Direct Propene Epoxidation over Gold-Titania Catalysts
Transcription
Direct Propene Epoxidation over Gold-Titania Catalysts
Direct Propene Epoxidation over Gold–Titania Catalysts: Kinetics and Mechanism PROEFSCHRIFT ter verkrijging van de graad van doctor aan de Technische Universiteit Eindhoven, op gezag van de rector magnificus, prof.dr.ir. C.J. van Duijn, voor een commissie aangewezen door het College voor Promoties in het openbaar te verdedigen op maandag 18 november 2013 om 14.00 uur door Jiaqi Chen geboren te Rudong, China Dit proefschrift is goedgekeurd door de promotor: prof.dr.ir. J.C. Schouten Copromotor: dr.ir. T.A. Nijhuis The Netherlands Organization for Scientific Research (NWO) is kindly acknowledged for providing an ECHO grant (700.57.044) to make the research described in this thesis possible. Chen, Jiaqi Direct Propene Epoxidation over Gold–Titania Catalysts: Kinetics and Mechanism Technische Universiteit Eindhoven, 2013 A catalogue record is available from the Eindhoven University of Technology Library ISBN: 978-90-386-3465-4 Copyright © 2013 by Jiaqi Chen Cover photo provided by Shutterstock, Inc. under the Single User Standard License Contents Summary vii 1 Introduction 1 1.1 Propene oxide and its production . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Direct epoxidation over gold based catalysts . . . . . . . . . . . . . . . . . . . 3 1.3 Microreactor system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1.4 Objectives and outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2 Kinetic study of the direct propylene epoxidation in the explosive regime 19 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.2.1 Catalyst preparation and characterization . . . . . . . . . . . . . . . . 22 2.2.2 Catalyst testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.3.1 Performance of microreactor and proof of concept . . . . . . . . . . . 24 2.3.2 Catalyst characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.3.3 Product formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.3.4 Propene epoxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.3.5 Water formation and link to epoxidation . . . . . . . . . . . . . . . . . 31 2.3.6 Effect of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 2.3.7 Relationship between selectivity, hydrogen efficiency, catalyst stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 2.4 Summarizing discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 iv CONTENTS Appendix 2.A Deactivation model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 Appendix 2.B Activation energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 3 Enhancement of catalyst performance: A study into gold–titanium synergy 55 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 3.2.1 Preparation of supports . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 3.2.2 Deposition of Au . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 3.2.3 Inversed incorporation of Ti onto Au/SiO2 . . . . . . . . . . . . . . . . 59 3.2.4 Catalytic testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 3.2.5 Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3.3.1 Characterization of the supports . . . . . . . . . . . . . . . . . . . . . . 60 3.3.2 Size of Au particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 3.3.3 PO formation and water formation . . . . . . . . . . . . . . . . . . . . . 64 3.3.4 Performance of the catalyts with inversely-grafted Ti . . . . . . . . . . 70 3.3.5 Effect of supports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 3.3.6 Propane formation and its suppression . . . . . . . . . . . . . . . . . . 73 3.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 3.4.1 Role of the support in enhanced PO productivity . . . . . . . . . . . . 73 3.4.2 Competition of epoxidation and water formation at the Au–Ti interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 3.4.3 Origin of the activity in propene hydrogenation . . . . . . . . . . . . . 79 3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 Appendix 3.A Tables of catalyst performance . . . . . . . . . . . . . . . . . . . . . . 80 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 4 Switching off propene hydrogenation in the direct epoxidation of propene 85 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 4.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 4.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 4.3.1 Activation of hydrogenation activity . . . . . . . . . . . . . . . . . . . . 88 4.3.2 Switching off propene hydrogenation by CO . . . . . . . . . . . . . . . 90 v CONTENTS 4.3.3 Probing the active site for propene hydrogenation . . . . . . . . . . . 97 4.4 Summarizing discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 5 How metallic is gold in the direct epoxidation of propene 115 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 5.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 5.2.1 Catalyst preparation and testing . . . . . . . . . . . . . . . . . . . . . . 118 5.2.2 Charaterization techniques . . . . . . . . . . . . . . . . . . . . . . . . . . 119 5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 5.3.1 Catalyst performance and characterization . . . . . . . . . . . . . . . . 120 5.3.2 CO adsorption on catalysts after epoxidation and regeneration . . . 121 5.3.3 CO adsorption on Au/Ti-SiO2 treated by O2 and H2 . . . . . . . . . . 130 5.3.4 CO adsorption on Au/TiO2 in the presence of H2 . . . . . . . . . . . . 132 5.3.5 C3 H6 adsorption on Au/TiO2 in the presence of CO . . . . . . . . . . 136 5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146 6 Conclusions and outlook 151 6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 6.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 List of Publications 157 Acknowledgements 159 About the Author 163 Summary Direct Propene Epoxidation over Gold–Titania Catalysts: Kinetics and Mechanism Propene oxide (PO) is a major bulk material used in chemical industry. The current main processes in operation, the chlorohydrin process and the hydroperoxide processes, are both complex and have a number of disadvantages. As an alternative, the direct epoxidation of propene to propene oxide using hydrogen, oxygen and propene is highly selective over a gold–titania based catalyst and has gained considerable attention in the past 15 years. However, one of the main hurdles for the industry application is the relatively low activity of such Au/Ti-support catalysts. Performing this reaction into the explosive region of the reactant mixture is supposed to be beneficial for a higher productivity since the formation of the reactive hydroperoxy intermediate from hydrogen dissociation is rate determining, bringing this process closer towards industrialization. The usage of a microreactor system for the direct epoxidation of propene over a goldtitania based catalyst system using a mixture of hydrogen, oxygen, and propene allows for the safe operation of the reaction in the explosive regime. A kinetic study was performed on the effect of the concentration of hydrogen, oxygen, and propene on the reaction rate as well as the catalyst deactivation and reactivation. A simple algebraic rate expression was developed, based on published kinetics, which provided the three reaction rate constants as a function of the feed gas concentrations. The observed rate dependency on the reactants for the epoxidation and the competing direct water formation is discussed in relation to the current mechanistic insights in literature. The formation rate of propene oxide is most dependent on the hydrogen concentration, in which the formation of an active peroxo species on the gold nanoparticles is the rate determining step. The deactivation is mainly caused by the consecutive oxidation of propene oxide. Oxygen favours the regeneration of the deactivated catalytic sites. Water formation and propene epoxidation are strongly correlated. Water is formed via two routes: through the active peroxo viii SUMMARY intermediate responsible for epoxidation and from direct water formation without involving this active intermediate. Improving the hydrogen efficiency, defined as the ratio between PO formation rate and total water formation rate, should distinguish between these two routes of water formation. The active peroxo intermediate in the epoxidation is competitively consumed by hydrogenation and epoxidation. The active gold site is blocked during deactivation, equally inhibiting epoxidation and water formation. The hydrogen efficiency should distinguish between these two routes of water formation. Based on these understandings, an increase in Au-Ti interface or a better synergy between Au and Ti should increase the hydrogen efficiency by mitigating direct water formation. These findings were validated in the catalyst optimization aiming at a catalyst with a better Au-Ti synergy. An enhanced productivity toward propene oxide in the direct propene epoxidation with hydrogen and oxygen over gold nanoparticles supported on titanium-grafted silica was achieved by adjusting the gold–titanium synergy. The ob−1 tained PO formation rates in this study (120–130 gPO · kg−1 cat h ) were approximately the same as the highest rates reported in literature. Highly isolated titanium sites were obtained by lowering the titanium loading grafted on silica. These active catalysts have low gold loadings of around or below 0.2 wt.%. The tetracoordinated Ti sites were attained by lowering the Ti loading to 0.2–0.3 wt.% (ca. 0.1 Ti/nm2 ). The tetrahedrally coordinated titanium sites were found to be favorable for attaining small gold nanoparticles and thus a high dispersion of gold. The improved productivity of propene oxide can be attributed to the increased amount of the interfacial Au–Ti sites. The active hydroperoxy intermediate is competitively consumed by epoxidation and hydrogenation at the Au–Ti interface. The PO/water ratio at the high PO rates obtained in this study ranged between 10 and 20 %. A higher propene concentration is favorable for a lower water formation rate and a higher formation rate of propene oxide. Under certain circumstances, propane formation may also happen or even prevail over the gold-titania catalysts. Propene hydrogenation was encountered during our study into the site synergy between gold and titanium using Ti-SiO2 as the support. The side reaction of propene hydrogenation over these gold–titania catalysts was studied in details. The addition of a small amount of carbon monoxide (10–1000 ppmv level) to the feed gas can completely switch off this propene hydrogenation, while at the same time also reducing the rate of direct water formation. The formation rate of propene oxide was not affected by the addition of carbon monoxide. An increase in the formation rate of carbon dioxide was observed when 1000 ppm carbon monoxide was added and the SUMMARY ix CO conversion was only 10 %. Hydrogenation of carbon monoxide to methane was not observed. The order of CO on this inhibiting effect is −1. Gold is not necessary for propane formation. The supports alone showed the same hydrogenation behavior as the catalysts: 1. enhancement of propene hydrogenation by oxygen; 2. peak activity at ca. 443 K in propene and hydrogen during temperature programmed reaction; 3. switching off by carbon monoxide with an order of −1. The coordination environment of titanium and surface hydroxyls may play an important role in propene hydrogenation. The most important place for PO formation is interfacial Au–Ti sites. Unraveling the oxidation state of gold is important to the understanding of the direct propene epoxidation on the gold–titania catalysts. Carbon monoxide was used as probe molecule in the infrared study to investigate the electron density of low-coordinated gold atoms on the gold–titania catalysts that are active in the direct propene epoxidation. The active gold sites were fully covered by reaction intermediates and deactivating species after the reaction. These species occupying the gold sites could not desorb even at 573 K. Calcination in oxygen removed the carbonaceous species on gold. The gold atoms were positively charged when oxygen was adsorbed on gold or at the interface. Reduction in hydrogen removed the adsorbed oxygen and the positively charged gold was reduced to its metallic form. When propene was adsorbed on the catalyst, gold atoms were negatively charged showing the carbonyl band as low as 2079 cm−1 . Carbon monoxide was replaced by propene on the catalyst surface and oxidation of carbon monoxide was suppressed by propene. Hydrogen significantly increased the coverage of carbon monoxide on the titania surface. The results from the infrared study provide a general scheme of electron transfer via gold on the gold–titania catalysts for the direct propene epoxidation. Introduction 1 1.1 Propene oxide and its production O Propene oxide (PO, CH3 CH CH2 ), also know as propylene oxide, methyloxirane, or 1,2-epoxypropane, is an important basic chemical used in the chemical industry and consumes over 10 % of all propylene produced [1]. Its major application is in the production of polyether polyols and propylene glycols, which are the starting materials for polyurethane and polyester production. The total production of propene oxide reached 7.7 million tons in 2012 after recovering from the 2008–2010 recession, and has grown by 25 % in over a decade when compared to 5.8 million tons in 1999 [1, 2]. The demand is expected to experience a positive development in general and is on high rise of 8 % especially in the emerging markets [2]. Contrary to the production of ethylene oxide, a direct process for the epoxidation is not yet available for propene. Two main routes are used in industry for PO production, the chlorohydrin process and the hydroperoxide processes. In the chlorohydrin process propene reacts in aqueous chlorine solution with hypochlorous acid to produce the propylene chlorohydrins (PCH). The chlorohydrins thereafter are dechlorinated, using a base (usually calcium hydroxide) to produce propene oxide. The hydroperoxide processes are a group of processes which use an alkyl-hydroperoxide as the intermediate oxidant to epoxidize propene producing propene oxide and an alcohol. In commercial operation, two main hydroperoxide processes, namely, the propene oxide–styrene monomer (SMPO) process and the propene 2 CHAPTER 1. INTRODUCTION oxide-tert-butyl alcohol (PO/TBA) process are applied. Approximately 60 % of the hydroperoxide plants use the SMPO process. In this process, ethylbenzene is oxidized by air to ethylbenzene hydroperoxide, which reacts with propene to produce propene oxide and 2-phenylethanol. After dehydration, 2-phenylethanol is converted to styrene. In the PO/TBA process, isobutane is oxidized to tert-butyl hydroperoxide (TBHP), which reacts with propene to produce propene oxide and tert-butyl alcohol (TBA). The third hydroperoxide process is developed by Sumitomo Chemical using cumene hydroperoxide, which is formed via oxidation of cumene, as the intermediate peroxide. The Sumitomo PO-only Cumene process produces no co-products since the formed α, α-dimethyl benzy alcohol is then simultaneously hydrogenated and dehydrated back to cumene for re-use. Propylene glycol 20% 5% 60–70% Polyether polyols Polyurathane (rigid foams) → refrigerator, insulator Polyurathane (flexible foams) → car seat, mattress, carpet Non foams → elastomer, adhesive, paint Unsaturated polyester resins → building materials Industrial use → antifreeze, lubricant Pharmaceuticals → toothpaste, cosmetics Propylene-based glycol ethers Other propoxylated compounds Figure 1.1: Uses for propene oxide [3, 4] The current main processes in operation, the chlorohydrins process and the hydroperoxide processes, are both complex and have a number of disadvantages. The chlorohydrins process produces large quantities of chloride/chlorine containing waste materials. For this reason no new chlorohydrins based plants have been built over the past 20 years [5]. The hydroperoxide processes (except Sumitomo PO-only Cumene process) produce a co-product (usually alcohols) in a quantity 2-3 times larger than that of propene oxide and, therefore, they are less flexible towards changing market demands. The recently commercialized process, the hydrogen peroxide process (HPPO) by BASF, Dow and Solvay as well as Evonik and Uhde, uses hydrogen peroxide to oxidize propene producing PO and water. Titanium silicalite (TS-1) zeolite is used as the catalyst and the epoxidation reaction is performed under mild temperature (< 100 o C) and high pressure 3 1.2. DIRECT EPOXIDATION OVER GOLD BASED CATALYSTS (ca. 25 bar) in liquid phase using methanol as a solvent and co-catalyst [6–9]. This process is more environmentally friendly and more flexible towards market demands. However, an on-site H2 O2 production via anthraquinone process is indispensable and a large amount of methanol needs to be separated and recycled. A high ratio of methanol is necessary for a stable operation to avoid formation of biphasic liquid [9]. Methanol in HPPO is generally considered as a co-catalyst[10], while Zwijnenburg et al. suggested that it may help PO desorption based on their study over a Au/TiO2 /SiO2 catalyst[11]. A comparison of commercial processes for propene oxide production is given in Table 1.1 1.2 Direct epoxidation over gold based catalysts The deficiencies of the mentioned chlorohydrin and hydroperoxide processes have spurred research into the production of PO as a single product derived from the oxidation of propene. Unlike ethylene epoxidation, the direct oxidation of propene with oxygen over silver catalyst is insufficiently selective due to a facile combustion via an allylic intermediate [12]. Considerable interest has arisen in the oxidation of propene to propene oxide in H2 /O2 mixtures after the discovery of gold–titania catalysts by Haruta et al. [13] in 1998. These promising catalysts are the catalysts consisting of gold nanoparticles smaller than 5 nm or sub-nano clusters on Ti-containing supports, e.g., TiO2 , TiO2 on SiO2 , and titanosilicates. They are typically used at a temperature of 50 – 200 o C, just above atmo−1 spheric pressure and a GHSV between 5000 and 10000 mL·g−1 cat h . The main advantage of these gold-based catalytic system is their high selectivity. The following reactions proceed on over the gold–titania catalysts in a mixture of propene, hydrogen and oxygen: O C3 H6 + H2 + O2 −→ CH3 CH CH2 + H2 O + 356.9 kJ (R1.1) 1 H2 + O2 −→ H2 O + 241.8 kJ 2 (R1.2) C3 H6 + 2 O2 −→ CH3 CHO + CO2 + H2 O + 826.5 kJ (R1.3) C3 H6 + H2 + O2 −→ CH3 CH2 CHO + H2 O + 450.9 kJ (R1.4) C3 H6 + H2 + O2 −→ CH3 COCH3 + H2 O + 480.7 kJ (R1.5) C3 H6 + H2 −→ C3 H8 + 125.1 kJ (R1.6) 4 Table 1.1: Comparison of commercial processes for PO production [3, 5–7, 9, 14–16] Process PCH PO PO/TBA PO/SM or SMPO Sumitomo Cumene HPPO (1910– ) (1969– ) (1974, 1980– ) (2003– ) (2008– ) 46 % 16 % 30 % 4% 4% Precursors Cl2 , H2 O i-butane ethylbenzene cumene H2 , O2 Intermediate HOCl, PCH t-butylhydroperoxide ethylbenzene hydroperoxide cumene Catalyst non-catalytic molybdenum naphthenate molybdenum naphthenate Share in capacity a H2 O2 hydroperoxide b Typical conditions b steam stripper, 0.1 MPa o PO selectivity (per t PO) b. only for epoxidation or silylated titania-on-silica containing Ti CSTR, 3.5 MPa CSTR or fixed bed o o fixed bed fixed bed o 3–4 MPa, 80–100 C <5 MPa, ca. 60 C ®3 MPa, <100 o C 87–90 % ¦95 % ∼ 95 % 95 % >90 % ¾ 4 t t-butanol ¾ 2.2 t styrene ∼ 1.5 t cumyl-alcohol ¾ 0.3 t H2 O ¾ 40 t H2 O CHAPTER 1. INTRODUCTION a. based on year 2009 (homogeneous) TS-1, MeOH top: 40 C; bottom: 100 C 120 C Co-products/Recycle ¾ 2 t chloride salts o mesoporous silica 5 1.2. DIRECT EPOXIDATION OVER GOLD BASED CATALYSTS 9 C3 H6 + O2 −→ 3 CO2 + 3 H2 O + 1926.3 kJ 2 (R1.7) The selectivity towards PO based on propene is generally higher than 90 mol%. The main side products are acetaldehyde, propionaldehyde, acetone, carbon dioxide and propane. Combustion of propene can only happen at higher reaction temperature, which depends on the catalyst, e.g., above 80 o C on Au/TiO2 , or above 200 o C for Au/TS-1. Propene hydrogenation is generally not encountered if the deposition–precipitation (DP) method is used for preparing the supported gold catalysts [13, 17, 18]. This side reaction has only raised concern in the recent two or three years when severe propene hydrogenation was also observed on gold catalysts prepared by DP method [19–21]. To obtain a PO selectivity above 95 % or even higher is not an issue, but it may sacrifice the activity which is already low. On the first investigated 1 wt% Au/TiO2 catalyst [13], propene conversion was below 1 % though with a selectivity of almost 100 %. According to the study by Haruta et al. [22, 23], the one-pass conversion of propene needs to reach ca. 10 −1 % at PO selectivity of 90 %, or equivalently a space-time-yield above 90 gPO · kg−1 at cat h −1 4000 mL·g−1 (10 vol% propene in feed), for an economically viable process in induscat h try. Besides, the hydrogen efficiency, defined as the amount of PO produced divided by the amount of water formed, should be increased up to 50 %. In literature, the hydrogen efficiency is generally below 30 %. Table 1.2 summarizes representative performance of gold–titania catalysts in the direct epoxidation (or hydro-epoxidation) of propene. At the beginning phase of catalyst development for the gold–titania system, low conversion of propene (typically <2 %), fast deactivation as well as relatively poor hydrogen efficiency (defined as the ratio between PO formation rate and total water formation rate) were the main hurdles towards industrial implementation. Tremendous work has been done to improve the catalyst performance by means of selecting and optimizing supports, optimizing the combination between the catalytic Au and Ti sites, or adding promoters. At present, the activity and stability of the Au/Ti-support catalyst have been improved to a commercially interesting range while efforts still need be done in enhancing the hydrogen efficiency (considering the production rate of PO and catalyst stability, 47 % in hydrogen efficiency is the best figure). Figure 1.2 gives a concise route map of the catalyst development. An in-depth review of the full progress in catalyst development till the year 2008 can be found in [24]. 6 Table 1.2: Direct propene epoxidation in hydrogen and oxygen: representative catalytic performance at atmospheric pressure Catalyst Temperature GHSVa o Conversion Selectivity Efficiency PO STYa ( C) −1 (mL·g−1 cat h ) of C3 H6 (%) of PO (%) of H2 (%) 1 wt% Au/TiO2 50 4000 1.1 > 99 35 11 1 wt% Au/TS-1 150 7000 1.1 > 99 5 18 0.1 wt% Au/Ti- 140 3000 8.4–7.8 95 unknown ∼ 35 j DP method [13] organic/inorganic [26] hybrid support 1 wt% Au/Ti-SiO2 150 9000 0.83 0.3–0.4 wt% Au/ 150 4000 5.0–9.8 87 7.1 18 90–95 27–35 130–150 200 7000 10 76 ® 30 134d 200 7000 8.8 81 ∼ 30 150 4000 8.5 91 35 f b titanosilicate c SiO2 , 300 m2 /g [27] mesoporous, Ti [22] content 2–6% TS-1 (Si/Ti ∼ 36) 0.05 wt% Au/ Ref. [25] trialkoxysilane 0.081 wt% Au/ Notes stable in 45 h, NH4 NO3 [28] treatment on TS-1 e 116 steady state [29] 64–80 silylation, 13–15 [23] TS-1 (Si/Ti = 36) titanosilicate (Si/Ti = 100 : 3) ppm (CH3 )3 N cofed CHAPTER 1. INTRODUCTION Ba(NO3 )2 -Au/ Catalyst Temperature GHSVa o ( C) 0.33 wt% Au/TS-1 200 Conversion Selectivity Efficiency PO STYa Notes −1 (mL·g−1 cat h ) of C3 H6 (%) of PO (%) of H2 (%) 7000 9.7 87 10 ± 5 132 (Si/Ti = 28) g Ref. carbon pearls in TS-1, [30] meso-scale defects 0.25 wt% Au/TS-1 200 8000 8.8 82 20 137 TS-1 treated by [31] NaOH for 1 h (Si/Ti = 48)h 0.05 wt% Au/TS-1 6.0 88 47 100 same as above 14.6 ∼ 70 ∼ 25 164 ionic liquid-enhanced 1.2. DIRECT EPOXIDATION OVER GOLD BASED CATALYSTS Table 1.2: (Continued) (Si/Ti = 48) h 1.0 wt% Au/TS-1 300 7000 (Si/Ti = 35) Au/TS-1 (Si/Ti = 99, 100) [32] immobilization of Au 200 14000 ∼5 a. generally, H2 /C3 H6 /O2 /balance= 1/1/1/7, 88–94 17 ± 2 158 ± 9 pH∼ 7.3 for DP, [33] Au loadings <0.1 wt% i −1 gPO · kg−1 cat h ; b. taken at 0.5 h; c. on a silylated and Ba-promoted catalyst before deactivation; d. steady state; e. similar to [22], but no direct quantification; f . increased from ∼ 17% by (CH3 )3 N cofeeding; g. steady state in 45 h; h. solid grinding method for gold anchoring; i. 9 catalysts; j, H2 /C3 H6 /O2 /balance= 75/6/5/14 7 8 CHAPTER 1. INTRODUCTION Figure 1.2: A brief route map of development of gold–titania catalysts for the direct propene epoxidation Mechanism The mode of operation of the gold–titania catalysts is assumed to be the production of a peroxide species on the gold nanoparticles, which epoxidize propene over a neighbouring titanium(IV) site [13, 17, 25, 27, 34]. Titanium(IV) based catalysts are well know for their capacity to epoxidize propene using organic peroxide, or hydrogen peroxide via the Ti(η2 -OOR) or Ti(η1 -OOH) intermediate as shown in Figure 1.3. Figure 1.3: Proposed intermediate for propene epoxidation in SMPO (left) and HPPO (right) processes [10, 35, 36] The gold nanoparticles that are active in the direct epoxidation of propene usually have a size within 2–5 nm and are prepared via the DP method. The fact that the oxidation reaction of propene to form PO requires the presence of both hydrogen and oxygen, as well as the fact that propene can be epoxidized very selectively by hydrogen peroxide over TS-1, creates the assumption that a peroxide species is involved in the re- 1.2. DIRECT EPOXIDATION OVER GOLD BASED CATALYSTS 9 action mechanism. Theoretical calculations showed that OOH or HOOH can be formed on small gold particles [37–39]. The hydro-peroxy species (HOOH or OOH) on gold was spectroscopically confirmed via an inelastic neutron scattering (INS) study by the group of Goodman [40]. Direct synthesis of hydrogen peroxide over supported gold and gold–palladium catalysts have been experimentally proved by Ishihara et al. [41] and by Edwards et al. [42–44], but the selectivity was not high. The DFT study by Ford et al. [45] explained that the preference of HOOH formation on gold is due to the weak Au−O bonding and the inefficient scission of O−O bond. Based on a combination of UV–vis and XANES study, the group of Oyama [34] proposed a mechanism for the propene epoxidation over gold supported on titanosilicates, in which the formation of HOOH is transferred to Ti(IV) site forming the active Ti−OOH intermediate as shown in Figure 1.4. However, proof of this ‘sequential’ mechanism is not flawless. An underlying assumption in their study is that propene does not affect HOOH formation and decomposition, which may not be true. A ‘simultaneous’ mechanism is proposed by the group of Delgass [46] based on their DFT study. They proposed that attack of HOOH to Ti-defect sites is energetically unfavourable and propene is likely adsorbed on Au–Ti interface sites attacking H−Au−OOH species. The most favourable site for propene adsorption on Au/TiO2 was also found to be Au–Ti interface as evidenced by the TPD study by Ajo et al. [47]. Figure 1.4: A ‘sequential’ mechanism proposed for propene epoxidation with hydrogen and oxygen over gold supported on titanosilicates. Reproduced from [34]. Another mechanism proposed by Nijhuis et al. [27] for Au/TiO2 suggests that a bidentate propoxy species formed on titania in adjacent to gold may be a reaction intermediate 10 CHAPTER 1. INTRODUCTION in the direct epoxidation of propene. Infrared spectroscopic information in their study indicated that propene is activated by gold nanoparticles and reacts with neighbouring Ti sites [27]. The formed bidentate propoxy species can be further oxidized to carbonate/carboxylate species leading to catalyst deactivation. This mechanism is depicted in Figure 1.5. The adsorption of propene on gold in this mechanism was evidenced by insitu XANES study on Au/SiO2 [48]. The participation of lattice oxygen as depicted in Figure 1.5 was latter investigated by a SSITKA study [49]. However, the role of support oxygen in forming the bidentate species could not be conclusively answered. Whether or not this bidentate species is the true reaction intermediate is still in debate. In the study by Mul et al. [50], the bidentate propoxy species is considered as a spectator formed from irreversible adsorption of propene oxide. Nevertheless, the effect on short-term deactivation by the carbonate/carboxylate species is generally accepted [27, 50, 51]. The catalyst activity after deactivation can be restored by calcination in oxygen. Rate determining H2 + O2 Au HOOH Au Au slow C3H6O Au C3H6 + Au C3H6 + Ti O Ti O fast + Ti O Ti O O CH2 CH CH3 OH OH O OH OH + fast OH OH O Au O O O Ti Ti Ti Ti O O + H2O O Au O OH C O2 o.a. CO2, CH3CHO + CH2CH3 O O O Ti Ti very slow OH OH O Ti O Ti O Deactivation Figure 1.5: The bidentate propoxy mechanism proposed for the direct propene epoxidation over Au/TiO2 by Nijhuis et al. [27]. Reproduced from ref [1]. Kinetic studies Table 1.3 summarizes the reaction orders in the power-rate-law (PRL) expression for gold-based propene epoxidaiton in hydrogen and oxygen. The ranges of dependency on hydrogen, oxygen and propene are 0.5–0.6, 0.2–0.3 and 0.2–0.3, respectively. Taylor et al. [52] used three Au/TS-1 catalysts to carry out the kinetic analysis. On average, 1.2. DIRECT EPOXIDATION OVER GOLD BASED CATALYSTS 11 the apparent activation energy in their study is 42 kJ/mol with a range between 35 and 55 kJ/mol for individual catalysts. However, this value is a bit higher than the apparent activation energy determined on a series of Au/TS-1 catalysts studied earlier in the same group [29], i.e., 25–36 kJ/mol. The mechanism they proposed to explain the fractional orders includes the HOOH formation on a single Au site (HOOH-Au) and the epoxidation of propene on an adjacent Ti site (C3 H6 -Ti). The rate-determining step (RDS) was proposed to be the epoxidation step between the two surface species, i.e., HOOHAu and C3 H6 -Ti. Lu et al. [53] performed the kinetic study over a barium-promoted gold catalyst supported on a mesoporous titanosilicate or Ti-TUD. Their explanation for the fractional dependencies differentiates from what Taylor et al. [52] proposed in two aspects: 1) the formation of HOOH involves two types of Au sites; 2) HOOH transfers onto a neighbouring Ti−OH forming Ti(H2 O)−OOH which adsorbs propene leading to Ti(H2 O)OOH(C3 H6 ). The RDS steps were proposed to be the formation of HOOH on gold and the transformation from Ti(H2 O)OOH(C3 H6 ) to Ti(H2 O)OH(PO). Based on data from Lu et al. [53], Bravo-Suárez et al. [54] performed a model validation of Langmuir– Hinshelwood (L–H) models (the ‘sequential’ or ‘simutaneous’ mechanism; single, two or three adsorption sites), the PRL model and a PRL/L–H hybrid model. They obtained the best fit by the hybrid model, rPO = kPO [H2 ] x [O2 ] y [C3 H6 ]/(c + [C3 H6 ]) [54], where c is a term representing the ratio between HOOH consumed by epoxidation and HOOH that decomposes [53]. Table 1.3: Dependencies in power-rate-law expressions from kinetic studies of different gold catalysts Catalyst Product Au/TS-1 PO Au-Ba/Ti-TUD PO H2 O CO2 b Au/TS-1 c PO rProd = kProd [H2 ] x [O2 ] y [C3 H6 ]z a Ea Notes x y z (kJ/mol) 0.60 ± 0.03 0.31 ± 0.04 0.18 ± 0.04 42 0.02–0.06 wt% Au Si/Ti= 143, 36 170 o C 0.54 ± 0.06 0.24 ± 0.06 0.36 ± 0.06 43 0.11 wt% Au 0.67 ± 0.07 0.16 ± 0.07 0.03 ± 0.07 51 2.4 wt% Ba 0.52 ± 0.14 0.26 ± 0.06 0.07 ± 0.06 80 Si/Ti= 100/3 150 o C 0.53 ± 0.02 0.26 ± 0.02 0.18 ± 0.04 n.d. 0.02 wt% Au 0.46 wt% Ti 180 o C Ref. [52] [53] [55] a. with 95 % confidence interval b. another form is rCO = kCO [H2 ]0.39±0.14 [O2 ]0.21±0.06 [PO]0.22±0.13 2 2 c. performed in an explosive mixture of H2 /O2 up to 40 vol% at atmospheric pressure in a membrane reactor 12 CHAPTER 1. INTRODUCTION The kinetic and DFT study by Barton and Podkolzin[39] for water formation over gold catalysts supported on silica, silicate-1 and TS-1 made a foundation for formulating HOOH formation in the direct propene epoxidation. The power law orders in their study are 0.7–0.8 on hydrogen, 0.1–0.2 on oxygen. The apparent activation energy was 37–41 kJ/mol. Their rate expression on water formation rH2 O = kKOOH KO2 PO2 1+ p p KH2 PH2 KH2 PH2 + KO2 PO2 + KOOH KO2 PO2 p KH2 PH2 PH2 p 1 + KH2 PH2 (1.1) is based on a two-site model described in the following proposed mechanism: O2 + ⋆ ←→ O2 ⋆ (R1.8) H2 + 2 ⋆ ←→ 2 H⋆ (R1.9) H2 + 2 ←→ 2 H O2 ⋆ +H ←→ HOO ⋆ + (R1.10) HOO ⋆ +H2 + −→ HOOH ⋆ +H (RDS) (R1.11) where two types of Au sites are involved, i.e., one for O2 , OOH and HOOH adsorption, one for dissociative adsorption of H2 . The dissociative adsorption of H2 to form HOOH is the rate-determining step. The water formation thus follows the cleavage of O−O bond in HOOH and the hydrogenation of OH. Nijhuis and Weckhuysen [56] also performed a kinetic study on water formation over gold catalysts supported on TiO2 , SiO2 and silicate1. The apparent activation energy in their study is 42–52 kJ/mol. 1.3 Microreactor system It has been shown by Oyama et al. [55] that if one increases the hydrogen and oxygen concentrations in the reactor, one can produce significantly larger quantities of propene oxide (> 10 % yield) without extensive catalyst deactivation. A problem with increasing the hydrogen and/or oxygen concentrations, however, is the explosion limits of this gas combination. The commonly used gas mixture of hydrogen/oxygen/propene of 10 vol% each is just outside of the explosion limits and increasing either the hydrogen or oxygen concentration results in a potentially explosive mixture. In the work of Oyama et al. [55], 1.3. MICROREACTOR SYSTEM 13 hydrogen and oxygen reached a 40 vol% composition via a membrane reactor module, in which hydrogen was gradually fed over the reactor length through inorganic membrane. The fact that the hydrogen conversion was almost 100 %, meant that although one would seemingly be operating within an explosive gas composition, but this was not the case in reality. It is clear, however, that the large increase in hydrogen/oxygen concentration in this membrane reactor system resulted in a dramatic increase in the propene oxide productivity, since the same catalyst in a normal flow reactor system with the typical 10 vol% hydrogen/oxygen/propene feed produced less than 2 % in conversion of propene oxide. Since microreactor technology first emerged as a scientific discipline in the 1990’s, a steady increase in the number of chemical reactions and physical changes that have been successfully performed in such miniature devices could be observed and commercial-scale applications are now available. Because of their inherently high surface-area-to-volume ratios, microreactors demonstrate order of magnitude improvements in heat and mass transfer rates, allowing highly efficient, compact and cost-effective devices to be created to carry out chemical and thermal reactions more safely, and with greater selectivity and conversion rates, higher yields, and improved product quality. In the epoxidation of propene, a significant advantage of the microreactor is its very small reactant inventory. The reactant gases are mixed just outside the microreactor unit, after which they immediately proceed to the multi-channel catalytic section. The small reactant volume makes it possible to work with a gas composition within the explosive region. For gas phase applications, typical reactor channels are 100–200 µm wide/deep and a few cm in length, resulting in a total volume of less than 1 µl. The energy liberated by an explosion in such a channel would be less than 10 mJ, which would not be able to affect the integrity of the microreactor. Aside from the microreactor being ‘explosionproof’, the excellent heat transfer rates of the reactor make it facile to run the reactor isothermally and prevent ranaways, which could be an explosion trigger. The potential of microreactors to operate safely with an explosive hydrogen-oxygen reaction mixtures has been demonstrated by Jensen et al. [57] in a study in which hydrogen peroxide was produced directly. The gold-catalyzed propene epoxidation requires only a relatively short contact time. Even conventional reactors are typically operated at a GHSV of around 10000 h−1 , which makes it a very suitable system for a microreactor. The maximum heat production in the reactor will be 30 W/ml, which can be easily removed from the microreactor system 14 CHAPTER 1. INTRODUCTION [58]. Thus, risks of a runaway are absent. 1.4 Objectives and outline The aim is to have an operational microreactor unit for the epoxidation early in the research program, which will be used for studies to determine the optimal conditions at which the propene epoxidiation will be carried out. For the microreactor, the plan is to use ‘conventional’ packed-bed reactor as depicted by Losey et al. [59], which then will limit the arduous development of a special in-situ catalyst preparation in the early phase. The microreactor will be used for kinetic studies over the entire concentration range (inside and outside of the explosive regions) as well as for studies to determine the catalyst stability. Based on the catalyst performance, further optimization of catalysts will be performed. The original final goal is to have a numbered-up microreactor system combined with an optimized catalyst, which can be connected to a designed micro separation unit, as a demonstration unit for the propene epoxidation with a much larger operational window. However, when the issue of competitive water formation and propene hydrogenation arose, which is embedded in the reaction mechnism, developing such an integrated micro unit became unrealistic. The research then switched to tackle the problems with side reactions and to gain more mechanistic insight of this catalytic reaction. The kinetic and stability studies over a 1 wt% gold catalyst supported on a dispersed Ti-SiO2 support is presented in Chapter 2. The competitive nature of propene epoxidation and water formation on the interfacial Au–Ti sites is revealed. A simplified deactivation model is developed and validated with experimental data. Chapter 3 focuses on catalyst optimization based on the insights gained in Chapter 2. Chapter 4 describes the general behaviour of propene hydrogenation over this catalytic system and provides the method to switch off propene hydrogenation. An infrared study over the gold–titania catalysts is presented in Chapter 5 trying to give second thoughts on the behaviour of gold in the direct epoxidation of propene. Final conclusions and outlook come in Chapter 6. References [1] T. A. Nijhuis, M. Makkee, J. A. Moulijn and B. M. Weckhuysen, Ind. Eng. Chem. Res., 2006, 45, 3447–3459. REFERENCES 15 [2] Merchant Research & Consulting. Ltd., Propylene Oxide (PO): 2013 World Market Outlook and Forecast up to 2018, abstract retrieved from http://mcgroup.co.uk/researches/ propylene-oxide-po, 2013. [3] J. Tsuji, J. Yamamoto, M. Ishino and N. Oku, SUMITOMO KAGAKU, 2006, 2006-I, 1–8. [4] Propylene Oxide, http://www.dow.com/propyleneoxide/app/index.htm, 2013. [5] S. T. Oyama, in Mechanisms in Homogeneous and Heterogeneous Epoxidation Catalysis, ed. S. T. Oyama, Elsevier, 2008, ch. 1, pp. 3–99. [6] G. Thiele (Degussa-Huls AG), US 6372924, 2002. [7] A. Forlin, G. Paparatto and P. Tegon (Dow Global Technologies Inc.), US 7138534, 2006. [8] Propylene oxide the Evonik–Uhde HPPO technology: Innovative, profitable, clean, retrieved from www.thyssenkrupp-uhde.de, 2013. [9] P. Bassler, H.-G. Göbbel and M. Weidenbach, Chem. Eng. Trnas., 2010, 21, 571–576. [10] M. G. Clerici and P. Ingallina, J. Catal., 1993, 140, 71–83. [11] A. Zwijnenburg, M. Makkee and J. A. Moulijn, Appl. Catal., A, 2004, 270, 49–56. [12] E. A. Carter and W. A. Goddard III, Journal of Catalysis, 1988, 112, 80–92. [13] T. Hayashi, K. Tanaka and M. Haruta, J. Catal., 1998, 178, 566–575. [14] P. Blankenstein, M. Crocker, C. J. G. van der Grift and J. J. M. van Vlaanderen (Shell Oil Company), US 7713906, 2010. [15] T. Seo and J. Tsuji (Sumitomo Chemical Company Ltd.), US 6646139, 2003. [16] H. A. Wittcoff, B. G. Reuben and J. S. Plotkin, in Chemicals and Polymers from Propylene, John Wiley & Sons, Inc., 2012, ch. 6, pp. 211–271. [17] E. E. Stangland, K. B. Stavens, R. P. Andres and W. N. Delgass, J. Catal., 2000, 191, 332–347. [18] A. Zwijnenburg, A. Goossens, W. G. Sloof, M. W. J. Crajé, A. M. Van der Kraan, L. J. De Jongh, M. Makkee and J. A. Moulijn, J. Phys. Chem. B, 2002, 106, 9853–9862. [19] S. T. Oyama, J. Gaudet, W. Zhang, D. S. Su and K. K. Bando, ChemCatChem, 2010, 2, 1582– 1586. [20] J. Gaudet, K. Bando, Z. Song, T. Fujitani, W. Zhang, D. Su and S. Oyama, J. Catal., 2011, 280, 40–49. [21] C. Qi, J. Huang, S. Bao, H. Su, T. Akita and M. Haruta, J. Catal., 2011, 281, 12–20. [22] A. K. Sinha, S. Seelan, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2004, 43, 1546– 1548. [23] B. Chowdhury, J. J. Bravo-Suárez, M. Daté, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2006, 45, 412–415. [24] A. M. Joshi, B. Taylor, L. Cumaranatunge, K. T. Thomson and W. N. Delgass, in Mechanisms in Homogeneous and Heterogeneous Epoxidation Catalysis, ed. S. T. Oyama, Elsevier, 2008, ch. 11, pp. 315–338. [25] T. A. Nijhuis, B. J. Huizinga, M. Makkee and J. A. Moulijn, Ind. Eng. Chem. Res., 1999, 38, 16 CHAPTER 1. INTRODUCTION 884–891. [26] P. Vogtel, G. Wegener, M. Weisbeck and G. Wiessmeier (Bayer Aktiengesellschaft), WO 2001041921, 2001. [27] T. A. Nijhuis, T. Visser and B. M. Weckhuysen, J. Phys. Chem. B, 2005, 109, 19309–19319. [28] L. Cumaranatunge and W. N. Delgass, J. Catal., 2005, 232, 38–42. [29] B. Taylor, J. Lauterbach and W. N. Delgass, Appl. Catal., A, 2005, 291, 188–198. [30] B. Taylor, J. Lauterbach and W. N. Delgass, Catal. Today, 2007, 123, 50–58. [31] J. Huang, T. Takei, T. Akita, H. Ohashi and M. Haruta, Appl. Catal., B, 2010, 95, 430–438. [32] M. Du, G. Zhan, X. Yang, H. Wang, W. Lin, Y. Zhou, J. Zhu, L. Lin, J. Huang, D. Sun, L. Jia and Q. Li, J. Catal., 2011, 283, 192–201. [33] W.-S. Lee, M. C. Akatay, E. A. Stach, F. H. Ribeiro and W. N. Delgass, J. Catal., 2012, 287, 178–189. [34] J. J. Bravo-Suárez, K. K. Bando, J. Lu, M. Haruta, T. Fujitani and S. T. Oyama, J. Phys. Chem. C, 2008, 112, 1115–1123. [35] J. K. F. Buijink, J. J. M. van Vlaanderen, M. Crocker and F. G. M. Niele, Catal. Today, 2004, 93-95, 199–204. [36] M. G. Clerici, Oil Gas Eur. Mag., 2006, 32, 77–82. [37] P. P. Olivera, E. M. Patrito and H. Sellers, Surf. Sci., 1994, 313, 25–40. [38] D. H. Wells Jr., W. N. Delgass and K. T. Thomson, J. Catal., 2004, 225, 69–77. [39] D. G. Barton and S. G. Podkolzin, J. Phys. Chem. B, 2005, 109, 2262–2274. [40] C. Sivadinarayana, T. V. Choudhary, L. L. Daemen, J. Eckert and D. W. Goodman, JACS, 2004, 126, 38–39. [41] T. Ishihara, Y. Ohura, S. Yoshida, Y. Hata, H. Nishiguchi and Y. Takita, Appl. Catal., A, 2005, 291, 215–221. [42] J. K. Edwards, B. E. Solsona, P. Landon, A. F. Carley, A. Herzing, C. J. Kiely and G. J. Hutchings, J. Catal., 2005, 236, 69–79. [43] J. K. Edwards, B. Solsona, P. Landon, A. F. Carley, A. Herzing, M. Watanabe, C. J. Kiely and G. J. Hutchings, J. Mater. Chem., 2005, 15, 4595–4600. [44] J. K. Edwards, A. F. Carley, A. A. Herzing, C. Kiely and G. J. Hutchings, Faraday Discuss., 2008, 138, 225–239. [45] D. C. Ford, A. U. Nilekar, Y. Xu and M. Mavrikakis, Surf. Sci., 2010, 604, 1565–1575. [46] A. M. Joshi, W. N. Delgass and K. T. Thomson, J. Phys. Chem. C, 2007, 111, 7841–7844. [47] H. M. Ajo, V. A. Bondzie and C. T. Campbell, Catal. Lett., 2002, 78, 359–368. [48] T. A. Nijhuis, E. Sacaliuc, A. M. Beale, A. M. J. van der Eerden, J. C. Schouten and B. M. Weckhuysen, J. Catal., 2008, 258, 256–264. [49] T. A. Nijhuis, E. Sacaliuc-Parvulescu, N. S. Govender, J. C. Schouten and B. M. Weckhuysen, J. Catal., 2009, 265, 161–169. REFERENCES 17 [50] G. Mul, A. Zwijnenburg, B. Van der Linden, M. Makkee and J. A. Moulijn, J. Catal., 2001, 201, 128–137. [51] A. Ruiz, B. van der Linden, M. Makkee and G. Mul, J. Catal., 2009, 266, 286–290. [52] B. Taylor, J. Lauterbach, G. E. Blau and W. N. Delgass, J. Catal., 2006, 242, 142–152. [53] J. Lu, X. Zhang, J. J. Bravo-Suárez, S. Tsubota, J. Gaudet and S. T. Oyama, Catal. Today, 2007, 123, 189–197. [54] J. J. Bravo-Suárez, J. Lu, C. G. Dallos, T. Fujitani and S. T. Oyama, J. Phys. Chem. C, 2007, 111, 17427–17436. [55] S. T. Oyama, X. Zhang, J. Lu, Y. Gu and T. Fujitani, J. Catal., 2008, 257, 1–4. [56] T. A. Nijhuis and B. M. Weckhuysen, Prepr. Pap.-Am. Chem. Soc., Div. Petr. Chem., 2007, 52, 292–295. [57] T. Inoue, M. A. Schmidt and K. F. Jensen, Ind. Eng. Chem. Res., 2007, 46, 1153–1160. [58] E. V. Rebrov, M. H. J. M. De Croon and J. C. Schouten, Catal. Today, 2001, 69, 183–192. [59] M. W. Losey, M. A. Schmnidt and K. F. Jensen, Ind. Eng. Chem. Res., 2001, 40, 2555–2562. Kinetic study of the direct propene epoxidation over Au/Ti-SiO2 in the explosive regime 2 This chapter is adapted from: T. A. Nijhuis, J. Chen, S. M. A. Kriescher, & J. C. Schouten. The direct epoxidation of propene in the explosive regime in a microreactor – A study into the reaction kinetics, Ind. Eng. Chem. Res., 2010, 49, 10479– 10485. J. Chen, S. J. A. Halin, J. C. Schouten, & T. A. Nijhuis. Kinetic study of propylene epoxidation with H2 and O2 over Au/Ti-SiO 2 in the explosive regime, Faraday Discuss., 2011, 152, 321–336. Abstract A kinetic study of the propene epoxidation with hydrogen and oxygen over a Au/TiSiO2 catalyst has been performed in a wide range of reactant concentrations including the explosive region in a microreactor. The observed rate dependency on the reactants for the epoxidation and the competing direct water formation is discussed in relation to the current mechanistic insights in literature. The formation rate of propene oxide is most dependent on the hydrogen concentration, in which the formation of an active peroxo species on the gold nanoparticles is the rate determining step. The deactivation is mainly caused by consecutive oxidation of propene oxide. Oxygen favours the regeneration of the deactivated catalytic sites. Water formation and propene epoxidation are strongly correlated. Water is formed via two routes: through the active peroxo intermediate responsible for epoxidation and from direct formation without involving this active intermediate. Improving the hydrogen efficiency should distinguish between these two routes of water formation. The active peroxo intermediate in epoxidation is competitively consumed by hydrogenation and epoxidation. The active gold site is blocked during deactivation. 20 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME 2.1 Introduction The direct epoxidation of propene to propene oxide (PO) by using the co-reactants hydrogen and oxygen, demonstrated for the first time by Haruta et al. [1], is highly selective over gold-titania based catalysts and has gained considerable attention in the past decade [2–10]. Although this catalytic reaction has been investigated extensively, the mechanism is still not well understood.The most important steps in the reaction mechanism proposed in literature include the formation of a hydroperoxy species on the gold nanoparticles, a reactive adsorption of propene on titania sites adjacent to gold, and the consecutive oxidation of propene by the hydroperoxy species to form propene oxide [11–13]. The highlights of this catalytic reaction are its exceptionally high selectivity towards propene oxide and the mild operating conditions. However, the main obstacles to its industrial application originate from the stability of the catalyst, the relatively low activity, and the insufficient hydrogen efficiency. Encouraging progress has been made over the past years concerning the three key factors that determine the catalytic performance, i.e. the size (or the morphology) of the gold nanoparticles [10, 14, 15], the state of Ti in the support [3, 8, 16–20], and the preparation method of the catalysts which plays an important role in the selective dispersion of the gold and its interaction with the Ti sites [6, 15, 21– 23]. Very high productivity of propene oxide has been reported by different groups and relatively stable performance has been achieved [5, 6, 8, 10, 15, 17]. However, this is usually at relatively high temperatures (150 – 200 o C), where the hydrogen efficiency is less. While the hydrogen efficiency of a certain catalyst can be mainly attributed to the synergy between the gold nanoparticles and the titanium sites in proximity, promoters [5, 8] and surface modification [24] may additionally depress the hydrogen consumption (or hydroperoxy decomposition) to some extent by adjusting the surface acidity. Further improvement in hydrogen efficiency seems difficult. To evade the issue of hydrogen efficiency in the direct epoxidation of propene, the Au-clusters–C3 H6 /O2 /H2 O system has come into view albeit the low activity and the selectivity [25–27]. The reported H2 efficiency of almost 100 % on a Au/TS-1 catalyst with a very low gold loading [23] reveals the potential in enhancing the H2 efficiency. The highest hydrogen efficiency (47 %) reported recently over the Au/TS-1 catalyst without any post-treatment, which still preserves a high conversion of propene, opens the door to a further step in catalyst development [15]. A kinetic study of the propylene epoxidation using H2 and O2 over a gold–titania 2.1. INTRODUCTION 21 based catalyst is important in revealing the reaction mechanism. However, two aspects are complicating: short-term deactivation due to build-up of carbonate/carboxylate species on active Ti-sites [11, 28, 29] and possible long-term changes in activity mainly caused by sintering of the gold nanoparticles [30, 31]. The former difficulty is reversible by burning the strongly adsorbed species and can be solved in a sense by choosing a relatively stable catalyst where Ti is highly dispersed and by operating at elevated temperatures. The latter situation is irreversible and the effect can be minimized by performing the catalytic testing in a period as short as possible or preparing catalysts which are stable (low Cl content, surface-stabilized Au nanoparticles). Another factor constraining a kinetic study is the explosive nature of the reactant mixture. Most reported studies were performed within the non-explosive region [12, 13]. An important exception is the work performed by the group of Oyama, who reported propene epoxidation in a packed-bed catalytic membrane reactor [32]. Feeding (part of) the hydrogen through a membrane allowed them to work safely with much higher hydrogen and oxygen concentrations (up to 40% each), which gave them a stable propene conversion of 10% at 80% propene oxide selectivity at 453 K. Their results showed fractional orders in a power-rate-law expression similar to what had been published previously [12, 13]. Although corresponding relations between hydrogen consumption and product selectivities as well as the formation rate of propene oxide can be observed among earlier studies [24, 33–35], further quantitative investigations have not been reported. In this work, a microreactor system is utilized to perform the propene epoxidation over a gold on titania–silica catalyst in an extended range of reactant concentrations including the explosive region. Microreactors have excellent potential for carrying out this reaction safely using a gas mixture which would be inside of the explosive region for a number of reasons. First, since microreactors have only a very small volume, the energy liberated from an explosion in such a channel would be less than 1 J, which would not be able to affect the integrity of the microreactor. Second, a microreactor allows for an excellent temperature control, which will dramatically diminish the risk of a runaway in the reactor. Most importantly, the fact that the characteristic length (diameter) of the reactor is smaller than the mean free path of the molecules implies that flame propagation inside the channels is not possible as the molecules transfer their energy to the wall instead of each other [36]. Different from previously published kinetic studies, the formation rate of propene oxide in this work is decoupled from the short-term deactivation, which uses the initial 22 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME activity of the catalyst extrapolated by using a simplified deactivation model based on the mechanism proposed by Nijhuis et al. [11]. In this study, quantitative description of the deactivation in a wider range of reactant concentrations will help to gain further insight into the mechanism of the propene epoxidation in hydrogen and oxygen. The relationship between the catalyst stability, the water formation during the epoxidation, and the propene epoxidation itself is thoroughly examined, which may provide mechanistic implication for improving the hydrogen efficiency in the future development of the catalyst. 2.2 Experimental 2.2.1 Catalyst preparation and characterization A catalyst consisting of 1 wt% gold on titania dispersed on silica was used, prepared in a manner similar to the method described by Nijhuis et al. [28]. A total of 15 g of dry silica support (Davisil 643, Aldrich, 300 m2 /g, pore size 150 Å, pore volume 1.15 cm3 /g) was dispersed in 250 mL of anhydrous isopropanol (Aldrich, 99.5 %) under a nitrogen atmosphere in a glove box. The slurry was stirred for 10 min and afterwards 0.70 mL of tetraethylorthotitanate (TEOT) (Aldrich, 97 %) was added. The amount of TEOT added was determined by calculating a theoretical titania coverage of 5 % monolayer (Ti/Si atom ratio) on the silica surface. The slurry was stirred for 30 min. The isopropanol was slowly evaporated at 318 K and 140 mbar in a rotary evaporator under nitrogen. After the isopropanol was removed, the powder was dried overnight at 353 K and subsequently calcined first at 393 K (5 K/min heating) for 2 h and then at 873 K (10 K/min heating) for 4 h. Gold was deposited on the catalyst support by a deposition–precipitation method using aurochloric acid and ammonia. A total of 10 g of support was dispersed in 100 mL of water. The pH of the slurry was adjusted to 9.5 by dropwise adding ammonia (2.5 wt%). A total of 575 mg of an acidic 30 wt% HAuCl4 solution (Aldrich) was diluted in 20 mL of demineralized water and was added dropwise to the support slurry over a 15 min period. While HAuCl4 solution was being added, the pH was kept at 9.5 using aqueous ammonia. After the addition of the gold solution, the slurry was stirred for one hour. The slurry was filtered and washed 3 times using 200 mL of water. The catalyst was dried overnight at 353 K and calcined first at 393 K (5 K/min heating) for 2 h and afterwards at 23 2.2. EXPERIMENTAL 673 K (10 K/min heating) for 4 h. Drying and calcination of the support and the catalyst were performed under atmospheric pressure in stationary air. Loadings of gold and titanium was determined by inductively coupled plasma atomic emission spectroscopy (ICP–OES). The average size of gold nanoparticles was determined by transmission electron microscopy (TEM). The coordination environment of grafted titanium was analyzed by UV–visible spectroscopy. 2.2.2 Catalyst testing Catalytic tests were performed using a microreactor system. The microreactor consisted of a stainless steel capillary tube (0.9 mm inner diameter), which was loaded with 20 mg of catalyst (sieved, 50 – 60 µm) and was operated at a gas feed rate of 3.33 NmL/min. The oxygen was mixed with the other two reactants shortly before the catalyst bed. Immediately after the catalyst bed, additional helium was added to the gas stream, diluting the gases to a composition outside the explosive region. In this manner, the total volume of a potentially explosive gas mixture was minimized. A normal mili-reactor (quartz, 6 mm inner diameter) was installed together with the microreactor. Both reactors had their own gas feeding section but were located in the same oven. The mili-reactor shared the efflux pipeline and GC with the microreactor. Samples can be simultaneously taken from both reactors via an automated 4-way valve. The mili-reactor was used to confirm the integrity and performance of the microreactor when necessary. Scheme 1: Scheme of the microreactor for kinetic study The catalytic experiments were performed in cycles. Prior to each test the catalyst was (re)activated by heating it to 573 K (10 K/min) in a 10 % oxygen in helium stream for 1 h. After cooling to the desired reaction temperature, the gas feed was switched to the desired composition and a 5 h catalytic test was performed. Experiments were performed at atmospheric pressure at varying feed concentrations ranging from 2 to 80 vol% for hydrogen and oxygen and 2 to 40 vol% for propene, with the remaining part 24 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME being helium. The “standard” feed concentration is 10 vol% for each reactant (hydrogen, oxygen and propene) with helium as balance. In the experiments in which the composition was varied, two reactants were kept at the concentration of 10 vol% with only the concentration of the third reactant being changed. A kinetic study over the full range of reactant concentrations was performed at 388 K, 403 K and 418 K. The long-term stability was examined before the kinetic study and was found to be excellent at 388 K. At the higher temperatures, however, a minor activity loss of 5 – 10 % compared with the initial activity was observed over a period of 20 days (around 50 cycles). In the kinetic study, the catalyst was replaced every week in order to minimize the long-term change in catalyst activity and the activity was monitored by repeating the reaction at the “standard” condition at the end of each series of catalytic cycles. Due to the similarity in trend of the results at different temperatures, the results at 403 K were mainly reported and discussed in detail. Product analyses were performed using a fast Interscience Compact GC system, equipped with a Porabond Q column and a Molsieve 5A column, capable of analyzing all products in 4 min. 2.3 Results and discussion 2.3.1 Performance of microreactor and proof of concept In Figure 2.1, the performance of the two reactors at the same reaction conditions is compared. It can be seen that the experiments performed in the two reactors under identical reaction conditions yielded identical results, and that repeated experiments in the same reactor after a catalyst reactivation also resulted in identical results. There was a concern about the ratio between the capillary diameter and the particle size (ca. 15 for the microreactor), which is relatively small and may lead to potential wall channeling and flow maldistribution [37, 38]. However, the results did not show much difference between the two reactors. Thus we exclude the possibility of wall channeling. The external and internal mass transfer limitations were also excluded. The data points in Figure 2.1 show a small degree of fluctuation, which is due to the valve switching between the two reactors when sampling. The kinetic study in the whole range of reactant concentrations was thus all performed in the microreactor. An experiment at a gas composition of 40 vol% hydrogen, 40 vol% oxygen and 20 vol% propene was performed as a first proof of the microreactor concept. The for- 25 2.3. RESULTS AND DISCUSSION 2 Conversion of C3H6 (%) run1, mili run1, micro run2, mili run2, micro 1.5 1 0.5 0 0 50 100 150 Time (min) 200 250 300 Figure 2.1: Performance of the capilary microreactor and the tubular milli reactor (1 wt % Au on Ti-SiO2 catalyst, gas feed 10 vol % propene, 10 vol % oxygen, 10 vol % hydrogen −1 in helium, 393 K, GHSV 10000 mL·g−1 cat h ) mation rate of PO at the elevated reactant concentrations is compared to the “classic” 10/10/10/70 composition as shown in Figure 2.2. It can be seen that performing this reaction under these explosive conditions in a microreactor system greatly improved the reaction rate of propene oxide by a factor of nearly 4 at the beginning. Although the catalyst experienced a relatively fast deactivation at higher reactant concentrations, the steadystate PO rate in the explosive mixture is still 2.5 times of the rate at the 10/10/10/70 composition. 2.3.2 Catalyst characterization The gold loading is 0.91 wt%, which is close to the target loading of 1 wt%. The titanium content on the catalyst is 1.29 wt%. TEM analysis showed a narrow particle size distribution for the gold nanoparticles, centered at 4.5 nm as seen in Figure 2.3. In the deconvoluted UV-vis spectrum, the bands at 202 and 224 nm are assigned to tetrahedral tetrapodal Ti and tetrahedral tripodal Ti, respectively, while the bands at 262 nm and 292 nm probably best assigned to penta- and hexacoordinated Ti structures from dinuclear Ti species. The grafted Ti is highly dispersed mainly in tetrahedral coordination with minor amount of penta- or hexacoordinatd Ti [39]. No adsorption was found above 300 nm. 26 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME −7 6 x 10 40% H 40% O 20% C H PO formation rate (mol⋅g−1 ⋅s−1) cat 2 2 3 6 10% H 10% O 10% C H 2 2 3 6 4.5 3 1.5 0 0 50 100 150 200 Time (min) 250 300 Figure 2.2: Enhanced PO formation rates in the explosive mixture of H2 , O2 and C3 H6 over a 1 wt% Au/Ti-SiO2 catalyst(5 %monolayer of Ti grafted on SiO2 , 393 K, atmo−1 spheric pressure, GHSV 10000 mL·g−1 cat h ) 2.3.3 Product formation In Figure 2.4, the formation rates of propene oxide and water in a single catalytic cycle at the chosen standard conditions are shown. Both formation rates show a relatively fast decrease within the first two hours. The rates continue to decrease, much slower, in the second half of the catalytic cycle. The performance between 150 min and 270 min is used for calculating the averaged H2 efficiency (defined as rPO /rH2 O ) and C3 H6 conversion for a single reaction cycle, which will be used for evaluating the catalyst performance. Different from the H2 efficiency and C3 H6 conversion, the PO selectivity is unchanged since the very beginning of a reaction cycle. The breakthrough curve (green dash) at the beginning of the reaction cycle shown in Figure 2.4 for the PO formation is due to the strong adsorption of PO on the catalyst surface [40]. This breakthrough is more obvious when the initial activity of PO formation is low, but it may be undetectable when the initial activity is high. The concomitant decrease in the formation rate of water with that of PO along time is commonly observed in literature for gold–titania catalysts [12, 24, 33–35]. The formation rate of water is the overall rate unless specified otherwise. The strong correlation between water formation and propene epoxidation during deactivation indicates that a shared site or a common reactive intermediate exists between these two reactions. Most likely, the deactivating species blocks this shared site or a site producing the common intermediate. Based on the current insights in literature on the mechanisms for both the epoxidation and the 27 2.3. RESULTS AND DISCUSSION 80 1.5 (a) (b) 224 60 1 Abs. Counts 292 0.5 202 0 200 262 40 20 0 300 400 500 Wavelength (nm) 600 1 2 3 4 5 6 7 Particle size (nm) 8 9 10 Figure 2.3: UV–visible spectrum of the support (a) and size distribution of gold nanoparticles (b) water formation [9, 12, 13, 41, 42], it is most likely that a peroxo species is this common intermediate. 2.3.4 Propene epoxidation A simplified model was developed to describe the time-dependent propene oxide formation in the catalytic cycles on the Au/Ti-SiO2 catalyst. This model is based on the reaction mechanism proposed previously by Nijhuis et al. [11, 28] as seen in Figure 1.5. The model is able to accurately describe the propene oxide formation rate in time, using three parameters: • rPO,0 : the extrapolated propene oxide formation rate at t = 0, i.e. the activity the catalyst would have had in the absence of deactivation; • kdeact : the deactivation rate constant; • kreact : the reactivation rate constant. The model is based on the following assumptions: 1. The propene oxide formation rate is proportional to the initial rate rPO,0 (without deactivation) multiplied with the fraction of active sites available on the catalyst (i.e., 1−θd , where θd is the fraction of the deactivated sites), according to Equation 2.1. 2. The catalyst deactivation proceeds via a consecutive reaction of propene oxide (or an intermediate directly correlated to the propene oxide formation rate) adsorbed CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME −7 6 −6 x 10 rPO rH2 O r̂PO x 10 2.4 3 1.2 1.5 0 0 50 100 150 200 Time (min ) 250 2 rH −1 (×)r PO,0 =3.38×10 −7 mol·g −1 cat ·s kdeact =1.31×10 3 mol−1 ·g cat kreact =5.19×10 −5 s−1 O 1.8 rPO (mol⋅g−1 ⋅s−1) cat 4.5 (mol⋅g−1 ⋅s−1) cat 28 0.6 0 300 Figure 2.4: Formation rates of propene oxide and water as a function of time: ◦, propene , fitted rate of propene oxide by Equation 2.5 (1 wt% Au on Ti-SiO2 , oxide; , water; gas feed 10 vol% hydrogen, 10 vol% oxygen, 10 vol% propene in helium, 403 K, GHSV −1 10000 mL·g−1 cat h ). on an active site, forming a deactivated site. The rate of this reaction is according to Equation 2.2. 3. The catalyst reactivation occurs via the desorption of the deactivating species from a deactivated site. The rate of this reaction is according to Equation 2.3. rPO = rPO,0 (1 − θd ) dθd rdeact = = kdeact rPO (1 − θd ) d t deactivating dθd rreact = = −kreact θd d t regenerating (2.1) (2.2) (2.3) The rate of formation of deactivated sites can be determined from a site balance for the deactivated sites: dθd dt = kdeact rPO (1 − θd ) − kreact θd (2.4) 29 2.3. RESULTS AND DISCUSSION By combining equations 1 and 4 and taking the initial conditions of rPO = rPO,0 and θd = 0 at t = 0, the equations can be solved analytically. The analytical solution is provided by Equation 2.5: 1 − exp (−At) rPO = rPO,0 1 − 1 a − exp (−At) a 1 kdeact rPO,0 with A = a − a and kreact kdeact · rPO,0 =a+ 1 a − 2 (a > 1) (2.5a) (2.5b) (2.5c) This 3-parameter model was fitted to the experimental cycles at different reaction conditions. An example of such a fit is shown in Figures 2.4 and 2.5. The model describes the time-dependent experimental results well. The highest error occurs on the fastest deactivating curve (normally at the highest hydrogen concentration, see discussion later), but errors from the experimental data are still within 10 % as shown in Figure 2.5. This model, however, is not yet a reaction mechanism based on the elementary reaction steps, but rather on a simplified mechanism describing the observed reaction rates well with parameters which lump a number of elementary steps. For this reason, the two rate constants are not only dependent on temperature, but also on the concentrations of the three reactants, viz., hydrogen, oxygen, and propene. The dependency of the three parameters on the reactant concentrations is shown in Figure 2.6. 30 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME −7 8 x 10 15 10 80% H2 7 80% H , fitted 0 5% H2 5 5% H2, fitted Residual (%) rPO (mol⋅g−1 ⋅s−1) cat 2 6 4 3 2 −10 −15 (b) 10 0 1 −10 (c) (a) 0 0 60 120 180 Time (min) 240 0 300 60 120 180 Time (min) 240 300 Figure 2.5: Illustration of accuracy of the deactivation model by Equation 2.5: (a) experimental and fitted rates of PO formation at H2 concentrations of 80 vol% and 5 vol%; residuals of the fitted data for (b) 80 vol% H2 and (c) 5 vol% H2 (1 wt% Au on Ti−SiO2 catalyst, gas feed 10 vol% oxygen, 10 vol% propene in helium, 403 K, GHSV 10000 −1 mL·g−1 cat h , in (b) and (c) residuals are given every third point, green lines are drawn for eye) −7 x 10 2400 6 1800 −1 kdeact (mol ⋅ gcat) ⋅s−1) rPO, 0 (mol⋅g−1 cat 8 4 H2 2 C3H6 (a) 0.2 0.4 0.6 Volume fraction 0.8 H2 600 O2 0 0 1200 O2 CH (b) 1 0 0 3 6 0.2 0.4 0.6 Volume fraction 0.8 1 −4 1 x 10 k react (s−1) 0.8 0.6 0.4 H2 0.2 O2 C3H6 (c) 0 0 0.2 0.4 0.6 Volume fraction 0.8 1 Figure 2.6: Initial activity rPO,0 (a), deactivation rate constant kdeact (b), and reactivation rate constant kreact (c) fitted from the deactivation model (Equation 5) at different −1 reactant concentrations (1 wt% Au on Ti-SiO2 , 403 K, GHSV 10000 mL·g−1 cat h , the concentration of one reactant is varied while the other two are fixed at 10 vol%). 2.3. RESULTS AND DISCUSSION 31 At low concentrations, the propene oxide formation rate is dependent on the concentrations of all three reactants, while at higher concentrations, the rate is only proportional to the hydrogen concentration (Figure 2.6a). This indicates that at higher reactant concentrations, only hydrogen is involved in the rate determining step. This observation is in agreement with the most common assumption in literature [2, 11–13] that the formation of peroxo species (the active oxidizing species) on the gold nanoparticles is the rate determining step. For this reaction a dissociative adsorption of hydrogen is required. In Figure 2.6b, it can be seen that the deactivation rate constant increases with the hydrogen concentration, while it decreases with the propene concentration and is only weakly dependent on the oxygen concentration. This is explained by propene oxide (or a precursor thereof) being oxidized further by the same peroxo species as is the active oxidant in the propene epoxidation. If the propene concentration is higher, the concentration of this species on the catalyst is lower, reducing the rate of the unwanted consecutive oxidation. If the hydrogen concentration is higher, this peroxo species is produced at a higher rate. The slightly mitigated deactivation at higher oxygen concentration will be discussed in Section 2.3.7. In Figure 2.6c, it can be seen that the catalyst reactivation rate is only dependent on the oxygen concentration. Nijhuis et al. [28] previously determined that the deactivating species are consecutive oxidation products of propene oxide, which can only leave the surface once they are completely combusted to produce CO2 . Our current results indicate that this consecutive oxidation to produce carbon dioxide is not by the hydroperoxy species, but rather by oxygen. 2.3.5 Water formation and link to epoxidation In Figure 2.4 it can be seen that the propene oxide production and the water formation show a similar pattern. This can be considered as an indication for a strong correlation between the water formation and the propene epoxidation. Since the amount of water produced is far larger than the stoichiometric amount produced from the epoxidation reaction, this similarity in the rate of formation cannot be explained by the fact that water and propene oxide would be the co-products of a single reaction. To examine this correlation further, the water formation rate as a function of the propene oxide formation rate for a number of catalytic cycles is plotted in Figure 2.7. It can be seen that the correlation between the water production and the propene 32 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME −6 −6 x 10 3 x 10 7.5 (a) (c) CH 3 2.5 (mol⋅g−1 ⋅s−1) cat 0.8 0.0 2− 2 0.4 1.5 rH 2 O 0.0 2 H rH 6 e tim 2− (mol⋅g−1 ⋅s−1) cat 3 2 4.5 O 6 1.5 1 0.5 0 0 1 2 r PO 3 4 (mol⋅g−1 ⋅s−1) 5 6 −7 x 10 cat 0 0 1 2 r PO 3 (mol⋅g−1 ⋅s−1) cat 4 5 −7 x 10 −6 3 x 10 (b) 2 8 1.5 0. 02 − 1 0 0 O 0.5 1 2 0. rH 2 O (mol⋅g−1 ⋅s−1) cat 2.5 2 3 −1 −1 rPO (mol⋅gcat⋅s ) 4 5 −7 x 10 Figure 2.7: Formation rate of water as a function of PO formation rate within each catalytic cycle when only the concentration of (a) hydrogen, (b) oxygen, (c) propene is changed while the other two are fixed at 10 vol% (1 wt% Au on Ti-SiO2 , 403 K, GHSV −1 10000 mL·g−1 cat h , balance in helium). oxide production is almost perfectly linear. The intercept of the linear correlation is not at the origin, which indicates that water is produced via two different routes, one correlated to the propene oxide formation, and one which is not. The water formation, which is directly correlated to the propene oxide formation, is tentatively assigned to the water produced out of a peroxo species (i.e. the species which is also responsible for the epoxidation of propene). The water formation which is not correlated to the propene oxide formation, is assigned to the direct water formation from the reaction between hydrogen and oxygen, not being formed via a peroxo species. These two routes for water formation are depicted in Scheme 1, which is resembling the scheme presented previously by Edwards et al. [43] for the direct formation of hydrogen peroxide over Au–Pd catalysts. It should be mentioned that even though the hydroperoxy species is 33 2.3. RESULTS AND DISCUSSION denoted as HOOH in Scheme 1, this species might also be OOH + H adsorbed in some other form [44, 45]. H2 H2 + O2 2 H2 O HOOH H2 O+ 1 2 O2 C3 H6 C3 H6 O+H2 O Scheme 2: Routes for water formation in propene epoxidation using hydrogen and oxygen. To more closely examine both the direct water formation and the water formation from the peroxo species, which is also responsible for the epoxidation, the linear correlation is used to fit each of the data sets according to Equation 2.6: rH2 O = c · rPO + d (2.6) The coefficient c represents the ratio between the change in the overall water formation rate and the loss in the PO formation rate on the active Au–Ti centers. The loss of activity for both water formation and propene expoxidation can only be attributed to a blockage on this active Au–Ti center. Occupation of Ti alone by the deactivating species does not necessarily result in this synchronous decrease in rates in the context of the sequential mechanism, i.e. HOOH forms first on gold and then spills over to Ti, since HOOH can decompose to water without Ti. Either gold is occupied in the deactivation, or the strong metal support interaction [46, 47] is so weakened by the deactivation of Ti sites that the formation of the active peroxo species is interrupted. The constant d represents the direct formation rate of water. Direct water formation can occur at gold sites near Ti via a route that does not involve a peroxo species responsible for epoxidation, but it can also occur at gold sites not in proximity to Ti. 34 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME x 10 −6 3 10 2.5 H2 1.5 c d (mol⋅g−1 ⋅s−1) cat 8 2 6 O2 C3H6 1 4 H 2 O2 2 0.5 (a) 0 0 C3H6 (b) 0.2 0.4 0.6 Volume fraction 0.8 1 0 0 0.2 0.4 0.6 Volume fraction 0.8 1 Figure 2.8: Direct formation rate of water not involved in epoxidation (a) and the ratio between the over all water formation rate and PO formation rate on the active Au–Ti sites (b) at different reactant concentrations (1 wt% Au on Ti-SiO2 , 403 K, GHSV 10000 −1 mL·g−1 cat h , balance in helium). In Figure 2.8, the values of c and d are shown as a function of the feed concentrations of hydrogen, oxygen, and propene. In Figure 2.8a, d is mainly a function of the hydrogen concentration in the feed and not dependent on propene and oxygen. The fact that parameter d is almost first order dependent on the hydrogen concentration and not dependent on the oxygen concentration (except for only mildly at very low oxygen concentration) suggests that the direct water formation is rate limited by either the hydrogen adsorption or the hydrogen dissociation on the gold nanoparticles. Interestingly, at lower oxygen concentrations the value of d seems to originate from a non-zero value, which is different for the situation for hydrogen. One possible explanation is that oxygen adsorption on gold is so much stronger than (dissociative) hydrogen adsorption, that only at much lower oxygen concentrations (beyond our current experimental capabilities) it would become limiting. The effect by propene on the parameter d is trivial, since the parameter d represents the direct water formation out of hydrogen and oxygen. In Figure 2.8b, it can be seen that parameter c is not dependent on the oxygen concentration, but positively dependent on the hydrogen concentration, and inversely dependent on the propene concentration. In Section 2.3.4, it has been discussed that the propene oxide formation is directly proportional to the formation of a peroxo species. The parameter c here is not representing the formation of a peroxo species on the catalyst, but the way this peroxo species is reacting further. Scheme 1 is showing the possible reactions for this peroxo species (for simplicity simply shown as HOOH). 2.3. RESULTS AND DISCUSSION 35 In Figure 2.8b, it can be seen that oxygen does not play a role in the reaction or decomposition of the peroxo species on the gold nanoparticles. This is in line with the expectation that oxygen is not directly involved in the rate determining step. The observed inverse proportionality of c on the propene concentration is in agreement with the reaction route of HOOH as shown in Scheme 1. Parameter c is actually inversely proportional to the efficiency with which the peroxo species is used for the desired epoxidation. A large value of c implies that a lot of water is produced compared to the amount of propene oxide produced. The inverse proportionality of c to the propene concentration can be easily explained by the fact that as the propene concentration is increasing, a larger fraction of the peroxo species will react with propene to produce propene oxide, and less will be lost by decomposition. The positive dependency of c on the hydrogen concentration can be explained by the fact that excessive hydrogen speeds up the loss of the peroxo species by hydrogenation to form two water molecules. This is in agreement with literature, where it is indeed concluded that in the direct hydrogen peroxide formation out of hydrogen and oxygen over gold nanoparticles, the largest loss of peroxide occurs via the hydrogenation of the peroxide produced [43]. As long as the oxidizing intermediate is formed, it will be consumed either by decomposition/hydrogenation to water or by epoxidation to PO and water. This is why even at very low hydrogen concentration the parameter c does not meet at the origin. 2.3.6 Effect of temperature Important information provided by the formation rates of water and propene oxide at different temperatures is the activation energy. Water formation and propene oxidation are two parallel reactions consuming the oxidizing peroxo intermediate originating from OOH. It might be inaccurate to perform a calculation of the activation energy of propene oxide formation separately. On the one hand, propene oxide (or consecutively oxidized species thereof) strongly adsorbs on the surface [3, 28, 48] resulting in an activation energy influenced by adsorption enthalpies and surface coverages. On the other hand, the competitive consumption of a reactive peroxo intermediate by epoxidation and water formation will cause the uneven distribution of this reactive intermediate towards one of these two reactions when the temperature increases, i.e. in r = Aexp(−Ea /RT ) f (C) the f (C) may change when T changes. The latter situation can be observed by a saturation and even a decrease in one of the two reaction rates when the effect of the strong product 36 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME adsorption is not obvious. x 10 −6 −7 5 x 10 (a) 473 K 453 K 445 K 407 K 394 K cat (mol⋅g−1 ⋅s−1) 3 (b) 2 2 HO 2 4 445 K 435 K 425 K 415 K 407 K 394 K 373 K r rPO (mol⋅g−1 ⋅s−1) cat 3 1 1 0 0 60 120 180 Time (min) 240 −12 300 0 0 1 2 −1 −1 rPO (mol⋅gcat⋅s ) 3 4 −7 x 10 2 −14 0 −1 ln −15 r H2 O,θd →1 r PO,0 r C3,0 −16 2.2 2.3 2.4 2.5 −1 1000 ⋅ 1/T (K ) µ 1 ¡ ¢ −1 ln r/mol · gcat · s−1 −13 drH2 O −1 drPO /selectivity ¶ (c) 2.6 −2 2.7 Figure 2.9: Formation rate of propene oxide as a function of time (a) and water formation rate as a function of PO formation rate (b) at different temperatures and (c) Arrhenius plots of the direct water formation rate, initial PO formation rate, modified initial rate of PO, and the ratio between water formation rate and modified PO formation rate at the active Au–Ti sites (1 wt% Au on Ti−SiO2 , gas feed 10 vol% hydrogen, 10 vol% oxygen, −1 10 vol% propene in helium, GHSV 10000 mL·g−1 cat h ). Table 2.1 gives the general performance of the catalyst at different temperatures. The time-on-stream performance at different temperatures is shown in Figures 2.9(a) and 2.9(b). The main side products are isomers of propene oxide, most probably, due to ring opening of propene oxide on acidic Ti4+ sites [3, 5, 48]. Propionaldehyde is primary among these side products, which is consistent with the study on the isomerization of propene oxide by Namuangruk et al. [49]. Since significant formation of CO2 appears at temperatures higher than 453 K, we consider these side products as formed by consecutive reaction of propene oxide at temperatures below (and including) 445 K. The overall formation rate of carbon oxygenates at t = 0 is denoted as rC3,0 . The relation between 37 2.3. RESULTS AND DISCUSSION the activation energy of water formation and that of epoxidation on the Au–Ti site is given by Equation 2.7, which is explained in more detail in Appendix, ln d rH2 O d rPO /Sel. −1 =− obs obs Ea,H O − Ea,PO 2 RT + const. (2.7) where Sel is the selectivity towards propene oxide. The Arrhenius plots of rH2 O,θd →1 (the parameter d in Equation 2.6), rPO,0 , and rC3,0 are presented in Figure 2.9c. The difference in overall activation energy between water formation and epoxidation at the Au–Ti site given by Figure 2.9c is 22±3 kJ/mol, while the overall activation energy of the direct water formation is 51±5 kJ/mol. The activation energy estimated by rPO,0 at the three lowest temperatures is 24 kJ/mol. These numbers are in agreement, i.e., the sum of 22 and 24 is close to 51. This leads to the conclusion that the water formation through Au–Ti is not much different from the water formation through the other Au sites. So for a high hydrogen efficiency, a low reaction temperature is desired. Equation 2.7 should satisfy the precondition that within a short period of time the amount of deactivating species accumulated on the active Au–Ti sites is far less than the amount of propene oxide formed. This should be checked by comparing the turn over frequency of deactivating species formation (TOFd ) and that of propene oxide formation (TOFPO ) based on the total number of Au–Ti sites. TOFd can be easily estimated from Equation 2.4 by kdeact rPO . At the standard condition at 403 K, kdeact rPO is estimated to be at the level of 5 × 10−4 s−1 . The total amount of gold on the catalyst used in this study is 4.6 × 10−5 mol · g−1 cat . The fraction of active gold atoms at corners and edges is roughly estimated to be 0.15 based on the mean particle size 4.5 nm [50, 51]. Assuming that each titanium alkoxide molecule forms three Ti–O–Si bonds on the silica surface, the chance that a gold atom on the edge contacts a titanium atom is therefore 0.15. The estimated TOFPO is at the level of 0.3 s−1 , which means that the deactivation is indeed 3 orders of magnitude lower. Actually, this precondition is expected to be satisfied, otherwise a complete deactivation within few minutes should be observed. Furthermore, since in our deactivation model, the deactivation and subsequent reactivation should produce complete oxidation products, thus we could also predict that the deactivation is much slower than the epoxidation based on the product composition. Stable performance is achieved starting from the temperature 453 K and the CO2 formation rate increases 38 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME significantly above that temperature (Figure 2.9a and Table 2.1). Table 2.1: General performance of 1 wt% Au on Ti–SiO2 in direct propene epoxidation at different temperatures and standard reactant concentrations T (K) 373 394 407 415 425 435 445 453 473 C3 H6 conversion (%) 0.56 0.73 0.91 1.06 1.26 1.45 1.62 2.05 3.13 PO >99 99 94 91 84 77 69 48 17 Selectivity (%) EA PA AC 0 0 0 0.7 0.1 0 1.8 3.8 0.3 2.0 6.2 1.1 2.5 9.8 3.9 3.2 13 5.7 4.2 17 7.7 4.9 28 12 8.4 42 17 CO2 0 0 0 0 0 1.8 1.2 6.9 16 H2 efficiency (%) 20 12 9.1 8.2 6.6 5.2 4.0 2.1 0.6 PO, propene oxide; EA, acetaldehyde; PA, propionaldehyde; AC, acetone Table 2.2: Power-law fitting of kinetic parameters fitted form Equation 2.5 at three temperatures T (K) 388 418 403 f rPO,0 kdeact rPO,0 kdeact rPO,0 kdeact rH O,0 2 d rH2 O /d rPO − 1 rH2 O,θd →1 f = k[H2 ] x [O2 ] y [C3 H6 ]z x y z 0.54 0.35 0.19 0.04 −0.10 −0.24 0.43 0.31 0.34 0.25 −0.11 −0.32 0.46 0.30 0.29 0.20 −0.10 −0.29 0.72 0.25 −0.06 0.26 0.85 0.02 0.05 −0.42 −0.10 R2 0.9835 0.9543 0.9516 0.9543 0.9781 0.9662 0.9922 95% confidence ∆x ∆y ∆z 0.03 0.03 0.05 0.02 0.02 0.03 0.05 0.05 0.07 0.03 0.03 0.05 0.03 0.03 0.05 0.02 0.02 0.03 0.03 0.03 0.04 0.9541 0.9919 0.03 0.03 0.03 0.03 0.05 0.05 2.3.7 Relationship between selectivity, hydrogen efficiency, catalyst stability Figure 2.10 gives the averaged H2 efficiencies and PO selectivities at different reactant concentrations. It can be seen that increasing the H2 concentration causes the H2 efficiency to decrease monotonously, while changing the C3 H6 concentration shows an inverse trend. This phenomenon makes it clear again that H2 and C3 H6 consume the same oxidizing peroxo intermediate competitively as we have discussed in Section 2.3.5. It is 39 2.3. RESULTS AND DISCUSSION interesting that O2 has a slightly positive effect on the H2 efficiency as it does on reducing the deactivation constant kdeact (see Figure 2.6b), since it is not directly involved in the rate determining step of water formation or epoxidation. This slight effect by oxygen on the H2 efficiency can also be observed from the work by Lu et al. [13]. Probably, oxygen may play a role as the stabilizer of excessive hydrogen in the form of OOH and may therefore slightly inhibit the redundant formation of the oxidizing peroxo species, which will subsequently either oxidize PO causing catalyst deactivation or directly decompose to water. 100 20 H2 O2 80 CH PO selectivity (%) H2 efficiency (%) 15 3 6 10 5 60 O2 20 C3H6 (b) (a) 0 0 H2 40 0.2 0.4 0.6 Volume fraction 0.8 1 0 0 0.2 0.4 0.6 Volume fraction 0.8 1 Figure 2.10: Hydrogen efficiency (a) and selectivity to propene oxide (b) at different −1 reactant concentrations (1 wt% Au on Ti−SiO2 , 403 K, GHSV 10000 mL·g−1 cat h , the concentration of one reactant is varied while the other two are fixed at 10 vol%). It can be seen that the PO selecitivity possesses the same trend as the H2 efficiency for different reactants when Figures 2.10b and 2.10a are compared. The PO selectivity is plotted against the H2 efficiency in Figure 2.11a. Interestingly, the trend shown in Figure 2.11a, in which the PO selectivity and hydrogen efficiency increase together, can also be obtained from the study by Huang et al. [15](see Figure 2.12) on their gold catalysts supported on the same TS-1(48), despite the fact that these catalysts have either different pre-treatment methods on the support or different gold loadings. The catalyst in this study has a titanium density of 2.7 × 10−4 mol · g−1 cat , which is far more than the amount of Ti that functions in epoxidation.† Propene oxide can easily react with these Lewis acidic sites and form mainly its isomers. A higher formation rate of water may increase the extent of hydrolysis of Ti–O–Si bonds, which yields a stronger Brønsted acidity on the catalyst surface favouring the ring opening of propene oxide. In Figure 2.11b, the deactivation rate constant kdeact is plotted against the H2 effi- 40 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME ciency at different reactant concentrations. It can be clearly seen that a higher H2 efficiency corresponds to a lower deactivation rate indicating that the deactivation process is closely linked to water formation, or more precisely, to the way how the (excessive) oxidizing peroxo species is used. A catalyst with a higher hydrogen efficiency should have a more stable performance and a better PO selectivity. A very important conclusion that can be drawn from Figure 2.11 is that a stable, selective and hydrogen efficient catalyst should be possible. PO selectivity (%) 100 95 H2 90 O2 CH 3 6 (a) 85 0 5 10 15 H2 efficiency (%) 20 2400 H kdeact (mol−1⋅ gcat) 2 1800 O 2 C3H6 1200 600 (b) 0 0 5 10 15 H efficiency (%) 20 2 Figure 2.11: Selectivity towards propene oxide (a) and deactivation rate constant (b) as a function of hydrogen efficiency at different reactant concentrations (1 wt% Au on Ti−1 SiO2 , 403 K, GHSV 10000 mL·g−1 cat h , the concentration of one reactant is varied while the other two are fixed at 10 vol%). 2.4 Summarizing discussion Deactivation on the catalyst in this study is mainly caused by consecutive oxidation of propene oxide as proposed by Ruiz et al. [29], while the water productivity reflects 41 2.4. SUMMARIZING DISCUSSION 100 Selectivity (%) 80 60 PO CO2 40 20 0 0 10 20 30 40 Hydrogen efficiency (%) 50 Figure 2.12: Relationship between product selectivities and the hydrogen efficiency on Au/TS-1 catalysts (adapted from data in the paper by Huang et al. [15]). the formation rate of the oxidizing peroxo species. The deactivating species in propene epoxidation was found to be carbonate/carboxylate adsorbed on active Ti sites [28]. Building up of carbonates on gold is also the cause of deactivation in CO oxidation on gold catalysts [52–54]. The concurrent decrease in water formation and epoxidation observed in this sutdy indicates that deactivating species block the active Au–Ti sites. The competing roles of propene and hydrogen in consuming the active peroxo intermediate indicate that a moderate hydrogen concentration is preferred for an acceptable propene conversion without much loss in hydrogen efficiency. Including the decomposition of the active peroxo intermediate into the rate expression based on a real mechanism [13] can well explain the saturation of PO formation at higher propene concentrations and the fractional order on propene in the power-rate-law expression, which is close to zero and normally within the range of 0.18–0.35 [12, 13, 32]. Most likely, hydrogen speeds up this decomposition by increasing surface coverage of dissociated hydrogen. This results in a rate expression for PO formation in the following form, rPO = kHOOH θOOH PH2 p 1 + KH2 PH2 kPO PC3 H6 p KH2 PH2 kPO PC3 H6 + kH2 O p 1 + KH2 PH2 (2.8) which has no essential difference from the rate expression proposed by Lu et al. [13], but may explain the generally observed lower order (0.55–0.60) on hydrogen in propene epoxidation than that in hydrogen oxidation (0.7–0.8) on gold–titania catalysts [2, 12, 13, 32, 41]. 42 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME Due to the parallel consumption of the active peroxo species by hydrogen and propene, a higher propene concentration will increase the utilization efficiency of this active intermediate towards propene oxide and suppresses the unwanted water formation, which accordingly enhances the catalyst stability. Higher oxygen concentrations favour the regeneration of the deactivated sites and might stabilize the excessive hydrogen alleviating the deactivation process. The elementary reaction steps happening at the Au–Ti interface can be described by the following reactions: O2 + ⋆ ←→ O2 ⋆ (R2.1) H2 + 2 △ ←→ 2 H△ (R2.2) O2 ⋆ +H△ ←→ HOO ⋆ +△ (R2.3) C3 H6 + HOO ⋆ +H2 + △ −→ PO + H2 O ⋆ +H△ (RDS) (R2.4) −− * PO + PO −) (R2.5) −− * C3 H6 C3 H6 + −) (R2.6) PO + HOO ⋆ +H2 + 2 △ −→ SÎ + H2 O ⋆ +H△ (R2.7) where ‘△’ denotes Au sites for hydrogen dissociation, ‘⋆’ denotes a second Au sites for O2 /OOH adsorption as proposed by Barton and Podkolzin [41], ‘’ denotes Ti in proximity to Au and ‘S’ is the deactivating species. For simplicity, adsorption of C3 H6 on gold and competing formation of water are not listed here. Further improvement in hydrogen efficiency will place a premium on the theoretical investigation into the pathway of water formation on gold. Barton and Podkolzin [41] proposed the HOOH pathway through which the O–O bond cleaves and two hydroxyls form. The study by Ford et al. [45] reinforced this perspective and suggested another energetically competitive pathway on Au(111) facets, in which OOHH is formed and O–O bond scission occurs leaving an oxygen atom on gold. Similar hydrogen-induced OOH dissociation on gold surface is also proposed in hydrogen-promoted CO oxidation [44] and in direct H2 O2 synthesis [55]. Recent DFT study by Li et al. [56] on propene epoxidation in oxygen and water on gold clusters suggested a pathway in which the scission of the oxygen bond in OOH on gold surface is preferred and the oxygen atom left epooxidizes propene. Considering our low activation energy of propene epoxidation and 2.5. CONCLUSIONS 43 the proposal by Joshi et al. [57] that there is an extra energy barrier for HOOH attacking Ti–OH, the reaction route on our catalyst might not be the sequential mechanism involving H2 O2 transfer on Au/titanosilicate catalysts [9]. In general, a lower reaction temperature is preferred in our system for a higher hydrogen efficiency, but this makes catalyst regeneration more difficult. An efficient activation of hydrogen on gold nanoparitles/clusters [58–61] as well as the synergy between Au and Ti is the key issue for a desirable performance. 2.5 Conclusions A kinetic study of propene epoxidation with hydrogen and oxygen over the Au/Ti–SiO2 catalyst has been performed over a wide range of reactant concentrations including the explosive region by utilizing a micro reactor system. Analysis of the dynamic deactivation process at different reactant concentrations showed that the formation rate of propene oxide is most dependent on the hydrogen concentration and that the formation of an active peroxo species on the gold nanoparticles is the rate determining step. Deactivation is mainly caused by the consecutive oxidation of propene oxide (or a precursor thereof). Higher hydrogen concentrations speed up the deactivation by increasingly forming the oxidizing peroxo species. When the propene concentration is higher, the concentration of this oxidizing species is lower by epoxidizing propene to form the desired propene oxide and therefore the deactivation is mitigated. Oxygen favours the regeneration of the deactivated sites. Water formation and epoxidation are strongly correlated. It can be concluded from our results that there are two routes for water formation, i.e. the water formation on the Au–Ti center through the active peroxo intermediate which is also responsible for epoxidation, and the direct water formation not related to epoxidation. Water formation and propene epoxidation on the active Au–Ti sites are two parallel reactions competitively consuming the same active peroxo intermediate. When the hydrogen concentration is higher, the hydrogenation of the peroxo intermediate to form water is more dominant instead of its consumption by epoxidation. Higher propene concentrations are preferred for the efficiency of utilizing this peroxo intermediate to form propene oxide. Oxygen has no influence on the direct water formation and does not affect the ratio of the water productivity and epoxidation rate on Au–Ti centers indicating that OOH is the true in- 44 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME termediate and that its reaction with hydrogen forming the active peroxo species is the rate determining step. Catalyst deactivation is caused by the blockage of the active Au–Ti center. Saturation of propene oxide formation is observed, which can be attributed to the nature of the parallel water formation and propene epoxidation consuming a common intermediate. The activation energy of propene epoxidation is 22 kJ/mol lower than the water formation on the Au-Ti center suggesting a low reaction temperature for propene epoxidation is favoured. A moderate hydrogen concentration combined with high propene and oxygen concentrations is preferred for a desirable performance of the catalyst, i.e. a higher hydrogen efficiency and consequently better stability and selectivity. 2.A Deactivation model Combining Equations 2.1 and 2.4 results in dθd dt = kdeact rPO,0 (1 − θd )2 − kreact θd (2.A.1) Equation 2.A.1 can be rewritten as dθd θd2 − 2+ kreact kdeact rPO,0 = kdeact rPO,0 d t (2.A.2) θd + 1 We denote a(a > 1) as the root of θd2 − 2 + kreact / kdeact rPO,0 θd + 1 = 0; then we have a+ 1 a =2+ kreact kdeact rPO,0 (2.A.3) Analytically solving Equation 2.A.2 results in 1 1 − exp − a − kdeact rPO,0 t a θd = 1 1 kdeact rPO,0 t a − exp − a − a a (2.A.4) Combining Equations 2.A.4 and 2.1 results in the solution Equation 2.5. Initial guesses for the least-squares fitting are based on the following physical meanings of parameters. 45 2.A. DEACTIVATION MODEL 1. The value of 1/a represents the coverage of deactivating species at steady-state. rPO, t=5h The initial guess for 1/a is 1 − ; rPO, t=0 2. Multiplying rPO,0 on both sides of Equation 2.A.1 results in d rPO,0 (1 − θd ) dt = −rPO,0 · kdeact rPO,0 (1 − θd )2 − kreact θd , which can be rewritten as d rPO dt = −kdeact rPO,0 · rPO (1 − θd ) + rPO,0 · kreact θd (2.A.5) Integrating Equation 2.A.5 leads to Z rPO d r = −kdeact rPO,0 rPO,0 Z t rPO (1 − θd )d t + 0 rPO = rPO,0 − kdeact rPO,0 Z Z t rPO,0 kreact θd d t 0 t rPO (1 − θd )d t + 0 Z t rPO,0 kreact θd d t. (2.A.6) 0 When t → 0, θd → 0; equation 2.A.6 can be simplified to rPO = rPO,0 − kdeact rPO,0 Then we plot rPO against Rt 0 Z t rPO d t. 0 rPO d t (the accumulative amount of PO produced by the time t), the slope at t = 0 should be very close to (−kdeact rPO,0 ). 3. The initial guess of rPO,0 is by fitting the data using r = a1 e−b1 t + a2 e−b2 t + c, where a1,2 , b1,2 and c are constants. This is explained in the following section. H2 + O2 + C3 H6 Au OOH⋆ + △ + H△ C3 H6 PO r s b i Figure 2.13: Schematic model of occupied sites on the Au–Ti center Ti 46 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME 1. −− * s s + −) C3 H6 adsorption, desorption 2. s −→ i PO formation 3. i −→ r PO oxidation, site deactivation 4. r −→ r + site regeneration 5. −− *i+ i −) PO desorption, re-adsorption The above reactions represent the transform of adsorbed species on the Ti sites in proximity to Au. The site balance gives: blank site C3 H6 site PO site deactivated site db dt ds dt di dt dr dt = −(k1 + k−5 )b + k−1 s = k1 b − (k−1 + k2 )s = k−5 b + k2 s + k5 i + k4 r − (k3 + k5 )i = k3 i − k4 r ′ where k−5 = k−5 PPO , k1 = k1′ PC3 H6 , k2 and k3 are most relevant to, strictly speaking, the instantaneous concentration of H2 . If the Au–Ti sites are fully occupied by C3 H6 , PO and the deactivating species under reaction conditions, and if the re-adsorption of PO on the active sites is neglectable compared with the adsorption of C3 H6 to the active sites, the following equations can be obtained: di = k2 s − (k3 + k5 )i dt dr = k3 i − k4 r dt s = 1 − i − r. Therefore, di dt dr dt = −(k2 + k3 + k5 )i − k2 r + k2 = k3 i − k4 r. The above two equations are re-written as x′ 1 x′ 2 = −(k2 + k3 + k5 )x 1 − k2 x 2 + k2 = k3 x 1 − k4 x 2 (2.A.7) (2.A.8) 47 2.A. DEACTIVATION MODEL with the initial conditions x 1 (0) = 0 and x 2 (0) = 0. If k2 and k3 are simply treated as a constant, after the Laplace transform one can obtain s x̃ − x (0) 1 1 s x̃ 2 − x 2 (0) = −(k2 + k3 + k5 ) x̃ 1 − k2 x̃ 2 + = k3 x̃ 1 − k4 x̃ 2 k2 s and subsequently (s + k + k + k ) x̃ + k x̃ 2 3 5 1 2 2 k3 x̃ 1 − (s + k4 ) x̃ 2 = = k2 + x 1 (0) s −x 2 (0). Solving the above equations yields (s + k2 + k3 + k5 )(s + k4 ) + k2 k3 x̃ 1 = k2 s + x 1 (0) (s + k4 ) − k2 x 2 (0). Substitution of the initial conditions yields x̃ 1 = k2 (s + k4 ) s(s2 + αs + β ) where α = k2 + k3 + k4 + k5 , β = k4 (k2 + k3 + k5 ) + k2 k3 . The roots of s(s2 + αs + β ) = 0 are s1,2 = 1 2 −α ± p α2 − 4β , s3 = 0. According to the residue theorem, one can obtain x1 = 3 P Res( x̃ 1 est , s j ) j=1 = k2 (s1 + k4 ) s1 (s1 − s2 ) e s1 t + k2 (s2 + k4 ) s2 (s2 − s1 ) e s2 t + k2 k4 s1 s2 (s − s j ) · k2 (s + k4 ) . (s − s1 )(s − s2 )(s − s3 ) Due to the time resolution of GC analysis, it is hard to get information of the very where Res( x̃ 1 est , s j ) = lim s→s j beginning period when the occupation of the clean Au–Ti sites by the formed PO occurs from the observed rates of PO formation. When fitting the time-on-stream rates of PO 48 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME formation, for simplicity one can fit the data only after the peak appears leaving the adsorption of PO on the Ti–Ti sites out of account. This treatment in data fitting requires a modification to the initial conditions, i.e., using x 1 (0) = f0 and x 2 (0) = 0 instead of x 1 (0) = 0 and x 2 (0) = 0, since the breakthrough information is lost/discarded. This is a reasonable assumption when the establishment of adsorption equilibrium of PO on Au–Ti sites is fast and this time is far shorter when compared with the duration of a single catalytic cycle. In another word, one can simply consider that the fitting using the modified initial conditions is valid from t 0 , for example, t 0 = 5 s. Therefore, the time-dependent coverage of PO on the Au–Ti sites is x1 = ( f0 s1 + k2 )(s1 + k4 ) s1 (s1 − s2 ) e s1 t + ( f0 s2 + k2 )(s2 + k4 ) s2 (s2 − s1 ) e s2 t + k2 k4 s1 s2 (2.A.9) Since the re-adsorption of PO on the active Au–Ti sites is neglected, the observed rate of PO formation (the descending part) is rPO = k5 x 1 · NAu−Ti (2.A.10) where NAu−Ti is the total number of Au–Ti sites. Combining equations 2.A.9 and 2.A.10 yields a general form for fitting the observed PO rates, rPO = a1 e−b1 t + a2 e−b2 t + c. (2.A.11) Equation 2.A.11 fits each individual set of experimental data very well. But due to the complexity of the coefficients, it is not easy to get the kinetic information directly. 2.B Activation energy The formation rate of water due to the decomposition or hydrogenation of the active peroxo intermediate, which is also responsible for epoxidation on the active Au–Ti centers, can be given by Equation 2.B.1, Au rH2 O,d = k1 θOOH PH2 θoAu · θHAu · Nactive (2.B.1) where Nactive = NAu−Ti (1 − θd ), NAu−Ti is the total number of Au–Ti sites, and θd is the fraction of deactivated sites. The rH2 O,d here is excluded from the water formed from the 49 2.B. ACTIVATION ENERGY epoxidation, i.e., from H2 + O2 + C3 H6 −→ PO + H2 O. The observed formation rate of propene oxide is given by Equation 2.B.2 with the assumption that one molecule of propene oxide may cause one deactivated Au–Ti center: Au rPO = k2 θOOH PH2 θoAu · θCTi H · Nactive − NAu−Ti · 3 6 dθd dt (2.B.2) deactivating Au The θOOH , θoAu , θHAu , and θCTi H in Equatons 2.B.1 and 2.B.2 are the coverage based on the 3 6 sites that remain active but not on the total Au–Ti sites. When the amount of deactivating species accumulated on the active Au–Ti sites is far less than the amount of propene oxide formed within a short period of time, the second term on the RHS of Equation 2.B.2 can be omitted: Au PH2 θoAu θCTi H · Nactive rPO = k2 θOOH 3 (2.B.3) 6 Since Eaobs = RT 2 (∂ ln r/∂ T )P , the difference between the overall activation energy of water formation and PO formation can be expressed by obs obs Ea,H − Ea,PO O 2 RT 2 rH2 O,d ∂ ln r PO = ∂T (2.B.4) P By integrating Equation 2.B.4, Equation 2.7 in Section 2.3.6 can be obtained. From Equations 2.B.1 and 2.B.3, we have rH2 O,d rPO ln rH2 O,d rPO = ln A1 A2 − = k1 · θHAu (2.B.5) k2 θCTi H 3 Ea,H2 O − Ea,PO RT 6 + ln θHAu − ln θCTi H 3 (2.B.6) 6 When the quasi equilibrium is reached on the active sites, θHAu and θCTi H can be given 3 6 by θHAu and θCTi H = 3 6 = p KH2 PH2 p 1 + KH2 PH2 KC3 H6 PC3 H6 1 + KC3 H6 PC3 H6 + KPO PPO (2.B.7) (2.B.8) 50 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME Thus we have ∂ ln θHAu ∂T = P = and ∂ ln θCTi H 3 ∂T 6 1 2 1 2 = (1 − θ Ti ) C H 3 6 P (1 − θHAu ) (1 − θHAu ) RT ∂T P ∆Hads,H2 ∆Hads,C3 H6 2 ∂ ln KH2 (2.B.9) RT 2 Ti − θPO ∆Hads,PO RT 2 (2.B.10) The difference in apparent activation energy can be described by 1 obs obs Ea,H − Ea,PO = Ea,H2 O − Ea,PO + (1 − θHAu )∆Hads,H2 − O 2 2 Ti (1 − θCTi H )∆Hads,C3 H6 + θPO ∆Hads,PO 3 6 (2.B.11) REFERENCES 51 References [1] T. Hayashi, K. Tanaka and M. Haruta, J. Catal., 1998, 178, 566–575. [2] E. E. Stangland, K. B. Stavens, R. P. Andres and W. N. Delgass, J. Catal., 2000, 191, 332–347. [3] G. Mul, A. Zwijnenburg, B. Van der Linden, M. Makkee and J. A. Moulijn, J. Catal., 2001, 201, 128–137. [4] A. Zwijnenburg, A. Goossens, W. G. Sloof, M. W. J. Crajé, A. M. Van der Kraan, L. J. De Jongh, M. Makkee and J. A. Moulijn, J. Phys. Chem. B, 2002, 106, 9853–9862. [5] B. Chowdhury, J. J. Bravo-Suárez, M. Daté, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2006, 45, 412–415. [6] L. Cumaranatunge and W. N. Delgass, J. Catal., 2005, 232, 38–42. [7] T. A. Nijhuis, T. Visser and B. M. Weckhuysen, Angew. Chem. Int. Ed., 2005, 44, 1115–1118. [8] A. K. Sinha, S. Seelan, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2004, 43, 1546– 1548. [9] J. J. Bravo-Suárez, K. K. Bando, J. Lu, M. Haruta, T. Fujitani and S. T. Oyama, J. Phys. Chem. C, 2008, 112, 1115–1123. [10] B. Taylor, J. Lauterbach and W. N. Delgass, Appl. Catal., A, 2005, 291, 188–198. [11] T. A. Nijhuis and B. M. Weckhuysen, Catal. Today, 2006, 117, 84–89. [12] B. Taylor, J. Lauterbach, G. E. Blau and W. N. Delgass, J. Catal., 2006, 242, 142–152. [13] J. Lu, X. Zhang, J. J. Bravo-Suárez, S. Tsubota, J. Gaudet and S. T. Oyama, Catal. Today, 2007, 123, 189–197. [14] J. Lu, X. Zhang, J. J. Bravo-Suárez, K. K. Bando, T. Fujitani and S. T. Oyama, J. Catal., 2007, 250, 350–359. [15] J. Huang, T. Takei, T. Akita, H. Ohashi and M. Haruta, Appl. Catal., B, 2010, 95, 430–438. [16] T. A. Nijhuis, B. J. Huizinga, M. Makkee and J. A. Moulijn, Ind. Eng. Chem. Res., 1999, 38, 884–891. [17] B. Taylor, J. Lauterbach and W. N. Delgass, Catal. Today, 2007, 123, 50–58. [18] T. Liu, P. Hacarlioglu, S. T. Oyama, M. Luo, X. Pan and J. Lu, J. Catal., 2009, 267, 202–206. [19] H. Yang, D. Tang, X. Lu and Y. Yuan, J. Phys. Chem. C, 2009, 113, 8186–8193. [20] Y. Liu, H. Yu, X. Zhang and J. Suo, Acta. Phys. Chim. Sin., 2010, 26, 1585–1592. [21] E. E. Stangland, B. Taylor, R. P. Andres and W. N. Delgass, J. Phys. Chem. B, 2005, 109, 2321–2330. [22] E. Sacaliuc-Parvulescu, H. Friedrich, R. Palkovits, B. M. Weckhuysen and T. A. Nijhuis, J. Catal., 2008, 259, 43–53. [23] J. Lu, X. Zhang, J. J. Bravo-Suárez, T. Fujitani and S. T. Oyama, Catal. Today, 2009, 147, 186–195. [24] B. S. Uphade, T. Akita, T. Nakamura and M. Haruta, J. Catal., 2002, 209, 331–340. 52 CHAPTER 2. KINETIC STUDY OF THE DIRECT PROPYLENE EPOXIDATION IN THE EXPLOSIVE REGIME [25] M. Ojeda and E. Iglesia, Chem. Commun., 2009, 45, 352–354. [26] J. Huang, T. Akita, J. Faye, T. Fujitani, T. Takei and M. Haruta, Angew. Chem. Int. Ed., 2009, 48, 7862–7866. [27] S. Lee, L. M. Molina, M. J. López, J. A. Alonso, B. Hammer, B. Lee, S. Seifert, R. E. Winans, J. W. Elam, M. J. Pellin and S. Vajda, Angew. Chem. Int. Ed., 2009, 48, 1467–1471. [28] T. A. Nijhuis, T. Visser and B. M. Weckhuysen, J. Phys. Chem. B, 2005, 109, 19309–19319. [29] A. Ruiz, B. van der Linden, M. Makkee and G. Mul, J. Catal., 2009, 266, 286–290. [30] N. Yap, R. P. Andres and W. N. Delgass, J. Catal., 2004, 226, 156–170. [31] S. C. Parker and C. T. Campbell, Top. Catal., 2007, 44, 3–13. [32] S. T. Oyama, X. Zhang, J. Lu, Y. Gu and T. Fujitani, J. Catal., 2008, 257, 1–4. [33] C. Qi, M. Okumura, T. Akita and M. Haruta, Appl. Catal., A, 2004, 263, 19–26. [34] A. K. Sinha, S. Seelan, M. Okumura, T. Akita, S. Tsubota and M. Haruta, J. Phys. Chem. B, 2005, 109, 3956–3965. [35] T. A. Nijhuis, E. Sacaliuc, A. M. Beale, A. M. J. van der Eerden, J. C. Schouten and B. M. Weckhuysen, J. Catal., 2008, 258, 256–264. [36] E. V. Rebrov, M. H. J. M. de Croon and J. C. Schouten, Catal. Today, 2001, 69, 183–192. [37] D. Mehta and M. C. Hawley, Ind. Eng. Chem. Proc. Des. Dev., 1969, 8, 280–282. [38] M. Winterberg and E. Tsotsas, AlChE J., 2000, 46, 1084–1088. [39] J. M. Fraile, J. I. García, J. A. Mayoral and E. Vispe, J. Catal., 2005, 233, 90–99. [40] T. A. Nijhuis, T. Q. Gardner and B. M. Weckhuysen, J. Catal., 2005, 236, 153–163. [41] D. G. Barton and S. G. Podkolzin, J. Phys. Chem. B, 2005, 109, 2262–2274. [42] T. A. Nijhuis, M. Makkee, J. A. Moulijn and B. M. Weckhuysen, Ind. Eng. Chem. Res., 2006, 45, 3447–3459. [43] J. K. Edwards, A. F. Carley, A. A. Herzing, C. J. Kiely and G. J. Hutchings, Faraday Discuss., 2008, 138, 225–239. [44] E. Quinet, L. Piccolo, F. Morfin, P. Avenier, F. Diehl, V. Caps and J.-L. Rousset, J. Catal., 2009, 268, 384–389. [45] D. C. Ford, A. U. Nilekar, Y. Xu and M. Mavrikakis, Surf. Sci., 2010, 604, 1565–1575. [46] D. W. Goodman, Catal. Lett., 2005, 99, 1–4. [47] M. Boronat and A. Corma, Dalton Trans., 2010, 39, 8538–8546. [48] F. Sun and S. Zhong, J. Nat. Gas Chem., 2006, 15, 45–51. [49] S. Namuangruk, P. Khongpracha, P. Pantu and J. Limtrakul, J. Phys. Chem. B, 2006, 110, 25950–25957. [50] A. Carlsson, A. Puig-Molina and T. V. W. Janssens, J. Phys. Chem. B, 2006, 110, 5286–5293. [51] C. J. Weststrate, A. Resta, R. Westerström, E. Lundgren, A. Mikkelsen and J. N. Andersen, J. Phys. Chem. C, 2008, 112, 6900–6906. [52] M. M. Schubert, A. Venugopal, M. J. Kahlich, V. Plzak and R. J. Behm, J. Catal., 2004, 222, REFERENCES 32–40. [53] Y. Denkwitz, B. Schumacher, G. Kuc̆erová and R. J. Behm, J. Catal., 2009, 267, 78–88. [54] F. Gao, T. E. Wood and D. W. Goodman, Catal. Lett., 2010, 134, 9–12. [55] R. Todorovic and R. J. Meyer, Catal. Today, 2011, 160, 242 – 248. [56] C. Chang, Y. Wang and J. Li, Nano Res., 2011, 4, 131–142. [57] A. M. Joshi, W. N. Delgass and K. T. Thomson, J. Phys. Chem. C, 2007, 111, 7841–7844. [58] A. G. Sault, R. J. Madix and C. T. Campbell, Surf. Sci., 1986, 169, 347–356. [59] T. V. Choudhary and D. W. Goodman, Appl. Catal., A, 2005, 291, 32–36. [60] A. Corma, M. Boronat, S. González and F. Illas, Chem. Commun., 2007, 43, 3371–3373. [61] M. Boronat, F. Illas and A. Corma, J. Phys. Chem. A, 2009, 113, 3750–3757. 53 Enhancement of catalyst performance in the direct propene epoxidation: A study into gold–titanium synergy 3 This chapter has been published as: J. Chen, S. J. A. Halin, E. A. Pidko, M. W. G. M. Verhoeven, D. M. Perez Ferrandez, E. J. M. Hensen, J. C. Schouten, & T. A. Nijhuis, Enhancement of catalyst performance in the direct propene epoxidation: A study into gold–titanium synergy, ChemCatChem, 2013, 5, 467–478. Abstract Enhanced productivity toward propene oxide in the direct propene epoxidation with hydrogen and oxygen over gold nanoparticles supported on titanium-grafted silica was achieved by adjusting the gold–titanium synergy. Highly isolated titanium sites were obtained by lowering the titanium loading grafted on silica. The tetrahedrally coordinated titanium sites were found to be favorable for attaining small gold nanoparticles and thus a high dispersion of gold. The improved productivity of propene oxide can be attributed to the increased amount of the interfacial Au–Ti sites. The active hydroperoxy intermediate is competitively consumed by epoxidation and hydrogenation at the Au–Ti interface. A higher propene concentration is favorable for a lower water formation rate and a higher formation rate of propene oxide. Propene hydrogenation, if occurring, can be switched off by a small amount of carbon monoxide. 56 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY 3.1 Introduction The direct epoxidation of propene to propene oxide (PO) using the co-reactants hydrogen and oxygen, demonstrated for the first time by Haruta et al. [1], is highly selective over gold–titania based catalysts and has gained considerable attention in the past fifteen years. This catalytic system requires highly dispersed gold nanoparticles or clusters and tetrahedral-coordinated Ti4+ sites in the support for a high PO productivity [2–6]. The most important steps in the reaction mechanism proposed in the literature include the formation of a hydroperoxy species on the gold nanoparticles, a reactive adsorption of propene on titania sites adjacent to gold, and the consecutive oxidation of propene by the hydroperoxy species to form propene oxide [7, 8]. The rate-determining step in this reaction is the formation of the hydroperoxy species on gold [9–11], which involves the dissociative adsorption of hydrogen, most likely, at low-coordinated gold atoms close to the Au–Ti interface [12–14]. The selectivity to PO from propene is generally high (> 90%), which makes this process attractive to the industry [15, 16]. One important issue hindering this process toward a commercial level is the unwanted hydrogen combustion. Besides the stoichiometric water formed in the epoxidation reaction, a large amount of hydrogen is directly converted into water resulting in a relatively low hydrogen efficiency. Though studies have addressed this problem, for example, by silylating the hydrophilic surface of the support or by introducing promotors and additives [17–20], the answer to an intrinsically higher hydrogen efficiency remains vague in most cases when a high PO yield is the main aim. Our previous study on the kinetics of this catalytic system has shown that the active hydroperoxy species is competitively consumed by epoxidation and hydrogenation at the Au–Ti interface [11]. Water is formed via two routes: through the active hydroperoxy intermediate responsible for expoxidation and from the direct formation not involving this active intermediate. Gold sites not neighbouring to Ti sites are inactive in the epoxidation and only produce water. Thus, in principle, the hydrogen efficiency may be improved by enhancing the synergy between gold and titanium to decrease the proportion of gold sites not adjacent to titanium. On the other hand, a catalyst with the same gold loading but having smaller gold nanoparticles shows a higher rate in hydrogen dissociation [12], which is preferred to speed up the rate-determining step, i.e., the formation of the active hydroperoxy species. Therefore, developing a catalyst with higher PO productivity and hydrogen efficiency should focus on achieving small gold nanoparticles and sufficient 3.2. EXPERIMENTAL 57 Au–Ti interface. In this study, the effect of the synergy between gold and titanium by adjusting metal loadings was investigated in terms of PO productivity and hydrogen efficiency. Titaniumgrafted silica (Ti-SiO2 ) was chosen as the support. Though titanium silicalite-1 (TS-1) is often preferred due to its hydrophobicity leading to less hydrogen loss, the zeolite needs special treatment to anchor gold efficiently [21], and a reproducible preparation method for the catalysts can be an issue [22]. One beneficial point using Ti-SiO2 as the support is that almost all the titanium sites are accessible to gold when the silica is mesoporous. Furthermore, the relatively simple preparation allows for an sufficient amount of support that can be prepared in one batch; thus the effect of gold loading can be easily and fairly compared. In this study, the relation between water formation and PO formation was systematically examined over a wide range of metal loadings and reactant compositions. Propene hydrogenation was encountered over catalyts using supports with short grafting time of titanium. The performance of these catalysts is compared and discussed. 3.2 Experimental 3.2.1 Preparation of supports The Ti-SiO2 supports were prepared by grafting Ti on silica. In a typical synthesis, 15 g of dry as-received silica (Davisil 643, 300 m2 /g, pore size 150 Å, pore volume 1.15 cm3 /g) was dispersed in 250 mL of anhydrous 2-propanol (Aldrich, 99.5 %) under a nitrogen atmosphere in a glove box. The slurry was stirred for 10 minutes and afterwards tetraethylorthotitanate (TEOT, Aldrich, 97 %) was added. The amount of TEOT added was determined by calculating the desired theoretical titania coverage of 1 – 10 % monolayer (Ti/Si atom ratio based on ca. 8 atom Si/nm2 ) on the silica surface. A coverage of 5 % monolayer on 15 g of silica corresponds to 0.7 mL TEOT in total, or 1 wt.% of Ti with respect to the amount of silica. The slurry was stirred for 30 min. The 2-propanol was evaporated at 333 K under vacuum in a rotary evaporator. The solvent was removed within 45 min. The powder was dried overnight at 353 K and subsequently calcined first at 393 K (5 K/min heating) for 2 hours and then at 873 K (10 K/min heating) for 4 hours. The prepared supports are denoted as Ti-SiO2 -x, where x stands for the nominal coverage of x % monolayer of Ti on silica surface. A blank support (Ti-SiO2 -0) was prepared following the above-mentioned procedure but without adding TEOT. 58 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY An additional Ti-SiO2 support with 1 % monolayer coverage of Ti was prepared on the as-received silica (Davisil 643) according to the above described procedure. The only difference in preparation is that the grafting time of titanium alkoxide was prolonged by controlling the evaporation temperature and pressure of 2-propanol (ca. 313 K and 140 mbar, respectively) [11]. The solvent was evaporated mildly in 3 hours. This support is denoted as Ti-SiO2 -1L, where ‘L’ stands for ‘prolonged’. The TS-1 zeolite was prepared following a method similar to the one described by Chen et al. [23]. An aliquot of 0.76 g of titanium butoxide (TBOT, Aldrich, 97 %) was mixed with 19.2 g of tetraehtylorthosilicate (TEOS, Aldrich, 98 %) at 273 K in an ice– water bath and stirred for half an hour. Afterwards, 15 mL of tetrapropylammonium hydroxide (TPAOH, Merck, 40 wt.% aqueous solution), previously mixed with 15 mL of de-ionized water, was added dropwise. The mixture was stirred for another half an hour at 273 K. The solution was then heated in an oil bath to 333 K and was stirred at this temperature for 2 hours. The crystallization was carried out in a PEEK-lined stainless steel autoclave at 448 K for 48 hours. The resulting white powder was separated by centrifugation, washed three times with plenty of de-ionized water, dried at 393 K overnight and calcined at 813 K for 6 hours. 3.2.2 Deposition of Au Gold was deposited on the support by a deposition–precipitation (DP) method using aurochloric acid and ammonia. In a typical synthesis, 2 g of support was dispersed in 100 mL of water. The pH of the slurry was adjusted to 9.5 by dropwise adding ammonia solution(2.5 wt%). The calculated amount of HAuCl4 solution (Aldrich, 30 wt.% aqueous solution) diluted in water (approximately 20 mL in total) was added dropwise within 15 min. The amount of gold solution added to the support was exactly the amount corresponding to the target loading of gold on the catalyst. The catalyst slurry was stirred for 1 hour after adding gold. During the preparation procedure, the pH of the slurry was kept between 9.4 and 9.5 by dropwise adding ammonia from time to time, while the slurry was vigorously stirred. The solid was then filtered (centrifuged for the Au/TS-1 catalyst) and washed 3 times with de-ionized water. The catalyst was dried overnight at 353 K and calcined first at 393 K (5 K/min heating) for 2 hours and afterwards at 673 K (10 K/min heating) for 4 hours. The prepared gold catalysts on the Ti-SiO2 -x supports are denoted as m-Au/Ti-SiO 2 -x, where m represents a nominal content of m wt.% of gold 3.2. EXPERIMENTAL 59 on this catalyst. After preparation, catalysts were stored in sealed amber bottles in an 2o C refrigerator. 3.2.3 Inversed incorporation of Ti onto Au/SiO2 A batch of 10 g Au/SiO2 was prepared following the DP method and calcination procedure described in section 3.2.2. The as-received silica (Davisil 643) was used as the support. The support was dispersed in 250 mL water instead of above-mentioned 100 mL in this batch. The gold loading was nominally 1 wt.%. 4 g of the obtained Au/SiO2 catalyst was dispersed in 250 mL anhydrous 2-propanol (Aldrich, 99.5 %) under nitrogen atmosphere. Thereafter TEOT was added to obtain 5 % or 10 % monolayer coverage of Ti on silica surface. The slurry was stirred in a glove box for 30 minutes. The solvent was evaporated in the rotating evaporator under vacuum at 333 K. The obtained catalysts were dried overnight in an oven at 353 K. No calcination was performed. These two catalysts with inversely-incorporated Ti are denoted as 1.0-Au/SiO2 -Ti-x (x = 5, 10). 3.2.4 Catalytic testing Catalytic tests were performed in a flow setup equipped with a fast Interscience Compact GC system (3 min analysis time) containing a Porabond Q column and a Molsieve 5A column in two separate channels, each with a thermal conductivity detector. 300 mg of catalyst was mounted into the tubular quartz reactor (6 mm inner diameter, 1.5 mm wall thickness). The catalytic performance was tested under the standard reaction condition in a 5-hour reaction cycle, typically with a gas feed rate of 50 mL min−1 in total (GHSV −1 10000 mL·g−1 cat h ), consisting of 10 vol% (each) of hydrogen, oxygen and propene (he- lium balance). Between each 5-hour reaction cycle, the catalyst was regenerated in 10 vol.% oxygen in helium at 573 K for 1 hour. The temperature of the catalyst bed was monitored by a K-type thermocouple tightly attached to the outer wall of the reactor. Carbon monoxide could be introduced with concentrations ranging from 70 ppm to 1 vol.% into the gas feed through an extra mass flow controller. The experiments with ammonia gas in the feed were carried out by passing the gas feed through an evaporator containing diluted ammonia solution at 273.6 K. Besides the 6-mm tube reactor, a microreactor system [11] was used for testing the 0.05-Au/Ti-SiO 2 -1 catalyst at high hydrogen and propene concentrations. The microreactor consisted of a stainless steel capillary tube (0.9 mm inner diameter), which was 60 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY loaded with 20 mg of catalyst and was operated at a gas feed rate of 3.33 NmL/min. The volumetric concentration of hydrogen or propene ranged from 2 vol.% to 80 vol.%, while the oxygen fraction was kept constant at 10 vol.%. The catalysts are generally prepared and tested within one month after the supports are prepared. 3.2.5 Characterization Diffuse reflectance UV–visible (DR UV–vis) spectra were recorded on a Shimadzu UV2401PC spectrometer using BaSO4 as a reference. Transmission electron microscope (TEM) images were recorded with a FEI Tecnai G2 Sphera transmission electron microscope at an acceleration voltage of 200 kV. The catalyst samples for TEM analyses were finely ground and suspended in drops of ethanol. The suspension was then deposited onto the TEM grid. The size distribution of gold nanoparticles was determined for each catalyst by counting all visible gold particles from around 20–40 images (at least 150 gold particles in total) at the magnification of 100 k. Loadings of gold and titanium were determined by inductively coupled plasma optical emission spectrometry (ICP-OES) using a Spectra CirosCCD system. In ICP analyses, gold was dissolved with aqua regia and grafted titanium was etched by 5 mol/L H2 SO4 solution. The H2 SO4 solution containing dissolved titanium was then diluted to 2.9 mol H2 SO4 /L for analyses. The TS-1 zeolite was dissolved in an equal-volume mixture of HNO3 , HCl and HF. The thermogravimetric (TG–DTG) profiles were measured on METTLER TOLEDO TGA/DSC STARe system with 20 mL/min nitrogen or helium flow as protective gas and 40 mL/min oxygen flow as reactive gas. The effluent was monitored by ThermoStarTM mass spectrometry system. 3.3 Results 3.3.1 Characterization of the supports Figure 3.1 shows the DRUV–vis spectra of different supports directly after calcination. The TS-1(Si/Ti ratio of 78, 1.0 wt.% of Ti) sample shows a main adsorption band at 210 nm indicating the tetrahedral coordination environment of Ti in the framework. The sample Ti-SiO2 -1L(see Experimental section) with a Ti loading of 0.22 wt.% has a very similar coordination environment of Ti to the TS-1 sample. The main part of the UV–vis band of Ti on this sample can be assigned to tetrahedral tripodal Ti species [24– 61 3.3. RESULTS 27]. A weak adsorption of Ti-SiO2 -1L at 250 nm may be assigned to the tripodal Ti site coordinated with an extra hydroxyl or water molecule coordinated. The supports Ti-SiO2 -x (x = 1, 2, 5, 10, x% monolayer in terms of Ti/Si on the surface) were prepared with a much shorter grafting time of Ti (see Experimental section). The intention of shortening the grafting time was to obtain more Ti-defects, which may be preferable for stabilizing gold nanoparticles [21, 28]. Compared to Ti-SiO2 -1L, the Ti-SiO2 -1 sample shows a broad adsorption at around 275 nm, which can be assigned to penta-coordinated Ti structures from TiO x moieties [24–27]. As seen from the Ti-SiO2 -x samples, increasing the loading of grafted Ti broadens the adsorption band to longer wavelengths in the region of 250 – 300 nm indicative of higher fraction of penta- and hexacooridnated Ti structures formed. (a) F E: Ti−SiO2−5 D: Ti−SiO2−2 C: Ti−SiO2−1 E Delta abs. B: Ti−SiO2−1L D Abs. F−E E−D D−C C−B B−A (b) F: Ti−SiO2−10 A: TS−1 C B A F−E E−D 0.2 D−C C−B 0.2 200 B−A 300 400 500 Wavelength (nm) 600 200 250 300 350 Wavelength (nm) 400 Figure 3.1: DRUV–vis spectra of the freshly calcined supports (a) and subtracted spectra from each two supports (b). The Ti contents (wt.%) of samples A–F are 1.00, 0.22, 0.27, 0.44, 1.07, and 1.78 respectively, determined by ICP analysis. Figure 3.2 shows the thermogravimetric (TG) and differential thermogravimetric (DTG) profiles of the fresh and the aged (stored in sealed bottle at RT for certain time) Ti-SiO2 supports. The as-received SiO2 is not fully hydroxylated. Estimated from the weight loss between 120 and 800 o C, the surface hydroxyl density of the as-received SiO2 is about 1.8 OH/nm2 , which is close to the literature data for this type of silica [29]. The hydroxylated SiO2 shows a main dehydroxylation peak (DTG, dashed line) above 500 o C as shown in Figure 3.2. Grafting of Ti was performed on the as-received SiO2 in our study. 62 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY x 10 0 100.5 −0.2 100.5 −0.2 100 −0.4 100 −0.4 99.5 −0.6 99.5 −0.6 99 98.5 −1.2 800 98 −0.8 SiO2 hydroxylated 50 120 120 200 400 600 o Temperature ( C) −3 101 x 10 0 100.5 −0.2 100.5 −0.2 100 −0.4 100 −0.4 99.5 −0.6 99.5 −0.6 99 98 DTG (%/s) x 10 0 98.5 −0.8 Ti−SiO2−1 fresh prepared 50 120 120 200 400 600 Temperature (oC) −1 99 98.5 −1.2 800 98 −0.8 Ti−SiO2−1 aged 8 months 50 120 120 200 400 600 Temperature (oC) −1 −1.2 800 −3 x 10 0 101 x 10 0 100.5 −0.2 100.5 −0.2 100 −0.4 100 −0.4 99.5 −0.6 99.5 −0.6 99 98.5 98 DTG (%/s) 101 −0.8 Ti−SiO2−5 fresh prepared 50 120 120 200 400 600 o Temperature ( C) Normalized weight (%) Normalized weight (%) −3 −1 99 98.5 −1.2 800 98 −0.8 Ti−SiO2−5 aged 7 months 50 120 120 200 400 600 o Temperature ( C) −3 −1 −1.2 800 −3 x 10 0 101 x 10 0 100.5 −0.2 100.5 −0.2 100 −0.4 100 −0.4 99.5 −0.6 99.5 −0.6 99 −0.8 Ti−SiO −10 98.5 98 2 fresh prepared 50 120 120 200 400 600 Temperature (oC) −1 −1.2 800 DTG (%/s) 101 Normalized weight (%) Normalized weight (%) −1.2 800 101 Normalized weight (%) Normalized weight (%) −3 −1 DTG (%/s) 50 120 120 200 400 600 o Temperature ( C) −1 DTG (%/s) 98 −0.8 SiO2 as received 99 DTG (%/s) 99 98.5 DTG (%/s) 101 DTG (%/s) −3 x 10 0 Normalized weight (%) Normalized weight (%) −3 101 −0.8 Ti−SiO −10 98.5 98 2 aged 8 months 50 120 120 200 400 600 Temperature (oC) −1 −1.2 800 Figure 3.2: Thermogravimetric (TG, blue solid line) and differential thermogravimetric (DTG, red dashed line) profiles of dehydration and dehydroxylation of SiO2 and Ti-SiO2 x supports ( heated at 10 K/min from 323 K to 1073 K, temperature stabilized at 393 K for 10 min; the hydroxylated SiO2 was obtained by hydrolyzing the as-received SiO2 in 373 K water for 48 h in an autoclave.) The Ti-SiO2 supports after calcination were slowly hydroxylated when stored sealed under ambient conditions. In general, the peak from the DTG curve at around 280 o C 63 3.3. RESULTS may be attributed to the hydrolyzed Si–O–Si bond in the aged supports, while the shoulder or peak at around 170 o C for the aged Ti-SiO2 -5 and Ti-SiO2 -10 is best attributed to the hydrolyzed Ti–O–Ti or Ti–O–Si bond. The support Ti-SiO2 -10 shows a much higher affinity to moisture, very likely due to an higher amount of Lewis acidic sites as well as defective Ti–OH on the surface. The re-hydroxylation of supports increases the coordination number of Ti even after re-calcination at 300 o C as indicated by the UV-vis spectra (not shown). 3.3.2 Size of Au particles The titanium loading affects both the coordination state of Ti as seen from Figure 3.1 and the size distribution of gold nanoparticles. Figure 3.3 summarizes the averaged gold particle sizes of catalysts with different gold and titanium loadings. From this summary it can be seen that the measured particle sizes are between 1 and 4 nm and that lower gold and titanium loadings generally lead to smaller gold nanoparticles. Detailed information on the size distribution and metal loadings determined by ICP are provided in the Appendix. bel ow d lim etect ion it Figure 3.3: A visual summary of the averaged sizes of gold nanoparticles on Ti-SiO2 -x (x = 1, 2, 5, medians are used for x = 1 due to skewness) Figure 3.4(e) shows the average sizes of gold nanoparticles with gold loading of 1 wt.% over the Ti-SiO2 -x (x = 1, 2, 5, 10) supports, which are 2.5 ± 0.9 nm, 2.9 ± 1.1 nm, 3.5±1.1 nm, and 6.0±1.4 nm respectively. As seen from these four samples, the average size of gold nanoparticles decreases when the titanium loading is lower. Figure 3.4(a–d) shows representative TEM images of these four samples. 64 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY Au 1 wt.% Ti 10ML.% Au 1 wt.% Ti 5ML.% 30 (e) Au 1wt.% Ti 10ML.% 20 10 0 40 Au 1wt.% Ti 5ML.% 20 nm (a) Au 1 wt.% Ti 2ML.% 20 nm (b) Au 1 wt.% Ti 1ML.% Counts 20 0 60 Au 1wt.% Ti 2ML.% 40 20 0 80 Au 1wt.% Ti 1ML.% 60 40 20 20 nm (c) 20 nm (d) 0 0 1 2 3 4 5 6 7 8 9 10 Particle size (nm) Figure 3.4: TEM images of catalysts with the gold loading of nominal 1 wt.% on the Ti-SiO2 -x supports and their size distributions of gold nanoparticles (In (e), the numberaveraged particle size and its ‘standard deviation’ are shown by a circle with an error bar; the median size is illustrated by the stem.) Figure 3.5(a–d) shows representative TEM pictures of catalysts with lower gold loadings. Figure 3.5(e) compares the size distribution of gold nanoparticles on the Ti-SiO2 -1 support. As seen from Figure 3.5(e), lower gold loadings lead to smaller gold particles. The particle sizes of catalysts with gold loadings lower than 0.2 wt.% were difficult to measure, because very few particles can be observed and our TEM is unable to detect subnanometer particles. Figure 3.6 shows TEM images of the 0.05-Au/TS-1 sample. The average crystal size is ca. 120 nm. Gold nanoparticles can rarely be observed on TS-1 and the size of those observed gold particles is around 3 nm. 3.3.3 PO formation and water formation The time-on-stream productivity of PO over the catalysts is generally stable at 433 K as shown in Figure 3.7(a). Compared to 0.5-Au/Ti-SiO 2 -5, the 0.5-Au/Ti-SiO 2 -1 and 0.5-Au/Ti-SiO 2 -2 catalysts showed a much higher activity. In general, the catalysts on supports with lower Ti loadings, which have a higher fraction of isolated Ti sites as 65 3.3. RESULTS Au 0.2wt.% Ti 5ML.% Au 0.2wt.% Ti 2ML.% 100 (e) Au 1wt.% Ti 1ML.% 50 (a) Au 0.5wt.% Ti 1ML.% 20 nm (b) Au 0.2wt.% Ti 1ML.% 0 120 Counts 20 nm Au 0.5wt.% Ti 1ML.% 80 40 0 90 Au 0.2wt.% Ti 1ML.% 60 30 20 nm (c) 20 nm (d) 0 0 1 2 3 4 5 6 7 8 Particle size (nm) 9 10 Figure 3.5: (a–d)TEM images of catalysts with lower gold loadings and (e) size distributions of gold nanoparticles on the Ti-SiO2 -1 support (the number-averaged particle size and its ‘standard deviation’ are shown by a circle with an error bar; the median size is illustrated by the stem) indicated by UV–vis spectra, gave a higher PO productivity as summarized by Figure 3.7(b). The 1.0-Au/Ti-SiO 2 -5 catalyst (ca. 1 wt.% of Au and Ti) gave its highest PO −1 rate at around 30 gPO · kg−1 at 433 K. In the series of catalysts over the Ti-SiO2 -5 cat h support, the PO formation rate decreased as the gold loading was lowered, while the PO rate reaches the maximum over the 0.5-Au/Ti-SiO 2 -Ti-2 catalyst on the Ti-SiO2 -2 support. Detailed performance and elemental analyses of all the catalysts at different temperatures are provided in the supplemental information. −1 PO formation rates higher than 100 gPO · kg−1 were achieved over the catalysts cat h using the Ti-SiO2 -1 support. Maximum PO formation rates obtained on the five catalysts over Ti-SiO2 -1 are plotted in Figure 3.8(a). What should be mentioned here is that these best rates for each catalyst were obtained at different tempratures ranging from 150 o C to 210 o C (Table 3.1). The highest PO rates are from the catalysts with gold loadings not higher than 0.2 wt.%. Suprisingly, the 0.05-Au/Ti-SiO 2 -1 catalyst with a gold loading as low as 0.05 wt.% also showed a very high activity, though no visible gold particles could be observed by TEM. The corresponding PO productivity per gram of Au is plotted in 66 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY 20 nm Figure 3.6: TEM images of the 0.05-Au/TS-1 sample −1 Figure 3.8(b) for these five catalysts. The activity per unit weight of gold (gPO · g−1 Au h ) increases significantly towards the low Au loadings. Table 3.1: Maximum PO formation rates obtained on the Au/Ti−SiO2 catalysts with different gold loadings over the 1% monolayera Ti−SiO2 and the corresponding selectivities Loading (Au wt%) 0.05 0.09 0.20 0.48 0.84 T (K) 483 465 457 437 424 rPO,max c h−1 ) (g · kg−1 cat 113 (226) 122 (135) 131 (67) 85 (18) 60 (10) Selectivity(%) PO C3 H8 26 69 23 73 38 53 45 40 67 7 H2 efficiency b (%) 7.8 8.1 5.4 4.0 2.2 a. 0.27 wt% Ti as determined by induced coupled plasma (ICP) analysis b. determined as rPO /(rH 2 O + rC H 3 8 ) −1 c. numbers in parentheses are in terms of gPO · g−1 Au h The formation rates of water and PO are correlated in the catalytic reaction. In general, they decrease concurrently when the active Au–Ti sites are blocked during the catalyst deactivation [11]. Since the amount of water produced per amount of PO is a determining factor for the viability of a process based on these catalysts, plotting the water formation rate against the PO rate for each single catalytic test is an informative 67 3.3. RESULTS −7 x 10 0.5−Au/Ti−SiO −1 (a) 2 (b) 0.5−Au/Ti−SiO2−2 5 0.5−Au/Ti−SiO −5 −1 −1 rPO (mol⋅gcat⋅s ) 2 4 3 2 1 0 0 60 120 180 Time (min) 240 300 Figure 3.7: (a) Time-on-stream formation rate of PO during a 5-hour catalytic test for three Au/Ti-SiO 2 catalysts: 0.5 wt.% Au, different Ti loadings; (b) summary of PO productivity for different Au/Ti-SiO2 catalysts. (H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, 433 K, GHSV −1 −7 −1 −1 −1 10000 mL·g−1 = 20.9 gPO · kg−1 cat h ; under this condition, 1 × 10 mol PO · gcat s cat h ; productivity averaged from 150 min to 270 min) method to compare the performance of catalysts. Figure 3.9(a) shows the trajectories of reaction rates for several catalysts in a single 5-hour cycle. The length of the trajectory reflects the short-term stability of a catalyst. The trajectory starts at the highest reaction rates of a catalyst and ends at the lowest rates after 5-hour deactivation. The hydrogen efficiency in terms of rPO /rH2 O can be easily calculated for each point. The slope of a trajectory represents the ratio of hydrogen consumed by water formation and by epoxidation at the Au–Ti interface when the blockage of the active site is the main cause of short-term activity change [11]. The experiments in this figure were performed at 3 different temperatures. On the left part of the diagram as shown in Figure 3.9(a), the activity of the catalysts with 1 wt.% gold loading is compared at 403 K, a relatively low temperature at which the deactivation is obvious. The 1.0-Au/Ti-SiO 2 -1 and 1.0-Au/TiSiO2 -2 catalysts show much higher rH2 O /rPO ratios when compared to the catalysts with higher Ti loadings. There are two possible reasons: 1) the smaller gold particles on the 1.0-Au/Ti-SiO 2 -1 and 1.0-Au/Ti-SiO 2 -2 catalysts have higher hydrogen dissociation rates while the epoxidation rate is relatively limited at the Au–Ti interface; 2) less likely, the gold sites not adjacent to Ti but producing water may also be blocked. The comparison of 0.2-Au/Ti-SiO 2 -2 and 0.5-Au/Ti-SiO 2 -2 at 433 K shows that a higher gold loading result in unwanted water formation since these two catalysts have the same support and similar gold sizes. By adjusting the Au and Ti loadings, the synergy between the two sites can be tuned. As seen from Figure 3.9(a), a much higher PO productivity can be achieved on 68 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY 150 250 Au efficiency (g ⋅g ⋅h ) −1 −1 PO Au 120 −1 −1 rPO, max (gPO⋅kgcat⋅h ) m−Au/Ti−SiO2−1 90 60 30 0.1 0.2 2 200 150 100 50 (a) 0 0 1.0−Au/Ti−SiO −5 (b) 0.5 Au loading (wt.%) 0.8 1 0 0 0.1 0.2 0.5 Au loading (wt.%) 0.8 1 Figure 3.8: (a)Maximum PO formation rates obtained on the Au/Ti-SiO2 -1 catalysts and (b) corresponding gold efficiency. (data see Table 3.1; 1.0-Au/Ti-SiO 2 -5 used as a refer−1 ence has a maximum PO rate of 31 gPO · kg−1 at 433 K and its actual gold loading is cat h 0.80wt.%; H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, GHSV 10000 mL·g−1 h−1 ; performance avercat aged from 150 min to 270 min in each 5-hour catalytic test) the 0.05-Au/Ti-SiO 2 -1 catalyst even at 483 K without increasing the rH2 O /rPO ratio much if compared with the performance of 1.0-Au/Ti-SiO 2 -5 at 403 K. In Chapter 2, it has been found that the active hydroperoxy species is competitively consumed by hydrogenation and epoxidaion at the Au–Ti interface. Figure 3.9(b) confirms this finding on the 0.05-Au/Ti-SiO 2 -1 catalyst, which showed no visible gold particles but having very high epoxidation activity. When the hydrogen concentration was raised, both formation rates of water and PO increased. However, hydrogenation of the hydroperoxy species is more dominant at higher hydrogen concentrations leading to further loss in the hydrogen efficiency. On the other hand, higher propene concentrations are favourable for suppressing water formation as demonstrated in Figure 3.9(b). The activity of the 0.05-Au/Ti-SiO 2 -1 catalyst was stable in each 5-hour test resulting in a compact cluster of points in the water–PO plot. The activity at the same condition (10 vol% for each reactant) is not overlapping in the two series of H2 and C3 H6 experiments as shown in Figure 3.9(b), which means that a slight long term change of the catalyst had occurred because of the gold sintering and/or the support hydroxylation. The results presented in Figure 3.9 indicate that the composition of catalysts (Au/Ti contents) and reactant concentrations should be optimized together for a better catalytic performance. 69 3.3. RESULTS x 10 8 6 (a) 433 K 403 K 483 K (b) 2% 20% 3% 1 (mol⋅g ⋅s ) 5% 6 Pr 1 cat 0.5-Au/Ti-SiO2-2 op en 1.0-Au/Ti-SiO2-1 8% e(+ 10% ) 8% 0.05-Au/Ti-SiO2-1 4 r 2 HO 5% 1.0-Au/Ti-SiO2-2 3% 0.2-Au/Ti-SiO2-2 2 2% Hy 1.0-Au/Ti-SiO2-5 dr e og n( ) 20% 30% 40% 50% 80% 1.0-Au/Ti-SiO2-10 0 0 1 2 r PO 3 4 1 1 (mol⋅g ⋅s ) 5 cat 6 7 x 10 Figure 3.9: (a) Water vs. PO formation rates for different catalysts at selected tempera−1 tures in a 5-hour catalytic test (H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, GHSV 10000 mL·g−1 cat h ); and (b) competitive effect of hydrogen and propene on water and PO formation over 0.05-Au/Ti-SiO 2 -1 at 483 K (varied hydrogen or propene concentrations, 10 vol% oxy−1 gen fixed, helium balance, GHSV 10000 mL·g−1 cat h , tested in the microreactor system, 5-hour test for each condition; the series of experiments changing C3 H6 concentrations were performed after the series of H2 ). 1.0−Au/SiO 2 60 40 Counts 20 0 1.0−Au/SiO −Ti−5 40 2 30 20 10 * 0 0 2 4 6 8 10 12 Particle size (nm) 14 Figure 3.10: Gold particle size distribution of the 1 wt.% Au/SiO2 and the reverselygrafted 1 wt.% Au/SiO2 -Ti-5 catalyst (∗: particles bigger than 14 nm are grouped in the column of 14 nm) 70 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY 3.3.4 Performance of the catalyts with inversely-grafted Ti In Figure 3.9(a), the activity of the catalysts with 1 wt.% of gold was compared. These catalysts have different gold particle sizes (Figure 3.4) in addition to different Ti loadings. In order to exclude the influence from the gold particle size, the method with which Ti was grafted onto Au/SiO2 was applied (see Experimental section). The resulting catalysts are denoted as 1.0-Au/SiO 2 -Ti-x. The 1 wt.% Au/SiO2 has an average gold particle size of 3.1 ± 0.8 nm. However, the inverse-grafting method was not successful to obtain an identical size distribution of gold particles. The resulting 1.0-Au/SiO 2 -Ti-5 catalyst has an average gold size of 4.8 ± 2.7 nm as shown in Figure 3.10. A large number of big gold particles was formed concluding that during Ti depostion gold particle size changed. Additionally, the obtained 1.0-Au/SiO 2 -Ti-x (x = 5, 10) catalysts both showed very high activity in propene hydrogenation (7% – 10% in propane yield), while the 1 wt.% Au/SiO2 had no activity in propene hydrogenation and negligible conversion (less than 0.02%) to PO at the same conditions given in Table 3.2. 3.3.5 Effect of supports The catalyst made on a support with grafted Ti can easily achieve high PO productivity −1 (greater than 100 gPO · kg−1 cat h ) at 473 K as listed in Table 3.3. The rPO /rH2 O ratio on these catalysts at 473 K ranges between 10% and 20% at the standard reaction conditions with 10% of each reactant. The gold loadings of these highly active catalysts are generally as low as 0.2 wt.%. The Ti loadings of these supports are lower than 0.3 wt.%. If the specific surface area of the silica chosen in this study is taken into account (300 m2 /g), the surface density of Ti is around 0.1 Ti/nm2 . Although the TS-1 zeolite used in this study has a much higher Ti loading than the Ti-SiO2 -1 and Ti-SiO2 -1L supports, the activity of 0.05-Au/TS-1 is very limited. The TEM images of the Au/TS-1 catalyst are provided in the supplemental information. Despite the difference in the UV–vis spectra, the most interesting difference between Ti-SiO2 -1 and Ti-SiO2 -1L is that a substantive amount of propane was formed over the catalysts on the Ti-SiO2 -1 support as shown in Table 3.3. The Ti-SiO2 -1L support was prepared with longer grafting time for Ti (3 hours instead of less than 45 min for TiSiO2 -x). The activity in propene hydrogenation over the Au/Ti-SiO2 -1 catalysts is more than two magnitudes higher than the Au/Ti-SiO 2 -1L catalysts. The 0.05-Au/TS-1 catalyst also has a low activity in propene hydrogenation. 3.3. RESULTS Table 3.2: Performance of the catalysts with inversely-grafted Ti at 433 K and the comparison with the catalyts prepared in the normal sequence a Sample ID b Ti loading (wt.%) Au loading (wt.%) Au particle size c (nm) 1.02 1.85 1.07 1.78 0.95 0.95 0.80 0.83 4.8 ± 2.7 (4.0) -d 3.5 ± 1.1 (3.5) 6.0 ± 1.4 (5.9) 1.0-Au/SiO2 -Ti-5 1.0-Au/SiO2 -Ti-10 1.0-Au/Ti-SiO2 -5 1.0-Au/Ti-SiO2 -10 Formation rate (×10−7 mol · g−1 s−1 ) cat PO C 3 H8 H2 O 0.38 11.89 19.74 0.28 9.05 20.88 1.47 0.45 28.95 0.73 8.44 14.40 PO yield (%) 0.31 0.23 1.18 0.59 Selectivity (%) PO C3 H8 3.1 95.2 2.9 93.0 64.1 19.5 7.7 87.9 H2 efficiency e (%) 1.2 0.9 5.0 4.9 −1 a. performance averaged from 150 min to 270 min in each 5-hour catalytic test; H2 : O2 : C3 H6 : He= 1 : 1 : 1 : 7, GHSV 10000 mL·g−1 cat h b. the catalysts with inversely-grafted Ti are denoted as 1.0-Au/SiO2 -Ti-x (x = 5, 10) c. median in brackets d. not determined e. determined as rPO /(rH 2 O + rC H 3 8 ) 71 72 Sample ID Ti loading (wt.%) Au loading (wt.%) 0.22 0.22d 0.22d 0.27 0.27e 0.27e 1.00 0.20 0.11 0.06 0.20 0.09 0.05 0.04 0.2-Au/Ti-SiO2 -1L 0.1-Au/Ti-SiO2 -1L 0.05-Au/Ti-SiO2 -1L 0.2-Au/Ti-SiO2 -1 0.1-Au/Ti-SiO2 -1 0.05-Au/Ti-SiO2 -1 0.05-Au/TS-1 Formation rate (×10−7 mol · g−1 s−1 ) cat PO C3 H8 H2 O 5.86 0.14 61.20 5.68 0.19 43.76 3.03 0.14 14.56 5.44 4.59 113.1 b 5.47 15.23 60.46 5.36 17.21 47.14 0.72 0.25 8.49 PO yield (%) 4.73 4.58 2.44 4.66 4.41 4.32 0.58 Selectivity (%) PO C3 H8 82.8 1.9 84.8 2.9 89.4 4.1 40.2 33.9 24.7 68.8 23.1 74.1 59.6 20.8 H2 efficiency c (%) CO2 4.9 0.0 0.0 9.4 1.9 0.2 0.0 9.6 12.9 20.6 7.2 8.3 8.2 −1 a. performance averaged from 150 min to 270 min in each 5-hour catalytic test; H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, GHSV 10000 mL·g−1 cat h , −1 −1 under this condition, 1 × 10−7 mol PO · g−1 = 20.9 gPO · kg−1 cat s cat h b. Hydrogen was fully combusted. c. determined as rPO /(rH 2 O + rC H 3 8 ) d. same support as 0.2-Au/Ti-SiO 2 -1L, Ti content assumed to be the same e. same support as 0.2-Au/Ti-SiO2 -1, Ti content assumed to be the same CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY Table 3.3: Performance of catalyts over Ti-SiO2 -1L, Ti-SiO2 -1 and TS-1 at 473 K a 3.4. DISCUSSION 73 3.3.6 Propane formation and its suppression Significant propane formation was encountered sepecially for the low-gold-loaded catalysts on the Ti-SiO2 -x supports in this study. Figure 3.11(a) summarizes the rC3 H8 /rPO ratio at 433 K over these catalysts, which ranges from 0.1 to 10. The catalysts with lower gold loadings show a higher selectivity towards propane. In this study, the first catalytic test for each catalyst produced less propane than the consecutive cycles. Therefore, the first catalytic cyle was not used when comparing the performance. The first 2-hour testing for the fresh 0.2-Au/Ti-SiO 2 -2 catalyst as shown in Figure 3.11(b) demonstrates such instability of propane formation. Such ascending trend in propene hydrogenation is general for the fresh catalysts in our study and will be described in detail in Chapter 4. A small amount of carbon monoxide in the feed can completely shut off propene hydrogenation while not affecting PO formation. In Figure 3.11(b), the inhibiting effect on propene hydrogenation by carbon monoxide is shown. After removing carbon monoxide, the activity in propene hydrogenation was restored. The propane formation can also be suppressed by ammonia. However, all the three reactions were inhibited in presence of ammonia as seen from Figure 3.11(b). After removing ammonia from the feed gas, only a portion of its original activity was restored. Detailed experimental results on propene hydrogenation will be given in Chapter 4. 3.4 Discussion 3.4.1 Role of the support in enhanced PO productivity The highest PO formation rates reported in literature using gold catalysts in the direct epoxidation of propene using hydrogen and oxygen are generally at the level of 120 −1 – 150 gPO · kg−1 [4–6, 18, 21, 22, 30], despite some differences in space velocity cat h under the testing conditions. These rates almost meet the industrial requirement for PO productivity [18]. Most of these records were obtained on catalysts using TS-1 as the support. The isolated tetrahedral-coordinated Ti4+ cations abundant in TS-1 make this material an ideal support for the hydro-epxoidation of propene. The TS-1 zeolite active in selective oxidation normally preserves a small crystal size (< 400 nm) facilitating to the use of the exterior surface [25, 31], especially in the case of hydro-epoxidation of propene when the gold nanoparticles are bigger than the micropores. The chemical environment 74 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY (a) −6 −7 x 10 2.5 PO C3H8 4 H2O 2 remove NH3 3 1.5 2 1 rH 2 H rPO or rC (mol⋅g−1 ⋅s−1) cat 1000 ppm NH3 3 8 (mol⋅g−1 ⋅s−1) cat 200 ppm CO 500 ppm CO O 10 5 1 0.5 remove CO (b) 0 0 2 4 Time (h) 6 0 8 Figure 3.11: (a) Ratio between the C3 H8 formation rate and the PO formation rate at 433 K of catalysts over the Ti-SiO2 -x (x = 1, 2, 5) supports and (b) effect of CO and NH3 on inhibiting C3 H8 formation on 0.2-Au/Ti-SiO 2 -2 at 423 K (H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, −1 GHSV 10000 mL·g−1 cat h ; in (a), performance averaged from 150 min to 270 min in each 5-hour test) of titanium in TS-1, to a great extend, is in the closed Ti(OSi)4 structure. The smooth surface of TS-1 crystals with little defective hydroxyls is also believed to be beneficial to the reduction of unnecessary decomposition of the hydroperoxy intermediate during the catalytic reaction. However, the limited amount of Ti–OH on the surface of TS-1 makes an efficient deposition of gold close to Ti very difficult, since the surface defects seems crucial for the interaction between gold and the support in the deposition–precipitation method [32, 33]. This dilemma encouraged researchers to develop synthesis methods roughening the surface of TS-1 and greatly improved PO productivity was achieved [21, 28, 30]. Using Ti-grafted silica as an alternative may overcome the problem in anchoring gold 3.4. DISCUSSION 75 in a cheaper way, since the grafted Ti should be in the form of an open structure on the surface. However, productivities obtained on gold catalysts supported on Ti-grafted silica in earlier studies were generally inferior when compared to that on TS-1, which may be attributed to the existence of higher number (penta or hexa-) coordianted Ti [34, 35]. In our study, we were aiming at a support mainly with highly isolated Ti4+ sites, which would be reflected in a narrow band in UV–vis spectra at around 200–210 nm as Ti(OSi)4 or at 220–230 nm as Ti(OH) x (OSi)3−x [24–27]. The principle of grafting Ti onto the silica surface is through the reaction between titanium alkoxide with surface silanols. The surface silanol density of the silica used in this study was 1.8 OH/nm2 as estimated from the weight loss during dehydroxylation. Though not fully hydroxylated, it is known for this type of silica (Davisil) that most of the silanol sites are paired across the entire silica surface [36]. For simplicity, if we assume that a titanium alkoxide molecule randomly reacts with one silanol first and that the silanol sites are evenly distributed in a small area, the distribution of Ti on the silica surface may look like how Figure 3.12 shows. Sites occupied by Ti in vicinity may end up in Ti–O–Ti connectivities after calcination and accordingly penta- or hexa-coordination environment. The Ti density on the silica surface calculated in our study is around 0.1, 0.2, 0.5 and 0.8 Ti/nm2 for the Ti-SiO2 -x (x = 1, 2, 5, 10) supports based on the actual Ti loadings and the specific area of the silica. The Ti-SiO2 supports with a Ti loading of 0.2 – 0.3 wt.% showed the most similar adsorption band to TS-1 in UV-vis spectra as seen in Figure 3.1. Indeed, −1 the highest PO rates obtained in this study (120 – 130 gPO · kg−1 cat h ) were from the Ti- SiO2 -1 and Ti-SiO2 -1L supports. Increasing the Ti loading may result in a full coverage of Ti on the silica surface as illustrated in Figure 3.12. Such a higher coverage of Ti can also be evidenced by the shoulder or peak at around 170 o C in the DTG curve during dehydroxylation as shown in Figure 3.2 for Ti-SiO2 -5 and Ti-SiO2 -10, which is known as the dehydroxylation temperature of Ti–OH [37]. It seems that highly isolated Ti4+ sites are also favourable for attaining smaller gold nanoparticles as indicated by Figure 3.4. In our preparation, the gold precursor was added into the slurry at a relatively high pH (9.4–9.5), which may lead to a fast formation of a hydroxo-gold complex interacting with Ti–OH [32, 33]. The decreasing trend in gold particle size on the Ti-SiO2 -x (x = 10, 5, 2, 1) supports but with the same gold loading may be considered as an indirect evidence of the interaction between gold precursors and defective Ti–OH sites. The TS-1 used in this study has a relatively high Ti loading (1 wt.%, or Si/Ti= 78). The 0.05-Au/TS-1 catalyst showed very poor activity when 76 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY compared to the catalysts with the same loading of gold on Ti-SiO2 as listed in Table 3.3. It is, very likely, due to the limited number of Ti-OH sites on the surface of TS-1 crystals so that efficient gold deposition to Ti could not occur. (a) (b) (c) (d) Figure 3.12: Schematic of Ti distribution on silica surface for different Ti-SiO2 -x supports (a–d corresponding to x = 1, 2, 5, 10; filled square: Ti, isolated or connected; empty circle: unoccupied Si–OH) 3.4.2 Competition of epoxidation and water formation at the Au–Ti interface In this study, the most active catalysts on Ti-SiO2 -1 and Ti-SiO2 -1L have gold loadings at around and lower than 0.2 wt.%. The gold particle sizes on these catalysts should be very small (not greater than 2 nm) inferred from the trend shown in Figure 3.5 and the fact that few nanoparticles were visible. The range of gold loadings of the highly active catalysts in this study is in accordance with those in literature reporting the highest PO formation rates [4–6, 21]. It has been shown that the turnover frequencies of hydrogen 3.4. DISCUSSION 77 dissociation are relatively constant over Au/TiO2 and independent of gold particle sizes [12]. The high activity in propene epoxidation over the low-loaded catalysts may be attributed to the increased number of interfacial Au–Ti sites from the high dispersion of gold. A good synergy between gold and titanium may, to a great extend, limit the direct water formation which occurs on the gold sites not adjacent to titanium as indicated by Figure 3.9(a). However, hydrogen, as a scavenger forming the active hydroperoxy species at the Au–Ti interface, also hydrogenates this active intermediate to water [11, 38]. Recent studies on hydrogen dissociation over Au/TiO2 suggests that hydrogen dissociation occurs at low-coordinated gold atoms near the metal–support interface [13, 14]. On the other hand, it has been shown in our earlier study that the inhibiting effect of propene on water formation over gold catalysts may be due to the propene adsorption and activation on gold [39]. The Au–Ti interface is also a site for propene adsorption. [40]. As demonstrated in Figure 3.9(b) and in Chapter 2, higher propene concentrations suppress water formation while increasing the PO formation rates. The active hydroperoxy intermediate at the Au–Ti interface is competitively consumed by epoxydation and hydrogenation, while higher propene concentrations do not inhibit the rate-determining step in the hydro-epoxidation of propene. This may explain why in an industrial example a higher propene concentration was used [15]. In conclusion, our findings suggest that a high propene concentration is preferred in real operation besides a moderate concentration for hydrogen. The hydrogen efficiency of catalysts with highest PO productivity in this study is not high enough. It is within the range of 10 – 20 % if propane formation is not counted, which is lower than the level at 20 – 30 % of Au/TS-1 catalysts with their best performance [21]. There may be still room for our catalysts to improve the hydrogen efficiency due to the nature of hydrophilicity in Ti-SiO2 since no sylilation had been performed. It is known for hydrophilic support that silylation can help reduce hydrogen consumption. An example can be seen from the work by Uphade et al. [17] as given in Figure 3.13. However, their study shows that the elimination of surface hydroxyls can decrease the direct water formation significantly, while the ratio of rH2 O /rPO at the Au–Ti interface was not much changed. In industry, a typical world scale PO plant can have capacity of 300, 000 t/a. PO pro−1 ductivites of state-of-the-art gold catalysts are generally above 150 gPO · kg−1 cat h , which is equal to the activity of our most active catalyst shown in Figure 7b. Simple calculations based on these two figures suggest that a new plant based on the hydro-epoxidation 78 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY 60 H2 conversion (%) 50 40 30 20 Au/Ti−MCM−48 10 Au/Ti−MCM−48(silylated) 0 0 1 2 3 C3H6 conversion (%) 4 5 Figure 3.13: An example of improvement in hydrogen efficiency by hydrophobic treatment on the catalyst support. Data adapted from the paper by Uphade et al. [17]. (Fig. 12 in ref [17]. The catalyst was 0.09 wt% Au/Ti-MCM-48 (Si/Ti= 50). Testing was performed at 423 K in a 10-mm tube reactor. H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, GHSV −1 4000 mL·g−1 cat h . The time-on-stream performance in the original study is plotted in the H2 O-vs-PO fashion here.) technology would require 228 tons of catalyst with gold as much as 456 kg (assuming a gold loading of 0.2 wt.%), which is still rather large. By further catalyst development and process optimization, the catalyst cost can be reduced further. Especially preparing catalysts in which the gold–titanium interaction is optimized can be a significant step forward. Even though we did not observe significant deactivation over a period of up to one month, the long-term stability of catalysts on a longer time scale was not examined. However, to an industrial process this longer term stability will be essential for the process economics. On the other hand, an intrinsically high efficiency in hydrogen utilization is indispensable to make this a viable process. Although PO productivity has been improved in this study, the PO/water ratio is still lower than 20%, which is inferior to the 30% level achieved on Au/TS-1 [21]. The hydrophilic nature of our support may be the reason. Previously in Chapter 2 and once again illustrated in Figure 3.9b, we demonstrated that operating the process with low hydrogen concentrations is also an efficient manner to increase the hydrogen efficiency, although this is at the expense of the catalyst productivity. 3.4. DISCUSSION 79 3.4.3 Origin of the activity in propene hydrogenation In this study, severe propene hydrogenation was observed on most catalysts. Propene hydrogenation was only supposed to happen on gold nanoparticles smaller than 2 nm in the hydro-epoxidation of propene over the gold–titania system [1]. This side reaction was rarely reported in earlier studies [41]. In the recent two years, however, severe propene hydrogenation has been reported by different groups over this catalytic system [42–44]. It was suggested by Qi et al. [42] that there is an strict boundary for the gold particle size (2 – 5 nm) to avoid propene hydrogenation and that contamination of alkali metals like sodium in TS-1 is the reason why propene hydrogenation hardly occurs on Au/TS-1. In the study by Oyama et al. [43, 44], the Au/TS-1 catalyst with an average gold particle size of 3 nm showed around 40% selectivity towards propane. After they treated this Au/TS-1 catalyst with NaCN solutions for leaching, Au+ appeared and an 100% selectivity to propane arose. In our study, the activity of propene hydrogenation does not seem to be dependent on the gold particle size. The catalysts with activity in propene hydrogenation in this study show gold particle sizes ranging from sub-nanometer up to 6 nm. On the catalyst using TS-1 as the support (0.05-Au/TS-1), propene hydrogenation was also observed. Discrepancies in our study and literature stimulated us to try to find the possible reasons for the propene hydrogenation. Blank experiments on the as-recieved SiO2 , the hydroxylated SiO2 and the Ti-SiO2 -0 (see Experimental section) showed no activity in any reaction, which excludes the influence of contaminants in the silica (0.005 wt.% Fe2 O3 and 0.07 wt.% Na2 O as analyzed by the distributor) and the possibility of strained siloxane [45] formed during support preparation and calcination. The gold oxidation state was analyzed for several catalysts in situ and gold was found in its metallic state (detailed in Chapter 4). However, there are common features for our catalysts in propene hydrogenation: 1) propene hydrogenation can be switched off by carbon monoxide at very low concentrations; 2) programmed reaction in propene and hydrogen shows an peak activity at around 170 o C (see Chapter 4). The anomalous behaviour in propene hydrogenation of the catalysts with inversely grafted Ti, when compared to Au/SiO2 , implies the role of Ti–OH in this side reaction, which has been addressed by Sykes et al. [46, 47]. On the other hand, the distinct activity in propene hydrogenation between catalysts supported on Ti-SiO2 -1L and Ti-SiO2 -1 suggests that the grafting time of Ti onto the silica plays an important role on the dispersion of Ti as mentioned by Gao et al. [48]. It appears more to us that certain type of TiO x on the support contributes to 80 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY the activity in propene hydrogenation over our catalysts. However, this cannot explain the propene hydrogenation activity of Au/MCM-41(Au< 0.01 wt.%) in the study by Qi et al. [42]. Our results strongly indicate that propane formation occurs via the catalyst support, but we cannot yet provide conclusive information on the reactive site. Further investigation is provided in Chapter 4. 3.5 Conclusions By adjusting the site synergy between Au and Ti, highly active catalysts in the direct epoxidation of propene were achieved by using Ti-grafted silica as the support in this study. These active catalysts have low gold loadings of around 0.2 wt.%. The tetracoordinated Ti sites were attained by lowering the Ti loading to 0.2 – 0.3 wt.% (ca. 0.1 Ti/nm2 ) on the silica. High dispersion of gold close to the tetra-coordinated Ti provides sufficient Au–Ti interface leading to the high activity of these catalysts. The obtained −1 PO formation rates in this study (120 – 130 gPO · kg−1 cat h ) are at the same level of the highest rates reported on Au/TS-1. The PO/water ratio at these high PO rates ranges between 10 % and 20 %. The active hydroperoxy species are competitively consumed by epoxidation and hydrogenation at the Au–Ti interface. High propene concentrations are favourable for a lower water formation rate and a higher PO formation rate. Propene hydrogenation, if occurring, can be switched off by introducing a small amount of carbon monoxide. 3.A Tables of catalyst performance Sample ID Loading (wt%) 1.0-Au/Ti-SiO2 -1 Ti Au 0.27 0.84 Au particle size (nm) a 2.5 ± 0.9 (2.4) Temperature Formation rate b PO yield Selectivity (K) (×10−7 mol · g−1 s−1 ) cat (%) (%) PO C3 H8 H2 O 403 1.43 1.31 43.62 433 2.69 0.21 108.3 1.09 7.76 18.43 H2 efficiency (%) PO C3 H8 CO2 1.16 49.6 45.4 0.0 2.17 46.9 3.8 21.7 2.0 0.88 12.2 86.9 0.0 4.2 3.2 0.5-Au/Ti-SiO2 -1 0.48 2.4 ± 1.5 (1.9) 403 433 4.06 3.62 98.40 3.27 43.6 38.9 5.5 4.0 0.2-Au/Ti-SiO2 -1 0.20 1.7 ± 0.6 (1.6) 403 1.10 14.56 10.66 0.89 7.0 92.6 0.0 4.4 433 2.84 15.62 26.43 2.29 15.2 83.6 0.0 6.7 473c 5.44 4.59 113.1 4.66 40.2 33.9 9.4 – 403 1.72 31.33 18.32 1.38 5.2 94.2 0.0 3.5 433 3.84 30.83 35.43 3.09 10.9 87.7 0.0 5.8 473 5.47 15.23 60.46 4.41 24.7 68.8 1.9 7.2 403 433 1.25 2.78 28.51 29.96 13.45 23.36 1.01 2.23 4.2 8.4 95.3 90.7 0.0 0.0 3.0 5.2 473 5.36 17.21 47.14 4.32 23.1 74.1 0.2 8.3 0.1-Au/Ti-SiO2 -1 0.05-Au/Ti-SiO2 -1 0.09 0.05 – – 3.A. TABLES OF CATALYST PERFORMANCE Table 3.A.1: Metal loadings, gold particle sizes and catalytic performance of catalysts on the Ti-SiO2 -1 support a. averaged size ± ‘standard deviation’, the ‘standard deviation’ is calculated by assuming a normal distrubution; the median is given in parentheses. −1 −7 −1 −1 −1 = 20.9 gPO · kg−1 b. H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, GHSV 10000 mL·g−1 cat h ; under this condition, 1 × 10 mol PO · gcat s cat h . c. hydrogen was fully combusted; propylene glycol was formed at this temperature. 81 82 Table 3.A.2: Metal loadings, gold particle sizes and catalytic performance of catalysts on the Ti-SiO2 -x (x = 2, 5, 10) supports Sample ID Au particle size (nm) Temperature (K) 403 433 403 433 403 433 403 433 403 433 403 433 403 433 403 433 403 433 403 433 1.0-Au/Ti-SiO2 -2 0.44a 0.90 2.9 ± 1.1 (2.8) 0.5-Au/Ti-SiO2 -2 0.42 0.47 2.9 ± 1.1 (2.8) 0.2-Au/Ti-SiO2 -2 0.20 2.4 ± 1.0 (2.3) 0.1-Au/Ti-SiO2 -2 0.10 – 0.05-Au/Ti-SiO2 -2 0.05 – 0.80 3.5 ± 1.1 (3.5) 0.5-Au/Ti-SiO2 -5 0.43 3.1 ± 1.2 (3.0) 0.2-Au/Ti-SiO2 -5 0.19 3.1 ± 1.1 (3.2)b 0.1-Au/Ti-SiO2 -5 0.12 – 0.83 6.0 ± 1.4 (5.9) 1.0-Au/Ti-SiO2 -5 1.0-Au/Ti-SiO2 -10 1.07 1.78 Formation rate (×10−7 mol · g−1 s−1 ) cat PO C3 H8 H2 O 1.28 2.94 1.57 3.72 1.39 3.32 0.62 1.35 0.34 0.69 0.85 1.47 0.96 1.24 0.59 1.01 0.33 0.62 0.56 0.73 1.09 0.70 1.60 1.40 2.30 2.67 1.72 1.80 1.75 2.83 0.37 0.45 0.29 0.17 0.32 0.41 0.56 0.80 5.61 8.44 26.84 90.07 20.25 63.48 9.59 26.75 2.34 6.60 2.10 5.06 9.93 28.95 7.55 12.79 2.84 6.71 1.52 3.82 5.71 14.40 PO yield (%) 1.03 2.37 1.26 3.00 1.12 2.68 0.50 1.09 0.28 0.56 0.68 1.18 0.78 1.00 0.48 0.82 0.26 0.50 0.45 0.59 a. a different batch of support was used for the 1.0-Au/Ti-SiO2 -2 catalyst; Ti loading of this support was determined by XPS. b. only 97 particles were counted from 40 TEM images. Selectivity (%) PO C3 H8 CO2 51.9 59.1 48.1 62.0 37.4 52.7 35.7 42.6 16.3 19.6 66.2 64.1 74.1 78.8 64.6 64.9 36.8 42.1 8.9 7.7 44.1 14.0 49.0 23.3 62.0 42.4 64.3 56.8 83.7 80.2 29.2 19.5 22.3 10.9 34.7 26.3 63.2 54.7 89.3 87.9 0.0 6.3 0.0 3.6 0.0 0.6 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 H2 efficiency (%) 4.6 3.2 7.2 5.7 11.7 11.3 17.9 16.1 8.9 8.8 8.2 5.0 12.3 9.6 18.7 14.2 15.7 13.4 4.9 3.2 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY Loading (wt%) Ti Au REFERENCES 83 References [1] T. Hayashi, K. Tanaka and M. Haruta, J. Catal., 1998, 178, 566–575. [2] G. Mul, A. Zwijnenburg, B. Van der Linden, M. Makkee and J. A. Moulijn, J. Catal., 2001, 201, 128–137. [3] C. Qi, T. Akita, M. Okumura and M. Haruta, Appl. Catal., A, 2001, 218, 81–89. [4] L. Cumaranatunge and W. N. Delgass, J. Catal., 2005, 232, 38–42. [5] B. Taylor, J. Lauterbach and W. N. Delgass, Appl. Catal., A, 2005, 291, 188–198. [6] J. Lu, X. Zhang, J. J. Bravo-Suárez, T. Fujitani and S. T. Oyama, Catal. Today, 2009, 147, 186–195. [7] T. A. Nijhuis and B. M. Weckhuysen, Catal. Today, 2006, 117, 84–89. [8] J. J. Bravo-Suárez, K. K. Bando, J. Lu, M. Haruta, T. Fujitani and S. T. Oyama, J. Phys. Chem. C, 2008, 112, 1115–1123. [9] B. Taylor, J. Lauterbach, G. E. Blau and W. N. Delgass, J. Catal., 2006, 242, 142–152. [10] J. Lu, X. Zhang, J. J. Bravo-Suárez, S. Tsubota, J. Gaudet and S. T. Oyama, Catal. Today, 2007, 123, 189–197. [11] J. Chen, S. J. A. Halin, J. C. Schouten and T. A. Nijhuis, Faraday Discuss., 2011, 152, 321–336. [12] T. Fujitani, I. Nakamura, T. Akita, M. Okumura and M. Haruta, Angew. Chem. Int. Ed., 2009, 48, 9515–9518. [13] M. Boronat, F. Illas and A. Corma, J. Phys. Chem. A, 2009, 113, 3750–3757. [14] I. X. Green, W. Tang, M. Neurock and J. T. Yates Jr., Angew. Chem. Int. Ed., 2011, 50, 10186– 10189. [15] S. S. Dhingra (Dow Chemical), US 2010/0076208, 2010. [16] R. G. Bowman (Dow Chemical), WO 2009/061623, 2009. [17] B. S. Uphade, T. Akita, T. Nakamura and M. Haruta, J. Catal., 2002, 209, 331–340. [18] A. K. Sinha, S. Seelan, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2004, 43, 1546– 1548. [19] B. Chowdhury, J. J. Bravo-Suárez, M. Daté, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2006, 45, 412–415. [20] B. Chowdhury, K. K. Bando, J. J. Bravo-Suárez, S. Tsubota and M. Haruta, J. Mol. Catal. A: Chem., 2012, 359, 21–27. [21] J. Huang, T. Takei, T. Akita, H. Ohashi and M. Haruta, Appl. Catal., B, 2010, 95, 430–438. [22] W.-S. Lee, M. C. Akatay, E. A. Stach, F. H. Ribeiro and W. N. Delgass, J. Catal., 2012, 287, 178–189. [23] L. Y. Chen, G. K. Chuah and S. Jaenicke, J. Mol. Catal. A: Chem., 1998, 132, 281–292. [24] B. Notari, Adv. Catal., 1996, 41, 253–334. [25] P. Ratnasamy, D. Srinivas and H. Knözinger, Adv. Catal., 2004, 48, 1–169. 84 CHAPTER 3. ENHANCEMENT OF CATALYST PERFORMANCE: A STUDY INTO GOLD–TITANIUM SYNERGY [26] J. M. Fraile, J. I. García, J. A. Mayoral and E. Vispe, J. Catal., 2005, 233, 90–99. [27] O. A. Kholdeeva, I. D. Ivanchikova, M. Guidotti, C. Pirovano, N. Ravasio, M. V. Barmatova and Y. A. Chesalov, Adv. Synth. Catal., 2009, 351, 1877–1889. [28] B. Taylor, J. Lauterbach and W. N. Delgass, Catal. Today, 2007, 123, 50–58. [29] D. Harrison, I. M. Coulter, S. Wang, S. Nistala, B. A. Kuntz, M. Pigeon, J. Tian and S. Collins, J. Mol. Catal. A: Chem., 1998, 128, 65–77. [30] M. Du, G. Zhan, X. Yang, H. Wang, W. Lin, Y. Zhou, J. Zhu, L. Lin, J. Huang, D. Sun, L. Jia and Q. Li, J. Catal., 2011, 283, 192–201. [31] A. J. H. P. van der Pol, A. J. Verduyn and J. H. C. van Hooff, Appl. Catal., A, 1992, 92, 113–130. [32] R. Zanella, L. Delannoy and C. Louis, Appl. Catal., A, 2005, 291, 62–72. [33] E. Sacaliuc-Parvulescu, H. Friedrich, R. Palkovits, B. M. Weckhuysen and T. A. Nijhuis, J. Catal., 2008, 259, 43–53. [34] E. Sacaliuc, A. M. Beale, B. M. Weckhuysen and T. A. Nijhuis, J. Catal., 2007, 248, 235–248. [35] C.-H. Liu, Y. Guan, E. J. M. Hensen, J.-F. Lee and C.-M. Yang, J. Catal., 2011, 282, 94–102. [36] K. G. Proctor, S. J. Markway, M. Garcia, C. A. Armstrong and C. P. Gonzales, in Colloidal Silica: Fundamentals and Applications, ed. H. E. Bergna and W. O. Roberts, CRC Press, 2006, ch. 32, pp. 385–396. [37] C. Morterra, J. Chem. Soc., Faraday Trans. I, 1988, 84, 1617–1637. [38] J. K. Edwards, A. F. Carley, A. A. Herzing, C. J. Kiely and G. J. Hutchings, Faraday Discuss., 2008, 138, 225–239. [39] T. A. Nijhuis, E. Sacaliuc, A. M. Beale, A. M. J. van der Eerden, J. C. Schouten and B. M. Weckhuysen, J. Catal., 2008, 258, 256–264. [40] H. M. Ajo, V. A. Bondzie and C. T. Campbell, Catal. Lett., 2002, 78, 359–368. [41] A. Zwijnenburg, A. Goossens, W. G. Sloof, M. W. J. Crajé, A. M. Van der Kraan, L. J. De Jongh, M. Makkee and J. A. Moulijn, J. Phys. Chem. B, 2002, 106, 9853–9862. [42] C. Qi, J. Huang, S. Bao, H. Su, T. Akita and M. Haruta, J. Catal., 2011, 281, 12–20. [43] S. T. Oyama, J. Gaudet, W. Zhang, D. S. Su and K. K. Bando, ChemCatChem, 2010, 2, 1582– 1586. [44] J. Gaudet, K. K. Bando, Z. Song, T. Fujitani, W. Zhang, D. S. Su and S. T. Oyama, J. Catal., 2011, 280, 40–49. [45] V. K. Rajagopal, R. D. Guthrie, T. Fields and B. H. Davis, Catal. Today, 1996, 31, 57–63. [46] E. C. H. Sykes, M. S. Tikhov and R. M. Lambert, Catal. Lett., 2002, 78, 7–11. [47] E. C. H. Sykes, M. S. Tikhov and R. M. Lambert, J. Phys. Chem. B, 2002, 106, 7290–7294. [48] X. Gao and I. E. Wachs, Catal. Today, 1999, 51, 233–254. Switching off propene hydrogenation in the direct epoxidation of propene over gold–titania catalysts 4 Part of this chapter has been published as: J. Chen, S. J. A. Halin, D. M. Perez Ferrandez, J. C. Schouten and T. A. Nijhuis, 2012, J. Catal. 285, 324–327. Abstract The study into site synergy of gold and titanium yielded catalysts with high performance in propene epoxidation as well as catalysts where propene hydrogenation prevails. The side reaction of propene hydrogenation over these gold–titania catalysts were studied in details. The addition of a small amount of carbon monoxide to the feed gas can completely switch off this propene hydrogenation, while at the same time also reducing the rate of direct water formation. The formation rate of propene oxide was not affected by addition of CO. The order of CO on this inhibiting effect is −1. Gold is not necessary for this side reaction. The supports alone showed the same hydrogenation behavior as the catalysts: 1) enhancement of propene hydrogenation by O2 ; 2) peak activity at ca. 443 K in propene and hydrogen during temperature programmed reaction; 3) switching off by CO with an order of −1. The coordination environment of titanium and surface hydroxyls may play an important role in propene hydrogenation. 86 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE 4.1 Introduction Gold nanoparticles or clusters supported on titanium-containing supports are capable of epoxidizing propene to propene oxide (PO) by using hydrogen and oxygen at a high selectivity [1–4]. However, propene hydrogenation may occur or even prevail under certain circumstances as reported in literature: the particle size of gold supported on TiO2 is smaller than 2 nm [1]; or, Ti-based oxides are used as supports while the gold size is out of the range of 2–5 nm [5]; and recently, supported gold (+1) cyanide particles were reported to show a high selectivity towards propene hydrogenation [6]. Significant propane formation is rarely reported when titanium silicalite-1 (TS-1) is used as the support. Qi et al. [5] proposed that the contaminant of sodium on the catalysts is responsible for inhibiting propene hydrogenation on Ti-based gold catalysts. In Chapter 2, we have examined the effect of reactant concentrations on the epoxidation reaction over a Au/Ti-SiO2 catalyst. Based on insights provided in Chapter 2, we performed a study concerning the synergy between gold and titanium sites by simply adjusting the metal loadings. The results on the catalyst optimization are given in Chapter 3. In literature, the trend for the gold–titania system is to go to low gold loadings with highly isolated Ti4+ sites [7–10]. These low-loaded catalysts generally have a higher PO productivity, although a higher reaction temperature (generally above 423 K) is needed. The gold particle size of these catalysts is smaller. As a result, the relative amount of Au– Ti interface may be higher, which is often seen as the active site [11–15]. The smaller particle size, however, makes these catalysts more susceptible to propane formation [1]. Sacaliuc et al. [16] showed that the epoxidation activity of Au/Ti–SBA-15 can be related to the differences in the amount of grafted Ti. The work by Delgass’s group [17] has shown that Ti-defects in TS-1 can boost the activity of Au/TS-1 in direct propene epoxidation. It is also known that Ti-defects on a silica surface can stabilize the supported gold nanoparticles or clusters [18, 19]. In our study, titanium was grafted onto a commercial silica using titanium alkoxide as precursor by the surface sol–gel method. Gold was then deposited to the calcined support by the deposition–precipitation method resulting in the Au/Ti-SiO2 catalysts. The aim of the study on the site synergy was to obtain a catalyst with a higher epoxidation activity and higher hydrogen efficiency. Although catalysts with higher epoxidation activities were obtained, significant propane formation was observed over almost all the catalysts we prepared with a short grafting time for Ti, which consequently reduced 4.2. EXPERIMENTAL 87 the hydrogen efficiency to a large extent. On the other hand, we found that introducing small amount of carbon monoxide in the reactant feed suppressed propene hydrogenation over these catalysts without affecting propene epoxidation. Here we provide the evidence for the inhibiting effect of carbon monoxide on propene hydrogenation in the direct epoxidation of propene over Au/Ti-SiO2 catalysts. The root cause of this side reaction, i.e. propene hydrogenation, is also investigated and discussed in this chapter. 4.2 Experimental The catalysts synthesized in Chapter 3, which are active in propene hydrogenation, are further investigated in this chapter. Thus the names of the catalysts follow the same nomenclature given in the previous chapter. Since the preparation method used for our catalysts does not introduce sodium, an amount of 300 mg of the 0.05 wt% Au/TiSiO2 (1% ML) catalyst (i.e. 0.05-Au/Ti-SiO 2 -1) was impregnated with sodium sulfate solution and was dried in vacuum at room temperature overnight. The resulting catalyst has a sodium loading of 0.5 wt%. The 0.05-Au/Ti-SiO 2 catalyst impregnated with 0.5 wt% sodium was used to validate the effect of sodium on suppressing propene hydrogenation in the direct propene epoxidation over gold–titania catalysts as proposed by Qi et al. [5]. The reason for using Na2 SO4 is based on the proposal by Mul et al. that highly acidic sites can be neutralized by a neutral salt while the active Ti site remains active [3]. Catalytic tests were performed in a flow setup equipped with a fast Interscience Compact gas chromatography system (3 min analysis time) consisting a Porabond Q column and a Molsieve 5A column in two separate channels, each with a thermal conductivity detector. The quartz reactor is located in an oven and has an inner diameter of 6 mm. When ’reaction cycle’ is mentioned, catalyst regeneration in 10 vol.% oxygen diluted in helium at 573 K for 1 hour was performed between each cycle. Each reaction cycle normally lasted for 5 hours. The X-ray photoelectron spectroscopy (XPS) measurements were carried out with a Kratos AXIS Ultra spectrometer, equipped with a monochromatic X-ray source and a delay-line detector (DLD). Spectra were obtained using the aluminium anode (Al Kα = 1486.6 eV) operating at 150 W. A U-shape Pyrex tubular reactor (6 mm inner diameter) with detachable valves on both ends was used for post-reaction catalysts. The fresh catalysts were activated for 4–5 hours at 423 K in flowing hydrogen, oxygen and propene 88 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE (diluted in helium) in the U-shape reactor. The temperature was maintained using an oil bath. The post-reaction catalysts were then flushed in helium and afterwards sealed in the U-shape reactor. The reactor was transferred into a nitrogen glove box free from oxygen (< 2 ppm) and moisture (< 0.5 ppm). The post-reaction samples for XPS measurements were prepared in the glove box. Transport of the samples from the glove box to the spectrometer was performed in an inert atmosphere by using a small nitrogenpurged chamber equipped with a magnetic arm. In this way, the post-reaction catalysts were kept ‘in-situ’. The XPS spectra of these post-reaction catalysts were used to compare with their counterparts prior to reaction. Infrared spectra were recorded with a Bruker Vertex 70v spectrometer at a 2 cm−1 optical resolution. The samples were pressed in self-supporting discs (diameter: 12.7 mm, ca. 7.9 mg cm−2 ). Progressive adsorption of carbon monoxide at 90 K was carried out by dosing carbon monoxide through a 50 µL gas-sample loop until a pressure of 1 mbar in the IR cell was reached. The IR cell was cooled in flowing liquid nitrogen. 4.3 Results and discussion 4.3.1 Activation of hydrogenation activity The fresh calcined catalysts showed a gradual activation in their activity of propene hydrogenation during the first catalytic cycle. Figure 4.1 shows such a phenomenon of the catalysts over the Ti-SiO2 -1 support. In the first cycle, the formation rate of PO started at a high level from the very beginning, while the activity in the propene hydrogenation experienced an activating period, whose duration mainly depended on the reaction temperature. In general, a higher reaction temperature led to a shorter activating period for propene hydrogenation. For the catalysts supported on Ti-SiO2 -1 and Ti-SiO2 -2, this activating period can only be seen in the first reaction cycle. Their activity in propene hydrogenation remained relatively stable in the sequential catalytic testings, between which the regeneration in 10 vol% O2 at 573 K was performed, despite that sometimes a very small increase (less than 10%) in propene hydrogenation could be observed along time. The gold particle sizes of the catalysts shown in Figure 4.1 are small with an average size less than 2 nm. This take-off phenomenon in propene hydrogenation is much more obvious on the catalysts over the supports with higher Ti loadings, e.g., Ti-SiO2 -5 and Ti-SiO2 -10. Figure 89 4.3. RESULTS AND DISCUSSION −7 −6 x 10 3 1.5 2 1.5 O or rC 3 4.5 H H2O 4.5 rH rPO (mol⋅g−1 ⋅s−1) cat PO C3H8 (mol⋅g−1 ⋅s−1) cat x 10 6 3 8 6 (a) 0 0 1 2 3 0 5 4 Time (h) −7 −6 x 10 4.5 H2O 3 1.5 2 1.5 O or rC 3 H PO C3H8 3 8 4.5 (mol⋅g−1 ⋅s−1) cat x 10 6 rH rPO (mol⋅g−1 ⋅s−1) cat 6 (b) 0 0 1 2 3 0 5 4 Time (h) −6 −5 x 10 x 10 1 H2O 1.2 0.6 H 2 0.4 rH rPO or rC 0.8 (mol⋅g−1 ⋅s−1) cat 0.8 PO C3H8 O 1.6 3 8 (mol⋅g−1 ⋅s−1) cat 2 0.4 0.2 (c) 0 0 1 2 3 Time (h) 4 0 5 Figure 4.1: Time-on-stream performance in the reaction cycle for catalyst activation: (a) 0.05-Au/Ti-SiO 2 -1 at 423 K; (b) 0.05-Au/Ti-SiO 2 -1 at 483 K; (c) 0.2-Au/Ti-SiO 2 -1 at 473 K (gas feed 10 vol% hydrogen, 10 vol% oxygen, 10 vol% propene in helium, GHSV 10000 −1 mL·g−1 cat h ). 90 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE 4.2 shows the activity in propene hydrogenation over the 1.0-Au/Ti-SiO 2 -10 catalyst, of which the average size of gold nanoparticles is 6.0 ± 1.4 nm (see Chapter 3 for details), in the first two reaction cycles. By the end of the second reaction cycle, the activity in propene hydrogenation had increased by one magnitude when compared to the activity at the very beginning of the first cycle. However, the formation rates of PO and water showed very little change in the first two cycles. The irrelevance between PO formation and propane formation over these catalysts implies that the two catalytic reactions may proceed over different sites. And this take-off phenomenon for propene hydrogenation indicates that this side reaction may be suppressed if the catalyst would remain its original status at the very beginning of the first reaction cycle. −6 1.2 x 10 0.8 0.6 rC H 3 8 (mol⋅g−1 ⋅s−1) cat 1 0.4 0.2 cycle 1 cycle 2 0 0 1 2 3 4 5 Time (h) Figure 4.2: Increasing activity in propene hydrogenation over the 1.0-Au/Ti-SiO 2 -10 catalyst during the first two catalytic cycles at 423 K (gas feed 10 vol% hydrogen, 10 vol% −1 oxygen, 10 vol% propene in helium, GHSV 10000 mL·g−1 cat h ). 4.3.2 Switching off propene hydrogenation by CO Carbon monoxide was found to be able to suppress propene hydrogenation while not affecting the epoxidation over the catalysts investigated in Chapter 3. In this section, the 0.05-Au/Ti-SiO 2 -1 catalyst is used as an example to illustrate such an effect of CO. Figure 4.3 gives the general performance of 0.05-Au/Ti-SiO 2 -1 at different temperatures. The conversion level of propene on this catalyst is very similar to the conversions reported by Qi et al. over their Au/Ti-TUD-1 (Ti/Si= 2/100 ) catalysts with very low gold loadings [5]. The conversion of propene on 0.05-Au/Ti-SiO 2 -1 decreased at higher temperatures, which was mainly due to the decreased activity in propene hydrogenation. At ca. 433 K, 91 4.3. RESULTS AND DISCUSSION 30 Conversion Yield Conversion, yield (%) 25 20 15 10 5 0 423 433 443 453 463 473 Temperature (K) 483 493 100 rest PO C3H8 Selectivity (%) 80 60 40 20 0 432 443 454 463 473 Temperature (K) 483 Figure 4.3: Catalyst performance of the 0.05-Au/Ti-SiO 2 -1 catalyst at different temperatures: (a) conversion of C3 H6 and yield to PO; (b) product selectivity (tested in 5-hour reaction cycles, gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in helium, GHSV −1 10000 mL·g−1 cat h , performance averaged between 150 min and 270 min in each cycle, 1-hour regeneration in 10 vol% O2 at 573 K between each cycle). the selectivity towards propane was as high as 90%. The activity in the propene epoxidation monotonously increased when the reaction temperature was raised. The selectivity towards oxygenates also increased at higher temperatures, which is in accordance with the behaviour of a normal catalyst which is only active in propene epoxidation as discussed in Chapter 2. In Figure 4.4, the effect of carbon monoxide addition on the formation rates of PO, water and propane is shown for the 0.05-Au/Ti-SiO 2 -1 catalyst. When the experiment started, the catalyst bed temperature increased sharply from ca. 463 K for over 10 K; due to the exothermic reactions. The temperature stabilized at around 473 K after about 1 h. At t = 2.5 h, the catalyst bed temperature was raised to 483 K to determine the CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE −6 −6 x 10 5 4 0.6 3 0.4 2 1 2 O H 3 8 0.8 or rC x 10 rH rPO (mol⋅g−1 ⋅s−1) cat 1 (mol⋅g−1 ⋅s−1) cat 92 PO C3H8 0.2 H2O 0 0 5 Temperature (K) 493 10 Time (h) 15 0 483 473 463 453 Catalyst bed Oven setting 0 5 10 Time (h) 15 Figure 4.4: Time-on-stream performance of the 0.05-Au/Ti-SiO 2 -1 catalyst (top) and the catalyst bed temperature (bottom) with and without CO addition (gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in helium, GHSV 10000 mL·g−1 h−1 , grayed area: 1 vol% cat CO introduced between 5 – 7.5 h, 10 – 12.5 h). temperature influence on the formation rate of propane. From t = 2.5 h till the end of the experiment, the oven temperature was kept constant. Carbon monoxide was introduced into the gas feed at a low concentration of 1 vol% during the time intervals between 5 h – 7.5 h, 10 h – 12.5 h. The formation rate of propane dropped immediately by two orders of magnitude without affecting the epoxidation reaction as carbon monoxide was added. Meanwhile, water formation was also suppressed. Accordingly, the catalyst bed temperature dropped by about 5 K. During the co-feeding of carbon monoxide, the hydrogen efficiency increased from 8 % to 17 %. The conversion of carbon monoxide was low at the level of 10 %. An increase in the formation rate of carbon dioxide was observed when carbon monoxide was added, but this was only the result of carbon monoxide conversion while not of propene combustion. Hydrogenation of carbon monoxide to methane was not observed. Removing carbon monoxide restored the catalyst activity in propene hydrogenation (7.5 h – 10 h, 12.5 h – 17.5 h). The result in the first 5 h showed that propane formed at a higher rate when the reaction temperature was lower on this catalyst. Therefore, the suppression of propane formation between 5 h – 7.5 h and 10 h 93 4.3. RESULTS AND DISCUSSION – 12.5 h can be completely attributed to carbon monoxide addition. −6 −6 x 10 x 10 5 PO C3H8 4 3 0.4 2 0.2 1 2 O H 3 8 0.6 or rC H2O rH rPO (mol⋅g−1 ⋅s−1) cat 0.8 (mol⋅g−1 ⋅s−1) cat 1 0 0 5 Temperature (K) 493 10 Time (h) 15 0 483 473 463 453 Catalyst bed Oven setting 0 5 10 Time (h) 15 Figure 4.5: Time-on-stream performance of the 0.05-Au/Ti-SiO 2 -1 catalyst impregnated with 0.5 wt% sodium (top) and the catalyst bed temperature (bottom) with and without CO addition (gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in helium, GHSV 10000 −1 mL·g−1 cat h , grayed area: 1 vol% CO introduced between 7.5 – 10 h). The effect of adding sodium to the 0.05-Au/Ti-SiO 2 -1 catalyst was also investigated to compare the effect of the carbon monoxide addition on suppressing the hydrogenation reaction to the recently published approach of adding sodium [5]. The results are shown in Figure 4.5. When the reaction started, the catalyst bed temperature increased instantly from ca. 463 K to 468 K. The catalyst bed temperature was raised to around 480 K at t = 2.5 h by increasing the oven temperature. Carbon monoxide (1 vol%) was introduced into the gas feed from t = 7.5 h till the end (t = 10 h). As seen from Figure 4.5, the existence of sodium has little effect on propene hydrogenation over our catalyst. However, the formation of PO and water was suppressed after sodium sulfate impregnation comparing to the original activity shown in Figure 4.5. It should be noted here that both sodium and sulfate might have great potential in modifying the catalyst acidity [20]. Figure 4.5 clearly shows that addition of carbon monoxide inhibited propene hydrogenation without affecting propene epoxidation over the sodium-impregnated catalyst while sodium had little effect on hydrogenation. The work by Haruta’s group [5] 94 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE proposed that propane formation could be reduced by sodium addition to the catalysts. Our results show that such a modification to the catalyst is apparently not as simple as it seems. Carbon monoxide addition, however, is a very simple and highly effective way to block propane formation (for the case where propene hydrogenation is unavoidable in the propene epoxidation under an atmosphere of hydrogen and oxygen), which was determined to be effective for all the catalysts in this study. −5 10 0.05% 0.1% CO CO 0.2% CO 0.3% CO 0.5% CO 0% CO −6 10 −7 10 PO C3H8 (a) H2O −8 10 0 2.5 5 7.5 Time (h) 10 12.5 −6 −6 x 10 x 10 5 4 0.6 3 0.4 2 1 2 O H 3 8 0.8 or rC (b) rH rPO (mol⋅g−1 ⋅s−1) cat 1 (mol⋅g−1 ⋅s−1) cat r (mol⋅g−1 ⋅s−1) cat 0% CO 0.2 PO CH 3 8 H2O 0 0 2.5 5 Temperature (K) 493 7.5 Time (h) 10 12.5 0 (c) 483 473 463 453 Catalyst bed Oven setting 0 2.5 5 7.5 Time (h) 10 12.5 Figure 4.6: Effect of CO concentrations (volumetric) on the formation rates of propene oxide, propane and water over the 0.05-Au/Ti-SiO 2 -1 catalyst: (a) formation rates in logarithmic scale; (b) formation rates in decimal scale; (c) catalyst bed temperature (gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in helium, CO introduced when required, −1 GHSV 10000 mL·g−1 cat h ). 95 4.3. RESULTS AND DISCUSSION Figure 4.6 shows the effect of CO concentrations on propene hydrogenation over the 0.05-Au/Ti-SiO 2 -1 catalyst. Over this catalyst while the addition of 0.05 vol% of carbon monoxide can already reduce 80% of propane formation on our catalysts, a concentration of 1 vol% for carbon monoxide may be needed to keep the ratio of propane and PO formation rates below 0.05. Such an effect of CO on propene hydrogenation is not only valid for the catalysts supported on Ti-SiO2 , but also true for the 0.05-Au/TS-1 catalyst studied in Chapter 3. Figure 4.7 shows how CO suppressed the propane formation during the direct epoxidation of propene over 0.05-Au/TS-1. The overal activity of this 0.05-Au/TS-1 catalyst was very low. The hydrogen efficiency was around 10 %. When no CO was introduced, the conversion of propene was 0.75 % at 473 K. The main side products were propane and propanal. The selectivity towards propane and propanal was 20 % and 17 %, respectively. When CO was introduced, the propane formation was completely suppressed, while the rates for PO and propanal were not affected. The concurrent decrease in water and propanal formation in the first 2 hours as shown in Figure 4.7 indicates that some strong acidic Ti−OH sites were passivated probably due to strong adsorption of reaction products. These acidic Ti−OH sites close to gold normally cause additional water formation and PO isomerization. −7 −6 0.8 0.8 0.6 0.6 PO C3H8 0.4 2 rH H rPO, rC 0.4 H2O (mol⋅g−1 ⋅s−1) cat x 10 1 O x 10 3 8 or rpropanal (mol⋅g−1 ⋅s−1) cat 1 propanal 0.2 0 0 0.2 1 2 3 4 Time (h) 5 6 7 0 Figure 4.7: Inhibiting effect of CO on propene hydrogenation over the 0.05-Au/TS-1 catalyst (at 473 K, gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in helium, 1 vol% −1 CO introduced between 2.5 h – 5 h, GHSV 10000 mL·g−1 cat h ) 96 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE −6 0.05−Au/Ti−SiO2−1 0.05−Au/Ti−SiO2−1 1.0−Au/Ti−SiO2−2 1.0−Au/SiO2−Ti−5 1.0−Au/Ti−SiO2−10 −7 10 rC H 3 8 (mol⋅g−1 ⋅s−1) cat 10 −8 10 1 10 2 3 10 10 4 10 PCO (ppm) Figure 4.8: Inhibition effect of CO on propene hydrogenation over different catalysts (gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in helium, CO concentration varied, −1 o o GHSV 10000 mL·g−1 cat h ; 200 C for 0.05-Au/Ti-SiO 2 -1, 130 C for 1.0-Au/Ti-SiO 2 -2, o 150 C for 1.0-Au/SiO 2 -Ti-5 and 1.0-Au/Ti-SiO 2 -10). There is no clear correlation between the activity in propene hydrogenation and the amount of gold or titanium loaded. On the other hand, the activity in propene hydrogenation can change during reactions or simply after storage for a long time and seems irrelevant to PO formation. The catalytic performance of the catalysts studied in Chapter 3 does not support the hypothesis that there is a strict boundary for the gold particle size as proposed in literature [5]. Adding sodium also had no influence on propene hydrogenation as shown in Figure 4.5. Nevertheless, CO can effectively suppress propene hydrogenation as long as it happens over the gold–titania catalysts. Meanwhile, it does not affect the reaction rate of PO formation. The order of CO on this inhibiting effect is determined to be −1.06 ± 0.13 over different catalysts as shown in Figure 4.8. The reaction rates of propene hydrogenation under conditions given in Figure 4.8 were very stable. Due to the exothermicity of the three reactions, i.e., water formation, PO formation and propene hydrogenation, the formation rates of propane under different CO concentrations were those at a constant temperature within an error of ±1 o C. In other words, they were those at relatively high CO concentrations (still ppm level) where the change in temperature due to the change in reaction rates became small. The conversion of CO to CO2 was not tracked for the experiments with CO concentrations (introduced) 97 4.3. RESULTS AND DISCUSSION below 103 ppm. However, the combustion of CO to CO2 was limited since CO could still be observed by GC in all cases. 4.3.3 Probing the active site for propene hydrogenation Oxidation state of gold and titanium There are concerns about the oxidation states of gold and titanium which may contribute to propene hydrogenation over the gold–titania catalysts: 1. oxidized gold, i.e., Au+ as proposed by Oyama et al. [6]; 2. reduced titanium, e.g., Ti3+ suggested by Sykes et al. [21, 22], hydride complexes of titanium proposed by Yermakov et al. [23]. In section 4.3.1, the activation of catalysts for propene hydrogenation has been described. If the activation period for propene hydrogenation is due to the change in oxidation states of gold and/or titanium, this change may be observed by comparing the fresh catalysts and their counterparts under reactions for a couple of hours. Pseudo in-situ XPS experiments were thus performed over three representative catalysts: 1.0-Au/SiO2 -Ti-5, 1.0-Au/TiSiO2 -10, and 0.2-Au/Ti-SiO 2 -1. Their catalytic performances are listed in the appendix of Chapter 3. The 1.0-Au/SiO 2 -Ti-5 catalyst was prepared by grafting Ti onto a 1 wt.% Au/SiO2 catalyst and gained activity in propene hydrogenation after grafting Ti. 8 7.5 PO C H 3 8 Yield (%) 6 4 2 1.06 0 0.25 0 1.0 wt% Au/Ti-SiO 2 1.0-Au/SiO -Ti-5 2 Figure 4.9: Comparison of catalytic performance between the 1 wt.% Au/Ti-SiO 2 catalyst investigated in Chapter 2 for kinetic studies and the 1.0-Au/SiO 2 -Ti-5 catalyst by inverse Ti grafting in Chapter 3 (423 K, gas feed 10 vol% H2 , 10 vol% O2 and 10 vol% C3 H6 in −1 helium, GHSV 10000 mL·g−1 cat h , performance averaged between 150 min and 270 min in a 5-hour reaction cycle) 98 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE 1.0−Au/SiO2−Ti−5 Au 1.0−Au/SiO2−Ti−5 4f7/2= 83.3 eV as prepared Ti as prepared (1.35) 2p3/2= 459.0 eV (2.85) ∆ = 3.68 eV 1.0−Au/SiO2−Ti−5 ∆ = 5.70 eV 1.0−Au/SiO2−Ti−5 4f7/2= 83.3 eV activated 2p3/2= 458.9 eV activated (1.49) (2.81) ∆ = 3.64 eV 4f7/2= 83.4 eV as prepared (1.48) ∆ = 3.71 eV 1.0−Au/Ti−SiO2−10 4f7/2= 83.4 eV activated (1.47) ∆ = 3.71 eV 0.2−Au/Ti−SiO2−1 x4 as prepared 4f7/2= 83.4 eV (1.90) 1.0−Au/Ti−SiO2−10 Intensity (arb. units) Intensity (arb. units) 1.0−Au/Ti−SiO2−10 ∆ = 5.68 eV 2p3/2= 459.0 eV as prepared (2.71) ∆ = 5.65 eV 1.0−Au/Ti−SiO2−10 2p3/2= 459.1 eV activated (2.76) ∆ = 5.65 eV 0.2−Au/Ti−SiO2−1 x4 as prepared 2p3/2= 459.5 eV (2.70) ∆ = 3.70 eV 0.2−Au/Ti−SiO2−1 activated x4 4f7/2= 83.4 eV (1.61) ∆ = 5.51 eV 0.2−Au/Ti−SiO2−1 activated x4 2p3/2= 459.5 eV ∆ = 3.70 eV 92 90 88 86 84 82 Binding energy (eV) 80 (2.99) ∆ = 5.50 eV 470 465 460 455 Binding energy (eV) 450 Figure 4.10: XPS spectra of Au 4f and Ti 2p for catalysts before and after reaction (‘as prepared’: fresh and flushed in helium at 423 K for 1 hour; ‘activated’: after 5-hour reaction in 2 % H2 , 2 % O2 and 10 % C3 H6 at 423 K and then flushed in helium at 423 K for 1 hour; spectra referenced to Si 2p line at 103.3 eV). Figure 4.10 shows the XPS spectra of Au 4f and Ti 2p lines for the selected catalysts before and after reaction. Trivial difference can be told from the comparison of Au 4f lines or Ti 2p lines. The gold atoms are determined to be Au0 and Ti is determined to be Ti4+ . The 1.0-Au/SiO 2 -Ti-5 catalyst showed an significant increase in intensity of Au 4f line after reaction, which is most likely due to the observed desorption of acetone (converted by gold from 2-propanol used for Ti grafting since no calcination was performed, see Experimental in Chapter 3 for details) and 2-propanol during heating up to the reaction temperature. The existence of oxidized gold is doubtful under the reaction atmosphere containing H2 and C3 H6 . Infrared spectroscopy using CO as probe molecule was also implemented to check if oxidized gold can be formed by calcination in O2 . 99 4.3. RESULTS AND DISCUSSION 0.04 (a) 1.0 O2 5 act vate ex s t s. 0.03 CO OH 0.02 CO ph s sor e 0.01 CO press re 0 2200 2150 aven m er (cm 2100 2050 1 0.04 ( ) 1.0 s. 0.03 O2 5 CO OH a ter calc nat on n O2 0.02 CO 4+ CO ph s sor e 0.01 CO press re 0 CO 2200 d+ CO 2150 aven m er (cm 0 2100 2050 1 Figure 4.11: Difference infrared spectra of progressive CO adsorption on the 1.0Au/SiO2 -Ti-5 catalyst at 90 K (the spectrum before CO dosing in each experiment was taken as the background for substraction; CO pressure: from 0.07 to 1.0 mbar, ca. 0.07 mbar interval; in ‘a’, the sample was activated ex-situ in a mixture of H2 /O2 /C3 H6 at 423 K and then dehydrated at 573 K in-situ; ‘b’ was performed after ‘a’, the same pellet was used, calcination was performed in-situ at 573 K and then evacuated at 573 K for 10 min.) Figure 4.11 shows the progressive CO adsorption on 1.0-Au/SiO 2 -Ti-5. For the activated catalyst, it seems that both Au and Ti sites were occupied by organic components and no adsorption was observed for CO. After calcination in O2 , the carbonaceous surface was cleaned. The band at 2179 cm−1 is assigned to CO adsorbed on Ti4+ . The broad band between 2100 and 2140 cm−1 is composed of contributions from CO on Auδ+ and Au0 . Progressive CO adsorption on 1.0-Au/Ti-SiO 2 -10 showed similar IR spectra to Figure 4.11. In Figure 4.2, the 1.0-Au/Ti-SiO2-10 catalyst showed continuous increase in hydrogenation activity in the first 2 reaction cycles, between which calcination in O2 was 100 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE performed. It is very unlikely that such a phenomenon can be attributed to oxidized gold. In-situ electron paramagnetic resonance spectroscopy (EPR, 9.4 GHz, 2.0 mW, RT or 110 K) was also performed to detect Ti3+ if there is any. However, no signal for paramagnetic species was observed for the activated catalysts. In summary, no obvious change in oxidation states of gold and titanium was observed. Effect of hydroxyls and O2 Although the root cause of propene hydrogenation as a side reaction remains obscure, there is a very important feature for the catalysts active for propene hydrogenation. Without oxygen, the time-on-stream performance in propane formation simply declines. Another consideration hinted by the work of Sykes et al. [21, 22] is that for their reduced TiO2 certain degree of hydroxylation is needed for the hydrogenation activity. If hydroxyls play a role in propene hydrogenation, temperature programmed reaction in C3 H6 and H2 only may be used to validate this hypothesis. The 0.05-Au/Ti-SiO 2 -1 catalyst is again used as the first example for illustration here. Figure 4.12 shows the results of an experiment designed to illustrate the effect of hydroxyls and O2 . In the first hour, the temperature of catalyst bed was raised from 303 K to 473 K with a ramping rate of 3 K/min. The activity in propene hydrogenation reached its peak at ca. 443 K and then started to decrease. CO was introduced in period III and propene hydrogenation was consequently switched off. In period IV, water vapor was introduced with an attempt to restore the activity in propene hydrogenation since the hydroxylation of Ti−O−Ti or Ti−O−Si was deemed important for this reaction. However, water vapor had no effect on restoring the declined hydrogenation activity. Thus, in period V, the reaction feed was replaced by 2 vol.% H2 and 2 vol.% O2 . Water was formed and the bed temperature increased significantly in period V. After the treatment in H2 and O2 , the hydrogenation activity of this catalyst was checked in the last hour (after t = 17.5 h). The hydrogenation activity was restored and showed the same pattern as the activity during t = 1–3 h. Attempts to restore the hydrogenation activity by treating with pure H2 or O2 were also performed. However, these attempts failed to restore the catalyst activity in propene hydrogenation. Figure 4.13 shows the temperature programmed reaction in H2 and C3 H6 but without O2 over different catalysts. No matter how active these catalysts are in propene hydrogenation, all of them showed a peak activity at 443 ± 5 K, a temperature where 101 4.3. RESULTS AND DISCUSSION −6 2.5 x 10 II III CO 0 2 4 6 8 10 12 Time (h) 14 16 18 20 0 2 4 6 8 10 12 Time (h) 14 16 18 20 I III CO II II III CO IV V II 1.5 1 rC H 3 8 (mol⋅g−1 ⋅s−1) cat 2 0.5 0 225 200 Temperature (oC) 175 150 125 100 75 50 25 Figure 4.12: Propane formation rate (top) and catalyst bed temperature (bottom) over the 0.05-Au/Ti-SiO 2 -1 catalyst in 10 vol% C3 H6 and 10 vol% H2 (Prior to the experiment, the catalyst was hydroxylated in 2 vol% H2 + 2 vol% O2 at 423 K for 1 hour and was then flushed in helium at 423 K for 1 hour, after which the catalyst was cooled down to 303 K. I: temperature programmed reaction, 3 K/min from 303 K to 478 K; II: no CO introduced; III: 1 vol% CO introduced; IV: no CO, 0.5 vol% water vapor introduced; V: no CO, re-hydroxylation in 2 vol% H2 + 2 vol% O2 instead of the reaction atmosphere −1 of 10 vol% H2 + 10 vol% C3 H6 . GHSV 10000 mL·g−1 cat h , helium balance. II, III, IV, V, constant oven setting temperature). either dehydroxylation or removal of water with a stronger adsroption occurred. Even for the 0.05-Au/TS-1 catalyst, whose activity in propene hydrogenation is very low, such a temperature-related feature can be observed. There are also some other interesting phenomena during propene hydrogenation without O2 when CO was introduced or in a case where the reaction started at a different temperature but after an identical pretreatment. Figures 4.14 and 4.15 show such phenomena using 1.0-Au/SiO 2 -Ti-5 as an example. In Figure 4.14, the activity in propene 102 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE 2.5 x 10 −6 x 10 1 o −8 −7 1 x 10 x 10 1 −7 200 ( C) o 0.5 2 150 0.6 0.6 0.05−Au/TS−1 0.4 0.4 2 100 0.2 0.2 50 50 60 0 180 120 0 0 60 Time (min) x 10 1 −6 1 o C3H8 1 0.2 (mol⋅g−1 ⋅s−1) cat H 0.4 100 2 2 3 8 2 rC 0.6 1.0−Au/Ti−SiO −2 O 150 rH Temperature 0.8 (mol⋅g−1 ⋅s−1) cat 0.8 H2O 3 0 200 ( C) C3H8 0.8 H2O Temperature 0.6 150 0.6 1.0−Au/Ti−SiO −10 2 0.4 0.4 100 0.2 50 0 x 10 1 o 200 ( C) rC H 3 8 (mol⋅g−1 ⋅s−1) cat 4 −7 x 10 2 x 10 −7 rH 5 0 180 120 Time (min) −7 (mol⋅g−1 ⋅s−1) cat 0 O 0 0.8 3 8 Temperature (mol⋅g−1 ⋅s−1) cat (mol⋅g−1 ⋅s−1) cat H 0.2 CH HO 3 8 rC 2 H 0.4 rC 100 1 0.6 (mol⋅g−1 ⋅s−1) cat 0.05−Au/Ti−SiO2−1 O 1.5 200 ( C) 0.8 rH (mol⋅g−1 ⋅s−1) cat Temperature 150 3 8 0.8 2 rH 3 8 HO O CH 2 0.2 50 60 120 Time (min) 0 180 0 0 60 120 0 180 Time (min) Figure 4.13: Temperature programmed reaction in hydrogen and propene for different catalysts (pretreatment: the catalyst was hydroxylated in 2 vol.% H2 + 2 vol.% O2 at 423 K for 1 hour and then flushed in helium at 423 K for 1 hour; reaction conditions: −1 10 vol.% H2 + 10 vol.% C3 H6 , helium balance, GHSV 10000 mL·g−1 cat h , ramping rate 3 K/min) hydrogenation showed the identical pattern to Figure 4.13 in the first 150 min, that is, a peak activity at ca. 443 K and then a continuous decay. When a very small amount of CO was introduced, the hydrogenation activity was suppressed to a low level but remained constant. There were some water formed in the period when 200 ppm and 500 ppm CO was introduced. The origin of this water is unclear since there is no O2 in the gas feed and the GC used in this study is also very sensitive to O2 at ppm level. Probably there was a trace amount of contaminant from air remained in the gas pipelines. In the other temperature programmed experiments where CO was co-fed from the very beginning, propene hydrogenation was suppressed at a very low level all the time (not shown). Figure 4.15 shows the activity evolution in another temperature programmed experiment over 1.0-Au/SiO 2 -Ti-5. After pretreatment in H2 and O2 , the reaction started at 403 K (130 o C) instead of 303 K. The reaction temperature was kept at 403 K (130 o C) for 30 min and then temperature ramping started. At 403 K, the activity also kept decreasing. Once the temperature started to increase, the reaction rate increased accordingly and 103 4.3. RESULTS AND DISCUSSION reached its maximum at ca. 443 K, after which decay started again. −7 −7 3.5 x 10 x 10 1 o C3H8 C3H6 an 10 H2 1.4 0.6 500 ppm CO 10 10 10 H2 0 H2 100 CO 0.7 0.2 0 ppm 200 ppm 500 ppm 50 0 C3H6 0.4 H2 0 60 120 180 240 Time me (min) (m n) 300 O 10 2.1 150 2 emperat re (mol⋅g−1 ⋅s−1) cat 0.8 H2O rC H 3 8 (mol⋅g−1 ⋅s−1) cat 2.8 rH ( C) 200 0 360 Figure 4.14: Effect of CO on propene hydrogenation without O2 (1.0-Au/SiO 2 -Ti-5, pretreatment: the catalyst was hydroxylated in 2 vol.% H2 + 2 vol.% O2 at 423 K for 1 hour and then flushed in helium at 423 K for 1 hour. reaction conditions given on graph, −1 GHSV 10000 mL·g−1 cat h ) −7 6 −7 x 10 x 10 1 o 0.8 H2O Temperature 0.6 0.4 100 2 2.4 150 rH 3.6 (mol⋅g−1 ⋅s−1) cat C3H8 O 200 ( C) rC H 3 8 (mol⋅g−1 ⋅s−1) cat 4.8 1.2 0.2 50 0 0 60 120 0 180 Time (min) Figure 4.15: Temperature programmed reaction in hydrogen and propene over 1.0Au/SiO2 -Ti-5. (pretreatment: the catalyst was hydroxylated in 2 vol.% H2 + 2 vol.% O2 at 423 K for 1 hour and then flushed in helium at 423 K for 1 hour; reaction condi−1 tions: 10 vol.% H2 + 10 vol.% C3 H6 , helium balance, GHSV 10000 mL·g−1 cat h , ramping rate 3 K/min) Both the experiments in Figure 4.14 and Figure 4.15 are highly reproducible in terms of the quantified reaction rate. But when one compares the reaction rates at 403 K(130 o C) or 443 K(170 o C) (where the activity reached its maximum) for the same catalyst in Figure 4.14 and Figure 4.15, they are quite different. It seems that propene hydro- 104 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE genation over these catalysts are very sensitive to the amount of hydroxyls present on the surface. The history of catalyst treatment heavily influences the activity in propene hydrogenation. Later in this chapter, the high reproducibility after identical pretreatment will also be shown. In order to validate the importance of dehydroxylation or dehydration at ca. 443 K to propene hydrogenation, another experiment was performed over the 1.0-Au/SiO 2 -Ti-5 catalyst for illustration. The results are given in Figure 4.16. The fresh catalyst showed no activity in propene hydrogenation at 423 K without hydroxylation or hydration in H2 and O2 . Pretreatment in 2 vol% water vapor only (diluted in helium) at 423 K showed no help either. Thus, the catalyst was activated in 2 vol% H2 and 2 vol% O2 at 423 K for 3 hours. The water formation rate increased along time and gradually stabilized. The change in water formation rate may be attributed to the gradual removal of adsorbed propoxy species due to 2-propanol used in grafting Ti on this catalyst (see Experimental in Chapter 3 for details). After drying in helium at 423 K, the activated catalyst still showed a relatively high activity in propene hydrogenation (ca. 5 % conversion of C3 H6 to C3 H8 at t = 210 min) as shown in Figure 4.16(a). The activity in propene hydrogenation continued to decrease in only C3 H6 and H2 . After O2 was co-fed, the formation rate of propane was increased as shown in Figure 4.16(a). However, after drying in helium at 573 K, the activated catalyst showed almost no activity in propene hydrogenation as shown in Figure 4.16(b). Co-feeding of O2 boosted propene hydrogenation. It is clear that hydrogen dissociation was enhanced by the presence of O2 and that the removal of hydroxyls or water had a negative impact on propene hydrogenation. However, still, the observation of simultaneous dehydroxylation or dehydration and decrease in hydrogenation activity can only be considered as an indirect evidence. Propene hydrogenation over the supports Since the dehydroxylation or dehydration is only relevant to the support, if gold nanoparticles are not the only source for hydrogen dissociation, propene hydrogenation should be observed on the supports only. This is true. Parallel to the experiment performed on the 0.05-Au/Ti-SiO 2 -1 catalyst as shown in Figure 4.12, the same experiment was performed over the Ti-SiO2 -1 support. The result is given in Figure 4.17. The Ti-SiO2 -1 support showed exactly the same catalytic features as the 0.05-Au/Ti-SiO 2 -1 catalyst in propene hydrogenation as shown in Figure 4.12. The support itself showed an even higher hy- 105 4.3. RESULTS AND DISCUSSION -6 3.0x10 10% H +10% C H + /s) (a) 2 6 2 o He, 150 C cat Formation rate (mol/g 3 o 10% O , 150 C, 2 hours -6 2% H 2.0x10 2 0.5 hours + 2% O H O 2 2 o 150 C, 3 hours 10% H 2 + 10% C H 3 6 o 150 C, 2 hours C H -6 1.0x10 3 8 PO x 5 0.0 0 60 120 180 240 300 360 420 t (min) -6 3.0x10 /s) (b) o Formation rate (mol/g cat He, 300 C H O 2 0.5 hours then 2% H -6 2 2.0x10 + 2% O 2 o 10% H +10% C H + cool down 2 150 C, 3 hours 3 6 o 10% O , 150 C, 2 hours 2 10% H 2 + 10% C H 3 6 o 150 C, 2 hours C H 3 -6 1.0x10 8 PO x 5 0.0 0 60 120 180 360 420 480 540 t (min) Figure 4.16: Effect of hydroxyls and O2 on propene hydrogenation (1.0-Au/SiO 2 -Ti-5, −1 GHSV 10000 mL·g−1 cat h ) drogenation activity at ca. 443 K or after restoring the activity by H2 /O2 treatment. Further check of the hydrogenation activity over different supports by the temperature programmed reaction in H2 and C3 H6 also confirmed the catalytic features shown in Figure 4.13 over the gold-containing catalysts. That is, a peak activity at ca. 443 K and activity decay when O2 is not present. It is clear that in the presence of O2 , H2 can dissociate on the support. Gold is not necessary for propene hydrogenation over the catalysts investigated. This explains why the propene epoxidation was not affected when propene hydrogenation was switched off. It seems that propene hydrogenation is merely catalyzed over the support, which is consistent with the findings by Sykes et al. [21, 22]. 106 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE −6 2.5 x 10 I III CO II II III CO II IV II III CO II 1.5 1 rC H 3 8 (mol⋅g−1 ⋅s−1) cat 2 0.5 0 0 2 4 6 8 10 12 14 16 18 20 22 24 Time (h) 0 2 4 6 8 10 12 14 16 18 20 22 24 Time (h) 225 200 Temperature (oC) 175 150 125 100 75 50 25 Figure 4.17: Propane formation rate (top) and catalyst bed temperature (bottom) over the Ti-SiO2 -1 support in 10 vol% C3 H6 and 10 vol% H2 (Prior to the experiment, the support was hydroxylated in 2 vol% water vapor in helium at 423 K for 2 hours and then in 2 vol% H2 + 2 vol% O2 at 423 K for 1 hour, after which the support was flushed in helium at 423 K for 1 hour and then cooled down to 303 K. I: temperature programmed reaction, 3 K/min from 303 K to 478 K; II: no CO introduced; III: 1 vol% CO introduced; IV: no CO, re-hydroxylation in 2 vol% H2 + 2 vol% O2 instead of the reaction atmosphere −1 of 10 vol% H2 + 10 vol% C3 H6 . GHSV 10000 mL·g−1 cat h , helium balance. II, III, IV, constant oven setting temperature). The order of CO on inhibiting propene hydrogenation was also checked over the Ti-SiO2 -10 support. The results are summarized in Table 4.1 and Figure 4.18. Over the Ti-SiO2 -10 support, the order of CO on propene hydrogenation gives −0.91, which is almost the same as the order on the gold-containing catalysts. The CO order was calculated based on low formation rates of propane to exclude the influence from the temperature. What should be mentioned here is that the highest rates shown in Figure 4.18 for each catalyst or support were already less than 20 % of the formation rates of 107 4.3. RESULTS AND DISCUSSION propane when there was no CO present. Table 4.1: CO order of the inhibiting effect on propene hydrogenation Sample ID 0.05-Au/Ti-SiO2 -1 0.05-Au/Ti-SiO2 -1 T (K) Reaction conditionsa 473 10/10/10 H2 /C3 H6 /O2 1.4 × 10 −6 −6 −1.26 rC H 3 8 (mol·g−1 h−1 ) cat b CO order −1.10 473 10/10/10 H2 /C3 H6 /O2 2.9 × 10 1.0-Au/Ti-SiO2 -2 403 10/10/10 H2 /C3 H6 /O2 1.4 × 10−7 −0.99 1.0-Au/SiO2 -Ti-5 423 10/10/10 H2 /C3 H6 /O2 1.1 × 10−6 −1.01 1.0-Au/Ti-SiO2 -10 423 Ti-SiO2 -10 473 c 10/10/10 H2 /C3 H6 /O2 1.2 × 10 −6 −0.94 10/10/0.2 H2 /C3 H6 /O2 3.3 × 10−6 −0.91 −1 a. GHSV 10000 mL·g−1 cat h . b. formation rate when CO was not introduced c. catalyst re-prepared after the support was aged for 9 months −6 0.05−Au/Ti−SiO2−1 0.05−Au/Ti−SiO2−1 1.0−Au/Ti−SiO2−2 1.0−Au/SiO2−Ti−5 1.0−Au/Ti−SiO2−10 Ti−SiO2−10 −7 10 rC H 3 8 (mol⋅g−1 ⋅s−1) cat 10 −8 10 1 10 2 3 10 10 4 10 PCO (ppm) Figure 4.18: Inhibition effect of CO on propene hydrogenation over the Ti-SiO2 -10 support (473 K, gas feed 10 vol% H2 , 10 vol% C3 H6 and 0.2 vol% O2 in helium, CO concen−1 tration varied, GHSV 10000 mL·g−1 cat h ; for the testing conditions for the gold catalysts, see Figure 4.8 or Table 4.1). What caused deactivation in propene hydrogenation If one looks at period II in Figure 4.17 or Figure 4.12 after restoring the activity by H2 /O2 treatment, a question may arise. That is: even though the formation rate of propane may be sensitive to the amount of hydroxyls or water on the surface, what causes a fast de- 108 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE activation at a constant temperature where the population of surface hydroxyl or water molecules should be relatively constant. In Figure 4.15, the deactivation can also be observed at a constant temperature below 443 K and is slower than the deactivation rate above 443 K. In order to answer this question, experiments were performed as shown in Figure 4.19 over the Ti-SiO2 -10 support. The original thought behind was that propene may form propoxy species on the surface and thus deactivates the hydrogenation activity as proposed for the propene hydrogenation over H-ZSM-5 [24, 25]; or that H2 may further reduce the active species causing deactivation; or that deactivation is only due to dehydroxylation or dehydration. The time-on-stream performance of cycles 2, 4 and 6 (interval flush with C3 H6 ) in Figure 4.19(b) confirmed that propene caused the deactivation in propene hydrogenation. It can be seen that flush in propene lowered the hydrogenation activity significantly. Another interesting phenomenon not shown here is that when CO and C3 H6 were co-fed during the interval flushing, the hydrogenation activity could not be lowered. Flush in pure helium or 10 vol% H2 as shown in Figure 4.19(a) did not lead to an obvious change in activity. This is in accordance with our experience that storage under ambient conditions or flush in inert atmosphere does not lower the hydrogenation performance of the activated catalysts or supports. Since CO can effectively suppress propene hydrogenation in a reversible way, it is very likely that CO adsorbs on or interacts with the active sites for propene hydrogenation much stronger than C3 H6 . But this contradicts the hypothesis that propene may form propoxy species causing deactivation in hydrogenation activity since this form of deactivation should be irreversible and should prevail over the reversible effect of CO. That is to say, the formation of propoxy species cannot explain at the same time these phenomena: the reversible effect of CO, a constant formation rate of propane when CO is fed as illustrated by Figure 4.14, and the deactivating effect of C3 H6 . Further investigation was performed by means of IR spectroscopy. Results are given in Figure 4.20. The Ti-SiO2 -10 support was first activated at 423 K to obtain a high activity in propene hydrogenation. The sample was then investigated in-situ by IR after prolonged dehydroxylation and contact with C3 H6 at 473 K, which mimicked the conditions in Figure 4.19(b). The extent of dehydroxylation at 473 K in vacuum for a prolonged period was very limited as seen from spectra a–c in Figure 4.20(A). However, after contacting with C3 H6 , dehydroxylation was greatly facilitated (spectra d, e, f) as evidenced by the decrease of intensity at 3480 cm−1 . Further evacuation in vacumm made no difference on the IR spectra as seen by the overlapped spectra f–h. On the 109 4.3. RESULTS AND DISCUSSION 1.6 x 10 6 1.2 0.8 rC H 3 8 (mol⋅gcat1⋅s 1) (a) 0.4 0 0 1.6 30 x 10 60 0 120 me (m n) 150 180 210 60 0 120 me (m n) 150 180 210 6 1.2 0.8 rC H 3 8 (mol⋅gcat1⋅s 1) (b) 0.4 0 0 30 Figure 4.19: Effect of flush in He, H2 and C3 H6 on the deactivation of propene hydrogenation over the Ti-SiO2 -10 support. Before each cycle, the support was calcined in 10 vol% O2 at 573 K for 1 hour, then was treated in 2 vol% H2 + 2 vol% O2 at 423 K for 1 hour, and then was kept at 423 K for 0.5 hour; afterwards, the temperature was raised to 473 K with a ramping rate of 3 K/min in He. During propene hydrogenation at 473 K, the feed gas was 10 vol% H2 + 10 vol% C3 H6 balanced by He. Flushing in He, or 10 vol% H2 in He, or 10 vol% C3 H6 in He was performed between t = 30 − 90 min and between t = 120 − 180 min. Cycle 5 in (a) was a cycle without interval flushing and was −1 used as a reference. GHSV was 10000 mL·g−1 cat h . All cycles were performed within a short period of time in a sequential order as indicated by the number. other hand, no adsorption in the CH stretching or bending region was observed, which excludes the formation of propoxy species on the support. It appears that the only effect of C3 H6 on the deactivation of hydrogenation activity is the accelerated dehydroxylation. This in turn confirms the importance of surface hydroxyls in propene hydrogenation over Ti-containing oxides as discussed in the previous section. 110 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE Adsorbance (a.u.) 0.05 0.05 e 3200 2800 2400 2000 1600 d a-c f,g,h (A) 3800 3600 3400 3200 3000 2800 -1 Wavenumber (cm ) 0.005 Adsorbance (a.u.) b c d e (B) f 3800 3600 3400 3200 3000 2800 1800 1600 1400 -1 Wavenumber (cm ) Figure 4.20: IR spectra of Ti-SiO2 -10 at 473 K. The support was pretreated ex-situ in 2 vol% H2 + 2 vol% O2 at 423 K for 1 hour and then transferred into IR cell. Spectra a: 5 min in vacuum; b: 30 min in vacuum; c: 60 min in vacuum; d: after c, in 17.5 mbar C3 H6 for 30 min and after subsequent evacuation; e: after d, in 17.5 mbar C3 H6 for 30 min and after subsequent evacuation; f: after e, in 17.5 mbar C3 H6 for 30 min and after subsequent evacuation; g: after f, 30 min in vacuum; h: after f, 60 min in vacuum. In section A, spectra f, g, and h overlapped. Section B gives the difference spectra of b–f subtracted from spectrum a. 4.4 Summarizing discussion It was observed in this study that the formation rate of propane kept increasing from zero until it leveled off during the activation of the catalysts while PO formation rate was stable at a high level from the very beginning as shown in section 4.3.1. The results in 4.4. SUMMARIZING DISCUSSION 111 this chapter imply that propene hydrogenation and epoxidation may happen on two different sites. The size effect of gold nanoparticles is undoubtedly important in hydrogen dissociation during the direct propene epoxidation [1, 5, 26], since Au–Ti4+ interface is indispensable for the direct propene epoxidation. Carbon monoxide might be competitively adsorbed on gold where hydrogen adatoms locate and thus abate the propene hydrogenation. It could also be the situation that Ti−OH induces hydrogen dissociation (on gold or with Ti) while carbon monoxide may poison the Ti−OH group probably by forming surface formate species, or the Ti site if the stability of possible titanium species at reduced oxidation states is an issue [27]. It has been observed by Sykes et al. [21, 22] that hydroxylated titania surface under partial reduction is capable of hydrogenating a terminal alkene. It has also been proposed that transition metals anchored on SiO2 can hydrogenate alkene probably by forming hydride complexes [23]. The catalytic performance of our catalyst, and one step further, of the hydroxylated support only, in hydrogen and propene but without oxygen clearly shows that the surface hydroxyl, most likely Ti– OH, plays an important role in propene hydrogenation and that gold is not of necessity for propane formation (section 4.3.3). The enhancement of propane formation by oxygen observed by Qi et al. [5] may be partly attributed to water formation in presence of oxygen and accordingly the hydration or hydroxylation of catalyst surface. Oxygen vacancy must be present on the supports in the study performed in this chapter, which facilitates the dissociation of H2 , since propene hydrogenation can be greatly enhanced by the presence of O2 and the activation process needs O2 . On the other hand, it is generally accepted that tripodal Ti4+ sites are favoured in liquid-phase epoxidation when concerning the coordination environment of titanium in silica-supported titanium catalysts [28–30]. Hydrothermal instability [31] of such tripodal Ti sites leads to the formation of bipodal and monopodal Ti sites on silica surface, which, combined with vicinal silanols, might contribute to propene hydrogenation in the complex system of propene, hydrogen and oxygen over gold–titania catalysts. Our experimental observations suggest that on the supports where Ti−O−Si or Ti−O−Ti linkage is not that resistant to hydrolysis, the role of support in propene hydrogenation should be taken into account. On the other hand, the temperature of ca. 443 K, at which the hydrogenation activity of the catalysts and supports investigated in this study reach the peak during TPR in only H2 and C3 H6 , is very close to the temperature for dehydroxylation of Ti−OH over TiO2 supported on SiO2 [32]. This temperature has also been observed as the dehyroxylation or dehydration temperature of the Ti-SiO2 supports in this study as shown in Chapter 3. 112 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE Primet et al. [33] indicated that for incompletely coordinated titanium atoms in TiO2 (the structure below), coordination-bonded water is removed at 423 K. Depending on the distance between the two ajacent Ti atoms, C3 H6 may facilitate the removal of the two hydroxyls as found in experiments shown in Figures 4.19 and 4.20. This dehydroxylation may be via the C− −C double bond forming a π complex with one OH and an interaction between the allylic hydrogen and the other –OH. However, checking with C2 H4 has not been performed. In this sense, the role of CO in switching off propene hydrogenation can also be the competitive adsorption on the Ti−OH or defective TiO x sites against C3 H6 . Regarding to the low hydrogenation activity on the 0.05-Au/TS-1 catalyst, there may be also some incompletely coordinated Ti atoms since the TS-1 support used here has a very small crystal size of 120 nm on average (see Figure 3.6). This structure shown above is also similar to a structure named strained siloxane in thermally activated SiO2 , which can also hydrogenate alkenes as found by Rajagopal et al. [34] at 473–623 K. In his study, it was found that silica gained hydrogenation activity after heating in Ar between 603 and 703 K for hours. However, blank experiments on pure silica (either hydroxylated or calcined) used in this study did not show any activity. Propene can also be hydrogenated by TiO2 P25 when TiO2 is reduced above 573 K. The hydrogenation activity on reduced TiO2 can also be suppressed when CO is present. But the appearance of peak activity at ca. 443 K on the Ti-SiO2 supports was not found on reduced TiO2 . It seems that defective TiO x and hydroxyls are necessary for propene hydrogenation over the Ti-SiO2 supports. Gold may play as an extra source for H2 dissociation but is not necessary for propene hydrogenation. 4.5 Conclusions Propene hydrogenation over the gold–titania catalysts can also proceed over the supports and can be switched off by CO without affecting the formation of propene oxide. The presence of O2 enhances propene hydrogenation over the catalysts and supports. Temperature programmed reaction in H2 and C3 H6 shows a peak activity at ca. 443 K, which corresponds to dehydroxylation or dehydration of the support. Activity decrease REFERENCES 113 in propene hydrogenation is due to accelerated dehydroxylation in the presence of C3 H6 . References [1] T. Hayashi, K. Tanaka and M. Haruta, J. Catal., 1998, 178, 566–575. [2] E. E. Stangland, K. B. Stavens, R. P. Andres and W. N. Delgass, J. Catal., 2000, 191, 332–347. [3] G. Mul, A. Zwijnenburg, B. Van der Linden, M. Makkee and J. A. Moulijn, J. Catal., 2001, 201, 128–137. [4] T. A. Nijhuis, B. J. Huizinga, M. Makkee and J. A. Moulijn, Ind. Eng. Chem. Res., 1999, 38, 884–891. [5] C. Qi, J. Huang, S. Bao, H. Su, T. Akita and M. Haruta, J. Catal., 2011, 281, 12–20. [6] J. Gaudet, K. Bando, Z. Song, T. Fujitani, W. Zhang, D. Su and S. Oyama, J. Catal., 2011, 280, 40–49. [7] L. Cumaranatunge and W. N. Delgass, J. Catal., 2005, 232, 38–42. [8] B. Taylor, J. Lauterbach and W. N. Delgass, Appl. Catal., A, 2005, 291, 188–198. [9] J. Lu, X. Zhang, J. J. Bravo-Suárez, T. Fujitani and S. T. Oyama, Catal. Today, 2009, 147, 186–195. [10] J. Huang, T. Takei, T. Akita, H. Ohashi and M. Haruta, Appl. Catal., B, 2010, 95, 430–438. [11] T. A. Nijhuis, T. Visser and B. M. Weckhuysen, J. Phys. Chem. B, 2005, 109, 19309–19319. [12] B. Chowdhury, J. J. Bravo-Suárez, N. Mimura, J. Lu, K. K. Bando, S. Tsubota and M. Haruta, J. Phys. Chem. B, 2006, 110, 22995–22999. [13] B. Taylor, J. Lauterbach, G. E. Blau and W. N. Delgass, J. Catal., 2006, 242, 142–152. [14] J. Lu, X. Zhang, J. J. Bravo-Suárez, S. Tsubota, J. Gaudet and S. T. Oyama, Catal. Today, 2007, 123, 189–197. [15] J. Chen, S. J. A. Halin, J. C. Schouten and T. A. Nijhuis, Faraday Discuss., 2011, 152, 321–336. [16] E. Sacaliuc, A. M. Beale, B. M. Weckhuysen and T. A. Nijhuis, J. Catal., 2007, 248, 235–248. [17] B. Taylor, J. Lauterbach and W. N. Delgass, Catal. Today, 2007, 123, 50–58. [18] W. Yan, B. Chen, S. M. Mahurin, E. W. Hagaman, S. Dai and S. H. Overbury, J. Phys. Chem. B, 2004, 108, 2793–2796. [19] W. T. Wallace, B. K. Min and D. W. Goodman, J. Mol. Catal. A: Chem., 2005, 228, 3–10. [20] X. Gao and I. E. Wachs, Catal. Today, 1999, 51, 233–254. [21] E. C. H. Sykes, M. S. Tikhov and R. M. Lambert, Catal. Lett., 2002, 78, 7–11. [22] E. C. H. Sykes, M. S. Tikhov and R. M. Lambert, J. Phys. Chem. B, 2002, 106, 7290–7294. [23] Yu. I. Yermakov, Y. A. Ryndin, O. S. Alekseev, D. I. Kochubey, V. A. Shmachkov and N. I. Gergert, J. Mol. Catal., 1989, 49, 121–132. [24] G. Spoto, S. Bordiga, G. Ricchiardi, D. Scarano, A. Zecchina and E. Borello, J. Chem. Soc., 114 CHAPTER 4. SWITCHING OFF PROPENE HYDROGENATION IN THE DIRECT EPOXIDATION OF PROPENE Faraday Trans., 1994, 90, 2827–2835. [25] A. Zheng, S.-B. Liu and F. Deng, Microporous Mesoporous Mater., 2009, 121, 158–165. [26] T. Fujitani, I. Nakamura, T. Akita, M. Okumura and M. Haruta, Angew. Chem. Int. Ed., 2009, 48, 9515–9518. [27] Y. Li, B. Xu, Y. Fan, N. Feng, A. Qiu, J. He, J. Miao, H. Yang and Y. Chen, J. Mol. Catal. A: Chem., 2004, 216, 107–114. [28] M. Crocker, R. H. M. Herold, A. G. Orpen and M. T. A. Overgaag, Dalton Trans., 1999, 3791– 3804. [29] J. K. F. Buijink, J. J. M. van Vlaanderen, M. Crocker and F. G. M. Niele, Catal. Today, 2004, 93-95, 199–204. [30] J. M. Fraile, J. I. García, J. A. Mayoral and E. Vispe, J. Catal., 2005, 233, 90–99. [31] J. Zhou, Z. Hua, X. Cui, Z. Ye, F. Cui and J. Shi, Chem. Commun., 2010, 46, 4994–4996. [32] S. Haukka, E. L. Lakomaa and A. Root, J. Phys. Chem., 1993, 97, 5085–5094. [33] M. Primet, P. Pichat and M. V. Mathieu, J. Phys. Chem., 1971, 75, 1216–1220. [34] V. K. Rajagopal, R. D. Guthrie, T. Fields and B. H. Davis, Catal. Today, 1996, 31, 57–63. How metallic is gold in the direct epoxidation of propene : An FTIR study 5 This chapter has been published as: J. Chen, E. A. Pidko, V. Ordomsky, M. W. G. M. Verhoeven, E. J. M. Hensen, J. C. Schouten, & T. A. Nijhuis, How metallic is gold in the direct epoxidation of propene: An FTIR study, Catal. Sci. Technol., 2013, DOI: 10.1039/C3CY00358B. Abstract Unraveling the oxidation state of gold is important to understand the role of gold in the direct propene epxoidation on the gold–titania catalysts. Fourier transform infrared study of low-temperature carbon monoxide adsorption was performed over Au/TiO2 and Au/Ti-SiO2 under atmospheres of the reaction mixture, oxygen, hydrogen, and propene, respectively. Data reveals that the active gold sites treated by the reaction mixture are fully covered by reaction intermediates and deactivating species. Oxidation at 573 K removes these carbonaceous species on gold. Oxygen adsorption at reaction temperatures leads to positively charged gold, which can be reduced to metallic gold in the presence of hydrogen. Propene plays as an electron donor to gold atoms resulting in negatively charged gold with the carbonyl band at 2079 cm−1 . The results in this study may provide a general scheme of electron transfer via gold on the gold–titania catalysts for the direct propene epoxidation. 116 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE 5.1 Introduction Gold nanoparticles in the direct propene epoxidation over gold–titania catalysts are considered to have two main functions. The first is the formation of the active oxidizing species, the hydro-peroxy species (HOOH, OOH), which has been spectroscopically confirmed via an inelastic neutron scattering (INS) study [1]. From a combination of in-situ UV–Vis and XANES study [2], it was suggested that the hydro-epoxidation of propene over gold–titania catalysts involves the formation of HOOH on gold and a sequential transfer of HOOH to Ti4+ forming the real active intermediate Ti−OOH. This route resembles the chemistry of propene epoxidation by H2 O2 over TS-1 in the HPPO process [3]. The second function of gold in the epoxidation of propene, which is less prominent, is that propene may be activated on gold nanoparticles [4], or in another sense, that the adsorption of propene on gold strongly affects the activity of gold in the formation of hydro-peroxy species, since the inhibiting effect of the alkene in the hydrogen oxidation on gold catalysts is generally observed [5]. The rate-determining step in the hydro-epoxidation of propene on gold–titania catalysts is considered to be the dissociative adsorption of hydrogen [6–8]. The study by Boronat et al.[9] suggests that the hydrogen dissociation over Au/TiO2 occurs only on low coordinated and neutral gold atoms at corner or edge sites of gold nanoparticles but not directly bonded to the support. The H2 –D2 exchange reaction studied by Fujitani et al.[10] on model Au/TiO2 (110) catalysts with gold nanoparticles of controlled sizes has shown a constant turnover frequency of HD formation when based on the perimeter length of gold nanoparticles, indicating that hydrogen dissociation on gold–titania catalysts may very likely be an interfacial phenomenon. Green et al.[11] studied lowtemperature hydrogen oxidation over Au/TiO2 by infrared (IR) spectroscopy and density functional theory (DFT) calculations. Their results suggest a pathway of hydrogen dissociation at the perimeter sites at the interface between Au and TiO2 , in which oxygen molecules adsorb on Ti5c perimeter sites and hydrogen dissociates into Au–H and Ti– OOH in the first step. Yang et al.[12] performed DFT study of hydrogen dissociation and diffusion at the perimeter sites of Au/TiO2 (110) and depicted two ways of hydrogen dissociation. In their study, the heterolytic dissociation of hydrogen with one H atom on gold and another on the bridge oxygen atom in the support is energetically more favourable, while the homolytic hydrogen dissociation mainly occurs after all Au–O–Ti sites are passivated into Au–O(H)–Ti. Using coadsorbed CO in an IR study of hydrogen dissociation 117 5.1. INTRODUCTION over Au/TiO2 at room temperature, Panayotov et al.[13] revealed that the most active sites for hydrogen dissociation are defect Au0 sites away from Au–O–Ti interface. Passivation of Au–O–Ti to Au–O(H)–Ti was also observed. The rate of hydrogen atom spillover 1/2 to the support was determined to be proportional to PH . The coadsorption (or com2 petitive adsorption) of CO does not change the chemistry of hydrogen dissociation and spillover, but it can suppress the initial rate of hydrogen dissociation by a factor of 2.6 as demonstrated by Panayotov et al. [13]. In summary, hydrogen dissociation occurs on low-coordinated Au0 atoms at edges and corners and the role of the support cannot be neglected over gold–titania catalysts. Another interesting phenomenon in the hydro-epoxidation of propene over gold– titania catalysts is that gold can catalyze the propene hydrogenation [14–17]. Although in our previous study [18], it has been demonstrated that the support itself can contribute significantly to the propene hydrogenation and that the presence of a small amount of CO (10 ppm to 1000 ppm) can switch off the propene hydrogenation, we cannot exclude contribution from gold. Gold-catalyzed alkene hydrogenation has long been known [19]. Specifically over the gold–titania system, the explanation to this side reaction can be that propene hydrogenation is sensitive to the structures of gold nanoparticles or clusters [14, 16, 20]. It can also be that Au+ exists and functions as proposed by Oyama et al. [17]. The Au+ in their study is in the form of Au(CN)–1 2 resulting from NaCN leaching of Au/TS-1 and has no activity in propene epoxidation, but is active in hydrogenation. Bond and Thompson[21] suggested that gold atoms with a low coordination number are more prone to be oxidized than those with a high coordination number. The findings by Oyama and coworkers[17] thus provide another mechanistic view towards the hydrogenation of propene, in which small gold clusters have probably oxidized interphase atoms. Nevertheless, Au+ should be neither easy to be produced nor stable in the reaction conditions for the direct propene hydrogenation where dihydrogen and propene are also present. Dekkers et al.[22] observed Au+ by pre-oxidizing Au/TiO2 in pure dioxygen at 573 K for one hour and found that Au+ was gradually reduced to Au0 in a CO–O2 mixture at room temperature. Venkov et al.[23] could not obtain Au+ by treating Au/TiO2 under oxygen at 573–773 K and only observed Au+ after oxidation in NO+O2 at 773 K. Thus it is doubtful that Au+ would exist under a H2 /O2 /C3 H6 mixture at 323–473 K. The most informative technique used in gold catalysis to determine gold oxidation state is using CO as the probe molecule to observe the frequency of ν(CO) bands by the IR spectrum. The sites on gold nanoparticles, where CO adsorbs, are also those gold 118 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE atoms with low coordination numbers at edges and corners, where all the chemistry is supposed to happen. Spectral regions of surface carbonyls on gold with different status are well summarized in reviews by Freund et al.[24] and Mihaylov et al.[25]. Though some of the bands overlap with CO interacting with Ti4+ and hydroxyls, they can be distinguished by progressive adsorption/desorption. The aim of this work is to have a general image of the oxidation state of gold in Au/Ti-SiO2 and Au/TiO2 which are active in the direct propene epoxidation. The catalysts were investigated by low-temperature CO adsorption after different pretreatment or in different atmospheres. We also have the two following questions left from previous studies [5, 18]. The first is that what may be the reason for the suppressing effect of propene adsorption on water formation over gold. The second is that to what extent can we observe the competitive adsorption of CO and H2 if the inhibiting effect of CO on propene hydrogenation is due to CO occupying gold. Co-adsorption of CO and H2 /C3 H6 was thus investigated and discussed in this study. 5.2 Experimental 5.2.1 Catalyst preparation and testing The catalyst consisting of 1 wt% gold on titania dispersed on silica, which has been fully investigated in our earlier study [8], was used in this study and denoted as Au/Ti-SiO2 (SiO2 , Davisil 643, Aldrich, 300 m2 /g, pore size 150 Å, pore volume 1.15 cm3 /g). The preparation of 1 wt% Au/SiO2 (Davisil 643, Aldrich) and 1 wt.% Au/TiO2 (P25, Degussa, 70% anatase, 30% rutile, 45 m2 /g) followed the deposition–precipitation method using ammonia described in our earlier study [4]. Safety concerns and suggestions on the possible formation of explosive fulminating gold when using ammonia have been addressed earlier [4]. Due to the limited amount of gold and ammonia, the risks are very minor in this study. A total of 2 g of the support was dispersed in 100 mL of water. The pH of the slurry was adjusted to 9.5 by dropwise adding ammonia (2.5 wt%). A total of 115 mg of an acidic 30 wt% HAuCl4 solution (Aldrich, 99.99% trace metal basis) was diluted in 20 mL of demineralized water and was added dropwise to the support slurry over a 15 min period. While HAuCl4 solution was being added, the pH was kept at 9.5 using aqueous ammonia. After the addition of the gold solution, the slurry was stirred for one hour. The slurry was filtered and washed 3 times using 200 mL of water. The catalyst 5.2. EXPERIMENTAL 119 was dried overnight at 353 K and calcined first at 393 K (5 K/min heating) for 2 h and afterwards at 673 K (10 K/min heating) for 4 h. Drying and calcination of the support and the catalyst were performed under atmospheric pressure in stationary air. Catalytic tests were performed in a flow setup equipped with a fast Interscience Compact GC system (3 min analysis time) containing a Porabond Q column and a Molsieve 5A column in two separate channels, each with a thermal conductivity detector. 300 mg of catalyst was loaded into the tubular quartz reactor (6 mm inner diameter) and tested with a gas feed rate of 50 mL min−1 in total consisting of 10 vol.% each for hydrogen, −1 oxygen, and propene with helium as the balance (GHSV 10000 mL·g−1 cat h ). The term ‘spent’ or ‘after reaction’ hereinafter means that the catalyst had been tested in the reaction mixture for epoxidation for at least 2 hours. The term ‘regenerated’ hereinafter means that the catalyst or sample had been calcined in 5 vol.% or 10 vol.% oxygen diluted in helium at 573 K for at least 30 min. When mentioned in an IR experiment, the regeneration was always conducted in-situ in 5 vol.% oxygen at 573 K for 30 min. 5.2.2 Charaterization techniques Transmission electron microscope (TEM) images were recorded with a FEI Tecnai G2 Sphera transmission electron microscope at an acceleration voltage of 200 kV. Loadings of gold and titanium were determined by inductively coupled plasma optical emission spectrometry (ICP-OES) using a Spectra CirosCCD system. In ICP analyses, gold was dissolved with aqua regia and grafted titanium was etched by 5 mol/L H2 SO4 solution. The H2 SO4 solution containing dissolved titanium was then diluted to 2.9 mol H2 SO4 /L for analyses. The X-ray photoelectron spectroscopy (XPS) measurements were carried out with a Kratos AXIS Ultra spectrometer, equipped with a monochromatic X-ray source and a delay-line detector (DLD). Spectra were obtained using the aluminium anode (Al Kα = 1486.6 eV) operating at 150W. A detachable U-shape Pyrex tubular reactor (6 mm inner diameter) with valves on both ends was used for post-reaction catalysts. The U-shape reactor can be heated using an aluminum heating jacket up to 573 K. The post-reaction catalysts were then flushed in helium and afterwards sealed in the U-shape reactor. The reactor was transferred into a nitrogen glove box free from oxygen (< 2 ppm) and moisture (< 0.5 ppm). The post-reaction samples for XPS measurements were prepared in the glove box. Transport of the samples from the glove box to the spectrometer was performed in an inert atmosphere by using a small nitrogen-purged chamber equipped with 120 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE a magnetic arm. In this way, the post-reaction catalysts were kept ‘in-situ’. XPS spectra were referenced to the C 1s line at 284.9 eV. Infrared spectra were recorded with a Bruker IFS 113v spectrometer at a 2 cm−1 optical resolution and accumulation of 128 scans (ca. 2 min of acquisition time per spectrum). The samples were pressed in self-supporting discs at 0.5 MPa (diameter: 12.7 mm, ca. 7.9 mg cm−2 ). The sample wafer was placed in a homemade doublewalled IR cell [26]. The high-vacuum (below 10−5 mbar) transmission IR cell was inhouse made and can be cooled by liquid nitrogen and heated by electric wire. The IR cell is connected to a gas dosing system, through which the progressive adsorption of adsorbates can be performed via a 5-µL or 50-µL sample loop on a Valco six-port valve. In a typical CO dosing experiment, the CO pulse was automatically carried out upon the acquisition of each spectrum (128 scans) finished. The stabilization of one spectrum was fast and normally within a few scans. An extra pair of inlet and outlet is located on top of the IR cell, through which in-situ calcination or pretreatment under different atmospheres can be performed. The gas for sample pretreatment can be dried by a liquid nitrogen trap when necessary. The catalysts were tested in the tube reactor confirming the activity before IR experiments. The catalyst discs were generally used for 3-4 sequential IR measurements. In-situ calcination in 5 vol.% oxygen diluted by helium at 573 K for 30 min followed by evacuation at 573 K for 30 min was carried out between each IR experiment. 5.3 Results 5.3.1 Catalyst performance and characterization The 1 wt.% Au/TiO2 and Au/Ti-SiO 2 catalysts were tested for propene epoxidation in hydrogen and oxygen. Their catalytic performance was listed in Table 5.1. The conversion of propene and the selectivity to PO were in the typical range for these two catalysts at these conditions [4]. The side products were mainly propionaldehyde, acetone, acetaldehyde and carbon dioxide. No propane formation was found. The size distribution of gold nanoparticles was narrow for both catalysts (spent), centered at 4 nm for Au/TiO2 and 4.5 nm for Au/Ti-SiO 2 . The catalyst stability was investigated for these two catalysts by repeatedly performing catalytic testing followed by 1-hour regeneration at 573 K in 10 vol.% oxygen. The results are shown in Figure 5.1. The time-on-stream performance 121 5.3. RESULTS Table 5.1: Charaterization and general performance of three gold catalysts Sample ID Au/Ti-SiO2 Au/TiO2 Au/SiO2 Au loading (wt.%) Ti loading (wt.%) Au size (nm) 0.91 0.93 0.95 1.29 – – 4.5±1.1 4.0±1.2 3.1±0.8 Temp. (K) 423 333 Catalytic performancea C3 H6 conv. PO sel. H2 eff. b (%) (%) (%) 1.32 85.5 7.1 0.20 98.0 15.9 – a. no activity in propene hydrogenation observed, activity taken at 2 h, H2 :O2 :C3 H6 :He= 1 : 1 : 1 : 7, −1 GHSV 10000 mL·g−1 cat h . b. determined as rPO /rH O. 2 of both catalysts was highly reproducible. The reason for these stability tests was to exclude a possible catalyst change in the IR experiments since each disc was re-used for 3–4 times. The effect of high vacuum at 573 K after calcination in oxygen in the IR cell on possible gold sintering was examined by repeatedly treating one Au/TiO2 sample under such condition before cooling down to 323 K in vacuum for 9 times. The average size of gold particles after such treatment increased slightly by less than 0.5 nm, but it was still well below 5 nm as shown in Figure 5.2 and the size distribution remained narrow. Thus sintering of gold nanoparticles in IR experiments for each pellet was not supposed to occur or had very limited effect on CO adsorption. 5.3.2 CO adsorption on catalysts after epoxidation and regeneration Figure 5.3 shows the changes in IR spectra during progressive CO adsorption on a clean Au/SiO2 sample at 90 K. The spectrum of the dehydrated sample before CO dosing is quite simple and similar to what can be observed for SiO2 . In the OH stretching region, the band at 3743 cm−1 is assigned to unassociated Si−OH. The band at 3715 cm−1 is assigned to the terminal silanol in a pair or chain of hydrogen-bonded hydroxyls (hydrogen perturbed OH), while the broad band centered at 3552 cm−1 is assigned to those within the hydrogen-bonded hydroxyl groups (oxygen perturbed OH) [27]. The adsorption at 3653 cm−1 is from inaccessible Si−OH. The adsorption at 1980 cm−1 and 1875 cm−1 is attributed to the skeleton vibrations of silica [28]. The 1640 cm−1 band is most likely to be assigned to the bending mode of some remaining adsorbed water. Upon adsorption, the CO molecules interact with terminal OH groups leading to a rise of the perturbed band at ca. 3550 cm−1 , which gradually shifted to lower wavelengths when CO coverage increased. CO also interacted with unassociated OH groups giving the decreased 122 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE x 10 −7 −6 4 (a) cycle 1 cycle 2 cycle 3 cycle 4 cycle 5 Au/Ti−SiO , PO 2 2 3 2 2 1 0 2 1 0 60 x 10 120 180 Time (min) 240 0 300 0 −7 60 120 180 Time (min) 240 300 −6 2 (c) x 10 cycle 1 cycle 2 cycle 3 Au/TiO2, PO (d) (mol⋅g−1 ⋅s−1) cat 1.5 1.5 1 2 O 1 cycle 1 cycle 2 cycle 3 Au/TiO2, H2O rH rPO (mol⋅g−1 ⋅s−1) cat 2 O 2 cycle 1 cycle 2 cycle 3 cycle 4 cycle 5 Au/Ti−SiO , H O rH rPO (mol⋅g−1 ⋅s−1) cat 3 x 10 (b) (mol⋅g−1 ⋅s−1) cat 4 0.5 0 0.5 0 30 60 Time (min) 90 120 0 0 30 60 Time (min) 90 120 Figure 5.1: Tests of activity stability for the 1 wt.% Au/Ti-SiO2 (top, a and b) and Au/TiO2 (bottom, c and d) catalysts in 10/10/10/70 H2 /O2 /C3 H6 /He mixture. GHSV −1 10000 mL·g−1 cat h , 423 K for Au/Ti-SiO 2 , 333 K for Au/TiO2 . A few hours of continuous testing in the reaction mixture followed by 1 hour calcination in 10 vol.% oxygen at 573 K constitutes one ‘cycle’. band at ca. 3746 cm−1 (shifted from 3743 cm−1 before CO was introduced) and lowered transparency at ca. 3610 cm−1 . The isosbestic point located at ca. 3680 cm−1 . Three carbonyl bands were observed after introducing CO. The band at 2158 cm−1 is known for CO interacting with OH groups. The band at 2136 cm−1 can be assigned to physisorbed CO [29, 30]. The 2136 cm−1 band is unlikely to be assigned to P-branch adsorption of gas phase CO. The contribution from gas phase CO at 1 mbar and 90 K was found to be very minor in our study. Besides, the R-branch component at higher frequencies is lacking as seen for Au/SiO2 . The band at 2099 cm−1 is assigned to Au0 −CO, which shifted from 2105 cm−1 at low CO coverage. A red shift of ν(CO/Au) with increased coverage of CO was generally observed and can be explained by the balancing between 123 5.3. RESULTS 50 (a) Counts 40 30 20 10 0 1 60 2 3 4 5 6 7 Particle size (nm) 8 9 10 2 3 4 5 6 7 Particle size (nm) 8 9 10 (b) Counts 50 40 30 20 10 0 1 Figure 5.2: Change in the size distribution of gold nanoparticles on the 1 wt.% Au/TiO2 catalyst after repeating 1-hour calcination in vacuum at 573 K in the IR cell and cooling down to 323K for 9 times: (a) before treatment; (b) after treatment. dipole–dipole coupling, which leads to a blue shift, and the chemical shift, which leads to a wavenumber decrease since the donation from the weakly antibonding 5σ orbital dominates in Au–CO interaction whereas the backbonding from gold atoms to 2π* CO orbital is lacking [31–33]. Progressive adsorption of CO on the spent and regenerated Au/Ti-SiO 2 sample is shown in Figure 5.4. Comparing to CO adsorption on Au/SiO2 , a new band above 2180 cm−1 appeared as seen from Figure 5.4, which can be assigned to CO adsorbed on Ti4+ ions [29]. The spent catalyst showed very weak adsorption of CO on both Ti4+ and Au. The weak band of linear carbonyl on Au0 at 2095 cm−1 was blue shifted by 12 cm−1 to 2107 cm−1 on the regenerated sample, on which propoxy species adsorbed at Au–Ti interface are supposed to be fully removed. After regeneration, the sample showed a much stronger adsorption of CO on gold indicating that during reaction the gold sites 124 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE m (A) a s. 0 0.1 3000 3610 (C) CO-OH 0.03 2099 3552 CO-Au0 m 0.02 2136 3653 3 43 1500 0.04 a m 0 n vacuum 0.01 a 0 .05 3800 2000 1 (B) 3 15 s. 2500 avenum er (cm 0.05 2158 3500 3700 3600 3500 avenum er (cm−1) 3400 2200 2150 2100 2050 Figure 5.3: Development of IR spectra during progressive CO adsorption on the 1 wt.% Au/SiO2 sample at 90 K: (A) wide-frequency range; (B) range of ν(OH); (C) difference spectra in the range of ν(CO), substracted from the spectrum ‘0’ before CO dosing. The following CO pressures after each dose were used (in mbar, spectra a–m): 0.06, 0.14, 0.23, 0.30, 0.36, 0.43, 0.50, 0.58, 0.66, 0.73, 0.81, 0.90, 0.99. Prior to low-temperature CO adsorption, the catalyst was calcined in 10 vol.% O2 at 573 K and then tested in 2 vol.% H2 and 2 vol.% O2 at 423 K in the flowing reactor. Afterwards the pellet was prepared ex-situ and transferred into IR cell for evacuation at 473 K for 1 hour. active for epoxidation were fully covered by strongly adsorbed species though the water formation rate remained high as shown in Figure 5.1. Upon dosing CO on the regenerated sample as shown in Figure 5.4(B), Auδ+ can be detected by the shoulder of ν(CO) at 2125 cm−1 [23, 34, 35]. The spectra of the spent and regenerated sample before CO dosing are given and compared in Figure 5.5. The difference spectrum given in Figure 5.5 confirmed the presence of strongly adsorbed propoxy on the catalyst surface. In Figure 5.5(A), it is clear that alkoxy adsorbed on the support cannot be fully removed by oxidation at 573 K 125 5.3. RESULTS 2158 (A) 2158 0.04 0.04 (B) CO-OH CO-OH 0.03 0.03 0 0 2200 2150 2100 Wavenumber (cm 1 2050 2107 2125 0.01 n' a' 2185 CO-Ti a CO-Ti4+ CO-Au0 2095 0.01 4+ 2136 n 2136 0.02 0.02 2181 Abs. CO-Au0 CO-Auδ+ 2200 2150 2100 2050 Figure 5.4: Progressive CO adsorption on the (A) spent and (B) regenerated 1 wt.% Au/Ti-SiO2 catalyst. The spectra were taken at 90 K and the following CO pressures after each dose were used (in mbar): 0.06, 0.12, 0.19, 0.26, 0.33, 0.40, 0.46, 0.53, 0.61, 0.67, 0.74, 0.82, 0.89, 0.96 for a – n; 0.06, 0.12, 0.18, 0.25, 0.32, 0.39, 0.45, 0.53, 0.62, 0.68, 0.76, 0.83, 0.90, 0.97 for a′ – n′ . The sample in vacuum was used as the background. as indicated by the remaining bands in the CH stretching region after regeneration. The difference spectrum between the spent and regenerated sample given in Figure 5.5(B) did show increased transparency by regeneration in the the CH stretching and bending regions, which was due to the removed propoxy species similar to what was observed by Mul et al. [36]. In the CH stretching region, the bands at 2989 and 2979 cm−1 are assigned to νas (CH3 ), while the bands at 2938, 2906, 2881, and 2860 cm−1 can be assigned to νas (CH2 ), ν(CH), νs (CH3 ), and νs (CH2 ), respectively [4, 36, 37]. A weak band at 2825 cm−1 was observed, which may be assigned to ν(CH) of surface formate [38]. In the CH bending region, the bands at 1459/1452, 1382, and 1347 cm−1 can be assigned to δas (CH3 )/δ(CH2 ), δs (CH3 ), and δ(CH), respectively. The bands at 1719 and 1697 cm−1 are assigned to C− −O stretching vibration from surface species containing carbonyl group. Since the sample of the spent catalyst was already evacuated at 573 K, the carbonyl group is mostly likely from dehydrogenation of intermediates such as Ti−O−CH2 CH(OH)CH3 or Ti−O−CH(CH3 )CH2 OH (preferred, however no obvious adsorption of ν(CH) at 2720–2750 cm−1 for aldehydes) by the ring opening of propene oxide on the surface [36, 39, 40]. Less intensive bands at 1682, 1626, 1591, and 1423 cm−1 may be attributed to surface carbonates and carboxylate [41]. The spectrum given in Figure 5.6A provides information on species adsorbed on the spent Au/TiO2 catalyst. Figure 5.6B–C shows CO adsorption on the spent and regen- 126 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE 2884 2979 2990 regenerated 2938 Abs. 1461 spent 1453 1725 1380 (A) 1800 1700 1600 1500 1400 1300 0.1 3500 3000 2000 1500 -1 1297 1347 1452 1423 1459 1591 1382 1682 1626 1697 1719 2825 2881 2938 2860 Abs. 2906 2989 (B) 2979 Wavenumber (cm ) .005 3000 2800 1800 -1 1600 1400 Wavenumber (cm ) Figure 5.5: IR spectra of the 1 wt.% Au/Ti-SiO2 sample after reaction and regeneration: (A) wide-frequency range; (B) difference spectrum of the spent subtracted from the regenerated. Both spectra were taken at 90 K in vacuum. The sample of the spent catalyst −1 (2 hours in 10/10/10/70 H2 /O2 /C3 H6 /He mixture at 423 K, GHSV 10000 mL·g−1 cat h , ex-situ) was evacuated in vacuum at 573 K in the IR cell for 1 hour and then cooled down to 90 K. After CO adsorption was performed, the same sample was then calcined in flowing 5 vol.% O2 at 573 K in-situ for 0.5 hour to clean the carbonaceous surface, evacuated at 573 K for 0.5 hour and then cooled down to 90 K. erated Au/TiO2 catalyst. The left inset in Figure 5.6A compares the IR spectra of the spent Au/TiO2 sample after evacuation at 573 K (spectrum b) and the same sample after regeneration (spectrum c). In the left inset, the regenerated sample clearly shows OH stretching bands above 3500 cm−1 . However, there are hardly OH stretching bands observed on the spent catalyst but evacuated at 573 K. These OH groups on a clean or regenerated catalyst sample can hardly be completely removed by evacuation at even 673 K. As for the spent catalyst, evacuation at 573 K eliminated all the surface hydroxyl 127 5.3. RESULTS d 1100 1050 1000 a 1147 1227 1331 1303 1375 1723 1674 2716 1544 1448 1150 1576 2869 2970 2905 2982 Abs. 2932 1200 1123 a-d 3500 3000 2500 2000 1500 1000 1020 a c 1053 (A) 1090 b 0.02 3200 2800 1600 1200 -1 108 185 0 2150 2100 Wavenumber (cm 1 CO Au0 107 1 7 1 CO β-Ti4+ a' 0 2200 ' 2050 1 a 3 1 8 0.02 09 0.02 9 CO CO Au0 14( 190 Abs. 0.04 Abs. 0.04 179 (C) CO β-Ti 0.06 CO β-Ti4+ 4+ (B) -Ti4+ 0.06 178 Wavenumber (cm ) 2200 2150 2100 Wavenumber (cm 1 2050 Figure 5.6: Difference spectrum of the spent and regenerated sample of Au/TiO2 (section A) and progressive CO adsorption on the spent (section B, pre-evacuated at 573 K) and the regenerated (section C) sample. The left inset in section A compares the IR spectra of the spent and regenerated sample. The spent catalyst was under the reaction conditions ex-situ at 333 K for 1 hour after a 5-hour catalytic testing and regeneration. The sample was then evacuated at 323 K and 573 K in the IR cell. All spectra were taken at 90 K. The following CO pressures after each dose were used (in mbar): 0.01, 0.03, 0.05, 0.07, 0.09, 0.11, 0.13, 0.16, 0.18, 0.20 for a – j in section B; 0.01, 0.03, 0.05, 0.06, 0.08, 0.10, 0.13, 0.16, 0.18, 0.21 for a′ – j′ in section C. In section A, spectrum a is difference spectrum between the sample evacuated at 573 K (spectrum b) and the regenerated sample (spectrum c); spectrum d is difference spectrum between the sample evacuated at 323 K (not shown) and the regenerated sample (spectrum c). The green spectrum ‘a−d’ in the right inset is the subtraction of ‘a’ by ‘d’. 128 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE groups indicating that probably some adsorbed monodentate propoxy species reacted with remaining Ti−OH during heating. The right inset in Figure 5.6A shows the change of the spent sample in the Ti–O–C/C–C stretching region during heating. After heating the spent sample in vacuum at 573 K, the intensity decrease of two bands at 1135 and 1090 cm−1 was observed. As seen from Figure 5.6A (spectrum a), the spent catalyst was fully covered by bidentate propoxy species, carboxylates as well as aldehyde adsorbed on the surface. The presence of a bidentate propoxy species on the spent catalyst is indicated by the pattern of bands in the CH stretching region at 2970 [νas (CH3 )], 2932 [νas (CH2 )], 2905 [ν(CH)], and 2869 cm−1 [νs (CH3 )] together with two sharp bands in the Ti−O−C and C−C stretching region at 1123 and 1090 cm−1 [4, 37]. The band at 1123 cm−1 can be assigned to Ti−O−C stretching, while for PO or 1,2-propanediol adsorbed on Au/TiO2 frequency of this band is ca. 20 cm−1 higher [36, 37]. The assignment for the 1090 cm−1 band can be ν(Ti−O−C) [4], or ν(C−C) [36, 42]. Aldehyde remaining on the surface is evidenced by the peak at 1723 cm−1 for C− −O vibration and the weak peak at 2716 −1 cm−1 for H−C− −O, while the peak at ca. 2820 cm is superimposed by bands of νs (CH3 ) and νs (CH2 ). Originally, the band at 1723 cm−1 was superimposed by a much stronger band at 1685 cm−1 of the spent sample evacuated at 323 K (not shown). However, the band at 1685 cm−1 disappeared after evacuation at 573 K, while the band at 1723 cm−1 remained intact. The strong and broad bands at 1435 cm−1 (overlapped with the sharp band at 1448 cm−1 for δas (CH3 )/δ(CH2 )) and 1544 cm−1 are assigned to νs (COO) and νas (COO) of surface acetate [43–45]. Additional bands at 1576 cm−1 and 1331 cm−1 are tentatively assigned to surface formate [45, 46]. The surface acetate and formate species are mainly from the oxidation of the strongly adsorbed bidentate propoxy species as proposed in literature [4, 37]. IR adsorption by acetate/formate is less pronounced on the spent catalyst evacuated at 323 K. After heating the sample to 573 K in vacuum, the increase of bands at 1435 cm−1 and 1544 cm−1 corresponded to the decrease of bands in the Ti−O−C/C−C stretching region between 1050 and 1200 cm−1 (the right inset in Figure 5.6A). It is more likely to assign the bands at 1140 and 1090 cm−1 to ν(Ti−O−C) since these two bands turned into O− −C−O vibrations and have the same width and reduced intensity. Probably lattice oxygen took part in the breakage of C−C bonds and consequent oxidation while heating up to 573 K in vacuum. After regeneration, a small amount of carbonate or carboxylates remained on the surface as indicated by strong adsorption at ca. 1540 and 1440 cm−1 (Figure 5.6 spectrum c), which were merely spectators on the surface since the catalyst activity was fully restored after regen- 129 5.3. RESULTS eration. As shown in Figure 5.6, on the spent Au/TiO2 catalyst there was limited adsorption of CO on both gold and Ti4+ similar to what was observed for the spent Au/Ti-SiO2 catalyst. The position of weak CO band located at 2096 cm−1 (shifted from 2108 cm−1 at low coverage) indicates that the accessible gold atoms are typical Au0 , although 2096 cm−1 is at the lower side for CO adsorption on Au0 [24, 25]. On the regenerated sample, CO adsorption on metallic gold is clearly featured by the sharp band at 2107 cm−1 , which was originally at 2123 cm−1 at low CO coverage. The red shift of ν(CO/Au0 ) is very typical for supported gold catalysts due to the fact that the backbonding from gold atoms to 2π* CO orbital is lacking [31–33]. A weak peak located at 2138 cm−1 upon CO adsorption can also be identified and it gradually evolved to lower frequencies at higher CO overages, seemingly saturated as the small band at 2127 cm−1 . The small band at 2127 cm−1 can be assigned to vibration of 13 CO arising from natural abundance, 4+ which interacts with 5-fold coordinated Ti atoms (Ti4+ 5c , or β-Ti ) [23, 29, 47]. This assignment can also be supported by Figure 5.13, which shows CO adsorption on bare TiO2 (P25). However, it is not unambiguous to distinguish the contribution from CO adsorbed on Auδ+ , which also gives a peak located between 2140 and 2125 cm−1 . The −1 −1 band of ν(CO) adsorbed on Ti4+ at 5c located at 2179 cm , which shifted from 2190 cm low CO coverage on the clean and dehydrated Au/TiO2 , while on the spent catalyst this band was at a bit lower frequency (2185 cm−1 ) at low CO coverage. On the regenerated Au/TiO2 the peak at 2214 cm−1 (shifted from 2229 cm−1 ) is assigned to ν(CO) on the 4+ stronger Lewis acid site Ti4+ [23, 29, 47]. 4c , or α-Ti The XPS survey spectrum given in Figure 5.7 also indicates the carbonaceous surface after the reaction. The O 1s, Ti 2p, and Au 4f XPS peaks of the spent catalyst are weaker than their counterparts after regeneration, while the C 1s peak is much stronger. Since the samples were kept in-situ, the surface carbon was most likely not from contamination. Peak fitting of the O 1s XPS spectra shows five components on the catalyst surface, which are located at BE 529.9, 531.2, 532.0, 532.8, and 533.6 eV, respectively. The components at 529.9 and 531.2 eV are assigned to oxygen from Ti−O−Ti and Ti−OH, respectively [48]. Combining the IR data by Figure 5.6 and their intensity change after regeneration, the components of the O 1s line at 532.0, 532.8, and 533.6 eV can be assigned to CO2– 3 or O− −O in ketone/aldehyde, and C−O−Ti species, respectively −C−O in carboxylates, C− [49, 50]. The Au 4f7/2 line is located at BE 83.4 eV for both the spent and regenerated catalyst and the spin doublet separation is 3.6 eV for the Au 4f line. The surface gold is 130 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE O 1s (A) Intensity (a.u.) Ti 2p Regenerated C 1s Spent Au 4f Ti 3s Ti 3p 800 700 600 500 400 300 200 100 0 Binding energy (eV) (B) Au 4f (C) 7/2 Regenerated Spent 92 O 1s 5/2 Intensity (a.u.) Intensity (a.u.) Au 4f Regenerated Spent 88 84 80 Binding energy (eV) 536 532 528 524 Binding energy (eV) Figure 5.7: XP survey spectrum (A), Au 4f photoelectron lines (B), and O 1s photoelectron lines (C) of the spent and regenerated Au/TiO2 catalyst. The spent catalyst was dehydrated in flowing helium at 473 K. Spectra were referenced to the C 1s line at 284.9 eV. simply in its metallic form [15, 51]. The difference in the intensity of the Au 4f lines from the spent and regenerated catalyst is much less profound than the difference observed from CO adsorption by IR. It can be concluded that XPS is not sensitive for detecting the active gold sites in our reaction since this method counts all the gold atoms at the surface rather than those specifically at edges and corners. 5.3.3 CO adsorption on Au/Ti-SiO2 treated by O2 and H2 The effect of O2 /H2 pretreatment on the gold oxidation state is shown in Figure 5.8. On the regenerated Au/Ti-SiO2 sample (Figure 5.8A), two bands can be distinguished: the sharp one at 2108 cm−1 and a shoulder at 2128 cm−1 . They are assigned to CO adsorbed 131 5.3. RESULTS 0.025 δ+ 1 8 CO Au CO Au0 1 1 c1 0 2150 2100 Wavenumber (cm 1 111 c10 185 0.005 0 2200 δ+ 1 8 CO Au 158 0.01 a10 a1 1 1 0.005 Abs. 0.01 (B) 0.02 0.015 158 185 CO Ti4+ Abs. 0.015 CO OH 0.02 108 (A) CO Au0 0.025 2200 2050 2150 2100 Wavenumber (cm 1 2050 0.025 0.015 a2-a1 2128 0.02 Abs. 0.002 0.01 b1 118 185 0.005 b2-b1 b10 158 Abs. (D) CO Au0 10 (C) c2-c1 0 2200 2150 2100 Wavenumber (cm 1 2050 2180 2160 2140 2120 2100 -1 Wavenumber (cm 2080 2060 ) Figure 5.8: Progressive CO adsorption on the Au/Ti-SiO2 sample after different pretreatment: (A) after regeneration at 573 K followed by 0.5 hour in vacuum at 573 K and quenching to 90 K; (B) after regeneration at 573 K, 0.5 hour in vacuum at 573 K, 0.5 hour in 5 vol% O2 at 423 K, 0.5 hour in vacuum at 423 K and quenching to 90 K; (C) after regeneration at 573 K, 0.5 hour in vacuum at 573 K, 0.5 hour in 5 vol% O2 at 423 K, 0.5 hour in vacuum at 423 K, 0.5 hour in 5 vol% H2 at 423 K, 0.5 hour in vacuum at 423 K and quenching to 90 K. Section (D) gives incremental adsorption between the second and first dose. The treated sample in vacuum at 90 K was used as background. CO pressure increased from 0.01 mbar to 0.20 mbar after 10 doses with increment of ca. 0.02 mbar per dose. on Au0 and Auδ+ , respectively. After pre-oxidation at 423 K (Figure 5.8B), the ν(CO/Au0 ) band became much weaker, while the ν(CO/Auδ+ ) almost remained the same. By further treating the pre-oxidized sample in H2 (Figure 5.8C), the ν(CO/Auδ+ ) band could not be clearly observed. The incremental adsorption of CO between the second and first dose (Figure 5.8D) indicates that on the reduced Au/Ti-SiO2 there were mainly Au0 sites and that the remaining small amount of Auδ+ may come from the interfacial Au−O−Ti. 132 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE As for the regenerated sample, there may be more Au−O−O−Ti sites at the perimeter as proposed in literature [35]. And on the pre-oxidized sample, it seems that Au0 atoms were covered by molecular oxygen even after evacuation at 423 K, which is in accordance with the high desorption temperature of O2 on small gold nanoparticles [52]. However, when Au/TiO2 was used, the disturbance of ν( 13CO) on Ti4+ to ν(CO) on Auδ+ obscured the band assignment and thus not discussed. On the other hand, treating Au/TiO2 in the same way as Figure 5.8 but at 323 K (the typical epoxidation temperature for Au/TiO2 ) showed no significant change in the adsorption intensity of CO on Auδ+ and Au0 when comparing to literature [23, 35], probably due to the low pre-oxidation temperature and low gold loading. 5.3.4 CO adsorption on Au/TiO2 in the presence of H2 Figure 5.9 shows the progressive CO adsorption on Au/TiO2 in the presence of H2 at 90 K. In the OH stretching region before CO dosing, three sharp bands at 3734, 3675, and 3420 cm−1 together with several weak shoulders at 3715, 3690, 3658, and 3648 cm−1 can be observed. During the H2 treatment at 323 K, trace amount of H2 O was formed as evidenced by the weak adsorption of δ(H2 O) at 1617 cm−1 . Upon CO dosing, a sharp band in the OH stretching region at 3648 cm−1 appeared together with a band for CO2 at ca. 2350 cm−1 . Surprisingly, the pressure in the IR cell continued to decrease till the 3rd or 4th dose. CO adsorbed on gold from the first 3 doses (via the 5 µL loop) was almost fully and instantly converted to CO2 . The CO2 vibration band reached it peak intensity after the 3rd dose and gradually decreased after subsequent CO dosage very likely due to the replacement of CO on the surface. The change in the OH stretching region after CO dosing was more difficult to explain. The band at 3734 cm−1 remained at its original position, which is generally assigned to impurities such as Si−OH [36, 47]. The weak shoulder at 3715 cm−1 shifted to 3706 cm−1 upon CO dosing and can be assigned to isolated Ti−OH of the anatase phase [53, 54]. The major band centered at 3679 cm−1 decreased its intensity as the CO coverage increased and shifted to the broad band at 3547 cm−1 due to the interaction with CO. This major band can be assigned to Ti−OH of anatase [47, 55]. The assignment of the 3425 cm−1 band was proposed to be from water molecules or hydroxy groups on TiO2 (rutile) [36, 55, 56]. The sharp band at 3648 cm−1 developed into the band at ca. 3500 cm−1 with a shift of 150 cm−1 as the CO coverage increased. A clear assignment of the 3648 cm−1 band is difficult in our case. This band 133 5.3. RESULTS (mbar) 0.04 (A) Pressure in cell 4.80 4.75 0 2 4 6 8 10 Abs. dose of CO 11 2 1 dose 0 90 K in 4.82 mbar H 3800 3600 3400 3200 2250 2000 1750 1500 2 1250 1000 -1 (C) 2103 2179 (B) 501 547 648 0.1 0.08 se 11 0.06 2127 0.04 0.02 se 1 2213 3492 Abs. Abs. 425 7 4 0.01 679 se 0 2165 2157 Wavenumber (cm ) dose 11 dose 1 0 3800 3700 3600 3500 3400 Wavenumber (cm 1 3300 3200 2200 2150 2100 Wavenumber (cm 1 2050 Figure 5.9: Progressive CO adsorption on Au/TiO2 in the presence of H2 : (A) wide frequency range, the inset shows the pressures in the IR cell after each dose of CO; (B) spectra in the ν(OH) region; (C) difference spectra in the ν(CO) region using the spectrum of dose 0 as the background. The regenerated Au/TiO2 sample was kept in 50 mbar H2 (pre-dried with liquid nitrogen trap) at 323 K for 10 min followed by quenching down to 90 K and adjusting the H2 pressure by evacuation to 4.82 mbar before CO dosing. All spectra were taken at 90 K. was very weak before CO dosing. Upon CO dosing in the presence of H2 , this band became very strong and sharp. Considering the decrease of cell pressure after CO dosing and the CO2 formation, it is clear that CO reacted with lattice oxygen in the presence of H2 forming CO2 and hydroxy groups and/or water. Water formation can be evidenced by the weak band at 3492 cm−1 upon CO dosing (red spectrum, dose 1 in Figure 5.9B), which can be assigned by ν(OH) bonded to water. On the other hand, from analysis of the difference spectra in the OH stretching region(not shown), the perturbation of a component at 3658 cm−1 upon CO dosing also contributed to the band increase at 3648 134 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE cm−1 but to a much lesser extent. The contribution of pre-adsorbed water to the bands at 3648 cm−1 , either by the reaction with CO or by the dissociation of water on coordination unsaturated Ti4+ sites, was excluded by the experiment shown in Figure 5.10. 3435 f 3500 a 2107 Abs. f 1616 a 0.02 -1 Wavenumber (cm ) Figure 5.10: Progressive CO adsorption on Au/TiO2 pre-covered by trace amount of H2 O: (a) in vacuum; (b–e) dose 1 to 4, in 0.01, 0.03, 0.05, 0.07 mbar CO respectively; (f) dose 10, in 0.21 mbar CO. The sample was pre-treated by partially dried H2 at 323 K and evacuated at 323 K for 1 hour. All spectra were taken at 90 K. In Figure 5.10, the coverage of water before CO adsorption was evidenced by the relatively stronger band of δ(H2 O) at 1616 cm−1 and a band of ν(OH) perturbed by H2 O at 3500 cm−1 . After CO adsorption, no significant change occurred to the 3648 cm−1 band. With the increased coverage of CO, the band for δ(H2 O) gradually shifted to 1645 cm−1 . The band at 3500 cm−1 gradually shifted to 3435 cm−1 . The band for ν(CO) on Au0 shifted from 2119 cm−1 at the low CO coverage and saturated at 2107 cm−1 . It seems that water is not necessary for the band at 3648 cm−1 although its decrease in intensity was observed concomitantly with outgassing of water molecules on TiO2 surface in other studies [56]. Figure 5.9 shows that the bands of ν(CO) at 2213(α-Ti4+ ), 2179(β-Ti4+ ), 2165(γTi4+ ), and 2103(Au0 ) cm−1 did not shift with the CO coverage in the presence of H2 . The band at 2157 cm−1 is assigned to ν(CO) interacting with OH groups. The small band at 2127 cm−1 is assigned to vibration of 4+ with β-Ti 13 CO (arising from natural abundance) interacting [23]. The bands at 2165 and 2157 cm−1 would only appear at very high CO pressures if there were no H2 in presence. After the experiment performed in Figure 5.9, the IR cell was immediately evacuated at 90 K and the progressive CO adsorption on this 135 5.3. RESULTS sample was repeated but without H2 in presence. The results are given in Figure 5.11. The pressure in the IR cell increased monotonously by each dose of CO with an average increment of ca. 0.02 mbar, which is significantly different from what was observed in Figure 5.9A, where CO and H2 reacted leading to a pressure decrease. The band at 3648 cm−1 remained. The bands at 2165 and 2157 cm−1 were not observed. It can be concluded that the presence of H2 significantly enhanced the adsorption of CO on TiO2 at low temperature. 3680 3648 2106 0.06 0 3700 3600 3500 3400 Wavenumber (cm−1) 3300 3200 2127 dose 11 2117 2213 2188 Abs. 0.04 0.02 3800 (B) 0.08 3425 Abs. 3735 0.01 2179 0.1 (A) 2200 2150 2100 Wavenumber (cm 1 dose 1 2050 Figure 5.11: CO adsorption on Au/TiO2 after removing hydrogen (to be compared with Figure 5.9, see text). The following CO pressures after each dose were used (in mbar): 0.00, 0.01, 0.03, 0.05, 0.07, 0.09, 0.11, 0.13, 0.16, 0.18, 0.20 for dose 1 to dose 11 (red to blue). All spectra were taken at 90 K. The same experiments were performed on the bare TiO2 (P25) sample for reference. The results are shown in Figures 5.12 and 5.13. The pressure in the IR cell increased monotonously after each CO dose when H2 was present as shown in Figure 5.12, which in turn confirmed that CO and H2 reacted on Au/TiO2 with lattice oxygen as indicated in Figure 5.9. The results shown in this study are consistent with what Widmann and Behm [57] revealed. They showed that only for Au on TiO2 can TiO2 be reduced by CO. The presence of H2 simply improved the CO coverage at low CO pressures featured by the stronger interaction of CO with hydroxy groups (3400 – 3600 cm−1 , 2159 cm−1 ) and the weak acidic γ-Ti4+ (2165 cm−1 ) as shown in Figure 5.12B/C. After the experiment in Figure 5.12 was done, the cell was fully evacuated and CO adsorption on TiO2 without H2 present was performed at the same temperature (Figure 5.13). The only and obvious difference in ν(CO) region as compared to Figure 5.11 is that no CO on gold was observed. The remaining peaks were exactly the same, which also confirms the assignment 136 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE (A) Pressure in cell (mbar) 0.04 5.30 5.20 5.10 0 2 4 6 8 10 Abs. dose of CO 9 2 1 dose 0 90 K in 5.12 mbar H 3800 3600 2 3400 3200 2250 2000 1750 1500 1250 -1 0.08 3432 (C) 0.06 Abs. Abs. 2165 2159 2179 (B) 3637 0.01 3729 0.1 3547 3520 3681 Wavenumber (cm ) dose 1 to 9 0.04 2213 0.02 2127 dose 0 dose 9 dose 1 0 3800 3700 3600 3500 3400 Wavenumber (cm−1) 3300 3200 2200 2150 2100 Wavenumber (cm 1 2050 Figure 5.12: Progressive CO adsorption on TiO2 (P25) in the presence of H2 : (A) wide frequency range, the inset shows the pressures in the IR cell after each dose of CO; (B) spectra in the ν(OH) region; (C) difference spectra in the ν(CO) region using the spectrum of dose 0 as the background. The TiO2 sample was calcined in O2 at 573 K and treated in 50 mbar H2 (pre-dried with liquid nitrogen trap) at 323 K for 20 min followed by quenching down to 90 K and adjusting the H2 pressure by evacuation to 5.12 mbar before CO dosing. All spectra were taken at 90 K. of the 2127 cm−1 band to 13 CO. In the OH stretching region showed no obvious sharp bands at 3648 and ca. 3420 cm−1 . 5.3.5 C3 H6 adsorption on Au/TiO2 in the presence of CO Figure 5.14 shows the IR spectra of C3 H6 adsorption on Au/TiO2 in the presence of CO at 230 K (above the boiling point of C3 H6 226 K at atmospheric pressure). The main adsorption band of ν(CO) on gold in 10 mbar CO located at 2108 cm−1 . Two weak bands 137 5.3. RESULTS 2179 3675 0.1 0.01 3644 2188 0.04 0.02 2213(2229) Abs. 3433 Abs. 0.06 2127(2139) 3727 0.08 0 3800 3700 3600 3500 3400 Wavenumber (cm 1 3300 3200 2200 2150 2100 Wavenumber (cm−1) 2050 Figure 5.13: CO adsorption on TiO2 after removing hydrogen. The following CO pressures after each dose were used (in mbar): 0.02, 0.04, 0.07, 0.11, 0.14, 0.18, 0.22, 0.26, 0.30 for dose 1 to dose 9 (red to blue). All spectra were taken at 90 K. at 2132 and 2069 cm−1 were also observed. The band at 2132 cm−1 can be assigned to CO on Auδ+ . The broad band at 2060 – 2070 cm−1 was also observed in several other studies when at relatively high CO pressures (e.g. 20 mbar) and temperatures (e.g., at room temperature) [58, 59]. In our experiment, this band was observed after a few minutes after the sample contacted with CO. It stabilized without further evolution. As proposed in literature, this weak band is more likely due to morphology change of gold nanoparticles (lowered coordination number of gold atoms) in CO but to a small extent in our case. The progressively dosed C3 H6 gradually replaced CO on the surface interacting with both Ti4+ and Au as indicated by the weakened CO bands in Figure 5.14A. It is known that C3 H6 forms a stronger π-complex with Ti4+ than CO [60], thus all CO adsorbed on Ti4+ was almost removed. It seems that C3 H6 has also a stronger interaction with gold than CO does. When there was 2.5 mbar C3 H6 present, the band of CO on gold red-shifted by 30 cm−1 to 2079 cm−1 . Interaction with C3 H6 does not lead to reconstruction of gold nanoparticles. This was confirmed by low temperature CO adsorption on the Au/TiO2 sample which had contacted with 50 mbar C3 H6 at 323 K for 30 min followed by evacuation. Thus, the red shift of ν(CO/Au) was more likely an effect of change in the electron density of gold atoms. The band at 2079 cm−1 may be assigned to ν(CO/Auδ− ) [61]. The presence of C3 H6 on the catalyst surface was confirmed by the IR spectra in the CH stretching and bending regions as given in Figure 5.14B. The details and assignment 138 2183 2108 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE 0.01 4+ CO-Ti 0.05 2132 2178 g 2069 a b 2200 2150 2100 2050 c d e f g 2079 Absorbance (a.u.) b CO-Au (A) 2150 2100 2050 2000 1950 -1 a - g 0.02 b - g 2894 2923 2856 0.01 1300 2950 1400 2977 a - g 1500 3009 1600 3081 1700 3059 3736 Absorbance (a.u.) 1373 1453 1632 3421 3648 3676 Wavenumber (cm ) (B) 3800 3600 3400 3200 3000 2800 -1 Wavenumber (cm ) Figure 5.14: Progressive C3 H6 adsorption on Au/TiO2 in the presence of CO at 230 K: (A) the CO stretching region; the spectra were shifted for clarity; spectra in the inset were not shifted; (B) the OH and CH stretching region; the inset shows C− −C stretching and CH bending region. Spectra (gas phase compensated): (a) in vacuum before CO and C3 H6 adsorption; (b) in 10.3 mbar CO after 20 min; (c) in 10.3 mbar CO + 0.5 mbar C3 H6 ; (d) in 10.3 mbar CO + 1.0 mbar C3 H6 ; (e) in 10.3 mbar CO + 1.5 mbar C3 H6 ; (f) in 10.3 mbar CO + 2.0 mbar C3 H6 ; (g) in 10.3 mbar CO + 2.5 mbar C3 H6 . 139 5.3. RESULTS g .005 Abs. Abs. .005 f e g c d c 3600 3400 3200 3000 Wavenumber (cm−1) 2800 2600 1800 1600 1400 1200 Wavenumber (cm−1) 1000 Figure 5.15: Difference spectra of progressive C3 H6 adsorption on Au/TiO2 in the presence of CO at 230 K (see Figure 5.14, spectrum b as the background). of bands from C3 H6 adsorbed on the Au/TiO2 sample are given in Figure 5.15 and Table 5.3. The decrease of the main bands at 3676 and 3648 cm−1 and increase of the broad band centred at 3410 cm−1 is attributed to the formation of hydrogen bond between the allylic hydrogen of C3 H6 and oxygen in Ti−OH [62]. The sharp band of ν(C− −C)at 1632 cm−1 is slightly lower than the gas-phase C3 H6 , indicating a weak π-bonding to Ti4+ and Au. Three bands above 3000 cm−1 were observed. The bands at 3081 and 3009 cm−1 can be assigned to νas (CH2 ) and ν(CH), respectively. The origin of the 3059 cm−1 is unclear. This band was not observed for C3 H6 adsorbed on TiO2 (anatase) [62]. We tentatively assign this 3059 cm−1 band as νas (CH2 ) of C3 H6 adsorbed at the Au–Ti interface or on gold. However, further investigation may be needed. After evacuation, no bands in the CH stretching region and for C− −C could be observed. The weak bands at 1576 and 1252 cm−1 (Figure 5.15) can be assigned to carbonates due to the presence of CO [41]. Since the co-adsorption of CO and C3 H6 was performed at a low temperature of 230 K, it is interesting to know how C3 H6 would compete with CO at temperatures where the epoxidation occurs. In Chapter 4, we demonstrated that CO can switch off propene hydrogenation over the gold–titania catalysts. However, it is not clear if the role of CO is on gold or the support or both. If the effect of CO is on gold, propene should not have much effect on CO oxidation if we assume all the chemistry happens on the low coordinated gold atoms close to the Au–Ti interface. 140 Table 5.2: Summary of frequency observed (cm−1 ) for CO adsorption for different catalysts in this study a Au/SiO2 Au/Ti-SiO2 S R 2181 2185 S Au/TiO2 R 2158 2136 2158 2136 2158 2136 2138 2127 2105 2099 2095 2125, 2128 2118, 2121 2107 a: S, spent; R, regenerated 2108 2096 2123 2107 2079 Ref. on Ti4+ of Ti-SiO2 on α-Ti4+ of TiO2 , at low CO coverage on α-Ti4+ of TiO2 , at high CO coverage on β-Ti4+ of TiO2 , at low CO coverage on β-Ti4+ of TiO2 , at high CO coverage on γ-Ti4+ of TiO2 with isolated OH groups on TiO2 with OH groups on SiO2 and/or Ti-SiO2 physisorbed CO not fully resolved, in Figure 5.6C, tentatively 13 CO on β-Ti4+ at low CO coverage 13 CO on β-Ti4+ on Auδ+ on Au0 , low CO coverage on Au0 , high CO coverage on Auδ− , C3 H6 present [29] [23, 29, 47] [23, 29, 47] [23] [23] [29, 30] [23, 29, 47] [23, 34, 35] [24, 25] [61] CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE 2185 2178 2229 2214 2190 2179 2165 2157 Assignment 141 5.3. RESULTS The effect of C3 H6 on CO oxidation over Au/TiO2 is shown in Figure 5.16. It can be seen that CO oxidation was suppressed by C3 H6 at either 303 K or 373 K. No significant deactivation was observed in the first 8 hours. However, after flushing in C3 H6 (period III) and further flushing in helium till C3 H6 concentration lower than 10 ppm, the activity in CO oxidation was severely reduced. Co-feeding of C3 H6 completely suppressed CO oxidation. At 373 K after C3 H6 flushing (11.5 – 13.5 h), the activity in CO oxidation gradually restored very likely due to desorption of adsorbed propene. However, after period IV when C3 H6 was removed from the CO/O2 stream (the last 1 hour), the activity could not be restored. It seems that propene formed a strongly adsorbed species at the Au–Ti interface when CO oxidation was proceeding (O adatoms available) and thus deactivated the catalytic site. The Au/TiO2 catalyst, after being tested at 303 K as shown in Figure 5.14a, was then heated to 333 K and tested for the direct propene epoxidation. The epoxidation activity at 333 K is given in Figure 5.17. There was only ca. 30% loss in the maximum PO formation rate. The trajectory of rH2 O against rPO was still on par with the original performance of the catalyst as indicated by Figure 5.17c. It can be concluded that the strongly adsorbed species formed from C3 H6 during CO oxidation can be removed by hydro-epoxidation. Table 5.3: IR frequencies (in cm−1 ) of C3 H6 and their assignment vibrations gas phase[63] in polyethylene[64] Ir4 /Al2 O3 [65] on TiO2 [62] this work νas (CH2 ) 3090 3075 3077 3090 3081, 3059 ν(CH) 3013 3008 3014 3010 3009 νs (CH2 ) 2992 2977 2979 2990 2977 νas (CH3 ) 2954, 2933 2958, 2933 2968 2960 2950, 2923 νs (CH3 ) ν(C−C) 2870 2912 2859 2930, 2870 2894, 2856 1652 1645 1640 1635 1632 δas (CH3 ) 1474, 1443 1449, 1443 1458, 1442 1460, 1440 1453, 1435 δ(CH) 1419, 990 1411, 988 1296 δs (CH3 ) 1378 1370 1376 δ(CH2 ) 1298 1293 1415 ρ(CH2 ) 1229 1169 1170 1171 ρ(CH3 ) 1171, 1045 1040 1049, 1007 1047 1415, 1010 1380 1373 1295 142 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE as c ncen ra n ( m) 5000 II I I II 4000 II III I IV I 3000 ×0.5 6 2000 1000 (a) 0 0 2 4 6 8 T me (h) 10 12 14 16 as c ncen ra n ( m) 5000 4000 3000 ×0.5 6 2000 1000 (b) 2 0 0 2 4 6 8 T me (h) 10 12 14 16 Figure 5.16: Effect of C3 H6 on CO oxidation over Au/TiO2 : (a) at 303 K; (b) at 373 K . I, 1500/4750 ppm of CO/O2 , helium balance; II, helium flush; III, 10000 ppm C3 H6 flush; IV, 1500/4750 ppm of CO/O2 , co-feeding 10000 ppm C3 H6 . 150 mg catalyst, GHSV −1 20000 mL·g−1 cat h . 5.4 Discussion The oxidation state of gold in the direct propene epoxidation can be summarized in the scheme given in Figure 5.18. Our data of propene adsorption on the Au/TiO2 catalyst suggests that the low-coordinated gold atoms become negatively charged in the presence of propene, which leads to a broad carbonyl band at a low frequency of 2079 cm−1 . The negatively charged gold has been also recently well described by Chakarova et al. on Au/SiO2 system by CO reduction [61]. Combining our previous work of propene adsorption on Au/SiO2 [5], we would not expect much difference for Au/Ti-SiO2 in terms of charge transfer over gold in different atmospheres. In our case for Au/TiO2 , contributions from propene adsorbed on gold and the support near the interface may both count 143 5.4. DISCUSSION 2 x 10 −7 −6 2 (a) x 10 (b) Au/TiO , PO Au/TiO , H O 2 2 (mol⋅g−1 ⋅s−1) cat rPO (mol⋅g−1 ⋅s−1) cat 1.5 1.5 1 rH 2 O 1 2 0.5 0 0.5 0 30 60 Time (min) 90 120 0 0 30 60 Time (min) 90 120 −6 1 x 10 (c) 0.6 0.4 rH 2 O (mol⋅g−1 ⋅s−1) cat 0.8 0.2 0 0 0.5 1 1.5 rPO (mol⋅g−1 ⋅s−1) cat 2 −7 x 10 Figure 5.17: The activity of Au/TiO2 in the direct propene epoxidation at 333 K after testing the effect of C3 H6 on CO oxidation as shown in Figure 5.16a. GHSV 10000 −1 mL·g−1 cat h , 10/10/10/70 H2 /O2 /C3 H6 /He mixture. (a) PO formation rate; (b) water formation rate; (c)relation between the PO and water formation rates. Grey dots are data of three reaction cycles from Figure 5.1 for reference. to increase the electron density on gold. But for Au/Ti-SiO2 , the contribution of charge transfer from the support and the role of the support oxygen may be less pronounced [66]. Propene adsorption on gold in this study was performed at a relatively low temperature (230 K) and in an atmosphere where oxygen was absent. However, under the real reaction conditions, the species adsorbed on gold are much more complicated. Oxygen molecules are activated on gold by charge transfer from reduced gold forming gold– oxygen complex [67, 68]. When hydrogen is present, OOH is formed [1]. The study by Bravo-Suárez et al. on propane oxidation in hydrogen and oxygen over gold catalysts suggests that even propane can be oxidized by Ti–OOH forming isopropoxy species 144 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE [69]. The real role of propene modifying the reactivity of gold may be complicated. On one hand, propene can pi-bonding to gold and reduce the water formation as shown by Nijhuis et al. in their XANES study [5]. The spectra shown in Figure 5.14 provide another evidence of the electron-donating ability from propene to gold, which seems to be even stronger than CO. On the other hand, oxidants adsorbed on gold may also react with propene forming strongly adsorbed species as proposed by Nijhuis et al.[5] and as suggested by the CO oxidation experiment cofed with propene in this study. Although propene adsorption on either gold single crystal surfaces or a model Au/TiO2 catalyst is found to be weak [70, 71], it may be not the case for a real catalyst. On a model Au/TiO2 (110) catalyst in the work by Ajo et al. [71], the most stable sites for propene adsorption were those at the edge of gold islands with a TPD peak at ca. 240 K, while for 2D Au islands the temperature was higher till 310 K. For a real Au/TiO2 catalyst active for the direct propene epoxidation, the TPD peak of propene desorption on gold and/or at the perimeter was found to be at 323 K [14]. TPD data over Au/Ti-SiO2 by BravoSuárez et al. suggests that the presence of gold and titania can significantly enhance the adsorption of C3 H6 and a desorption temperature higher than 450 K was reported by them [72]. However, it would be more interesting to know if propene and oxygen can cooperatively adsorb on gold just as CO and oxygen on gold clusters [73, 74]. In the coadsorption of CO and oxygen on gold clusters, it is proposed that CO acts as the electron donor and oxygen as the acceptor, both not competing for adsorption sites but cooperating [73, 74]. Our finding on the stronger electron-donating ability of propene than CO on gold might provide a starting point to this question. The desorption temperature for O2 on gold is generally high as reported in literature. In the early work by Hayashi et al. [14], the TPD curve of Au/TiO2 which had been treated at 573 K in O2 showed that O2 desorbed at above ca. 500 K. The desorption of recombined O adatoms was reported to be 470 K on Au(100) and above 500 K on Au(111) [70], and even higher for thin gold nanoparticles supported on TiO2 [52]. The results in Figure 5.8 of Au/Ti-SiO2 treated by O2 and H2 also indicate that the removal of O2 needs a relatively high temperature. Evacuation at 573 K seems not enough for us to fully remove oxygen on gold for Au/Ti-SiO2 . The presence of O2 on gold, or preferably at the interface [75], makes the gold atoms electron deficient. The Auδ+ sites can be reduced by hydrogen to Au0 as shown in Figure 5.8. The CO adsorption on spent catalysts, as shown in Figures 5.4 and 5.6, indicates that the gold sites during reaction are fully covered by deactivating species and reac- 5.5. CONCLUSIONS 145 Figure 5.18: Schematic of gold oxidation state under conditions for propene epoxidation tion intermediates. Very limited gold sites were accessible for the probe molecule CO, even if the Au/Ti-SiO2 catalyst still had half its activity in PO formation after deactivation. CO adsorbed on these remaining gold sites shows the typical band location for Au0 . However, these CO vibration bands show a long tailing shape towards lower wavenumbers indicating certain degree of reduction of the supports due to the adsorbed organic species and thus less electron transfer from gold to the support. IR using CO as the probe molecule may be not the preferred method to detect gold oxidation state under the real reaction conditions due to the fully covered surface. When using other bulk methods like XANES or XAFS one may also need to pay attention to the small contribution from the low-coordinated gold atoms if the gold particle size is not small enough. A very interesting phenomenon observed in this study is that the presence of H2 greatly increased the amount of CO adsorbed on the TiO2 support at low temperature, as shown in Figures 5.9 and 5.12. This may provide one explanation for the preferential oxidation (PROX) of CO over Au/TiO2 with aid of H2 . On Au/TiO2 , the metal–support interface is considered to be the active site for O2 dissociation [75, 76]. It is also found that the diffusion of CO adsorbed on TiO2 to the Au–Ti interface contributes more to CO oxidation over Au/TiO2 at low temperature [76]. Thus, besides a facile dissociation of O2 in the presence of H2 , H2 may also promote the adsorption of CO on the support and consequently increase the reaction rate. 5.5 Conclusions On both Au/TiO2 and Au/Ti-SiO 2 catalysts investigated after use in the direct propene epoxidation, the active gold sites can hardly be detected by CO molecules. The low- 146 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE coordinated gold atoms are most likely occupied by propoxy species and carboxylates, which are most abundant on the catalyst surface after the reaction. Those gold atoms which can still be detected by CO show a very weak band at 2095 cm−1 indicating their metallic state. XPS spectra show that the oxidation state of surface gold atoms is metallic before and after reaction. However, XPS is not a sensitive tool to pinpoint those lowcoordinated atoms at the perimeter of gold nanoparticles. Calcination in oxygen at 573 K restores the catalyst activity by removing the deactivating species and carbonaceous spectators. Oxygen can strongly adsorb on gold or at the perimeter of gold nanoparticles over the Au/Ti-SiO 2 catalyst at the reaction temperature for propene epoxidation (e.g., 423 K), leading to positively charged gold atoms as evidenced by the carbonyl band on gold at ca. 2125 cm−1 . Reduction in hydrogen at the reaction temperature removes adsorbed oxygen on the catalyst and restores the low-coordinated gold atoms to electron-neutral state, which give carbonyl bands at 2100 - 2110 cm−1 . Over the Au/TiO2 catalyst, it has been observed that the presence of hydrogen at low temperatures can significantly enhance CO adsorption on the support, which may help explain the role of hydrogen in the preferential oxidation of CO. The low-coordinated gold atoms become negatively charged in the presence of propene. Propene shows a stronger chemisorption than CO on both gold and Ti4+ . Progressive propene adsorption leads to the weakening and broadening of CO band on gold with a red shift of ca. 30 cm−1 towards 2079 cm−1 . This implies a strong interaction between propene and gold through electron donation, which may explain the suppressed water formation over the gold catalysts when propene is present. References [1] C. Sivadinarayana, T. V. Choudhary, L. L. Daemen, J. Eckert and D. W. Goodman, JACS, 2004, 126, 38–39. [2] J. J. Bravo-Suárez, K. K. Bando, J. Lu, M. Haruta, T. Fujitani and S. T. Oyama, J. Phys. Chem. C, 2008, 112, 1115–1123. [3] M. G. Clerici and P. Ingallina, J. Catal., 1993, 140, 71–83. [4] T. A. Nijhuis, T. Visser and B. M. Weckhuysen, J. Phys. Chem. B, 2005, 109, 19309–19319. [5] T. A. Nijhuis, E. Sacaliuc, A. M. Beale, A. M. J. van der Eerden, J. C. Schouten and B. M. Weckhuysen, J. Catal., 2008, 258, 256–264. [6] B. Taylor, J. Lauterbach, G. E. Blau and W. N. Delgass, J. Catal., 2006, 242, 142–152. REFERENCES 147 [7] J. Lu, X. Zhang, J. J. Bravo-Suárez, S. Tsubota, J. Gaudet and S. T. Oyama, Catal. Today, 2007, 123, 189–197. [8] J. Chen, S. J. A. Halin, J. C. Schouten and T. A. Nijhuis, Faraday Discuss., 2011, 152, 321–336. [9] M. Boronat, F. Illas and A. Corma, J. Phys. Chem. A, 2009, 113, 3750–3757. [10] T. Fujitani, I. Nakamura, T. Akita, M. Okumura and M. Haruta, Angew. Chem. Int. Ed., 2009, 48, 9515–9518. [11] I. X. Green, W. Tang, M. Neurock and J. T. Yates Jr., Angew. Chem. Int. Ed., 2011, 50, 10186– 10189. [12] B. Yang, X.-M. Cao, X.-Q. Gong and P. Hu, Phys. Chem. Chem. Phys., 2012, 14, 3741–3745. [13] D. A. Panayotov, S. P. Burrows, J. T. Yates and J. R. Morris, J. Phys. Chem. C, 2011, 115, 22400–22408. [14] T. Hayashi, K. Tanaka and M. Haruta, J. Catal., 1998, 178, 566–575. [15] A. Zwijnenburg, A. Goossens, W. G. Sloof, M. W. J. Crajé, A. M. Van der Kraan, L. J. De Jongh, M. Makkee and J. A. Moulijn, J. Phys. Chem. B, 2002, 106, 9853–9862. [16] C. Qi, J. Huang, S. Bao, H. Su, T. Akita and M. Haruta, J. Catal., 2011, 281, 12–20. [17] J. Gaudet, K. K. Bando, Z. Song, T. Fujitani, W. Zhang, D. S. Su and S. T. Oyama, J. Catal., 2011, 280, 40–49. [18] J. Chen, S. J. A. Halin, D. M. Perez Ferrandez, J. C. Schouten and T. A. Nijhuis, J. Catal., 2012, 285, 324–327. [19] A. S. K. Hashmi and G. J. Hutchings, Angew. Chem. Int. Ed., 2006, 45, 7896–7936. [20] J. Chou, N. R. Franklin, S.-H. Baeck, T. F. Jaramillo and E. W. McFarland, Catal. Lett., 2004, 95, 107–111. [21] G. C. Bond and D. T. Thompson, Gold Bull., 2000, 33, 41–50. [22] M. A. P. Dekkers, M. J. Lippits and B. E. Nieuwenhuys, Catal. Lett., 1998, 56, 195–197. [23] Tz. Venkov, K. Fajerwerg, L. Delannoy, Hr. Klimev, K. Hadjiivanov and C. Louis, Appl. Catal., A, 2006, 301, 106–114. [24] R. Meyer, C. Lemire, Sh. K. Shaikhutdinov and H.-J. Freund, Gold Bull., 2004, 37, 72–124. [25] M. Mihaylov, H. Knözinger, K. Hadjiivanov and B. C. Gates, Chem. Ing. Tech., 2007, 79, 795– 806. [26] E. J. M. Hensen, D. G. Poduval, D. A. J. M. Ligthart, J. A. R. van Veen and M. S. Rigutto, J. Phys. Chem. C, 2010, 114, 8363–8374. [27] Characterization and Chemical Modification of the Silica Surface, ed. E. F. Vansant, P. Van Der Voort and K. C. Vrancken, Elsevier, 1995, pp. 59–77. [28] A. Burneau and J. P. Gallas, in The Surface Properties of Silicas, ed. A. P. Legrand, John Wiley and Sons, 1998, ch. 3. Hydroxyl Groups on Silica Surfaces, pp. 147–234. [29] K. Hadjiivanov, B. M. Reddy and H. Knözinger, Appl. Catal., A, 1999, 188, 355–360. [30] D. A. Boyd, F. M. Hess and G. B. Hess, Surf. Sci., 2002, 519, 125–138. 148 CHAPTER 5. HOW METALLIC IS GOLD IN THE DIRECT EPOXIDATION OF PROPENE [31] J. France and P. Hollins, J. Electron. Spectrosc. Relat. Phenom., 1993, 64–65, 251–258. [32] F. Boccuzzi, S. Tsubota and M. Haruta, J. Electron. Spectrosc. Relat. Phenom., 1993, 64–65, 241–250. [33] C. Ruggiero and P. Hollins, J. Chem. Soc., Faraday Trans., 1996, 92, 4829–4834. [34] F. Boccuzzi, A. Chiorino and M. Manzoli, Mater. Sci. Eng., C, 2001, 15, 215–217. [35] I. X. Green, W. Tang, M. McEntee, M. Neurock and J. T. Yates, JACS, 2012, 134, 12717–12723. [36] G. Mul, A. Zwijnenburg, B. Van der Linden, M. Makkee and J. A. Moulijn, J. Catal., 2001, 201, 128–137. [37] A. Ruiz, B. van der Linden, M. Makkee and G. Mul, J. Catal., 2009, 266, 286–290. [38] F. Meunier, D. Reid, A. Goguet, S. Shekhtman, C. Hardacre, R. Burch, W. Deng and M. Flytzani-Stephanopoulos, J. Catal., 2007, 247, 277–287. [39] J. Gong, D. W. Flaherty, T. Yan and C. B. Mullins, ChemPhysChem, 2008, 9, 2461–2466. [40] S. Y. Liu and S. M. Yang, Applied Catalysis A: General, 2008, 334, 92–99. [41] M. A. Bollinger and M. A. Vannice, Appl. Catal., B, 1996, 8, 417–443. [42] W.-C. Wu, C.-C. Chuang and J.-L. Lin, J. Phys. Chem. B, 2000, 104, 8719–8724. [43] G. A. M. Hussein, N. Sheppard, M. I. Zaki and R. B. Fahim, J. Chem. Soc., Faraday Trans. 1, 1989, 85, 1723–1741. [44] V. Sanchez Escribano, G. Busca and V. Lorenzelli, J. Phys. Chem., 1990, 94, 8939–8945. [45] F. P. Rotzinger, J. M. Kesselman-Truttmann, S. J. Hug, V. Shklover and M. Grätzel, J. Phys. Chem. B, 2004, 108, 5004–5017. [46] M. Haruta, S. Tsubota, T. Kobayashi, H. Kageyama, M. J. Genet and B. Delmon, J. Catal., 1993, 144, 175–192. [47] G. Busca, H. Saussey, O. Saur, J. C. Lavalley and V. Lorenzelli, Appl. Catal., 1985, 14, 245–260. [48] B. Erdem, R. A. Hunsicker, G. W. Simmons, E. D. Sudol, V. L. Dimonie and M. S. El-Aasser, Langmuir, 2001, 17, 2664–2669. [49] G. P. López, D. G. Castner and B. D. Ratner, Surf. Interface Anal., 1991, 17, 267–272. [50] J. Stoch and J. Gablankowska-Kukucz, Surf. Interface Anal., 1991, 17, 165–167. [51] M. P. Casaletto, A. Longo, A. Martorana, A. Prestianni and A. M. Venezia, Surf. Interface Anal., 2006, 38, 215–218. [52] V. A. Bondzie, S. C. Parker and C. T. Campbell, Catal. Lett., 1999, 63, 143–151. [53] M. Primet, P. Pichat and M. V. Mathieu, J. Phys. Chem., 1971, 75, 1216–1220. [54] H. Lin, J. Long, Q. Gu, W. Zhang, R. Ruan, Z. Li and X. Wang, Phys. Chem. Chem. Phys., 2012, 14, 9468–9474. [55] P. Du, A. Bueno-López, M. Verbaas, A. R. Almeida, M. Makkee, J. A. Moulijn and G. Mul, J. Catal., 2008, 260, 75–80. [56] C. Deiana, E. Fois, S. Coluccia and G. Martra, J. Phys. Chem. C, 2010, 114, 21531–21538. [57] D. Widmann and R. J. Behm, Angew. Chem. Int. Ed., 2011, 50, 10241–10245. REFERENCES 149 [58] T. Diemant, Z. Zhao, H. Rauscher, J. Bansmann and R. J. Behm, Top. Catal., 2007, 44, 83–93. [59] M. Boronat, P. Concepción and A. Corma, J. Phys. Chem. C, 2009, 113, 16772–16784. [60] A. A. Efremov, Yu. D. Pankratiev, A. A. Davydov and G. K. Boreskov, React. Kinet. Catal. Lett., 1982, 20, 87–91. [61] K. Chakarova, M. Mihaylov, S. Ivanova, M. A. Centeno and K. Hadjiivanov, J. Phys. Chem. C, 2011, 115, 21273–21282. [62] A. A. Efremov and A. A. Davydov, React. Kinet. Catal. Lett., 1980, 15, 327–331. [63] R. C. Lord and P. Venkateswarlu, J. Opt. Soc. Am., 1953, 43, 1079–1085. [64] J. G. Radziszewski, J. W. Downing, M. S. Gudipati, V. Balaji, E. W. Thulstrup and J. Michl, JACS, 1996, 118, 10275–10284. [65] A. M. Argo, J. F. Goellner, B. L. Phillips, G. A. Panjabi and B. C. Gates, JACS, 2001, 123, 2275–2283. [66] T. A. Nijhuis, E. Sacaliuc-Parvulescu, N. S. Govender, J. C. Schouten and B. M. Weckhuysen, J. Catal., 2009, 265, 161–169. [67] J. A. van Bokhoven, C. Louis, J. T. Miller, M. Tromp, O. V. Safonova and P. Glatzel, Angew. Chem. Int. Ed., 2006, 45, 4651–4654. [68] N. Weiher, A. M. Beesley, N. Tsapatsaris, L. Delannoy, C. Louis, J. A. van Bokhoven and S. L. M. Schroeder, JACS, 2007, 129, 2240–2241. [69] J. J. Bravo-Suárez, K. K. Bando, J. Lu, T. Fujitani and S. T. Oyama, J. Catal., 2008, 255, 114–126. [70] K. A. Davis and D. W. Goodman, J. Phys. Chem. B, 2000, 104, 8557–8562. [71] H. M. Ajo, V. A. Bondzie and C. T. Campbell, Catal. Lett., 2002, 78, 359–368. [72] J. J. Bravo-Suárez, J. Lu, C. G. Dallos, T. Fujitani and S. T. Oyama, J. Phys. Chem. C, 2007, 111, 17427–17436. [73] W. T. Wallace and R. L. Whetten, JACS, 2002, 124, 7499–7505. [74] A. Prestianni, A. Martorana, F. Labat, I. Ciofini and C. Adamo, J. Phys. Chem. B, 2006, 110, 12240–12248. [75] M. Boronat and A. Corma, Dalton Trans., 2010, 39, 8538–8546. [76] I. X. Green, W. Tang, M. Neurock and J. T. Yates, Science, 2011, 333, 736–739. Conclusions and outlook 6 6.1 Conclusions The usage of a microreactor system for the direct epoxidation of propene over a gold– titania-based catalyst system using a mixture of hydrogen, oxygen, and propene allows for the safe operation of the reaction in the explosive regime. The kinetic study performed over a very wide range of reactant concentrations in the capillary reactor as discussed in Chapter 2 provides the mechanistic insights for both catalyst development and optimization of the operation. The formation rate of propene oxide is most dependent on the hydrogen concentration, confirming that the dissociative adsorption of hydrogen is the rate determining step in the direct epoxidation of propene. The formation of an active hydroperoxo species on the gold nanoparticles, which is responsible for the propene epoxidation, is thus influenced by this rate determining step. The dependencies of PO formation rate over the 1 wt% Au/Ti-SiO 2 catalyst investigated are ca. 0.5, 0.3 and 0.3 for hydrogen, oxygen and propene, respectively. Although performing the direct epoxidation of propene at higher reactant concentrations can increase the formation rate of PO further, a higher hydrogen concentration also plays an adverse role in the utilization efficiency of hydrogen, which leads to much more water formation than the extra PO that can be gained. The active hydroperoxo species responsible for PO formation is competitively consumed by hydrogenation and epoxidation at the Au–Ti interface. The apparent activation energy for the water formation at the Au–Ti interface is determined to be 22 kJ/mol higher than the PO formation, which implies that a lower reaction temperature 152 CHAPTER 6. CONCLUSIONS AND OUTLOOK Table 6.1: Qualitative guideline in reactant concentrations for a better catalyst performance H2 ↑ O2 ↑ C3 H6 ↑ PO productivity ++ + + Hydrogen efficiency −− +/0 + Catalyst stability − 0 + is preferred for a better hydrogen efficiency. Besides the water formation via the hydrogenation of the active hydroperoxo species responsible for PO formation, there is an extra route for hydrogen oxidation, i.e., so called ‘direct water formation’. The direct water formation happens very likely on gold not adjacent to a Ti(IV) site. A higher propene concentration is beneficial to reducing both the hydrogenation of the active hydroperoxo species at the Au–Ti interface and the direct water formation. In regard to the catalyst deactivation, it is found that the deactivating species blocks the Au–Ti interfacial sites and diminishes both water formation and PO formation. The rate of deactivation and the hydrogen efficiency can be correlated. A low hydrogen efficiency usually corresponds to a higher deactivation rate. A lower hydrogen concentration or a higher propene concentration has a positive effect on improving the hydrogen efficiency. An enhanced productivity of propene oxide was achieved by adjusting the gold– titanium synergy over catalysts supported on titanium-grafted silica. The original idea was to reduce the amount of gold sites that are not in proximity to a Ti(IV) site to mitigate the unwanted direct water formation. Highly isolated titanium sites were obtained by lowering the amount of titanium grafted on silica. The tetrahedrally coordinated titanium sites were found to be favourable for attaining small gold nanoparticles and thus −1 a high dispersion of gold. A PO productivity as high as 100gPO · kg−1 can be easily cat h achieved on gold catalysts supported on the Ti-SiO2 with mostly tetrahedrally coordinated Ti(IV) sites. The improved productivity of propene oxide can be attributed to the increased amount of the interfacial Au–Ti sites and less deactivation. There is no need for a high gold loading to obtain a high PO productivity. A gold loading smaller than 0.2 wt% is sufficient for high PO formation rates as long as enough Au–Ti interfacial sites can be created by a dedicated method of gold deposition. The active hydroperoxy intermediate is competitively consumed by epoxidation and hydrogenation at the Au–Ti interface, which has been confirmed again on the catalysts with invisible gold clusters. A 6.1. CONCLUSIONS 153 higher propene concentration is favourable for a lower water formation rate and a higher formation rate of propene oxide. The hydrogen efficiency achieved on the catalysts with high PO productivity was with the range of 10–20 %. Under certain circumstances, propane formation may also happen or even prevail over the gold–titania catalysts under the conditions for the direct propene epoxidation. Propene hydrogenation was encountered during our study into the site synergy between gold and titanium using Ti-SiO2 and TS-1 as the supports. It is found that propene hydrogenation can be completely suppressed by adding a small amount of carbon monoxide while the propene epoxidation was not affected. Further investigation showed that the support itself plays an important role in propene hydrogenation, since the gold particle size varied between 1–6 nm on different catalysts and gold is not necessary for this side reaction. Catalysts with similar gold/titanium loadings can give different performance in propene hydrogenation. Gold on the catalysts producing propane only can hardly be detected by CO adsorption while an in-situ XPS study showed that the gold is in its metallic form. The supports alone showed the same hydrogenation behavior as the gold–titania catalysts: 1) enhancement of propene hydrogenation by dioxygen; 2) peak activity at ca. 443 K accompanying the process of dehydroxylation during temperature programmed reaction in only propene and hydrogen ; 3) switching off by CO with an order of −1. The activity in propene hydrogenation deactivates by flushing in propene and can be restored by a treatment in a mixture of hydrogen and oxygen. The hydroxylated catalyst can hydrogenate propene in the absence of oxygen but the dehydroxylated one cannot. An investigation by infrared spectroscopy confirmed that the effect of propene flush on the loss of activity in propene hydrogenation is merely related to a loss of surface hydroxyls. Although a few papers argue that oxidized gold may be responsible for propene hydrogenation [1, 2], unraveling the general oxidation state of gold is important to the understanding of the direct propene epoxidation on the gold–titania catalysts. Carbon monoxide was used as probe molecule in the infrared study to investigate the electron density of low-coordinated gold atoms on the gold–titania catalysts that are active in the direct propene epoxidation. The active gold sites were fully covered by reaction intermediates and deactivating species after the reaction. These species occupying the gold sites could not desorb even at 573 K. Calcination in oxygen removed the carbonaceous species on gold. The gold atoms were positively charged when oxygen was adsorbed on gold 154 CHAPTER 6. CONCLUSIONS AND OUTLOOK Table 6.2: Typical operating conditions of an industrial EO (ethylene oxide) reactor and state-of-the-art performance for a PO process based on gold catalysts EO [3] PO 0.15–0.40 0.10 O2 0.05–0.09 0.10 H2 – 0.10 Molar fraction at inlet C2 H4 /C3 H6 One-pass conversion 7–15 % 10 % Selectivity 80 % 90 % Space velocity 4500–7500 h−1 4000 mL·g−1 h−1 cat 3 Space-time yield 0.13–0.26 g/h/cm 0.10 g/h/gcat Temperature 500–550 K 423–473 K Pressure 1.0–2.2 MPa 0.1 MPa or at the interface. Reduction in hydrogen removed the adsorbed oxygen and the positively charged gold was reduced to its metallic form. When propene was adsorbed on the catalyst, gold atoms were negatively charged showing the carbonyl band as low as 2079 cm−1 . Carbon monoxide was replaced by propene on the catalyst surface and oxidation of carbon monoxide was suppressed by propene. Hydrogen significantly increased the coverage of carbon monoxide on the titania surface at low temperatures. 6.2 Outlook After more than a decade of research, the direct propene epoxidation over gold–titania catalysts has progressed significantly in terms of PO productivity and short-term stability over different supports as summarized in Table 1.2. The research goal for this catalytic system set by the group of Haruta [4] is to have one-pass conversion of 10 %, selectivity of 90 % and a hydrogen efficiency of 50 %. However, an economic feasibility study needs to be solidified based on comparison with PO/TBA, SMPO and HPPO processes just as one performed by Ghanta et al. [5] for a new PO process. From the engineering point of view, the direct PO process based on the gold catalyst will be no much difference from the existing EO production based on the silver catalyst, nor different from other gas-phase oxidation processes based on fixed-bed reactor technology. A high propene concentration will be implemented to overcome the issue of hydrogen efficiency and safety when using the hydrogen/oxygen/propene mixture. The product stream would contain a significant 6.2. OUTLOOK 155 amount of unreacted gas, including propene, hydrogen, and oxygen, which needs to be recycled. Water and PO would be very likely condensed at a relatively low temperature in the first step of gas separation since the boiling point (at atmospheric pressure) of PO is only 307 K. To overcome the energy consumption by refrigeration, the reactor and the first separator would be preferably operated at a few bars. The following separation step would be water–PO separation, likely, by fractional distillation. Nevertheless, the bottleneck of the whole process still remains at the catalyst development phase to address more efficient utilization of hydrogen and long-term stability. It is also very important for the industry to experimentally determine the operation window outside the explosive region. In our earliest kinetic study, which is not included in this thesis, the catalyst activity for water formation doubled after storage of a few months, but there was no deactivation in PO formation observed. Possible cause of this problem may be attributed to hydroxylation of the support, which may be supported by Figure 3.2. Considering the relatively large amount of water co-produced in the direct epoxidation of propene, one may pay additional attention to improving the hydrogen efficiency to avoid a long-term change of catalyst in a humid environment at relatively high temperatures. Hydrophobic supports are preferred. The highest records in the hydrogen efficiency are exclusively from the TS-1 support as seen in Table 1.2. Thus developing a type of TS-1 crystal with a specific flat shape may be helpful to enhance gold dispersion. Another option for a hydrophobic support is hybrid organic/inorganic Ti-containing materials [6, 7], in which the silanols are replaced or passivated by alkoxy groups. This kind of hybrid supports, when used in the direct propene epoxidation, showed a high PO yield approaching 10 % with very limited deactivation for as long as 10 days [6] due to the absence of surface hydroxyls. Although not much data about hydrogen efficiency has been reported on gold catalysts supported on this type of supports, it may be interesting to expand work into this area. The competitive hydrogenation and epoxidation at the Au–Ti interface consuming the active hydroperoxo species are very similar to what is happening in the direct synthesis of hydrogen peroxide over Au–Pd catalysts [8]. It was found that hydrogenation of hydrogen peroxide over the Au–Pd catalysts can be switched off [8]. This may also provide alternatives to improve the hydrogen efficiency since hydrogen peroxide is often supposed to be the intermediate in the direct propene epoxidation over gold–titania catalysts. 156 CHAPTER 6. CONCLUSIONS AND OUTLOOK References [1] S. T. Oyama, J. Gaudet, W. Zhang, D. S. Su and K. K. Bando, ChemCatChem, 2010, 2, 1582– 1586. [2] J. Gaudet, K. K. Bando, Z. Song, T. Fujitani, W. Zhang, D. S. Su and S. T. Oyama, J. Catal., 2011, 280, 40–49. [3] H. Kestenbaum, A. Lange de Oliveira, W. Schmidt, F. Schüth, W. Ehrfeld, K. Gebauer, H. Löwe, T. Richter, D. Lebiedz, I. Untiedt and H. Züchner, Ind. Eng. Chem. Res., 2002, 41, 710–719. [4] A. K. Sinha, S. Seelan, S. Tsubota and M. Haruta, Angew. Chem. Int. Ed., 2004, 43, 1546–1548. [5] M. Ghanta, D. R. Fahey, D. H. Busch and B. Subramaniam, ACS Sustainable Chem. Eng., 2013, 1, 268–277. [6] P. Vogtel, G. Wegener, M. Weisbeck and G. Wiessmeier (Bayer Aktiengesellschaft), WO 2001041921, 2001. [7] C. Qi, J. Huang, S. Bao, H. Su, T. Akita and M. Haruta, J. Catal., 2011, 281, 12–20. [8] J. K. Edwards, B. Solsona, E. Ntainjua N, A. F. Carley, A. A. Herzing, C. J. Kiely and G. J. Hutchings, Science, 2009, 323, 1037–1041. List of Publications Journal publications • Chen, J., Pidko, E.A., Ordomsky, V., Verhoeven, M.W.G.M., Hensen, E.J.M., Schouten, J.C., Nijhuis, T.A. (2013). How metallic is gold in the direct epoxidation of propene : An FTIR study. Catalysis Science & Technology, DOI: 10.1039/c3cy00358b. • Chen, J., Halin, S.J.A., Pidko, E.A., Verhoeven, M.W.G.M., Perez Ferrandez, D.M., Hensen, E.J.M., Schouten, J.C., Nijhuis, T.A. (2013). Enhancement of catalyst performance in the direct propene epoxidation : A study into gold-titanium synergy. ChemCatChem, 5, 467– 478. • Chen, J., Halin, S.J.A., Perez Ferrandez, D.M., Schouten, J.C., Nijhuis, T.A. (2012). Switching off propene hydrogenation in the direct epoxidation of propene over gold–titania catalysts. Journal of Catalysis, 285, 324–327. • Chen, J., Halin, S.J.A., Schouten, J.C., Nijhuis, T.A. (2011). Kinetic study of propylene epoxidation with H2 and O2 over Au/Ti-SiO2 in the explosive regime. Faraday Discussions, 152, 321–336. • Nijhuis, T.A., Chen, J., Kriescher, S.M.A., Schouten, J.C. (2010). The direct epoxidation of propene in the explosive regime in a microreactor – A study into the reaction kinetics. Industrial & Engineering Chemistry Research, 49, 10479–10485. The author also contributed to the following publications outside the scope of this thesis: • Wu, C., Chen, J., Cheng, Y. (2010). Thermodynamic analysis of coal pyrolysis to acetylene in hydrogen plasma reactor. Fuel Processing Technology, 91, 823–830. • Chen, J., Cheng, Y. (2009). Process development and reactor analysis of coal pyrolysis to acetylene in hydrogen plasma reactor. Journal of Chemical Engineering of Japan, 42, 103– 110. 158 PUBLICATIONS • Chen, J., Cheng, Y., Xiong, X., Wu, C., Jin, Y. (2009). Research progress of coal pyrolysis to acetylene in thermal plasma reactor. Chemical Industry and Engineering Progress (China), 28, 361–367. • Cheng, Y., Chen, J.Q., Ding, Y.L., Xiong, X.Y., Jin, Y. (2008). Inlet effect on the coal pyrolysis to acetylene in a hydrogen plasma downer reactor. Canadian Journal of Chemical Engineering, 86, 413–420. • Cheng, Y., Chen, J., Ding, S. (2007). Controlled synthesis of nano-sized TiO2 powders using high-temperature vapor phase process. Journal of Chemical Industry and Engineering (China), 58, 2103–2109. Conference presentations • Chen, J., Nijhuis, T.A., Schouten, J.C. (2012). The role of support in propene hydrogenation over gold-titania catalysts. The 6th International Conference on Gold Science Technology and its Applications (GOLD 2012), 5–8 September, 2012, Tokyo, Japan. [Poster] • Chen, J., Ordomskiy, V., Schouten, J.C., Nijhuis, T.A. (2012). Probing the active site of propene hydrogenation in the direct propene epoxidation over gold–titania catalysts. The 15th International Congress on Catalysis (ICC 2012), 1–6 July, 2012, Munich, Germany. [Oral] • Chen, J., Nijhuis, T.A., Schouten, J.C. (2012). The inhibition effect of carbon monoxide on propene hydrogenation over gold-titania catalysts. The 13th Netherlands’ Catalysis and Chemistry Conference (NCCC XIII), 5–7 March 2012, Noordwijkerhout, the Netherlands. [Poster] • Chen, J., Nijhuis, T.A., Schouten, J.C. (2011). Kinetic study of direct propylene epoxidation over Au/Ti-SiO2 in the explosive regime. The 12th Netherlands’ Catalysis and Chemistry Conference (NCCC XII), 28 February – 2 March, 2011, Noordwijkerhout, the Netherlands. [Oral] • Chen, J., Halin, S.J.A., Nijhuis, T.A., Schouten, J.C. (2011). Kinetic study of the epoxidation of propene over gold–titania catalysts into the explosive regime. The 22nd North American Catalysis Society Meeting (NAM22), 5–10 June, 2011, Detroit, USA. [Poster] • Chen, J., Nijhuis, T.A., Schouten, J.C. (2010). Direct epoxidation of propylene to propene oxide in a microreactor. The 11th Netherlands’ Catalysis and Chemistry Conference (NCCC XI), 1–3 March, 2010, Noordwijkerhout, the Netherlands. [Poster] • Chen, J., Nijhuis, T.A., Schouten, J.C. (2010). Direct epoxidation of propylene to propene oxide in a micro reactor. Netherlands Process Technology Symposium 2010 (NPS-10), 25–27 October, 2010, Veldhoven, the Netherlands. [Oral] Acknowledgements Finally the research project came to an end where I am so grateful to all the people who have helped me, encouraged me and inspired me through the whole journey. I enjoyed the great time here in Eindhoven and learned a lot from all of you. I would like to thank my promoter Prof.dr.ir. J.C. Schouten. Jaap, thanks for providing me such a great opportunity to pursue my doctor’s degree in the group SCR. You were always helpful and open-eared in our progress meetings. I appreciate your trust and always constructive suggestions which lead me on the way towards the end of the project. My special thanks are for my daily supervisor Dr.ir. T.A. Nijhuis. Xander, I benefited a lot from the fruitful discussions with you, from your expertise, enthusiasm and optimism. You showed me into the wonderful world of catalysis since the beginning of the project. You are so peaceful in mind, patient, always approachable and never put pressure on your students. I really appreciate your saying of being prepared to be an independent researcher. Xander, many thanks for your guidance all the way. Besides, I would also like to thank my former daily co-supervisor, Prof.dr. Evgeny Rebrov. Evgeny, although you’ve been promoted and left Eindhoven to Belfast, thank you very much for your scientific input through discussion and progress meetings in the early phase of my project. I’m very grateful to the skillful craftsmen in the workshop. Without the timely completion of small or big changes on my setup with high quality by you guys, I would not be able to implement my ideas and would not have enough and reliable data to complete my thesis. Erik, you are so professional, efficient and always willing to help. Thank you so much for all the efforts you made. I would also like to thank Dolf, Theo, Chris and Madan for your help whenever I bothered you guys in the workshop. In addition, I want to thank Anton Bombeeck, our technical coordinator. Anton, I bothered you also quite a lot and thanks for your help with making orders and hunting treasures in the storage room. 160 ACKNOWLEDGEMENTS My heartfelt thanks are given to Peter, Carlo and Marlies. Peter, thank you so much for all the help that you had offered me. Carlo, thanks a lot for the automation work done by you and I really appreciate your professionalism on safety. Marlies, you always helped me solve the GC problems in the first place, many thanks. I would like to thank people who helped me with different spectrometers and analytical tools. I am very grateful to Prof.dr.ir. Emiel Hensen for allowing me to use the IR spectrometer in your group. I want to express my sincere gratitude to Dr. Evgeny Pidko for the time and energy that you spent in my IR experiments and for your tolerance to my mistakes. I am also much obliged to Tiny Verhoeven for performing all the XPS analyses and for your helpful discussion. And again, Carlo Buijs, thank you for taking countless TEM and SEM pictures for me. As well as my lab mate, Serdar, thank you very much for your time on my TEM samples. In addition, I would also like to thank Prof.dr.ir. René Janssen for EPR measurements. I’d like to thank my two master students, Stefanie Kriescher and Sander Halin. Thank you for your contribution to the experimental work of this thesis. Both of you are diligent not only in academic research, but also in sports, where you impressed me by persistence and self-discipline, as well as your achievements. And Denise, it is very kind of you to help me arrange all the administrative documents. Whenever I needed help, you were always there and gave me directly the answer. I want to express my deep gratitude to you for all the things you’ve done for me all the years. I’d also like to thank John, Mart and Martin. Thank you all for making SCR a splendid place for research and you helped make it possible for me to meet so many excellent and young researchers from all over the world. I spent a lot of enjoyable time with my colleagues in SCR. Ma’moun, Marco and Christine, my dear office mates in STW 1.38, thank you for the memorable time we spent together and also thanks for all the help you provided me in either research or daily life. Dulce and Serdar, my dear lab mates, I really enjoyed our discussion and chat in the lab. As well as Jack, thanks a lot for your patient explanation about your delicate setup even when you were busy in the last phase of your PhD. Vitaly, I really appreciate your insightful input in IR experiments and your exemplary role of hard working. The same gratitude is given to my lovely colleagues for the pleasant time in SCR: Bianca, Bruno, Carlos, Chattarbir, Emila, Faysal, Fernanda, Frans, Gregory, Halabi, Ivana, Joost, Jovan, Kevin, Lara, Lidia, Maria, Michiel, Minjing, Narendra, Nopi, Oki, Paola, Patrick, Roman, Slavisa, Stijn, Shohreh, Tom, Violeta. I’d also like to thank the people from the third ACKNOWLEDGEMENTS 161 floor: Chaochao, Leilei, Sami, Guanna, Michel. Thank you all for the assistance when using the equipment in your group. There are many people who have enriched my life here in the Netherlands. The name list is not short: Gao Yang, Harrison, Donglin, Jinbao, Wang Qi and Chuanbo, Yue Jun couple, Qingling and Maarten, Pengzhao, Sijun and Jijing, Yongjian couple, Xixi, Danqing, Gu Xi, Xiaoran, Changwen couple, Xueqing couple, Junlin and Wang Juan, Piming couple, Delei, Lianghui, Wu Jing, Xu Wei couple, Jogesh, Dries, Elham, Lou Xianwen, Yu Fangli, Zhang Yanmei, Gao Lu, Ma Zhe, Wenhao. It is destiny to meet you guys in the boundless huge crowd here in the Netherlands. I really cherish this experience. Special thanks are given to Shih-chin and Te-hui. You showed me endless love and care and how life can always be positive and definite. I also want to thank David and Prisca, Sandy and other saints met in Eindhoven for sharing and encouragement. Furthermore, I am very grateful to my former supervisor Prof. Yi Cheng. Without your encouragement, I would not take the first step to embrace the challenging but also wonderful experience in the Netherlands. As well as sister Gong Weiping, thank you for all the support during my stay in Beijing and your hospitality in Brussels. I also want to thank my college mates, Baoxiang, Chen Zhen, Chen Bo and Hangzhou for the fellowship in the past years. Huihui, my beloved wife, it is lucky for me to have you, your continuous loving and caring. You are magic and can always make my troubles disappear. I don’t know how to express my gratitude to you in words. You and I will grow together and I am forever in your debt, that’s for sure. My dear parents, thank you so much for your understanding and support, your continuous teaching and sharing. I also want to express my deep gratitude to my parents in law. Thank you for your trust and support. I appreciate all the support, understanding and encouragement from my family members and friends. 我深 深地感谢我的爸爸妈妈,谢谢你们这些年来的关心、支持和分享。 我同样感谢我的 岳父和岳母,谢谢你们的信任和支持。最后,我真诚地感谢一直以来家人和朋友们 对我的鼓励和支持。 About the Author Jiaqi Chen (陈家琦) was born on January 27, 1984, in Rudong, Jiangsu Province, China. After finishing the secondary school in 2001, he started his study in the department of Chemical Engineering at Tsinghua University, Beijing. In 2005, he worked on the topic ‘Controlled synthesis of titanium dioxide nanoparticles in a flame reactor’ for his graduation project under the supervision of Prof. Yi Cheng and obtained his bachelor’s degree in July. Afterwards, he continued with the master programme in the group of reaction engineering in the same department. In 2008, he obtained his master’s degree after working on a joint project ‘Coal pyrolysis to acetylene in a hydrogen plasma reactor’ with Xinjiang Tianye Group. After having spent 7 years in Beijing, he started off to Eindhoven, the Netherlands. From 2008 to 2012, he worked as a PhD student on ‘propylene epoxidation over gold catalysts’ in the Laboratory of Chemical Reactor Engineering in Eindhoven University of Technology under the supervision of Prof.dr.ir. J.C. Schouten and Dr.ir. T.A. Nijhuis. His PhD project involved mainly the kinetic study, catalyst development and mechanistic investigations into the bi-functional gold–titania system. Since October 2012, he has worked as a researcher in Shell Global Solutions in Amsterdam.