion was observed upon removal of the water vapor. A similar cyclic trend was observed over the Pd/SZ:

2% CoOx/CeO2 dual-catalyst bed. NOx and CH4 conversions dropped from 94% to 73% and 53% to 22%, respectively. It is apparent that the dual-catalyst bed sample containing

2%CoOx/CeO2 showed better NOx conversion compared to that containing 10%

CoOx/CeO2 whereas hydrocarbon conversion trend was opposite to NOx conversion. The higher NOx conversion and lower hydrocarbon conversion in 2% CoOx/CeO2 containing dual-catalyst bed compared to 10% CoOx/CeO2 containing dual-catalyst bed could be due to lower hydrocarbon oxidation, and hence more hydrocarbons being available for NO2 reduction. A similar trend was observed for higher hydrocarbons (not shown). Both C2H6 and C3H8 conversions were adversely affected, but also reversible, in the presence of water vapor. However, conversions of both C2H6 and C3H8 were more than 90% even the presence of water at 450 oC, while the CO conversion was 100% in the presence of water indicating that CO oxidation was unaffected during the wet cycle.

A noticeable decrease in CH4 conversion in the presence of water indicated that

CH4 combustion was significantly affected by water vapor. The inhibition effect of water on CH4 combustion is reported in several studies in the past over Pd-based catalysts [233-

235]. It is believed that either the formation of Pd(OH)2 phase, which is inactive for CH4 combustion [233], or water inhibition during CH4 oxidation [234, 235] could result in a decrease of CH4 conversion in the presence of water vapor.

165

Figure 11.5: Reversibility of the effect of water vapor on (a) NOx (b) CH4 conversion over Pd/SZ: CoOx/CeO2=8:1 at different cobalt loadings 2%CoOx/CeO2 (,) and

10%CoOx/CeO2 (,) under simulated lean exhaust. Reaction conditions: [NO2]=180 ppm, [CH4]=1737 ppm, [C2H6]= 208 ppm, [C3H8]= 104 ppm, [CO]= 650 ppm, [CO2]=

o -1 6.5%, [O2]=10%, [H2O]= 0% or 7%; 450 C; 1 atm; GHSV : 32,000 h

166

11.3.4 Kinetics of Hydrocarbon Oxidation

Hydrocarbon oxidation plays an indirect role in determining NO2 SCR performance. It is desirable to remove hydrocarbons from the engine exhaust via combustion. However, only the remaining hydrocarbons, which have not reacted during

NO2 SCR, should be used for the oxidation process. Unselective direct oxidation of hydrocarbons could result in a considerable drop in NO2 SCR due to lack of available hydrocarbons for the NO2 SCR reaction. In the dual-catalyst bed, it was expected that oxidation catalyst would play a major role in determining the hydrocarbon oxidation; thereby indirectly affecting the NO2 SCR. Therefore kinetic studies of hydrocarbon oxidation were conducted to understand the role of oxidation catalyst, individually and in a dual-catalyst bed composition. Figures 11.6, 11.7 and 11.8 show the fractional hydrocarbon conversions (CH4, C2H6 and C3H8) in the temperature range between 300-

o 500 C over 2% CoOx/CeO2, Pd/SZ and dual-catalyst bed, respectively. The reactor was treated either as a differential or integral depending upon the fractional conversion of hydrocarbons. It should be noted that kinetic parameters for C3H8 oxidation could not be derived, due to the fact that almost 100% C3H8 conversion was achieved under the operating conditions regardless of the choice of the catalyst. This would indicate that the oxidation of C3H8 occurs more readily compared to both CH4 and C2H6.

Figure 11.6a shows rate of CH4 oxidation (mol/gcat. min) as a function of CH4

o concentration in the temperature range of 400-500 C over 2% CoOx/CeO2. Fractional conversion of CH4 (XCH4) was less than 0.2, therefore differential reactor kinetics were

XCH # FCH 0 employed and the reaction rate was calculated using "r = 4 4, CH 4 W 167

! where W is the catalyst weight (g) and F is the initial CH flow rate (mol/min). The CH 4, 0 4

following Power-law model was assumed for the kinetic analysis:

r k (C )a (C )b ! (1) " CH 4 = # CH 4 # O2

3 where C and C are concentrations (mol/cm ) of CH4 and O2, respectively. CH 4 O2

! However, the oxygen concentration is approximately 100 times higher than that of the

! methane! concentration and hence the above Power-law model could be reduced to

r k* (C )a where, k* k (C )b (2) " CH 4 = # CH 4 = " O2

Using a linear regression analysis, the kinetic parameters for CH4 oxidation over 2%

! CoOx/CeO2 were found! as activation energy (Ea)= 79 ± 13 KJ/mol and reaction order “a”

w.r.t. CH4 = 1.0 ± 0.3. The inset of Figure 11.6a shows the parity plot comparing

calculated reaction rates (Rate Cal) vs. experimental reaction rates (Rate Exp).

For the kinetic analysis of C2H6 oxidation over 2%CoOx/CeO2, integral reaction

kinetics were employed as the fractional conversion of C2H6 (XC2H6) was greater than 0.2.

Figure 11.6b shows the plot of XC2H6 vs. W/FC2H6,0,, where FC2H6, 0, is initial flow rate of

ethane (mol/min). The reaction rate was calculated by applying the differential analysis

as follows:

dX "r = C2H6 (3) C2 H 6 W d( F ) C2 H 6 ,0

A reduced Power-law model (equation 2) was used for data fitting and linear regression

! analysis was applied to obtain the kinetic parameters for C2H6 oxidation over

2%CoOx/CeO2 as followed, activation energy (Ea)= 45 ± 3 KJ/mol and reaction order “a”

w.r.t. C2H6 = 0.7 ± 0.08. The inset in Figure 11.6b shows the parity plot comparing

168 calculated reaction rates (Rate Cal) vs. experimental reaction rates (Rate Exp). Kinetic analysis confirmed that the oxidation of C2H6 occurs more readily than the oxidation of

CH4 over 2% CoOx/CeO2.

Figures 11.7a and 11.7b illustrate the influence of W/FA0 on CH4 and C2H6 conversions respectively, during oxidation reaction over only the reduction catalyst,

Pd/SZ. The reactor was treated as an integral reactor and reaction rates were obtained using equation 3. A reduced Power-law model (equation 2) was used for data fitting and linear regression analysis was performed to obtain the kinetic parameters for CH4 and

C2H6 oxidation over Pd/SZ. These results have been summarized in Table 11.1.

Activation energies for CH4 and C2H6 oxidation over Pd/SZ were 95 ± 14 KJ/mol and 50

± 23 KJ/mol, respectively which was in agreement with results found by Ribeiro et al.

[235] where the kinetic study of methane oxidation over a Pd supported catalyst was reported. The higher activation energy barrier for hydrocarbon oxidation over Pd/SZ compared to 2% CoOx/CeO2 indicates that the direct oxidation of hydrocarbons over

Pd/SZ is difficult to achieve. As will be discussed in section 11.3.5, Pd/SZ takes part primarily in the activation of CH4 rather than its direct oxidation. Here, CH4 is activated via hydrogen abstraction to form –CH3 or -CH2 species over Pd/SZ, instead of being directly oxidized to form CO2.

Figures 11.8a and 11.8b show a similar kinetic study of CH4 and C2H6 oxidation over a dual-catalyst bed sample (Pd/SZ: 2% CoOx/CeO2). The synergetic effect between reduction and oxidation catalysts significantly decreases the activation energy barrier.

The activation energies for CH4 and C2H6 oxidation over dual catalyst bed were found to be 66 ± 5 KJ/mol and 35 ± 11 KJ/mol, respectively. This observation is supported using

169

DRIFTS data presented in section 11.3.5, where it was observed that the oxidation of

CH4 was more achievable over 2% CoOx/CeO2 after its initial activation by Pd/SZ.

170

CH4 Oxidation C2H6 Oxidation

Ko a E (kJ/mol) Ko a E (kJ/mol) 8 3 2%CoOx/CeO2 2.76×10 1.0 ± 0.3 79 ± 13 6.2×10 0.7 ± 0.08 45 ± 3

Pd/SZ 1.61×107 0.7 ± 0.19 95 ± 14 7.5×102 0.6 ± 0.17 50 ± 23

Pd/SZ+ 6.2×103 0.6 ± 0.06 66 ± 5 1.49 0.4 ± 0.12 35 ± 11 2%CoOx/CeO2

Table 11.1: Summary of kinetic parameters during CH4 and C2H6 oxidation over oxidation, reduction and dual catalyst bed catalyst

171

Figure 11.6: (a) Rate of methane oxidation vs. methane concentration and (b) fractional ethane conversion vs. W/FA,0 over 2%CoOx/CeO2. Inset: comparison of power-law model predictions for hydrocarbon oxidation rate against experimentally determined rates over

2%CoOx/CeO2

172

Figure 11.7: (a) Fractional methane conversion Vs. W/FA0and (b) fractional ethane conversion Vs. W/FA0 over Pd/SZ. Inset: comparison of power-law model predictions for hydrocarbon oxidation rate against experimentally determined rates over Pd/SZ

173

Figure 11.8: (a) Fractional methane conversion Vs. W/FA0 and (b) fractional ethane conversion Vs. W/FA0 over Pd/SZ: 2%CoOx/CeO2= 8:1. Inset: comparison of power-law model predictions for hydrocarbon oxidation rate against experimentally determined rates over Pd/SZ: 2%CoOx/CeO2= 8:1

174

11.3.5 Investigation of Surface Species during CH4-TPD

The investigation of surface species formed as a result of CH4 adsorption was examined using in situ DRIFTS. Figure 11.9 shows in situ DRIFT spectra collected during methane temperature desorption (TPD) over pre-oxidized 2% CoOx/CeO2 , Pd/SZ and a 1:8 dual-catalyst bed consisting of the two afore-mentioned catalysts. Figure 11.9a illustrates the higher wavenumber region of 2000-3800 cm-1 while Figure 11.9b depicts the lower wavenumber region of 900- 2000 cm-1. Band assignments for the respective samples are summarized in Table 11.2.

The DRIFT spectra obtained from both the dual-catalyst bed and reduction catalyst exhibited distinct bands, which were absent in the 2% CoOx/CeO2 spectra. In both instances, a large negative peak is evident at 1392 cm-1, which is indicative of sulfated zirconia and is attributed to the asymmetric S=O stretching mode [236, 237]. An additional broad band at c.a. 1250 cm-1 can be assigned to S-O vibrations. Additional bands can also be attributed to sulfate species on both Pd/SZ and dual-catalyst bed samples as shown by the negative band centered at c.a. 1175 cm-1 which becomes more

2- negative with increasing temperature. This band corresponds to the SO4 ion and symmetric stretch of O=S=O [238, 239] In all catalysts a broad band is apparent at around 1610-1630 cm-1, assigned to the bending vibration of molecularly adsorbed water

δ(HOH). This band becomes more intense over the Pd/SZ and the dual-catalyst bed, signaling an increase in water formation with increasing temperature. This increase coincides with the S-O and O=S=O features becoming more negative, which suggests a strong interaction of water with the surface sulfate species on sulfated zirconia support, as

175 reported previously [225-227, 240-243]. A weak shoulder at 1302 cm-1 was observed on the CoOx/CeO2 oxidation catalyst. This is a characteristic band of adsorbed methane, which is also evidenced in the reduction catalyst (1304 cm-1), but to a lesser extent [244].

This is perhaps not surprising in view of the fact that Pd species are known to promote H- abstraction, resulting in methane being converted to methyl species [245].

From Figure 11.9a, it is evident that all three samples have well resolved bands in

-1 the hydroxyl region. Two negative bands at 3720 and 3615 cm for CoOx/CeO2 and at

3730 and 3633 cm-1 for Pd/SZ, were identified as type I and type II OH groups, respectively [170, 246]. These bands were not as well resolved in the dual-catalyst bed sample, however the negative band at 3622cm-1 can be assigned to type II OH vibrations

[170]. These negative bands may represent interactions between surface hydroxyl species and adsorbed methane. In addition to the negative features, well resolved OH bands, due to molecularly adsorbed water, were also observed at around 3649 cm-1 and 3510 cm-1 in

-1 the case of the CoOx/CeO2, dual-catalyst bed samples and 3700 cm in the Pd/SZ sample.

Chen et. al [244] have previously reported similar bands occuring in the hydroxyl region due to the adsorption of methane on a ZSM-5 catalyst. It was reported that the strong band at 3510 cm-1 may be due to a shift of the band at c.a. 3620 cm-1 as a result of perturbation of surface bridging hydroxyl groups by adsorbed methane [244]. A broad positive band was observed, for all three samples, in the range of 3600-3000 cm-1 owing to weakly H-bonded hydroxyl groups on the support and is characteristic of OH vibrations associated with physically adsorbed H2O [241]. Both Pd/SZ and dual-catalyst bed samples exhibited additional OH bands. The negative peak evident at 2272 cm-1 can

176 be attributed to terminal OH species interacting with surface lewis acid sites, on the sulfated support..

Bidentate formate species were evidenced in both CoOx/CeO2 and dual-catalyst bed samples. The band observed at approximately 2835 cm-1 was attributed to surface

-1 CH species while the band at 2918 cm could have contributions from CH2 stretching vibrations and νas (OCO) vibrations. An additional small peak was also observed at 2704 cm-1 due to the overtone vibration of 2δ(C-H) [152, 170, 173]. These peaks were not observed in the DRIFT spectra of Pd/SZ; however bands arising from a CH3 asymmetric

-1 -1 stretch and CH3 deformation modes were apparent at 2958 cm and 1448 cm , respectively. A small weak feature at 2727 cm-1 was also evident in both the Pd/SZ and dual-catalyst bed samples, (Figures 11.9a- ii and iii), which may be ascribed to the asymmetric C-H stretch of surface formate species [247, 248].

The IR band observed at 2035 cm-1 in spectra over Pd/SZ and the dual-catalyst bed has previously been reported as corresponding to bridged-bonded CO over reduced

Pd (110) [249-251]. This peak decreases with increasing temperature, however the development of additional IR bands at 1846 cm-1 and 1130cm-1, above 250oC, is observed. These peaks may be attributed to multi-coordinated forms of CO on Pd0 [168,

243, 252]. This type of carbonyl formation and adsorption on Pd may possibly be explained by the reaction of gasous CH4 with the pre-oxidized surface of the Pd particles.

Pd/SZ catalyst, and to a weaker extent, the dual catalyst bed show negative peaks around

-1 2343 and 2320 cm , corresponding to CO2 doublet. This suggests some adsorbed CO2 species on the surface, which leads to negative bands due to background subtraction.

These features become more positive over the dual catalyst bed following a similar trend

177 that is observed with temperature over the CoOx/CeO2, signaling formation of CO2.

However no evidence of CO2 formation was observed in the CH4-TPD reaction over the reduction catalyst alone (Pd/SZ). These observations would suggest that CH4 is only partially oxidized to CO in the presence of the reduction catalyst. However it may be

o postulated that at temperatures above 250 C the mobility of oxygen atoms from the CeO2 support, in both the dual catalyst bed and oxidation catalyst itself, is sufficiently high to enable re-oxidation of the catalyst surface allowing the complete oxidation of CO and

CH4 to occur.

The interaction between CO and/ or CO2 with surface methyl species leads to the formation of various adspecies including, but not limited to, carbonate and formate type intermediates chemisorbed on the support. For this reason, the region between 1600-

1300cm-1 is particularly difficult to interpret, which is perhaps more apparent in the spectra of CoOx/CeO2 and dual-catalyst bed samples. The most intense bands in this zone can be ascribed to bidentate carbonates at c.a. 1563 and 1363 cm-1 which are particularly evident in the lower temperature spectra, Figure 11.9b-i). As the temperature is increased above 200oC, the intensity of these bands decrease at the expense of bands at approximetaly 1540 cm-1 and 1336 cm-1. These features are characteristic of bidentate formate species [170] and correspond to asymmetric and symmetric v(oco) modes. The intensity of these bands increase with increasing temperature and is most likely a result of the interaction between OH and CO species. These bands, however, are not observed in the Pd/SZ or dual catalyst bed spectra, (Figure 11.9b ii and iii), and are most likely masked as a result of the strong negative band associated with sulfated zirconia. The positive peak centered at approximately 1130 cm-1 in both Pd/SZ and dual-catalyst bed

178 samples can be attributed to monodentate carbonate species. This is confirmed by the weak band at 1502 cm-1 and 1506 cm-1 for the respective samples, which can also be assigned to monodentate carbonates [168]. The presence of carboxylate species over all samples cannot be ruled out due to the presence of a peak at c.a. 1500-1510 cm-1. This band is usually accompanied by a strong peak around c.a. 1560 cm-1, which is well

-1 evident in CoOx/CeO2. However, the band at 1560 cm is not well resolved for Pd/SZ or the dual catalyst bed due to the presence of molecularly adsorbed water at 1620 cm-1.

-1 The band observed at 1211 cm in the spectra of CoOx/CeO2, along with those at c.a. 1420 cm-1, which are also evidenced in the high temperature dual-catalyst bed spectra, are attributed to hydrogen carbonate species. The presence of a strong band due to adsorbed water makes it difficult to resolve a possible band around 1608 cm-1, which would be expected in hydrogen carbonate species. Bands at 1047 and 1024 cm-1 may also be attributed to hydrogen carbonate species [170], however it may also be possible that these bands correspond to νS-O stretching mode(s), particularly in the dual catalyst bed and Pd/SZ samples [238, 253].

By studying the adsorption of CH4 on both the reduction and oxidation catalysts independently, it is evident that each serves a different role in the mixed-bed system.

From Figure 11.9 it is clear that in the spectra of both the oxidation catalyst and the dual catalyst bed catalyst, bands attributed to surface CH and CH2 species are present. These bands are not, however, observed in the spectra obtained from the Pd/SZ reduction catalyst. What is apparent though is the presence of CH3 stretching and deformation modes. This would suggest that the Pd reduction catalyst promotes the C-H bond dissociation in CH4, ultimately activitating CH4 and allowing for its further oxidation.

179

This can be illustrated in the spectra of the dual catalyst bed sample, Figures 11.9a) i and

11.9b) i. In these spectra there is again no evidence of CH3 bands, but only CH and CH2, which would imply that the methly species that are generated are immediately oxidized to either carbonate or formate species. These, in turn, are further oxidized to CO2, as evidenced by the CO2 doublet in spectra of both the dual catalyst bed and oxidation catalyst.

This study helps to elucidate the roles of both the oxidation and reduction components of a dual-catalyst system. It is also clear from this study that there is a synergystic effect in the dual catalyst bed, in that the Pd/SZ reduction catalyst dissociates a C-H bond in methane and these CHx species subsequently participate in NOx reduction or directly oxidized by the oxidation catalyst. Secondly it is evident that formation of

CO2 is over the CoOx/CeO2 oxidation catalyst and the formation of CO over Pd/SZ may be a result of surface hydroxyl species interacting with methane or other hydrocarbons.

180

Figure 11.9: In situ DRIFT spectra collected during CH4-TPD over i) 2% CoOx/CeO2, ii)

Pd/SZ and iii) 1:8 Dual-catalyst bed 2%CoOx/CeO2-PdSZ a) high wavenumber region b) low wavenumber region

181

Wavelength (cm-1) 2% CoOx/CeO2 Pd/SZ Dual catalyst bed CH stretch 2839 2835 2-δ (C-H) 2704 2704 CH2 stretch 2916 2918 CH species from formate 2727 2727 CH3 asymmetric stretch 2958 deformation mode 1448 Gas phase CH4 1302 1304

OH Bands Type I 3720 3730 - Type II 3615 3633 3622 Physically adsorbed H2O 3600-3000 3600-3000 3600-3000 Molecularly adsorbed water 3649 3700 3649/3701 3510 3510 Bending vibration δ(HOH) 1638 1620 1620 Acidic terminal OH 2272

CO Linear 2035 2021 Bridged 1846 1846 1130 1130 CO2 2339/2316 2343/2330 2343/2320 O=S=O asymmetric 1392 1392 symmetric 1176 1174 v (s-o) 1246 1252

Bidentate carbonates 1563 1363 Monodentate carbonates 1502 1506 1331 1326 Bidentate formates 1543 1540 1540 1336 Carboxylates 1563 1559 1518 1302 1304 1304 Hydrogen carbonate 1211 1417 1427 1049 1047 1047 1022 1024 1026

Table 11.2: Summary of DRIFTS study during CH4-TPD over oxidation, reduction and dual catalyst bed catalyst

182

11.4. Conclusions

The investigated dual-catalyst system has been shown to achieve high NOx conversions during CHx-SCR in lean burn conditions, which can be further improved by optimizing the ratio of reduction-to-oxidation catalyst in the bed composition. The choice of cobalt loading in the oxidation catalyst is shown to play a crucial role in determining the overall performance of the dual-catalyst bed. The lower cobalt loading in the oxidation catalyst in a dual-catalyst bed led to a higher NOx conversion. This improvement in NOx conversion could be associated with lower hydrocarbon oxidation and hence increased availability of hydrocarbons for NOx reduction. A similar trend was also observed in the presence of water vapor. Furthermore, water inhibition effect was reversible on NOx reduction and CHx oxidation regardless of the choice of oxidation catalyst used. Both the kinetic and DRIFTS studies confirmed that the reduction catalyst,

Pd/SZ activated the CH4 molecule via hydrogen abstraction to form CH3 and CH2 species, which subsequently either participate in NOx reduction or directly oxidized by the oxidation catalyst.

* Information in this chapter is adapted/taken from P. Gawade, A.M.C Alexander, R.

Clark and U.S. Ozkan, Catalysis Today, Submitted (2012).

183

CHAPTER 12

EFFECT OF VARIOUS ENGINE EXHAUST PARAMETERS TO IMPROVE THE

HYDROTHERMAL STABILITY DURING HYDROCARBON-SCR

12.1 Overview of Improving Hydrothermal Stability

The primary focus of this chapter is to examine the hydrothermal stability of the dual catalyst bed, and improve its hydrothermal durability by studying the effect of the engine exhaust parameters. Two parameters were investigated, namely hydrocarbon concentration and reaction temperature. Both cyclic as well as time-on-streams experiments were conducted under various engine exhaust compositions and different reaction temperatures to establish whether water tolerance of the dual catalyst bed could be further improved. The higher concentration of hydrocarbon mixture in the simulated engine exhaust was seen to assist the water tolerance of the dual-catalyst bed. The detailed study was conducted to isolate and identify the primary component in hydrocarbon mixture that contributed more towards water tolerance. This study revealed that ethane had more prominent of an effect than methane or propane for improving the hydrothermal stability of the catalyst bed. Moreover, the effect of reaction temperature confirmed shift in operating temperature window in the presence of water vapor, as

184 higher reaction temperatures were seen to significantly improve the hydrothermal stability of the dual-catalyst bed.

12.2 Experimental and Reaction Studies

Palladium supported over sulfated zirconia was prepared using one-pot sol-gel method with 0.3% Pd loading. The details of catalyst preparation can be found elsewhere

[223, 224]. The oxidation catalyst 2% CoOx/CeO2 was prepared through a wet impregnation method using water as the solvent. The ceria support, for oxidation catalyst, was prepared using the precipitation method. The details of both ceria support and the oxidation catalyst can be found elsewhere [159, 229].

Steady state reactions were conducted in a ¼” O.D. stainless steel packed bed reactor. The schematic of the reaction system is illustrated in Figure 11.1. A physical mixture of reduction (Pd/SZ) and oxidation (2% CoOx/CeO2) catalysts were packed inside the reactor tube and held centrally using quartz wool plugs. The ratio of reduction catalyst to oxidation catalyst was maintained at 8:1 (by wt). Inert quartz powder was mixed with dual-bed in order to maintain a constant catalyst bed volume. Interested readers may find the optimization of the dual-catalyst bed elsewhere [254]. The experiments reported in this chapter were conducted at gas hourly space velocity (GHSV) of 32,000 h-1 unless otherwise mentioned. Studies have also been conducted at higher

GHSV (in the 45,000-60,000 h-1 range - not shown).

12.2.1 Effect of Hydrocarbon Concentration

185

Reaction studies reported in this chapter were carried in the presence of simulated engine exhaust conditions as outlined below: 180 ppm NO2; 2050 ppm to 3050 ppm CHx;

650 ppm CO; 6.5% CO2; 10% O2, and (0-10%) water vapor. It should be noted that for all the experiments reported in this submission, only CHx and water vapor concentrations were varied; other components in the feed such as, NO2, CO, CO2 and O2 were held constant. It is worthwhile to note that CHx mixture was composed of CH4 (∼85% vol.),

C2H6 (∼10% vol.) and C3H8 (∼5% vol.) unless otherwise stated. Hence the concentrations of CH4, C2H6, and C3H8 varied, with respect to each other, from (1740-2592 ppm), (208-

305 ppm) and (104-152 ppm), respectively in the feed.

In order to understand the effect of hydrocarbon concentration under wet exhaust, cyclic experiments were conducted in the presence of 7% water vapor over dual-catalyst bed at 450 oC. It should be noted that each cycle (dry or wet) was kept on-line for approximately 45-60 minutes unless and otherwise mentioned. CHx concentration was varied from 2050 ppm to 3050 ppm while keeping the space velocity constant (32,000 h-

1). Time-on-stream experiments were conducted over 40 h in the presence of 10% water vapor at 450 oC.

In order to establish whether hydrocarbons other than methane aided the hydrothermal stability of the catalysts, various experiments were carried under the wet exhaust. This was achieved by initially comparing a methane only hydrocarbon feed against a mixed hydrocarbon feed, comprising of the hydrocarbon composition outlined in the previous section. For instance, a dual-catalyst bed was tested under simulated engine exhaust in 10% water vapor on time-on-stream over 40 h, as described above,

186 however CH4 concentration was increased to 2050ppm to match the total hydrocarbon

-1 concentration of the CHx gas feed, whilst maintaining the space velocity at 32,000 h

A similar set of studies were conducted to that outlined earlier, however separate hydrocarbon feed streams were utilized. For instance, in these experiments either CH4

(2050 ppm), C2H6 (1025 ppm) or C3H8 (683 ppm) reductant was introduced along with other feed gases as previously described. Here, the concentration of the various hydrocarbon reductants was carefully chosen on equal carbon basis in order to rule out any artifact due to extra carbon in the stream. From these studies, ethane was shown to have the most significant effect on water tolerance and hence further studies were conducted in excess ethane. In these experiments, an extra 100 ppm C2H6 was externally injected into the simulated engine exhaust containing 2050 ppm CHx, as described earlier.

12.2.2 Effect of reaction temperature

Both cyclic and time-on-stream experiments were conducted at different reaction temperatures under wet exhaust. Cyclic experiments were performed at

425 oC, 450 oC and 475 oC and were carried in the presence of 7% water vapor and

2500 ppm CHx concentration, whilst keeping the space velocity and other parameters constant. Time-on-stream experiments were performed at 450 oC, 475 o o C and 500 C in the presence of 10% water vapor whilst keeping the CHx (2050 ppm) concentration and space velocity constant.

187

12.3 Results and Discussion

12.3.1 Effect of Hydrocarbon Concentration

Figure 12.1 demonstrates the reversible effect of water vapor on dual-catalyst bed composed of Pd/SZ and 2% CoOx/CeO2 at two different hydrocarbon concentrations.

Both, NOx and CH4 conversion were reversibly affected in the presence of 7% water vapor. This observation was in line with our previous reports [223, 224, 254] on the dual- catalyst bed as well as studies reported by Quincoces, et al. [220] on Pd-Co/SZ catalyst.

From previous work [223, 224], it was demonstrated that the reduction catalyst was significantly affected in the presence of water vapor while the negative effect of water vapor on the oxidation catalyst component was negligible. Here, temperature programmed techniques was used to demonstrate that water significantly inhibited NO and CH4 adsorption on Pd/SZ while NO2 adsorption was essentially unaffected. It was reported that competition between NO and water vapor results in the blocking of active sites for NO adsorption, consequently preventing reduction of NO to N2.

A drop in NOx conversion was observed in the presence of 7% water vapor regardless of CHx concentration. However the negative effect of water vapor decreased considerably by increasing the CHx concentration from 2050 ppm to 3050 ppm in the feed as illustrated in Figure 12.1a. During the third wet cycle NOx conversion was observed to be 65% for CHx=2050 ppm, 78% for CHx=2500 ppm and 90% for CHx=3050 ppm. This preliminary work indicates that by increasing the hydrocarbon concentration

NOx activity improves significantly in the presence of water vapor. This may be associated with the fact that more hydrocarbons are available in the feed stream to

188 overcome water inhibition of the reaction. It should be noted that under the given conditions that both C2H6 and C3H8 conversions were greater than 90%, however both were also affected reversibly, to some extent, in the presence of water vapor, regardless of hydrocarbon concentration. Moreover, CO conversion remained affected under the wet conditions, indicating that CO oxidation was independent of exhaust conditions over the dual-catalyst bed.

During cycling type experiments CH4 conversion was also observed to decrease considerably, although reversibly, under simulated wet exhaust as shown in Figure 12.1b.

Unlike in the NOx conversion, higher hydrocarbon concentrations did not appear to have a significant effect on CH4 conversion in the wet feed For example, CH4 conversion decreased from 57% to 20-23%, regardless of CHx concentration in the feed during the

3rd wet cycle. Several studies [178] (and reference therein) [233-235] have previously reported the negative effect of water vapor on methane combustion over Pd-based catalyst which was either associated with formation of an inactive Pd-phase [233] or water inhibition during methane oxidation [234, 235]. For instance, Cullis et al. [233] suggested that the competition between methane and water for surface active sites lead to the formation of the Pd(OH)2 phase from PdO, which is an inactive phase for methane oxidation. This claim was further supported by Burch et al. [255]. In their findings, the loss of water from Pd(OH)2 was a rate limiting step rather than the activation of C-H bond in methane. Kinetic studies on methane oxidation over Pd-catalyst, reported by

Ribeiro et al. [235], reveal the inhibitory effect of water during methane oxidation as well. In their findings, the reaction order with respect to water was negative one during methane oxidation. In addition, some of our earlier work [223] indicate that water

189 competes with methane adsorption on the catalyst surface rather than preventing the methane oxidation reaction taking place. With this in mind it may be possible that under wet conditions, only a limited amount of methane is adsorbed onto the catalyst surface which can subsequently react further either being consumed during the SCR reaction or methane oxidation. Methane oxidation contributes significantly to methane conversion, rather than its consumption during the SCR reaction this is mainly due to the fact that the

NO2 concentration is much lower in the feed compared to the hydrocarbon concentrations. Moreover, it will be discussed in the subsequent sections that not only methane, but ethane and propane also contribute considerably to NO2 reduction. Hence, the reported methane conversion is majorly a product of methane oxidation reaction rather than its contribution in NO2 reduction.

190

Figure 12.1: Effect of hydrocarbon concentration () CHx=2050 ppm, () CHx=2500 ppm and () CHx=3050 ppm on (a) NOX and (b) CH4 conversion during NO2 reduction in the presence of H2O. Reaction conditions: [NO2]= 180 ppm, [CHx]= varying, [CO]=

o -1 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=7% ; 450 C ; 1 atm ; GHSV ̴ 32,000h 191

Figure 12.2 shows the time-on-stream performance of dual-catalyst bed in 10% water at two different hydrocarbon mixtures concentrations. An initial drop in NOx

(Figure 12.2a) and methane (Figure 12.2b) conversions was observed during the first 10 h of the experiment, stabilizing afterwards. By increasing the CHx concentration from 2050 ppm to 2500 ppm NOx conversion was improved from 43% to 55% at 40 hours on stream whereas, CH4 conversion was stabilized around 16-17% regardless of hydrocarbon concentration in the feed. These results suggest that by increasing the mixed hydrocarbon concentration by 18%, a further 12% increase in the NOx conversion could be achieved.

In addition, it is worthwhile to note that methane conversion remained reasonably stable

+2 around 17% over 40 h. Typically, dispersed Pd sites are active for NOx reduction, while formation of PdO aggregates results in direct methane combustion [186]. Adelman and Sachtler [256] suggested the possibility of PdO formation via hydrolysis of Pd+2 ions.

Ohtsuka et al.[257] have reported the formation of PdO via agglomeration of palladium in the presence of water vapor over Pd/MOR catalyst, as result, methane conversion was observed to increase over the time. The similar observation was reported by Quincoces et al. [220] over Pd/SZ during NO reduction in the presence of 6% water. However, no such phenomenon was observed in the present study, indicating (indirectly) the lack of PdO formation in the presence of water vapor.

192

Figure 12.2: Effect of hydrocarbon concentration on hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion

(b) CH4 conversion. ()[CHx]=2050 ppm, ()[CHx]= 2500 ppm. Reaction conditions: [NO2]= 180 ppm, [CHx]= varying, [CO]= 650 ppm, [CO2]= 6.5%,

o -1 [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ; GHSV 32,000h̴

193

Earlier, it has been shown that having a higher concentration of mixed hydrocarbons in the feed assisted the NO2 reduction in wet exhaust. This led on to further studies investigating the effect of each hydrocarbon, methane, ethane and propane independent of each other in the wet exhaust. Methane is the major component of natural gas engine exhaust, accounting for approximately 80-90% (by vol.). Hence time-on- stream experiments were conducted with methane as the only hydrocarbon in the feed, its performance was then compared with the equivalent CHx concentration. Figures 12.3a and 12.3b show NOx and CH4 conversion, respectively in 10% water vapor. An initial drop was observed in both NOx and CH4 conversion during the first 10 h of the experiment, stabilizing afterwards, this is a common feature of the time on stream experiments. After 40 h on-stream, NOx conversion was 35% under CH4 only (2050 ppm) feed as compared to 43% in a CHx (2050 ppm) feed. This indicates that higher hydrocarbon species such as, ethane and propane can help to achieve higher NOx conversions than methane alone in wet exhaust conditions, thus aiding water tolerance of the catalyst.

In addition to this, a slight drop in methane conversion was observed from 21% to

17%, when only CH4 was present in feed as opposed to CHx. This phenomenon can be explained by the competitive adsorption between methane and other hydrocarbons

(ethane/propane). These other hydrocarbons compete with methane for adsorption on the catalyst surface, resulting in lower methane oxidation in their presence. This was confirmed in a separate study (not shown), in which we observed that methane conversion decreased during methane oxidation in the presence of higher hydrocarbon species.

194

Figure 12.3: The effect of either a methane only hydrocarbon feed or a mixed higher hydrocarbon feed on the hydrothermal stability of a dual catalyst bed Pd/SZ and

CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b) CH4 conversion.

() [CH4]= 2050 ppm, ()[CHx]= 2050 ppm. Reaction conditions: [NO2]=180 ppm,

o [CHx]= varying, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ;

GHSV 32,000h̴ -1 195

Although the CH4 Vs. CHx study demonstrated that higher hydrocarbon species played a significant role during NO2-SCR, possibly more so than methane in wet exhaust, it was unclear whether this was a result of more ‘total carbon’ in the feed or whether ethane and propane had more reducing capability than methane to aid water tolerance. To address these questions, experiments were conducted in 10% water vapor with single hydrocarbons, methane, ethane or propane, in which the hydrocarbon concentration was kept on an equal carbon-basis, 2050 ppm, 1025 ppm, and 683 ppm respectively. As can be observed from Figure 12.4, ethane had the highest NOx conversion of 54% after 40 hours on stream which was significantly higher than methane

(35%) and in particular propane (20%). The much lower NOx conversion observed by propane was mainly due to its direct oxidation, which is evident in Figure 4b in which almost 100 % CHx conversion was achieved, as a result there was an insufficient amount of propane for NO2 reduction to occur. Our previous report [254] suggests that a reduction catalyst assists the methane activation via hydrogen abstraction, resulting in the formation of CH/CH2 species. These species subsequently either participate in the NOx reduction reaction or are directly oxidized with the help of oxidation catalyst. This methane activation via hydrogen abstraction was an essential step for methane participation in either the SCR process or the combustion reaction. Methane is considered as the hardest molecule to activate via H-abstraction followed by ethane and then propane. Therefore, it was plausible that enough ethane molecules were activated by the reduction catalyst to participate in SCR process, resulting in better NOx conversion, while excess activation of propane molecule might have resulted in direct combustion, thereby performing poorly during NO2 reduction.

196

Figure 12.4: The effect of different hydrocarbons, as a function of equal carbon basis, on the hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion (b) CH4 conversion. () [CH4]= 2050 ppm, () [C2H6]= 1025 ppm, ()[C3H8]= 683 ppm. Reaction conditions: [NO2]=180

o ppm, [CHx]= varying, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ; GHSV 32,000h̴ -1 197

Finally, in order to verify the claim that ethane could significantly improve the hydrothermal stability by injecting small amounts (100 ppm) of extra ethane externally into the simulated engine exhaust composition as shown in Figure 12.5. Here, all other parameters such as engine composition, reaction temperature and space velocity were maintained constant and 100 ppm ethane was mixed just before the reactor inlet. An improvement in NOx conversion, from 43% to 53%, after 40 h on-stream was observed with extra ethane addition as shown Figure 6.

198

Figure 12.5: The effect of supplementary ethane addition on the hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust.

(a) NOx conversion (b) CH4 conversion. () [CHx]= 2050 ppm, () [CHx]= 2050 ppm + 100ppm C2H6. Reaction conditions: [NO2]=180 ppm, [CHx]= varying, [CO]=

o -1 650 ppm, [CO2]= 6.5%, [O2]=10%, [H2O]=10% ; 450 C ; 1 atm ; GHSV 32,000h̴ 199

12.3.2 Effect of Reaction Temperature

Previously it has been reported [223, 254] that under dry reaction conditions an

o increase in the reaction temperature, beyond 450 C, results in a drop in the NOx conversion. This observation was mostly associated with significant hydrocarbon combustion above 450 oC and as a consequence there was an insufficient supply of hydrocarbons for NO2 reduction. However, the presence of water vapor was seen to cause a shift in the operating temperature window. This shift towards the higher temperature could be associated with decrease in water adsorption on catalyst surface at higher reaction temperature. As discussed earlier, water was seen to compete with NO and CH4 for surface active sites, resulting in decrease in both NOx and methane conversion [223].

It is plausible that an increase in reaction temperature probably results in less competition from water, offering better chance for NO and CH4 adsorption on the catalyst surface, consequently improving NOx and methane conversion. In order to prove this argument, the preliminary cyclic experiments were conducted in 7% water vapor over dual-catalyst bed at various reaction temperatures as shown in Figure 12.6. It was observed that

o optimum temperature window for NO2 reduction was between 425 – 450 C under dry conditions, as the maximum NOx conversion was achieved around 96%. However, the

o further temperature increase decreased the NOx conversion to 90% at 475 C under dry feed. This observation could be associated with significant increase in hydrocarbon combustion as mentioned above. On the contrary, NOx conversion remained constant

(75-78%) during the 3rd wet-cycle of, indicating the shift the beginning in temperature window shift. Furthermore, methane conversion improved from 11% to 38% with increase in temperature from 425 oC to 475 oC.

200

Figure 12.6: Effect of reaction temperature () 425oC () 450oC, and () 475oC on

(a) NOX and (b) CH4 conversion during NO2 reduction in the presence of H2O.

Reaction conditions: [NO2]=180 ppm, [CHx]= 2500 ppm, [CO]= 650 ppm, [CO2]=

-1 6.5%, [O2]=10%, [H2O]=7% ; 1 atm ; GHSV 32,000h̴

201

Time-on-stream experiments were conducted in 10% water vapor to further evaluate the dual-bed performance in the temperature window of 450 oC to 500 oC as shown Figure 12.7. After 40 h on-stream, it was observed that NOx conversion improved from 45% to 57%, when the reaction temperature was increased from 450 oC to 475-500 oC. Furthermore, a significant improvement in methane conversion was observed, as methane conversion increased from 17% to 52%. This shift in operating temperature window in the wet run is in agreement with study reported by Quincoces et al. [220] over

Pd/SZ and Co-Pd/SZ catalysts. In their findings, a significant drop in NO and methane conversion was observed below 500 oC in 6% water during NO reduction with methane.

However, the increase in reaction temperature in wet exhaust was seen to improve both,

NO and methane conversion. In addition, the current findings are in-line with our previous work [223] in which we have demonstrated that water adsorption decreased with increase in temperature and water was probably unable to compete with reactants and/or intermediate species at high temperatures.

Therefore, time-on-stream at various reaction temperatures in wet exhaust validated two things: first, the temperature window shifted to higher temperature from

o o 450 C to 475-500 C. Second, a significant improvement in NOx as well as methane conversion can be achieved over the long periods if reactions are conducted within a narrow temperature window of 475-500 oC. It should be noted that reactions were not conducted above 500 oC in wet exhaust for two reasons: first, the exhaust coming from natural as engines under consideration would not go above 500 oC. Second, we have found that above 500 oC, hydrocarbon combustion dominated significantly (even in wet

202 stream), hence poor NO2 reduction performance was observed due to lack hydrocarbon availability.

203

Figure 22.7: The effect of temperature on the hydrothermal stability of a dual catalyst bed Pd/SZ and CoOx/CeO2 under simulated lean exhaust. (a) NOx conversion

o o o (b) CH4 conversion. () 450 C, () 475 C, () 500 C. Reaction conditions:

[NO2]=180 ppm, [CHx]= 2050 ppm, [CO]= 650 ppm, [CO2]= 6.5%, [O2]=10%,

-1 [H2O]=10% ; 1 atm ; GHSV 32,000h̴ 204

12.4 Conclusion

This study shows that it is possible to improve the hydrothermal stability of the dual-catalyst bed by modifying engine exhaust parameters under the lean conditions. The increase in hydrocarbon concentration in the feed was seen to improve

NOx conversion considerably. This improvement was associated with the fact that more hydrocarbon molecules were available to overcome water inhibition of the reaction. The reaction studies confirmed that ethane had the highest reactivity for NOx SCR and injecting a small amount (~100 ppm) ethane can improve the water tolerance of the dual- catalyst bed significantly. Finally, a shift in reaction temperature window was observed in wet exhaust, as the higher reaction temperatures were seen to improve both NOx and methane conversions.

* Information in this chapter is adapted/taken from P. Gawade, A.M.C Alexander, R.

Silver and U.S. Ozkan, In Preparation (2012).

205

CHAPTER 13

SUMMARY AND FUTURE WORK

13.1 Water Gas-Shift

13.1.1 Fe-based Catalysts

WGS remains an essential part of integrated gasification combined cycle (IGCC) for hydrogen production, connecting gasification process and fuel cell operations. Our previous work on Cr-free Fe-based catalyst showed that Fe-Al-Cu formulation can be a possible replacement for the current commercial Fe-Cr-Cu catalyst, which suffers from several operational drawbacks as, mentioned earlier. However, in spite of excellent WGS performance of Fe-Al-Cu, catalyst preparation method was tedious, required critical pH, several washing steps. Therefore, in the present work, “single-step” sol-gel preparation method using propylene oxide as a gelation agent was reported to obtain highly active and stable Fe-Al-Cu catalyst.

Reaction studies in the simulated coal syngas over this Fe-Al-Cu showed that the amount of Cu-loading played a crucial role in determining the WGS activity and stability.

The optimized the ratio of Fe/Cu was found to be 5. XRD study along with surface area

206 analysis showed that the optimum amount of copper in iron oxide matrix facilitated the formation of WGS maghemite phase and improved the surface area of Fe-based catalysts.

The excess Cu containing sample showed the CuO segregation, which resulted in particle sintering, consequently poor WGS performance. The capillary XRD was used to study reduced Fe-Al-Cu in order to avoid catalyst oxidation during sample transfer. This study confirmed the formation of WGS active phase, magnetite (resulted from reduction of maghemite) in all samples. The optimized Cu containing sample did not show any metallic Cu formation, unlike excess Cu-containing Fe-Al-Cu sample. XPS study was used to investigate the oxidation state of iron and copper species over pristine as well as reduced samples. Cu 2p envelope of the optimized sample showed a contribution from

Cu+2 environment over the pristine sample, most of which was retained during the reduction process, indicating the stabilization of copper by iron oxide matrix. However, the excess Cu-containing sample demonstrated higher amount of lower oxidation state copper (Cu+1/Cu0) under the reducing atmosphere, which was attributed to un- incorporated copper in iron oxide matrix. Hence deprived of the stabilization, this free

CuO was prone to reduction and sintering at elevated temperatures and harsher reducing atmosphere. Finally, in situ DRIFTS along with temperature programmed re-oxidation and reaction studies under excess steam indicated the possibility of redox mechanism over Fe-Al-Cu.

Although the reaction studies over these Fe-Al-Cu was carried under simulated coal syngas, the effect of sulfur poisoning needs to be investigated. The preliminary study in the presence of 50 ppm H2S showed that Fe-Al-Cu prepared through propylene oxide suffers from sulfur poisoning. In spite of deactivation due to H2S, our formulation still

207 showed better performance than the commercial Fe-Cr catalyst. Therefore, further investigation is required to understand the sulfur poisoning issues over Fe-Al-Cu samples. Some of our initial catalyst evaluation showed that copper; especially segregated copper was interacting with sulfur. Therefore, it is important to develop Cu- containing catalyst without segregation. Ternary mixtures of copper such as CuFe2O4,

CuMn2O4 and CuAl2O4 could be the possible candidates for such application. In these structures, a considerable amount of copper can be included without segregation from other species. Therefore, such Cu-ternary mixtures need to be investigated for WGS system.

13.1.2 Cu-based Catalysts

Copper supported over two different morphologies of ceria, nanoparticles (NP) and nanorods (NR) were investigated for WGS application. Steady state experiments confirmed that Cu/CeO2-NP prepared through wet-impregnation method using ethanol as a solvent demonstrated excellent WGS performance, which was comparable to our Fe-

Al-Cu catalyst. On the contrary, Cu/CeO2-NR showed poor WGS performance, regardless of the choice of impregnation medium. XRD study revealed the segregation of

CuO species from ceria matrix in case of Cu/CeO2-NR. Unlike the nano-rod-supported catalysts, diffraction peaks associated with a separate CuO phase were either non-existent or very weakly resolved over these catalysts suggesting the presence of smaller CuO particles. In situ XRD under the reduction atmosphere confirmed the formation of metallic copper through the reduction of segregated CuO in nanorods-based samples.

208

Temperature programmed reduction along with N2O chemisorption technique demonstrated well-dispersed fine copper particles in close interaction with the ceria support over Cu/CeO2-NP, whereas bulk-like isolated particles of copper were observed over the Cu/CeO2-NR sample. Finally in situ DRIFTS indicated the presence of a redox mechanism, which was facilitated by the higher reduction-oxidation potential of small, well-dispersed copper species in Cu/CeO2-NP.

Although it has been established in this study that Cu/CeO2-NP was better than nanorods based samples, other morphologies of ceria need to be investigated. The following figure (14.1) shows the different sizes and shapes of ceria prepared through hydrothermal-precipitation method. Hence these morphologies should be investigated and optimized for WGS application. Furthermore, only 5%Cu loading has been reported in the current submission. Therefore, copper loading in these Cu/CeO2 nano- morphologies needs to be optimized. Finally, the effect of calcium doping in ceria matrix should be evaluated for WGS application. Our research group (Song et al. [258]) has shown in the past that calcium doping in ceria matrix for Co/CeO2 catalyst improves the oxygen mobility, thereby offering better hydrogen yields for ethanol steam reforming application. Along those lines, there is a strong possibility that calcium doping in

Cu/CeO2 catalyst might assist WGS reaction as the reaction proceeds via redox mechanism over these samples.

209

Figure 14.1: Various nano-morphologies of ceria prepared through hydrothermal- precipitation (Image Courtesy: Bing Tan)

13.2 Preferential Oxidation of Carbon Monoxide

PROX offers an efficient and economical way to purify the hydrogen stream for

PEM fuel cell application. Our group [151] in the past investigated various potential commercial supports for Co such as ZrO2, CeO2, Al2O3, SiO2 and TiO2 for their PROX performance. In that study, ZrO2 and CeO2 emerged as promising candidates for PROX application. In subsequent years [158], in-house nano-sized ceria was prepared using tributyl amine/ethylene glycol, which showed a better PROX performance than its commercial counterpart. Although our Co/CeO2 showed promising PROX performance, catalyst preparation involved the use of flammable/toxic chemicals. The present work is an extension of our previous studies on PROX over Co/CeO2 catalysts. The catalyst preparation method used here offers a simpler and cleaner way to prepare high surface area nanoparticles of CoOx/CeO2. The effect of Co-loading is examined and optimized under hydrogen rich conditions as well as in the presence of CO2 and water vapor.

210

Steady state reaction study under dry PROX conditions showed that 2%

o CoOx/CeO2 could be a better choice than 10% CoOx/CeO2 at 175 C, as it offered similar

CO conversion (above 90%) and better O2 selectivity. The kinetics of CO and H2 oxidation demonstrated increase in activation energy barrier for both CO and H2 oxidation with decreased in cobalt loading. Furthermore, higher activation energies for

H2 oxidation compared to CO oxidation showed that H2 oxidation was more temperature sensitive and became more significant at higher PROX temperatures. The presence of

CO2 and water in PROX feed adversely affected the catalyst performance, which was more prominent at lower cobalt content catalysts and low reaction temperatures. The cobalt phase was identified as Co3O4 and was found to be stable in the reducing environments present in PROX reactions. A stable time-on-stream performance over

CoOx/CeO2 along with catalyst characterization using several analytical techniques indicated no significant contributions from lower valency cobalt species, including metallic cobalt under the reducing environment. Finally, in situ DRIFTS study under

PROX feed revealed the formation of carbonate species at room temperature, which were converted to more stable polydentate carbonates, and formate type species with increased temperature.

Future work should focus on more stringent PROX feed conditions, involving higher water and CO2 content. Therefore, more detailed and long-term stability experiments should be conducted in the presence of water and CO2. The stability of

Co3O4 phase is a crucial component in this system. Although the present study indicates no significant contributions from lower valency cobalt species, the possibility of a partial reduction of cobalt species if the catalyst was kept on-line for much longer period or

211 under higher reaction temperatures cannot be ruled out. Therefore catalyst characterization is required for longer period under reaction conditions. The role of ceria morphology should be evaluated thoroughly for PROX applications. It is a well- established fact that CO oxidation can significantly improve by using different morphologies of ceria. Also some work in PROX area as well over Au/CeO2 catalysts hints the possibility ceria morphology effect.

13.3 Selective Catalytic Reduction

The selective catalytic reduction (SCR) of NOx species by hydrocarbons is an attractive and promising alternative to conventional after-treatment technologies. This type of application is particularly suited to natural gas engines, in which unburned hydrocarbons, particularly methane, are already present in the exhaust stream and capable of acting as the reducing agent. Our previous work in this area has established that dual- catalyst bed is capable of three distinct functions specifically, NOx reduction, CO oxidation and hydrocarbon combustion. The dual-catalyst approach for lean-burn exhaust after-treatment, takes advantage of the stronger oxidizing potential of NO2 compared to NO, which in turn helps to utilize the reducing capability of unburned hydrocarbons in the exhaust. The present work focuses on multifunctional role of metal oxide supported cobalt oxidation catalyst, namely CoOx/CeO2, in a dual-catalyst system.

The experiments are conducted in realistic simulated lean burn natural gas engine exhaust to evaluate the catalyst performance. In addition, the issues such as water tolerance and improving the hydrothermal stability have been addressed as well.

212

The efficiency of the oxidation catalyst to perform oxidation of NO to NO2 is fundamental to the selective reduction of NO to N2. The formation of NO is largely a result of the partial reduction of NO2 during NO2-SCR, and it is required to be converted back into NO2. Therefore NO to NO2 oxidation over CoOx/CeO2 with different Co loadings was evaluated under steady state conditions. The results indicated that the NO2 yield was largely independent of cobalt loading on the ceria support. Although NO2 yield during NO oxidation was independent of choice of the oxidation catalyst, a significant improvement in NOx conversion was observed in the presence of low cobalt containing oxidation catalyst in dual bed during CHx-SCR. This observation could be associated with lower hydrocarbon oxidation and hence more availability of hydrocarbons for NOx reduction. Furthermore, the reaction study in the presence of water vapor showed the reversible inhibition effect on NOx reduction and CHx combustion regardless of the choice of oxidation catalyst used. Both the kinetic and DRIFTS studies confirmed that the reduction catalyst, Pd/SZ, did not directly participate in the oxidation of CH4, but instead activated the CH4 molecule via hydrogen abstraction to form –CH3 or -CH2 species which were then subsequently oxidized by the oxidation catalyst.

After establishing the fact that water inhibition effect on the optimized dual –bed was reversible; we looked into improving the hydrothermal stability of the system. Cyclic as well as time-on-stream steady state reactions were conducted in the presence of 7 to

10% water vapor at different hydrocarbon concentration and reaction temperatures. This study indicated that higher hydrocarbon concentration in the feed considerably improved the NOx conversion, which could be associated with more availability of hydrocarbons to

NOx reduction. Also, the increase in reaction temperature seemed to assist the

213 hydrothermal stability, which could be due to less water adsorption on the catalyst surface at high temperatures.

One of the major challenges in commercializing the powder catalyst is to coat the catalyst on the monolith support. The conventional way is to prepare the catalyst slurry with addition of the binder in an industrial solvent. This binder usually assists to adhere the catalyst on the monolith cores. However, this approach did not work in our case, as the addition of the finished catalyst product with binder in a solvent resulted in activity loss (not shown). Therefore an un-conventional method was developed to incorporate the commercial binder in the catalyst. This binder-added catalyst was then tested in a powder form in the reactor system under various reaction conditions (not shown). A comprehensive reaction studies confirmed that binder addition in the catalyst assisted the

NOx reduction performance in dry as well as wet conditions. In addition, this binder added catalyst demonstrated better sulfur tolerance in the presence of SO2 as well (not shown). It should be noted that binder containing catalyst performance has not been included in this current submission in order to protect the intellectual property (IP).

Interested reader may contact Dr. Umit S. Ozkan (Advisor) for further correspondence.

Our preliminary efforts have established that catalyst can be coated on the monolith using the commercial binder through un-conventional method. Therefore, future work should involve testing of the catalyst coated monolith cores in the simulated engine exhaust. So far only one type of the binder has been investigated and it has shown promising results. Therefore, different types of commercial binders need to be investigated and their loadings should be optimized. Furthermore, this submission only includes kinetic study for hydrocarbon oxidation. This knowledge needs to be extended

214 to understand the kinetics of NO2 reduction with hydrocarbons. Some of the preliminary work on this project has demonstrated that (not shown) the addition of higher hydrocarbons (especially, ethane) in exhaust stream assisted the NOx reduction. Hence the further detailed investigation is required along this line.

215

References

[1] E.I.A. (EIA):, Annual Energy Review 2006, DOE/EIA 0384 (2006). [2] M. Parikka, Biomass Bioenergy 27 (2004) 613-620. [3] G. Fischer, L. Schrattenholzer, Biomass Bioenergy 20 (2001) 151-159. [4] M. Hoogwijk, A. Faaji, R. van den Broek, G. Berndes, D. Gielen, W. Turkenburg, Biomass Bioenergy 25 (2003) 199-133. [5] J.L. Rangel Costa, G.S. Marchetti, M.C. Rangel, Catalysis Today 77 (2002) 205- 213. [6] J. Souza, M. Rangel, Reaction Kinetics and Catalysis Letters 83 (2004) 93-98. [7] M. Sohrabi, A. Irandoukht, Reaction Kinetics and Catalysis Letters 80 (2003) 303-309. [8] C. Rhodes, G.J. Hutchings, Physical Chemistry Chemical Physics 5 (2003) 2719-2723. [9] R.J. Gorte, S. Zhao, Catalysis Today 104 (2005). [10] E.B. Quadro, M.L.R. Dias, A. Maria, M. Amorim, M.C. Rangel, Journal of the Brazilian Chemical Society 10 (1999) 51-59. [11] Q. Liu, W. Ma, R. He, Z. Mu, Catalysis Today 106 (2005) 52-56. [12] J.C. Gonzalez, M.G. Gonzalez, M.A. Laborde, N. Moreno, Applied Catalysis 20 (1986) 3-13. [13] C. Rhodes, B.P. Williams, F. King, G.J. Hutchings, Catalysis Communications 3 (2002) 381-384. [14] V. Idakiev, A.D. Mihajlo, B. Kanev, A. Andreev, Reaction Kinetics and Catalysis Letters 33 (1987) 119-124. [15] A. Andreev, V. Idakiev, D. Mihajlova, D. Shopov, Applied Catalysis 22 (1986) 385-387.

216

[16] J. Tsagaroyannis, K.J. Haralambous, Z. Loizos, Materials Letters 28 (1996) 393. [17] G.C. Araújo, M.C. Rangel, Catalysis Today 62 (2000) 201-207. [18] S. Natesakhawat, X. Wang, L. Zhang, U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 260 (2006) 82-94. [19] L. Zhang, X. Wang, J.-M.M. Millet, P.H. Matter, U.S. Ozkan, Applied Catalysis A: General 351 (2008) 1-8 [20] L. Zhang, J.-M.M. Millet, U.S. Ozkan, Applied Catalysis A: General 357 (2009) 66-72. [21] A.L.M. da Silva, J.M. Assaf, Studies in Surface Sciences and Catalysis 167 (2007) 213-218. [22] P. Djinovic, J. Batista, A. Pintar, Catalysis Today 147S (2009) S191-S197. [23] O. Ilinich, W. Ruettinger, X. Liu, R. Farrauto, Journal of Catalysis 247 (2007) 112-118. [24] J.B. Ko, C.M. Bae, Y.S. Jung, D.H. Kim, Catalysis Letters 105 (2005) 157-161. [25] L. Li, Y. Zhan, Q. Zheng, Catalysis Letters 118 (2007) 91-97. [26] L. Li, Y. Zhan, Q. Zheng, Y. Zheng, C. Chen, Y. She, X. Lin, K. Wei, Catalysis Letters 130 (2009) 532-540. [27] Y. Li, Q. Fu, M. Flytzani-Stephanopoulos, Applied Catalysis B: Environmental 27 (2000) 179-191. [28] A. Pintar, J. Batista, S. Hocevar, Journal of Colloid and Interface Science 307 (2007) 145-157. [29] J.A. Rodriguez, J. Graciani, J. Evans, J.B. Park, F. Yang, D. Stacchiola, S.D. Senanayake, S. Ma, M. Perez, P. Liu, J.F. Sanz, J. Hrbek, Angewandte Chemie International Edition 48 (2009) 8047-8050. [30] Y. She, Q. Zheng, L. Li, Y. Zhan, C. Chen, Y. Zheng, X. Lin, International Journal of Hydrogen Energy 34 (2009) 8929-8936. [31] T. Shishido, M. Yamamoto, D. Li, Y. Tian, H. Morioka, M. Honda, T. Sano, K. Takehira, Applied Catalysis A: General 303 (2006) 62-71. [32] X. Wang, J.A. Rodriguez, J.C. Hanson, D. Gamarra, A. Martinez-Arias, M. Fernandez-Garcia, Journal of Physical Chemistry B 110 (2006) 428-434. 217

[33] H. Yahiro, K. Murawaki, K. Saiki, T. Yamamoto, H. Yamaura, Catalysis Today 126 (2007) 436-440. [34] P. Djinovic, J. Batista, J. Levec, A. Pintar, Applied Catalysis A: General 364 (2009) 156-165. [35] S. Pradhan, A.S. Reddy, R.N. Devi, S. Chilukuri, Catalysis Today 141 (2009) 72- 76. [36] G. Jacobs, U. Graham, E. Chenu, P.M. Patterson, A. Dozier, B.H. Davis, Journal of Catalysis 229 (2005 ) 499–512. [37] G. Jacobs, E. Chenu, P.M. Patterson, L. Williams, D. Sparks, G. Thomas, B.H. Davis, Applied Catalysis A: General 258 (2004) 203-214. [38] A. Luengnaruemitchai, S. Osuwan, E. Gulari, Catalysis Communications 4 (2003) 215–221. [39] H. Zou, X. Dong, W. Lin, Journal of Natural Gas Chemistry 14 (2005) 29-34. [40] M.A. Arias, A.B. Hungaria, M. G, D. Gamarra, Applied Catalysis B: Environmental 65 (2006) 207-216. [41] E. Moretti, M. Lenarda, L. Storaro, A. Talon, R. Frattini, S. Polizzi, E. Castellon, A. Lopez, Applied Catalysis B: Environmental 72 (2007) 149-156. [42] T. Caputo, L. Lisi, R. Pirone, G. Russo, Applied Catalysis A: General 348 (2008) 42-53. [43] A. Luengnaruemitchai, S. Osuwana, E. Gulari, International Journal of Hydrogen Energy 29 (2004) 429 – 435. [44] N.L.T. Nga, C. Potvin, G. Djega-Mariadassou, L. Delannoy, C. Louis, Topics in Catalysis 42-43 (2007) 91-94. [45] L. Kundakovic, M. Flytzani-Stephanopoulos, Journal of Catalysis 179 (1998) 203-221. [46] M. Qi, M. Flytzani-Stephanopoulos, Industrial and Engineering Chemistry Research 43 (2004) 3055-3062. [47] T. Bunluesin, R.J. Gorte, G.W. Graham, Applied Catalysis B: Environmental 15 (1998) 107-114. [48] S. Hilaire, X. Wang, T. Luo, R.J. Gorte, J. Wagner, Applied Catalysis A: General 215 (2001) 271-278. 218

[49] G. Jacobs, E. Chenu, P.M. Patterson, L. Williams, D. Sparks, G. Thomas, B.H. Davis, Applied Catalysis A: General 258 (2004) 203-214. [50] E.S. Bickford, S. Velu, C. Song, Catalysis Today 99 (2005) 347-357. [51] G. Avgouropoulos, T. Ioannides, Applied Catalysis B: Environmental 67 (2006) 1-11. [52] L. Kundakovic, M. Flytzani-Stephanopoulos, Applied Catalysis A 171 (1998). [53] A. Bumajdad, J. Eastoe, A. MAthew, Advances in Colloid and Interface Science 147-148 (2008) 56-66. [54] H. Mai, L. Sun, Y. Zhang, R. Si, W. Feng, H. Zhang, H. Liu, C. Yan, Journal of Physical Chemistry B 109 (2005) 24380-24385. [55] P.X. Huang, F. Wu, B.L. Zhu, X.P. Gao, H.Y. Zhu, T.Y. Yan, W.P. Huang, S.H. Wu, D.Y. Song, Journal of Physical Chemistry B 109 (2005) 19169-19174. [56] C. Pan, D. Zhang, L. Shi, Journal of Solid State Chemistry 181 (2008) 1298– 1306. [57] C. Pan, D. Zhang, L. Shi, J. Fang, European Journal of Inorganic Chemistry (2008) 2429–2436. [58] Z. Yang, K. Zhou, X. Liu, Q. Tian, D. Lu, S. Yang, Nanotechnology 18 (2007) 185606. [59] K. Zhou, X. Wang, X. Sun, Q. Peng, Y. Li, Journal of Catalysis 229 (2005) 206- 212. [60] Tana, M. Zhang, J. Li, H. Li, Y. Li, W. Shen, Catalysis Today 148 (2009) 179- 183. [61] G. Yi, Z. Xu, G. Guo, K.-i. Tanaka, Y. Yuan, Chemical Physics Letters 479 (2009) 128-132. [62] R. Si, M. Flytzani-Stephanopoulos, Angewandte Chemie International Edition 47 (2008) 2884-2887. [63] H.E. Kissinger, Analytical Chemistry 29 (1957) 1702-1706. [64] J.W. Niemantsverdriet, Spectroscopy in Catalysis: An Intorduction 2nd Edition (2000). [65] M. Zayat, D. Levy, Chemistry of Materials 12 (2000) 2763-2769.

219

[66] H. Cui, M. Zayat, D. Levy, Journal of Sol-Gel Science and Technology 35 (2005) 175–181. [67] A.E. Gash, T.M. Tillotson, J. Joe H Satcher, J.F. Poco, L.W. Hrubesh, R.L. Simpson, Chemistry of Materials 13 (2001) 999-1007. [68] W. Dong, C. Zhu, Journal of Materials Chemistry 12 (2002) 1676–1683. [69] M. Anilkumar, R. Pasricha, V. Ravi, Ceramics International 31 (2005) 889- 891. [70] D. Segal, Journal of Materials Chemistry 7 (1997) 1297. [71] S.R. Dhage, S.P. Gaikwad, P. Muthukumar, V. Ravi, Materials Letters 58 (2004) 2704-2706. [72] S.R. Dhage, R. Parsicha, V. Ravi, Materials Research Bulletin 38 (2003) 1623- 1628. [73] A.O. Souza, M.C. Rangel, Reaction Kinetics and Catalysis Letters 79 (2003) 175-180. [74] X. Liang, X. Wang, J. Zhuang, Y. Chen, D. Wang, Y. Li, Advanced Functional Materials 16 (2006) 1805. [75] M.L. Kundu, A.C. Sengupta, G.C. Maiti, Journal of Catalysis 112 (1988) 375. [76] M.A. Edwards, D.M. Whittle, C. Rhodes, A.M. Ward, D. Rohan, M.D. Shannon, G.J. Hutchings, C.J. Kiely, Physical Chemistry Chemical Physics 4 (2002) 3902- 3908. [77] J.F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, Handbook of X-ray photoelectron spectroscopy, Perkin Elmer Corporation Physical Electronics Division, Eden Prairie, 1992. [78] D.C. Frost, A. Ishitani, C.A. McDowell, Molecular Physics 24 ( 1972) 861-877. [79] D.E. Mencer, M.A. Hossain, R. Schennach, T. Grady, H. McWhinney, J.A.G. Gomes, M. Kesmez, J.R. Pargad, T.L. Barre, D.L. Cocke, Vacuum 77 (2004) 27- 35. [80] C.-K. Wu, M. Yin, S. O’Brien, J.T. Koberstein, Chemistry of Materials 18 (2006) 6054-6058. [81] N.S. McIntyre, D.G. Zetaruk, Analytical Chemistry 49 (1975) 1521-1529.

220

[82] G. Jacobs, P.M. Patterson, L. Williams, E. Chenu, D. Sparks, G. Thomas, B.H. Davis, Applied Catalysis A: General 262 (2004) 177-187. [83] G. Jacobs, G.M. Patterson, U.M. Graham, D.E. Sparks, B.H. Davis, Applied Catalysis A: General 269 (2004) 63-73. [84] C.J.G. van der Grift, J.W. Geu, M.J. Kappers, J.H. van der Maas, Catalysis Letters 3 (1989) 159-168. [85] G. Smit, N. Strukan, M. Craje, K. Lazar, Journal of Molecular Catalysis A 252 (2006) 163-170. [86] C. Rhodes, G.J. Hutchings, A.M. Ward, Catalysis Today 23 (1995) 43-58. [87] R.M. Van Natter, J.S. Coleman, C.R.F. Lund, Journal of Molecular Catalysis A: Chemical 292 (2008) 76-82. [88] J.E. Kubsh, J.A. Dumesic, AIChE Journal 28 (1982) 793. [89] C.R. Lund, J.E. Kubsh, J.A. Dumesic, in: R.K. Graselli, J.F. Bradzil (Eds.), Solid state Chemistry in Catalysis, American Chemical Society, Washington DC, 1985, p. 19. [90] G.C. Chinchen, J.E. Kubsh, J.A. Dumesic, Applied Catalysis 12 (1984) 97-103. [91] R.L. Keiski, O. Desponds, Y.F. Chang, G.A. Somorjai, Applied Catalysis A 101 (1993) 317-338. [92] J.R. Jensen, T. Johannessen, H. Livbjerg, Applied Catalysis A: General 266 (2004) 117-122. [93] P. Gawade, B. Mirkelamoglu, B. Tan, U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 321 (2010) 61-70. [94] P. Burroughs, A. Hamnett, A.F. Orchard, G. Thornton, Journal of the Chemical Society-Dalton Transactions 1686 (1976) 1686-1698. [95] A. Pfau, K.D. Schierbaum, Surface Science 321 (1994) 71-80. [96] M. Romeo, K. Bak, J. El Fallah, F. Le Normand, L. Hillaire, Surface and Interface Analysis 20 (1993) 508-512. [97] X. Tang, B. Zhang, Y. Li, Y. Xu, Q. Xin, W. Shen, Applied Catalysis A: General 288 (2005) 116-125. [98] L. Li, Y. Zhan, Q. Zheng, Y. Zheng, X. Lin, D. Li, J. Zhu, Catalysis Letters 118 (2007) 91-97. 221

[99] M.-F. Luo, Y.-J. Zhong, X.-X. Yuan, X.-M. Zheng, Applied Catalysis A: General 162 (1997) 121-131. [100] H. Zou, X. Dong, W. Lin, Applied Surface Science 253 (2006) 2893-2898. [101] P. Bera, A.L. Camara, A. Hornes, A. Martınez-Arias, Journal of Physical Chemistry C 113 (2009) 10689–10695. [102] D. Gamarra, C. Belver, M. Fernandez-Garcia, A. Martinez-Arias, Journal of American Chemical Society 129 (2007) 12064-12065. [103] M. Daturi, C. Binet, J.C. Lavalley, G. Blanchard, Surface and Interface Analysis 30 (2000) 273-277. [104] C.S. Polster, H. Nair, C.D. Baertsch, Journal of Catalysis 266 (2009) 308-319. [105] T. Venkov, K. Hadjiivanov, A. Milushev, D. Klissurski, Langmuir 19 (2003) 3323-3332. [106] M. Krumpelt, T.R. Krause, J.D. Carter, J.P. Kopasz, S. Ahmed, Catalysis Today 77 (2002) 3-16. [107] J.L. Ayastuy, A. Gil-Rodríguez, M.P. González-Marcos, M.A. Gutiérrez-Ortiz, International Journal of Hydrogen Energy 31 (2006) 2231 – 2242. [108] C.Y. Huang, Y.Y. Chen, C.C. Su, C.F. Hsu, Journal of Power Sources 174 (2007) 294-301. [109] B. Atalik, D. Uner, Journal of Catalysis 241 (2006) 268-275. [110] T. Ince, G. Uysal, A.N. Akin, R. Yildirim, Applied Catalysis, A: General 292 (2005) 171-176. [111] E.-Y. Ko, E.D. Park, K.W. Seo, H.C. Lee, D. Lee, S. Kim, Catalysis Letters 110 (2006) 275-279. [112] S.H. Oh, R.M. Sinkevitch, Journal of Catalysis 142 (1993) 254-262. [113] F. Mariño, C. Descorme, D. Duprez, Applied Catalysis B: Environmental 54 (2004) 59-66. [114] H.-S. Roh, H.S. Potdar, K.-W. Jun, S.Y. Han, J.-W. Kim, Catalysis Letters 93 (2004) 203-207. [115] A. Wootsch, C. Descorme, D. Duprez, Journal of Catalysis 225 (2004) 259-266. [116] H. Liu, L. Ma, S. Shao, Z. Li, A. Wang, Y. Huang , T. Zhang, Chinese Journal of Catalysis 28 (2007) 1077-1082. 222

[117] E.-Y. Ko, E.D. Park, K.W. Seo, H.C. Lee, D. Lee, S. Kim, Catalysis Today 116 (2006) 377-383. [118] G. Uysal, A.N. Akin, Z.I. Onsan, R. Yildirim, Catalysis Letters 108 (2006) 193- 196. [119] C. Pedrero, T. Waku, E. Iglesia, Journal of Catalysis 233 (2005) 242-255. [120] J. Yan, J. Ma, P. Cao, P. Li, Catalysis Letters 93 (2004) 55-60. [121] E. Simsek, S. Oezkara, A.E. Aksoylu, Z.I. Onsan, Applied Catalysis, A: General 316 (2007) 169-174. [122] A. Manasilp, E. Gulari, Applied Catalysis B: Environmental 37 (2002) 17-25. [123] G. Avgouropoulos, M. Manzoli, F. Boccuzzi, T. Tabakova, J. Papavasiliou, T. Ioannides, V. Idakiev, Journal of Catalysis 256 (2008) 237-247. [124] G. Avgouropoulos, J. Papavasiliou, T. Tabakova, V. Idakiev, T. Ioannides, Chemical Engineering Journal (Amsterdam, Netherlands) 124 (2006) 41-45. [125] M.M. Schubert, A. Venugopal, M.J. Kahlich, V. Plzak, R.J. Behm, Journal of Catalysis 222 (2004) 32-40. [126] M.J. Kahlich, H.A. Gasteiger, R.J. Behm, Journal of Catalysis 182 (1999) 430- 440. [127] H. Wang, H. Zhu, Z. Qin, F. Liang, G. Wang, J. Wang, Journal of Catalysis 264 (2009) 154-162. [128] D. Cameron, R. Holliday, D. Thompson, Journal of Power Sources 118 (2003) 298-303. [129] W.Y. Yu, W.S. Lee , C.P. Yang, B.Z. Wan, Journal of the Chinese Institute of Chemical Engineers 38 (2007) 151-160. [130] C. Rossignol, S. Arrii, F. Morfin, L. Piccolo, V. Caps, J.L. Rousset, Journal of Catalysis 230 (2005) 476-483. [131] M.M. Schubert, V. Plzak, J. Garche, R.J. Behm, Catalysis Letters 76 (2001) 143- 150. [132] J. Papavasiliou, G. Avgouropoulos, T. Ioannides, Applied Catalysis B: Environmental 66 (2006) 168-174. [133] A. Martı´nez-Arias, A.B. Hungrı´a, G. Munuera, D. Gamarra, Applied Catalysis B: Environmental 65 (2006) 207-216. 223

[134] E. Moretti, M. Lenarda, L. Storaro, A. Talon, R. Frattini, S. Polizzi, E. Rodrı´guez-Castello´n, A. Jime´nez-Lo´pez, Applied Catalysis B: Environmental 72 (2007) 149-156. [135] Y.-Z. Chen, B.-J. Liaw, W.-C. Chang, C.-T. Huang, International Journal of Hydrogen Energy 32 (2007) 4550-4558. [136] Y.Z. Chen, B.J. Liaw, J.M. Wang, C.T. Huang, International Journal of Hydrogen Energy 33 (2008) 2389-2399. [137] G. Marban, A.B. Fuertes, Applied Catalysis, B: Environmental 57 (2005) 43- 53. [138] P. Ratnasamy, D. Srinivas, C.V.V. Satyanarayana, P. Manikandan, R.S. Senthil Kumaran, M. Sachin, V.N. Shetti, Journal of Catalysis 221 (2004) 455–465. [139] H. Vidal, J. Kašpar, M. Pijolat, G. Colon, S. Bernal, A. Cordón, V. Perrichon, F. Fally, Applied Catalysis B: Environmental 27 (2000) 49-. [140] Y.Z. Chen, B.J. Liaw, H.C. Chen, International Journal of Hydrogen Energy 31 (2006) 427-435. [141] Y.Z. Chen, B.J. Liaw, C.W. Huang, Applied catalysis A: General 302 (2006) 168- 176. [142] C.-W. Tang, M.-C. Kuo, C.-J. Lin, C.-B. Wang, S.-H. Chien, Catalysis Today 131 (2008) 520-525. [143] C.-W. Tang, C.-C. Kuo, M.-C. Kuo, C.-B. Wang, S.-H. Chien, Applied Catalysis, A: General 309 (2006) 37-43. [144] F. Balikci, C. Guldur, Turkish Journal of Chemistry 31 (2007) 465-471. [145] J. Jansson, Journal of Catalysis 194 (2000) 55-60. [146] J. Jansson, A.E.C. Palmqvist, E. Fridell, M. Skoglundh, L. Oesterlund, P. Thormaehlen, V. Langer, Journal of Catalysis 211 (2002) 387-397. [147] J. Jansson, M. Skoglundh, E. Fridell, P. Thormahlen, Topics in Catalysis 16/17 (2001) 385-389. [148] P. Thormahlen, M. Skoglundh, E. Fridell, B. Andersson, Journal of Catalysis 188 (1999) 300-310. [149] Y. Teng, H. Sakurai, A. Ueda, T. Kobayashi, International Journal of Hydrogen Energy 24 (1999) 355-358. 224

[150] C. Kwak, T.-J. Park, D.J. Suh, Applied Catalysis, A: General 278 (2005) 181- 186. [151] Z. Zhao, M.M. Yung, U.S. Ozkan, Catalysis Communications 9 (2008) 1465- 1471. [152] M.M. Yung, Z. Zhao, M.P. Woods, U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 279 (2008) 1-9. [153] Q. Guo, Y. Liu, T. Caputo, L. Lisi, R. Pirone, G. Russo, Reaction Kinetics and Catalysis Letters 92 (2007) 19-25. [154] L.E. Gómez, I.S. Tiscornia, A.V. Boix, E.E. Miró, Applied Catalysis A: General 401 (2011) 124– 133. [155] K. Omata, Y. Kobayashi, M. Yamada, Catalysis Communications 6 (2005) 563- 567. [156] P. Broqvist, I. Panas, H. Persson, Journal of Catalysis 210 (2002) 198-206. [157] K. Omata, T. Takada, S. Kasahara, M. Yamada, Applied Catalysis, A: General 146 (1996) 255-267. [158] M.P. Woods, P. Gawade, B. Tan, U.S. Ozkan, Applied Catalysis B: Environmental 97 (2010) 28-35. [159] P. Gawade, B. Mirkelamoglu, U.S. Ozkan, Journal of physical Chemistry C 114 (2010) 18173-18181. [160] H. Song, B. Mirkelamoglu, U.S. Ozkan, applied Catalysis A: General 382 (2010) 58-64. [161] Z. Zhao, R. Jin, T. Bao, X. Lin, G. Wang, Applied Catalysis B: Environmental 110 (2011) 154-163. [162] S. Monyanon, S. Pongstabodee, A. Luengnaruemitchai, Journal of the Chinese Institute of Chemical Engineers 38 (2007) 435-441. [163] H. Song, U. Ozkan, Journal of Molecular Catalysis A: Chemical 318 (2010) 21- 29. [164] J. Haber, L. Ungier, Journal of Electron Spectroscopy and Related Phenomenon 12 (1977) 305-312. [165] J.P. Bonelle, J. Grimllot, A. D'Huysser, Journal of Electron Spectroscopy and Related Phenomenon 7 (1975) 151. 225

[166] J.C. Vickerman, I.S. Gilmore, Surface Analysis: The Principle Techniques 2nd Edition Wiley (2009). [167] J.-Y. Luo, M. Meng, X. Li, X.-G. Li, Y.-Q. Zha, T.-D. Hu, Y.-N. Xie, J. Zhang, Journal of Catalysis 254 (2008) 310-324. [168] M. Daturi, C. Binet, J.C. Lavalley, G. Blanchard, Surface and Interface Analysis 30 (2000) 273-277. [169] M. Daturi, C. Binet, J.C. Lavalley, A. Galtayries, R. Sporken, Physical Chemistry Chemical Physics 1 (1999) 5717-5724. [170] O. Pozdnyakova, D. Teschner, A. Wootsch, J. Kroehnert, B. Steinhauer, H. Sauer, L. Toth, F.C. Jentoft, A. Knop-Gericke, Z. Paal, R. Schloegl, Journal of Catalysis 237 (2006) 1-16. [171] T. Takeguchi, S. Manbe, R. Kikuchi, K. Eguchi, T. Kanazawa, M. Shinichi, W. Ueda, applied Catalysis A: General 293 (2005) 91-96. [172] C. Binet, M. Daturi, J.C. Lavalley, Catalysis Today 50 (1999) 207. [173] H. Kalies, N. Pinto, G.M. Pajonk, D. Bianchi, Applied catalysis A: General 202 (2000) 197. [174] A. Badri, C. Binet, J.C. Lavalley, Journal of Chem. Soc. Faraday Trans. 92 (1996) 4669-4673. [175] E. Finocchio, M. Daturi, C. Binet, J.C. Lavalley, G. Blanchard, Catalysis Today 52 (1999) 53-63. [176] S. Roy, A. Baiker, Chemical Reviews 109 (2009) 4054-4091. [177] R.M. Heck, Catalysis Today 53 (1999) 519-523. [178] P. Gelin, M. Primet, Applied Catalysis B: Environmental 39 (2002) 1-37. [179] V.I. Parvulescu, P. Grange, B. Delmon, Catalysis Today 46 (1998) 233-316. [180] R. Burch, J.P. Breen, F.C. Meunier, Applied Catalysis B: Environmental 39 (2002) 283-303. [181] M. Iwamoto, H. Hamada, Catalysis Today 10 (1991) 57-71. [182] M. Iwamoto, H. Yahiro, Catalysis Today 22 (1994) 5-18. [183] M. Iwamoto, H. Yahiro, Y. Yuu, S. Shundo, N. Mizuno, Shokubai 32 (1990) 430- 433.

226

[184] W. Held, A. Koenig, T. Richter, L. Puppe, Society of Automotive Engineers, [Special Publication] SP SP-810 (1990) 13-20. [185] B.J. Adelman, T. Beutel, G.D. Lei, W.M.H. Sachtler, Journal of Catalysis 158 (1996) 327-335. [186] A. Ali, W. Alvarez, C.J. Loughran, D.E. Resasco, Applied Catalysis B: Environmental 14 (1997) 13-22. [187] K.A. Bethke, C. Li, M.C. Kung, B. Yang, H.H. Kung, Catalysis Letters 31 (1995) 287-299. [188] R. Burch, S. Scire, Applied Catalysis B: Environmental 3 (1994) 295-318. [189] B.K. Cho, Journal of Catalysis 155 (1995) 184-195. [190] B.K. Cho, J.E. Yie, Applied Catalysis B: Environmental 10 (1996) 263-280. [191] P. Ciambelli, P. Corbo, M.C. Gaudino, F. Migliardini, D. Sannino, Topics in Catalysis 16/17 (2001) 413-417. [192] C.M. De Correa, F. Cordoba C, F. Bustamante L, Studies in Surface Science and Catalysis 130B (2000) 1469-1474. [193] C. Descorme, P. Gelin, C. Lecuyer, M. Primet, Applied Catalysis B: Environmental 13 (1997) 185-195. [194] C. Descorme, P. Gelin, C. Lecuyer, M. Primet, Journal of Catalysis 177 (1998) 352-362. [195] M. Ogura, E. Kikuchi, Studies in Surface Science and Catalysis 101 (1996) 671-680. [196] Z. Liu, S. Woo, Catalysis Reviews 48 (2006) 43–89. [197] B. Adelman, R. Rohe, R. Christopher, Ricardo Inc. CRC Project No. AVFL-7 (2002) C02 -2814. [198] J.A. Sullivan, J. Cunningham, Applied Catalysis B: Environmental 15 (1998) 275-289. [199] J.O. Petunchi, W. Keith Hall, Applied Catalysis B: Environmental 2 (1993) L17-L26. [200] A. Kubacka, J. Janas, E. Wloch, B. Sulikowski, Catalysis Today 101 (2005) 139- 145.

227

[201] M.D. Amiridis, T. Zhang, R.J. Farrauto, Applied Catalysis B: Environmental 10 (1996) 203-227. [202] J. Suzuki, S. Matsumoto, Topics in Catalysis 28 (2004) 171-176. [203] J.A.Z. Pieterse, H. Top, F. Vollink, K. Hoving, R.W. Brink, Chemical Engineering Journal 120 (2006) 17-23. [204] J.P. Breen, R. Burch, Topics in Catalysis 39 (2006) 53-58. [205] H. Hong He, X. Xiuli Zhang, Q. Wu, C. Changbin Zhang, Y. Yunbo Yu, Catal Surv Asia 12 (2008) 38-55. [206] H. He, J. Wang, Q.C. Feng, Y.B. Yu, K. Yoshida, Applied Catalysis B: Environmental 46 (2003) 365-370. [207] I. Balint, A. Miyazaki, K.-i. Aika, Applied Catalysis B: Environmental 37 (2002) 217-229. [208] G.R. Bamwenda, A. Obuchi, A. Ogata, J. Oi, S. Kushiyama, K. Mizuno, Journal of Molecular Catalysis A: Chemical 126 (1997) 151-159. [209] G.R. Bamwenda, A. Obuchi, A. Ogata, J. Oi, S. Kushiyama, K. Mizuno, Reaction Kinetics and Catalysis Letters 63 (1998) 53-59. [210] L.E. Lindfors, K. Era¨nen, F. Klingstedt, D.Y. Murzin, Topics in Catalysis 28 (2004) 1-4. [211] H. He, Y.B. Yu, Catalysis Today 100 (2005) 37-47. [212] Y.-H. Chin, A. Pisanu, L. Serventi, W. Alvarez, D.E. Resasco, Catalysis Today 54 (1999) 419-429. [213] B. Azambrea, L. Zenbourya, P. Da Costab, S. Capelac, S. Carpentierc, A. Westermanna, Catalysis Today 176 (2011) 242-249. [214] N. Li, A. Wang, Z. Liu, X. Wang, M. Zheng, Y. Huang, T. Zhang, Applied Catalysis B: Environmental 62 (2006) 292-298. [215] H. Zhang, N. Li, L. Li, A. Wang, X. Wang, T. Zhang, Catalysis Letters 141 (2011) 1491–1497. [216] V. Indovina, M.C. Cristina Campa, F. Pepe, D. Pietrogiacomi, S. Tuti, Applied Catalysis B: Environmental 60 (2005) 23-31. [217] C.J. Loughran, D.E. Resasco, Applied Catalysis B: Environmental 7 (1995) 113-126. 228

[218] H. Ohtsuka, T. Tabata, T. Hirano, Applied Catalysis B: Environmental 28 (2000) L73-L76. [219] H. Ohtsuka, Catalysis Letters 87 (2003) 179-186. [220] C.E. Quincoces, S. Guerrero, P. Araya, M.G. Gonzalez, Catalysis Communications 6 (2005) 75-80. [221] A.A. Rubert, C.E. Quincoces, A.-S. Mamede, D. Gazzoli, C. Cabello, M.G. Gonza´lez, Catalysis Communications 9 (2008) 1096-1100. [222] E. Long, X. Zhang, Y. Li, Z. Liu, Y. Wang, M. Gong, Y. Chen, Journal of Natural Gas Chemistry 19 (2010) 134. [223] B. Mirkelamoglu, U.S. Ozkan, Applied Catalysis B: Environmental 96 (2010) 421-433. [224] B. Mirkelamoglu, M. Liu, U.S. Ozkan, Catalysis Today 151 (2010) 386-394. [225] E.M. Holmgreen, M.M. Yung, U.S. Ozkan, Applied Catalysis B: Environmental 74 (2007) 73–82. [226] E.M. Holmgreen, M.M. Yung, U.S. Ozkan, Journal of Molecular Catalysis A: Chemical 270 (2007) 101-111. [227] E.M. Holmgreen, M.M. Yung, U.S. Ozkan, Catalysis Letters 111 (2006) 19-26. [228] M.M. Yung, E.M. Holmgreen, U.S. Ozkan, Journal of Catalysis 247 (2007) 356– 367. [229] P. Gawade, B. Mirkelamoglu, U.S. Ozkan, Applied Catalysis B: Environmental In Review (2011). [230] G. Djega-Mariadassou, M. Boudart, Journal of Catalysis 216 (2003) 89. [231] G. Djega-Mariadassou, Catalysis Today 90 (2004) 27. [232] B. Rotavera, A. Kumar, S. Seal, E.L. Petersen, Proceedings of the Combustion Institute 32 (2009) 811-819 [233] C.F. Cullis, T.G. Nevell, D.L. Trimm, Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases 68 (1972) 1406- 1412. [234] K. Fujimoto, F.H. Ribeiro, M. Avalos-Borja, E. Iglesia, Journal of Catalysis 179 (1998) 431-442.

229

[235] F.H. Ribeiro, M. Chow, R.A. Dalla Betta, Journal of Catalysis 146 (1994) 537- 544. [236] E. Escalona Platero, M. Penarroya Mentruit, Catalysis Letters 30 (1995) 31. [237] G. Larsen, E. Lotero, R.D. Parra, L.M. Petkovic, H.S. Silva, S. Raghavan, Applied Catalysis A: General 130 (1995) 213-226. [238] C. Morterra, G. Cerrato, F. Pinna, M. Signoretto, G. Strukul, Journal of Catalysis 149 (1994) 181. [239] F. Babou, G. Coudurier, J.C. Vedrine, Journal of Catalysis 152 (1995) 341-349. [240] B. Mirkelamoglu, U.S. Ozkan, Applied Catalysis B-Environmental 96 (2010) 421-433. [241] C. Morterra, G. Cerato, S. Di Ciero, Applied Surface Science 126 (1998) 107. [242] L.L. Sheu, H. Knozinger, W.M.H. Sachtler, Journal of Molecular Catalysis 57 (1989) 61. [243] C. Schild, A. Wokaun, Journal of Molecular Catalysis 63 (1990) 223. [244] L. Chen, L. Lin, Z. Xu, T. Zhang, Q. Xin, P. Ying, G. Li, C. Li, Journal of Catalysis 161 (1996) 107-114. [245] A. Yamaguchi, E. Iglesia, Journal of Catalysis 274 (2010) 52-63. [246] C. Li, Y. Sakata, T. Arai, K. Domen, K.I. Maruya, T. Onishi, J.Chem.Soc. Faraday Trans. 1 85(4) (1989) 929-943. [247] S. Bertarione, D. Scarano, A. Zecchina, V. Johánek, J. Hoffmann, S. Schauermann, J. Libuda, G. Rupprechter, H.-J. Freund, Journal of Catalysis 223 (2004) 64-73. [248] O.S. Alexeev, D.-W. Kim, B.C. Gates, Journal of Molecular Catalysis A: Chemical 162 (2000) 67-82. [249] R. Hicks, H. Qi, M. Young, R. Lee, Journal of Catalysis 122 (1990) 280. [250] E. Garwoski, C. Feumi-Jantou, N. Mouaddib, M. Primet, applied Catalysis A: General 109 (1994) 277. [251] A.M. Bradshaw, F. Hoffman, Surface Science 72 (1978) 513. [252] L.F. Liotta, G.A. Martin, G. Deganello, Journal of Catalysis 164 (1996) 322-333. [253] T. Yamaguchi, T. Jin, K. Tanabe, Journal of Physical Chemistry 90 (1986) 3148-3152. 230

[254] P. Gawade, A.-M.C. Alexander, R. Clark, U.S. Ozkan, Catalysis Today (2012) In Review. [255] R. Burch, F.J. Urbano, P.K. Loader, Applied Catalysis A: General 123 (1995) 173-184. [256] B.J. Adelman, W.M.H. Sachtler, Applied Catalysis B: Environmental 14 (1997) 1-11. [257] H. Ohtsuka, T. Tabata, Applied Catalysis B: Environmental 21 (1999) 133- 139. [258] H. Song, U.S. Ozkan, journal of physical Chemistry A 114 (2010) 3796-3801.

231

APPENDIX A

List of Acronyms

WGS: Water-gas shift

HT-WGS: High temperature water-gas shift

LT-WGS: Low temperature water-gas shift

SG: Sol-gel

GC: Gas chromatography

BET: Brunauer-Emmett-Teller

BJH: Barret-Joiner-Halenda

TPR: Temperature programmed reduction

TPO: Temperature programmed oxidation

XPS: X-ray photoelectron spectroscopy

XRD: X-ray diffraction

DRIFTS: Diffuse reflectance infra-red Fourier transform

TCD: Thermal conductivity detector

TEM: Transmission electron microscopy

CeO2-NP: Ceria nanoparticles

CeO2-NR: Ceria nanorods

PROX: Preferential oxidation of carbon monoxide

PEM: Proton exchange membrane

XANES: X-ray absorption near edge

232

WHSV: Weight hourly space velocity

GHSV: Gas hourly space velocity

SZ: Sulfated zirconia

SCR: Selective catalytic reduction

233

APPENDIX B

Sample Calculations

Agilent micro-GC 3000A or Shimadzu GC-14 equipped with TCD detector was

used to determine the gas stream concentrations. The concentrations of gases such as O2,

N2, CH4, CO, CO2, H2O and H2 was determined using the following the relation between

response factor (RF) and integrated peak area of the signal.

(Concentration of component i ) RF = i (Peak area of component i )

The concentrations of NO and NO2 were measured directly using the Thermo-42i- € HL chemi-luminescent NOx analyzer

The conversions and selectivity of the components was calculated using the

following relationships:

(Inlet Conc. of A − Outlet Conc. of A) Conversion of component A = (Inlet Conc. of A)

(Outlet Conc. of CO2 − Inlet Conc. of CO2) O2 selectivity to CO2 = 2 × (Inlet Conc. of O2 − Outlet Conc. of O2)

234

The following is the sample calculation for NOx-SCR project from the raw data acquired using Thermo-42i-HL and Agilent micro-GC 3000A series

RF O2 0.257 CO2 0.809 CH4 0.282 N2O 0.917 C2H6 0.842 C3H8 0.560 CO 0.214

Areas NOx Readings O2 CH4 C2H6 C3H8 CO CO2 N2O NO NO2 NOx Feed 385674 6169 262 238 3063 82595 0 7.1 176.2 183.3 300oC 382281 6160 261 127 0 81768 0 144.5 1 145.5 350oC 379924 6152 210 80 0 83487 0 135.8 1 136.8 400oC 380036 5797 47 0 0 83418 0 74 0.5 74.5 450oC 374622 3517 0 0 0 85177 0 59.9 2 61.9 500oC 369560 693 0 0 0 85447 0 103 6 109

Concentrations Conversions (%) O2 (%) CH4 C2H6 C3H8 CO CO2 (%) N2O NO NO2 NOx C-bal NOx CH4 C2H6 C3H8 CO Feed 9.899 1739.04 220.6 133.3 655.48 6.6803 0 7.1 176 183.3 300 9.812 1736.5 219.8 71.12 0 6.6134 0 145 1 145.5 97.84 20.62 0.146 0.38 46.64 100 350 9.751 1734.25 176.8 44.8 0 6.7524 0 136 1 136.8 99.58 25.37 0.276 19.8 66.39 100 400 9.754 1634.17 39.57 0 0 6.7468 0 74 0.5 74.5 98.78 59.36 6.03 82.1 100 100 450 9.615 991.442 0 0 0 6.8891 0 59.9 2 61.9 99.78 66.23 42.99 100 100 100 500 9.485 195.357 0 0 0 6.911 0 103 6 109 98.95 40.53 88.77 100 100 100

235

APPENDIX C

Sample calculation for methane/ethane oxidation kinetics

Part a: Methane oxidation over 2% CoOx/CeO2 (Differential reactor)

XCH # FCH 0 "r = 4 4, CH 4 W

Where W is the catalyst weight (g) and F is the initial CH flow rate (mol/min). The CH 4, 0 4 ! following Power-law model was assumed for the kinetic analysis:

r k (C )a (C )b ! " CH 4 = # CH 4 # O2

3 Where C and C are concentrations (mol/cm ) of CH4 and O2, respectively. CH 4 O2

! However, the oxygen concentration is approximately 100 times higher than that of the

! methane! concentration and hence the above Power-law model could be reduced to

r k* (C )a where, k* k (C )b " CH 4 = # CH 4 = " O2

r k* e(−Ea / RT ) (C )a − CH 4 = 0 × × CH 4 ! ! −E ln(−r ) = ln(k* ) + ( a ) + a ln(C ) CH 4 0 RT CH 4 €

€ Using a linear regression analysis, the kinetic parameters for CH4 oxidation over 2%

CoOx/CeO2 were found as activation energy (Ea)= 79 ± 13 KJ/mol and reaction order “a”

w.r.t. CH4 = 1.0 ± 0.3 236

Temp © Temp (K) (mol/min.gcat) CA (mol/cm3) ln (rate) 1/T ln (Ca) 400 673 2.19969E-06 1.5889E-08 -13.0272 0.001486 -17.9577 400 673 4.86092E-06 3.0913E-08 -12.2343 0.001486 -17.2921 400 673 6.2203E-06 3.7816E-08 -11.9877 0.001486 -17.0905 400 673 6.52601E-06 4.6139E-08 -11.9397 0.001486 -16.8916 450 723 4.02585E-06 1.4584E-08 -12.4228 0.001383 -18.0433 450 723 9.72183E-06 2.8228E-08 -11.5411 0.001383 -17.3829 450 723 1.16366E-05 3.4591E-08 -11.3614 0.001383 -17.1797 450 723 1.27573E-05 4.2247E-08 -11.2694 0.001383 -16.9797 500 773 1.25341E-05 1.2745E-08 -11.2871 0.001294 -18.1781 500 773 2.38855E-05 2.4911E-08 -10.6422 0.001294 -17.5079 500 773 2.7674E-05 3.0665E-08 -10.495 0.001294 -17.3001 500 773 3.23774E-05 3.7449E-08 -10.338 0.001294 -17.1003

Part b: Ethane oxidation over mixed-bed (Integral reactor)

Plot of XC2H6 vs. W/FC2H6,0,, where FC2H6, 0, is initial flow rate of ethane (mol/min). The

slope at each data point was used to derive the reaction rate. The reaction rate was

calculated by applying the differential analysis as follows:

dX "r = C2H6 C2 H 6 W d( F ) C2 H 6 ,0

The following Power-law model was assumed for the kinetic analysis:

! r k (C )a (C )b − CH 4 = × C2 H 6 × O2

3 Where C and C are concentrations (mol/cm ) of C2H6 and O2, respectively. C2 H 6 O2

€ However, the oxygen concentration is approximately 1000 times higher than that of the

€ ethane concentration! and hence the above Power-law model could be reduced to

r k* (C )a where, k* k (C )b − C2 H 6 = × C2 H 6 = " O2

r k* e(−Ea / RT ) (C )a − C2 H 6 = 0 × × C2 H 6 € ! 237 €

−E ln(−r ) = ln(k* ) + ( a ) + a ln(C ) C2 H 6 0 RT C2 H 6

A linear regression analysis was applied to obtain the kinetic parameters for C2H6

€ oxidation over 2%CoOx/CeO2 as followed, activation energy (Ea)= 45 ± 3 KJ/mol and

reaction order “a” w.r.t. C2H6 = 0.7 ± 0.08.

Temp © Temp (K) (mol/min.gcat) CA (mol/cm3) ln (rate) 1/T ln (Ca) 300 573 6.2488E-07 2.143E-09 -14.2857105 0.0017452 -19.9610808 300 573 8.2345E-07 4.3272E-09 -14.0097589 0.0017452 -19.2583449 300 573 8.9327E-07 5.3079E-09 -13.9283809 0.0017452 -19.0540695 300 573 9.6908E-07 6.4528E-09 -13.8469163 0.0017452 -18.8587553 350 623 1.0572E-06 1.7619E-09 -13.7598722 0.00160514 -20.1568589 350 623 1.6354E-06 3.5602E-09 -13.3236344 0.00160514 -19.4534611 350 623 1.8599E-06 4.3121E-09 -13.1949881 0.00160514 -19.2618424 350 623 2.1155E-06 5.5165E-09 -13.066205 0.00160514 -19.0155204 400 673 1.5766E-06 1.2302E-09 -13.3602281 0.00148588 -20.5161059 400 673 2.599E-06 2.5905E-09 -12.860371 0.00148588 -19.7714251 400 673 3.0118E-06 3.45E-09 -12.7129634 0.00148588 -19.4848931 400 673 3.4907E-06 3.9734E-09 -12.565399 0.00148588 -19.3436324 450 723 1.8188E-06 6.9478E-10 -13.2173442 0.00138313 -21.0874222 450 723 3.1544E-06 1.6209E-09 -12.666712 0.00138313 -20.2402553 450 723 3.7105E-06 2.0171E-09 -12.5043308 0.00138313 -20.0216142 450 723 4.3655E-06 2.6037E-09 -12.341777 0.00138313 -19.7663143 500 773 1.4902E-06 3.0085E-10 -13.4166169 0.00129366 -21.9244002 500 773 2.7035E-06 7.8937E-10 -12.8209628 0.00129366 -20.9597809 500 773 3.2227E-06 1.0302E-09 -12.6453046 0.00129366 -20.693524 500 773 3.8422E-06 1.3238E-09 -12.4694597 0.00129366 -20.4427445

238