+ All Categories
Home > Documents > A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance...

A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance...

Date post: 10-Sep-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
255
A spectroscopic characterization of the structure of supported metal catalysts Citation for published version (APA): Martens, J. H. A. (1988). A spectroscopic characterization of the structure of supported metal catalysts. Technische Universiteit Eindhoven. https://doi.org/10.6100/IR282367 DOI: 10.6100/IR282367 Document status and date: Published: 01/01/1988 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 31. Dec. 2020
Transcript
Page 1: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

A spectroscopic characterization of the structure of supportedmetal catalystsCitation for published version (APA):Martens, J. H. A. (1988). A spectroscopic characterization of the structure of supported metal catalysts.Technische Universiteit Eindhoven. https://doi.org/10.6100/IR282367

DOI:10.6100/IR282367

Document status and date:Published: 01/01/1988

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 31. Dec. 2020

Page 2: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

A Spectroscopic Characterization of the

Structure of Supported Metal Catalysts

J. H. A. Martens

Page 3: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

A Spectroscopic Characterization of the

Structure of Supported Metal Catalysts

Page 4: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- Ill -

A Spectroscopic Characterization of the

Structure of Supported Metal Catalysts

F:en Spectroscopische Karakterisering van de

Struktuur van Gedragen MetaalkaJalysaJoren

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de

Technische Universiteit Eindhoven, op gezag van

de rector magnificus, prof. dr. F .N. Hooge, voor

een commissie aangewezen door het college van

dekanen in het openbaar te verdedigen op

dinsdag 22 maart 1988 te 14.00 uur

door

J. H. A . Martens

geboren te Elsloo

Page 5: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- IV -

Dit proefschrift is goedgekeurd door de pronwtoren:

prof. dr. R. Prins

en

prof. dr. ir. D. C. Koningsberger

The research reported in this thesis has been carried out at the Laboratory of Inorganic Chemistry and Catalysis at the Eindhoven University of Technology and has been supported by the Nether­lands Foundation for Chemical Research (SON) with financial aid from the Netherlands Organization for the Advancement of Pure Research (ZWO).

Page 6: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- v -

aan mijn vader

en aan mijn moeder

aan Angeliene

Page 7: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- VI -

Contents

Chapter 1 Introduction 1 1.1 Catalysis 1 1.2 Heterogeneous Catalysis 2 1.3 Characterization of Supported Catalysts 7 1.4 Scope of this Thesis 8 1.5 References 8

Chapter 2 Experimental Techniques 11 2.1 Catalyst Preparation 11

2.1.1 Introduction 11 2.1.2 Pore Volume Impregnation 13 2.1.3 Ion Exchange 14 2.1.4 The Urea Method 15

2.2 Temperature Programmed Reduction 15 2.3 Hydrogen Chemisorption 17 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21 2.7 Laser Raman Spectroscopy 23 2.8 ASED-MO Computations 25 2.9 EXAFS 28

2.9.1 Basic Principles 28 2.9.2 Fourier Transformations 34 2.9.3 Reference Compound and 37

Calculating Spectra 2.9.4 Data Analysis 38 2.9.5 Experimental Method 43

2.10 References 44

Chapter 3 The Preparation of y -Al20 3 supported Monometallic 47 Rh and Pt and Bimetallic Rh-Pt Catalysts 3.1 Introduction 47 3.2 Experimental 48

3.2.1 NMR and Laser Raman Experiments 48 3.2.2 Adsorption Experiments 49 3.2.3 TPR of Rh, Pt and Rh-Pt/ A1 20 3 50

Page 8: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- VII -

3.3 Results and Discussion 51 3.3.1 NMR and Laser Raman Experiments 51 3.3.2 Adsorption Experiments 56 3.3.3 TPR of Rh, Pt and Rh-Pt/ Al 20 3 67

3.4 Conclusions 71 3.5 References 72

Chapter 4Ferric Iron in Reduced Si0 2 Supported 73 Fe-Ru and Fe-Pt Catalysts Evidence from Mt>ssbauer Spectroscopy and Electron Spin Resonance 4.1 Introduction 73 4.2 Experimental 75 4.3 Results and Discussion 76 4.4 Conclusions 81 4.5 References 81

Chapter 5Controlled Oxygen Chemisorption on an 83 Alumina Supported Rhodium Catalyst The Formation of a Metal-Metal Oxide Interface Determined by EXAFS 5.1 Introduction 83 5.2 Experimental 84 5.3 Results 86 5.4 Discussion 90

5.4.1 Rh/Al 20 3 after Reduction and Evacuation 90 5.4.2 A Model for the Oxidation of Metal Particles 95 5.4.3 Rh/ Al 20 3 after Oxygen Admission at 100 K 101 5.4.4 Rh/ A1 20 3 after Warming up to 300 K 102 5.4.5 General Remarks 104

5.5 Conclusions 108 5.6 References 109

Chapter 6 The Structure of the Metal-Support Interface 111 in Rh/ A1 20 3 Determined with the AS ED-MO Method 6.1 Introduction 111 6.2 Theoretical Method 112 6.3 Description of the Model and Results 114

Page 9: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- VIII -

6.4 Discussion 117 6.5 Conclusions 120 6.6 References 121

Chapter 7 The Structure of Rh/Ti0 2 in the Normal and the 123 SMSI State as Determined by EXAFS and HRTEM 7.1 Introduction 123 7 .2 Experimental 127

7.2.1 Catalyst Preparation 127 7.2.2 EXAFS Measurements 129 7.2.3 HRTEM Experiments 130

7.3 Results 132 7.3.1 Analysis of the EXAFS Spectra 132 7.3.2 Characterization with H RTEM 145

7.4 Discussion 147 7.4.1 Rh/Ti02 after Reduction at 473 K 147 7.4.2 Rh/Ti0 2 after Reduction at 723 K 151 7.4.3 Evacuation at 623 K 159

7.4.3.1 Rh/ Al20 3 159 7.4.3.2 Rh/Ti0 2 160

7.4.4 Oxygen Admission at 100 K 160 7.4.4.1 Rh/ Al20 3 160 7.4.4.2 Rh/Ti0 2 161

7.4.5 Oxygen Admission at 300 K 162 7.4.5.1 Rh/A120 3 162 7.4.5.2 Rh/Ti0 2 164

7.4.6 Different Rh0-Tin + Contributions 165 7.4.7 Comparison with Literature Data 167

7.5 Final Conclusions 170 7.6 References 173

Chapter 8 Strong Metal-Support Interactions 177 in Rh/Ti0 2 Prepared with Ion Exchange 8.1 Introduction 177 8.2 Experimental 179

8.2.1 Catalyst Preparation 179 8.2.2 EXAFS Measurements 179

8.3 Results 180

Page 10: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Chapter 9

- IX -

8.4 Discussion 184 8.4.1 Rh/Ti0 2 after Reduction at 494 K 184 8.4.2 Rh/Ti0 2 after Reduction at 623 K 187 8.4.3 Rh/Ti02 after Reduction at 773 K 190 8.4.4 General Remarks 191

8.5 Final Conclusions 194 8.6 References 195

EXAFS Evidence for Direct Rh0-Tan+ Bonding 197 and Coverage of the Metal Part icles in a Rh/Ta20 5 Catalyst in the SMSI State 9.1 Introduction 197 9.2 Experimental 199

9.2.1 Catalyst Preparation 199 9.2.2 EXAFS Measurements 200

9.3 Results 201 9.3.1 Reference Compounds 201 9.3.2 Analysis of the EXAFS Spectra 203

9.4 Discussion 207 9.4.1 Rh/Ta20 5 after Reduction at 523 K 207 9.4.2 Rh/Ta20 5 after Reduction at 858 K 210 9.4.3 Rh/Ta20 5 after Admission of 0 2 215

in the S MS I State 9.5 Final Conclusions 9.6 References

216 217

Chapter 10 Concluding Remarks 219

Summary 225

Samenvatting 231

Dankwoord 237

Curriculum Vitae 240

List of Publications 241

Page 11: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 1 Chapter 1

Chapter 1

Introduction

1.1 Catalysis

Chemistry is an area of vital importance in todays society, and

within chemistry. catalysis is a mainstay. not only in many indus­

trial applications. but also in numerous processes in the chemistry of life. Catalysis performs a key role in processes such as the

conversion of crude oil into a wide scale of useful products, in the preparation of many nutrients, in the conversion of coal via syn­

thesis gas into products like alcohols, gasoline, and in the removal

of noxious components from exhaust gases.

Catalysis is the science of accelerating chemical reactions that under normal conditions proceed only slowly or not at all. The rate

of a chemical reaction can be controlled by a few parameters only:

temperature, pressure and composition. lr:i addition, the choice of a suitable catalyst may change the reaction pathway . As a conse­

quence, the overall reaction rate may be increased and/or new path­

ways, and therefore new products, may become feasible. A chemi­

cal reaction is the result of a collision of two or more molecules or

atoms. The function of a catalyst is merely to capture the partici­pants of the reaction, to bring them in close contact and thus to

guide them through some reaction pathway . The combination of

catalyst and reactant( s) dictates the pathway. Two characteristics

are important in describing a catalyst. First of all its activity. that

is the rate at which the products are generated. The higher the

activity, the better the catalyst. Secondly, there is the selectivity. In most cases, a catalyst produces a (wide) range of products, some

of which are useful and others are not . By selectivity in general we

mean the fraction of (useful) products, usually expressed as a

Page 12: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Introduction page 2

percentage . A high selectivity indicates that the catalyst produces

mainly the desired products.

Catalysts are known in may varieties, but in principle they can

be classified into two categories : homogeneous and heterogeneous catalysts. Homogeneous catalysts can be mixed perfectly with the

reactants, i.e., up to molecular scale . Homogeneous catalysis mainly occurs in the liquid phase. The reactants and the catalyst

are liquids or dissolved in the liquid phase. All the enzymes at work

in our body are homogeneous catalysts . In heterogeneous catalysis.

the catalyst. the reactants and the products are in separate phases and therefore the mixing is far from perfect. The catalyst is usually

a solid and the reactants are liquids or gases. The automotive catalyst is an example of a heterogeneous catalyst. The reactants,

hydrogen carbon monoxide and nitric oxides, and the products, water, carbon dioxide and nitrogen, are gaseous while the active

catalyst is a combination of several precious metals supported on

some (inert) support.

1.2 Heterogeneous Catalysis

In this thesis we will describe heterogeneous catalysts contain­ing precious metals. We have already met one application of such

catalysts : the use of noble metals in automotive catalysis. Another

important use of metals is in the Fischer Tropsch synthesis. In this

process, carbon monoxide and hydrogen are combined to long-chain hydrocarbons and oxygen-containing organic molecules like alcohols

( 1-3). The way Fischer Tropsch catalysts and automotive catalysts work is very similar. The metals are capable of adsorbing the reac­

tants and of splitting them into smaller fragments or atoms. On the surface of a suitable metal, sometimes with the help of addi­

tives. these fragmented molecules rearrange, in the case of Fischer Tropsch synthesis to useful products and in the case of automotive

Page 13: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 3 Chapter 1

catalysis to harmless products . These products desorb and leave

the surface of the catalyst as gases. Here we have encountered on of the most important aspects of heterogeneous catalysis : only the

surface of the metals is exposed to the (gaseous or liquid) reactants and products. Therefore, only the metal atoms in the surface are

active in the process . The atoms directly beneath the surface

atoms may still play a (minor) role , while all the other atoms in the bulk are 'useless' . To give an impression of this 'waste' : in a metal

sphere or ball of size 1 mm, only one in about 1.3 million atoms is a

surface atom ; in a metal particle of one micron , one in .approxi­mately 1.3 thousand atoms is in the surface of the particle and

therefore potentially useful in a heterogeneous reaction. For inex­pensive and commonly available metals, this is hardly an objection .

For precious noble metals which are not only very expensive, but also sometimes very rare, this is worth consideration. The general

aim therefore is to reduce the size of the metal particles and conse­quently to use the metal as efficiently as possible. This can be

achieved by 'dispersing' the metal on an inert support material . A commonly used support is aluminum oxide, Al 20 3, known as

alumina. It has a typical surface area of several hundreds of square meters per gram . The surface area in a few grams of alumina (a tea spoon full) would typically cover the area of a football field . Several techniques are available to implant large amounts of 'highly dispersed ' metal particles on this surface. The size of these parti­cles is measured in nanometers (1 nm= 10-9 m) or in Angstroms

(1 A= 10-10 m). In most cases, the fraction of metal· atoms in the surface is close to unity and in the majority of cases above one half. Typically, the activity in Fischer Tropsch synthesis of one gram of alumina loaded with one percent rhodium exceeds the activity of

one gram pure rhodium powder by one or two orders of magnitude, while the price of the alumina supported catalyst is lower by about

two orders of magnitude.

Very small metal particles may not be metallic : their chemical

properties, i.e. , the properties of the surf ace atoms, may deviate from the properties of the surface atoms in larger metal particles

Page 14: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Introduction page 4

(4-6). Another important characteristic of small metal particles is,

that they are more easily influenced than larger metal particles . A

striking example is a phenomenon discovered in 1978 (7-9) . For

metal particles supported on oxides of transition metals. such as

titania (Ti0 2), vanadia (V 20 3) and tantalum oxide (T a20 5), two

different 'states' are accessible. There is a 'normal' state, in which

the properties of the metal particles are comparable to the proper­

ties of the same metal particles supported on inert oxides like

alumina (A1 20 3) and silica (Si0 2) . The surface area of such parti­

cles can be estimated by measuring the amount of gas that adsorbs

on the surface of the particles. The other state is known as the

Strong Metal-Support Interaction (SMSI) state. The properties

differ markedly from the properties of the 'normal' metal particles.

Most pronounced is the decrease of the amount of gas that the

metal particles can adsorb . In either state, however, the basic

structure of the particles is the same and the metal particles can be

brought from the normal into the SMSI state and vice versa. After

reduction at low temperatures the particles are in the 'normal' state

and a subsequent reduction at high temperatures induces the SMSI

state. After oxidation at mild temperatures (up to 500 K) and a

subsequent reduction at low temperature. the 'normal' state is

restored. Catalysts that can be brought into the SMSI state have

one characteristic in common : in the temperature regime up to

800 K the support can be reduced to a suboxide and this suboxide

can be re-oxidized to the original oxide. Thus, SMSI state and pres­

ence of suboxides go together. The reduction of the support is in

general catalyzed by the metal particles and is limited to the direct

environment of the metal particles. These reduced suboxides have

in general different properties than the original oxide . They may

have semi-conducting or even meta II ic properties (7-9), or they may

have an enhanced mobility ( 13-20). These special properties are

thought to be responsible for the SMSI state. Several models have

been proposed to explain the SMSI state. In the first model, already

adopted by Tauster and his co-workers (7-9), the origin for this

SMSI state is thought to be an enhanced interaction between metal

particles and supporting oxide, due to the electronic properties of

Page 15: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 5 Chapter 1

the suboxide . In one of the first papers on SMSI (JO) even a direct bonding between metal atoms and Tin+ ions in the support was

suggested. The formation of alloys is another model that may explain the incapability of adsorbing gases in the SMSI state ( 1-

3.11 >. In this model it is assumed that during the reduction of the

support (for example, Ti02) metallic titanium is formed which may diffuse into the metal particles and form alloy particles. It is known that the gas adsorption capacity of these alloys is very low ( 12). The third and last model that is important in explaining the SMSI

state is the coverage model. It is assumed that reduced support species have an increased mobility and may cover the metal parti­

cles, thus blocking the surface of the metal particles and decreasing their adsorption capacity. In many publications, evidence for cover­age has already been reported ( 13-20). In chapters 7, 8 and 9 of this thesis, we will discuss the SMSI phenomenon and its for origin

rhodium catalysts supported on Ti02 and Ta20 5.

In the above discussion, it was assumed that pure metals were used to guide reacting molecules along their reaction pathway. This pathway can be modified by introducing additives on, in and/or

beneath the surface of the metal particles . Like the support materi­als mentioned above, these additives influence the properties of (the

surface atoms of) the metal particles. We can discern two classes of additives : promoters and (other) metals. The difference between the two classes can be found in their activities towards the desired process in the absence of their host metal. A promoter is,

on its own, incapable of catalyzing the desired reaction. Combined with some active host metal, however, the activity of the host may

be increased significantly. Promoters are found among alkali (Li, Na, K) ( 21-26), rare earth and some transition metal or metal

oxides (V20 3, Mo0 3, Th02) (27,28) . The other class of additives is in fact the class of active metals itself. The intention here is to

'combine' the properties of two (or more) active metals. There are cases known where the combination of two metals is 'better' than the 'sum' of the two monometallic cases, better in terms of activity and/or selectivity. For example, for the Fischer-Tropsch synthesis,

Page 16: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Introduction page 6

an increase in methanol and ethanol selectivities for Fe-Rh/Si02 catalysts ( 29) and an increase in ethane and propene selectivities

for Co-Rh catalysts (JO) has been reported. The activity of bime­

tallics has been the subject of many studies . The major problem is to prepare metal particles that contain both metals and to verify

that indeed alloy formation has taken place. In most studies, TP R

is used to investigate .the formation of alloy particles. In preparing

bimetallic catalysts, two metal precursors are used. These precur­

sors may have a different reducibility. Thus, in Temperature Pro­grammed Reduction (T PR), they will be reduced separately. How­

ever, when they are co-impregnated, already during the impregna­

tion and subsequent drying step, particles containing both precur­

sors may have been formed and in general, these particles will be

reduced at the the reduction temperature of the component that is reduced most easily. This component, once reduced, can adsorb

and dissociate hydrogen and thus can catalyze the reduction of the

second component. In that case, TP R may indeed point to the for­

mation of bimetallic particles. Using TPR, it was found that in Cu- Ni/Si02 alloy particles were formed (3/ ,32). Evidence has also

been found for an intimate contact between the two constituent

metals in Pt-Re/ Al20 3 catalyst <33,34). Oxidation at high tempera­

tures caused segregation of platinum oxide and rhenium oxide. In

(35-37), the formation of bimetallic Co-Rh catalysts supported on

A1 20 3, Ti02 and Si0 2 has been described. It was found that rho­

dium aided the reduction af cobalt and that bimetallic particles were

formed. During oxidation, segregation occurred, but this did not

lead to the formation of monometallic particles during a subsequent

reduction. For Co-Rh/Ti02 (36), it was found that CoRh 20 4 was formed and that this mixed oxide was covered with Co30 4. Such a

segregation may also occur in metallic particles : the component with the lowest sublimation enthalpy will be present preferably in

the surface of the metal particles. For Co-Rh catalysts, it has been

reported that the outer shell of the alloy particles was enriched in cobalt ( 37 ). Segregation may be enhanced by the gas atmosphere;

this is known as gas induced surface enrichment. Clearly, bimetallic

catalysis is a delicate subject; the structure and composition of the

Page 17: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 7 Chapter 1

alloy particles can be affected in numerous ways . In chapters 3 and

4, some aspects of bimetallic Rh-Pt/ Al 20 3, Fe-Ru/Si0 2 and Fe­Pt/Si02 catalysts will be discussed.

1.3 Characterization of Supported Catalysts

In order to control and to steer the properties of supported metal particles , it is of prime importance to know and to under­

stand their structure. Once their structure has been related to the catalytic action of the catalyst, one may be able to develop 'better'

catalyst systems in a scientific manner. In determining the struc­ture of a heterogeneous catalyst, the support is the major obstacle.

The active metal is present on the 'internal' surface, 'inside' the porous support material. The number of metal particles visible with

an ordinary light microscopy is only a small fraction of the total amount of metal particles present in the specimen. This is a major disadvantage of characterization techniques that use radiation that cannot penetrate the support. In this thesis, we will describe the

use of radiation of high enough energy to penetrate the support and thus to reach the metal particles. EXAFS uses high energy X-ray

radiation, ESR and NMR use radio- and microwaves and IVlossbauer uses gamma radiation. However, the amount of support exceeds

the amount of metal by one or two orders of magnitude. As a result, the signal-to-noise ratio and the separation of signals may become a draw back. Another way 'around the support' is to use gases which penetrate in the pores, reach the metal particles and are adsorbed on the surface of the particles (hydrogen chemisorp­tion) or which in some way react with the metal particles (tempera­

ture programmed reactions such as reduction and oxidation). In chapter 2 these techniques will be discussed in more detail.

Page 18: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

lntrod uction page 8

1.4 Scope of this Thesis

In this thesis we will focus on determining the structure of

supported noble metal catalysts . Several interesting catalyst sys­

tems have been studied. Chapter 2 describes the techniques used

to prepare and study the catalysts . In chapter 3, preparational

aspects of alumina-supported monometallic rhodium and bimetallic

rhodium-platina catalysts will be discussed. Chapter 4 deals with

the intriguing presence of ferric iron (Fe3+) in Si02 supported bime­

tallic Fe-Ru and Fe-Pt catalysts, the existence of which has not

been realized for a long time. In chapter 5, the structure of an

alumina-supported rhodium catalyst during an oxidation process is

described . Chapter 6 deals with a computational approach of sup­

ported rhodium catalysts . In chapters 7, 8 and 9 EXAF S studies of

catalysts that suffer from metal-support interactions will be

highlighted : Rh/Ti0 2 and Rh/Ta20 5. In chapter 10, the results are

discussed in a wider context and where possible interrelated .

1.5 References

1. Fischer, F.; Tropsch , H. Brennstof Chemie 1923, 4, 276

2. Fischer , F.; Tropsch , H. Brennstof Chemie 1924, 5, 201

3. Fischer, F.; Tropsch , H. BrennstofChemie 1924, 6, 217

4. Yao, H. C. ; Yu Yao, Y F : Otto, K. J . Catal . 1978, 45, 120

5. Graydon , W F.; Langan , M. D. J . Catal. 1981 , 69 , 180

6. Hugues , F.; Besson , B.; Basset, J. M. J . Chem. Soc ., Chem. Comm. 1980, 719

7. Tauster, S. J. ; Fung, S. C.; Garten, R.L. J . Am. Chem. Soc . 1978, 100, 170.

8. Tauster, S. J .; Fung , S C. J. Catal. 1978, 55, 29.

Page 19: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 9 Chapter 1

9. Tauster , S J.; Fung, S C : Baker . R T. K : Horsley . J. A . Science

(Washington, DC) 1981, 211 . 1121.

10. Horsley. J. A . J . Am. Chem. Soc . 1979. 101 . 2870

11 . Beard , B C; Ross , P N. J . Phys. Chem . 1984. 90 , 681 1.

12. Brewer, L. "Phase Stability in M etals and Alloys" : Rudman . P.;

Stringer , J.; Jaffee, R , ed.; MacGraw-Hill: New York . 1967; pp 39-61.

13. Meriaudeau , P.; Dutel , J. F : Dufaux , M .; Naccache C " Studies of

Surface Science and Catalysis" 1982, 11 .

14. Belton , D. N.; Sun , Y -M : White , J. M . ]. Phys . Chem. 1984, 88 , 1690.

15. Belton, D. N. ; Sun , Y -M.; White , J. M. J . Phys . Chem . 1984, 88 ,

5172.

16. Simoens, A J.; Baker , R. T. K.; Dwyer. D. J.; Lund . C R. F : Madon. R. J. J . Catal 1984, 86. 359

17. Chung, Y M .; Xiong. G.; Kao, CC. J. Catal . 1984, 85 , 237 .

18. Sadeghi , H. R. ; Henrich, V. E. J. Catal . 1984, 87, 279.

19. Sun , Y -M .; Belton. D. N.; White , J. M . J . Phys . Chem . 1986, 90, 5178.

20. Ko, G. S.; Gorte , R. J. J . Catal 1984, 90, 59.

21. Dry , M . E. "Catalysis"; Anderson , J. R.; Boudart , M ., Eds.; Springer

Verlag, 1981, Vol. I, p. 159

22. Anderson, R. B. "Catalysis "; Emmet , P. H , Ed .; Reinhold , New

York, 1956, Vol. IV , p. 123.

23 . Kikuzono, Y.; Kagami , S.; Naito, S.; Onishi, T ., Tamaru, K . Far .

Disc. Chem. Soc. 1981 , 72, 135.

24. Vedage , G. A .; Himelfarb, P B.; Simmons , G. W .; Klier , K. Solid

State Chem . 1985 (ACS Symposium series 279, Graselli , R. K.; Braz­

dil. J. C.; Eds. )

25. Mori , T .; Masuda , H.; Imai , H.; Miyamoto, A .; Niizuma, H.; Hattori . T. ; Murakami , Y J. M olec . Catal . 1984, 25, 263.

26. Mori , T. ; Miyamoto, A. ; Takahashi , N.; Niizuma, H.; Hattori , T. ; Murakami , Y. J. Catal . 1986, 102, 199.

27. Mori, T.; Miyamoto, A .; Takahashi , N.; Fukagaya, M .; Hattori , T. ; Murakami , Y. J . Phys . Chem. 1986, 90, 5197.

Page 20: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Introduction page 10

28. Ichikawa, M.; Shikakura, K.; Kawai , M. "Heterogeneous Catalysis Related to Energy Problems" , Proc. Symp. Dalian , China . 1982 , A-08-1.

29. Bhasin , M. M. ; Bartley, W. J .; Ellgen, D. C. ; Wilson , T. P J . Catal . 1978 , 54 , 120.

30. Villiger , P.; Barrault, J.; Barbier , J .; Leclerq , G.; Maurel , R. Bull. Soc . Chim. Fr . 1979, 1-413.

31 . Robertson , S. D.; McNicol , B. D.; de Baas , J . H.; Kloet , S C. ; Jen­kins , J . W J. Catal . 1975 , 37, 424.

32. Jenkins, J . W ; McNicol, B. D.; Robertson , S D. Chem. Tech 1975, 7, 316.

33. Wagstaff , N.; Prins , R J . Catal . 1979, 59, 435.

34. Wagstaff , N.; Prins , R J . Catal . 1979, 59, 445.

35. van ·t Blik, H. F. J .; Prins , R. J . Catal . 1986, 97, 188

36. Martens , J . H. A. ; van ·t Blik, H. F. J .; Prins , R. J . Catal . 1986, 97,

200

37. van 't Blik, H. F. J .; Koningsberger , D. C. ; Prins, R. J. Catal . 1986, 97, 210

Page 21: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 11 Chapter 2

Chapter 2

Experimental Techniques

2.1 Catalyst Preparation

2.1.1 Introduction

For the properties such as activity, selectivity and stability of the eventual catalyst, the method of preparation is of crucial impor­

tance . Prime consideration- always is to keep the size of the metal particles within acceptable ranges . The smaller the metal particles,

the more active sites per gram of metal. Several techniques are known to bring forth small metal particles . The choice of the

preparation method depends on the metal and the support to be used . We will describe three preparation methods which were used

to prepare the catalysts that will be discussed in this thesis . The active material is in our case always a metal. It is impossible to

directly dispers a metal homogeneously on a porous support. In most cases and in all cases discussed in this thesis, a precursor,

usually a metal salt, is dissolved in a solute. This. solution can penetrate into the pores of the support and enable the metal precur­

sor to be deposited on the internal surface area of the support. This is the part of the preparation in which the three methods differ. In

paragraphs 2.1.2 , 2.1.3 and 2.1.4 we will discuss this deposition of metal precursor for the three methods separately.

After fixing the metal precursor onto the support, the solvent is removed by filtering and drying and the precursor remains in the pores of the support. Sometimes not only the metal salt but in

addition some residue originating from the solvent stays behind as well. To discard of this, a calcination step may be introduced, i.e.,

Page 22: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 12

the precursor catalyst is heated in air to a few hundred degree cen­

tigrade. After calcination or drying, the metal precursor is brought

into the active. metallic state by a reduction in hydrogen . After

reduction, the catalyst is highly active and exposing the catalyst without further precaution to air would result in a process known as

'run away oxidation'. Already during the start of the oxidation pro­

cess enough heat would be evolved to allow the oxidation process

to continue in an uncontrollable way. The temperature of the parti­

cles would reach a level at which the mobility of the particles is

high enough to start a sintering process, in which several smaller

metal particles join to form larger metal particles. This process is

of course to be avoided. Therefore, direct after reduction , the active

catalyst is flushed with nitrogen in order to remove all hydrogen.

Thereafter , oxygen is added carefully to the nitrogen feed . The

oxygen content is increased slowly up to a level of 20%. This pro­

cess is known as passivation : a careful and controlled oxidation of

the metal particles. For noble metal particles, passivation is in gen­

eral limited to the surface of the metal particles . The active metal is then covered by an oxide layer, is made 'passive' and can be stored

in air to await further study. A reduction in hydrogen at low tem­

peratures is in general enough to restore the catalyst to its active

state.

In the discussion above, the first and most important part of

the preparation, the introduction of the metal precursor onto the internal surface area of the support, has been skipped. We will now

continue to describe this essential part for three methods . The methods are the pore volume impregnation method, the ion

exchange method and the urea method.

Page 23: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 13 Chapter 2

2.1.2 Pore Volume Impregnation

The pore volume impregnation method is the method which is most frequently used in catalyst preparation because of the elegant simplicity of the method. The quantity of metal salt needed to prepare the catalyst is dissolved in the exact amount of solute needed to fill the pores of the support. We illustrate this by the fol­lowing example, in which we prepare 5 g of an alumina supported catalyst loaded with 2 wt% rhodium. The alumina used has an internal surf ace area of 180 m2 g-1 and a pore volume of 0.6 ml g- 1

.

The precursor is RhCl3·3H 20 with a molecular weight of 263.3 g. 5 g of the eventual catalyst will contain 4.9 g A1 20 3 and 0.1 g Rh, which is equivalent to 0.2559 g RhCl 3·3H 20 and has to be dissolved in 4.9*0.6 = 2.94 ml water. This solution is added slowly to the alumina, which is stirred vigorously in order to distribute the dis­solved precursor evenly over the support. Because of capillary forces, the solution is soaked up quickly in the pores of the support. It is important that the pores are just filled. When too little solvent is used, the precursor is spread only on part of the surface area of the support, and this may result in larger metal particles. When too much solvent is used, part of the precursor will end up on the rela­tively small external surface area of the support, which may result in a few but very large metal particles outside the pores. Because the pores have to be filled precisely, th is method is also known as the incipient wetness technique : when the pores are.just filled, on the verge of 'flowing over', the support starts to feel wet. After impregnating or incipiently wetting the support, the catalyst precur­sor is, as described above, dried carefully, calcined if necessary,. reduced in hydrogen and finally passivated.

Page 24: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 14

2.1.3 Jon Exchange

A technique more sophisticated than the pore volume impreg­nation method is the ion exchange technique. In this method, cations are fixed to the internal surface area of the support. An example is the ammonia ion NH 4 + which absorbs readily on sup­ports like titania, Ti02• An ammonium solution (NH 40H) is added to the support and allowed to adsorb. After the adsorption process, the support is filtered off. This support, saturated with ammonia, is added to a solution of a metal salt, e. g, an aqueous solution of Rh ( N 0 3h It is essential that the metal ions in the solution are present as cations . In the case of Rh(N0 3b, rhodium is present as IRh(OH)n(H 20)6.nJ( 3·n)+ complexes. These (positively charged) complexes exchange readily with the absorbed N H4 + ions on the support. In this way, an equal spread of the metal precursor over the support can be achieved. If RhCl3 were used as metal precursor, rhodium would be present as (H 30+ +) [RhCl3(0H)n(H 20) 3_nJn­complexes and these negatively charged complexes will not exchange with the N H4 + ions on the support.

In this thesis, we will encounter an example in which no specific counter ion has been adsorbed on the support to exchange with a metal complex. The 4 wt% Rh/Ti02 catalyst described in chapter 7 has been prepared by exchanging Rh(N03h with the pro­tons present in the surface hydroxyl groups in Ti0 2; the Rh/Ti02 catalyst described in chapter 8 has been prepared by ion exchange

. . using ammonia.

After the metal precursor has been exchanged, the support is filtered off and dried. In case ammonia has been used, a calcination step is usually applied to remove the ammonia left behind on the support. The dried or calcined catalyst is then reduced in hydrogen and finally passivated.

Page 25: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 15 Chapter 2

2.1.4 The Urea Method

The urea method is founded on a controlled raise of the pH of

a solution of the metal precursor and urea (CO(NH 2)i) in which the

support is suspended ( 1,2). The solution is stirred vigorously. An

adequate temperature is selected ( ± 365 K) at which the urea

decomposes slowly according to the following reaction :

As a result, the pH value of the solution slowly increases as the urea decomposes. At a certain pH value a metal hydroxide starts

to form and precipitate on the support. Because the urea decompo­

sition rate and thus the liberation of OW groups can be controlled

by regulating the temperature, the precipitation of the metal precur­

sor can be controlled elegantly. As a result, the method provides a

homogeneous spread of metal hydroxide over the support, which is

a good starting point for obtaining highly dispersed metal particles .

After drying, a calcination step is essential in order to remove the

urea left behind in the sample. Finally , calcination is followed by

reduction and passivation.

2.2 Temperature Programmed Reduction

Temperature programmed reactions are used frequently to

study the chemical behavior of supported metal catalysts. Most

important among them is Temperature Programmed Reduction

(TP R). Other examples are Temperature Programmed Desorption

(TPD), Oxidation (TPO) and Sulfiding (TPS). The principles are

the same for all temperature programmed reaction techniques. As

the temperature of the catalyst is increased, some reaction of the

Page 26: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 16

active phase with the gas atmosphere is studied. In TPR. the redu­

cibility of a sample , e. g, an oxidized metal catalyst or a catalyst

precursor, takes place. The catalyst is flushed with a mixture of

4% H2 in N2 and the temperature of the sample is increased at a

constant heating rate of 5 K min-1. The following reaction schemes

illustrate the reduction processes that may proceed in case of a pre­

cursor catalyst which has been prepared using RhCl 3, or in case of

an oxidized rhodium catalyst :

2 RhCl 3 + 3/2 H2 +::± 2 Rh + 3 HCI

Rh20 3 + 3 H2 +::± 2 Rh + 3 H20

By monitoring the consumption of hydrogen, one can get informa­

tion on the reduction process. The hydrogen uptake as a function

of temperature is usually denoted as the 'TPR profile ' of the sam­

ple . The temperatures at which the reduction proceeds reveal what

substance is being reduced and can be used as a 'fingerprint' . The

amount of hydrogen consumed provides the reaction stoichiometry

and/or the degree of reduction at a certain temperature during the

process.

Just as the other temperature programmed techniques, TP R is

used most often as a fingerprint technique. For detailed descrip­

tions, we refer to the review by Hurst (3) and a few of the earliest

papers on TPR (4-11 ). The apparatus we used has been described

extensively by Boer et al. ( 12).

Page 27: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 17 Chapter 2

2.3 Hydrogen Chemisorption

In supported metal catalysis , the amount of metal atoms exposed to the gas atmosphere is one of the most important characteristics of the catalyst. The dispersion of a catalyst is used to quantify this and is defined as the fraction of metal atoms in contact with the gas atmosphere . Adsorption and desorption tech­niques can be used to estimate the dispersion. These techniques use a selected gas that adsorbs only on the metal particles and not on the support. Dependent on the specific technique used, the amount of gas that adsorbs or desorbs is measured. The hydrogen chemisorption technique used to measure the dispersion of the catalysts discussed in this thesis is extensively described in ( 13,14).

Briefly, the experimental procedure comprised the following steps. A catalyst sample, dried , calcined, passivated or oxidized, is reduced

in situ in 100% H2 at the desired temperature. After the reduction procedure. the sample is evacuated at some elevated temperature, in general 473 K. The chemisorption cell contains two sections, of which the exact volume is known. The first compartment is used as a reference chamber, the second contains the (now reduced and evacuated) catalyst sample. A stop cock separates the two sec­tions . After reduction and evacuation , a known amount of hydro­gen is admitted into the reference chamber. After the valve between the reference and catalyst section is openeq . hydrogen starts to adsorb on the catalyst . In most cases, adsorption is an activated and therefore slow process. To circumvent this, the catalyst is temporarily heated , usually to 473 K, to speed up the adsorption process. After cooling down and equilibrating, the amount of gaseous hydrogen can be computed by measuring the

pressure and the amount of adsorbed hydrogen can be calculated. The experiment is then continued in the desorption mode. The stop cock between the two chambers is closed and the hydrogen pres­sure in the reference chamber is lowered. After opening the stop cock, hydrogen desorbs from the catalyst because of the lower pres­sure, and the amount can again be computed by measuring the

Page 28: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 18

equilibrium pressure. Thus. the ratio of the amount of adsorbed hydrogen and the amount of metal present. the H/M value. is moni­tored as a function of pressure. The linear part of this desorption isotherm is extrapolated to zero pressure in order to nullify small errors in the volume of the catalyst section and to eliminate adsorp­tion of hydrogen on the support.

As indicated in ( /4), this H/M value is not a direct measure for the dispersion. but can be used to compare catalysts and is therefore very useful as a fingerprint. However. since in the same paper the hydrogen chemisorption method has been calibrated with an independent technique. EXAFS. we can estimate very accurately. dispersion and particle sizes from the H/M values .

2.4 Electron Spin Resonance Spectroscopy

In case a catalyst contains paramagnetic centers. Electron Spin Resonance is a useful technique to study these centers. For an iso­lated electron. a paramagnetic center. two states are accessible, one with spin +1/z (O'), the other with spin -1/z (y) . In the absence of a magnetic field these two states are degenerated : E( O') = E( y). A magnetic field removes this degeneracy. the two states differ in energy by

!iE = £(0') - E(y) - gel3H 12.4.1]

in which 13 is the Bohr rnagneton and H is the magnitude of the magnetic field . The electron g-factor, ge, for an isolated electron in a spin-only case. is equal to 2.0023. Transitions between the two levels O' and 13 can be generated by a suitable electromagnetic radia­tion. In practice. the frequency of the electromagnetic radiation is kept constant while the magnetic field is varied linearly with time.

Page 29: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 19 Chapter 2

When the photon energy h ii equals the difference in energy of the

two states, transitions between ex and {3 may occur. This 1s

accompanied by absorption of the electromagnetic radiation.

In practice, electrons are never isolated and we have to expand

the theory to perturbed unpaired electrons. In case of a free elec­

tron, the electron spin is solely responsible for the electrons mag­

netic moment and the Hamiltonian H can be written as

H gf3(H·S) [2.4.2]

In relevant cases, however, the electron 'moves' in an orbit 'around' a nucleus and this gives rise to an orbital angular momentum cou­

pling (f3 (H ·L )) and a spin-orbit coupling (.\ (L ·S)). Conse­

quently, the Hamiltonian can be written as

H = g{3(H·S) + f3(H·L) + .\(L·S) [2.4.3]

The complication induced by the latter two phenomena can be cir­

cumvented be defining an effective spin Hamiltonian Herr which

operates only on fictive spin states :

Herr = f3 (H ·g err"S:) [2.4.4]

The effects therefore manifest themselves in the value of the

effective g-tensor g err· Measuring the effective g-values therefore

provides information on the magnitude of the orbital angular

momentum and the spin-orbit coupling. Two important properties

of the g-tensor need to be mentioned. The electron spin is coupled

to its orbital momentum and therefore to the lattice in which the

atom or ion is situated. Hence, relaxation phenomena are related to the effective g-value. Enhanced relaxations may be accompanied by

large deviations of g err from 2.0023. The second property to

Page 30: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 20

mention 1s the symmetry of the g-factor. A crystal field with spherical symmetry gives rise to an isotropic g-value. For an axial field along the z-axis. two g-values, gxx = gYY and g

11• may be

observed.

In Chapter 4, an ESR study of Fe3+ ions will be described. Fe3+ has a 3d5 configuration and in a high spin case in an octahedral crystal field. Fe3+ has five unpaired electrons. Whenever the site symmetry deviates slightly from perfect octahedral. which is usually the case, this gives rise to a very characteristic ESR signal centered at g = 4.2 ( 15-17).

2.5 Nuclear Magnetic Resonance Spectroscopy

For Nuclear Magnetic Resonance, the basic principles are the same as those for ESR. While in ESR transitions between electron spin states are observed, in NM R transitions between nuclear spin states are studied. For nuclei with I= +1/z in a magnetic field, again two states are accessible : one with the nucleus' magnetic moment parallel and the other anti-parallel to the external magnetic field. When irradiated with a suitable electromagnetic radiation, transitions between those states can be generated and an absorp­tion of the electromagnetic radiation can be observed. As . for ESR, the frequency of the electromagnetic radiation is kept constant and the absorbance is monitored as a function of the external magnetic field. The magnetic field which the nucleus experiences is in general not equal to the applied magnetic field : the electrons surrounding the nucleus modify the external magnetic field. Thus, the shift in NMR spectra gives information on the chemical environment of the nucleus under study and is therefore called the chemical shift. For example, the chemical shift for Pt in H2PtCl6 is different from the chemical shift of Pt in Na2Pt(OH) 6 ( 18,19) and can therefore, although both salts are present in the same sample, be used to

Page 31: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 21 Chapter 2

study these salts separately. In chapter 3 we will encounter an

example in which 195 Pt NM R is used to estimate the amount of Pt

present in H2PtCl6 crystallites in an impregnated and dried catalyst

sample.

2.6 Mossbauer Spectroscopy

In 1957, Rudolf L. Mossbauer demonstrated that nuclei can

resonantly absorb gamma rays which originate from similar nuclei

decaying from excited states ( 20-22). The basic principles are

explained in Figure 2.1a. In this example, 57Co is used as a source

and decays according to the scheme in Figure 2.1. The 14.4 keV

emission can be used to generate transitions in 57 Fe nuclei in the

sample between I = 1h and I = %. In practice, the energy of the

emitted gamma quanta differs slightly from the energy needed to

excite the 57 Fe nucleus under study. To circumvent this, the source

is give a velocity and because of the Doppler effect, the emitted

energy is modulated slightly. The absorption therefore, is measured

as a function of the Doppler velocity of the source and peak posi­tions and shifts are reported in mm s-1

. This is ii lust rated In

Figure 2.1b. There are three hyperfine interactions which are respon­

sible for the fact that the 57 Fe nuclei in the sample absorbs quanta

of a different energy than those emitted by the excited 57 Fe nuclei

in the source . The first is the isomer shift (J.S.) which is a meas­

ure for the electron density around the nuclei under study . The iso­

mer shift is caused by the Coulomb interaction between the posi­

tively charged nucleus and the negatively charged s-electrons,

whose wave functions overlap with the nucleus. The isomer shift

gives information on the oxidation state of the iron . The second

hyperfine interaction is the quadrupole splitting. In its ground state,

the 57 Fe nucleus has a spherical charge distribution and therefore no quadrupole moment. In the excited state, however, the nucleus has

an ellipsoidally shaped charge distribution and therefore has a

Page 32: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 22

Figure 2.1 Mossbauer Spectroscopy

(a) Basic Principles

(b) Experimental set up

(c) Three basic Mossbauer spectra : (1) no hyperfine interactions (the spectrum of stainless steel), (2) the influence of quadru­pole splitting (the spectrum of sodium nirtoprusside) and (3) magnetic hyperfine splitting (the spectrum af a-Fe) .

57Co

~lectron b CJ \apture

.. source sample detector

9% 91%

137 keV 123 keV

- - -'.----- I = 3/ 2

--~-I=1/2 (14.4 keV---

____L _ _ __,__

Doppler velocity I I I I I I I I I I I I I I I I I I I I I

-10 -5 0 5 10 c

~ (1)

w (21

(3)

Page 33: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 23 Chapter 2

positive electric quadrupole moment . In case the nucleus experi­ences an electric field gradient , two orientations are allowed for the quadrupole moment and a splitting of the excited level is observed (see Figure 2.1c) . The splitting between the two excited levels t::i..£ 0

is proportional to the electric field gradient at the nucleus ; The third interaction is a magnetic hyperfine splitting. The magnetic moment of the excited nucleus will react on any external magnetic field, including the magnetic field induced by the surrounding electrons . Two orientations are accessible for the ground state and four for the excited level. Thus, a splitting of the ground level and the excited level is observed (see Figure 2.1c) . Therefore , for a nucleus with a magnetic moment , eight transitions are possible , two of which are forbidden, leaving six transitions.

Thus , isomer shift , quadrupole splitting and magnetic hyperfine splitting are used to identify the environment of 57Fe nuclei in a sample . Mossbauer spectroscopy can of course be used for other elements as well. Examples are Ir and Ru. In chapter 4 we will discuss an example in which the state of iron in Si02 supported bimetallic Fe-Ru and Fe-Pt catalysts is studied using Mossbauer spectroscopy.

2. 7 Laser Raman Spectroscopy

In Raman spectroscopy , the em1ss1on spectra of excited molecules are studied. A laser is used as a light source. The electric field E of light will give rise to a redistribution of the charge in a

polarizable molecule and thus induce a dipole, the dipole moment 71 being equal to

[2.7 .1]

Page 34: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 24

in which O' is the polarizability of the molecule. For light with fre­quency v 0• the dipole moment is equal to

µ - 0'£0sin(21Tv0 t) [2.7.2)

When during the absorption process the excited molecule decays to a state different from the ground state, a state with an internal vibration with frequency vi, the polarizabil ity O' will oscillate accordingly :

[2 .7.3)

and therefore

µ [O'o + 13 sin(217' vi t ) )Eosin(217' v0 t ) [2.7.4)

- O'oEosin(21Tvo t)

+ 1hf3E0 ,cos(21T(vo-vi)t) - cos(21T(v0+vi)t)]

The excited molecule will therefore emit radiation with frequencies equal to v0 (Rayleigh scattering), v0 - vi and v0 + vi, the Stokes and Anti-Stokes lines respectively. The frequency shifts in Raman therefore correspond directly to the frequency of the induced vibra­tion. As in infrared (IR) spectroscopy, these frequencies are very specific for the molecule under study. In chapter 3 we will discuss the Raman spectra of the PtCl6

2- unit. The basic unit of this

molecule is an octahedron and three of its vibrations are Raman active. Figure 2.2 shows the PtCl6

2- unit. For centro-symmetric

systems, the vibrations that are inactive in Raman spectroscopy, are active in IR and vice versa. However , IR turned out to be inca­pable of_ detecting any vibrations in the PtCl6

2- unit of impregnated

and dried Pt/ A1 20 3 catalysts. Details of the Raman spectroscopy experiments wil I be given in chapter 3.

Page 35: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 25 Chapter 2

Figure 2.2 The structure of H2PtCl 6

Cl Cl I / Cl I / Cl

Cl-Pt-Cl ········ ·· Cl--Pt-Cl /I /I Cl .Cl Cl Cl

I / er I / er c1 Cl-Pt-Cl ···········C l--Pt - -Cl

/ I. / I Cl Cl Cl Cl I / Cl I / Cl

- Pt-· · C 1--Pt---C 1 / I ,, I Cl Cl Cl ,,

I/Cl I/Cl Cl C 1----Pt---C 1 · ··· C 1- Pt-C l

c1/ I c1/ I Cl Cl

2.8 ASED-MO Computations

ASED is an acronym for atom superposition electron delocali­zation. The ASED-Molecular Orbital theory (23) has been applied in

numerous and diverse studies to predict structures, reaction mechanisms and vibrational and electronic properties . The theory

uses for input data the ionization potentials and valence state Slater orbital exponents for the constituent atoms (24-28) . These parame­

ters are sometimes altered, particularly in treatments of ionic hetero nuclear molecules, to ensure reasonably accurate calculations. The

electronic charge density function of a molecule or solid can be par­titioned into components in any number of arbitrary ways. The

ASED-MO theory is based on a partitioning of the density function

Page 36: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 26

into free atomic components and the rest. The atomic components

are spherically symmetric in a field-free space and are centered on

the nuclei . They follow the nuclei 'perfectly' . The remainder

changes its shape depending on the geometry. With respect to the nuclei. it is 'non-perfectly following' . Figure 2.3a shows an example

of perfectly following atomic densities Pa and Pb for atoms a and b

in a diatomic molecule and a nonperfectly following charge density

Pnpf . The interaction energy is made up of a repulsive part which

can be computed from the perfectly following charge densities and an attractive part, which can be calculated from the nonperfectly following charge density. In Figure 2.3b both repulsive and attrac­

tive energy terms and their sum, the interaction energy are shown schematically . The first assumption in the ASED method is, that the atoms are first instantaneously superimposed. Based on this, the repulsive energy term follows from

[2.8.1]

in which Z is the nuclear charge, p the atomic charge density func­tion, R the coordinate of the nuclei and r the coordinate of the elec­

trons. E R is repulsive because nuclear repulsion energy is greater

than the attractive energy between nucleus b and Pa. The non­perfectly following energy term is attractive because of the concen­

tration of charge in the internuclear region due to bonding. The attractive energy term is given by

z Rfb r (. R . ) d J i - - b oo J Pnpf

1 • b dR; I R;- r I drdR; [2.8.2]

It is impossible to evaluate equation [2.8.2] . However , the non­perfectly following or attractive energy term is due to electron delo­

calization and is roughly equal to the difference in atomic and

Page 37: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 27 Chapter 2

Figure 2.3 ASED-MO terms for a diatomic molecule .

(a) The perfectly and non perfectly following charge densities for a diatomic molecule AB

(b) The attractive energy term. the repulsion term and the in­teraction energy as a function of interatomic distance.

a b

/

/

R a-b

Etot E

npf

ER

molecular orbital energies. Thus, E att is successfully approximated by

!).£mo

'° n Eab £., I I

[2.8.3)

which is a summation over the molecular orbitals i; n; is the orbital occupation number (0, 1 or 2), Eia and Eib are the atomic orbital

energies (in practice, VSIP) and Eiab the molecular orbital energies. The molecular orbital energies are the solutions of the (extended Hi.ickel) Hamiltonian with

Page 38: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 28

HC1a - -(\ISJP)r [2.8.4) ll

Ha.a - 0 [2.8.5) lj

H/2j 1125(Hcza + Hbb) 5 ab c-0.13R • ll J J lj [2.8.6)

in which S;'1/' is the overlap integral of a and b. R the internuclear

distance. Finally. the total energy is the sum of ER and E att :

Etot - L, ER(a .b) + Eau a > b

[2.8.7]

For a detailed discussion, we refer to ( 23 ,29 ). In chapter 6 we will describe the results of this kind of calculation for 10 atom rho­

dium metal clusters supported on y-Al 20 3. From these calcula­

tions, we will be able to derive information about the binding of

rhodium metal particles to the alumina support, the binding energy

and the structure of the metal- support interface.

2.9 Extended X-ray Absorption Fine Structure

2.9.1 Basic Principles

EXAFS is an acronym for Extended X-ray Absorption Fine

Structure. It refers to the 'wiggles', the fine structure that can be

observed in the X-ray absorption spectra of condensed phases at

the high energy side of absorption edges. Let us focus on a 1s elec­

tron which is subjected to monochromatic X-ray radiation. When

the photon energy tiw is lower than the binding Eb energy of the 1s

electron, absorption will not take place. When the photon energy

equals the binding energy. the 1s electron may be excited and this

Page 39: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 29 Chapter 2

gives rise to a sharp increase in the x-ray absorption. The probabil­

ity of exciting the electron is given by

[2.9.1]

in which l/Ji and l/Ji are the electron initial and final state. When we

assume that the photon has its electric field polarization in the z­direction, a simple deduction ( 30) leads to the formula for the

attenuation coefficient µ. :

µ. - [2.9.2]

in which w is the frequency of the photon, c the speed of light, Na

Avogadro's number and p(E t) the density of final states.

At photon energies above the electron binding energy, the tran­sition probability and therefore the attenuation coefficient slowly

decrease with increasing photon energy. A typical example is the case of a monoatomic gas such as krypton. Figure 2.4a gives an

ideal absorption spectrum of such an unperturbed transition .

In case the excited atom is surrounded by neighboring atoms, the absorption changes . The neighboring atoms have a pronounced influence on the final state and therefore on the attenuation

coefficient. Because of the neighboring atoms, some final states are favored and others are disfavored with respect to the unperturbed case where no nearest neighbors are present. The result of this is illustrated in Figure 2.4b : at the high energy side of the adsorption

edge of rhodium foil the attenuation coefficient µ. shows an oscilla­tory behavior which damps out slowly .

Page 40: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 30

Figure · 2.4 The influence of neighboring atoms on the X-Ray absorption spectrum

(a) An unperturbed adsorption spectrum

(b) The X-Ray absorption spectrum of the K-edge of rhodium foil

1

E (e V) 24000

E (e V) 25000

With scattering theories, we can describe the final state in an understandable way and derive a mathematical expression for the fine structure in the absorption. In our model, the outgoing part of

the wave function l/Jf will scatter from the neighboring atoms. The scattered state will interfere with the outgoing state and depending on the relative phases of both states, the final state, being the sum of outgoing and scattered state, is amplified in case both states are in phase and will be attenuated in case both states are out of phase. The electron wavelength in the outgoing and scattered state is given by

h (2.9.3)

Both outgoing and scattered state are in phase when

Page 41: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 31

R n A. 2

Chapter 2

[2 .9.4]

in which R is the interatomic distance between absorbing and

scattering atom. Thus, when the photon energy equals

E - hw [2.9.5)

both outgoing and scattered state will be in phase. the final state

will be augmented and absorption will increase relative to the unper­

turbed state. When we convert the energy scale to a wave vector (k) scale using

k 2.J'i.TTm ,/tiw-£. h I

(2.9.6]

we will find maxima in the attenuation coefficient at

k n 1T

R (2.9.7}

The maxima, and therefore the minima and the nodes in µ will occur at constant k-intervals of fj,k = TT/ R. Therefore, EXAFS

spectra are always represented in k-space rather than in energy space.

We will not further explore the physics of the EXAFS phenomenon. In the above discussion the basic principles are

explained sufficiently to give a good insight into the spectra which

will be discussed in the next chapters. A rigorous derivation of a

more complete EXAFS formula includes phase shift functions, back­

scattering amplitude, the lifetime of the final state, the mean free path of the electron, disorder and finally, corrections can be made

Page 42: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 32

for approximations induced by the small atom method used and for

multiple scattering phenomena ( 30). Such a derivation is clearly

out of place in this thesis and we will therefore conclude this intro­

duction by discussing a general formula for the EXAFS function

x (k) :

x(k) = L,Aj(k)sin(2kRj + ct>j(k)) J

(2.9.8]

The amplitude function Aj (k) is given by :

Clearly, X(k) is a sine function, modulated by an amplitude func­

tion. In the argument of the sine function we find the interatomic distance R . between absorbing atom and scattering atom j and a

J phase shift function ct> j (k ). The amplitude function Aj (k) is

governed by the number of neighbors in shell j (Nj ), their distance

with respect to the absorbing atom (Rj ), a backscattering function

Fj (k), SJ which corrects for relaxations in the absorbing atom

and multi-electron excitations, and two exponential terms which account for disorder (the term with o 2) and for the main free path

of the electron (>-e ).

The backscattering amplitude F (k) indicates 'how well' the neighboring atom performs in scattering the excited electron. F (k ) is therefore element specific for the scattering atom. In general, heavier atoms scatter more and therefore have a higher backscatter­

ing amplitude. The backscattering amplitude is also a function of the wave vector k and F (k ) for high-Z elements may have a very . characteristic k-dependence . In Figure 2.5a, F (k) for 0, Ti, Rh and Ta are given. The data have been taken form (]J ).

Page 43: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 33 Chapter 2

Figure 2.5 The k-dependence of F (k ) and cf> (k )

(a) The backscattering amplitude for 0. Ti . Rh and Ta

(b) The backscattering phase shift function for 0. Ti. Rh and Ta

- Ta ··· ··· Rh ---- Ti -0

a b 12

. I' 9 ~ :, \

. . ... . .. . . ·.· . . 6

0.5 -----3

0 ------------------------ ---3

5 10 15 5 10 15

k r,J-11 k [ ,J -11

Just as F (k ) , the phase shift function cf> (k ) is element

specific. However, in contrast to F (k ) . cf> (k ) is a function of both absorbing and scattering atom. The contributions of absorbing and

scattering atom to the total phase shift are independent <31-33) :

[2.9.10)

cf>~ is the contribution of the absorbing atom A and cb~ the contri­

bution of the scattering atom B to the total phase shift function :

The constant 8 is equal to unity for K and L1 edges and is zero for L11 and L111 edges. In Figure 2.5b, the phase shift functions for the

absorber-scatterer pairs Rh-0, Rh-Ti, Rh-Rh and Rh- Ta are given.

As we will see in the next sections, F (k) and c/>(k) perform a unique role in identifying the neighboring atom . And by determin­ing the interatomic radii and the coordination numbers, EXAFS 1s

clearly a very powerful technique in analyzing local structures.

Page 44: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 34

2.9.2 Fourier Transformations

In general, more than one shell of neighbors is present around

an atom and this makes the analysis of the EXAFS spectrum more

complicated . Therefore, Fourier transforms are used frequently to

separate the contributions of these neighboring shells of atoms . The mathematical formulation for a Fourier transformation is

[2.9.11]

A Fourier transform is no more than a mathematical tool

which 'sorts' the information in a frequency spectrum in a more

convenient way. The result of the Fourier transform given above is

a radial distribution function en (r), a function that has maxima at

radii that are related to the coordination distances R; in the original

function. In Figure 2.6a the EXAFS function of rhodium foil and its

Fourier transform are shown. In the Fourier transform, the first and

higher order shells are clearly visible. In case a peak in a Fourier

transform is clearly separated from neighboring contributions (like the first and main peak in the Fourier transform in Figure 2.6b), an

inverse Fourier transform can be used to extract a single shell EXAFS function. In case two or more peaks are included in the

back-transformation range, a two or more shell EXAFS function 1s

obtained. An inverse Fourier transform is given by

rmax

x· (k) - 1 J e (r )e - 2ikr dr kn .J2.7r . . n

r1nm

[2.9.12)

In Figure 2.6c, the result of an inverse Fourier transform over an ,._ range from 1.2 to 3.3 A, i.e. the first shell, is given.

Page 45: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 35 Chapter 2

Figure 2.6 EXAFS functions and Fourier transforms

(a) The EXAFS function of rhodium foil

(b) k1-weighted Fourier transform of X (k ) of rhodium foil

(c) The inverse Fourier transform of the k1-weighted Fourier transform of X (k ) of rhodium foil

(d) k1-weighted Fourier transform of X (k) of rhodium foil, corrected for phase and amplitude of Rh.

*10-2 *10- 1

12 5 ~~-~~~--,.--~---,

a b B

4 0

0

-4

-B -5 0 5 10 15 20 25 0 2 6

* 10-2

B 4

c d 4

0 0

-4

-B - 4 _,________.,_---+-~~-+--~---< 0 6 12 18 24 0 2 4 6

k r A- 1 J R [A]

Backscattering amplitude and phase shift functions can com­

plicate Fourier transforms. In Figure 2.6b, a k 1-weighted Fourier

transform of the EXAFS function of rhodium foil is given. The first peak is clearly asymmetric, for the magnitude as well for the ima­

ginary part of the Fourier transforms. This is caused by a k­

dependence in both functions. Usually, the phase shift function has

Page 46: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 36

a linear k-dependence like

<t> (k ) - <t>o + <J>k [2.9.13]

The argument of the sine in the function can then be written as

2kR + <i>(k) - 2kR + <i>o + <i>k

- 2k (R + ~ ) + <i>o

[2.9.14]

and consequently, the Fourier transform 'peaks' at a distance of

R + <t>/2. However, when the phase shift function is known, we

can correct for this by multiplying x (k) with e- i ¢(k I prior to the

Fourier transformation. The effect of the backscattering amplitude is only of importance when F (k) shows one or more extrema .

That is the case for high-Z scatterers . In that case, the EXAFS function is 'modulated' by a more or less periodic function, and the

corresponding peak in the Fourier tran~form will be accompanied by

one or more sidelobes. When F (k) is known , we can circumvent this by dividing x (k) by F (k) prior to the Fourier transformation . Thus, a phase and backscattering amplitude corrected Fourier

transform is given by the following equation :

e~(r) k "'°' -i¢(k)

_1_Jx(k)e kne2ikrdr J21T k . F (k)

mm

[2.9.15]

In Figure 2.6d, the result of a phase and backscattering amplitude corrected and k 1-weighted Fourier transform of the EXAFS function of rhodium foil is given. Clearly, the first peak is now centered at

the correct Rh-Rh distance (2.687 A) and has no sidelobe any more . Corrected Fourier transforms look less complicated than uncorrected

transforms and are therefore used whenever possible in this thesis .

Page 47: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 37 Chapter 2

2.9.3 Reference Compounds and Calculating Spectra

The mainstay of analyzing experimental data is to calculate an EXAFS functions that resembles the measured function as accurate as possible. In addition to the parameters N, R, a 2, we need to know F (k ) and <t> (k ) to calculate an EXAF S spectrum. Both these function can be extracted from the EXAFS function of a suit­able reference compound. For example, for a Rh-Rh EXAFS func­tion, we extract FRh(k) and <t>Rh-Rh(k) from the EXAFS spectrum of rhodium foil and for a Rh-0 EXAFS function, we extract F 0 (k) and <t>Rh-o{k) from the EXAFS spectrum of Rh 20 3. The procedure is in fact very simple but very sensitive towards the choice of Fourier transform ranges and should therefore be carried out with great care. Assuming we have correctly extracted x (k ) from the raw data (see the next paragraph), we perform a Fourier transform over the largest possible range in k-space, selecting krnin and krnax in nodes of the EXAFS function to avoid cut-off effects. An inverse Fourier transform containing only but completely the desired shell yields the EXAFS function from which the amplitude function Aj (k) and the argument of the sine function (2kR + <t>(k )) can be extracted. Using crystallographic data (N and R) we can derive the phase shift function and a normalized amplitude function

F. (k) A (k) kR 2 N

[2.9.16]

This amplitude function has the same general behavior as the backscattering amplitude, but in addition contains the disorder term, the main free path term and SJ (k). We will assume that the latter two terms will be the same for the samples under study and the reference compound. We will account for disorder by introduc­

ing the term e- 2k

2t:.<Y

2 in which tia 2 is the Debye-Waller factor, the

Page 48: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 38

disorder in the sample relative to the disorder in the reference com­

pound. Thus, an EXAFS function can be calculated using

x (k)

in which F • (k ) and <b (k ) are the experimentally determined back­scattering amplitude and phase shift functions.

2.9.4 Data Analysis

Analyzing EXAFS spectra is a very delicate procedure. An excellent review has been published by Sayers in ( 34). The reliabil­

ity of the resulting structural parameters depends up to a high

degree on the accuracy with which data reduction and data analysis have been carried out. There are many pitfalls in the data reduction

and data analysis procedure. For example, subtracting a back­ground which has not been optimized thoroughly will result in incorrect coordination numbers and Debye Waller factors and possi­

bly in wrong coordination distances. Even a small glitch or a jump

in a spectrum will induce effects in a Fourier transform and thus, may give rise to artificial peaks in the radial distribution function.

Before the actual procedure of analyzing the data, a data

reduction procedure is needed to extract the EXAFS function from

the experimentally determined X-ray absorption spectrum. Depen­dent on the software which controlled the experiments at the sta­

tion where the measurements have been preformed, the monochro­mator position may have to be converted to an energy scale and the

read out of the two ion chambers may have to be converted to absorbance. In Figure 2.7, as an example, the X-ray absorption

spectrum is given for the 4 wt% Rh/Ti0 2 sample to be discussed in

Page 49: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 39 Chapter 2

Figure 2.7 The Victoreen curve (dashed line) and the background (dot­

ted line) used to extract X (k) from the experimental data

15

5

' '

1. 0

0.8

0 ...._..__..__,__L.......JL.......J--L___J.--'---1

23000 24000

Energy (e V) -0 500 1000

E - E0

chapter 7. From this spectrum a Victoreen curve is subtracted

(the dashed line in Figure 2.7). After the Victoreen curve has been subtracted, the inflection point of the edge is chosen to represent the actual edge position. After a Victoreen curve has been sub­

tracted, the gradient in the data is small. Therefore, at this stage any glitches or jumps have to be removed carefully from the spec­trum. During the next step, a smooth background is constructed to fit loosely the data at the high energy side of the edge. This back­ground is supposed to represent the unperturbed absorption spec­trum, i. e., the spectrum in case the absorbing atom is not sur­

rounded by neighboring atoms. The dotted line in Figure 2.7 represents the background. This background is subtracted from the

experimental data to give a preliminary EXAFS function. The back­ground is constructed using a cubic spline routine. The parameters

are chosen such that the background does not contain any EXAFS

oscillations and that the resulting preliminary EXAFS function con­tains 'as little background' as possible (34). A Fourier transform of the resulting spectrum and the derivative of the background are

used to optimize the cubic spline parameters. Slow background oscillations present in the EXAFS function manifest themselves in a Fourier transform as peaks below 1 or 1.5 A. In case EXAFS oscil­lations are present in the background, the main peak(s) in the

Page 50: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 40

Fourier transforms will have decreased in intensity and these oscil­

lations can be observed in the derivate of the background . Clearly.

in the ideal case the contribution in the Fourier transform below 1 A. is negligible and the major peaks still have their maximum attain­able intensity .

After the background is subtracted. the preliminary EXAFS

function is normalized with respect to the height of the edge and

has therefore a per-atom dimension . This allows a quantitative

analysis of the spectrum. The data reduction procedure is finally

concluded by converting the energy scale to a wave vector (k) scale

using equation (2.9.6) and taking Ei equal to the inflection point.

Since the inflection point is not a correct measure for the actual edge position but the only objective alternative at hand, a small

correction on the edge position will have to be included in calculat­

ing EXAFS spectra. This correction 11£0 thus comprises the error

in the edge position of the reference compound used to calculate the

spectrum and the error in the edge position of the sample under

study .

The data analysis procedure consists of calculating EXAFS

spectra that resemble the measured spectra as accurately as possi­ble . The mainstay of this process is the comparison of the calcu­

lated and experimentally determined spectra. In this respect,

Fourier transforms are preferred above the data in k-space, espe­cially when more shells . are present. In addition, corrected Fourier

transforms give a good insight in the effect of changing the parame­ters used to calculate the spectra. For example, a wrong coordina­

tion distance will give a peak at the wrong position and an incorrect

value of 11£0 will result in a different symmetry in the imaginary

part of the Fourier transforms, provided the correct reference has

been used to correct the Fourier transform . When the coordination

number is too low or the Debye Waller factor (!io 2) is too high , the

resulting peak in the Fourier transform is lower than the

corresponding peak in the experimental data . Here we encounter a

first problem : N and !io 2 seem to be interrelated. at least in a

Page 51: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 41 Chapter 2

Fourier transform . There is an infinite number of combinations of

N and ~a 2 that give the same peak height in the Fourier

transform. However, N and ~a 2 both have a different k­

dependence : a higher coordination number will increase the ampli­

tude of the EXAFS function over the whole k-range, a lower Deb ye

Waller factor will result in an increase in amplitude especially at

higher k-values, in other words, the damping of the EXAFS f unc­

tion is less for a lower ~a 2. There is of course only one combina­

tion of N and ~a 2 that will fit the experimental data best and this

combination can be found as follows. There are two restrictions.

One is that the contribution to be optimized should be separated

reasonably from the other contributions in the spectrum and the

second is that R and ~Eo are chosen correctly. We start by deter­

mining N as a function of ~a 2 using a k 1-weighted Fourier

transform and assuring that the peak height of the calculated and

measured spectrum are the same over the whole ~a 2 range. This is

done easiest by taking a few fixed values for ~a 2 and determining

the corresponding coordination number. For a fixed value of ~a 2,

N is directly proportional to the peak height and an incorrect choice

of N can be corrected easily. (For a fixed value of N, ~a 2 is not

directly proportional to the magnitude of the peak in the Fourier

transform, see equations [2 .9 .9) and (2.9.17]) We repeat this pro­

cedure, now using a k 3-weighted Fourier transform. Since in a k 1-

and k3-weighted Fourier transform the data are weighted differently

with respect to k and because N and ~a 2 have a different k­

dependence, we will find two curves for N as a function of !J.a 2.

The intersection point of the two curves represents the combination

of N and ~a 2 that will result in the best fit over the whole range in k-space.

When there is one dominant contribution and one or more

smaller contributions in the spectrum, we can use the difference file

technique. We calculate and optimize a one-shell EXAFS spectrum

that matches the main peak in the Fourier transform of the experi­

mental data best. We then subtract this calculated spectrum from the experimental data and analyze the difference spectrum, possibly

Page 52: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 42

using once more the difference file technique. In this way , all con­

tributions are analyzed. This technique has been introduced first by van Zon ezal. ( J S).

We can expand this technique to a recurrent optimization pro­

cess. We subtract the calculated EXAFS function(s) that. matches

the difference spectrum (spectra) from the experimental data and

use this new difference file to optimize the major contribution in the

experimental data. We thus initiate a recurrent process that, when

carried out correctly , will converge to a set of parameters that yields

an EXAFS function that accurately resembles the measured spec­

trum. However, care should be taken that the final set of parame­

ters does not represent a ' local minimum '.

Using these procedures, the EXAFS spectra in the next

chapters have been analyzed. The details of the analysis procedure

depend highly on the system under study and thus , for details we

refer to the experimental part of the chapters 5, 7, 8 and 9.

Figure 2.8 The experimental set up at station 9 .2 at the SRS in Dares­

bury .

Mo no-chr omat or

Wiggl er X-rays

~ ·~ ~ Samp l e

Stor age e

~ D D D ring

I C1 IC2

Page 53: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 43 Chapter 2

2.9.5 Experimental Metfwd

A reliable analysis can only be done on high quality data and

high quality data can only be collected when the experimental set

up during the measurement has been carried out with utmost care.

The set-up as used on station 9.2 at the Synchrotron Radiation Source (SRS) in Daresbury, where most of the spectra discussed in

the next chapters have been measured , is shown schematically in Figure 2.8 . The synchrotron was operated at 1.8 or 2.0 GeV and

the ring current ranged from 100 to 300 mA. From the wiggler magnet (or from a bending magnet), a primary X-ray beam enters

the monochromator chamber . IP contains a whole spectrum of wavelengths. The beam leaving the monochromator, I 0, is highly

monochromatic : the wavelength obeys the Bragg-relation :

2d sin{ 8) (2.9 .18]

in which /... is the wavelength. d the d-spacing in the Crystals of the monochromator and e the angle between the incident beam and the crystals . The photon energy is in general given by

E = 12398.52 /...

[2.9.19)

Thus. taking into account only the first order reflection (n = 1). we find for the photon energy in I 0 :

E - 12398.52 2dsin(8)

[2 .9 .20]

{When d is in A. E is in eV). The crystals used at station 9.2 were two Si [220] crystals with a d-spacing of 1.916 A. A measure for the absorbance of the sample is given by ln(J 0/1). This is however

Page 54: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 44

not an absolute measure for the attenuation coefficient µ.. but is

directly proportional to it . We will therefore use In (I 0/ I) as a

measure for µ. .

An optimum signal-to-noise ratio in µ. can only be achieved by

selecting optimum absorbances for the two ion chambers 11 and 12

and for the sample. The gas filling of the first ion chamber is such

that it absorbs 20% of the incoming beam I 0, the second ion

chamber absorbs 80% of the photons leaving the sample. The

thickness of the sample is chosen such that the absorbance (µ.) at

the edge is 2.5, i. e ., the sample absorbs about 8% of the photons

leaving the first ion chamber , 11. We refer to (36) for more details .

So far the description of the experimental set-up. The actual

data collection is now rather straightforward . The catalyst is

pressed into a thin , self supporting wafer with an absorbance of 2.5

at the edge and mounted in an in situ cell. In this · cell, in situ treatments can be carried out prior to the measurement. The pre­

treatments of the different samples are described in the relevant

chapters. After treating the sample in the in situ cell. the cell is

placed in the beam between the two ion chamber and after a correct

alignment. the absorption spectrum can be measured . Usually . the

spectrum is recorded while the sample is cooled with liquid nitrogen

to about 100 K. The reference compounds are measured in exactly

the same way as the catalyst samples.

2 .10 References

1. Hermans, L. A. M._; Geus , J . W. "PreJX1.ration of Catalysts II" : Del­mon, B.; Grange, P.; Jacobs , P. A., Eds .; Elsevier , Amsterdam 1983, p. 113

2. Geus, J . W "Prepa.ratiOn of Catalysts III" : Poncelet , G.; Grange , P.; Jacobs , P . A., Eds .; Elsevier, Amsterdam 1983, p. 1

Page 55: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 45 Chapter 2

3. Hurst, N. W.; Gentry, S. J.; Jones, A. Catal. Rev., Sci. Eng. 1982, 24, 233

4. Robertson, S D.; McNicol, B. D.; de Baas, J. H.; Kloet, S. C.; Jen­

kins, J. W J. Catal. 1975, 37, 424.

5. Jenkins, J. W.; McNicol, B. D.; Robertson, S. D. Chem. Tech 1975, 7, 316.

6. Wagstaff, N.; Prins, R. J. Catal. 1979, 59, 435.

7. Wagstaff, N.; Prins, R. J. Catal. 1979, 59, 445.

8. Vis, J. C.; van 't Blik, H. F. J.; Huizinga, T.; van Grondelle, J.; Prins, R. J. Catal. 1985, 95, 333

9. van "t Blik, H. F. J.; Prins, R. J. Catal. 1986, 97, 188

10. Martens, J. H. A.; van 't Blik, H. F. J.; Prins, R. J. Catal. 1986, 97,

200

11. van 't Blik, H. F. J.; Koningsberger, D. C.; Prins, R. J. Catal. 1986, 97, 210

12. Boer, H.; Boersma, W. J.; Wagstaff, N. Rev. Sci. Inst. 1982, 53, 349

13. Kip, B. J.; van Grondelle, J.; Martens, J. H. A.; Prins, R. Appl. Catal.

1986, 26, 353

14. Kip, B. J.; Duivenvoorden, F. B. M.; Koningsberger, D. C.; Prins, R'.

J. Catal. 1987, 105, 26.

15. Castner. T.. Newell. G. S .. Holton. W. C.. Slichter. C. P. J. Phys. Chem. 32. 668 (1960).

16. Wickman. H. H .. Klein. M. P .. Shirley. D. A. J. Phys. Chem.42. 2113 (1965).

17. Dowsing. R. D .. Gibson. J. F. J. Phys. Chem. 50. 294 (1969).

18. Mehring, M. "High Resolution NMR Spectroscopy in Solids";

Springer, 1976

19. Harris, R. K.; Mann, B. E. "NMR and the Periodic Table";

Academic Press, 1978

20. Mossbauer, R. L. Z. Physik 1958, 151, 124

21. Mossbauer, R. L. Naturwissenschaften 1958, 45, 538

22. Mossbauer, R. L. Z. Naturforsch. 1959, 149, 211

23. Anderson, A. B. J. Phys. Chem. 1975, 62, 1187

Page 56: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Experimental page 46

24. Richardson, J. W.; Nieuwpoort, W. C.; Powel, R. R.; Edgel, W. F. J. Phys . Chem. 1962, 36, 1057

25. Clementi, E. ; Raimondi, D. L. J. Phys. Chem. 1963, 38, 2686

26. Basch, H.; Gray , H.B. Theor. Chim. Acta 1966, 4 , 367

27. Lotz, F. W J . Opt. Soc. Am. 1970, 60, 206

28. Moore, C E. Atomic Energy Levels; NBS Circ. no. 467; National

Bureau of Standards ; U.S. Government Printing Office; Washington, DC, 1958

29. Anderson, A. B.; Grimes, R. W; Hong, S. Y. J. Phys . Chem. 1987,

91,4245

30. Stern , A . E, "X-Ray Absorption"; Koningsberger, D. C. ; Prins, R ..

Eds.; John Whiley & Sons, 1987; Chapter 1, p. 3

31. Teo, B. K.; Lee, P A. J. Am. Chem. Soc. 1979, 101, 2815.

32. Citrin, P. H.; Eisenberger, P.; Kincaid, B. M. Phys. Rev. Let . 1976,

36, 1346

33. Sinfelt, J. H.; Via, G. H.; Lytle, F. W.; Greegor, R. B. J . Chem. Phys. 1980, 72(9), 4832

34. Sayers, D. E. "X-Ray Absorption" ; Koningsberger, D. C.; Prins, R.,

Eds.; John Whiley & Sons, 1987; Chapter 6, p. 211

35. van Zon, J. B. A. D.; Koningsberger, D. C.; van 't Blik , H. F. J.; Sayers , D. E. J . Chem . Phys. 1985, 12, 5742.

36. Heald, S., "X-Ray Absorption"; Koningsberger, D. C.; Prins, R., Eds.; John Whiley & Sons , 1987; Chapter 3, p. 87

Page 57: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 47 Chapter 3

Chapter 3

The preparation of y-Al20 3 supported Monometallic Rh and Pt and Bimetallic Rh-Pt Catalysts

3.1 Introduction

The introduction of the metal precursor onto the internal sur­

face area of the support, the very first step in the preparation of a

supported metal catalyst, is of vital importance. This first step dic­

tates the lower limit of the size of the eventual metal particles, and

in addition, it may be the prelude for the formation of bimetallic

particles. Here we will describe the pore volume impregnation

method and the adsorption method to prepare RhCl3 and H2PtCl6 catalyst precursors supported on y-Al 20 3. For the H2PtCl6 sam­

ples, we used 195Pt NMR and Laser Raman spectroscopy to follow the growth of H2PtC16 crystallites during the pore volume impregna­

tion method. From these experiments it became evident that above

a loading of 0.18 mmol H2PtC16 per gram of Al 20 3 crystals of

H2PtCl6 started to form. This confirmed the suggestion in the

literature that H2PtCl6 can adsorb as single entities on the Al20 3 support at low loadings ( 1 ,2). Furthermore, we studied the adsorp­

tion of H2PtCl6 and RhCl3 on y -Al 20 3. Furthermore, we will

describe TPR profiles of RhCl3, H2PtCl6 · and (RhCl3 + H2PtCl6)

supported on Al20 3 in order to follow the formation of bimetallic

particles.

Page 58: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh. Pt and Rh-Pt/ Al 20 3 page 48

3.2 Experimental

3.2.1 NMR and Laser Raman Experiments

In all experiments we used y-Al 20 3 from Ketjen (000-1.5E)

with a internal surface area of 180 m2 g-1 and a pore volume of

0.65 ml g- 1. The radius of the y-Al 20 3 particles was approximately

0.1 mm. H2PtC1 6/A1 20 3 samples were prepared containing 0.087,

0.169. 0.215. 0.256. 0.297 and 0.343 mmol H2PtCl 6 per g Al 20 3 using the pore volume impregnation method. The samples were

dried at room temperature for 24 h and at 393 K for another 24 h.

The l\IMR experiments were carried out on a Bruker CXP300

spectrometer under Magic Angle Spinning conditions in a Beams­

Andrew rotor. For each spectrum, 4096 Bloch decays were accumu­

lated at room temperature and Fourier transformed to give the

resulting spectrum.

For the Raman experiments, the dried samples were pressed

into circular tablets which were mounted in a holder and rotated at

an angle of 45° with respect to the laser beam, in order to avoid damage to the sample by heating and possibly burning. The mono­

chromator was aligned with respect to the reflected beam. thus at

an angle of 45° to the sample. Using a dye liquid, the laser was

tuned to a wavelength of 5140 A and a power of 90 mW. The spec­

tra were recorded at room temperature, at energies between 50 and

500 cm -l away from the energy of the incident light.

Page 59: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 49 Chapter 3

3.2.2 Adsorption Experiments

For the adsorption experiments, two solutions containing

RhCl3 (15.5 and 31.2 mM) and two solutions containing H2PtCl6 (9.76 and 17.5 mM) were prepared. Various amounts of Al20 3,

ranging from 0.1 to 1.4 g, were added to a fixed amount of these

solutions (20 ml for the lower concentrated and 10 ml for the higher

concentrated solutions). After allowing the metal salts to adsorb

on the support for four days, the equilibrium concentrations of

RhCl3 and H2PtCl6 in the solution were measured. From these con­

centrations, the amount of metal salt adsorbed on the support could

be calculated. For the set of lower concentrated solutions, the pH value of the solution in equilibrium was measured as well. The con­

centrations of RhCl 3 and H2PtC16 were measured according to the

following procedure . 5 ml of a concentrated HCI solution and 5 ml

of a 1 M SnCl2 solution were added to 5 ml of the metal salt solu­

tion. This mixture was heated on a water bath to the boiling point.

Highly colored [Rh2C14(SnC1 3) 4( and [PtC12(SnC13ht complexes were formed, and their extinction coefficients were measured at

A. = 471 nm for the rhodium complex and at A. = 401 nm for the platinum complex. For the rhodium complex, the extinction

coefficient is linearly dependent on the concentration only between 4 and 20 mM and for the platinum complex between 3 and 12 mM.

Therefore, in some cases the amount of 5 ml metal salt solution

mentioned above had to be diluted with distilled water in order to

assure that the conc~ntration of the final solution was in the desired range. For both complexes, the extinction coefficient was calibrated.

In order to establish the influence of the acidity of the solution on the adsorption process, the same procedure, using the lower con­

centrated solutions, was followed again. but in this case, after the A120 3 had been added, and during the adsorption process, the pH value of the solution was maintained at 1.9 ± 0.2 using a 0.1 M HCI

solution.

Page 60: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al 20 3 page 50

We also prepared 5 solutions of 20 ml for both the rhodium (15 .5 mM) and platinum (9.76 mM) salt and added about 0.1 g Al 20 3 to the solution. Directly after the addition of A1 20 3, the pH

value was brought to values ranging from 0.5 to 5.0. After allowing the adsorption process to come to equilibrium, the pH value and the amount of adsorbed metal salt were measured . We repeated these experiments, by bringing the pH of the solution to values ranging from 0.1 to 3.0 before the addition of the Al 20 3. After a week, the pH value and the amount of adsorbed metal salt were measured .

Finally, in order to establish whether there was any competi­tion between RhCl 3 and H2PtC1 6 during the adsorption process , 0.6 g Al20 3 was added to a solution of 10 ml containing 31.2 mM RhCl 3 and 17.5 mM H2PtCl6. After a week, the rhodium and platinum con­tent of the solution and the support were measured using the pro­cedure described above.

3.2.3 TPR of Rh, Pt. and Rh-Pt / Al20 3

Six bimetallic Rh-Pt/y -Al 20 3 precursor catalysts were prepared using the pore volume impregnation method . For all the samples. the RhCl 3 and H2PtCl6 loadings were 0.100± 0.005 mmol g- 1 Al 20 3. Four of these samples were prepared using equimolar solutions of rhodium chloride (RhCl3·3H 20) and a platinum chloride. The acidity of the solution and the chloride con­tent were increased by using respectively PtCl4·5H 20 and H2PtCl6·3H 20, dissolved in distilled water, and H2PtCl6 dissolved in 0.5 and 1.0 M HCI. 0.5 M HCI was equivalent with 3 HCI molecules per rhodium or platinum atom, 1.0 M HCI with 6 HCI molecules per rhodium or platinum atom. These samples will be denoted as RPA1, RPA2, RPA3 and RPA4, respectively.

Page 61: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 51 Chapter 3

The other two samples were prepared using equimolar

amounts of RhCl3·3H 20 and H2PtCl6·3H 20 dissolved 1n either

CH 30H or C2H50H . The sample impregnated with CH 30H is

denoted as RPA5, the sample impregnated with C2H50H as RPA6.

After impregnating and carefully drying, TP R experiments were car­

ried out as described in chapter 2, using a heating rate of 5 K min -1.

Drying of the samples impregnated with CH 30H and C2H50H gave

problems. Drying at elevated temperatures (up to 500 K for 10 h,

even in vacuum) was insufficient to remove the alcohol completely.

The only effective procedure we found was to dry the sample at

393 K for a few hours and then for at least two weaks at room tem­

perature. Obviously, CH 30H and C2H50H adsorbed strongly on the

y-Al 20 3 and desorption was very slow.

3.3 Results and Discussion

3.3.1 NMR and Laser Raman Experiments

In Figure 3.1a, . the NMR spectra of four supported H2PtCl6

samples are shown. The chemical shift indicated that the peak

observed originated from Pt in IPtC16f octahedra ( 3,4). Apart from

this, no other peaks were observed. As a measure for the intensity

of the peaks, we used the peak height (in arbitrary units), multi ­

plied by the width of the peak at half height (in arbitrary units).

These intensities are tabulated in Table 3.1 and plotted in Figure

3 .1b as a function of H2PtCl6 loading . As indicated in Figure 3.1b.

a straight line can be constructed through these points and this line

intersects the abscissa at a loading of 0.16 ± 0.01 mmol H2PtCl6 per g Al 20 3. Obviously, H2PtCl6 crystals start to form only above

this limit. Below this limit, the platinum salt are adsorbed strongly on the support, as will be demonstrated by the adsorption experi­

ments . We could not observe other peaks in the spectrum that

Page 62: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al20 3

Figure 3.1 Results of the NMR experiments .

1

2

3

4

(a) NMR spectra of H2PtCldAl20 3 with loadings of :

1 : 0.215 mmol g- 1, 2 : 0.256 mmol g- 1

3 : 0.297 mmol g- 1, 4 : 0.343 mmol g- 1

(b) NMR intensities as a function of H2PtC1 6 loading

400 b

200

I

I I

I I

page 52

I

0-=---~~~~'-=---=-~~~~---,," M Q2 M

Loading-

Table 3.1 NMR data

H2PtCl6 Peak Peak Intensity loading height width (h*w)

(mmol gr 1 (a .u.) (a .u.) (a .u.)

0.215 12.0 10.0 120 0.256 17.5 9.5 166 0.297 40.0 7.5 300 0.343 57 .0 6.5 371

possibly could be attributed to such adsorbed complexes. A possi­

ble reason might be the following. It is known that small devia­tions in the chemical environment of platinum result in enormous

chemical shifts and this alone makes it difficult to detect them.

Page 63: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 53 Chapter 3

Therefore, platinum chloride complexes which are adsorbed on the

support and which most probably have exchanged one or more

chloride ions for an oxygen ion or a hydroxyl group of the support,

will have a different chemical shift than [PtCl6t complexes. Evi­dence for such an exchange was found by Lagarde (5), who

observed that after impregnating a y-Al20 3 support with H2PtCl6,

the platinum complexes had lost approximately one Cl- ligand

(NPt-CI = 5.0 ± 0.5) . The loading of that sample, prepared via pore volume impregnation, was low enough to assure that no cry­

stalline H2PtCl6 had been formed (1.5 wt% Pt, which is equal to .

0.077 mmol per gram Al20 3; the surface area of the Al20 3 was

240 m2 g-1). In addition, we expect that the variation in the chemi­

cal environment of these strongly adsorbed complexes may be (rela­tively) large, which will give rise to a large spread in chemical shifts and therefore to low and broad peaks, which are difficult to detect.

In Figure 3.2a, the Raman spectra of a H2PtCl6 solution and of

three supported H2PtCl6 samples are shown. In these spectra, the

Raman active frequencies of the [PtC16]2- units are clearly visible.

The line width of the peaks around 340 and 320 cm -l could not be

measured accurately. Therefore, the peak height was taken as a measure for the intensity. These 'intensities' are collected in Table

3.2 and depicted in Figure 3.2b. The general behavior of the inten­sities as a function of loading is the same for the three bands.

Above a loading of about 0.18 ± 0 .02 mmol H2PtCl6 g-1 Al20 3, the

intensity increases more with increasing H2PtCl6 loading than below

that loading. This can be explained as follows. Below the limit of 0.18 mmol H2PtCl6 g-1 Al20 3, H2PtCl6 adsorbed as single entities on

the support and these adsorbed platinum complexes were not per­

fect [PtC16f octahedra, as could be concluded from the shift in the

frequencies in the spectra (cf. Table 3.2). The Raman frequencies

are related to the vibrational force constants. Since these frequen­cies changed only slightly, the force constants had not changed

drastically, which indicates that the [ PtC 16]2- octahedra were

slightly deformed or had exchanged one (or more) chloride ions for

-0 H- groups of the support. As discussed for the NM R

Page 64: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al20 3 page 54

Figure 3.2 Results of the Raman experiments .

a

1

(a) Raman spectra of H2PtCICi and H2PtCl<i/Al20 3 (spectra are not to scale) :

1 : H2PtCl6 in aqueous solution,

2 : 0.169 mmol g- 1, 3 : 0.256 mmol g-1

, 4 : 0.343 mmol g-1

(b) Raman intensities as a function of H2PtCICi loading

340 cm-1 : solid line, 320 cm-1

: dashed line, 150 cm-1 : dotted

line

2

b

~ I

Q /

,•

~~ ,

/

>- ,i' (/) 0

J

c .O" (1) • _.,...--- ,, c .··

I

4 ,,

. ..0 ... ·

0 " 0 "

0 " 400 100 0.0 0.2 0.4

-Wavenumber (cm·1 ) Loading --

experiments, this will be accompanied by a broadening of the bands and therefore, the peak heights will be lower. Above the limit of 0.18 mmol H2PtCl6 g-1 Al20 3, the frequencies of the three bands corresponded precisely with the frequencies of pure H2PtCl6, which indicated that (perfect) H2PtCl6 crystals had started to form. The solid line in Figure 3.2b depicts the intensity of the bands of pure H2PtCl6 in the samples and the dotted line the estimated contribu­tion of the adsorbed platinum complexes.

Page 65: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 55 Chapter 3

Table 3.2 Raman data

H2PtCl6 Frequency Peak loading height

(mmol g- 1) (cm- 1) (a.u .)

0.087 338 0.42 324 0.30

0.169 343 1.18 325 0.96 154 0.35

0.256 343 1.53 319 1.34 159 0.48

0.343 343 3.36 316 3.04 157 1.09

The main conclusion from the NMR and Raman experiments is

that up to loadings of 0.16-0.18 mmol H2PtCl6 g- 1 Al20 3 platinum

complexes were adsorbed on this support, and were probably molec­

ularly dispersed. These complexes differed from perfect [PtCl6)2-

octahedra and had probably exchanged one or more Cl- ligands for

oxygen ions or hydroxyl groups from the support. Above this limit,

H2PtCl6 crystals had started to form. This limit is far above the

amounts of H2PtClri normaliy used to prepare Pt/ Al 20 3 catalysts.

A metal loading of 2 wt% corresponds to a loading of 0.1 mmol H2PtCl6 per g Al 20 3. After reduction, these catalysts are

never atomically dispersed . Typical particles sizes for Pt/A1 20 3

catalysts are about 8 A and these particles contain about 10 plati­

num atoms (6). Thus, the formation of these larger metal particles

in the final form of the catalyst must have taken place during the

reduction procedure. It should be noted, that these results were obtained for a y-Al 20 3 with a surface area of 180 m2 g-1. For other

aluminas, the results may be different.

Page 66: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al20 3 page 56

3.3.2 Adsorptwn Experiments

Table 3.3 Adsorption isotherms for RhCl 3 and H2PtCl6 on Alp3

Aip3 [M]a [M/5] 0 pH Aip3 [M]a [M/5] 0 pH

f g) (g)

a. RhCI, (10 ml. 31.2 mM) b. H1PtCI" (10 ml. 17.5 mM) 0.6196 11.0 0.326 0.6527 2.77 0.226 1.0394 0.117 0.299 1.1057 0.038 0.158 1.1333 0.071 0.274 1.2092 0.024 0.145 1.2478 0.022 0.250 1.3352 0.059 0.131

c. Rh Cl, (20 ml. 15.5 mM) d. H1PtCI" (20 ml. 9.76 mM) 0.1031 13.1 0.47 3.0 0.1023 8.5 0.24 2.6 0.5868 2.8 0.433 3.2 0.5927 2.7 0.238 2.7 0.9362 0.30 0.324 3.4 0.9106 0.66 0.200 2.9 1.0052 0.22 0.302 3.5 1.0097 0.37 0.186 3.2 1.1067 0.144 0.276 3.6 1.0993 0.24 0.173 3.4 1.1522 0.089 0.266 3.6 1.1429 0.173 0.168 3.5 1.1905 0.084 0.258 3.6 1.2018 0.128 0.161 3.6 1.2605 0.047 0.244 3.7 1.2450 0.103 0.155 3.7 1.3917 0.018 0.222 3.7 1.2939 0.072 0.150 3.7

a [M] is the concentration ( mM) of the metal salt in the solution above

Alp3 in equilibrium

b [M/S] is the metal-salt loading (mmol g- 1) on the Alp3 support

In order to obtain information of the state of the complexes adsorbed on the support, we performed adsorption experiments to

establish adsorption isotherms for RhCl 3 and H2PtCl 6 on Al 20 3.

Table 3.3 summarizes the results. The maximum amount of RhCl3 that was adsorbed was 0.326 ± 0.009 mmol g- 1 for the more concen­trated and 0.47± 0.1 mmol g-1 for the less concentrated RhCl3 solu­

tion. For H2PtCl6, these amounts were 0.225 ± 0.003 and 0.24± 0.07 mmol g-1

, respectively. (l\Jote that for the experiments where the equilibrium concentration was lower, the uncertainty in the loading will be lower as well. The uncertainty in the concentra­

tions ranged from 2 to 4%. For concentrations > 1 mM. the

Page 67: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 57 Chapter 3

uncertainty is about 4%. for concentrations > 0.1 mM ± 3%. for

concentrations > 0 .01 mM ± 2%. Thus, the uncertainties in the

maximum attainable coverages are highest. This rjtiight suggest

that the uncertainty in the actual maximum attainabl.e coverages is

high . But the uncertainties in the second figures in li able 3.3a, b, c

and d are much lower . These uncertainties are about 0 .002-0.004) .

During the experiments with the lower concentrated solutions, the

pH value of the solution in equilibrium was measured. At the max­

imum attainable coverage, the pH value of the solution was 3.0 for

the RhCl 3 solution and 2.6 for the H2PtCl 6 solution . At low metal

salt coverages, the pH value was about 3.6. A1 20 3 served as a

buff er agent and stabilized the pH value to 3.6 ± 0.1 over a wide

concentration range.

Table 3.4 Adsorption capacities for RhCl3 and H2PtCl6 at low pH values

Al20 3 [M)a [M/S]b pH HCI added (g) (ml , 0.1 M)

a. RhCl3 (20 ml. 15.5 mM)

0.6857 10.6 0.056 1.7 5.2 0.8756 9.86 0.056 1.8 6.1 1.0113 9.62 0.051 1.7 6.6 1.1526 9.22 0.048 1.7 7.6 1.2100 9.00 0.051 1.9 7.3 1.2987 8.88 0.047 2.0 7.7

b. H2PtCl6 (20 ml. 9.76 mM)

0.9188 6.66 0 033 1.7 6.7 1.0061 6.54 0.032 1.7 7.0 1.1007 6.19 0.036 1.7 7.2 1.2031 6.02 0.036 1.8 7.3 1.3350 5.85 0.036 2.0 7.5

a [M] is the concentration (mM) of the metal salt in the solution above A1p3 in equilibrium

b [M/S] is the metal-salt loading (mmol g- 1) on the Al20 3 support

Page 68: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al 20 3 page 58

In Table 3.4, the results of the adsorption experiments at a

constant pH value of 1.7-2.0 are summarized. From these results,

it is clear, that at this low pH value the maximum attainable cover­

age was 0.056 mmol g-1 for RhCl3 and 0.036 for H2 PtCl6 . This is

almost an order of magnitude lower than the maximum attainable

coverages reported above.

Table 3.5 Adsorption of RhCl 3 and H2PtCl6 on Al20 3 as a function of

acidity

A1p3 pH-values [Mt [M/SJ0 A1p3 pH-values [M]a [M/SJ

0

(g) start end start end

1. HCI after Al')O~ la. RhCI, (20 ml. 15.5 mM) lb. H?PtCI,, (20 ml. 9.76 mM)

0.098 0.5 0.7 15.3 0.07 0.093 0.4 0.7 9.8 0.00 0.112 1.4 1.5 13.8 0.31 0.105 1.7 3.5 9.2 0.23 0.113 2.5 3.4 12.8 0.50 0.114 2.8 3.4 8.6 0.32 0.111 3.5 3.5 12.1 0.61 0.113 4.2 3.5 8.2 0.38 0.118 4.5 3.8 10.3 0.83 0.127 5.2 3.9 7.5 0.45

2. HCI before Al?01

2a. RhCI, (20 ml. 15.5 mM) 2b. H7PtCl6 (20 ml. 9.76 mM)

0.138 0.3 0.8 15.9 0.00 0.245 0.1 0.8 9.8 0.00 0.102 1.3 1.6 15.9 0.02 0.164 0.9 1.1 9.9 0.03 0.189 2.3 2.7 11.2 0.51 0.157 1.8 3.1 8.3 0.28 0.159 3.1 3.8 10.8 0.58 0.267 2.9 3.8 5.4 0.30

a [ MJ is the concentration ( mM J of the metal salt in the solution above A1p3 in equilibrium

b [M/SJ is the metal-salt loading (mmol g- 1) on the A1p3 support

Table 3.5 summarizes the results of the adsorption experi­

ments which were performed using different starting acidities.

From these results, it is clear that acidifying the solution had a

major impact on the adsorption capacity of the Al 20 3. It did not

matter whether the pH value of the solution was adjusted before or

after the addition of Al 20 3. The reason is, that the process of buffering the solution to a pH value of 3.6 was rather slow. Typi­

cally, it took several hours for the A1 20 3 to adjust the pH value of

the solution to 3.6. To bring the solution (before or after the

Page 69: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 59 Chapter 3

addition of the A120 3) with HCI to the values listed in Table 3.5 was a matter of a few minutes . During that short period of time,

· the Al20 3 had no time to buffer the pH value of the solution.

Therefore, the two sets of experiments can be regarded as the same. Note, that in case the starting acidity was lower than about

1.5, the Al20 3 did not manage to stabilize the pH value to

3.6 ± 0.1. The amount of Al 20 3 has been chosen low in order to

ensure that complete coverage would be obtained in all cases.

Therefore, the loadings in Table 3.5 reflect the maximum adsorption

capacity as a function of equilibrium pH value. The obvious con­

clusion is, that the higher the starting pH value, the more RhCl3 or

H2PtCl6 can be adsorbed on the A120 3.

At this point, it is easy to understand why a lower amount of

metal salt was adsorbed when a more concentrated solution of

these. salts was brought in contact with Al20 3 (cf. Table 3.3). The

more concentrated solutions obviously has a lower pH value and

therefore less metal salt was adsorbed from these solutions.

A similar pH-dependence for the adsorption of H2PtCl6 on y­Al203 has been reported by Heise et al. (7 ). They found a max­

imum in the adsorption capacity at pH 3.5-4.0. In that study, the penetration depth was measured of the platinum complexes in y­A203 pellets mounted above the solution and in contact with that

solution. The penetration depth they reported (0.2-0.5 mm) is larger than the radius of the Al20 3 particles we used (0.1 mm).

Thus, there is no reason to assume any depletion of the solvent in the pores during our experiments.

In order to explain the results of these adsorption experiments,

we must focus on the reactions which may take place during the

adsorption process. For example, for RhC13, the following reactions are of importance :

Page 70: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh , Pt and Rh- Pt/ A120 3 page 60

RhCl3 + 3 H20 -+ RhC13(H 20)3 (3.1)

Rhc13(H2o h + H2o +:t !Rhc13(H 20)20Hr + H3o+ 13.21

We assume that in aqueous solutions the majority of the oxy­

gen ions from the support which are exposed to the solvent are sur­

face -0 H groups. In acid solutions , such as solutions of RhCl3 and

H2PtCl 6, these hydroxyl groups will be partially protonated :

l-OHJIH30+J

1-0Ht)

101 I [-OH2+] I og I-OH]

OH

13.3)

[3.4)

+pH

Cl

[3.5)

. I / Cl H20-/ Rti-OH2

Cl +H20

0

Thus, in the acid solutions of RhCl3 and H2PtCl6, the Al20 3 surface

is positively charged and the rhodium complexes are negatively

charged or are neutral. Adsorption of the complexes to the surface

of the support will therefore take place, due to Coulomb forces. A possible mechanism for the adsorption of rhodium complexes is

shown in equation [3.5]. The amount of protonated surface hydroxyl groups can be calculated from the pH value of the solution

and the pK0 value which is assumed to be equal to 3.6, the pH

Page 71: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 61 Chapter 3

value of solutions stabilized by Al 20 3. In (7 ), Heise et al. found

that at pH= 8, the A1 20 3 surface was electrically neutral. Thus, at this pH, the fraction of protonated sites will be close to zero and

the fraction of unprotonated sites will be close to unity . Assuming that the amounts of protonated and non-protonated sites vary with

the pH of the solution as reported by Bowers (8), we can estimate

that the amount of protonated and unprotonated surface hydroxyl

groups will equal each other at approximately pH= 4-5, which is close to the value of 3.6 we assumed . Using equation (3 .4J, we can

estimate the amount of protonated and non-protonated sites on the Al 20 3 surface.

Table 3.6 Amount of protonated and non-protonated surface hydroxyl

groups as function of pH value

pH f -OH 2+ f -OH N -OHt N-oH

(mmol g-1) (mmol g-1)

3.6 0.50 0.50 2.02 2.02 3.0 0.75 0.25 3.03 1.01 2.5 0.92 0.08 3.72 0.32 2.2 0.96 0.04 3.88 0.16 2.0 0.97 0.03 3.92 0.12 1.8 0.98 0.02 3.96 0.08

f -OH fraction non-protonated surface hydroxyl groups

f -OHt fraction protonated surface hydroxyl groups

N -OH number of non-protonated surface hydroxyl groups

N -OHt number of protonated surface hydroxyl groups

In Table 3.6, the fraction of protonated and non-protonated surface hydroxyl groups and the estimated amounts of these groups

are presented . The total amount of surface hydroxyl groups has been estimated assuming that all Al20 3 crystal faces exposed were

(111) faces, that the radius of oxygen ions is 1.4 A and that the

Page 72: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al 20 3 page 62

Al 20 3 had an internal surface area of 180 m2 g-1. With these

assumptions, we calculated the amount of surface hydroxyl groups

to be 4.04 mmol per g Al 20 3. This corresponds to 13.5 surface

hydroxyl groups per square nm. As is to be expected, the amount

of protonated surface hydroxyl groups increases with increasing aci­

dity. The amount of asdorbed RhCl 3 was found to decrease drasti­

cally at lower pH values. This must be due to the fact that upon

complexation of the rhodium complex by an -OH ligand from the

suppdrt, a water molecule is replaced by an oxygen ion ligand and a

hydronium ion is formed (cf. equation [3.5))

-OH+ H20+ RhCl3(H20h = -OH+ [RhCl3(H20)20Hf + H30+ [3.6]

# ~ J 11 ~ Hp++ er+ H20 + H30+ + 2 H30~ + H3 0++ Cl-+ H•+

r H20 H20 2- r OH 1-OH

l/OH2 I ,..,,.OH2 l/OH2 I l/OH2 ! Cl-Rh-Cl Ci-Rh-Cl Cl-Rh-Cl I Cl-Rh-Ci· l H20 ~ c( I c( I lH2~I !

0 0 0 j

In scheme [3.6] a more complete reaction network is depicted, in which we assumed that adsorption takes place on a non­

protonated surface hydroxyl group. In the solution, there is an

equilibrium between the complexes [RhCl3(H 20)J] and (H 3o+ +) [RhCl3(H 20)iOHr. On the surface of the support, four different

complexes can exist : [-O-RhC1 2(H 20)J], [-O-RhC13(H 20) 2r, [-O­

RhC13(H20)0Hf and [-O-RhC12(H 20)ioHr. These complexes result from adsorption of the two complexes in solution; during

adsorption, a Cl- ion, a H20 molecule or a OW ion of the complex

in solution has been exchanged. From these six species, the

[RhC13(H 20)3] complex in solution is the only stable species at high Cl- concentrations and low pH values. Thus, under these

Page 73: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 63 Chapter 3

circumstances. adsorption will be small. On the other hand. at

higher pl I values. but still in acid media. and lower Cl - concentra­

tions. the l-O-RhCl?(H 20)J] complex adsorbed on the support will

be the most stable species and. hence. under these circumstances.

the coverage of metal salt on the support will be relatively high.

We will now focus on the adsorption of platinum complexes.

Cl Cl 2-

I / Cl I /Cl Cl -:fit-- Cl Cl-Pt-Cl

Cl I ;:::! c(I +H 30+ 13 .8] OH 2

0

OH I I

For H2PtCl6• we propose the reaction scheme depicted in 13.7]

and IJ.8). The exchange of a Cl- ion by a OW ion or a H20

molecule has been reported by Cox (9), v.d. Berg (JO) and by Bol­

man ( 11 >. Since in our studies the processes took place in add solu­

tions. we assumed that in the majority of cases a H20 molecule

rather than an OW ion will be exchanged. Reaction IJ .8] is compar­

able to reaction IJ.5] . Another possible mechanism is the direct

exchange of a chloride ion in the platinum complex with a surface

hydroxyl group :

In ( 1 ), a slightly different model has been described in which a

Al3+ ion 'aided' the adsorption of a 1PtCl6]2- unit to the surface of

the Al 20 3. We reject this model based on the following considera­

tions :

(i) with l\IMR. we detected no 1PtC16( units in the cases where

only adsorption had occurred, and

Page 74: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh- Pt/ Al20 3 page 64

Cl 2-

Cl 2-

I /Cl Cl-Pt-- Cl

l/c1 Cl-Pt-Cl

c(I /

(3.9] H20 +:! Cl H3o++c1-Cl

0 OH

(ii) at lower pH values, the amount of dissolved Al20 3 should

increase and therefore the amount of Al3+ in the solution as

well. This would , according to the model in ( l) lead to a

higher coverage, which is in contradiction with our observa­

tions.

(3.10] 2- CL Cl

2-

c1~[_-c1 Pt

I \ 0 0

OH OH I I

Another model for adsorption of [PtC16j2- to a support surface

has been reported by Le Page ( 12) and has been used by Castro

el al. ( 13) and is given by 13.10]. However, when, like in [3.10], only

-OH groups would be responsible, this process is very unlikely to

occur since the amount of -0 H groups is low and the amount of

-OH groups which have a neighboring -OH group will be lower by

an order of magnitude at pH values lower than 3.6. We therefore

conclude that equations 13.5] and (3.8], or [3.5] and [3.9], sufficiently

describe the processes that take place during adsorption of RhCl3 and H2PtCl6. The main conclusion from these experiments is that

unprotonated surface hydroxyl groups play a key role in the

Page 75: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 65 Chapter 3

adsorption process.

· In ( 14), Heise er al. proposed a different model to explain the

adsorption of !PtC16f on y-Al 20 3. They showed that the thick­

ness of the electrical double layer and the platinum ion activity

could be related to the adsorption. When the electrical double layer

decreases in thickness because of an increase in the concentration

or valnce or the electrolytes, the amount of adsorbed [PtC16]2-

decreases also. Alternatively, when the platinum ion activity

increases, the affinity of [PtC16]2- towards the solution increases and

therefore, the amount of adsorbed !PtC16)2- decreases. The expla­

nation of Heise et al. is based on a macroscopic model, while we

focussed on a molecular-microscopic model. Their explanation does not account for real bonding on a molecular scale, while our expla­

nation with equilibria between complexes in solution and bonded to

the surface does not account for long-range effects induced by the

electrical double layer and the platinum ion activity. In a sense, the

two models are complementary . But real chemical bonding of a

metal complex to the support has to be invoked in order to explain

why many metal complexes cannot simply be washed away from

the support.

In the NMR and Raman experiments, we found that above a

certain H2PtCl6 loading H2PtCl6 crystals had started to form. We

therefore assumed that below that loading platinum complexes were

adsorbed as single entities on the support. With the adsorption

experiments we found that indeed y-Al20 3 can adsorb H2PtCl6 (and

RhC13) and that the adsorption capacity was limited. Care should

be taken, however, in comparing these results . The NMR and

Raman experiments were performed with dried catalyst samples

prepared with the pore volume impregnation method , while the

adsorption experiments were actually performed in solution. During

the drying step in the pore volume method, the solvent is removed

slowly from the pores and, thus, the remaining solution slowly

retracts to the inner parts of the support. Re-crystallization may

occur during this drying step, even when before drying the metal

Page 76: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al 20 3 page 66

salt was adsorbed in a mono-disperse way . Consequently. in the

pore volume impregnated samples used in the NMR and Raman

experiments, only part of the support surface may have been util­

ized and therefore , the loading of 0.18 mmol per g Al20 3 reflects a

lower limit when we want to compare the results of the pore

volume impregnation method with the adsorption experiments .

The simultaneous adsorption experiment, in which 0.6 g of

Al 20 3 was added to 10 ml of a solution containing 31.2 mM RhCl 3

and 17.5 mM H2PtCl6, showed that 0 .212 mmol RhCl3 and 0.138

mmol H2PtCl 6 could be adsorbed per gram Al 20 3. The amount of

A1 20 3 used assured that these figures reflect the maximum attain­

able coverages . Since the starting solutions were more concentrated

than any of the solutions mentioned before, we assume that the pH of the solution after equilibrium was established , was lower than

the lowest pH value of the solutions in Table 3.3. Both the adsorp­

tion capacities for RhCl 3 and H2PtCl6 decreased . Strikingly, how­

ever, the sum of the two adsorption capacities, 0 .35 mmol g-1, was

about equal to the maximum attainable coverage for RhCl 3 in Table

3.3a, 0.33 mmol g-1. This might suggest, that there were two kinds

of adsorption sites, sites on which only RhCl3 can adsorb and sites

on which both RhCl 3 and H2PtCl6 can adsorb. Conformation of this

idea can be found in Tables 3.3c, 3.3d, 3.4 and 3.5 : At any given

pH value, the maximum attainable coverage for RhCl3 was always

higher than that for H2PtCl6, even when we compare RhCl3 and H2PtCl6 solutions in which the RhCl3 and H2PtCl6 concentrations

are comparable (cf. Table 3.3b and 3.3c) . From these results, how­

ever, no conclusions can be drawn about the different adsorption

sites. There is no doubt , that the acidity of the sites plays an

important role, for the more acidic sites will loose a proton more

easily, in other words , will be 'neutral', not protonated , at low pH values, whereas the more basic sites will already be protonated .

Page 77: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 67 Chapter 3

3.3.3 TPR of Rh, Pt and Rh-Pt/ Al20 3

Figure 3.3 TPR profiles of :

:::i .;i c .Q a. E :J ,,, c 0 u

c a.> O!

8 "() >-

I

(a) RhC13/ Al 20 3, (b) H2PtC16/ Al20 3

(c) RPA1, (d) RPA2, (e) RPA3, (f) RPA4, (g) RPA5, (h)

RPA6

a e

b

c g

d h

500 700 700

T(K) T(K)

The TPR profiles of the samples RPA1 to RPA6 and the TP R profiles of RhCl3/ A1 20 3. PtC14/ A1 20 3 and H2PtC16/ A1 20 3 are presented in Figure 3.3. In Table 3.7, the results of the TPR experi­ments are summarized. Note, that the metal loading was 0.10 mmol g-1 in all cases. This loading is lower than the limit above which crystalline H2PtCl6 started to form according to the NMR and Raman experiments (0.16-0.18 mmol g-1

). Although this value was obtained for H2PtCl6, we observed that the adsorption capacity for RhCl 3 was always higher than for H2PtCl6 and there­fore we expect that the loading above which crystalline RhCl3 starts to form is at least equal or even higher than the value reported for

Page 78: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ Al20 3 page 68

H2PtCl6. Hence, we assume that in the samples in which RhCl 3,

PtCl 4 and H2PtCl6 were dissolved in water (RPA1 and RPA2), the

salts were adsorbed as single , isolated entities on the support. In

the cases where HCI has been added to the solution (RPA3 and

RPA4), we may expect that the adsorption capacity of the support

has decreased and thus, that crystalline material may have been

formed.

Table 3.7 TPR Data

Sample Metal Salts Solute Tmax

RPA1 RhCl3 PtCl4 Hp 375 RPA2 RhCl3 H2PtCl6 H20 415 RPA3 RhCl3 H2PtCl6 0.5 M HCI 395 RPA4 RhCl3 H2PtCl6 1.0 M HCI 375 RPA5 RhCl3 H2PtCl6 CH pH 345 500 RPA6 RhCl3 H2PtCl6 C2H50H 365 510

RhCl3 H20 415 PtCl4 H20 510

H2PtCl6 H20 505

Tmax is the temperature in the TPR profile of the maximum hydrogen up­take.

For the monometallic samples. RhC1 3/ Al 20 3 reduced at 415 K,

PtC1 4/ Al 20 3 at 510 K and H2PtC16/ A1 20 3 at 505 K. For the bime­tallic samples, imprengated with H20 , the reduction took place in

one step at temperatures where monometallic RhCl 3 was reduced.

For the samples impregnated with HCI solutions , the hydrogen

uptake at higher temperatures increased (cf. Figure 3.3). Using the

results from the adsorption experiments, we can explain these

findings as follows. Going from RPA1 to RPA4 , the acidity of both the solut ion and , because of the increasing chloride content, of the

Al 20 3 increased. From the NMR and Raman experiments we may

Page 79: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 69 Chapter 3

conclude that for RPA1 and RPA2 no crystalline RhCl 3 or crystal­

line H2PtCl 6 was present. Therefore, on these samples. the rhodium

and platinum complexes were adsorbed as isolated complexes on

the support. After the reduction process, the H /M value deter­

mined with hydrogen chemisorption for RPA2 was 0 .77 . Using the

calibration described in (6), we can estimate that the particles were

about 15 A in diameter and contained about 50 atoms . Obviously,

during the reduction process, the atomically dispersed rhodium and

platinum complexes have been reduced and have grown or sintered

to larger metal particles. In order to explain why RhCl 3 and H2 PtCl6

were reduced in one step, we propose the following model. During

the reduction process, the adsorbed RhCl 3 complexes were reduced

first. The metallic rhodium atoms that were formed had only very

little interaction with the support and therefore had a significant mobility. These mobile rhodium metal atoms can 'diffuse' over the

support and, in addition, adsorb and dissociate hydrogen. Therefore,

when they encounter an unreduced rhodium or platinum complex

during this 'random walk'. they can catalyze the reduction of that

complex. Thus, during the reduction process, the average particle

size gradually increases due to the accumulation of additional metal

atoms, until the metal particles are too large to have any significant

mobility. or until all the RhCl 3 and H2 PtCl 6 is reduced. For the more acid solutions, RPA3 and RPA4, we may expect that not only

monodispersed rhodium and platinum complexes have been formed,

but crystalline material as well. These crystallites may 'capture'

diffusing small metal clusters and become reduced. But once they

are reduced, these larger crystallites have no significant mobility and

thus smother the reduction process . This is most pronounced for R PA4 : after a sharp increase, due to the reduction of monodisperse

RhCl3 and the start of the reduction of larger crystallites, the hydro­

gen uptake decreased immediately. The reduction of the remaining

crystallites was not catalyzed but depended on the reducibility of the particles and thus on their (surface) composition : crystallites

containing more RhCl3 were reduced more easily than crystallites containing more H2PtCl6. This explains the tailing behavior of the

reduction profile for RPA3 and RPA4. In the TPR profile of RPA4,

Page 80: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh, Pt and Rh-Pt/ A1 20 3 page 70

some hydrogen uptake was still visible in the region were pure

H2PtCl6 is reduced. This indicates that some monometallic H2PtCl6 crystallites were present. The main conclusion is that the reduction

process was dominated by mobile metal clusters that once formed

catalyzed the reduction of the whole sample, if their mobility was

not reduced drastically. When crystalline material was present, the

reduction process was s.mothered.

The TP R profiles of the two alcohol-impregnated samples

were completely different. There was a peak in the hydrogen

uptake at very low temperatures, below the reduction temperature

of RhCl 3, and a peak exactly at the reduction temperature of

H2PtCl6. This indicates that RhCl3 and H2PtCl6 were separately

present on the support. In the experimental part we noted, that the

alcohols adsorbed strongly on the support. Therefore, we may

assume that there were no sites left for RhCl 3 and H2PtCl6 to

adsorb and thus, that the metal salts were not adsorbed as single

complexes. Hence, during the drying procedure, only crystalline

RhCl3 and H2PtCl6 is formed. However, the hydrogen uptake of the

two separate peaks did not agree with the amount of RhCl3 and

H2PtC16 present in the sample, although the total hydrogen uptake

agreed very well. The hydrogen uptake in the low temperature peak

was too high for the reduction of RhCl3 only. Apparently, some

H2PtCl6 had been incorporated in the RhCl3 crystallites. Why RhCl3 and H2PtCl6 crystallized separately is not yet clear. The RhCl3 crys­

tallites were reduced below the temperature where RhC13/ Al20 3 is

reduced. We believe that this effect is due to the presence of

CH 30H or C2H50H in the crystallites replacing crystallization water.

Pure RhCl3 contains 3 molecules crystallization water per RhCl3.

Because these samples were prepared using CH 30H and C2H50H

and RhCl3.3H 20 (see experimental), during recrystallization part of

the crystallization water may have been replaced by CH 30H or

C2H50H. Because these molecules are substantially larger than

H20, they perturb the structure, and thus may decrease the reduci­

bility of the particle. This effect is most pronounced for the CH 30H sample, because CH 30H is smaller than C2H50H and thus

Page 81: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 71 Chapter 3

could replace crystallization water more easily. For the reduction

peak of the H2PtCl6 crystals we did not observe a lowering of the

reduction temperature. This can be explained by the fact that

H2PtCl 6 already has a very weak crystal structure and the sites in

the structure for the crystallization water are fairly large (see Figure

2.2 in chapter 2). Thus, replacing H20 for CH 30H or C2H50H will

have no major influence on the structure and on the reducibility of

the H2PtCl6 crystallites.

3.4 Conclusions

From the NMR and Raman experiments we conclude that dur­

ing the pore volume impregnation method H2PtCl6 can adsorb as

single, isolated complexes on the Al20 3 support up to a loading of

about 0.16 mmol g-1. Above this limit, H2PtCl 6 crystals are formed.

The adsorption experiments indicated that for RhCl 3 the same

behavior could be expected. The maximum attainable mono­

disperse coverage was even larger for RhCl 3 than for H2PtCl 6. This

maximum attainable coverage was dependent on the pH of the solution, with a decreasing adsorption capacity with decreasing pH. During the reduction process, the monodisperse rhodium complexes were reduced first and because of their mobility they could catalyze

the reduction of the rest of the metal complexes. The presence of larger crystallites smothered the reduction process . When an

alcohol was used as solute, no monodisperse adsorption had taken

place. During the drying process, RhCl 3 and H2PtC16 were

separated . Because there were no monodisperse rhodium complexes

and therefore no mobile rhodium atoms formed during the reduction

process, the RhCl 3 and H2PtCl6 crystallites were reduced separately.

Page 82: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Preparation of Rh. Pt and Rh-Pt/ Al20 3 page 72

3 .5 References

1. Santecesaria, E.; Carra, S .; Adima , I. Ind . Eng . Chem ., Prod . Res . Dev. 1977, 16(1), 41

2. Santecesaria, E.; Gelose, D.; Carra, S. Ind. Eng. Chem .. Prod. Res. Dev. 1977, 16( 1 ), 45

3. Mehring, M. "High Resolution NMR Spectroscopy in Solids"; Springer, 1976

4. Harris, R. K.; Mann, B. E. "NMR and the Periodic Table"; Academic Press, 1978

5. Lagarde , P .; Murata, T.; Vlaic, G.; Freund, E.; Dexpert, H. J. Catal . 1983, 84, 333

6. Kip, B. J.; Duivenvoorden, F. B. M.; Koningsberger, D. C.; Prins, R. J. Catal. 1987, 105, 26.

7. Heise, M. S.; Schwarz, J. A. J . Col. Int. Sci. 1985, 107, 237

8. Bowers, A. R.; Huang, C. P .; J . Col. Int. Sci. 1985, 105 ( 1),197

9. Cox, L. E.; Peters, D. E.; Wehry, E. L. J. lnorg . Nucl . Chem. 1972,

34, 297

10. v.d . Berg, G.H.; Rijnten, H. T. "Preparation of Catalysts II"; Del­

mon, B.; Grange, P .; Jacobs, P . A., Eds.; Elsevier, Amsterdam 1979,

p. 285

11. Botman , M. J. P., thesis, State University of Leiden (Netherlands),

1987.

12. Le Page, J. F. Catalyse de Contact, Technip, Paris, 1978, p. 589

13. Castro, A. A.; Scelza, 0 . A.; Benvenuto, E. R.; Baronetti, G. T.; de

Miguel, S . R.; Parera, J. M: "Prepa.ration of Catalysts Ill"; Pon­

celet, G.; Grange, P.; Jacobs, P . A., Eds.; Elsevier, Amsterdam 1983,

p. 47

14. Heise, M. S .; Schwarz, J. A. J. Col. Int. Sci. 1985, 113, 55

Page 83: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 73 Chapter 4

Chapter 4

Ferric Iron in Reduced Si02 Supported Fe-Ru and Fe-Pt Catalysts

Evidence from Mossbauer Spectroscopy and Electron Spin Resonance

4.1 Introduction

Supported bimetallic catalysts consisting of iron and one of the more noble group VIII metals M (M = Ru, Rh, Pd, Ir and Pt) have extensively been studied by Mossbauer spectroscopy ( 1-11 ). In gen­eral, the Mossbauer spectra of reduced Fe-M/Si02 catalysts contain two contributions, one due to an Fe-M alloy and the other to a doublet with an isomer shift (IS) of about 0.65 mm s- 1 relative to sodium nitroprusside and a quadrupole splitting (QS) in the range of 0.6 - 1.0 mm s-1

. These parameters are entirely characteristic of high-spin Fe3+, and several authors ( 5-11) have made th is assign­ment. Garten ( 1 ), Lam et al. ( 2), Vannice et al. ( J) and Garten and Sinfelt (4), on the other hand, favor the interpretation that the doublet in the Mossbauer spectra of reduced Fe-M /Si02 and Fe­M/ Al 20 3 catalysts corresponds to zero-valent iron atoms in the sur­face of the Fe-M alloy particles. The high isomer shift is explained by the assumption that the electron density for surface iron atoms is lower than for bulk iron atoms ( 2). The Mossbauer spectra of reduced Fe-Rh /Si02 and Fe-Ir /Si02 catalysts, measured in sizu at 4 K, however, do not support the interpretation in terms of zero­valent surface iron but are in agreement with the assignment of the

Page 84: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Fe3+ in Fe-Pt and Fe-Ru/Si02 page 74

doublet to Fe3+ (9,J J ).

From the viewpoint of Mossbauer spectroscopy the assign­

ment of the doublet with the parameters as given above to zero­

valent iron seems unlikely and interpretation in terms of Fe3+ would

be preferred. From a chemical point of view, however, it is not

readily apparent why substantial amounts of ferric iron should be

stabilized in the presence of a noble metal. which in general facili­

tates the reduction of the less noble component, iron. In most

Fe/Si02 and Fe/ Al20 3 catalysts, iron can be reduced to at least the Fe2+ state ( J ,JS), although in some cases, such as the promoted

ammonia or Fischer-Tropsch synthesis catalysts small amounts of

Fe3+ are also observed (JO.JS). In conclusion, the presence of ferric

iron in reduced Fe-M/Si02 catalysts, as deduced from Mossbauer

spectroscopy, seems somewhat unexpected and confirmation by

another in situ technique would be highly desirable.

Electron Spin Resonance (ESR) is very sensitive in detecting

F 3+ . d b 1 · d . . F 3+ . h h e ions an can e app 1e zn sztu. e ions ave a very c arac-

teristic ESR signal centered at g = 4.2 whenever the site symmetry deviates slightly from the perfectly octahedral or tetrahedral sym­

metry ( J2-J4). Trivalent iron has been the subject of many ESR

studies and the corresponding g = 4.2 ESR signal cannot be mis­

taken for divalent or zerovalent iron.

In this note we report ESR and Mossbauer results of reduced

SiOrsupported Fe-Ru and Fe-Pt. These catalysts represent the combination of iron with the least noble and the most noble metal

in the Fe-M/Si02 series. The ESR experiments confirm that both

catalysts contain ferric iron, in amounts comparable to those deter­

mined by Mossbauer spectroscopy.

Page 85: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 75 Chapter 4

4.2 Experimental

Catalysts were prepared by impregnating the Si02 support

(Cab-0-Sil, EH-5, 310 m2 g-1) with aqueous solutions of

Fe(N03)J.9H 20 and RuCl 3.3H 20 or H2PtCl6.6H 20 under frequent stirring, until the incipient wetness point was reached. The Fe­

Ru/Si02 catalyst contained 0.46 wt% iron and 4.15 wt% ruthenium: the Fe-Pt/Si02 catalyst 0.28 wt% iron and 4.72 wt% platinum. The iron was 10% enriched in the isotope 57 Fe. Catalysts

were dried in air at 295 K for a week , at 330 K for 24 h and at 400 K for 72 h. The catalysts were reduced at 400 K for 0.5 h and sub­

sequently at 725 K for 6 h in the Mossbauer in situ reactor.

Mossbauer spectra were measured at room temperature with a

constant acceleration spectrometer . Doppler velocities are reported with respect to the isomer shift of sodium nitroprusside at 295 K. After measuring the Mossbauer spectra , the catalysts were pas­sivated in air at 295 K and transferred to an in situ ESR sample

holder, described in ( 16). The samples were reduced in flowing hydrogen at 725 K. It was checked that Mossbauer spectra of the

catalysts after passivation and rereduction at 725 K are identical to those obtained after the first reduction treatment.

The X-band ESR spectra were recorded with a Varian E-15 spectrometer equipped with an Oxford Instruments ESR-9 continu­ous flow cryostat. In order to quantitatively determine Fe3+ concen­

trations we measured ESR spectrum intensities of the two reduced catalysts and of a reference compound with a known Fe3+ concen­tration (A1 20 3 CK300, Ketjen : 0.03 wt% Fe3+) at different tem­

peratures between 4 and 80 K.

Page 86: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Fe3+ in Fe-Pt and Fe-Ru/Si02 page 76

4.3 Results and Discussion

Mossbauer spectra of the reduced Fe-Ru/Si02 and Fe-Pt/Si02 catalysts are shown in Figure 4.1. The spectra have been analyzed by

computer to determine the l\llossbauer parameters of the iron compounds

present and their spectral contributions: see Table 4.1 for the results. The

spectrum of Fe-Ru/Si02 consists of two quadrupole doublets, one charac­

teristic of iron in hep Fe-Ru (17 ,18) and the other of high-spin Fe3+. The

spectrum of Fe-Pt/Si02 has been fitted with two doublets as well. One is

identical to the doublet reported for iron in an ordered tetragonal Fe-Pt

alloy (19 ), the other doublet is characteristic for high-spin Fe3+. As Table

4.1 shows, the contribution of Fe3+ to the Mossbauer spectra at 295 K of

reduced Fe-Ru/Si02 and Fe-Pt/Si02 is in the order of 80%. This number

should be considered as a lower limit for the actual Fe3+ content, because a

previous study of the Fe-Rh/Si02 system has shown that the recoilless

fraction of Fe3+ is considerably smaller than that of zero-valent iron in the

Fe-Rh alloy ( 19) Hence, the actual Fe3+ content of the reduced Fe-Ru and

Fe-Pt catalysts may well exceed 80%.

Figure 4.2 shows the ESR spectra of reduced Fe-Ru/Si02 and Fe-Pt/Si0 2 and of the Fe3+-containing Al 20 3 reference compound. All spectra show the characteristic Fe3+ spectrum at g = 4.2. The

presence of ferric iron has thus been established. The amount of ferric iron in both catalysts has been obtained by measuring the

ESR intensity at different temperatures. The spectral intensity fol­

lows from the formula

I - H (WPP )2 12.9.1]

in which I is the intensity, H the peak-to-peak height of the spec­

trum (corrected for receiver gain) and W the peak-to-peak line width of the spectrum in Gauss. PP

Page 87: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 77 Chapter 4

Figure 4.1 Mossbauer spectra of reduced Fe-Ru and Fe-Pt/Si02 ca­

talysts, measured in situ under H2 at 295 K

Ci) I ..... .4 c:: :::J 0 (.)

(!)

0 .s >- 1.3

..... "ii) c:: (!) ..... c::

1 :5 Fe-Ru I Si02

-5

nFeRu nFe3+

0

1: 5 Fe-Pt/Si02

nFePt nFe3+

5 -5 0 Doppler Velocity (mm;s)

3.8

3.6

5

Table 4.1 Mossbauer parameters of Fe in Fe-Ru/Si02 and Fe-Pt/Si02

after reduction in H2 at 725 K

Mossbauer Parameters

IS QS % Assigned Catalyst -! mm s- 1 to mm s

Fe-Ru 0.27 0.19 16 Fe3+ in Fe-Ru

0.69 0.71 84 Fe3-'-

Fe-Pt 0.56 0.43 17 Fe0 in Fe-Pt

0.69 0.76 83 Fe3+

Figure 4.3 shows the calculated reciprocal intensity for the three samples as a function of temperature. At temperatures above 10 K the magnitude of 1 I I depends linearly on T. Above 60 K

Page 88: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Fe3+ in Fe-Pt and Fe-Ru/Si02 page 78

Figure 4.2 ESR spectra of the reduced Fe-Ru and Fe-Pt/Si02 catalysts

measured in situ at 4 K. Spectrum c corresponds to the 0.03 wt% Fe3+-in-Al20 3 reference. For clarity, the curves have

been shifted : the arrows correspond tog = 4 .2. The intensi­ties are not to scale .

a

c

lOOG

a Fe-Ru/Si02 b Fe-Pt/Si02

c Al20 3

b

c

saturation occurs, giving deviation from the linear dependence . Since the slope of the linear part of the 11 I curve is inversely pro­portional to the Fe3+ concentration, the latter follows from the for­mula

c, - I ~: II i c, [2.9.2]

in which C is the concentration of Fe3+ in wt%, D the bulk density of the sample and S the slope of the linear part in the reciprocal-

Page 89: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 79 Chapter 4

Figure 4.3 Reciprocal intensities of the g = 4.2 ESR lines versus tem­

perature for the reduced Fe-Ru and Fe-Pt/Si02 catalysts and

the 0.03 wt% Fe3+-in-Al 20 3

:i ~ -.5 ...... ,...

1.0

a Fe-Ru/Si02

b Fe-Pt/SiC>i

c Al20 3

50 T(K)

100

intensity versus temperature plot . The indices s and r denote sam­ple (catalyst) and reference compound (the 0.03 wt% Fe3+-in­

Al203). Table 4.2 summarizes these results.

The ESR analyzes confirm that Fe-Ru/Si02 and Fe-Pt/Si02 catalysts contain substantial amounts of ferric iron which survives reduction in H2 at 725 K, notwithstanding the presence of a noble

metal. For the Fe-Ru/Si02 catalyst, both Mossbauer spectroscopy and ESR indicate that at least 80% of the iron is in the ferric state.

For Fe-Pt/Si0 2, on the other hand, the Fe3+ contents as deter­mined by Mossbauer spectroscopy and ESR are 83 and 40%,

respectively. It should be noted that ESR detects Fe3+ provided that these ions are not antiferromagnetically ordered as in the com­mon bulk iron(JJJ)oxides, Fe20 3 and FeOOH. Also, the intensity of the g = 4.2 signal may depend slightly on the deviation of the site

Page 90: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Fe3+ in Fe-Pt and Fe-Ru/Si02 page 80

Table 4.2 ESR results and comparison with Mossbauer results

Bulk density Slope wt% Fe3+ Sample

ml/g •10- 1 (1) (2) (3)

Al20 3 1.49 1.400 0.03 Fe-Ru 2.33 0.181 0.36 80 84 Fe-Pt 2.33 0.617 0.11 40 83

(1.) : wt% Fe3+ for the catalyst as determined by ESR (2) : percentage of iron present as Fe3+ as determined by ESR (3) : Contribution of Fe3+ to the Mossbauer spectra at 295 K

symmetry from octahedral or tetrahedral. Therefore, the amounts of Fe3+ calculated from the ESR intensities should be considered to be semi quantitative. Nevertheless, they prove that substantial amounts of Fe3+ are present in reduced Fe-Ru and Fe-Pt/Si02 catalysts and that these amounts are of the same magnitude as those determined by Mossbauer spectroscopy.

The fact that relatively large amounts of Fe3+ can be detected

by ESR is in agreement with the conclusion based on Mossbauer spectra that the ferric iron is present in a highly dispersed state, in close contact with the Si02 support (8,9,JO). The reason why unre­duced iron occurs predominantly as Fe3+ in bimetallic Fe-M/Si02 catalysts and as Fe2+ in most monometallic catalysts is probably due to differences in dispersion. According to Guczi ( S), the iron and the noble metal impede each others migration over the support during reduction and maintain each other in a state of high disper­sion. In this view, the formation of ferrous iron in Fe/Si02 catalysts is accompanied by sintering of the iron to some extent. Experiments to test the validity of this explanation are in preparation.

Page 91: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 81 Chapter 4

4.4 Conclusions

With Mossbauer spectroscopy, the presence of ferric iron in

reduced Fe-Ru/Si02 and Fe-Pt/Si0 2 has been observed. However, from the Mossbauer spectroscopy point of view, their might be

doubts about this assignment. In ESR, Fe3+ ions cannot be mis­taken with other Fe ions. Therefore, the ESR experiments clearly

indicated that the assignment made by Mossbauer was indeed correct. Moreover, the amounts of ferric iron as determined by ESR

agreed very well with the amount as found with Mossbauer. There­fore, these observations make the model in which iron and the noble

metal maintain each other in a highly dispersed state, a state in which iron cannot be reduced to ferrous iron, very Ii kely.

4.5 References

1. Garten, R.L. 11 Mossbauer Effect Methodology 11 I. J. Gruverman, Ed.; Vol. 10, p. 69, Plenum, New York, 1976.

2. Lam, Y. L.; Garten, R. L. 11 Proceedings, the 6th Ibero-American Sym­posium on Catalysis " Rio de Janeiro, 1978

3. Vannice, M.A.; Lam., Y. L.; Garten, R. L. Advan. Chem. 1979, 178,

15.

4. Garten. R. L., Sinfelt, J. H.J. Cata!.. 1980, 62, 127.

5. Guczi, L. Cata!.. Rev. - Sci. Eng. 1981, 23, 329.

6. Minai, Y.; Fukushima, T.; Ichikawa, M.; Tominaga, T. J. Radioanal.

Nucl. Chem., Lett. 1984, 87, 189.

7. Niemantsverdriet, J. W.; van der Kraan, A. M.; van Loef, J. J.; Del­gass, W. N. J. Phys. Chem. 1983, 87, 1292.

8. Niemantsverdriet, J. W.; Aschenbeck, D. P.; Fortunato, F: A.; Del­gass, W. N. J. Mol. Cata!.. 1984, 25, 285.

Page 92: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Fe3+ in Fe-Pt and Fe-Ru/Si02 page 82

9. Niemantsverdriet, J. W.; van der Kraan , A. M.; Delgass , W. N. J . Catal . 1984, 89, 138.

10. Niemantsverdriet, J. W.; van Kaam, J. A. C.; Flipse, C. F. J.; van der Kraan, A. M. J. Catal. 1985, 96, 58.

11. Niemantsverdriet, J. W.; van der Kraan, A. 1\/1 . Surf. Interface Anal.

1986,9,221.

12. Castner, T.; Newell, G. S.; Holton, W. C. ; Slichter , C. P. J . Phys. Chem. 1960, 32, 668.

13. Wickman, H. H.; Klein, M. P.; Shirley , D. A. J. Phys . Chem. 1965, 42, 2113.

14. Dowsing, R. D.; Gibson, J. F. J . Phys. Chem. 1969, 50, 294.

15. Tops~ , H.; Dumesic, J . A.; M~rup, S. "Applications of Mossbauer Spectroscopy" Cohen , R. L., Ed. , Vol. II , p. 55, Academic Press, New York, 1980.

16. Konings , A. J. A.; van Dooren, A. M.; Koningsberger, D. C.; de Beer, V. H. J .; Farragher , A. L.; Schuit, G. C. A. J. Catal. 1978, 54, 1.

17. Rush , J. D.; Johson, C. E.; Thomas, M. F. J. Phys. Chem. 1976, 6,

2017 .

18. Williams ; J . M.; Pearson, D. I. C. J . Phys. (Paris) 1979, C6, 401.

19. Barholomew, C. H.; Boudart , M. J. Catal. 1973, 29, 278.

Page 93: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 83

Chapter 5

Controlled Oxygen Cherrisorption on:an Alumina Supported Rhodium Catalyst.

Chapter 5

The Fonnation of a Metal-Metal Oxide Interface Determined by EXAFS.

5.1 Introduction

Oxidation of bulk metals is a process that is in general well understood ( 1,2). In contrast, the oxidation of supported metal

catalysts is less well understood. In technological applications, heterogeneous catalysts need to be regenerated several times and

oxidation is an important step during the regeneration process of supported metal catalysts. Therefore, oxidation of small metal par­

ticles is a process that needs to be understood better. One of the techniques which can be used to study the reduction and oxidation behavior of a catalyst is Temperature Programmed Reduction (TPR) and Oxidation (TPO) (3-10). Reduction of metal catalysts is in general a fast process and, hence, T PR is a very sensitive tech­nique in describing the reducibility of supported metal catalysts ( 3-6). Oxidation is in general a slow process, limited by diffusion of oxygen or metal ions through the oxidic skin around the metal par­ticles, after the oxidation process has started ( 1,2,7-10). Conse­quently. the usefulness of TPO is limited.

EXAFS is a technique that has proved to be very adequate in studying supported metal catalysts ( 11,12). Here, we will present the results of an EXAFS study in which we followed the oxidation of small rhodium metal particles (25 ± 5 .A) at temperatures of

Page 94: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ A1 20 3 page 84

100 K up to 300 K. In this temperature regime, oxidation is not

complete. A careful oxidation at room temperature is in general

known as passivation (7-10.13-16). · As a result of the incomplete­

ness of the oxidation, a cherry-like situation develops in which a

small metallic kernel is covered with metal oxide ( 16).

For highly dispersed catalysts, a careful analysis of the EXAFS spectra of the fully reduced sample usually indicates metal-oxygen

. distances to be present in the range of 2.6- 2.8 A ( 12,17-24).

These distances have been ascribed to 0 2- neighbors of Rh0 atoms

in the interface between metal particle and supporting oxide. Two

observations confirm this assignment . The first is that these dis­

tances are always found when oxidic type supports are used (Al20 3 (12,17-21). Ti0 2 (22,23) and even X-zeolite (24)), and regardless of

the metal (Rh (12,17,18,22,23), Pt (19), Ir (20,21), Pd (24)), the

distance is always in the range of 2.6 - 2.8 It Secondly, the relative

number of 0 2- neighbors decreases with increasing rhodium particle

size ( 17 ), indicating that these Rh-0 distances are not related to

the bul.k of the metal particles but to the surface or the metal­

support interface. In this study we will present additional evidence

that indeed metal-to-metal oxide interfaces can be detected with

EXAFS and that in these interfaces metal-to-oxygen distances are

present in the range of 2.6 - 2.8 A.

5.2 Experimental

A 1.9 wt% Rh/ Al 20 3 catalyst was prepared by incipient wet­

ting of the Al 20 3 support (000-1.5E, Ketjen : surface area 180 m 2 g- 1• pore volume 0.65 ml g- 1

) with an aqueous solution of

RhClf3H 20 (Drijfhout) . After impregnation, the catalyst precursor

was carefully dried (24 h at room temperature, 12 h at 395 K, heat­ing rate 5 K min-1

) . The dried catalyst was pressed into a thin self­

supporting wafer with an absorbance (µ.x) of 2.5 at the rhodium

Page 95: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 85 Chapter 5

K-edge. assuring an optimum signal-to-noise ratio in the rhodium

EXAFS spectra . The catalyst wafer was first reduced in an EXAFS in situ cell at 623 K (heating rate 5 K min-1

, 5 ml H2 min- 1) and

after reduction, the catalyst was evacuated at the same tempera­

ture. After this, an EXAFS spectrum was recorded with the sample

at 100 K. At the same temperature, a small amount of oxygen was

admitted and after 10 min a second EXAFS spectrum was recorded .

Thereafter, the sample was allowed to warm to 300 K under oxygen

and after 10 min at 300 K, the catalyst was quickly cooled down to

100 K and a third EXAFS spectrum was recorded. The EXAFS

spectra were recorded at the synchrotron radiation source (SRS) in

Daresbury, U.K. The storage ring was operated at 2.0 GeV, the ring

current was in the range of 100-300 mA.

In order to analyze the EXAFS spectra of the catalyst, EXAFS

spectra of reference compounds are needed. From these, back-

scattering amplitude (F(k)) and phase shift ( <b(k)) functions have

to be extracted. With these functions , we can calculate EXAFS

spectra for the catalyst samples . In the calculations we have four

independent parameters, the coordination number N, the coordina­

tion distance R, the Debye Waller factor /:J,.o 2 which accounts for

disorder and "£o, which allows for a correction on the position of the

adsorption edge. By changing these parameters, EXAFS spectra that fit the measured spectra best are calculated. Rhodium foil was

used as a reference for the Rh0-Rh0 contributions and Rh20 3 as a reference for Rh0-02

- and Rh 3+-0 2- contributions. The thickness of

the rhodium foil was 20 µm (corresponding to an absorbance

µ x = 1.4). For Rh20 3 a wafer with an absorbance of 2.5 was

prepared by mixing and crushing 70 mg of Rh 20 3 with 30 mg of

A1 20 3. The EXAFS spectra of the reference compounds were

recorded with the sample at room temperature.

Page 96: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 86

5.3 Results

The EXAFS functions, x(k), were obtained from the X-ray absorption spectra by subtracting a Victoreen curve, followed by a cubic spline background removal ( 25). Normalization was performed by division to the height of the edge . In Figures 5.la , 5.lb and 5.lc , the raw EXAFS spectra of the catalyst after reduction and evacua­tion. after oxygen admission at 100 K and after warming to 300 K are shown .

For a complete description of the analysis procedure we refer to ( 17 ,23 .26 ,n ). Briefly , the analysis consisted of the following steps which led to a recurrent optimization process. For the reduced and evacuated sample and for the sample oxidized at 100 K, an Rh­Rh EXAFS function was calculated using F( k) and <b< k) obtained from rhodium foil . The parameters N. R, D.a 2 and Eo were chosen to give the best agreement in r-space of the main peaks in a k 3

-

weighted Fourier transform. In Figures 5.ld and 5.le the k 1-

weighted Fourier transforms of the experimental data and the calcu­lated Rh-Rh EXAFS functions are shown. Corrections were made in the Fourier transformation for the k-dependence in backscattering amplitude and phase shift using F(k) and <b< k) of the rhodium foil ( 17). From Figures 5.ld and 5.le it is obvious that apart from a Rh-Rh contribution , other contributions are present. Note that these differences are less obvious in a k3-weighted Fourier transform . The calculated Rh-Rh EXAFS spectrum was subtracted from the meas­ured spectrum . The resulting difference spectrum contained two Rh-0 contribut ions . A two shell Rh-0 EXAFS function (using F( k)

and <f>( k) from Rh20 3) and its Fourier transform were calculated to give the best agreement with the difference spectrum and with the k 1-weighted Fourier transform of the difference spectrum (corrected for the k-dependence in the Rh-0 phase shift). In Figures 5.2a and 5.2b the k 1-weighted Rh-0 corrected Fourier transforms of the difference spectra and the calculated two-shell Rh-0 spectra are shown . Th is calculated two-shell Rh-0 EXA F S function was again

Page 97: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 87 Chapter 5

subtracted from the experimental data , in order to obtain a

difference spectrum with a major Rh- Rh contribution and almost no

Rh-0 contributions. Again a best fitting Rh-Rh EXAFS function

was calculated, subtracted from the experimental data in order to

once more optimize the Rh-0 contributions. In Figure 5.2d and 5.2e

the k 1-weighted Rh-Rh corrected Fourier transforms of the experi­

mental data and the Fourier transforms of the sum of the calculated

Rh- Rh and two shell Rh-0 EXAFS functions are shown . The results

of this analysis procedure are presented in Table 5.1.

For the sample that has been oxidized at 300 K, the procedure

was different. Because in this case the main contribution originated

from oxygen neighbors, a Rhn+ _0 2- EXAFS spectrum using F(k)

and cb(k) from Rh20 3 was calculated first to give the best agree­

ment in the k 1-weighted Rh-0 corrected Fourier transform with the

experimental data. Figure 5.1f shows the k 1-weighted Rh-0

corrected Fourier transform of the experimental data and the calcu-

Figure 5.1 Raw EXAFS data of

(a) Rh/A120 3 after reduction and evacuation at 623 K

(b) Rh/A120 3 after oxygen exposure at 100 K

(c) Rh/Alp3 after oxygen admission at 300 K

Imaginary parts of the Fourier transforms of the original EX­

AFS spectra in Figure 5.1a-5.1c (solid lines) and calculated

Rh-Rh EXAFS spectra (dashed lines). The Fourier transforms

are k 1-weighted and corrected for the Rh-Rh phase shift and

backscattering amplitude. The Fourier transform ranges in k­

space are indicated in brackets.

(d) Rh/Al 20 3 after reduction and evacuation at 623 K (2 .97 -

14.58)

(e) Rh/Al20 3 after oxygen exposure at 100 K (3.26 - 12.03)

(f) Rh/Alp3 after oxygen admission at 300 K (2.55 - 12.45)

In Figure 5.1f the dashed line represents the calculated dom­

inant Rh3+-0 2- contribution, this Fourier transform is k 1

-

weighted and corrected for Rh-0 phase shift .

Page 98: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 88

* 10-2

5

a d 1

0 -0

-5 -1

0 5 10 15 0 3 6 * 10-2

5 1

b :~ e ;:

0 0

-5 -1 0 5 10 15 0 3 6 * 10-2 * 10- 1

5

c 1 f

0 0

-1 - 5 __..._~~--+--~~_.____._~~__,___,

0 5 10 15 0 3 6

k rA. -11 R [A]

Page 99: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 89 Chapter 5

Table 5.1 Final results from EXAFS data analysis

Coordination Distance /)..(J L D

Treatment NN number tAJ (* 10- 3 A- 2i {al (a)

R623-E623 Rh 5.6 0.2 2.635 0.005 3 0 0.6 03 2.055 0.01 1 0 2.0 0.2 2.73 0.01 2.5

R623-E623 Rh 4.1 0.2 2.63 0.005 3.4 -0100 0 1.4 0.2 2.055 0.01 1

0 1.4 0.2 2.77 0.05 2.5

R623-E623 Rh 1.9 0 .3 2.645 0.01 4.4 -0100 0 3.6 0.3 2.03 0 .01 4.7 -0300 0 1.7 0.3 2.76 0.03 6

Nearest l\leighbor 1\11\1 =

R = E =

0=

Reduction in H2 at the temperature indicated

Evacuation at the temperature indicated

Admission of oxygen at the temperature indicated

(al

1 2 1

1 2 1

2 2 3

Estimated overall (experimental +systematic) error

Eu D

(eV) (a)

1.7 1 8.4 2

-4.1 2

7.9 1 -2.5 1 -2.9 1

4 2 -0 .5 2 4 2

(a) b

!).a 2, the De bye Waller factor , is a measure for the disorder and E0

is a correction on the edge position: see ref ( 17) for more details .

lated Rh-0 2- EXAFS function. The remaining difference spectrum

could be fitted with a combination of two contributions, a Rh0-Rh0

and a Rh0-02- EXAFS function. The major contribution in the

difference spectrum originates from oxygen scatterers . Therefore, a k 1-weighted Rh-0 phase corrected Fourier transform rather than a

k 3-weighted Fourier transform (corrected for Rh-Rh phase and

backscattering amplitude) was used to optimize the different contri­

butions in the EXAFS spectrum. Since in this Fourier transform an

incorrect phase shift function is used for the (small) Rh0-Rh0 contri­

bution , the accompanying Rh- Rh peak is shifted and coincides with

the peak originating from the Rh0-0 2- bond of about 2.76 A. The

fact that the peak at the right hand side of the Rh3+-0 2- contribu­

tion in Figure 5.1f is indeed the result of two contributions (Rh0-

Page 100: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al20 3 page 90

Rh0 and Rh0-o2- ). becomes clear from a careful study of Figure

5.2c. In Figure 5.2c. the magnitude of the Fourier transform of the original data (solid line) and of the Fourier transforms of a calcu­lated Rh3+-0 2

- + Rh0-Rh0 EXAFS spectrum (dotted line) and of a calculated Rh3+ -02

- + Rh0-02- EXAFS spectrum (dashed line) are

presented. In the Fourier transform of the former calculated EXAFS function. a strong destructive interference is visible in the region between both peaks, whereas in the latter a slightly constructive interference is observed. In the same region in the Fourier transform of the original data, a slightly destructive interference is present. Therefore, we conclude that the right-hand side peak in Figure 5.2c is the result of the sum of two contributions, namely Rh0-Rh0 and Rh0-0 2

- . In a k 1-weighted Rh-Rh phase and backscattering ampli­tude corrected Fourier transform, all contributions are separated, but , as mentioned above, this Fourier transform is not suitable to optimize the dominant low-Z scatterer contribution. Figure 5.2f shows once again the k1-weighted Rh-0 corrected Fourier transforms of the experimental data and the calculated {Rh-02

- + Rh0-02

- + Rh0- Rh0) EXAFS functions . The results of this analysis

are also presented in Table 5.1.

5.4 Discussion

5.4.1 Rh/ Al20 3 after Reduction and Evacuation

After reduction and evacuation of the catalyst at 623 K the major contribution in the EXAFS spectrum is from rhodium neigh­bors. In the difference spectrum (see Figure 5.2a) two contributions are present which both originate from oxygen neighbors.

Page 101: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 91 Chapter 5

* 10-2

3-~-~~~--~-~

a 1 d

-3 -1 6 0 3 6 0 3

* 10-2 * 10- 1

5 8

b e

- 5 ..l--------'----'--___,__ _ __,_______,__ -8 -'--~~__,_ _ _,__~~~-0 3 6 0 3 6

* 10-2 * 10-2

10

5

·I \

12

c ,_

f

0 .J______.~__._ _ _,__..i=::..::.~~ - 1 2 .l--------'----'--___,__ _ __,_______. _ ___.

0 3 6 0 ·3 6

R [A] R [A]

Page 102: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 92

Figure 5.2 Imaginary parts of the Fourier transforms of the difference spectra (original EXAFS spectra minus th e calculated Rh-Rh EXAFS function s. solid lines) and the Fourier transforms of the calculated two shell Rh-0 EXAFS functions (dashed lines) . The Fourier transforms are k1-weighted and corrected for Rh-0 phase shift. The Fourier transform ranges in k­

space are indicated in brackets.

(a) Rh/A1p3 after reduction and evacuation at 623 K (3.46 -8.33),

(b) Rh/Alp3 after oxygen exposure at 100 K (2 .54 - 8.00)

(c) Magnitude of the Fourier transform of

solid line : Rh/ Alp3 after oxygen admission at 300 K (2.55 -12.45) ,

dotted line : Calculated Rh h- 0 2- + Rh 0-Rh 0 EXAFS function ,

dashed line Calculated Rh 3+-02- + Rh 0-02

- EXAFS func­

tion . (See text for further details)

Fourier transform of the original EXAFS spectra (solid lines) and the Fourier transforms of the calculated best fitting EX­AFS spectra (using the parameters presented in Table 5.1 . The Fourier transforms are k1-weighted and corrected for Rh-Rh phase s hift and backscattering amplitude The Fourier transform ranges ink-space are indicated in brackets .

(d) Rh/A1p3 after reduction and evacuation at 623 K (2 .97 -

14 .58)

(e) Rh/Alp3 after oxygen exposure at 100 K (3.26 - 12.03)

(f) Rh/Alp3 after oxygen admission at 300 K (2 .55 - 12.45)

The rhodium atoms and ions in the sample have on the aver­

age 5.6 rhodium nearest neighbors. The coordination distance of

2.635 A points to zerovalent rhodium atoms having zerovalent rho­

dium neighbors; the contribution originates from rhodium atoms in

metallic particles . From (28) we can estimate that the particles con­

tain about 15 atoms on the average, and are roughly 10 A. in diame­

ter. However. the results underestimate the size of the metal

Page 103: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 93 Chapter 5

particles . For , in the EXAFS spectrum, an additional contribution from oxygen neighbors at 2.055 A is present and this contribution arises from oxidic type Rh-0 2

- bonds . From TPR studies , which are usually carried out in 5% H2 in some inert gas, it is evident that at 623 K reduction in 100% H2 should have been complete ( 17 ). Dur­ing the evacuation procedure at high temperature however, the hydroxylated A120 3 surface looses water and at the high evacuation temperature, the metal particles may possibly be partly reoxidized by H20. Another possibility is that a small leakage of air into the cell has occurred during the evacuation procedure. Anyway , it is obvious that the metal particles were partly oxidized after the reduction treatment, during the evacuation procedure. In bulk Rh20 3 the Rh3+ -0 2

- distance is 2.05 A and each rhodium ion has 6 oxygen neighbors. We found a Rh-0 2

- coordination number of 0.6. The coordination numbers measured with EXAFS are averaged coordina­tion numbers, i.e .. they are averaged over all rhodium atoms and ions present in the sample. The zerovalent rhodium atoms in the metal particles will not have oxygen neighbors at an oxidic type dis­tance. Therefore, the EXAFS oscillations in the X-ray absorption spectrum before the normalization procedure, originating from these rhodium atoms will be proportional to the number of neighbors on metallic Rh-Rh distances The rhodium ions in an oxidic phase in the catalyst will have only oxygen neighbors in their first coordination shell at about 2.05 A. The EXAFS oscillations originating from these rhodium ions will be proportional to the number of oxygen neighbors at 2.05 A. In the normalization procedure, we divide these EXAFS oscillations by the height of the edge. However, both rho­dium metal atoms (in the metal particles) and rhodium ions (in the oxidic phases) will contribute to the height of the edge jump. Therefore, the ampl itude of the oscillations in the normalized EXAFS function which originate from metallic rhodium atoms will be proportional to the number of neighbors at metallic distances divided by the total number of rhodium atoms and ions present in

the sample : A ~e~~l =A ,~e~11a1n metal/ (n meta1+n oxidic ) and therefore. N meas - N real l l I (n +11 ) ·1n wh"1ch A meas and N meas meta l - meta l meta l meta l oxidic • meta l meta l

are the measured amplitude of the EXAFS oscillations and the

Page 104: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 94

measured coordination number. based on these oscillations. A ,~~~11;i 1 and N r';~~ 1;i 1 are the real amplitude and the real coordination number in the EXAFS function . n nwlJI and n oxidi, are the number of rhodium atoms in the metal particles and rhodium ions in the oxidic phase that contribute the the adsorption. Thus , the coordination numbers measured by EXAFS are always fractional coordination numbers . Thus, when we want to determine the real or corrected coordina­tion numbers in order to be able to estimate particle sizes, we will have to correct fot this . Therefore, we need to know the fraction

of rhodium atoms present in the metal particles or the fraction of rhodium ions present in the oxidic phase . Using the following pro­cedure, we can estimate the fraction of rhodium ions in the oxidic phase. In bulk Rh 20 3 each rhodium ion has six oxygen neighbors at 2.05 A. That indicates, that the rhodium ions in the oxidic phase in the catalyst sample will have up to six oxygen neighbors . As a lower limit, we estimate that these rhodium ions will have at least four oxygen neighbors . Thus, we assume that N ~;~i( is 4-6. the

measured N ~~~iz is 0.6. thus. we find that fox= 0.6/6 - 0.6/4. This means that 15 to 10% of the rhodium atoms is oxidized (0.6/4 = 0.15 > fox > 0.6/6 = 0.10, f ox is the fraction of rhodium atoms that has been oxidized) . Because the Rh-Rh coordination number of 5.6 is also an average coordination number . the Rh-Rh coordination number of rhodium atoms ·present in metallic particles

t b t d f II . N real - N rneils / f . h" h mus e correc e as o ows . rnet;il - me.till metill in w IC

f metill = 1 - f ox = 0.85-0.90 As a consequence, the average Rh­Rh coordination number is 6.4 ± 0.2 and the metal particles will consist approximately of 25 atoms, their diameter is about 15 A. Also, although the catalyst had been reduced at high temperature (623 K), after evacuation the sample was partly oxidized , as can be concluded from the presence of the Rh-0 2

- distance of 2.05 A.

One more contribution needs to be discussed , the 2.73 A Rh-0 contribution (cf. Table 5.1). This contribution arises from oxygen ions in the surface of the support, which are neighbors to zerovalent rhodium atoms in the metal-support interface. This type of rhodium-to-oxygen contribution has frequently been reported

Page 105: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 95 Chapter 5

( 12,17-24). Like the Rh-Rh contribution, this Rh0-02- contribution

must be corrected for the presence of rhodium cations. The

corrected coordination number is 2.0/(1 - fox) = 2.3 ± 0.1. How­

ever , this number does not yet represent the number of oxygen

anions in contact with rhodium atoms in the metal-support inzer­face, because for three dimensional particles only p~rt of the rho­

dium atoms will be present in this interface. To be able to deduce

the real number of neighboring oxygen ions for a rhodium atom in

the metal-support interface we need to know the size and the shape

of the metal particles. The size is obtained from the Rh-Rh coordi­

nation number , but no information on the shape is available. Alter­

natively, if we know the real number of oxygen neighbors for a

interfacial rhodium metal atom, we can determine the shape of the

metal particles . In the following we assumed that the rhodium metal

particles rest on a [111] face of the y-Al20 3 support. The size of

the rhodium atoms is about equal to the size of the oxygen ions

(the Rh-Rh distance is 2.635 A, the 0 2--02- distance is 2.80 .A), the

real Rh0-02- coordination number of a rhodium atom in the metal

support interface is 3.0. With this knowledge we can make a model

for the size and shape of the metal particles.

5.4.2 A Model for the Oxidation of Metal Particles

In th is paper we will try to describe the oxidation process of

small metal particles . At low temperatures, oxygen readily adsorbs

and even partially dissociates on rhodium metal (29). Considering

the fact that we study very small metal particles, oxidation may be

a relatively fast process. In the following, we will assume that the

metal atoms in the particle with the least negative binding energy

(i. e. atoms which are the most coordinative unsaturated) will be

oxidized first. We have calculated the cohesive energy of every atom in the metal particle by summation of the Lennard-Jones energy

contributions of all the other atoms in the metal particle. The

Page 106: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 96

Lennard-Jones binding energy is given by

E(R) = Al-1- - Ji_ R12 R6 (5.4.1]

where R is the interatomic rhodium-rhodium distance. The constant A in the expression is not important when one compares binding energies of the same type and has therefore been taken equal to unity. B determines the interatomic distance at which the minimum in the binding energy occurs and has been given the value of 0.005989 A.6. This value corresponds to Rmin = 2.635 A, which is equal to the Rh-Rh distance in the metal particles (see Table 5.1). Thus, the relative binding energy of atom i in the metal particle has been calculated by using

E= L i ;"= j

1 R.12

l j

5.989 10- 3

RiJ (5.4.2]

In most cases, the R i) 12 term is negligible compared to the

Ri)6 term . Therefore, a model in which only the number of nearest rhodium neighbors is taken as a measure for the binding energy of a given atom gives nearly the same result as the model described above, which takes into account the long range interactions as well. In Table 5.2 the calculated binding energies and the number of rho­dium nearest neighbors is given during a hypothetical oxidation pro­cess in which the rhodium atoms, one by one, become 'oxidized ' to Rh . Once a rhodium atom is oxidized, it will preferentially be sur­rounded by 0 2

- ions. Hence, it will not contribute to the EXAFS Rh0-Rh0 coordination number and it will also not contribute to the binding energy of the remaining rhodium atoms in the metal parti­cle. After an atom has been oxidized (i.e., removed from the actual cluster) , the new Rh0-Rh0 coordination number, the binding energy and the number of rhodium neighbors for each rhodium atom in the remaining particle were re-calculated.

Page 107: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 97 Chapter 5

Table 5.2 Lennard-Jones binding energies (E , a u.) and number of rho­dium neighbors (N) for each rhodium metal atom during a hypothetical oxidation process

Atom Number of atoms oxidized : nrs 0 1 2 3 4 5 6 7 8 9

1 -E 4.1 3.2 N 4 3

2, 3 -E 6.0 5.9 5.0 5.0 5.0 5.0 5.0 5.0 5.0 5.0 N 6 6 5 5 5 5 5 5 5 5

4 -E 4.1 4.1 4.0 3.2 N 4 4 4

5 -E 4.0 N 4

6 -E 8.7 7.8 6.9 6.9 6.9 6.9 6.9 6.8 6.0 5.9 N 9 8 7 7 7 7 7 7 6 6

7 -E 9.1 9.1 9.0 9.0 9.0 8.9 8.9 8.8 8.8 8.7 N 9 9 9 9 9 9 9 9 9 9

8 -E 8.7 8.7 8.7 7.8 6.9 6.9 6.9 6.8 6.8 6.0 N 9 9 9 8 7 7 7 7 7 6

9 -E 4.0 4.0 4.0 N 4 4 4

10.13 -E 5.9 5.0 4.9 4.9 4.9 4.0 4.0 4.0 N 6 5 5 5 5 4 4 4

11 ,12 -E 8.8 8.8 8.7 8.8 8.7 7.9 7.8 6.9 6.0 6.0 N 9 9 9 9 9 8 8 7 6 6

14 -E 4.1 4.1 4.1 4.1 4.1 N 4 4 4 4 4

15 -E 5.2 5.2 5.2 5.2 5.2 4.3 3.4 N 5 5 5 5 5 4 3

16 -E 4.1 4.1 4.1 4.1 4.1 4.0 N 4 4 4 4 4 4

17 -E 6.1 5.9 5.0 5.0 5.0 5.0 5.0 5.0 4.9 4.9 N 6 6 5 5 5 5 5 5 5 5

18 -E 7.1 7.1 7.0 7.0 6.9 6.9 6.9 6.9 6.9 6.9 N 7 7 7 7 7 7 7 7 7 7

19 -E 6.1 6.1 6.1 5.9 5.0 5.0 5.0 5.0 5.0 4.9 N 6 6 6 6 5 5 5 5 5 5

20 -E 6.0 5.2 5.0 5.0 5.0 4.9 4.9 4.9 4.0 4.0 N 6 5 5 5 5 5 5 5 4 4

21 ,22 -E 9.0 9.0 8.9 8.9 8.9 8.8 8.8 8.8 8.6 8.6 N 9 9 9 9 9 9 9 9 9 9

23 -E 6.0 6.0 6.0 5.2 5.0 5.0 4.9 4.9 4.9 4.0 N 6 6 6 5 5 5 5 5 5 4

24 ,26 -E 6.1 6.0 6.0 6.0 6.0 5.1 5.1 4.9 4.0 4.0 N 6 6 6 6 6 5 5 5 4 4

25 -E 7.2 7.2 7.2 7.2 7.2 7.0 6.8 6.0 5.9 5.8 N 7 7 7 7 7 7 7 6 6 6

Page 108: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20~ page 98

Figure 5.3 26 Atom rhodium metal particle (fee structured) on a [111] y-Al 20 3 surf ace.

· .. ... · ·· ........ ·· . : ·· ..... · · .... ·· ' .· ··· ... · ·.\

. ,,l,_'{,L,,.I . L,,,.,,\,,L, )., i

........ ::-......... :.-:"' ... :.::-.~.:··.:"' ......

. ::-............... / ....... .

' i .... ·····\'··-·.::7' /.\ ... ,,./,c..::..._,,_/,,-: .. :.,,,/,...,· ... : .. "<:·//.:: ..... \i,...,'·''····\.:// ....... .

·. . ... r .............. · · ........... /\ ./\ ........... .J ..... \ ........... ... .1... .. .

Q Rh atom

........... ) 0 2- in Al20

3 ·· .......... ·

In Figure 5.3, a 26 atom metal particle resting on a [111] y­Al203 face is drawn. The averaged Rh- Rh coordination number is

6.38. The particle is assumed to be grown epitaxially on an alumina

(111] crystal face, which according to (JO) is the most stable crystal

face. In the 26 atom metal particle in Figure 5.3, atom number 5 and atom number 9 have the lowest binding energy. The atoms

with the next lowest binding energy are atoms 1 and 4. With

fox = 0.10 - 0.15, it follows that about 3 rhodium atoms in this

particle were in the oxidic phase. Removing (arbitrarily) atom 5

from the particle induces a new situation in which atom 1 has the

lowest binding energy in the 25 atom metal particle (see Table 5.2, column under '1'). After subsequently removing atom 1. atom 9 has

(again) the lowest binding energy. When atom 9 is removed as well,

the Rh- Rh coordination number averaged over all 23 atoms and 3 ions equals to 5.54 (i. e. , the coordination number as it would have

Page 109: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 99 Chapter 5

been measured with EXAFS), in excellent agreem~nt with the experimentally observed value of 5.6. However , after removing atom 9, atom 4 obtains a very low binding energy (Table 5.2 . column under '3') . We therefore assume that when atom 9 is oxidized, atom number 4 will immediately be oxidized as well . As a consequence, the situation after reduction and evacuation will be represented by an average of two situations, one in which two atoms (atom 1 and 5 or 9 and 4) and one in which four atoms (atoms 1, 5, 9 and 4) are oxidized . The situation in which three atoms are oxidized is unlikely to occur. In Figure 5.4a and 5.4b, a bare 24 and 22 atom metal par­ticle are shown. For these metal particles , the calculated Rh0-0 2

-

coordination number is on the average equal to 1.5 (1.38 and 1.62 respectively; note that these are already average coordination numbers), which is too low compared to the measured value of 2.0. However, up to now we have only considered 0 2

- anions of the sup­port which are in contact with the remaining metallic rhodium atoms . But the rhodium oxide phase which has been formed may well be in contact with the remaining metal particle and this would explain the higher Rh0-0 2

- coordination number. In Figure 5.4c and 5.4d, a possibility is presented in which several oxygen ions take the place of the oxidized rhodium atoms. Using that configuration, the calculated and averaged Rh0-02

- coordination number is 2.0. We have assumed that the oxidation process leads to stoichiometric Rh 20 3 and thus, in Figure 5.4 (as well as in Figures 5.5 and 5.6) three 0 2

- ions are incorporated for every two rhodium atoms which are assumed to be oxidized to Rh3+. The Rh3+ ions are not shown in these Figures. They most probably reside in the octahedral or tetrahedral sites present between the oxygen ions from the support and the oxygen ions in the rhodium oxide phase.

By varying the assumptions, we found that is was impossible to reproduce the measured Rh0-02

- coordination number of 2.0 with situations which are principally different from the situation depicted in Figures 5.4c and 5.4d . In all cases, it was essential that .rhodium oxide was in contact with the metal particle. We therefore assume that Figures 5.4c and 5.4d give a fairly accurate description of the

Page 110: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 100

Figure 5.4 Model of the rhodium metal particles of y-A20 3 after evacua-tion .

(a) Bare 24 atom rhodium metal particle

(b) Bare 22 atom rhodium metal particle

(c) 24 Atom rhodium metal particle in contact with Rhp3

(d) 22 Atom rhodium metal particle in contact with Rhp3

a

\ ..... ,,X., ... > y

/' ;,,.,,. ........ l~: ...... _ f '---~._... ........ __,.........__/ ' '::-.......... ·· ' "····~. y

b !,.... ........ !,.......- \l/ ............ [,· ......... ,... ............ !.( ............ 1./ ........... J •. / ...... .

\ ........... , ,, ......... ........ , ·.' .. ",· -.····. , ... ·.·.··- .··· ·. , "' ~ ,,. ..... r -. ,,,. \ • ., .......... : . .";. ......... ....-'·:~·,.__, ...... ~1:.\ ... , ...... / y

QRh atom

metal particles in the catalyst after reduction and evacuation . As a

consequence, the 26 atom metal particle in Figure 5.3 will give a good representation of the average metal particle in our Rh / A1 20 3 sample directly after reduction.

c

d

Page 111: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 101 Chapter 5

5.4.3 Rh/ Al20 3 after Oxygen Admission at JOO K

After admission of oxygen to the catalyst at 100 K, the

EXAFS spectrum changed drastically (see Figure 5.1). The Rh0-Rh0

contribution drastically diminished and the contribution of the

oxidic type of Rh-0 2- bonds increased in magnitude (cf. Table 5.1).

Clearly, oxidation has occurred.

We continue to describe the oxidation process as indicated in

the foregoing section. Assuming that the rhodium ions in the oxide

phase have 4 to 6 oxygen neighbors, we find that

0.23 < fox < 0.35, indicating that 6 to 9 rhodium atoms have been

oxidized. Thus, the remaining metal particles contain 17 to 20 rho­

dium atoms. Using the same procedure as described above, we find

that successively atoms (5 , 1, 9, 4.) 14. 15 and 16 will be oxidized

(i. e. 'removed' from the metal particle) . The calculated and aver­

aged Rh0-Rh0 coordination number is 4.46, which is higher than the

measured value of 4.1 . In this situation , atoms 10 and 13 both have

a low binding energy . Removing only atom 10 or 13 gives a 19

atom metal particle with a Rh 0-Rh0 coordination number of 4.15, in

very good agreement with the measured value. However. because of

their low binding energy, it is more likely that both atoms 10 and 13

are oxidized at the same time . In that situation, the Rh0- Rh 0 coordi­

nation number is 3.85. Again. we assume that reality is an average

of two or possibly three situations . In all cases, atoms 5, 1, 9, 4, 14, 15 and 16 are oxidized. In addition. (a) no atoms. (b) one atom (10 or 13) or (c) both atoms 10 and 13 may be oxidized . The average

Rh0-Rh0 coordination number will be somewhere between 3.85 and

4.46. With the two situations most likely to occur, situations (a)

and (c), in a one-to-one ratio , the average · Rh0-Rh0 coordination

number would be equal to 4.1, and would be in excellent agreement

with the experimental data.

Page 112: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al20 3 page 102

The calculated Rh0-02- coordination numbers (averaged over

all 26 rhodium atoms and ions) of the 19. 18 and 17 bare atom

metal particles (Figures 5.5a. 5.5b and 5.5c) are 1.04. 0.93 and 0.81

respectively. They are too low compared to the measured value of 1.4 . Figures 5.5d, 5.5e and 5.5f give a representation of a possible

arrangement of rhodium oxide around a remaining 19. 18 and 17 rhodium metal particle after 7, 8 and 9 atoms have been oxidized.

The calculated Rh0-Rh0 coordination number (4 .15. the average over 4.46. 4.15 and 3.85 for the 19, 18 and 17 atom metal particle)

agrees very well with the measured value (4.1), and the calculated and measured Rh0-02

- coordination numbers (1.5 and 1.4 respec­

tively) also agree nicely. Clearly. at least part of the rhodium oxide is in contact with the metal particle. Obviously, the interface

between rhodium metal particle and rhodium oxide is rather stable

under the experimental conditions.

5.4.4 Rh/ Al20 3 after warming up to 300 K under oxygen

After oxygen admission at 100 K, the sample was warmed to 300 K under oxygen atmosphere . The resulting spectrum (Figure

5.1c) was completely different form the two other spectra. The most dominant contribution is now from oxidic Rh-0 2

- bonds and

almost no Rh0-Rh0 contribution is left (see Table 5.1) .

From the measured Rh-0 2- coordination number, we calculate

that 0.60 <fox < 0.90. However, because oxidation has progressed that far, we expect that the number of oxygen neighbors in the rho­

dium oxide phase will be close to 6 and therefore the fraction of oxi­dized rhodium atoms will be close to 0.6. That means that about 16

atoms will be oxidized and about 10 atoms will remain metallic. Our

model calculations indicate that during the oxidation as observed,

atoms number 17, 19. 20, 23, 24, 25 and 26 will be oxidized as well.

The calculated Rh0-Rh0 coordination number of the remaining .

Page 113: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 103

a

b

c

Q Rh atom

Chapter 5

c· y ....• ( .. ( . . _ •.•• ,.···. ./ .";,'-. .. : ...... :.:·: ......

........

{ y / ···~ ....... /.:·:.. .. ·~ ··~·""'/\ ... . ...

) \.,_/}---···+·. (,....:""'(' ,,_.....,._,.--.,.....,,-.__,,..........,

.····( \ ..

. ·._,7-~

..... \

/::·· .. ~:>,<~>< .. /, /·' .. ) .. ) , /.( .. / x

_)

d

e

f

Page 114: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 104

Figure 5.5 Model of the rhodium metal particles of y -Al 20 3 after oxida-tion at 100 K

(a) Bare 19 atom rhodium metal particle

(b) Bare 18 atom rhodium metal particle

(c) Bare 17 atom rhodium metal particle

( d) 19 Atom rhodium metal particle in contact with Rhp 3

( e) 18 Atom rhodium metal particle in contact with Rhp 3

(f) 17 Atom rhodium metal particle in contact with Rhp 3

particle , averaged over all 26 rhodium atoms , is 1.85, is good agree­

ment with the measured value of 1.9. (The Rh0-Rh0 coordination

number averaged only over the 10 atoms in the remaining metallic

kernel of the particle, the corrected coordination number, is 4.8.)

We once more found it imperative that at least part of the rhodium

oxide formed is in contact with the metal particle and compared to the situation in the preceding paragraph, the rhodium metal-to­

rhodium oxide interface has even increased in magnitude. In Figure

5.6a, a bare 10 atom metallic kernel is shown . For this metal parti­

cle, if not covered with rhodium oxide, the Rh0-0 2- coordination

number is 0 .81, which is about a factor two lower than the meas­

ured value (1.7) . Figure 5 .6b represents a plausible arrangement of

rhodium oxide around the 10 atom metallic kernel.

5.4.5 General Remarks

In the discussion above, we used the Rh0-Rh0 coordination

numbers determined with EXAFS to estimate the size of the metal

particles and the Rh-02- coordination numbers to estimate the

amount of oxidized rhodium. The Rh0-02- coordination numbers

were used to estimate the extent of the interface between metal and

oxide. The accuracy of the latter coordination number, however, is lower than that of the others. The reason for this is shortly as fol­

lows. We used a Rh0-Rh0 absorber-scatterer pair (R = 2.687 , the

Page 115: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 105 Chapter 5

Figure 5.6 Model of the rhodium metal particles of y-A1p 3 after oxida­tion at 300 K.

(a) Bare 10 atom rhodium metal particle

(b) 10 Atom rhodium metal particle in contact with Rhp3

a

Q Rh atom

Rh-Rh distance in bulk rhodium) to extract F(k) and (/>(k), and

used these to calculate Rh0-Rh0 EXAFS spectra with R = 2.63-2.645 A, very close to the reference . The Rh-02

- EXAFS functions

(with R = 2.03-2.055 A) were calculated using the absorber­scatterer pair Rh 3+ -0 2

- (R = 2.05 A). The coordination distances in

calculated EXAFS function and reference compound differ only

slightly and, therefore, also this calculation is reliable. However, to

use the same absorber-scatterer pair to calculate Rh0-0 2- EXAFS

spectra with R = 2.7 A is not correct. Mainly because of the longer

distance, the calculated coordination number underestimates the

real coordination numbers . The extent of underestimation may be

as much as 20% ( 17,3 0 . However , this does not affect our main

conclusion that during oxidation a rhodium-rhodium oxide interface is formed. Since the real Rh0-0 2- coordination numbers will be

higher than the values reported in Table 5.1, the metal-oxide inter­faces will even be larger and the coverage by rhodium oxide on the

metal particles will be more complete.

b

Page 116: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ A1 20 3 page 106

In the above discussion it was assumed that energetic con­

siderations played the key role during the oxidation process. How­

ever, as oxidation proceeds , the metal particles become covered with

rhodium oxide. As a consequence, diffusion will sooner or later

become the rate limiting step in the oxidation process. In the situa­

tion described above, we assumed that rhodium atoms with the

lowest binding energy will be oxidized. The binding energy calcula­

tions indicate that metal atoms which become oxidized have a (rela­

tive) binding energy less than 4.1 (or less than 5 rhodium nearest

neighbors) . In the metal particles in Figure 5.3, 5.4 and 5.5, all rho­

dium atoms which have this low binding energy, and which are to

be oxidized in the next step, are exposed to the gas atmosphere and

therefore, a kinetic limitation of the oxidation process is most

unlikely. In Figure 5.6, a 10 atom metal particle is shown. In this

metal particle, 6 rhodium atoms have a binding energy lower than

4.1 (and less than 5 rhodium neighbors). Therefore, these are can­

didates to be oxidized in the next stage. They are, however, (partly)

covered with rhodium oxide and oxidation may therefore be ham­pered by diffusion. The three top rhodium atoms still have a bind­

ing energy of 4.9, far above the limit (4.1) we used in order to

establish whether an atom will be oxidized. Thus, the situation in

Figure 5.6b may represent a stable situation in which oxidation will

proceed only after elevating the temperature . Obviously, up to tem­

peratures of about 300 K. oxidation is not limited by diffusion under

the experimental conditions : the oxidic skin is too small to screen

the metallic kernel from the surrounding gas phase.

In most cases, during oxidation, the rhodium atom that was to

be oxidized in a next stage of oxidation had a binding energy in the range of 4 .0 . In some cases, removing one rhodium atom resulted in

a severe decrease in binding energy of one of its direct nearest rho­

diu m neighbors. In those cases , we assumed that this weakly bound

rhodium atom would be oxidized immediately. We never encoun­

tered situations in which more than two atoms would be oxidized at

the same time. Therefore , we conclude that oxidation has to be a rather smooth and straightforward process.

Page 117: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 107 Chapter 5

Even after oxidation (passivation) at 300 K, some clean metal

surface is still exposed (see Figure 5.6b): the Rh0-0 2- coordination

number was too low to account for complete coverage. The binding

energy of the three rhodium atoms on top is 4 .9 and they are there­

fore rather stable. This may very well explain why passivated metal

particles are very easily reduced : the metallic kernel adsorbs and

dissociates hydrogen even at low temperature and because of the

dissociated (activated) hydrogen, reduction of the neighboring oxide

phase is very fast and can proceed at temperatures far below the

reduction temperature of the bulk oxide.

The rhodium metal particles and rhodium oxide phase have

been grown epitaxially on the Al20 3 [111] crystal face, which is not unlikely for small metal particles. nor for oxide particles. It has been

shown that rhodium oxide grows epitaxially on rhodium [111). According to Castner and Somorjai ( 32) Rh20 3 [0001 J fits on Rh

1111) when the rhodium oxide unit cell expands by about 5%. The

strain induced by this mismatch has only a very small influence on

the cohesive energy calculated by using the Lennard-Jones poten­

tials. Rh20 3 has the same corundum structure as A1 20 3. Epitaxial

growth of rhodium oxide on the alumina support is also not

unlikely. Altogether, the assumptions made in the model described

above, epitaxial growth of rhodium particles on the alumina support

and epitaxial growth of rhodium oxide on both rhodium metal parti­

cles and alumina support, seem justified .

We used only the rhodium atoms in the metal particle in calcu­

lating the Lennard-Jones binding energies of the rhodium atoms:

the influence of neighboring oxygen ions has been neglected. The

additional contribution of oxygen neighbors is expected to be small

and not to influence the results significantly. At the moment we are

studying the binding of rhodium metal particles on Al20 3 surfaces using the AS ED-MO method described by Anderson et al. (33) and

we expect to be able to quantify the contribution of oxygen ions to

the binding energy in the near future .

Page 118: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ A1 20 3 page 108

In the discussion above we described the oxidation of a 26 atom rhodium metal particle. We applied the same procedure to metal particles ranging from 22 to 30 atoms per particle in order to find out whether the observed trends also hold for these particles. In all cases we found no major differences, and the results presented above and the conclusions to be presented below are representative for the cases we have studied.

5.5 Conclusions

Using Rh0-Rh0, Rh0-02- and Rh-02- coordination numbers determined with EXAFS. a model has been derived to describe the oxidation of small rhodium metal particles supported on A1 20 3.

After reduction, the metal particles were about 15 A in diameter and contained about 26 rhodium atoms; after evacuation, the particles were partly oxidized, either by water formed during evacuation or by leaking in of oxygen. and the metal particles contained about 22 to 24 atoms. After oxygen admission at 100 K, oxidation had pro­ceeded and the metal particles contained about 17 to 19 rhodium atoms. After warming the catalyst under oxygen to 300 K, the metal particles contained only about 10 rhodium atoms . The Rh0

-

02- coordination numbers indicated that in all cases the rhodium oxide formed was in contact with the metal particles . The remain­ing metallic kernel was partly covered with rhodium oxide. We therefore conclude, that under the experimental conditions, the rhodium-rhodium oxide interface is more or less stable. Our main conclusion is that metal-support interfaces indeed can be detected with EXAFS and that the distance between interfacial zerovalent rhodium atoms and divalent oxygen atoms is about 2.7 A, in accor­dance with literature data.

Page 119: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 109 Chapter 5

5 .6 References

1. Cabrera, N. "Semiconductor Surface Physics" ; Kingston , R.H ., ed .; J. Wiley & Sons, New York 1960.

2. Hauffe, K. "The Surface Chemistry of Metals and Semiconductors" ;

Gatos , H.C. , Ed .; J. Wiley & Sons , New York 1960.

3. Robertson , S. D.; McNicol , B. D.; de Baas , J. H.; Kloet, S. C. ; Jen­kins, J. W J. Catal . 1975 , 37 , 424.

4. Jenkins , J. W.; McNicol, B. D.; Robertson , S. D. Chem. Tech 1975, 7, 316.

5. Wagstaff, N.; Prins , R. J. Catal . 1979, 59, 435.

6. Wagstaff, N.; Prins , R. J . Catal . 1979, 59 , 445.

7. Vis , J. C.: van 't Blik, H. F. J.; Huizinga, T.; van Grondelle, J.; Prins , R. J. Catal . 1985, 95 , 333

8. van 't Blik , H. F. J.; Prins, R. J . Catal . 1986, 97, 188

9. Martens . J. H. A. ; van 't Blik, H. F. J.; Prins , R. J. Catal . 1986, 97 , 200

10. van 't Blik , H. F. J. ; Koningsberger , D. C.; Prins , R. J. Catal. 1986, 97, 210

11. Sinfelt , J. H.; Via , G. H. ; Lytle , F. W. J. Chem. Phys. 1917, 67 , 3831 .

12. van Zon , J. B. A. D. ; Koningsberger , D. C. ; van 't Blik, H. F. J.; Prins ,

R.; Sayers , D. E. J . Chem. Phys. 1984 , 80 , 3914.

13. Uchijima, T. ; Herrmann , J. M .; Inoue, Y.; Burwell. R. L. Jr.; Butt , J. B.; Cohen, J. B. J. Catal . 1977, 50 , 464.

14. Koboyashi , M .; Inoue . Y ; Takahashi , N.; Burwell , R. ,L_ Jr .; Butt , J. B.; Cohen, J. B. J. Catal . 1980, 64, 74.

15. Nandi, R. K.; Georgopoulos , P ; Cohen , J. B.; Butt , J. B.; Burwell , R.

L. Jr. J . Catal . 1982, 77 , 421 .

16. Huizinga , T ; van Grondelle, J.; Prins , R. A ppl . Catal. 1984, 10, 199

17. van Zon , J. B. A. D.; Koningsberger , D. C. ; van 't Blik, H. F. J. ;

Sayers, D. E. J . Chem. Phys . 1985, 12, 5742.

Page 120: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

0 2 Chemisorption on Rh/ Al 20 3 page 110

18. Koningsberger , D. C. ; van Zon , J . B. A. D.; van 't Blik , H. F. J .; Man­sour, A N.; Visser , G. J .; Prins , R.; Sayers , D. E.; Short, D. R.;

Katzer , J. R. J. Chem. Phys. 1985, 89 . 4075

19. Koningsberger , D. C. ; Sayers , D. E. Solid St . Jon . 1985, 16, 23

20. van Zon. F. B. M .; Visser , G. J .; Koningsberger , D. C. 9th Interna­

tional Congress on Catalysis , Calgary . 1988. to be published

21. Koningsberger, D. C.: Duivenvoorden , F. B. M.; Kip, B. J.; Gates , B. C. "EXAFS and Near Edge Structure"; Lagarde, P.; Raoux , D.; Petiau, J Eds.; Les Editions de Physique , 1986; vol. 1, p. C8-255.

22 . Koningsberger, D. C.; Martens .. J. H. A.; Prins, R.; Short, D. R.;

Sayers , D. E. J. Phys . Chem. 1986, 90 , 3047 .

23. Martens , J. H. A.; Prins , R.; Zandbergen, H.; Koningsberger , D. C.;

accepted for publication in J . Phys . Chem ..

24 . Moller , K .; Bein , T. "EXAFS and Near Edge Structure "; Lagarde, P.; Raoux, D.; Petiau , J. Eds .; Les Editions de Physique , 1986; vol. 1,

p. C8-231.

25. Cook, J. W.; Sayers, D. E. J. Appl. Phys. 1981, 52, 5024 .

26. Duivenvoorden, F. B. M. ; Koningsberger, D. C.; Uh, Y. S.; Gates , B. C. ]. Am. Chem . Soc . 1986, 108, 6254

27 . van 't Blik, H. F. J .; van Zon , J . B. A. D.; Huizinga, T. ; Vis , J. C.;

Koningsberger , D. C. ; Prins , R. J . Am. Chem. Soc. 1985, 107, 3139.

28. Kip, B. J. ; Duivenvoorden , F. B. M .; Koningsberger, D. C. ; Prins , R. J . Catal . 1987, 105 , 26.

29. Matsushima , T. Surf. Sci . 1985, 157, 297

30. Knozinger , H.; Ratnasamy , P. Catal . Rev .-Sci . Eng. 1978 , 17(1), 31.

31. Stern, E. A .; Bunker , B. A.; Heald, S. M . Phys. Rev . B 1980, 21 ,

5521.

32. Castner, D. G. ; Somorjai , G. A. Appl. Surf. Sci. 1980, 6 , 29.

33. Anderson, A. B.; Ravimohan, Ch.; Mehandru, S. P Surf. Sci. 1987,

183, 438.

Page 121: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 111 Chapter 6

Chapter 6

The Structure of the Metal-Support Interface in Rh/ Al20 3 Determined with the ASED-MO Method

6.1 Introduction

In Extended X-ray Absorption Fine Structure ( EXAFS) spectra

of the metal edge of fully reduced supported metal catalysts.

metal-oxygen contributions are present . These contributions have

been ascribed to metal atoms in the metal -support interface having

divalent oxygen ions of the support as neighbors . This assignment

was based on the following observations . Firstly , this contribution

is always observed when oxidic type of supports are used (A1 20 3,

Ti02. X-zeolite) . and regardless of the metal , the corresponding

metal-to-oxygen bond distance is always in the range of 2.6-2 .8 A ( 1-7 ). Secondly , the relative number of oxygen neighbors decreases

with increasing metal particle size ( 2), indicating that these contri­

butions are not related to the bulk of the metal particles. Hence, it

was concluded that the metal-oxygen contributions originated from

the metal-support interface. The metal-oxygen bond length (2.6-

2.8 A, depending on support and metal) was explained assuming

that the metal atoms and oxygen ions could be regarded as hard

spheres with radii equal to the radius of the metal atom and the

radius of divalent oxygen ions respectively , both in the range of

1.3-1 .4 A. Since EXAFS is the only technique that has been able to identify these contributions , confirmation by another, independent

technique seemed desirable. In this study we will present theoreti­cal evidence for these metal-oxygen coordination in the metal­support interface.

Page 122: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Rh/A120 3 characterized with AS ED-MO page 112

6.2 Theoretical Method

We used the atom superpos1t1on and electron delocalization

molecular orbital (ASED-MO) method to study the binding of rho­

dium metal particles to y-Al 20 3. The theory has been outlined in

chapter 2 and has been described extensively in the literature (8,9).

Therefore, we will only give a brief summary of the basics of the

method. The theory is based on a partioning of the electronic charge density function (p) of a molecule, or in this case a cluster,

into rigid free atoms parts (Pi) which center on the nuclei and fol­

low them perfectly, and a non-perfectly following or bond charge

density function (Pnpd :

atoms

L Pi + Pnpf !6.2.l]

The force on nucleus b that is due to Pa and nucleus a is repulsive

because the repulsion component of this force increases more

rapidly than the attractive force of Pa due to the penetration of the

Pb charge cloud. The force on nucleus b due to Pb is zero. Hence,

as the force on nucleus b that is caused by atom a is integrated, a

repulsive energy ER• is obtained. The Pnpf density will cause an

attractive force on nucleus b and thus, its integral is an attractive

energy, Enpf· The binding energy curve is then the sum of two

nonzero components :

E !6.2.2]

However, since Pnpf is not a know function, Enpf cannot be obtained

from an integral of the force it causes on a nucleus . Nevertheless,

the electron delocalization energy can be well approximated by a molecular orbital delocalization energy in most instances (8,9) .

Page 123: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 113 Chapter 6

E ::::::: ER + ~mo [6.2.3]

were b..£ 1110 is the total one-electron molecular orbital 1energy minus the one-electron atomic orbital energies. The appropriate one­electron Hamiltonian is a function of experimentally determined valence state ionization potentials (VSIP), valence Slater atomic orbital overlap integrals and internuclear distances . For aluminium and oxygen, these data have been taken from the Al20 3/Ni and Al 20 3/Pt studies by Anderson et al. ( 10,11 ). For rhodium, these data have been taken from the compilation of Lotz ( J 2), from the work of Basch ( 13) and the work of Moore ( 14). These parameters are listed in Table 6.1. The ionization potential for the rhodium 4d orbital had to be adjusted in order to obtain physically relevant results . The Fermi level for the bare y-Al20 3 cluster was -10.60 eV. For the bare rhodium duster, with the 4d VSIP equal to the value found in the literature, -9.56 eV, the fermi level was -9.24 eV. As a result, when the total energy of the metal cluster plus the Al 20 3 support was calculated, there was a net charge of approxi­mately + 10 on the rhodium metal particle due to the difference in Fermi levels, independent of the distance between metal particle and support cluster. This is not only a physically irrelevant situa­tion, but in addition it did not give rise to a situation in which bonding between metal particle and support cluster occurred. The rhodium 4d VSIP has therefore been taken equal to -10.95 eV. In that case, the metal particle far away from the support cluster had a zero net charge. For more negative values for the rhodium 4d VSIP , the metal cluster obtained a negative net charge of -10 and also in that case, no situations were found in which bonding occurred between rhodium metal particle and Al20 3 support clus ter.

Page 124: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Rh/ Al 20 3 characterized with AS ED-MO page 114

Table 6.1 Parameters used in the calculations principal quantum

Atom

Rh s p

Al s p

0 s

p

number n, Slater exponent ~ (a .u.) , ionization potential IP (eV) and the coefficients for the double~ d orbitals .

s, p d n ' IP n c1 'I C2 , 2 IP

5 2.315 -8.09 5 2.100 -4.57 4 0.5823 4.29 0.6405 1.97 -10.95 ~ 1.521 -12.62 .., 3 1.504 -7.99 2 1.746 26.48 2 1.727 11 .62

6.3 Description of the Model and Results

We used the AS ED-Mo method to calculate the total energy of

a y-Al 20 3 cluster of 54 oxygen atoms , 32 aluminium and 12 hydro­

gen atoms. A 10 atom rhodium metal particle was places on this

cluster. The A1 20 3 cluster has been composed using the structural

data from Lippens (JS) and Knozinger ( 16). The rhodium metal

particle had an fee structure and it consisted of 7 atoms in the

metal support interface and 3 atoms on top. Both clusters are

shown in Figure 6.1a. The rhodium-rhodium distance in the metal

particles has been taken equal to the oxygen-oxygen distance in the

[111) Al 20 3 surface plane (2 .80 A) , which is almost equal to the

rhodium-rhodium distance in bulk rhodium (2 .687 A) . As a result , the metal particle could be fitted epitaxially on the support cluster .

From Figure 6 .1a it can be seen that the interfacial rhodium atoms

may rest on different sites. When we only take the oxygen ions

into account , we can discern one-fold (r on-top), two-fold and

thre~fold coordinated sites . These sites are illustrated in Figure

6.1b. Because of the symmetry, these sites are the same for all

seven interfacial metal atoms. The on-top positions 1 and 5 in Fig­

ure 6.1b are in theory the same. In practice , they differ slightly

Page 125: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 115 Chapter 6

because of the limited dimensions of the support cluster. In addi­

tion, there is one two-fold and there are three-fold coo~dinated sites.

In one of these two three-fold sites, there is an oxygen ion of the

second layer directly beneath the site (position 4 in Figure 6 .1b) , for

the other three-fold site this is not the case (position 2 in Figure

6.1b).

Figure 6.1 Model for the A1p3 support cluster and the rhodium metal

particle

(a) the support cluster and the rhodium particle

(b) a schematic representation of the five sites · 1 = on-top , 2 = 3-fold, 3 = 2-fold , 4 = 3-fold and 5 =on-top.

Rh

1 2 3 4 5

b

Page 126: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Rh/ Al 20 3 characterized with AS ED-M 0 page 116

The total energy of the system has been computed as a f unc­tion of the height of the cluster above the Al 20 3 support cluster . The range in height was chosen such that the Rh-0 bond lengths ranged from 2.0 to 3.0 A. Each curve contained 10 points. The value of the total energy of the system with the particle at 10 A above the cluster was taken as zero energy . In all cases, th is energy was equal to the sum of the total energy of the bare metal particle and the bare support cluster. The min ima in these Lennard-Jones type of curves represent energetically stable situations and the accompany­ing binding energy and Rh-0 bond length could be calculated from these minima. Such curves have been calculated for the metal par­ticle above the five sites shown in Figure 6.1b.

In practice, the surface of the y-Al20 3 support is a hydroxy­lated surface ( 15.16). In fact , the chemical formula for y-Al 20 3 is Al20 3·nH 20 . Therefore, 12 protons have been incorporated in the Al20 3 support cluster. These protons were assumed to be located on top of oxygen ions in the first layer, thus resembling surface hydroxyl groups. The 0 -H bond length has been obtained by vary­ing the 0-H bond length between 0.8 and 1.5 A for the bare support cluster. A sharp minimum was found at a bond length of 0.95 A. There are several possibilities to distribute these 12 protons over the 27 oxygen ions in the surface of the support cluster. We have studied three cases. In situation A, no protons were present under the rhodium metal particle. The protons were situated on the edge of the support cluster . In situation B, the protons were distributed evenly over the support, such that the rhodium atoms in the metal­support interface had both bare oxygen ions and hydroxyl groups as direct neighbors . In situation C, the protons were located on the oxygen ions in the middle of the surface of the support cluster and the rhodium atoms in the metal-support interface had only hydroxyl groups as nearest neighbors.

Page 127: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 117 Chapter 6

Thus, we have calculated total energy curves as a function of height of the rhodium particle above the support cluster for 15 cases, in the three situations A. B and C as mentioned above. and above the 5 sites depicted in Figure 6.1b. The results of these cal­culations are presented in Table 6.2 .

6.4 Discussion

From the results in Table 6.2, it is obvious that the most stable arrangement is the one in which the interfacial rhodium atoms rest in 3-fold sites of surface hydroxyl groups (sites 2 and 4 in situation C) . When we compare this to the metal particle in a 3-fold site in situation A, it is obvious that the interaction between metal particle and support cluster has decreased drastically : the total binding energy decreased from 146.3 to 85 .9 kcal mol-1 {for site 2) . In situation A, the fully dehydroxylated surface, on-top sites are favored. The binding per Rh-0 bond in an on-top site in situation A is very strong : 20.8 kcal moi-1

. This bond strength is almost equal to the sum of the three bond strengths in sites 2 and 4 in situation C, the fully hydroxylated surface : 20.9 and 20.5 kcal mol-1. Obviously, the protons have a pronounced influence on the binding of the metal particle to the support. When the inter­facial rhodium atoms are situated on top of hydroxyl groups, the protons are situated in between the oxygen ions and the rhodium atoms . This situation is clearly very unfavorable : the binding energy of a rhodium atom on top of a bare oxygen ion is approxi­mately 20.8 kcal mol- 1

• on top of a hydroxyl group the binding energy has decreased to 11.3 kcal mol 1

-. Above 3-fold sites, how­ever, the situation is reversed : the binding per rhodium atom and per Rh-0 bond increased when a bare oxygen ion was replace by a hydroxyl group. The increase is such that 3-fold sites on a hydroxylated surface are even preferred above on-top sites of a dehydroxylated surface . Note, that the binding energy per Rh-0

Page 128: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Rh/ A1 20 3 characterized with AS ED-MO page 118

Table 6.2 Final results of the ASED-MO calculations on a 10 atom Rh cluster above a A120 3 support cluster with 54 0 , 32 Al and 12 H atoms .

site 1 site 2 site 3 site 4 site 5 on-top 3-fold 2-fold 3-fold on-top

A Hmin 2.09 1.68 1.66 1.70 2.09 R Rh- 0 2.09 2.34 2.17 2.35 2.10 E 101 -145.3 -85.9 -135.0 -87.1 -142.7 £Rh -20.8 -12.3 -19.3 -12 .4 -20.4 ERh-0 -20.8 -4.1 -9.6 -4.1 -20.4

B Hmin 2.51 1.79 1.85 1.80 2.61 RRh-0 2.51 2.41 2.32 2.42 2.61 E 101 -65.1 -112 .8 -116.7 -115.4 -61.7 £Rh -9.3 -16.1 -16.7 -16.5 -8 .8 ERh- 0 -9.3 ·-5.4 -8.3 -5.5 -8.8

c Hmm 2.72 1.96 2.15 1.99 2.72 RRh-0 2.72 2.54 2.57 2.57 2.72 E101 -78.9 -146.3 -127 .5 -143 .5 -78.5 £Rh -11.3 -20.9 -18.2 -20 .5 -11.2 E Rh-0 -11 .3 -7.0 -9.1 -6.8 -11.2

H min the height of the metal particle above the support cluster in the minimum of the energy-height curve

E,01 The net binding energy in kcal mor1 in the minimum of the energy-height curve (£101 = E,0 ,(H =H min) - E,0 ,(H = 10A))

RRh-0 the rhodium-oxygen bond length

E Rh the binding energy in kcal per mol interfacial rhodium atoms

(=E101/7)

E Rh- 0 the binding energy in kcal per mol rhodium-oxygen bonds (=E 101/7n in which n= 1, 2 or 3, for 1-fold , 2-fold and 3-fold sites respectively)

A All 12 hydrogen atoms on the edge of the Al20 3 cluster : no -OH ions underneath the rhodium metal particle.

B All 12 hydrogen atoms evenly distributed over the Alp3 support cluster

C All 12 hydrogen atoms in the middle of the Al20 3 support cluster : only -OH ions underneath the rhodium metal particle

Page 129: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 119 Chapter 6

bond in an on-top site is still larger th an the binding energy per

Rh-0 bond in a three-fold site. However, since in a three-fold site

three Rh-0 bonds per interfacial rhodium atom are present, the

three-fold sites are favored above the on-top sites.

For supported metal catalysts . metal-oxygen distances in the

metal-support interface ranging from 2.6-2.8 A have been reported.

For Rh/ Al 20 3 catalysts it has been found that the rhodium atoms in the metal-support interface have two to three oxygen neighbors

( 2) . Thus, the results from this AS ED-MO study agree nicely with

the results from the EXAFS studies. We also found that on a fully

dehydroxylated surface on-top sites were preferred and that the

accompanying Rh-0 bond length has decreased to 2.1 A. It is very difficult to study catalysts under fully dehydroxylated conditions

with EXAFS, since the H20 vapor pressure under high vacuum con­

ditions in an EXAFS cell is high enough to sustain a fully hydroxy­

lated y -Al20 3 surface. However, in the EXAFS study on lr/A120 3 ( 17), in which the catalyst had been evacuated at high tempera­

tures, a decrease in lr-0 distance in the metal-support interface was

reported : after the evacuation procedure the lr-0 distance had con­

tracted from 2.6 to 2.15 A. This is of course in very good agree­

ment with the Rh-0 bond length calculated for the meta l particle on a fully dehydroxylated surface.

It is clear that protons play a key role in binding a rhodium

metal particle to the support. What the influence of the protons

exactly is , is a matter that has to be studied more closely . The

intention of this study merely was to show that indeed the Rh-0 distances that have been detected before with EXAFS are

trustworthy. The model which has been used up to now to explain

these Rh-0 distances , which assumed that the rhodium atoms and

oxygen ions can be regarded as hard spheres, however , is shown to be incorrect.

Page 130: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Rh/ Al 20 3 characterized with AS ED-MO page 120

Although the theoretical results agree nicely with the experi­

mental data. their reliability still has to be ascertained. The results may be model-dependent. For instance. the position of the protons

is a matter that may be of importance. They may even be placed inside the Al20 3 support cluster . Or we could leave them out .

Secondly. the rigid structure of the metal particle and the cluster can be questioned. It might be better to optimize the position of

each rhodium atom separately and the position of each proton separately. Unfortunately, such an optimization would be virtually

impossible with this method. In addition to this model-dependence, the method · does not take electrostatic and van der Waals forces

into account . Although the support cluster and the metal particle were neutral and van der Waals forces may be relatively small,

these forces may have· an effect of the final results. Because of these uncertainties, and also because of the semi-empiric nature of

the ASED-MO method, the results presented above should only be used in a semi-qualitative, or better qualitative way . In this con­

text, we think that not too much attention should be paid to small differences in energies and bond lengths. But the dramatic

influence of hydroxylation on the position of the rhodium atoms and on the rhodium-oxygen bond length is expected to be independent of the approximations in the theory.

6.5 Conclusions

Using the ASED-MO theory, we have shown that the binding

of a rhodium metal particle to an y-Al20 3 support cluster is governed by the hydroxylation state of the surface of the support. Energetically, three-fold sites on a fully hydroxylated surface are favored. The accompanying Rh-0 bond length is about 2.5-2.6 A. This is very close to the values reported with EXAFS for such metal-oxygen distances and with the fact that according to EXAFS

rhodium atoms in the metal-support interface have approximately 2

Page 131: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 121 Chapter 6

to 3 oxygen neighbors in Rh/ Al20 3 catalysts. For a fully dehy­

droxylated A120 3 surface. the rhodium atoms in the metal-support interface prefer on-top sites: the accompanying Rh-0 bond length

decreases to 2.1 A..

6 .6 References

1. van Zon, J. B. A . D.; Koningsberger , D. C. ; van 't Blik , H F. J .; Prins,

R.; Sayers , D. E. J. Chem. Phys . 1984, 80, 3914.

2. van Zon, J. B. A. D.; Koningsberger, D. C.; van 't Blik, H. F. J .; Sayers, D. E. J . Chem. Phys. 1985, 12, 5742.

3. Koningsberger, D. C. ; van Zon, J . B A. D. ; van 't Blik, H. F. J.; Man­sour, A . N.; Visser , G. J.; Prins , R. ; Sayers, D. E.; Short , D. R.; Katzer , J. R. J. Chem. Phys . 1985, 89 , 4075

4. Koningsberger , D. C. ; Duivenvoorden , F. B. M .; Kip, B. J .; Gates , B. C. "EXAFS and Near Edge Structure"; Lagarde, P.; Raoux, D.; Petiau, J. Eds.; Les Editions de Physique, 1986; vol. 1, p. C8-255.

5. Koningsberger, D. C.; Martens, J. H. A.; Prins, R.; Short, D. R.;

Sayers , D. E. J. Phys. Chem. 1986, 90, 3047.

6. Martens, J. H. A. ; Prins, R.; Zandbergen, H.; Koningsberger, D. C.;

accepted for publication in J. Phys . Chem ..

7. Moller, K.; Bein, T. "EXAFS and Near Edge Structure "; Lagarde, P.; Raoux, D.; Petiau , J. Eds.; Les Editions de Physique, 1986; vol. 1, p. C8-231.

8. Anderson , A . B. J. Phys. Chem. 1975, 62, 1187

9. Anderson, A . B.; Grimes, R. W.; Hong, S. Y. J. Phys . Chem. 1987, 91 , 4245

10. Anderson , A. B.; Mehandru , 5 . P ; Smialek, J. L. J. Electrochem. Soc .

1985, 132, 1695

11. Anderson, A . B.; Ravimohan , Ch .; Mehandru , S. P Surf. Sci . 1987, 183,438

Page 132: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Rh/A1 20 3 characterized with ASED-MO page 122

12. Lotz . F. W. J. Opt . Soc. Am. 1970, 60 . 206

13. Basch , H. ; Gray , H. B. Th.ear . Chim. Acta 1966. 4 , 367

14. Moore , C. E. Atomic Energy Levels ; NBS Circ . no. 467 ; National

Bureau of Standards, U.S. Government Printing Office: Washington ,

DC, 1958

15. Lippens, B. C. Thesis , Delft, 1961

16. Knozinger, H.; Ratnasamy, P Cata!.. Rev .-Sci. Eng . 1978, 17(1), 31

17. Kampers , F. W. H.; Sayers, D. E.; Koningsberger, D. C. to be pub­

lished

Page 133: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 123

Chapter 7

The Structure of Rh/Ti02 in the Nonnal and the SMSI State

as Determined by EXAFS and H RTEM

7 .1 Introduction

Chapter 7

In heterogeneous metal catalysis the support is used to provide

a large surface area to facilitate the preparation of well dispersed

catalysts and to prevent sintering of the small supported metal par­

ticles, in order to preserve their state of high dispersion. It is often found that support materials modify the chemical reactions of the

metal catalyst. Examples are shape selectivity induced by a zeolitic

support and bif unction al catalysis of metal particles dispersed on an

acidic support. where the metal component catalyzes the

hydrogenation/dehydrogenation reactions_ and the acidic support facilitates isomerization of olefinic compounds. In addition, the sup­

port may have a more direct influence on the chemical properties of

supported metal particles, especially after reduction at high

(> 650 K) temperature. Thus, it is well known that for metals dispersed on certain transition metal oxides, the capacity to adsorb

hydrogen or carbon monoxide drastically diminishes when the

catalyst is reduced at high, rather than low temperatures ( < 650 and usual > 450 K) even though the particle size remains

unchanged ( J-3). Non-transition metal oxides like Al20 3 and Si02 do not influence the capacity to adsorb gasses; the decrease in

adsorption after reduction at high temperature of the metal particles dispersed on these supports can be accounted for by sintering.

Another interesting phenomenon is observed in chemical processes.

Page 134: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 124

When metal particles are dispersed on transition metal oxide sup­

ports, their properties m chemical reactions such as (de)hydrogenation and Fischer-Tropsch synthesis differ markedly

from those traditional oxides, like A1 20 3 and Si02.

A clear distinction between the two classes of support materi­

als can be based on their reducibility. Oxides like Al 20 3 and Si02

are hard to reduce, while transition metal oxides like Ti0 2 and

Ta 20 5 can be reduced to suboxides at moderate temperatures. These suboxides are thought to be responsible for the inability of

the metal particles to adsorb gasses after reduction at high tem­perature. Most of these suboxides are known to have semiconduct­ing properties . In the first reports dealing with this phenomenon ( J-6), these semiconducting suboxides were thought to have a

strong (electronic) influence on the supported metal particles and the phenomenon was labeled SMSI, an acronym for strong metal­

support interaction. Thus, SMSI refers to the state of inability of supported metal particles to adsorb hydrogen and carbon monoxide,

invoked by a reduction at temperatures where the support is known to be at least partially reduced.

Oxidation at mild temperatures (> 450 K) restores both the original oxide and the original properties (i.e. the properties when

reduced at mild temperatures) of the metal particles, demonstrating that suboxide formation and inability to adsorb hydrogen and car­bon monoxide are related ( 1,6-8). Hence. the SMSI state can be

removed in under (slightly) oxidizing conditions at elevated tem­peratures.

Since the first discovery, many studies have been devoted to

SMSI. In addition to the model based on an electronic effect several other explanations for SMSI have been suggested. The most impor­

tant of them is the coverage model (9-17 ). Reduced transition metal oxides can wet metal, in contrast to unreduced oxides . Thus, it has

been suggested that after reduction at high temperature, the subox­ides cover the metal particles and consequently reduce their

Page 135: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 125 Chapter 7

capacity to adsorb gasses.

SMSI has so far been studied mostly by using model systems

like metal films deposited on oxidized titanium (Ti02), or TiOx

deposited on a metal. The techniques frequently used in these stu­

dies are 'surface sensitive', such as Auger and XPS, and in the

majority of cases these studies report coverage . These techniques ,

however, are not truly surface sensitive. In XPS even up to five

layers of the sample can contribute to the spectrum and the results

of sputtering can be questioned because of the destructive character

of the technique. Kelley et al. ( 18), using EELS and other surface

sensitive techniques, did not report complete coverage. Their results

pointed rather to 'electron structural changes' in both metal and support .

Recently, evidence for alloy formation under SMSl-like condi­

tions has been reported by Beard and Ross ( 19). In the alloy forma­

tion model it is assumed that part of the Ti0 2 supporting oxide is

reduced to metallic Ti and forms an alloy with the supported metal

particles. For each noble metal (M) at least three stable titanium

alloys are known : M3 Ti, MTi and MTi 3. It has been shown that

alloying can reduce the hydrogen and carbon monoxide adsorption

capacity as well (20) and alloying may therefore be another plausi­

ble explanation for SMSI.

Up to now, no hard evidence in favor of any of the models to

explain SMSI in real catalysts has been presented in literature. The

model of covered metal particles, though, has been accepted most

widely. In addition, suggestions have been made that coverage alone

cannot explain the anomalous properties of the supported metal

particles under SMSI conditions, and it has been said that an elec­

tronic influence of the covering oxide on neighboring (bare) metal

sites might also play an important role ( 10-13,16,17).

Page 136: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 126

EXAFS has proven to be an excellent tool to investigate the

local environment around metal atoms in a supported metal catalyst

(21,22) . Since only metal atoms in the metal-support interface will

be sensitive to changes in the supporting oxide and because only

surface metal atoms will be sensitive to (changes in) coverage, it is

evident that highly dispersed catalysts must be used . Furthermore,

since the contribution to EXAFS spectra of the low-Z atoms of the

support (0 2- and Ti4+) will be low, high quality data are a prere­

quisite as well. In a preceding study (23), we presented the results

of an EXAFS study of the structure of the rhodium metal particles

in a Rh/Ti0 2 catalyst. This highly dispersed 2.85 wt% Rh/Ti0 2 catalyst was studied after reduction at low temperature and high

temperature, the latter leading to the SMSI state. After reduction at

low temperature, the Rh-Rh coordination number was 3.2, proving

that the metal particles were very small indeed. The rhodium atoms

in the metal-support interface had oxygen neighbors at 2.75 A. These oxygen neighbors originated from the support. From this it

was concluded that the metal particles rested on a (001] anatase

crystal face. When reduced at higher temperature, in the EXAFS

spectrum a 3.4 A Rh- Ti contribution could be detected, indicating

that the (001] anatase crystal face was reduced as well. Since the

Rh0-02- coordination number hardly changed upon reduction at

higher temperature and no other type of oxygen neighbors could be

detected, we concluded that with EXAFS no evidence for coverage

was found.

In the following, we will present the results of a consecutive EXAFS study of a highly dispersed 4 wt% Rh/Ti0 2 catalyst in the

'normal' and the 'SMSI' state . Because of the higher metal loading,

a better signal-to-noise ratio could be realized. High Resolution

Transmission Electron Microscopy ( H RT EM) has been used to ver­

ify the average metal particle size determined by EXAFS. Since in

our earlier EXAFS study we found no evidence for coverage, we

have complemented these experiments by investigating the the

catalyst after reduction at higher temperature (723 K compared to

673 K in the previous study) and we have studied the influence of

Page 137: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 127 Chapter 7

oxygen on the metal particles in the normal and the SMSI state.

After reduction at 723 K and evacuation at 623 K, the Rh/Ti02

catalyst was exposed to oxygen at liquid nitrogen temperature and

at room temperature and after each exposure an EXAFS spectrum was recorded. The same experiments were carried out on a

Rh/ Al 20 3 catalyst in order to study oxygen adsorption on 'normal' rhodium metal particles .

7 .2 Experimental

7 .2.1 Catalyst Preparation

In order to obtain highly dispersed catalysts, a high surface

area support is imperative. Since surface areas of commercially available Ti02 are low, in the range of 10 - 50 m 2g-1

, we have

prepared our own Ti02 support according to the following pro­

cedure. A solution of 8 ml of Ti(OC3H7) 4 in 200 ml ethanol was

added slowly, dropwise, to 4 I of a 1:1 well stirred mixture of ice

and distilled water. The precipitated Ti(OH) 4 was filtered off,

washed with distilled water and dried for 24 h at room temperature, 1 h at 363 K (heating rate 1 K min-1

) and finally 12 h at 393

(heating rate 1 K min-1) . The sample was powdered and calcined

for 3 hat 923 K (heating rate 5 K min-1) . The uncai"cined sample

had a surface area of 700 m2g-1. After calcination the surface area

had decreased to 130 m2g-1 and the pore volume was 0.65 ml g-1.

Calcining for longer than 3 hours at 923 K did not affect surface

area, nor pore volume.

A 4 wt% Rh/Ti02 catalyst was prepared by adding 5 ml of an

aqueous solution of Rh(N03h6H 20 (90 mg mr1) to 3 g of Ti0 2.

After 48 hours , the Ti02 with adsorbed Rh3+ was filtered off,

washed, filtered off and dried as described above for the Ti(OH) 4

Page 138: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 128

prec1p1tate (heating rate 5 K min-1). The dried catalyst was cal­

cined at 623 K for 3 h (heating rate 5 K min-1). This sample was

used as starting material for further experimentation .

The H/Rh value as determined by hydrogen chemisorption for

the calcined sample after reduction at 525 K. was found to be 1.2. Recently we published an empirical calibration of hydrogen chem­

isorption with EXAFS results for several supported rhodium. plati­

num and iridium catalysts (24). According to this calibration a

H/Rh value of 1.2 corresponds to a EXAFS Rh-Rh coordination

number in the range of 7 to 8. For half-spherical particles , this

coordination number corresponds to particles containing roughly 40 to 60 atoms and thus to rather large metal particles . However. as

will be discussed , our · EXAFS results and High Resolution

Transmission Electron Microscopy prove that the metal particles

are very small with five o.r six metal atoms per particle and a parti­

cle size of about 7 A. Two explanations for this contradiction are

possible. In the first place, the rhodium metal particles in the

present sample are much smaller than the (rhodium) metal particles in the most highly dispersed catalyst in ( 24). Their chemistry ,

therefore, may deviate from the behavior of the larger metal parti­

cles in hydrogen atmosphere as described in ( 24). Because of their

size, the small rhodium metal particles approach the quantum size limit and consequently may not be able to adsorb as much hydro­

gen per surface metal atom as a larger (fully metallic) particle.

Another plausible explanation is that during reduction at 525 K the

formation of TiOx suboxides has already started and that the influence of SMSI is reflected in a low adsorption capacity of these

very small rhodium particles. After reduction at 723 K, the H/M value decreased to 0.2, indicating that the catalyst is in the SMSI

state.

In temperature programmed reduction, a sharp peak at 330 K

was observed for the calcined catalysts. After oxidation at 773 K, the temperature programmed reduction profile contains one single

and sharp peak at 350 K. According to Vis et al. (25), a TPR

Page 139: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 129 Chapter 7

reduction peak at 330 - 350 K corresponds to highly dispersed

metal particles.

1 .2.2 EXAFS Measurements

EXAFS spectra were recorded at the synchrotron radiation

source (SRS) in Daresbury, United Kingdom. The storage ring was

operated at 1.8 or 2.0 GeV, the ring current was in the range of 100 - 300 mA. The samples were pressed into thin self supporting

wafers. The thickness of the wafers was chosen to give an absor­bance (µ.x) of 2.5, assuring an optimum signal-to-noise ratio. The

pressed samples were mounted in an in situ EXAFS cell, enabling in situ treatments and measurements in different gas atmospheres.

The EXAFS spectra of the rhodium K-edge were recorded with the sample at approximately 100 K. EXAFS spectra of the reference

compounds were recorded at room temperature.

The experiments on the catalyst were carried out in two series. In all cases the heating rate was 5 K min -1

. After each pretreat­ment an EXAFS spectrum was recorded. The samples were cooled

using liquid nitrogen, the temperature of the samples was approxi­mately 100 K. In the first series, the catalyst was reduced at 473 K for 0.5 h and subsequently at 723 K for 1 h. In the second series, a fresh (calcined) sample was reduced at 723 K for 1 h and subse­

quently evacuated at 623 K for 2 h. After recording the EXAFS

spectrum, a small amount of 0 2 was admitted to the evacuated

sample at 100 K. After recording an EXAFS spectrum the sample was allowed to warm up to room temperature and after 10 min at

room temperature an EXAFS spectrum was recorded at liquid nitro­gen temperature. The same experiments have been carried out with

a Rh/A1 20 3 catalyst after reduction at 623 K.

Page 140: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 130

Phase shifts and backscattering amplitudes from reference

compounds were used to calculate EXAFS spectra and to correct in

the Fourier transformations for the k-dependence in both phase

shifts and backscattering functions . Rhodium foil was used as a

reference for Rh-Rh contributions , Rh20 3 for Rh-0, Rh Ti alloy for

Rh-Ti contributions in the EXAFS spectra . The Rh Ti alloy was

prepared by arc melting equimolar amounts of rhodium and

titanium. The structure and homogeneity were checked by X-ray

diffraction and micro probe analysis. After careful powdering and

sieving, 43 mg of the alloy was mixed and crushed with 32 mg of

Al 20 3 and pr~ssed into a self supporting wafer with absorbance

µx =2.5. For Rh 20 3, a supporting wafer with an absorbance of 2.5

was prepared in the same way (70 mg Rh20 3 and 30 mg Al 20 3). A

rhodium foil was chosen with a thickness of 20 µm thick, µx =1.4.

7.2.3 HRTEM Experiments

Before the HRTEM experiments, the 4 wt% Rh/Ti0 2 catalyst

was reduced at 773 K in a 4% H2 in Ar mixture (heating rate

5 K min-1). After the reduction treatment the catalyst was care­

fully passivated at room temperature in a 4% 0 2 in He mixture . In

order to study the passivated sample in the electron microscope. the

catalyst was suspended in methanol. A droplet of this suspension

was put on a carbon coated Formvar holey copper grid. For the

H RTEM recordings , an objective aperture of 7 nm -l was used . The

photographs were taken with a magnification of 5.105 and an expo­sure time of 1 sec.

An electron microscope image is formed from two contribu­tions, the absorption contrast and the phase contrast. Absorption

contrast originates primarily from an intensity deficiency caused by

the exclusion of a number of electrons with scattering angles larger

than the aperture used. Secondly, inelastically scattered electrons

will have wave lengths different from the incident and elastically

Page 141: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 131 Chapter 7

scattered electrons and will contribute only non-constructively to

the image. Phase contrast arises from constructive interference of diffracted and undiffracted electrons within the aperture used .

When low scattering and thin supports are used. metal parti­cles can be detected best at focus values very close to zero. Since at

this focus phase contrast is almost zero, absorption contrast, due to the higher electron scattering amplitude of the metal particles, is

most pronounced . However, .near zero focus, the resolution of the absorption contrast is approximately twice the point resolution of

the microscope and the particle size determinations will have about the same uncertainty . Since the point resolution of the microscope

used is approximately 2.4 A, the uncertainty in the particle sizes is about 5 A, provided no image calculations are carried out. The best

resolution is obtained near Scherzer focus, where, for thin specimen, phase contrast dominates the absorption contrast. Here, the uncer­

tainty is about the point resolution of the microscope. Compared to zero focus, where absorption contrast dominates, the visibility of

the metal particles is worse at Scherzer focus. We have taken a number of photographs at different focus. Such a through-focus

series can help to identify the presence of metal particles (at zero focus) and to determine their size (at Scherzer focus). Therefore, the uncertainty in the metal particle size determined in this way 1s about the point resolution of the microscope, which is 2.4 A.

In case of very small metal particles, it is difficult to distin­guish them from artefacts. There is also the possibii"ity that very

small particles are overlooked. Partial in situ sintering of the metal particles with a very intense electron beam can help in this respect.

By comparing the total volumes of the metal particles before and after sintering, an estimation can be obtained of the metal particle

size before sintering. The in situ sintering experiment has been per­formed by taking out the condensor aperture and focusing the elec­

tron beam on the agglomerate of Ti02 crystallites and rhodium par­ticles. The degree of sintering can be regulated by the focus of the

condensor lens .

Page 142: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 132

7.3 Results

7.3.1 Analysis of the EXAFS spectra

The EXAFS functions (x(k)) were obtained from the X-ray

absorption spectra by subtracting a Victoreen curve, followed by a

cubic spline background removal (26). Normalization was performed by division to the height of the edge. In Figure 7.1, the raw EXAFS

functions of the Rh/Ti0 2 catalyst after reduction at 473 and at 723 K and the raw EXAFS functions of the Rh/ Al 20 3 and Rh/Ti02 catalysts after evacuation at 623 K and after 0 2 admission at 100 and 300 K are shown .

The spectra of the reference compounds, used to .obtain phase

shifts and backscattering functions, were processed in the same way as the catalyst samples. To obtain phase sh if ts and back­scattering amplitudes, the EXAFS spectra of the reference com­

pounds were Fourier transformed over the largest possible range in

k-space. To avoid cut-off effects, kmin and kmax were chosen in nodes of the EXAFS function . Table 7.1 presents the Fourier

transform ranges and the crystallographic data for the reference

compounds (27 ). In the Fourier transforms of the EXAFS functions of the rhodium foil and Rh 20 3, the Rh-Rh and Rh-0 peaks are clearly separated from higher coordination shells. An inverse transformation over a limited range in r-space gave the EXAFS

spectra for the single shell Rh-Rh and Rh 3+-02- absorber-scatterer

pairs . Since the Fourier transforms have not been corrected for the

k-dependence in phase shift and backscattering amplitude, the transforms contain side lobes. These side lobes were included in the

inverse transformation range. The required phase and backscatter­ing amplitude have been derived from these spectra .

Page 143: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 133 Chapter 7

* 10-2 * 10-2

3 5

a b f--j

::r:: -0 0 CJ

-3 -5 0 5 10 0 5 10 * 10-2 * 10-2

5 5

c d

~ ::r:: CJ

0 0

-5 -5 0 5 10 15 0 5 10

* 10"'"2 * 10-2

5 5

e f f--j

::r:: 0 0 CJ

-5 -5 0 5 10 15 0 5 10 * 10-2 * 10-2

5 5

g h f--j

::r:: 0 0 CJ

-5+-'-~~-+-"~~-t--'~~~ -5-+-'~~~,__._~~_.____.~~

0 5 0 10 15 0 5 0 10 k [A- 1] . k [A- 1]

Page 144: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 134

Figure 7 .1 Raw EXAFS data of

(a) Rh/Ti02 after reduction at 473 K

(b) Rh/Ti02 after reduction at 723 K

(c) Rh/A1p3 after reduction and evacuation at 623 K

(d) Rh/Ti02 after reduction at 723 K and evacuation at 623 K

(e) Rh/A1p3 after oxygen exposure at 100 K

(f) Rh/Ti02 after oxygen admission at 100 K

(g) Rh/A1p3 after oxygen admission at 300 K

(h) Rh/Ti02 after oxygen exposure at 300 K

Table 7 .1 Crystallographic data and Fourier transform ranges for the reference compounds

Rb Fourier transformation

a

b

(

d

Compound NNa

Rh foil Rh

Rhp3 0

RhTi Rh Ti

Nearest Neighbor

Coordination Distance (A)

Coordination Number

NC

2.687 12

2.05 6

2.949 4 2.676 8

Weighting factor in Fourier transformation

nd k-range

3 2.16 - 24.0

1 2.35 - 20.0

3 2.83 - 16.6 1 3.00 - 15.0

e On these data , no direct inverse transformation has been applied .

Crystallographic data derived from (27)

r-range

1.42 - 3.00

0.00 - 2.10 e

1.06 - 2.59

For the Rh T i alloy this procedure was more complex. In the Fourier transform, two contributions are present . The first contribu­tion is from 8 titanium neighbors at 2.676 A and the second from 4

Page 145: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 135 Chapter 7

rhodium neighbors at 2.949 A. (27 ). Since these contributions over­

lap, they could not be separated by an inverse Fourier transform over a window in r-space . A Rh-Rh contribution was calculated

using the phase shift and backscattering amplitude obtained from the rhodium foil. The best agreement in r -space in the region of the

main Rh-Rh peak in the k 3-weighted Fourier transform of the meas­ured and calculated C:XAFS function was obtained with the follow­

ing Rh-Rh parameters N = 4.0 , R = 2.95. !:::.a 2 = 0.0042 and !:::.£ 0 = -1. (N is the coordination number, R the coordination dis­

tance, !:::.a 2 the De bye Waller factor , a measure for the disorder and !:::.E 0 is a correction on the edge position; see ref ( 28) for more

details.) This calculated Rh-Rh EXAFS was subtracted from the

experimental data and the difference spectrum was used to obtain the Rh- Ti phase shift and backscattering amplitude. The Fourier transform of the difference spectrum showed one single peak at

2.16 A.. The Fourier transform has not been corrected for the k­

dependence in phase shift and backscattering amplitude . Because of

the phase shift, the main peak in the Fourier transform has shifted

from the real coordination distance. Inverse transformation over the r-range 1.06-2.59 resulted in the Rh- Ti EXAFS function , from which the Rh-Ti phase shift and backscattering amplitude could be

obtained.

The spectra of the different catalyst samples as presented in Figure 7.1 may contain several contributions. Because of the higher

backscattering amplitude of the high-Z elements, the contribution from rhodium neighbors will be dominant. Other contributions are to be expected from oxygen and possibly from titanium neighbors.

The information of the low-Z elements such as oxygen and titanium is limited in k-space to about k = 7 or 8 A.- 1

. For high-Z elements such as rhodium, the information extends, depending on coordina­

tion number, up to k = 11 or 13 ;...-1 and for high coordination numbers even up to k = 15 f...-: 1

. A k 3-weighted Fourier transform emphasizes the high k part of the EXAFS spectrum , therefore

strongly enhances the high-Z element information relative to the low-Z scatterer information and has thus been used to separate

Page 146: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 136

high- and low-Z scatterer contributions .

For a detailed description of the data analysis procedure, we

refer to (28-30). Briefly, the analysis consisted of the following

steps . An Rh-Rh EXAFS function was calculated. The parameters

N, R, /),.a 2 and f),.£0 were chosen to give the best agreement in r­space with the main peaks in the k 3-weighted Fourier transform of

the calculated and the measured EXAFS function. This calculated

Rh-Rh EXAFS spectrum was subtracted from the measured spec­

trum. In most cases, the resulting difference spectrum contained up

to 3 or 4 different contributions. Since it was impossible to further

separate these contributions, a 2-, 3- or 4-shell EXAFS spectrum

was calculated and optimized to model the difference spectrum in

k-space as well as the k 1-weighted Fourier transform in r-space (the

Fourier transform was corrected for Rh-0 phase shift obtained from

Rh20 3). In order to calculate Rh-0 and Rh-Ti EXAFS spectra,

phase shifts and backscattering amplitudes obtained from Rh20 3

and Rh Ti were used. For each contribution, a separate spectrum

was calculated. These calculated spectra were added to give the

resulting 2-. 3- or 4- shell Rh-0.Ti EXAFS spectrum.

The resulting calculated Rh-0. Ti EXAF S function was sub­

tracted from the original spectrum, in order to start an optimization

cycle. Since this new difference spectrum contained mostly Rh-Rh

information, both the k 1- and k 3-weighted Fourier transforms and

the data in k-space could be used to calculate the best fitting Rh­

Rh EXAFS. This improved Rh-Rh EXAFS spectrum was subtracted

from the original data and the resulting difference spectrum was

analyzed to further optimize the Rh-0 and (if present) Rh-Ti

parameters . In this way a cyclic optimization process was started.

The procedure of subtracting a calculated spectrum to separate the

information of the low-Z scatterers from the high-Z scatterer was

followed until the parameters N, R, /),.a 2 and /),.£ 0 for each contri­

bution became constant. Figure 7.2 shows the k 3-weighted Rh-Rh

corrected Fourier transform of the original data and the calculated Rh-Rh EXAFS for all samples are shown. The observed differences

Page 147: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

1--Lt_

I-­Lt_

1--LL

page 137 Chapter 7

a b

-12 -30 0 3 6 0 3 6

90 30

c d

- 0 -0

-9 0 +-----'----'----4----'-----'----1 - 3 0 -l----L---'-------1-- --'----'-----l

0 3 6 0 3 6

e f

-35 -30 0 3 6 0 3 · 6 * 10-2

12 25

g h

0 0

-12+-----'-- --'-----4----'-----'----I -25 .l-----'-- --'---=-------<'-----'---'---..__-I 0 30 6 0 30 6

R ~] R ~]

Page 148: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 138

Figure 7 .2 Imaginary parts of the Fourier transforms of the original EX­

AFS spectra (solid lines) and calculated Rh-Rh EXAFS spec­

tra (dashed lines). The Fourier transforms are k3-weighted

and corrected for the Rh-Rh phase shift and backscattering

amplitude. The Fourier transform ranges in k-space are indi­

cated in brackets.

(a) Rh/Ti02 after reduction at 473 K (3.33 - 10.24 .&.-1)

(b) Rh/Ti02 after reduction at 723 K (2 .82 - 10.13 .&.-1)

(c) Rh/A1p3 after reduction and evacuation at 623 K (2.97 -14.58 A- 1

)

(d) Rh/Ti02 after reduction at 723 K and evacuation at 623 K

(i.85 - 10.06 ,&.-1)

(e) Rh/A1p3 after oxygen exposure at 100 K (3.26 - 12.03 .&.-1)

(f) Rh/Ti02 after .oxygen admission at 100 K (2.84 - 10.06 .&.-1)

(g) Rh/Alp3 after oxygen admission at 300 K (2.55 - 12.45 .&.-1)

In this case, the dashed line represents the dominant Rh3+-

02- contribution, the Fourier transform is k 1-weighted and

corrected for Rh-0 phase shift.

(h) Rh/Ti02 after oxygen exposure at 300 K (3.22 - 10.18 .&.-1)

are due to low-Z scatterer contributions present in the original data.

Figure 7.3 shows the k 1-weighted Rh-0 corrected Fourier transform of the final difference spectrum (original data minus calculated Rh­

Rh EXAFS, which contains oxygen and. if present, titanium contri­

butions) and the calculated Rh-0. Ti EXAFS functions for all sam­

ples . Finally, Figure 7.4 shows the kt-weighted Rh-Rh corrected

Fourier transform of the original EXAFS function and the EXAFS

function obtained by adding the calculated Rh-Rh and Rh-0.Ti EXAFS functions. In a kt -weighted Fourier transform the low-Z

scatterer is much more pronounced than in a k3-weighted Fourier

transform (as in Figure 7.2). From Figures 7.3 and 7.4 it is evident

that the calculated spectra reproduce the measured spectra extremely well, which emphasizes the reliability of the results.

Page 149: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 139 Chapter 7

* 10-2 *10-2

4 4

a b

I-ll_ 0 0

-4 -4 0 3 6 0 3 6

* 10-2 * 10-2

3 3

c d

I--:. -0 -0 ll_

-3 -3 0 3 6 0 3 6

* 10-2 * 10-2

5 3

e f

I-ll_ 0 0

-5 -3 0 3 6 0 3 6

* 10-2 * 10-2

7

10 g h . . I-

..... ,\

ll_ 0 5

-7-1------L..~-'--+~-'----'~~

3 0

R [A] 6 0 3 0

R [A] 6

Page 150: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 140

Figure 7.3 Imaginary parts of the Fourier transforms of the difference spectra (original EXAFS spectra minus the calculated Rh-Rh EXAFS functions , solid lines) and the Fourier transforms of the calculated Rh-0.Ti EXAFS functions (dashed lines) . The Fourier transforms are k1-weighted and corrected for Rh-0 phase shift. The Fourier transform ranges in k-space are indi­cated in brackets.

(a) Rh/Ti0 2 after reduction at 473 K (2.56 - 8.73 .A- 1)

(b) Rh/Ti02 after reduction at 723 K (2.95 - 8.39 .A-1)

(c) Rh/ A1p3 after reduction and evacuation at 623 K (3.46 -8.33 ,A-1)

(d) Rh/Ti02 after reduction at 723 K and evacuation at 623 K (2.94 - 8.48 A- 1

)

(e) Rh/Alp3 after.oxygen exposure at 100 K (2 .54 - 8.00 A- 1)

(f) Rh/Ti02 after oxygen admission at 100 K (2.97 - 8.53 A- 1)

(g) solid line : Rh/ Alp3 after oxygen admission at 300 K (2.55 -12.45 A- 1

)

and Fourier transforms of two calculated EXAFS functions :

dotted line : Rh 3+-02- + Rh0-Rh0

,

dashed line : Rh 3+ -02- + Rh0-02

-.

(See text for further details)

(h) Rh/Ti02 after oxygen exposure at 300 K (2 .87 - 8.92 A- 1)

The structural parameters obtained in this way are presented

in Table 7.2. The contributions of the Rh-0 and Rh-Ti absorber­

scatterer pairs in the difference spectra are sometimes small and

could be due to artefacts induced by an incorrectly calculated Rh-Rh

EXAFS. In order to ensure that these contributions are indeed real,

and to ensure that the set of parameters we obtained is indeed the

solution that led to the best fit with the experimental data, for all

Rh/Ti0 2 catalyst samples Rh-Rh EXAFS spectra were calculated

for which N and 1::3.a 2, or R and 1::3.E 0, were varied over a large range.

1::3. a 2 was varied concurrently with N so as to give a constant mag­nitude in the k 3-weighted Fourier transform, while 1::3.£ 0 was varied

concurrently with R in order to prevent the main peak in the Fourier

Page 151: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 141 Chapter 7

* 10-1 * 10-1

3 7

a b

I- ,, LL -0 ., -0

·:

-3 -7 0 3 6 0 3 6 * 10-1 * 10-1

14 7

c d

i- · LL -0

-14 -7 0 3 6 0 3 6 * 10-1 * 10-1

8 7

e f

I-LL 0 -0

-8 -7 0 3 6 0 3 . 6 * 10-2 * 10-1

12 7

g h

I-LL 0 0

-12 -+--~-~--~~----< - 7 +--~-~--r-~~---t 0 3 0 6 0

R [A] 3

R [A] 6

Page 152: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti02 page 142

Figure 7.4 Fourier transform of the original EXAFS spectra (solid lines) and the Fourier transforms of the calculated best fitting EX­AFS spectra (using the parameters presented in Table 7.2. The Fourier transforms are k1-weighted and corrected for Rh-Rh phase shift and backscattering amplitude. The Fourier transform ranges in k-space are indicated in brackets.

(a) Rh/Ti02 after reduction at 473 K (3.33 - 10.24 .&.-1)

(b) Rh/Ti02 after reduction at 723 K (2 .82 - 10.13 .&.-1)

(c) Rh/ Alp3 after reduction and evacuation at 623 K (2.97 -14.58 .&.- 1

)

(d) R.h/Ti02 after reduction at 723 K and evacuation at 623 K (2 .85 - 10 06 .&.- 1

)

(e) Rh/A1p3 after oxygen exposure at 100 K (3 .26 - 12.03 A-1)

(f) Rh/Ti02 after oxygen admission at 100 K (2.84 - 10.06 .&.-1)

(g) Rh/ Al20 3 after oxygen admission at 300 K (2 .55 - 12.45 A-1)

(h) Rh/Ti02 after oxygen exposure at 300 K (3.22 - 10.18 .&.- 1)

transform of the calculated EXAFS from shifting with respect to

the main peak in the Fourier transform of the original data. In all

cases, the difference spectra contained the same features and the

same contributions; only the magnitude of the contributions varied

slightly with varying Rh- Rh parameters. In all cases, the data

analysis procedure as described above led to the parameters as

presented in Table 7.2. The experimental errors have been

estimated by slightly varying the parameters in Table 7.2. Small

deviations from the 'best fit ' in the Fourier transform were not

allowed. In this way. a good approximation for the experimental errors could be established. The overall errors which are presented

in T abel 2, are the sum of this experimental error and an estimation

for the systematic error. The systematic error is induced by the

procedure of analyzing the data and incorporates errors induced by phase and magnitude transferribility, electron main free path , etc.

Page 153: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 143 Chapter 7

Coordination Distance acr 2 ll Eo D

Treatment NN number !Al (* 10-3 ft.-2) (eV) (a\ (al (a) (a}

1. Rh/TiO)

R473 Rh 2.5 0 .2 2.687 0.005 8 1 2.6 1 0 1.3 0.4 2.075 0.01 2 1 -1 1 0 1.0 0.3 2.78 0.01 2 1 -7 2

R473-R723 Rh 3.4 0.2 2.634 0.005 1.6 1 7.5 1 0 1.9 0 .3 2.60 0.01 7 2 -2 1 Ti 2.8 0 .4 3.41 0.03 3.5 1 10 2 Ti 2.8 0 .4 4.39 0 .05 3.5 1 10 2

R473-R723- Rh 3.4 0 .2 2.646 0.005 1 1 3.8 1 -E623 0 1.8 0.3 2.61 0.01 9 2 -2.0 1

Ti 2.7 0.3 3.43 O.Q3 4.5 1 10 2 Ti 2.4 0.4 4.32 0.05 4.5 1 10 2

R723-E623- Rh 3.4 0.3 2.637 0.005 1.5 1 5.1 1 -0100 0 1.0 0 .3 2.09 0.01 6.5 2 -5 1

0 1.2 0.3 2.61 0.01 9.8 2 -5 1 Ti 3.3 0.4 3.43 0.03 1.2 1 8 2 Ti 2.8 0 .4 4.33 0.05 4.5 1 10 2

R723-E623- Rh 3.4 0.2 2.63 0.01 2.9 2 2.4 2 -0100-0300 0 2.2 0.3 2.06 0.02 6.5 2 -3 2

0 1.5 0.3 2.75 0.02 9.8 2 5 2 Ti 2.5 0.3 3.48 0.03 1.2 2 10 3 Ti 1.4 0.4 4.36 0.05 4.5 2 10 3

2. Rh/A110 3

R623-E623 Rh 5.6 0.2 2.635 0.005 3 1 1.7 1 0 0.6 0.3 2.055 0.01 1 2 8.4 2 0 2.0 0 .2 2.73 0.01 2.5 1 -4.1 2

R623-E623 Rh 4.1 0.2 2.63 0.005 3.4 1 7.9 1 -0100 0 1.4 0.2 2.055 0.01 1 2 -2.5 1

0 1.4 0.2 2.77 0.05 2.5 1 - -2.9 1

R623-E623 Rh 1.9 0.3 2.645 0.01 4.4 2 4 2 -0100-0300 0 3.6 0 .3 2.03 0.01 4.7 2 -0.5 2

0 1.7 0.3 2.76 0.03 6 3 4 2

Page 154: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh / Ti02 page 144

Table 7.2 Final results from EXAFS data analysis

R Reduction in H2 at the temperature indicated

E Evacuation at the temperature indicated

0 = Admission of oxygen at the temperature indicated

(a) = Estimated overall (experimental+ systematic) error b

t:.a 2, the Debye Waller factor , is a measure for the disorder and 11Eo is a correction on the edge position : see ref ( 29) for more details .

Figure 7 .5 H RTEM micrograph of the Rh /Ti02 catalyst after reduction

at 773 K en subsequent passivation. The arrows indicate me­

tal particles. The micrograph is taken near Scherzer focus .

Therefore, the vis ibility of the metal particles is not very

good , but the uncertainty in their size is best.

Page 155: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 145 Chapter 7

7.3.2 Characterization with HRTEM

The metal particles were observed to have a very uniform and

narrow size distribution. In Figure 7.5, a micrograph near Scherzer

focus , the metal particles size was determined to be between 7 and

8 A, the uncertainty being about 2 to 3 A. in very good agreement

with the results from EXAFS, based on the Rh-Rh coordination

number of 3.2 - 3.4 (see Discussion). It can be seen that the metal

particles seem to favor positions at the edges of the surface planes

and not on flat surfaces. The number of metal particles observed on

the micrograph is roughly in agreement with the estimated occupa­

tion based on the .size of the Ti02 crystallites, the 4 wt% metal load and a metal particle size of about 7 or 8 A.. The in situ sinter­

ing experiments (see Figure 7.6) indicate that indeed the major part

of the metal particles is visible. During sintering by exposing the

catalyst to an intense electron beam, large rhodium metal particles

were formed. The d-spacings in these particles indicated that all

metal particles were fee rhodium; alloy particles in which Rh and Ti

are ordered (RhTi3, RhTi or Rh3Ti) have not been observed. The

same sintering experiments on Ti02 supported Ir catalysts resulted

in lr3 Ti alloy formation . Clearly, for rhodium catalysts, alloy forma­

tion in the SMSI state is unlikely to occur.

During the HRTEM experiments, several Rh/Ti02 catalysts

have been studied extensively after reduction at high temperature.

In these experiments no sign of even the slightest coverage has

been found. Image calculations (JJ) show that a monolayer of TiOx

on top of a rhodium particle containing 5 rhodium atoms should be

visible. The visibility depends on the defocus of the microscope, the

atomic configuration of the monolayer and the thickness and orien­

tation of the support. Before transferring the sample to the micro­

scope, the sample will always be passivated. On larger rhodium par­

ticles (other Rh/Ti0 2 catalysts have been investigated as well) islands of TiOx will be formed on top of the metal particles during

passivation. Since these islands will be several atomic layers thick,

Page 156: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/ Ti0 2 page 146

Figure 7.6 HRTEM micrograph of the Rh/ Ti02 catalyst after reduction

at 773 K en subsequent passivation and in situ sintering.

they must be visible with HRTEM. Although investigations were performed to image these in particular , such islands were never

observed. In conclusion, we feel that coverage is unlikely to occur for Rh / Ti02 catalysts.

In Figure 7.7 it is shown that the crystallographic surface plane

exposed most is evidently the [101) crystal face. Other planes exposed are [001) and [103). In Figure 7.7 an example is given. The particles on these micrographs are found mainly on edges of the

Ti0 2 crystallites and on [101) crysta l faces ; only few particles were found on [001) crystal faces.

Page 157: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 147 Chapter 7

Figure 7.7 HRTEM micrograph of the Rh / Ti02 catalyst after reduction

at 773 K en subsequent passivation . Several exposed crystal

faces are indicated .

7.4 Discussion

7 .4.1 Rh/ Ti02 after Reduction at 473 K

After reduction at 473 K and cooling down under H2, the aver­

age Rh-Rh coordination number is 2.5 (Table 7.2), demonstrating

that the metal particles were highly dispersed. The Rh-Rh distance

was found to be equal to the Rh- Rh bulk distance : 2.687 A. In

addition to rhodium nearest neighbors , two different rhodium­

oxygen contributions could be discerned. These contributions arise

from approximately 1.3 oxygens at 2.07 A and 1.0 oxygen at 2.78 A. The first Rh-O distance is only slightly longer than the Rh3+ -02

-

distance in bulk Rh 20 3 (2.05 A). This indicates that reduction was

incomplete and that part of the rhodium was still in the calcined,

oxidized state. This might seem to contradict our TPR results,

which indicated that reduction was complete at 400 K. There is ,

however , a slight difference between the two reduction treatments .

In TP R, the H2 mixture is forced to flow through the catalyst bed,

whereas in the in situ EXAFS cell H2 flows along and not through

Page 158: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti02 page 148

the self supporting sample wafer. Thus, in TPR the removal of

water out of the catalyst bed is by convection and consequently

much faster than during the in situ reduction in the EXAFS cell,

where water is removed only by diffusion. In the EXAFS cell, there­fore, reduction may take longer to complete. In our earlier paper

(23), we found that the catalyst was completely reduced after

reduction at 473 K. This can readily be explained by the lower

heating rate (2.5 K min-1 vs. 5 K min-1) and the longer reduction

treatment (2 h vs. 0.5 h) used in our earlier study.

Reduction of supported noble metal oxide particles is known to

be a fast process, limited only by nucleation, not by particle size

( 25). As a consequence, once the reduction of an oxide particle has started (nucleation) the· reduction process is rapidly completed. We

expect the particles to be either fully metallic or fully oxidized.

Therefore, in the catalyst reduced at 473 K, metal particles as well

as oxide particles will exist. Since the information in EXAFS is aver­

aged over all rhodium atoms and ions present in the sample, the

actual Rh- Rh coordination number for the rhodium atoms in the

metal particles is higher than the measured coordination number

(32). We will assume that the Rh-0 coordination number in the

oxide particles is 6, the same as in bulk Rh20 3 (32). This assump­

tion ts reasonable, even for small oxide particles, since Rh3+ ions may have 0 2

- ions from the support as one or more of the six oxy­

gen nearest neighbors . The fraction of rhodium present in the oxidic

form is then 1.3/6 = 0.22 and the fraction of rhodium atoms in the

metal particles is 0.78. This means that the actual Rh-Rh coordina­

tion number for the rhodium atoms in the metal particles is equal to

2.5/0.78 = 3.2.

As in earlier papers ( 28,32), we ascribe the oxygen neighbors

at 2.78 J.. to oxygen ions from the supporting oxide. The rhodium atoms in the metal-support interface have oxygen ions from the

support as nearest neighbors . The radius of zerovalent rhodium is about 1.34 A, the radius of divalent oxygen ions is about 1.4 A.. One

may therefore expect a Rh 0-02- di stance of about 2. 7 4 J.., in good

Page 159: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 149 Chapter 7

agreement with the reported value of 2.78 A. Since the oxygen ions

at 2.78 A are only nearest neighbors to rhodium atoms in the metal­

lic particles, the Rh0-02- coordination number must be corrected in

the same way as we have corrected the Rh0-Rh 0 coordination

number: the corrected value is 1.0/0.78=1 .3. To find the real

number of oxygen neighbors for each interfacial rhodium metal

atom, this value has to be divided by the fraction of metal atoms in

the metal-support interface ( 28).

It is often found that background subtraction is difficult in

EXAFS spectra of catalysts in which the metal component is

(partly) oxidized. Using the cubic spline method. there is no unam­

biguous criterion for a correct background subtraction. For metallic

catalysts, the choice of the smoothing parameter in the cubic spline

routine should be such that the magnitude of the Fourier transform

in the region below 1 A is low. For oxid ic samples as the present

one, this can be achieved, but it is always observed , that the first peak in the Fourier transform, in this case the Rh3+-0 2

- contribu­

tion, has slightly decreased in intensity. The second peak, in gen­eral, has not decreased in intensity. Therefore, the Rh3+-02

- coordi­

nation number will be slightly underestimated and · so will be the

corrected Rh0-Rh0 and Rh0-02- coordination numbers. In general,

the error will not exceed 20% .

In the analysis described above, we have used the phase shift

and backscattering amplitude obtained from Rh 20 3 to calculate the Rh0-0 2

- EXAFS functions. Since in Rh 20 3 the absorber-scatterer

pair is Rh 3+-02- with a coordination distance of 2.05 A, it is

incorrect to use data from Rh20 3 EXAFS functions to calculate 2.78 A Rh0-02

- EXAFS functions in which the coordination distance is

considerably larger and the valence state of one of the members of

the absorber-scatterer pair is different . It has been shown (28,33) that in such cases the calculated distances are reliable. The Rh0-0 2

-

coordination numbers however, are underestimated. There is no indication for the degree of underestimation. The actual average

Rh0-0 2- coordination number must therefore be higher than the

Page 160: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 150

above reported value of 1.3.

Figure 7 .8 Small metal particles

(a) four-atom metal particle

(b) five-atom metal particle

(c) seven-atom metal particle

(d) eight-atom metal particle

At this point, we are in a pos1t1on to estimate the size of the

metal particles. In Figure 7.8, examples of four, five, seven and

eight-atom metal particles are shown. In the four-atom metal parti­cle, each metal atom has three direct rhodium nearest neighbors.

The average Rh- Rh coordination number for this particle therefore

is 3.0. In the five-atom metal particle, each interfacial atom has

three nearest neighbors, two interfacial atoms and the top atom.

The top atom has four direct nearest neighbors , the four interfacial

atoms. The average number of rhodium nearest neighbors in this

particle is therefore 3.2 and the average diameter is about 6.5 A.. The average coordination number for the seven and eight atom

Page 161: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 151 Chapter 7

metal particles is 4.0 and their diameter is about 8.5 A. Since the

observed Rh-Rh coordination number is about 3.4 ± 0.3, we con­

clude that the metal particles in the Rh/Ti0 2 catalyst contain

roughly 5 or 6 metal atoms. The results of th.e HRTEM characteri­zation, a very uniform and narrow particles size distribution around

7 A, are in good agreement with this conclusion. Obviously, even for very small metal particles, the EXAFS coordination number very

. accurately determines the metal particle size. These results, how­

ever, leave some room for a discussion about a particle size distribu­tion. The EXAFS results give a Rh-Rh coordination number of 3.4 ± 0.3. This corresponds to inetal particles containing from 4 up to

6 or 7 or even 8 atoms. The particle diameter determined with

HRTEM is about 7.5 ± 1.5 A. This corresponds to metal particles

containing roughly 3 up to about 9 atoms per particle. Because the average particle size as determined by both EXAFS and HRTEM

points to particles containing about 5 atoms, we assume the metal

particles contain between 4 and 8 metal atoms. The average metal

particle will contain 5 atoms. Even explanations in t~rms of bimo­dal size distributions are possible. In that case, a minor part of the

Rh atoms is present in particles containing one or twp metal atoms and the major part in larger particles, which contain bn the average

7 or 8 atoms per particle. We believe, however, that the latter situa­

tion is unlikely.

7 .4.2 Rh/Ti02 after Reduction at 723 K

After reduction at 723 K, we found no evidence for the pres­

ence of rhodium oxide particles. Apparently, at this temperature the

reduction was complete. The Rh-Rh coordination number was 3.4, in good agreement with the corrected (.and possibly underestimated)

value of 3.2 reported above. This indicates that during the high temperature reduction no sintering had taken place. The Rh-Rh

coordination distance had decreased markedly . For non-SMSI

Page 162: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 152

catalysts such as Rh/ Al 20 3, such a decrease in the Rh-Rh distance

has been observed after reduction and evacuation at high tempera­

ture (34) . After removal of adsorbed hydrogen, the remaining sur­

face metal atoms contract in order to compensate for the loss of hydrogen nearest neighbors. Obviously, since the measurement was

performed in an H2 atmosphere at 100 K. the metal particles in the

Rh/Ti0 2 catalyst after reduction at 723 K did not chemisorb H2.

Clearly, the metal particles were in the SMSI state. In our earlier

Rh/Ti0 2 study we reported the same observations.

The presence of the long distance Rh-0 contribution is as

expected. However, the bond length between interfacial rhodium

atoms and supporting 0 2- ions has decreased from 2.78 to 2.60 A.

The decrease in Rh-Rh distance, 0.053 A, indicates a decrease in the

average rhodium atomic radius of 0.027 A. This decrease, however,

is too small to account for the observed decrease of 0.18 A in the

Rh0-0 2- bond. Although it is obvious that this decrease is the result

of a change in the metal-support interaction, it is not evident what

kind of interaction has caused this decrease. In our earlier study we reported a decrease in the Rh0-02

- distance of 0.04 A. That catalyst

was reduced at a temperature 50 degrees below the reduction tem­

perature of the present catalyst (673 and 723 K, respectively) . The

metal-support interaction induced by the high temperature reduc­tion is obviously stronger when the catalyst is reduced at higher

temperatures.

In the SMSI state, two additional contributions were detected,

titanium neighbors at 3.41 and 4.39 A. These contributions could not be detected after the low temperature reduction. In our previous

paper (23), we reported that a 3.4 A Rh-Ti contribution was present

after reduction at 673 K. The better signal-to-noise ratio of the

present experiments and the enhanced metal-support interaction obviously enable us to distinguish even more contributions. To

explain these contributions, we need to take a closer look at the supporting oxide, anatase Ti02. The H RTEM micrographs show

that the [101] anatase crystal face is exposed most. HRTEM also

Page 163: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 153 Chapter 7

showed that the majority of the metal particles was present on the

edges of the Ti02 crystallites, the rest of the metal particles was present predominantly on [101] crystal faces. It is therefore not evi­

dent what the structure of the (average) crystal face is on which the particles rest. The majority of the crystal faces are 1101] faces.

It is not unlikely, that the crystal face on the edge of the crystal­lites. between two [101] crystal faces, is indeed a [001] type crystal

face . Surface energy calculations have shown that the [001] crystal face is indeed a stable crystal face (JS). We therefore assume that the metal particles rest on [001] and [ 101] type crystal faces.

In Figure 7.9a and 7.9b, both a [001] and (101] anatase crystal

face are shown. Both crystal faces consist of a two dimensional rec­tangular array of oxygen ions, with as many octahedral sites as oxy­gen ions. Half of these octahedral sites is occupied by Ti4+ ions.

We assume that the five (or eight) atom metal particle (see Figure

7.8) rests on these crystal faces. Figure 7.9c and 7.9d show the most plausible arrangement for this metal particle on both faces;

Figure 7.9 Ti02 crystal faces

(a) 1101] anatase crystal face

(b) [001] anatase crystal face

( c) [101] anatase crystal face with a five atom metal particle

(d) 1001] anatase crystal face with a five atom metal particle

(e) [101] anatase crystal face with a five atom metal particle after

reduction

(f) 1001] anatase crystal face with a five atom metal particle after

reduction

Ri = 2.7 A R2 = 2.0 A R3 = 3.4 A R4 = 4.3 A

(Rh0-o2-)

(Rh0-Ti)

(Rh0-Ti)

(Rh0-Ti)

Page 164: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 154

a b

c d

e f

eTin+

Page 165: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 155 Chapter 7

four metal atoms are interfacial atoms and each interfacial atom has four oxygen nearest neighbors . The average Rh0-0 2

- coordination

number in this arrangement is 3.2. Earlier we argued that our

method of calculating Rh0-0 2- coordination numbers underestimates

the real Rh0-02- coordination numbers. The apparent discrepancy

between the expected and the measured value of 1.9 is therefore acceptable.

In the models we have described , we assumed that in a first approximation the rhodium atoms of the metal particles are

situated on lattice-oxygen positions, that is, positions which would

have been occupied by oxygen ions, if another Ti0 2 layer had been

deposited on the crystal. Therefore, it does not matter whether the metal particles rest on [101] or [001] or any other crystal face.

Because the rhodium atoms occupy oxygen-equivalent positions, the Rh-0 2

- distances (and Rh-Ti distances, these will be discussed

later) will be the same, regardless of the crystal face on which the particles rest. The coordination distances are model-independent.

The coordination numbers , however, may vary , but only slightly . Another consequence of this assumption is, that the metal particles

grow epitaxially on the supporting oxide, which is not unlikely for very small metal particles . The metal particles in Figure 7.8a and

7.8c, which have been grown epitaxially on [111] type crystal faces, also fit on the [001] and [101] type crystal faces. Thus, the four, five, seven and eight atom metal particles in Figure 7.8 represent possible structures of the metal particles. For the sake of clarity,

however , we will focus only on the five atom metal particle on a

[001] or [101] type anatase crystal face . One more detail needs closer attention. Anatase is build of slightly distorted octahedra . As a result, in pure anatase, two 0 2

- -02- distances are encountered . If

the Rh atoms are situated on 0 positions, one should expect the same two Rh-0 2

- distances as well. We found however only one

Rh-0 2- distance. This can be accounted for as follows. In the [101] plane (see Figure 7.9a), the lower 0 2

- ions, the ions underneath Ti4+ ions , are shifted alternately slightly to the right and slightly to the left with respect to their position as shown in Figure 7 .9a. This

Page 166: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 156

has very little influence on the Rh-0 2- distances because the dis­

placement is perpendicular to the distance. For metal particles

situated on [101.] type crystal faces we thus expect only one Rh-0 2-

distance with possibly a increased disorder around that distance. From Table 7.2, it is indeed clear that the disorder in the Rh0-0 2

-

distances is larger than any other disorder (note that the disorder is

expressed relative to the disorder in the reference compounds) . The

0 2- ions in a ideal (001] plane (see Figure 7.9b) are situated alter­

nately a little above and a little underneath the horizontal plane.

But if we assume that the outermost [001] plane of 0 2- ions has

relaxed in such a way that all 0 2- ions are in the same plane, only

one Rh-0 2- distance is present with possibly an increased disorder.

A closer look at ou; model in Figure 7.9 reveals three Rh0-Ti4+

distances : R2 = 2.1 , R3 = 3.4 and R4 = 4.3 A. The average coordi­

nation numbers for these contributions are respectively 0.4 , 1.6 and

3.2. (The latter coordination number is higher than one might derive from Figure 7.9. In Figure 7.9, however , only the Ti4+ ions in the

top layer are shown . The Ti4+ ions in the second layer will also con­

tribute to the 4.3 A Rh0-Ti4+ contribution . Another 'hidden' 4.3 A Rh0-Ti 4+ contribution arises from the top rhodium atom in the five

metal atom particle and the Ti4+ ion directly underneath the top

atom). These coordination numbers are low and one would expect

their contributions , especially for the longer distance, not to be

detectable by EXAFS. The EXAFS amplitude is in a first approxi­

mation inversely proportional to the square of the distance. The

values of N/R 2 for the three contributions are 0.09, 0.15 and 0.16,

respectively. Because of the Debye Waller factor, however , these

values will change . For comparison, 1\J/R2 for the Rh0-02- contribu­

tion is 0.44. Since the backscattering amplitudes for the Rh-Ti and

Rh-0 contributions do not differ much in magnitude (the Ti back­scattering amplitude is slightly larger) , it is evident that the Rh0-

Ti4+ contributions will be hard to detect , compared to the Rh0-0 2-

contribution . Because of the Debye Waller factor, the 3.4 A contri­

bution is expected to be the most dominant of the three Rh0-Ti4+

contributions . In Figure 7.3a, the difference spectrum for the 473 K

Page 167: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 157 Chapter 7

reduced catalyst, a small contribution around 3.4 A can indeed be

observed. But this contribution is too small to allow a reliable analysis.

In our previous paper , we already reported the presence of a 3.4 A Rh-Ti4+ coordination in the SMSI state catalyst . In the 723 K

reduced sample of the present study , both the 3.4 and 4.3 A Rh­Ti4+ contributions are indeed observed, but the coordination

numbers are higher than anticipated . The explanation for the enhanced presence of these Rh-Ti4+ contributions in the SMSI state

can be found in the reducibility of the Ti0 2 supporting oxide. Dur­ing the reduction process, oxygen is removed from the Ti0 2 surface

in the form of water . This process is centered around the metal par­ticle, which provides reactive hydrogen atoms and thus catalyzes

this reduction process. After a few oxygen ions have been removed , the second [001] or [101] crystal plane becomes exposed. Bare Ti3+

ions from the first (outermost) crystal face remain on top of the 0 2

- ions of the second plane. This situation is energetically very

unstable and these Ti3+ ions will migrate to empty ocrtahedral sites in the second oxygen layer and possibly the third or even further

crystal planes . What we see here, is the formation of a TiOx subox­ide and especially the formation of part of a shear plane . The five atom metal particle is positioned in the same way on the second crystal planes as it was on the outermost plane before reduction,

but now the interfacial rhodium atoms have more Ti 3+ and Ti4+ nearest neighbors at 3.4 and 4.3 A (see Figure 7.9e,f). This increase

in Ti nearest neighbors explains our EXAFS results. We thus con­clude that after reduction at high temperature, the metal particles

rest on partially reduced Ti02.

Our first conclusion from this analysis is that the metal parti­

cles are very small indeed. The average rhodium metal particle in the 4 wt% Rh/Ti0 2 catalyst contains about 5 rhodium atoms. The diameter of these particles is about 6.5 A. This was confirmed by HRTEM . This technique revealed even more information. After

exposing the sample to an intense electron beam, which resembles

Page 168: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 158

reduction at high temperature , the metal particles had sintered to

large fee rhodium metal particles; the presence of any ordered alloy

of Rh and Ti could be ruled out. The same experiments on Ir /Ti02 resulted in lr3 Ti alloy particles. Thus, it is unlikely that in the

Rh/Ti0 2 catalyst in the SMSI state alloy formation takes place.

This in in accordance with the results resported by Beard and Ross ( 19). They found that upon heating a platinum catalyst supported

on carbon and impregnated with TiCl4 Pt3 Ti alloy particles were formed . For iridium and platinum alloy formation may play a role in

the SMSI state, but for rhodium catalysts this is not the case.

Secondly, we found with EXAFS no evidence for coverage, and

HRTEM confirmed this. In the HRTEM study, this catalyst and

other Rh/Ti0 2 catalysts with larger metal particle sizes were stu­

died extensively after reduction at high temperatures and subse­quent passivation . In these studies no sign of even the slightest

coverage was found . It is known , that after passivation and subse­quent reduction at low temperature the normal hydrogen adsorption

capacity of the metal particles is partly , but not completely restored (7 ). Therefore, if coverage occurs after reduction at high tempera­

ture and suppresses the hydrogen chemisorption by decreasing the exposed metal surface area, passivation will not completely remove

the covering oxide . Since in our HRTEM experiments on Rh/Ti0 2 catalysts we never observed covering, we conclude that coverage is

unlikely to occur for Rh/Ti0 2.

In the following we will discuss oxygen adsorption experiments

which have been performed in order to investigate whether the metal surface of the rhodium particles in the SMSI state is blocked

by a covering TiOx (sub)oxide . We exposed the catalyst in the in situ cell to pure 0 2 at 100 K. Before oxygen admission the

catalyst had been evacuated at 623 K. Subsequently the tempera­ture was raised to room temperature and another EXAFS spectrum

was recorded . We performed the same experiments on a Rh/ Al 20 3 catalyst (reduced at 623 K) in order to compare oxygen adsorption properties of rhodium metal particles in the 'SMSI state' and the

Page 169: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 159 Chapter 7

'normal state' (A1 20 3 supported). For the sake of clarity, we will

discuss alternately the results of the experiments on the Rh/ Al20 3 and Rh/Ti02 catalysts.

7 .4.3 Evacuation at 623 K

After reduction and evacuation of the Rh/ A1 20 3 catalyst, the major contribution is from rhodium neighbors with a Rh-Rh coordi­nation number of 5.6. This indicates that the metal particles con­tain about 15 atoms and are roughly 10 A in diameter (24). Com­pared to the bulk value, the Rh-Rh coordination distance has decreased by 0.052 A. As we have described in the foregoing this contraction is the result of the evacuation procedure. Because of the removal of adsorbed hydrogen the surface metal atoms contract in order to compensate for the loss of neighbors . A Rh0-02

- contribu­tion is also present, originating from neighboring 0 2

- ions in the metal-support interface. The small contribution at 2.06 A has to be ascribed to the presence of rhodium oxide. The maximum tempera­

ture during the reduction treatment (623 K) was high enough to obtain complete reduction of the Rh/A120 3 catalyst (28). However,

it is possible that partial oxidation has taken place due to formation of water during the evacuation procedure or because of a small inleakage during or after the evacuation treatment.

Page 170: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 160

7.4.3.2 Rh/Ti02

After reduction at 723 K of a fresh Rh/Ti0 2 sample followed

by evacuation at 623 K, the EXAFS spectrum (in k-space as well in

r-space after Fourier transformation) resembles very closely the

EXAFS spectrum measured directly after reduction. All differences are within the indicated experimental errors . The Rh-Rh distance is

still significantly smaller than the bulk value (2.687 A) and almost equal to the Rh-Rh distance in the evacuated Rh/ A1 20 3 catalyst.

Thus, we have confirmed that in the SMSI state the rhodium metal particles do not adsorb H2.

7 .4.4 Oxygen Admission at 100 K

After admission of 0 2 to the Rh/ Al 20 3 catalyst at 100 K, the

EXAFS spectrum changed drastically . The Rh-Rh contribution diminished, as is clearly reflected in the decreased amplitude of the EXAFS function at higher k-values (k > 7 A-1

• see Figure 7.1e). Obviously, after 0 2 admission at this temperature all the metal par­ticles are partially oxidized. This situation is different from the par­

tially reduced Rh/Ti0 2 catalyst, where some of the particles are already fully reduced and others were still oxidized . It is impossible to calculate the size of the remaining metal kernel because now this metallic kernel is covered with Rh20 3 or possibly another form of

rhodium oxide and the Rh 3+-0 2- coordination number in the cover­

ing shell of oxide will be different from six . At lower k-values, the

influence of the pronounced presence of the two Rh-0 contributions

is obvious. These qualitative conclusions are confirmed by the quantitative analysis (cf. Table 7.2). The decrease in the Rh-Rh

Page 171: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 161 Chapter 7

coordination number and the increase of the Rh3+-0 2- coordination

number indicate that the oxidation process has started. Note, that

the high k-value part of the EXAFS function is very sensitive for

changes in the Rh-Rh contribution (i.e. average metal particle size) and the low k-value part is sensitive for low-Z scatterer contribu­

tions.

7 .4.4.2 Rh/Ti02

The results for the Rh/ Al 20 3 catalyst presented above are in

sheer contrast to the results for oxygen admission to the Rh/Ti0 2 catalyst. At higher k-values, the EXAFS spectrum of the Rh/Ti0 2 catalyst after 0 2 admission at 100 K resembles very closely the

spectra of the reduced and the evacuated samples (cf. Figures 7.1d

and 7.1f). No changes in the Rh 0-Rh0 parameters could be detected.

Obviously, for the Rh/Ti0 2 catalyst, the basic structure of the

metal particles remained intact and oxidation has not taken place.

At lower k-values there are small differences. This becomes clear

from the Fourier transform of the difference spectrum, Figure 7.3f.

Apart from the Rh-0 and Rh- Ti contributions discussed before, a Rh-0 contribution at 2.09 A. is clearly visible with a coordination

number of 1.0. This Rh-0 contribution must be different, however,

from that of the sample reduced at 423 K. In the EXAFS spectrum

of that sample, the 2.06 A. Rh-0 contribution was assigned to Rh 3+-0 2- absorber-scatterer pairs, present in Rh 20 3 particles. But in

the spectrum taken after oxygen admission (cf. Figure 7.1f) no

differences with the spectrum taken after evacuation (Figure 7.1d)

can be observed at k > 6 A.-1. In this region, EXAFS is very sensi­

tive to changes in the Rh-Rh contribution, therefore neither the

Rh-Rh coordination number nor the metal particle size have

changed. The results of the detailed data analysis confirm this (see

Table 7 .2). Clearly, oxidation of the rhodium particles on the Ti02 support has not taken place. The only possibility left is to ascribe

the 2.09 A. Rh-0 contribution to oxygen adsorbed on the surface of

Page 172: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 162

the metal particles. The atomic radius of zerovalent rhodium in the

metal particles is 2.64/2 = 1.32 and the radius of covalent oxygen

is about 0.73. The expected Rh0-o0 distance (2.05 A) is in good

agreement with the calculated distance of 2.09 A.

A very important result of this analysis is that the metal parti­

cles in the SMSI state are capable of adsorbing oxygen. In order to

be able to adsorb oxygen, the metal particles must be bare,

uncovered, or at least not fully covered with a TiOx suboxide. Even

though oxygen was able to adsorb on the metal particles, oxidation

did not take place. in contrast to the Rh/ Al20 3 catalyst. This

suppressed oxidation can only be the result of an electronic

influence from the TiOx .suboxide.

7 .4.5 Oxygen Admissi.on at 300K

After oxygen admission at 100 K, the Rh/ Al20 3 catalyst was

warmed up to room temperature. The EXAFS spectrum (Figure 7.1g) differed completely from the spectra of the Rh/ A120 3 catalyst

after evacuation and oxygen admission at 100 K (Figures 7.1c and

7.1e). At high k-values, almost no high-Z scatterer EXAFS is visi­ble, while at low k-values. the low-Z scatterer contribution differs markedly from the low k- value part of the two preceding spectra.

The main reason for this is a decrease in the Rh-Rh coordination number and an increase in the oxidic Rh3+-02

- contribution.

Because of the high Rh3+-0 2- contribution, the analysis of this ·

spectrum was different from the procedure as described above. The

major contribution to the spectrum originated from oxygen scatter­

ers and therefore a k 1-weighted Rh-0 phase corrected Fourier

transform rather than a k 3-weighted Rh- Rh phase and

Page 173: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 163 Chapter 7

backscattering corrected Fourier transform was used to calculate

the different contributions in the EXAFS spectrum. Since in this

Fourier transform an incorrect phase shift function has been used

for the (small) Rh0-Rh 0 contribution, the accompanying Rh-Rh peak

is shifted and coincides with the peak originating from the Rh0-0 2-

bond of about 2.76 A. The fact that the peak at the right hand side of the Rh3+-0 2

- contribution is indeed the result of two (Rh0-Rh0

and Rh0-0 2-) contributions is indicated if Figure 7.3g, in which the

magnitude of the Fourier transform of the original data and of the

Fourier transforms of the calculated EXAFS spectra of Rh3+-02- +

Rh0-Rh0 and of Rh3+-02- + Rh0-02

- are presented. In the Fourier

transform of the former calculated EXAFS function, a strong des­

tructive interference is visible in the region between both peaks, in

the latter Fourier transform there is a slightly constructive interfer­

ence. In the same region in the Fourier transform of the original

data, there is a slightly destructive interference. Therefore, we

must conclude that the right-hand side peak in Figure 7 .3g is the

result of the sum of two contributions, namely Rh0-Rh0 and Rh0-

02-. In a k 1-weighted Rh-Rh phase and backscattering amplitude

corrected Fourier transform, all contributions are separated, but, as mentioned above, this Fourier transform is not suitable to optimize

the dominant low-Z scatter contribution. Because of this, the errors in the parameters of the Rh-Rh and Rh0-0 2

- contributions used to

calculate the best fitting spectrum, are larger than the errors in the

same parameters of the other EXAFS spectra of the Rh/ A1 20 3

catalyst.

As argued before (see 7.4.1), the Rh 3+-02- coordination

number is not very accurate. The value of 3.6 m:.ist be considered

as a lower limit. The Rh 0-Rh0 and the Rh0-0 2- coordination

numbers on the other hand are more reliable. The Rh0-Rh0 contribu­tion had diminished markedly, while the contribution of the long

distance Rh0-0 2- coordination number increased relative to the Rh­

Rh coordination number. From this we conclude that the remaining

metal kernels of the rhodium particles are covered with rhodium

oxide. The presence of rhodium oxide on top of the metal particle

Page 174: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti02 page 164

creates an extra interface in which zerovalent rhodium is in contact

with an oxide. In this interface, new Rh0-02- bonds will be present.

This proves that EXAFS is capable of detecting coverage. It impli­

cates, that coverage of titania supported metal particles in the SMSI state is even more unlikely than we have suggested up to now .

7 .4.5.2 Rh/ Ti02

When the Rh/Ti0 2 catalyst after oxygen admission at 100 K

was warmed up to room temperature, the EXAFS spectrum at

higher k-values still re~embled quite closely the spectrum of the

sample after evacuation. This indicates that the basic structure of

the metal particle size had not changed and that the formation of

rhodium oxide had not yet taken place. Detailed analysis confirmed

that the Rh-Rh coordination number remained constant. Compared

to the spectra after evacuation and after oxygen admission at 100 K, the differences in the EXAFS spectrum at lower k-values

were much more pronounced. The main reason for this is an

enhanced influence from the short distance Rh0-o0 contribution.

Obviously, since oxidation had still not taken place, the metal parti­cles had adsorbed more oxygen . This will be explained in the fol­

lowing paragraph.

The contributions from the support ions (Ti and 0 2-) all

relaxed to longer distances . For the Rh0-Ti contributions, this may seem to be in contradiction with the indicated error . However, the

errors indicated in Table 7.2 are overall errors . They are the sum of

the experimental error and the systematic error . The experimental

error is the result of (in)accuracies during analyzing the data. The

systematic error is the result of the procedure of analyzing the data

and may be regarded as being equal for comparable contributions.

For example, the quoted error of 0.03 in the Rh0-Ti distance of 3.43

A in the sample after oxygen admission at 100 K indicates that the

real Rh0-Ti distance has a 95% probability of being in the range of

Page 175: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 165 Chapter 7

3.40 to 3.46 A. When comparing two distances, however, the sys­

tematic error can be ruled out and only the experimental error

should be taken in account. To a good approximation, the experi­

mental error in the Rh-Ti distances is about 0.01 A. Thus, the Rh0-

Ti distance in the sample after oxygen admission at 300 K is in

fact significantly larger than the distance found after oxygen admis­sion at 100 K. The Rh0-02- contribution relaxed to 2.75 A, which is

very close to the Rh0-02- distance for the catalyst in the normal

state. The Rh0-Ti contributions relaxed to 3.48 and 4.36 A, indicat­

ing that the metal-support interaction has weakened. Both Rh0-Ti

coordination numbers have decreased . The decrease in the coordi­

nation number of the 3.48 A contribution is almost within the

experimental error, but the decrease in the 4.36 A Rh0-Ti contribu­tion is more pronounced. Clearly, Ti ions in the vicinity of the metal

particles have disappeared. This indicates that the TiOx suboxide

around the metal particles has started to re-oxidize. This is, of

course, in agreement with the fact that the metal-support interac­

tion has decreased in strength, the SMSI state has been removed partly and it may explain the fact that the metal particles have

adsorbed more oxygen.

7.4.6 Different Rh0-Ti Contributions

In the preceding paragraphs, we discussed two Rh-Ti contribu­

tions present in the EXAFS spectra. Based on these findings, on

literature data and on HRTEM, we concluded that the rhodium

metal particles in Rh/Ti0 2 probably rest on anatase (001] and (101] crystal faces. In a model in which a five atom metal particle rests

on a [001] or [101] anatase crystal face, the distances found with EXAFS are indeed present. It is evident that EXAFS will not be able

to detect the distances longer than 4.4 A. However, the short 2.1 A Rh-Ti distance should be detectable by EXAFS but we have not

been able to straightforwardly detect this Rh- Ti contribution. One

Page 176: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 166

of the most important reasons is that especially in the normal state, the expected coordination number for this 2.1 A bond is low. Con­sider for example the particle on the [001] plane : two rhodium atoms have each one Ti neighbor; the coordination number there­fore is 2/5=0.4. We may expect that this coordination number will increase for the metal particle in the SMSI state by a factor of two (or less). Thus, the maximum coordination number one may expect for this contribution is 0.8 and therefore still too small to be detect­able by EXAFS. Compare for example Figures 7.3b and 7.3d. The Rh 0-02

- coordination number is 1.9. This contribution is well above the noise level (about 1 *10- 2

) but decreasing it by a factor 2 brings it already very close to the noise level. Therefore, a Rh0-02

- coordi­nation number of 1 would be difficult to detect and even more difficult to analyze. Consequently, a Rh0-Ti contribution with a coordination number less than about 1 is almost impossible to detect or analyze. However, in the imaginary parts of the Fourier transforms of the difference spectra of the Rh/Ti02 catalyst in the SMSI state (Figures 7.3b and 7.3d) small deviations between theory and experiment are visible around 2 A. These deviations could not be completely eliminated, but adding a 2.1 A Rh-Ti contribution improved the agreement between calculated and experimental spec­tra . Because of the low coordination number, there remains some uncertainty about the scatterer. It was not possible to unambigu­ously identify the scatterer as oxygen or titanium , although titanium as neighbor resulted in a better fit. It is evident that at the low-R side in the Fourier transform of the EXAFS spectrum, a small contribution is present and it is likely that this contribution is due to a Rh-Ti coordination.

Page 177: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 167 Chapter 7

Figure 7.10

(a) Magnitude of the k1 -weighted Fourier transforms of

solid line : EXAFS spectrum of the Rh Ti alloy

dotted line : Calculated Rh-Ti EXAFS function

dashed line : Calculated Rh-Rh EXAFS function

(b) Reference spectra of

solid line : Calculated Rh-Ti EXAFS function

dotted line Rh-Ti EXAFS function from inverse Fourier

transform 0. 15 -,----,------r--~--~

0.10

0.05

-RhTi ··········Ti -----·Rh

..-

'i /~', f\

t\ ' '

! \

,/ .. " ,, ~

~· ··· .

0. 00 +---'-----+----'--__;_-"! 0 2

R [A] 4

0.10

0.05

0 . 00

-0 . 05

-0 . 10

-0 .15 2

7 .4. 7 Comparison with Literature Data

-Calcul . ···········Inv. FT

14

Recently, similar EXAFS spectra of a Rh/Ti0 2 catalyst in the SMSI state were reported by Sakelson et al. (36). They ascribed the

major peak in the magnitude of the Fourier transform at the low r­side of the dominant Rh-Rh contribution to a Rh-Ti contribution, while our results point to a Rh0-02

- contribution for the same peak

in the Fourier transform. We call the assignment of this peak to Rh-Ti into question for the following reasons :

( i) The above mentioned contribution, referred to as Rh-X, resembles closely the peak in the Fourier transform of the EXAFS functions of Rh/ A1 20 3 catalysts, which has unambigu­ously been ascribed to a Rh-0 contribution (28,32).

Page 178: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti02 page 168

Furthermore, we observed the same Rh-X contribution in the

'normal' as well as in the SMSI state of the catalyst. Since in the normal state there is no reason to assume Rh-Ti coordina­

tion, the assignment of the Rh-X peak to oxygen neighbors in the metal-support interface is much more likely.

(ii) In the fitting procedure described in (36), only titanium as a

neighbor was taken into account; oxygen was not tried as a possible neighbor. The fitting procedures on our spectra yield,

in terms of sum-of-least-squares and variance, better results with oxygen than with titanium as scatterer in the Rh-X con­tributio·n.

(iii) The Rh-Rh and Rh-Ti contributions in the RhTi reference

compound overlap both in the k 1- and in the k 3-weighted

Fourier transforms. A rough approximation indicates that the overlap is at least 25%. Also, the well known side lobe of the

Rh-Rh peak (due to the k-dependence in phase and back­scattering functions) is hidden under the Rh-Ti peak. In Figure

7.10a, the k 1-weighted Fourier transforms of the calculated Rh-Ti (N=8, R=2.676) and Rh-Rh (N= 4, R = 2.949) EXAFS spectra are shown, demonstrating their overlap. Thus, inverse transformation as performed in (36), using a window to

separate both contributions, definitely results in incorrect phase and backscattering functions for the Rh-Ti absorber­scatterer pair. In Figure 7.10b the Rh-Ti EXAFS function is shown as we have calculated it and the same contribution

which has been achieved by an inverse Fourier transform over the window 0.00 - 2.38 A.. It is obvious that especially at lower k-values the differences are substantial.

(iv) Sakelson et al. stated that after a k 1 -weighted Fourier

transform, the Rh-Rh and Rh-X contribution can be separated

much better (36). However, in a k 1-weighted Fourier transform , the low-Z scatterer information is much more pro­nounced than in a k 3-weighted Fourier transform. As a result, both Rh- Rh and Rh-X peaks still overlap, and the overlap is even larger . Because of the larger overlap, there is a strong destructive interference in the region between the two peaks.

Page 179: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 169 Chapter 7

This can best be seen in the imaginary part of the Fourier

transform . Because of the anti- phase behavior of the ima­ginary part of the separate contributions. there is a strong

destructive interference and the magnitude of the Fourier transform shows a sharp minimum between both peaks. In a

k 1-weighted Fourier transform, the Rh-Rh and Rh-Ti contribu­tions are therefore separated even less clearly than in a k 3

-

weighted Fourier transform. This can also be seen in Figure

7.10, where the magnitude of the k 1-weighted Fourier

transform of the Rh Ti alloy is compared to the separate Rh-Ti

and Rh-Rh contributions. Because of the destructive interfer­

ence, the sum of the two separate contributions (solid line) is lower in magnitude than the respective contributions at their

maximum. A careful inspection also reveals that the peak maxima are shifted slightly. Also for this reason , it is incorrect

to use a window in a k 1-weighted Fourier transform to

separate the Rh-Rh and Rh-X contributions.

(v) It is insufficient to use a k\(k) plot to show the result of the

best fit (36), since the k 3 term strongly enhances the high-Z Rh-Rh contribution which makes the plot less sensitive to the

low-Z Rh-X contribution. A plot of x<k) is necessary to com­

pare experimental and calculated data.

Rh-Ti phase and backscattering functions have in a first approximation the same general k-dependence as the Rh-0 phase

and backscattering functions . This is as expected, because of the

low-Z character of both oxygen and titanium (37 ). As a result, it is

always possible to fit a Rh-0 EXAFS function with Rh-Ti parame­ters , and vice versa. A difference in backscattering amplitude can

be compensated with a particular (i.e. wrong) choice of coordination

number and Debye-Waller factor. Because of a different phase func­

tion, however , such a fit will result in incorrect coordination dis­tances. Clearly , to discriminate between Rh- Ti and Rh-0 contri bu­

tions, a very careful data analy s is procedure is required. The pro­cedure which we have used in our analysis procedure is superior to

the procedure described in (36). We believe that the assignment of

Page 180: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 170

the Rh-X contribution to oxygen neighbors is correct and that the

conclusion drawn by Sakelson et al. about RhTi alloy formation in

the SMSI state, is not justified.

7 .5 Final Conclusions

From tbe observed neighboring Ti ions, we conclude that in

the SMSI state the metal particles rest on a TiOx suboxide. This has already been suggested in the first reports on SMSI ( 1,3). In the

SMSI state, the metal p<,1rticles are incapable of adsorbing H2 and as

a consequence the Rh-Rh distance has contracted by about 0.05 A. The Rh0-02- distance in the metal-support interface has contracted

as well, by about 0.18 A. Th is is the result of an interaction between metal and support. In the SMSI state, the rhodium metal

particles are capable of adsorbing 0 2 at 100 and 300 K. Therefore, and because we found no evidence for a covering suboxide with

EXAFS nor with HRTEM , we conclude that coverage is not com­plete or is even absent in Rh/Ti0 2. Oxidation of the metal particles

in the SMSI state, however, is suppressed, even at 300 K. This

again must be the result of an electronic metal-support interaction

and cannot be explained by coverage.

Another interesting phenomenon observed has been observed

with EXAFS. When warming the oxygen exposed Rh/Ti0 2 catalyst to room temperature, the TiOx suboxide in the vicinity of the metal

particles starts to reoxidize. At the same time, the metal-support interaction weakens . This proves that indeed the suboxide under­neath and around the metal particles plays a key role in the anomalous properties of the supported rhodium metal particles in ·

the SMSI state. At this point , however, it is not clear what the role of the suboxide is . A clue for further studies may possibly be found

in the contribution present around 2.1 A in the EXAFS spectra of the Rh/Ti02 catalyst in the SMSI state.

Page 181: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 171 Chapter 7

During the past years, the majority of the literature dealing

with SMSI reported coverage (9-17 ). Since the results of this study

do not support complete coverage, we will focus on covering in the

following. Covering of metal particles by an oxide can take place in

two ways. The covering oxide may have a chemical interaction, or

may have no interaction at all with the metal particles. When there

is an interaction between covering oxide and metal particle, the

oxide will 'wet' the metal surface. Most authors see this kind of

covering as the origin for the decrease in adsorption capacity of

metal particles in the SMSI state. Such a covering would greatly

enhance the number of 0 2- neighbors for the metal atoms in the

surface of the metal particles . Since in such a case there is an

attractive interaction between oxide and metal particle (and because

the state of coverage has been formed at high temperatures), the

disorder in these Rh0-02- bonds that are formed in the SMSI state

is expected to be small. Therefore, EXAFS should be able to detect

such bonds. For instance , in the 5 atom metal particle in Figure 7.9,

each rhodium atom is a surface atom. In the case of complete cov­

erage, the interfacial rhodium atoms should have seven (three more

than in the 'normal' state) 0 2- neighbors and the top rhodium atom

should have eight new 0 2- nearest neighbors. Therefore, the Rh0-

02- coordination number should increase at least by a factor of 2 or

3. For the Rh/ Al 20 3 catalyst reduced at 623 K and oxidized at

room temperature, the Rh0-Rh0 coordination number decreased by a

factor 3, the Rh0-02- coordination number by a factor 1.2 (cf. Table 7.2). Relative to the Rh0-Rh0 coordination number, the Rh0

-

02- coordination number increased by a factor of 2.5. From this we

concluded that the metallic kernel of the oxidized particles was

covered with rhodium oxide and it proves that EXAFS can detect a

covering oxide. For the Rh/Ti02 catalyst in the SMSI state we cer­

tainly did not find any evidence for such an increase in Rh0-02-

coordination number . In this study, after reduction at 723 K. the

Rh0-02- coordination number increased from 1.3 ± 0.3 to 1.9 ± 0.3. In our previous study, we found a small decrease in the Rh0-0 2-

coordination number . The analysis of the EXAFS spectrum of the

Rh/Ti02 catalyst reduced at 473 K was hampered by the presence

Page 182: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 172

of Rh20 3. Therefore. the Rh0-02- coordination number for this

catalyst is rather inaccurate (after correction. this coordination

number is 1.3) and. in our view, too low. Therefore. the apparent

increase in the Rh 0..,o2- coordination number when increasing the

reduction temperature is insignificant and certainly too small to jus­

tify the conclusion that the metal particles are tightly covered with

TiOx . Based on these Rh0-02- coordination numbers. one could

state that perhaps a small (partial) covering of the rhodium metal

particles has taken place. It is also possible that a very loose cover­

ing occurs: this induces only a small number of additional 0 2-

neighbors. In the following paragraph we will discuss the effects of

such a loose covering.

When there is no interaction between metal and support. wet­

ting and tight packing of the oxide around the metal particles will

not take place. Coverage then simply means the physical presence

of oxide crystallites over the metal particles. The spread in the

Rh0-0 2- distances will then be large, which makes it impossible for

EXAFS to detect the oxide (38). Based on the EXAFS results . we

therefore cannot exclude the presence of a loosely bound TiOx

suboxide on top of the metal particles. Such a coverage does not

have to be complete. but may still be able to suppress adsorption of

hydrogen and carbon monoxide .

From the oxygen adsorption experiments we learned that sur­

face metal atoms are exposed . Our results therefore exclude com­

plete coverage for rhodium metal particles supported on Ti02 in the

SMSI state. The metal particles may. however, be partially covered

with TiOx if this covering oxide does not tightly adhere to the metal

surface. But such a loosely packed covering oxide. which does not

interact with the metal particles. cannot be responsible for the

observed behavior of the SMSI state catalyst during oxygen expo­

sure.

Page 183: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 173 Chapter T

We feel that a loose coverage may occur in the SMSI state and

may also explain the difference in catalyst performance between

metals supported on the traditional supports like A1 20 3 and on

reducible supports . (see for example ref (39 )) . It is our conviction,

however. that coverage alone cannot explain the large decrease in

adsorption capacity (usually one order of magnitude) of the rho­

dium metal particles in the SMSI state in the discussed Rh/Ti02 catalyst. From this and many other studies in this field (7 ,10-13,16-20) it is clear that SMSI is a complex phenomenon and may

vary from system to system. It may well prove impossible to con­

dense this complexity into one model which cari explain all the

results presented up to now in the literature on SMSI. Rather, the

explanation for SMSI may vary from system to system. The

present study indicates that coverage is a phenomenon unlikely to

occur in Rh/Ti02. Although , based on the EXAFS results , coverage

cannot completely be excluded, it is evident that if a covering oxide

is present, the interaction between this covering oxide and the sur­

face metal atoms is weak. But such a loose packed weakly interact­

ing oxide cannot account for the fact that oxidation was greatly

suppressed when the surface metal atoms were exposed to oxygen.

Therefore, an electronic perturbation is the most likely explanation

for the anomalous properties of these small rhodium metal particles

supported on Ti02 in the SMSI state.

7 .5 References.

1. Tauster , S. J .; Fung, S. C. ; Garten , R.L. J. A m . Chem. Soc. 1978, 100, 170.

2. Tauster , S J.; Fung, S. C. J. Catal. 1978, 55, 29.

3. Tauster , S. J .; Fung, S. C. ; Baker , R. T. K .; Horsley , J . A. Science (Washington , D .C.) 1981 , 211, 1121.

4. Horsley , J. A. J . Am. Chem. Soc . 1979, 101 , 2870.

Page 184: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 174

5. Baker, R. T. K.; Prestridge, E. B.; Garten, R. L. J. Catal. 1979, 56,

390.

6. Baker, R. T. K.; Prestridge, E. B.; Garten, R. L. J. Catal. 1979, 59,

293.

7. van 't Blik, H. F. J.; Vriens, P. H. A.; Prins, R. ACS Symposium

1985, 298, 60:

8. Meriaudeau, P.; Ellestad, 0. H.; Dufaux, M.; Naccache C. J. Catal.

1982, 75, 243.

9. Meriaudeau, P.; Dutel, J. F.; Dufaux, M.; Naccache C. 11

Studies of Surface S~ience and Catalysis 11 1982, 11.

10. Belton. D. N.; Sun, Y. -M.; White, J. M. J. Phys. Chem. 1984, 88,

1690.

11. Belton, D. N.; Sun, Y.. -M.; White, J. M. J. Phys. Chem. 1984, 88,

5172.

12. Simoens, A. J.; Baker, R. T. K.; Dwyer, D. J.; Lund, C. R. F.; Madon, R. J. J. Catal. 1984, 86, 359.

13. Chung, Y. M.; Xiong, G.; Kao, C. C. J. Catal. 1984, 85, 237.

14. Sadeghi, H. R.; Henrich, V. E. J. Catal. 1984, 87, 279.

15. Sun, Y. -M.; Belton, D. N.; White, J. M. J. Phys. Chem. 1986, 90, 5178.

16. Ko, G. S.; Gorte, R. J. J. Catal. 1984, 90, 59.

17. Raub, G. B.; Dumesic, J. A. J. Phys. Chem. 1984, 88, 660.

18 Kelley, M. J.; Short, D. R.; Swartzfager, D. E. J. Catal. 1983, 20, 235.

19. Beard, B. C.; Ross, P. N. J. Phys. Chem. 1984, 90, 6811.

20. Brewer, L. 11 Phase Stability in Metals and Alloys 11; Rudman, P.;

Stringer, J.; Jaffee, R., ed.; MacGraw-Hill : New York, 1967; pp 39-61.

21. Sinfelt, J. H.; Via, G. H.; Lytle, F. W. J. Chem. Phys. 1977, 67, 3831.

22. van 't Blik, H. F. J.; van Zon, J. B. A. D.; Huizinga, T.; Konings­berger, D. C.; Prins, R. J. Phys. Chem. 1983, 87, 2264.

23. Koningsberger, D. C.; Martens, J. H. A.; Prins, R.; Short, D. R.; Sayers, D. E. J. Phys. Chem. 1986, 90, 3047.

24. Kip, B. J.; Duivenvoorden, F. B. M.; Koningsberger, D. C.; Prins, R. J. Catal. 1984, 105, 26.

Page 185: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 175 Chapter 7

25. Vis, J. C.; van 't Blik, H. F. J.; Huizinga, T.; van Grondelle , J .: Prins, R. J. Mal. Catal. 1984, 25, 367.

26. Cook, J W.; Sayers, D. E. J. Appl. Phys . 1981 , 52 , 5024 .

27 Crystal structures :

Rh metal Wycklwff Crystal Structures 1963, 1, 13.

Rh 20 3 Structure Reports 1974, 40a, 301.

RhTi Structure Reports 1964, 29 , 13.

28. van Zon, J. B. A. D.; Koningsberger, D. C. ; van 't Blik, H. F. J.;

Sayers, D. E. J. Chem. Phys . 1985, 12, 5742.

29. Duivenvoorden, F. B. M.; Koningsberger , D. C.; Uh, Y S.; Gates, B. C. J . Am. Chem. Soc. 1986 , 108, 6254.

30. van 't Blik, H. F. J. ; van Zon, J. B. A. D.; Huizinga, T.; Vis, J. C.;

Koningsberger, D. C.; Prins, R. J. Amer . Chem . Soc. 1985 , 107, 3139.

31. Zandbergen, H. W.; Martens, J. H. A . (to be published)

32. Koningsberger, D. C.; van Zon, J. B. A . D.; van 't Blik, H. F. J.; Prins,

R.; Mansour, A . N.; Sayers, D. E.; Short, D. R.; Katzer, J. R. J. Phys . Chem. 1985, 89, 4075 .

33. Stern, E. A .; Bunker , B. A.; Heald , S. M . Phys . Rev . B 1980, 21,

5521.

34. van 't Blik, H. F. J.; van Zon, J . B. A. D.; ~oningsberger, D. C.; Prins,

R. J . Mal . Catal . 1984, 25, 379.

35. Woning, J.; Santen , R. A. Chem . Phys . Leu. 1983, 101 (6), 541.

36. Sakelson, S.; McMillan, M.; Haller, G. L. J . Phys . Chem . 1985 , 90 ,

1733.

37. Teo, B. K .; Lee , P. A. J. Am. Chem. Soc. 1979, 101, 2815.

38. Eisenberger , P.; Brown, G. S. Solid State Comm. 1979, 29, 481.

39. Levin, M. E.; Salmeron, M.; Bell, A. T.; Somorjai, G. A. Far. Symp. Chem. Soc. 1986, 21, paper 10.

Page 186: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

EXAFS and HRTEM of Rh/Ti0 2 page 176

Page 187: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 177 Chapter 8

Chapter 8

Strong Metal-Support Interactions in Rh/Ti02 Prepared with Ion Exchange

8.1 Introduction

The special properties of small metal particles supported on reducible transition-metal oxides have attracted considerable atten­

tion during the last decade. As Tauster et al. already pointed out in their first papers ( 1-3), the properties of these catalysts after reduc­tion at low temperature resemble the properties of such metal parti­

cles supported on inert, unreducible support materials like Al 20 3 and Si02. After reduction at high temperature, however, the pro­

perties change drastically. Most pronounced is the decrease in hydrogen and carbon monoxide adsorption capacity, suggesting that

the exposed metal surface has decreased due to sintering. Electron Microscopy revealed, however, that the metal particle size remained

virtually unaffected during the high temperature reduction. Several models have been proposed to explain this discrepancy. The first model, adopted by Tauster et al. ( 1-3) and later by many other authors, assumes a strong electronic interaction between metal par­

ticle and the reduced support. Another model assumes covering of the metal particles by reduced support oxide species to be responsi­ble for the SMSI state of the catalyst (4-7 ). Alloy formation is

another model proposed to explain the incapability of adsorbing

hydrogen and carbon monoxide by metal particles in the SMSI state ( 1-3,8). In a very recent EXAFS study (9). Sakelson et al. reported

alloy formation for a Rh/Ti02 catalyst in the SMSI state.

Page 188: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 178

In two recent EXAFS studies ( 10,11) and in chapter 7 of this

thesis, we reported on the structure of a Rh/Ti02 catalyst in the

normal and in the SMSI state. In these studies, we reported that

metal particles in the SMSI state rested on reduced Ti02 and also, that there was no evidence for alloy formation or for coverage of the

metal particles by reduced TiOx species. Consequently, we con­

cluded that the decrease in hydrogen adsorption capacity and the

suppression of oxidation after reduction at 723 K was the result of

an electronic interaction between metal and support. On the other hand, Sakelson in the group of professor Haller (Yale University)

(9) reported alloy formation in the SMSI state, based on EXAFS

results for a similar Rh/Ti02 catalyst in the SMSI state. This

catalysts had been prepared with ion exchange, measured at CHESS

and analyzed with their own data analysis procedure. Thus. there was a serious problem. The results of Sakelson, alloy formation,

contradicted the results of our study, which clearly indicated no

alloy formation. The difference between their and our Rh/ Ti02 catalyst was the preparation method. The Rh/Ti0 2 sample in (9)

was prepared with ion exchange using ammonia, reduced and finally calcined, while the sample in our studies ( 10,l J) was prepared with

ion exchange without ammonia and was directly calcined. However,

also the EXAFS analysis procedure used by Sakelson differed from

the analysis procedure we used. Hence, the difference between the

two studies may be either the result of a different preparation

method or of a different analysis procedure. We reached an agree­

ment with professor Haller that we should try to solve this contrad­

iction. Professor Haller supplied us the raw EXAFS data of the

Rh/Ti02 sample of the Yale University which were used for the study in (9), in order to subject them to our data analysis pro­

cedure. In this chapter we will describe the results of the analysis

of these EXAFS data from Yale University using the Eindhoven

data analysis procedure.

Page 189: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 179 Chapter 8

8.2 Experimental

8.2.1 Catalyst Preparation

The catalyst was prepared at the Yale University using ion

exchange. The Ti0 2 was first immersed in a NH 40H solution at

pH= 11 for 15 h: this solution was stirred continuously. The Ti02 was then washed with distilled water until a pH of 7.5 was

obtained, filtered and dried for 24 h at 373 K. The ammonia treated

Ti0 2 was placed in a round bottom flask with water in a ratio of

60 cm3 g-1 Ti0 2. The temperature was raised to about 323 K and,

while vigorously stirring, 1 cm3 of a Rh(N03h solution was added

per gram Ti0 2. The solution contained 6.6 mg Rh cm-3 and was

added over a period of 5 h. Subsequently, the solution was allowed

to stir overnight at room temperature. The resulting catalyst was

centrifuged and washed several times with hot distilled water,

allowed to dry at room temperature for several days and then dried

for 5 h at 383 K. After reduction in flowing hydrogen at 773 k for

2 h and oxidation at 623 k for 2 h, the catalyst was stored in a

desiccator. The metal loading measured by HCI extraction and

atomic absorption was 0.47% and the H/Rh determined at Yale University with hydrogen chemisorption was 1.00± 0.05.

8.2.2 EXAFS Measurements

The X-ray absorption spectra of the rhodium K-edge of the

samples were measured by the Yale group at the Cornell high

energy synchrotron source (CHESS). All spectra were measured

with the sample cooled to liquid nitrogen temperature . The sam­ples were pressed into thin self supporting wafers of approximately

20 mg cm-2. Two of these wafers were placed into a cell which

Page 190: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 180

allows in situ reduction. The design of the cell is such that in case

of inleakage, the influx of gases into the catalyst section consisted

of pure helium. The hydrogen flow was high enough to ensure no significant dilution with helium. X-Ray absorption spectra of the

reference compound were measured as well. The EXAFS spectra of

the reference compounds, rhodium foil, Rh02 and RhTi were also

measured by the Yale group at CHESS. These EXAFS spectra were

used to extract backscattering amplitude and phase shift functions

for the Rh-Rh, Rh-0 and Rh- Ti absorber-scatterer pairs; using the Eindhoven data analysis procedure.

8.3 Results

The extraction of backscattering amplitudes and phase shift

functions for the absorber-scatterer pairs Rh- Rh, Rh-0 and Rh-Ti

has been described extensively in chapter 7. The crystallographic

data for the reference compounds were taken from ( 13) and are

summarized, together with Fourier transform ranges in Table 8.1. Briefly, the analysis procedure was as follows. The EXAFS func­

tion of the reference compound was Fourier transformed over the indicated range in k-space. An inverse Fourier transform over a lim­

ited range in r-space yielded the EXAFS function for the mentioned

absorber-scatterer pairs. From these functions, backscattering

amplitude and phase shift functions could be derived. For the Rh Ti reference compound, first a best fitting Rh- Rh contribution was cal­

culated and this EXAFS function was subtracted from the experi­

mental data . Froms this difference spectrum, the desired Rh-Ti

backscattering amplitude and phase shift function were acquired as

described above .

Page 191: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 181 Chapter 8

Table 8.1 Crystallographic data and Fourier transform ranges for the reference compounds

Rb Fourier transformation

a

b

(

d

Compound NN;i

Rh foil Rh 2.687

Rh02 0 1.963

Rh Ti Rh 2.949 Ti 2.676

Nearest Neighbor

Coordination Distance (A)

Coordination Number

Ne d n

12 3

6 1

4 3 8 1

Weighting factor in Fourier transformation

Crystallographic data were obtained from ( 12)

k-range r-range

2.81 - 14.50 1.52 - 3.02

2.49 - 16.98 0.00 - 1.84

2. 73 - 15.40 2.86 - 13.96 1.10 - 2.56

The procedure of analyzing the Rh/Ti0 2 EXAFS data has been described in chapter 7 and in ( 10, 11 ). Briefly, the procedure comprised the following steps. For all spectra , the Rh-Rh contribu­tion was dominant. A k 3-weighted Fourie~ transform was used to

obtain a first estimate for the Rh- Rh parameters (N , R , !:!:.a 2 and t:J.£ 0) . Using these parameters , a Rh- Rh EXAFS function was cal­

culated and subtracted from the experimental data. The resulting difference file was used to analyze the residual Rh-0 and Rh- Ti con­

tributions . The calculated best fitting Rh-( O+ Ti) EXAFS function

was subtracted from the experimental data and the resulting

difference spectrum, now containing mostly Rh-Rh contribution, was used to optimize the Rh-Rh parameters . In this way, a recurrent optimization process was started that converged to the final set of parameters as presented in Table 8.2. In Figure 8.1 are shown the raw EXAFS spectra , their k 1-weighted Fourier transforms and the Fourier transforms of the best fitting Rh-Rh

Page 192: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 182

Table 8.2 Final results from EXAFS data analysis

Treat- Coordination Distance A<r L D AEo D

ment NN number (Al fal (* 10-3 A,-2) (eV)

(a) (a) (a)

R494 Rh 2.8 0 .2 2.63 0.01 5.0 1 -105 2 0 1.2 0.2 2.00 0.01 5.0 2 -5.0 2 0 2.2 0.3 259 0.02 5.0 2 5 ~0 2

Ti 1.3 0.3 3.54 0.05 3.0 2 3.0 2

R628 Rh 4.6 0.2 2.668 0.005 9.8 2 2.2 1 0 0.7 0.3 2.06 O.D1 5.0 2 -18 4 0 0.9 0.2 2.77 0.02 5.0 2 -13 4

Ti 1.1 0 .3 353 0.05 5.0 2 15 5

R773 Rh 4.9 0 .2· 2.650 0.005 5.3 2 5.0 1 0 2.0 0.3 2.51 0.01 2.0 2 4.5 3

Ti 2.0 0 .3 3.51 0.05 5.2 2 25 2

R : Reduction in H2 at the temperature indicated

a Estimated overall (experimental+ systematic) error b

fj,a 2, the Debye Waller factor , is a measure for the disorder and E0 is a

correction on the edge position .

contributions . In Figure 8.2, the k 1-weighted Fourier transforms of the difference spectra and the best fitting Rh- (O+ Ti) EXAFS func­tions are shown as well as the k1-weighted Fourier transforms of

the raw data and the best fitting Rh-( Rh+O+ T i) EXAFS functions . Clearly, even in the difference files (Figures 8.2a-c) , the calculated spectra are in excellent agreement with the measured spectra, demonstrating the reliability of the results .

Page 193: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 183

*10-2

3

-3

0 *10-2

2

-2

a

5 10 15

b

Chapter 8

* 10- 1

4 -.---~-~----,---r----,.--..

d

-4 0 2 4 6

* 10-1

5

e

-5 ~---'---4------'-~-...__----l 0 5 10 15 0 2 4 6

* 10-2 * 10- 1

5 -r--"r---Y--r-T""--r-1r-r--r-T-r-.,...-,----y--,--, 7~~-~~-~-~~

c f

-5 4--1-_,__..__._--+--''--'-~~..__.__,__..__. - 7 +------'---+-------'---+----'--~

0 5 10 15 0 2 4 6

k r A-1 J R [A]

Page 194: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 184

Figure 8.1 Raw EXAFS data for •

(a) Rh /Ti02 after reduction at 494 K

(b) Rh/Ti02 after reduction at 628 K

( c) Rh/Ti02 after reduction at 773 K

Imaginary parts of the k1-weighted Rh-Rh corrected Fourier

transforms of the raw EXAFS functions (solid lines) and the

best fitting calculated Rh- Rh contributions (dotted lines) for

(Fourier transform ranges are indicated in brackets) •

(d) Rh/Ti02 after reduction at 494 K (3.15 - 9.95 A-1)

(e) Hh/Ti02 after reduction at 628 K (3.38 - 10.74 A- 1)

(f) Rh/Ti02 after reduction at 773 K (3.38 - 10.83 A- 1)

8 .4 Discussion

8 .4.1 Rh/Ti02 after Reduction at 494 K

The Rh-0 contribution at 2.0 A points to the presence of unre­

duced rhodium oxide. Using the procedure described in ( 14), we

estimate that about 20% of the rhodium was not reduced . Since

the coordination numbers measured with EXAFS are average coor­

dination numbers, we have to correct the measured Rh-Rh coordina­

tion number in order to obtain the actual size of the metal particles.

Using the procedure described in ( 14), we find that the corrected

coordination number N< is equal to 2.8/0.8 = 3.5. Hence, the metal

particles were very small and of about the same size as those

reported in chapter 7 and in ( 10,11) (N = 3.3, particle diameter typ­

ically about 7 A containing approximately 4 to 7 rhodium atoms) .

Page 195: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 185 Chapter 8

* 10-2 * 10-1

3...-------.--.-------..--.-----.-~ 4 .,....--.,...--.-----.----r--..------.

a d

-3+-~--t-~--+----'-~ -4+-~--+-~--1-----'--'-~

0 2 4 6 0 2 4 6 *10-2 *10- 1

5~~-~~-~~-~

2 e

-2 -5

0 2 4 6 0 2 4 6 * 10-2 * 10-1

3 7

c f

-34"--~--+-~--+----'-~ -7+-~--t-~--+----'-~

0 2 4 6 0 2 4 6

R [A] R [A]

Page 196: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti02 ex ion exchange page 186

Figure 8.2 Imaginary parts of the k1-weighted Rh-Rh corrected Fourier transforms of the difference files (raw EXAFS minus calculat­ed Rh-Rh contribution , solid lines) and the best fitting calcu­lated Rh-(0+ Ti) contributions (dotted lines) for (Fourier transform ranges are indicated in brackets) :

(a) Rh/Ti02 after reduction at 494 K (4.00- 9.00 .&.- 1)

(b) Rh/Ti02 after reduction at 628 K (3.33 - 8.75 .&.-1)

(c) Rh/Ti02 after reduction at 773 K (3.38 - 9.00 .&.-1)

Imaginary parts of the k1-weighted Rh-Rh corrected Fourier transforms of the raw EXAFS functions (solid lines) and the best fitting calculated Rh-(Rh+O+ Ti) contributions (dotted lines) for (Fourier transform ranges are indicated in brackets)

(d) Rh/Ti02 after reduction at 494 K (3.15 - 9.95 .&.-1)

(e) Rh/Ti02 after reduction at 628 K (3.38 - 10.74 .&.-1)

(f) Rh/Ti02 after reduction at 773 K (3.38 - 10.83 .&.- 1)

Apart from the Rh- 0 contribution at 2.0 A, a second Rh-0

contribution at 2.59 A was observed. This contribution can be ascribed to rhodium metal atoms having oxygen anions as neigh­

bors . This kind of metal-oxygen contribution has been reported

before (see chapter 5 and ref 14-17). The oxygen ions may ori­

ginate from the support and, in this case, also from unreduced rho­

dium oxide which is in contact with the rhodium metal particles

(see chapter 5 and ref 15) . This can be concluded from the follow­ing observations . First of all. the accompanying Rh0-02

- coordina­

tion number is relatively high . The corrected Rh0-0 2- coordination

number is 2.2/0.8 = 2.8. For similar metal particles we reported a Rh0-0 2

- coordination number of 1.9 ( 1 J). Secondly. the Rh-Rh dis­

tance (2.630 A) was contracted significantly with respect to the dis­

tance in bulk rhodium metal (2.687 A) and the Rh-Rh distance of the sample reduced at 628 K (2.668 A) . Such a decrease in Rh-Rh

distance has been ascribed to the absence of adsorbed hydrogen on the surface of the metal particles ( 18). This may be induced by the

SMSI state or may be the result of the presence of oxide species on

Page 197: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 187 Chapter 8

top of the metal particles (see chapter 5 and ref ( 15)) . Although

some neighboring Ti ions could already be observed at 3.54 A, we

assume that the catalyst was not in the SMSI state because the

number of neighboring Ti ions was relatively low (in chapter 7 and in ref ( 11 ), coordination numbers ranging from 2.5 to 3.3 were

reported for the sample in the SMSI state). In addition, the reduc­

tion temperature was rather low (SMSI behavior is observed usually

after reduction at temperatures above 600 K) . As reported in

chapter 7 and in ref ( 11 ), the Ti neighbors at 3.5 A are Ti ions in

the support underneath the metal particles. The coordination

number of 1.3 found in this study corresponds nicely with the value

based on the model described in chapter 7 and in ref ( 11) (for a 5

atom metal particle resting on unreduced Ti02, a Rh0-Ti coordina­

tion number of about 1.5 may be expected). Since the measured and expected Rh0-Ti coordination numbers in the normal state

agree very well. we conclude that the particles fit well on the sup­

port according to the model described in chapter 7 and in ref ( 11)

(see Figure 8.3) . Since the catalyst is not in the SMSI state, we conclude that the absence of hydrogen on the surface of the metal

particles, leading to a decrease in the Rh-Rh distance, is the result

of rhodium oxide species on top of or in contact with the metal par­

ticles. In conclusion, after reduction at 494 K the situation was as

follows : about 80% of the rhodium was reduced . The metal parti­

cles contained about 5 atoms and rested on unreduced Ti0 2.

Finally, the particles were covered or at least in contact with unre­

duced rhodium oxide.

8.4.2 Rh/ Ti02 after Reduction at 628 K

After reduction at 628 K, the Rh-Rh coordination number has increased and the Rh3+-02

- contribution at 2 A. has decreased. The

estimated amount of unreduced rhodium oxide is about 10%. Therefore, the corrected Rh-Rh coordination number 1s

Page 198: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 188

Figure 8.3 A 13 atom metal particle with N=4.9 , resting on a anatase

1001] face .

ORh 0Ti4

+

( .Ti 4+ under Rh

4.6/0.9 = 5.1. corresponding to metal particles containing about 15 rhodium atoms (I 2). The amount of rhodium oxide still present

after reduction at 494 K was about 20%, while the metal particles contained approximately 4 to 7 rhodium atoms . This means that

for every metal particle, or for every 4-7 rhodium atoms in the metallic state , there will be one or two rhodium ions in the oxidic

state. After complete reduction of this rhodium oxide phase, the metal particles would, on the average, contain about 5 to 9 atoms,

which is clearly not the case : the corrected coordination number of 5.1 points to particles with 15 metal atoms . Obviously, some

sintering has occurred during further reduction. Since the catalyst had been pre-reduced at 773 K in the preparation, one might con­

clude that during subsequent reduction experiments, no sintering should occur at reduction temperatures lower than 773 K. Neverthe­

less, we observed sintering even at 628 K. This can be explained as described by Wang and Schmidt for Ir /Si02 (/9 ) . After pre­

reduction at 773 K, metal particles are formed. During subsequent calcination, rhodium oxide is formed . This oxide phase may break

up into several smaller oxide particles. These smaller oxide parti­cles are close together, but not necessarily in intimate contact. Upon reduction at lower temperatures, these smaller oxide particles

Page 199: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 189 Chapter 8

will be reduced to small metallic particles . Since these metal parti­

cles are not far apart, sintering will take place below the tempera­ture of 773 K. After the sintering process has stopped, the metal

particle size wil I be equal or even larger than the metal particle size after the first reduction treatment at 773 K.

The Rh-Rh distance in the metal particles after reduction at 628 K (2 .668 A) has relaxed towards the Rh-Rh bulk distance

(2.687 A), indicating that there was no more coverage of rhodium oxide and that the metal particles had adsorbed hydrogen. Hence,

also this catalyst was not in the SMSI state. Because of the larger metal particles, the Rh0-02

- and Rh0-Ti coordination numbers can­

not be compared directly to the corresponding coordination numbers for the smaller metal particles . Therefore, we take as a model a 13

atom rhodium metal particle which has 9 atoms in the metal­support interface and 4 atoms on the top. In Figure 8.3, this metal

particle on a anatase [001) crystal face is shown. The 9 interfacial metal atoms constitute a 3*3 square and can be fitted adequately

on the Ti02 support (see the models in Figure 7.9 in chapter 7). The calculated coordination number for this particle is 4.9. When the Ti0 2 is not reduced, we may expect a Rh0-02

- coordination number of 2.5 and a Rh0-Ti coordination number of 1.5. Both coor­

dination numbers measured are significantly lower. Hence, we con­clude that the metal particles did not fit 'perfectly' on the support .

This may be the result of the sintering process : from the coordina­tion numbers we may estimate that two to three rhodium particles

of about five atoms each, have sintered to one particle of about 15 atoms. Because of the low temperature, the particles may not have

reorganized completely and therefore. the structure of these parti­cles may be imperfect and will not fit adequately on the support.

Confirmation for this can be found in the Debye-Waller factor , the disorder for the Rh-Rh contribution (see Table 8.2) . For the sample

reduced at 628 K, the Debye-Waller factor is higher by about a fac­tor of two compared to the samples reduced at 494 and 773 K. Concluding, after reduction at 628 K, the catalyst was not in the SMSI state. The particles had sintered and were covered with

Page 200: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 190

adsorbed hydrogen . The reduction process was not complete .

8.4.3 Rh/Ti02 after Reduction at 773 K

After reduction at 773 K, there was no evidence for unreduced

rhodium oxide. Clearly, reduction was complete. The Rh-Rh coor­

dination number was 4.9, in excellent agreement with the corrected

coordination number measured after reduction at 628 K and indicat­

ing that no additional sintering had occurred during reduction between 628 and 773 .K. The Rh 0-0 2

- and Rh0-Ti coordination

numbers both increased significantly. In addition, the Rh0-0 2- coor­

dination distance decreased markedly. We ascribed this decrease to

the interaction between metal and support (see chapter 7 and ref ( 11 ). Because of this enhanced interaction and because of the

higher reduction temperature, the metal particles had reorganized

(the Debye-Waller factor had decreased) and fitted better on the

Ti0 2 support. The Rh0-Ti coordination number is higher than for a

metal particle on unreduced Ti0 2 (from the model described above,

this was estimated to be about 1 .5). Obviously, as reported in

chapter 7 and in ref ( 11 ), the particles rested on reduced Ti0 2. In

addition, the Rh-Rh distance has again decreased . Hence, the sam­ple was in the SMSI state and the metal particles had not adsorbed

hydrogen. We found no titanium neighbors at metallic distances.

Thus, alloy formation can be ruled out as an explanation for the

SMSI state. Also, the measured Rh0-0 2- coordination number (2.0)

was lower than the expected coordination number for the 13 atom

metal particle ( ± 2.8), indicating that the particles were not covered

with reduced TiOx. The fact that the measured coordination

number was lower than the expected coordination number has been

explained in ( 11) as being the result of the fact that the reference

Rh-0 distance (1.963 A) was shorter than the calculated Rh-0 dis­tance (2.51-2.77 A) , giving rise to an error in the mean free path

term in the EXAFS equation ( 16). Hence, we cannot but conclude

Page 201: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 191 Chapter 8

that an electronic interaction is responsible for the SMSI state.

8.4.4 General Remarks

The results presented above agree nicely with the results of

our earlier paper ( 11 ). However, there are differences. First of all,

even after reduction at 628 K, the sample was still not completely

reduced. The reason for this may be found in the preparation method. The support was immersed in water and stirred vigorously

several times. Consequently, the support has been powdered very

finely. As a result , after exchanging ammonium groups with rho­

diu m nitrate, the support had to be centrifuged in order to separate

it from the solution . When such a finely powdered support is

pressed into a wafer , the wafer may be very compact and gases will

diffuse only slowly through the sample. Consequently, hydrogen will

diffuse only slowly into the catalyst wafer and the water formed

during the reduction process will diffuse only very slowly out of the

catalyst sample. Hence , the reduction process will continue very

slowly and unreduced rhodium oxide may be found in the sample

even at relatively high reduction temperatures.

Another difference can be found in th~ Rh0-Ti contributions. In

the present study we found only one contribution clearly present :

the 3.5 A contribution. This distance is within the experimental

error equal to the distances reported in (10,JJ). In (JJ) , however ,

an additional contribution at 4.3 A was found . This contribution

has not been observed clearly in the present sample. In the difference spectra in Figure 8.2a-c , however , small differences at

about 4.3 A are still visible and these could be fitted with Rh0-Ti

contributions, but the accompanying coordination numbers were

Very small , N < 0.4, in contrast to 1 .4 < N < 2.8 in ( 11 ). Although the particles were larger, we would still expect this contri­

bution to be present . However, as argued above, the reduction pro­cess may have been delayed . As a consequence. the reduction of

Page 202: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 192

the support may have been limited to the direct environment of the

metal particle , whereas in ( 11 ), the support was reduced to a larger

extent . This is the result of the more open structure of the support

used in ( 11) (the Ti02 in ( 11) has not been stirred using a mag­

netic bar) . Hence, the reduction process of both metal particle and

support could proceed more easily and thus. the SMSI state could

be invoked at lower temperatures . Consequently. the binding of the

metal particles to the (reduced) support was strong already at low temperatures. This explains why we did not observe any sintering

after reduction at 773 K in chapter 7 and in ( 11) (for that sample,

no pre-reduction has been applied), whereas in this study sintering

was already observed at 628 K, notwithstanding the higher metal

loading (4 wt% in ( J J) and 0.47 wt% in this study).

Now we come to the most important part of this study.

Clearly, our data analysis procedure produces results completely

different from the results obtained by the procedure followed by

Sakelson. The explanation for this discrepancy is in fact very sim­

ple. Sakelson used a k 3-weighted Fourier transform to study the

data. The transform range was in all cases choosen from k=3.0 to

15.0 A_-l and this is the major reason for the differences in the

results. Above k=11 A-1 the noise level increases drastically. For

the samples reduced at 494 and 628 K even a few glitches and jumps had to be removed from the raw data in order to obtain reli­

able EXAFS functions. Anyhow, above k=11 A.- 1, the data are

unreliable . This can be seen in Figure 8.4, in which we have plotted

the EXAFS function of the sample reduced at 773 K, weighted by

k 3 and the k 3-weighted EXAFS function which fitted these data

between k = 3.38 and 10.83 A-1 and was calculated using the param­

eters from Table 8.2. Note. that these functions are in fact the

functions that are transformed in a k 3-weighted Fourier transform.

The differences are clearly visible and these differences are the result of the noise in the spectrum above k=11 A_-l. As a result , a Fourier transforms up to k=15 A_-l will result in a strongly distorted

radial distribution function and will inevitable result in incorrect parameters.

Page 203: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 193 Chapter 8

Figure 8.4 k3*CHI for the experimental data of the sample reduced at 773 K (solid line) and the calculated best fitting EXAFS func­

tion (dotted line)

13

9

-----..._ 5 ~ '---

f-j 1

J: lJ * -3

('T)

~ -7

-11

-15 0 5 15

k

In general, next to the potential errors indeuced by choosing

wrong Fourier transform ranges, many more pittfals are present in

the data analysis procedure. One can think of removing glitches

and jumps and background subtraction. Another very important

aspect is imposed by the reference compounds. Even more care

should be taken during the data analysis procedure, for backscatter­

ing amplitudes and phase shift functions obtained from reference

compounds are especially sensitive towards incorrect background

subtraction and Fourier transform ranges. And unreliable reference

spectra will only add to the unreliability of the analysis of the spec­tra of the catalyst samples. Obviously, analyzing EXAFS spectra

should be carried out with the utmost caution and only a data

analysis performed with such utmost care will result in reliable

parameters.

Page 204: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSI in Rh/Ti0 2 ex ion exchange page 194

8.5 Final Conclusions

After reduction at 494 K, reduction was not complete and very

small metal particles were formed, containing about 5 atoms. Because of their small dimensions, these particles fitted easily on

the support. Unreduced rhodium oxide covered the metal particles.

After reduction at 628 K, still 10% of the rhodium oxide was not

reduced: the particles were not in the SMSI state. The binding of the metal particles to the support was relatively weak. Conse­

quently, the ·metal particles had sintered and the larger metal parti­

cles did not fit perfectly on the support. After reduction at 773 K.

the reduction process was complete and the catalyst was in the SMSI state: the supp.ort underneath the metal particles was

reduced . Because of the stronger binding of the metal particles to the support in the SMSI state, the metal particles had rearranged

with respect to those after reduction at 628 K and fitted more per­fectly on the support . Because of the deferred reduction process,

the SMSI state was invoked only at relatively high temperature. In the SMSI state no evidence for alloy formation nor for coverage was

found . Thus, we conclude that an electronic interaction between metal particle and reduced support is responsible for the SMSI state

with particles of the size as studied in this chapter .

The differences between the results of the data analysis fol­lowed by Sakelson in (9) and the data analysis procedure used at the Eindhoven University could be ascribed to the increased noise

level at higher k-values and an incorrect choice of Fourier transform ranges .

Page 205: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 195 Chapter 8

8 .6 References.

1 Tauster . S J; Fung, S C.; Garten, R.L. J. Am Chem . Soc . 1978. JOO. 170

2 Tauster . S J.; Fung. S. C ] . Catal. 1978 . 55, 29

3 Tauster . S. J; Fung, S C. ; Baker , R T. K.; Horsley, J . A Science

{Washington, D.C.) 1981. 211, 1121 .

4. Meriaudeau, P.; Dutel , J F ; Dufaux. M.; Naccache C "Studies of Surface Science and Catalysis" 1982. 11 .

5 Belton . D. N : Sun . Y. -M; White , J . M J . Phy.1. Chem . 1984. 88,

1690.

6 Simoens. A. J.; Baker. R. T. K: Dwyer , D J; Lund , C RF ; Madon . R. J. J. Catal . 1984 . 86. 359.

7. Sadeghi. H. R.; Henrich , V E. J . Catal. 1984. 87, 279.

8 Beard , B. C. ; Ross . P. N. J. Phys. Chem . 1984. 90 . 6811 .

9. Sakelson , S .; McMillan , M.; Haller, G. L. J . Phys . Chem . 1985, 90, 1733

10. Koningsberger, D. C.; Martens , J. H. A.; Prins, R. ; Short, D R : Sayers, D. E. J . Phys . Chem. 1986, 90 , 3047.

11. Martens, J . H. A ; Prins, R. ; Koningsberger, D. C. J . Phys . Chem.,

accepted for publication .

(chapter 7 of this thesis)

12. Kip, B. J.; Duivenvoorden, F. B. M.; Koningsberger , D. C.; Prins , R J. Catal. 1987, 105 , 26.

13. Crystal structures •

Rh metal Wyckhoff Crystal Structures 1963. 1 . 10.

Rh0 2 Structure Report s for 1968 1975 , 33a , 271 .

RhTi Structure Reports f or 1964 1972 , 29 , 13

14 Koningsberger, D. C.; van Zon , J . B. A D; van 't Blik, H.F . J .; Man­sour. A N ; Visser, G. J .; Prins , R; Sayers , D. E.; Short, D. R.; Katzer , J . R. J . Chem . Phys . 1985 , 89, 4075

15. Martens , J H. A ; Prins , R ; Koningsberger , D. C. J . Phys . Chem.,

submitted for publication

Page 206: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

SMSl _in Rh/Ti02 ex ion exchange page 196

(chapter 5 of this thesis)

16. van Zon, J. B. A. D.; Koningsberger, D. C.: van ·t Blik, H. F. J.; Sayers, D. E. J. Chem. Phys. 1985, 12, 5742.

17. Koningsberger, D. C.; Duivenvoorden, F. B. M.; Kip, B. J .; Gates, B. C. "EXAFS and Near Edge Structure"; Lagarde, P.; Raoux, D.; Petiau, J. Eds.; Les Editions de Physique, 1986; vol. 1, p. C8-255 .

18. van 't Blik, H. F. J.; van Zon, J. B. A. D.; Koningsberger, D. C.; Prins, R. J. Mol. Catal . 1984, 25, 379.

19. Wang, T.; Schmidt, L. D. J. Catal . 1980, 66 , 301

Page 207: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 197 Chapter 9

Chapter 9

EXAFS Evidence for Direct Rh0 -Ta"+ Bonding and Coverage of the Metal Particles

in a Rh/Ta20 5 Catalyst in the SMSI State

9.1 Introduction

Since the discovery of the phenomenon of Strong Metal­Support interaction (SMSI) in the late seventies ( J-3), a lot of research effort has been devoted to the behavior of various metal catalyst systems in this SMSI state and to the search for the expla­nation of the anomalous properties of such catalysts. Shortly, the SMSI state can be defined as follows : after reduction at lower tem­peratures, the properties of metal particles supported on selected transition metal oxide supports can be classified as 'normal'. The

capacity to adsorb hydrogen and carbon monoxide is one of those properties. After reduction of the catalyst at higher temperatures, the capacity to adsorb hydrogen and carbon monoxide diminishes drastically. This reduced adsorption capacity is always observed for such catalyst systems and is therefore generally used to define the state of a catalyst.

In two recent studies ( 4,5) and in chapters 7 and 8, we reported on the structure of a titania supported rhodium catalyst in the normal and the SMSI state. From these studies it became evi­dent that in the SMSI state alloy formation (which is one of the proposed explanations for SMSI) had not taken place. Although it was found that coverage of the metal particles by reduced support species (coverage or decoration, another explanation for SMSI that has been proposed in the literature) was unlikely to occur, coverage

Page 208: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 198

could not be excluded completely. Oxygen adsorption experiments

indicated that oxidation was suppressed, but that the metal parti­cles were exposed to the gas atmosphere and were covered with

absorbed gaseous oxygen ( 5). A Rh/ A120 3 catalysts become oxi­dized under these conditions (see chapter 5). Thus, if the metal par­

ticles in the Rh/Ti02 sample in chapter 7 were covered with a TiOx suboxide, this coverage was not complete and could not explain the

behavior of the catalyst in the SMSI state. In the EXAFS spectra

of the catalyst in the SMSI state, contributions from neighboring

titanium ions at 3.4 and 4.3 A were present. It was concluded that

the metal particles rested on reduced titania. Since this was the

only change in the EXAFS spectra after reduction at high tempera­

ture, it was suggested that there was an interaction between the

supporting oxide underneath the metal particles and the metal parti­cles themselves. In a model of the structure of small rhodium metal

particles on (001J and (101J anatase crystal faces after reduction at high temperature, Rh0-Ti"+ coordination distances were observed

which were equal to those present in the EXAFS spectra ( 5 ). How­ever, in this model titanium ions at about 2 A are present as well.

The number of titanium neighbors at 2 A is lower than the number of titanium neighbors at 3.4 and 4.3 A by more than a factor 2.

Because of this , and because of the fact that titanium has a rela­tively low backscattering amplitude, the contribution of these

titanium ions is expected to be small. In practice, in the Fourier

transform of the EXAFS spectra of the rhodium catalyst in the

SMSI state, a small contribution was present at the low r-side of

the main contributions. Although this contribution could be fitted

with a Rh0-Ti"+ contribution at about 2 A, the contribution was too low to allow a reliable analysis.

In the present study we have tried to observe these support

cations at short distance by modifying the Rh/Ti0 2 system. The

coordination number is a quantity which we do not have in hand. However, by choosing another cation than that of titanium, we can

manipulate the backscattering amplitude and use it to our advan­

tage. In this study we will describe the results of an EXAFS study

Page 209: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 199 Chapter 9

of a Ta 20 5 supported rhodium catalyst. We choose Ta 20 5 because

it is known to be an SMSI support ( 2) and, more importantly, because tantalum has a backscattering amplitude which is higher by

a factor 2 to 4 than that of titanium at higher k-values . In Figure 2.5 (chapter 2), the backscattering amplitudes for Rh, Ta and Ti

according to Teo and Lee (6) are shown . Clearly, at higher k-values, the contribution of tantalum is much more pronounced than that of

titanium. If tantalum ions are present at short distances in the SMSI state, we may expect that EXAFS will be capable of detecting such contributions.

9.2 Experimental

9.2.1 Catalyst Preparation

A high surf ace area T a20 5 support was prepared according to

the following procedure. 20 g T aCl 5 was dissolved in a 100 ml con­centrated HCI solution. The resulting solution was added carefully

to a mixture of 4 I distilled water and ice, which was acidified with HCI to pH= 0.0. Approximately 300 ml of a NH 40H solution was

added dropwise to the TaCl 5 solution during a period of 100 min while vigorously stirring the solution . After all the ammonia had

been added to the TaCl5 solution, the pH had increased to 6.0. After stirring for another 100 min, the precipitated Ta(OH)s was filtered off and washEd several times with distilled water. thereafter it was carefully dried at 393 K for 24 h (heating rate 2 K min-1),

cooled down to room temperature. powdered, dried again as

described above, and finally calcined for 1 hr at 873 K (heating rate 2 K min -l). The resulting 7 g T a20 5 had a surface area of approxi-

2 -1 mately 100 m g .

Page 210: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 200

A 3 wt % Rh/T a20 5 catalyst was prepared using the urea

method (7,8). 350 ml distilled water was stirred and heated to

365 K. 3 g of the high surface area Ta20 5 was added and the solu­

tion was acidified with 8 N HCI to pH = 2.5. Then 0.53 g of urea (a tenfold excess based on the amount of RhC1 3) was added and

finally 0.2254 g of RhCl3. At 365 K the urea slowly decomposed

and the pH of the solution increased very slowly. At a pH value of

approximately 4, Rh( OH h started to precipitate. After 10 h the

catalyst precursor was filtered off and dried as described above for

the T a20 5 support. In order to remove the remainings of the urea,

the sample was calcined at 923 K, pre-reduced in hydrogen at 773 K

and finally oxidized at 573 K. This sample was stored for further

use. Temperature programmed reduction experiments indicated that

reduction was complete· at 470 K when using 4% H2 in N2. Hydro­

gen chemisorption measurements after reduction at 523, 773 and

873 K gave H/Rh values- of 0.93, 0.14 and 0.06, respectively.

9.2.2 EXAFS Measurements

The dried, calcined, reduced and subsequently oxidized catalyst

was pressed into a thin self supporting wafer. The thickness of the

wafer was such as to give an absorbance (µx) of 2.5 at the rho­

dium K-edge (23219.8 eV). assuring an optimum signal-to-noise

ratio in the rhodium EXAFS spectra. The wafer was mounted in an

EXAFS cell which enabled in situ pretreatments in different gas

atmospheres at temperatures ranging from 100 to 873 K. The sam­

ple, once mounted in the cell, was reduced in 100 % H2 at 523 K for

1 h (heating rate 5 K min-1). After cooling with liquid nitrogen to

100 K an EXAFS spectrum was recorded with the sample still under

H2. Thereafter, the sample was reduced in H2 at 858 K for 15 min (heating rate 5 K min-1

), cooled down to 100 K and a second

EXAFS spectrum was recorded. Finally, the sample was evacuated

at 523 K for 1 h, oxygen was admitted to the sample at 100 K and

Page 211: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 201 Chapter 9

another EXAFS spectrum was recorded, with the sample still under

oxygen atmosphere. The EXAFS spectra of the reference com­

pounds, rhodium foil , Rh 20 3• RhCl3• Rh3 Ta alloy. Ta powder and

TaCl5, were recorded at 100 K as well. The absorption spectra were recorded at the synchrotron radiation source (S RS) in Dares bury.

U.K. The ring was operated at 2.0 GeV and with ring currents from 100 to 300 mA.

9.3 Results

9.3.1 Reference Compoonds

The backscattering amplitude F (k ) and the phase shift func­

tion </> (k) which are necessary for analyzing the EXAFS data have been obtained from reference compounds. Following the same pro­

cedure as described in chapters 2, 7 and 8, we extracted F (k ) and

</> (k) for Rh-0 contributions using Rh20 3, for Rh-Cl contributions using RhCl3, for Ta-Ta contributions using Ta powder and for Ta­

Cl using TaCl5. In Table 9.1 all the relevant information concerning the references is given. The crystallographic data were obtained from (9).

Extracting F (k ) and </> (k ) for Rh-Ta contributions was not as straightforward as described above. In the most obvious refer­

ence compound, the Rh3Ta alloy, rhodium has rhodium neighbors · and tantalum neighbors . . In the Fourier transform of the EXAFS

spectrum of the alloy, however, the Rh-Rh and Rh-Ta peaks over­

lapped almost completely and ex fr acting F (k ) and </> (k ) for Rh-Ta was impossible. We therefore choose to compose F (k ) and </> (k) using other reference compounds. The basic principle is the follow­ing. The backscattering amplitude is only function of the scattering

atom (6). Therefore, F (k) is the same for the absorber-scatterer

Page 212: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 202

Table 9.1 Crystallographic data and Fourier transform ·ranges for the reference compounds

Compound Edge NNa

Rh foil Rh . K Rh

Rh70 1 Rh. K 0

RhCI ~ Rh. K Cl

Ta powder Ta . L 111 Ta

TaCI~ Ta . L111 Cl

a : Nearest Neighbor b : Coordination Distance (A) t : Coordination Number ·

Rb NC no

2.687 12 3

2.05 6 3

2.31 6 1 2.863 8 3 2.37 6 3

d : Weighting factor in Fourier transformation

Crystallographic data were obtained from (9)

Rh K-edge : 23219.8 eV

Ta L1u-edge : 9881.0 eV

Fourier transformation k-range r-range

2.90-25.48 1.80·2.90 2.64-22.17 0.68-2.12 3.00-20.15 0.00-2.33 2.95-16.97 2.14-3.48 2.44-15.87 0.18-1 .88

pairs A- B and B- B. Thus, we used F (k) from the Ta-Ta contribu­

tion to represent F(k) for Rh- Ta. F(k) for Ta-Ta has been

extracted from the L111 EXAFS spectrum of tantalum powder. The

phase shift function <f>A-B (k) can be written as (6)

<t>A-B(k) = <t>t_(k) + <t>~(k) - 07T

In this equation, <f>J.(k) is the contribution of the absorbing atom A

and <f>s(k) of the scattering atom B to the phase shift in the EXAFS function. For K and L1 edges, withs-symmetry, the factor S is equal to unity , for L11 and L111 edges, with p-symmetry, S is zero.

We measured the Rh K-edge EXAFS spectrum of RhCl3 and the Ta

L111 EXAFS spectra of Ta powder and of TaCl5• From these EXAFS

functions we extracted ¢Rh-c1(k ), <l>Ta-Ta(k) and ¢Ta-c1(k ). A linear

Page 213: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 203 Chapter 9

combination of these three functions yielded the desired phase shift

function for Rh-Ta contributions (6,10,J J ), as is shown below. Note, that the factor TT remains in the expression, so the resulting

phase shift function can indeed be used for Rh K-edge EXAFS spectra (even when for Ta powder and T aCl 5 the Ta K or L1 edges

were measured, this would still be the case).

¢Rh-c1(k) + ¢Ta-Ta(k) - ¢Ta-o(k) = ¢~h(k) + <f>t1(k) - TT

+ <f>fa(k ) + ¢f a(k )

- ¢f a(k ) - <f>ti(k )

= <f>~h(k) + <f>fa(k) - TT

= <f:>Rh-Ta(k)

9.3.2 Analysis of the EXAFS Spectra

Our procedure of analyzing the EXAFS spectra has already been described extensively in the literature (5,12-14) and in

chapters 2, 7 and 8. Briefly, the analysis comprised the following steps. Using the cubic spline routine (JS), a smooth background

was subtracted from the experimental data. The resulting EXAFS function was then normalized by division to the height of the edge.

Then, using the backscattering amplitude F (k) and the phase shift function </>(k) of suitable reference compounds, EXAFS spectra

containing one or more shells were calculated. In calculating EXAFS spectra, for each shell of neighbors, four parameters can be varied :

the coordination number N, the coordination distance R, the Debye

Waller factor !::.a 2 to account for any disorder, and E-o, which allows a small correction on the edge position . By varying these parame­ters one tries to make the calculated EXAFS spectra resemble the

Page 214: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 204

measured spectra as accurately as possible. In two previous studies

(4,5) we described a recurrent optimization process in order to

separately analyze the contributions from high-Z and low-Z scatter­

ing neighbors. Since in this study the contribution of the low-Z

scatterer (oxygen) turned out to be very small and the remaining

contributions originated from high-Z scatterers (rhodium and tan­talum), this procedure could not be used. Moreover, the Rh-Rh and

Rh-Ta contributions overlapped in the Fourier transform (see Figure 9.2f), making it impossible to use the difference file technique. As a

result, the analysis we used was a single step multiple shell

approach wi.th four parameters (N, R. !J,.a 2 and Eo) for each shell to

optimize : we calculated a Rh-Rh EXAFS function and two Rh-Ta

EXAFS functions, added them and Fourier transformed the result- .

ing spectrum in order to compare with the (Fourier transform of the) measured data. Because of the high-Z character of the main

contributions (Rh and Ta), the use of k3-weighted Fourier transforms was essential. Differences between the two spectra were minimized by varying the Rh-Rh and Rh-Ta parameters. The

results of this analysis procedure are presented in Table 9.2.

In Figure 9.1a, the raw EXAFS data for the sample reduced at 523 K and the calculated best fitting Rh-Rh EXAFS function are

shown. The imaginary part is shown in Figure 9.1b and the magni­

tude of the Fourier transforms of these EXAFS functions in Figure

9 .1c. From Figure 9.1, it is obvious that apart from the Rh-Rh con­tribution, no other contribution is present in the spectrum. In Fig­

ures 9.2a and 9.3a the raw EXAFS spectra and the calculated best fitting Rh-Rh EXAFS functions for the samples after reduction at 858 K and oxygen admission are shown . In Figures 9.2b and 9.3b

are shown the imaginary parts of the k 3-weighted Fourier

transforms of the measured data and calculated Rh-Rh EXAFS functions . Figures 9.2c and 9.3c show the magnitude of these

Fourier tranforms. The differences in Figures 9.2b, 9.2c, 9.3b and 9.3c at the left hand side of the main Rh- Rh peak are due to the neighboring tantalum ions. We tried to fit these differences with rhodium and oxygen neighbors as well, but the fits resulted in

Page 215: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 205 Chapter 9

Table 9.2 Final results from EXAFS data analysis

Treat- Coordination Distance /j.(T L D

ment NN number (AJ (* 10-3 .A-2) {al fa) (a)

R523 Rh 7.9 0.2 2.658 0.005 7.4 1

R858 Rh 7.9 0.2 2.650 0.005 7.0 1 Ta 1.6 0.5 1.7 0.3 5.4 2 Ta 0.8 0.2 2.0 0.1 5.4 2

R858 Rh 7.9 0 .2 2.650 0.005 6.9 1 -E523 Ta 1.0 0.5 1.7 0.3 5.6 2 -0100 Ta 1.0 0.2 2.1 0.1 5.6 2

R

E

0 :

Reduction in H2 at the temperature indicated

Evacuation at the temperature indicated

Admission of oxygen at the temperature indicated

Estimated overall (experimental + systematic) error

Eo ()

(eV) (a)

5.5 1

7.8 1 -10 3

-6 3

7.0 1 -20 4 -15 4

a b r:w 2, the Debye Waller factor , is a measure for the disorder and E0 is

a correction on the edge position.

physically irrelevant parameters : for Rh-Rh , coordination distances

of about 1 A and for oxygen coordination numbers higher that 10. In addition, the resulting fit was worse than the fit with tantalum

neighbors. In Figures 9.2d, 9.2e and 9.2f are shown the raw data

and the calculated best fitting Rh-(Rh+ Ta) EXAFS f1,mctions, the

imaginary parts of their k 3-weighted Fourier transforms and the

magnitude of the Fourier transform of the raw data and of the three

separate contributions. Clearly, the agreement at the left hand side

of the main Rh-Rh peak in the Fourier transform is better . The

Fourier transforms of the EXAFS spectra were complicated by the

k-dependence in F (k) and <!>(k ). Therefore. the transforms were

corrected for F (k) and c/>(k) from rhodium foil, the reference for the Rh-Rh contribution, which was the major contribution in all

spectra. As a result, in the Fourier transforms, the Rh-Rh contribu­

tions ' peaked ' at the correct Rh-Rh distance and the imaginary

Page 216: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 206

Figure 9.1 Rh/Ta20 5 after reduction at 523 K •

(a) Raw data (solid line) and calculated Rh-Rh EXAFS funct ion (dotted line)

(b) Imaginary parts of the k3-weighted Fourier transform of the raw data (solid line) and calculated Rh-Rh EXAFS function (dotted line)

(c) Magnitude of the k3-weighted Fourier transform of the raw data (solid line) and calculated Rh-Rh EXAFS function (dot­ted line)

* 10-2 30 b 5 ~~~~~~..--;-.,~~~~~

a

-30

40 c

20

-5 0 5 10 15 20 25

k [A-1 J 0

0 2 4 6

R [AJ parts of the Fourier transforms were more or less symmetric. The

peaks corresponding to other minor contributions can shift to seem­ingly longer or shorter distances and can be asymmetric. However, since the same correction has been applied to measured and calcu­

lated data, the calculated coordination radii and coordination numbers represented those in the sample as accurately as possible.

Page 217: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 207 Chapter 9

9.4 Discussion

9.4.1 Rh/ Ta20 5 after Reduction at 523 K

According to the TPR experiments, reduction of the sample

was complete at 523 K. A careful analysis of the EXAFS spectrum

confirmed this . From Figures 9.1a, 9.1b and 9.1c it is obvious that

apart from a Rh-Rh contribution, no other contribution was present.

The deviations in Figure 9.'.lb and 9 .1c are very small: the deviation

at the right hand side of the main peak could not be fitted with a

Rh- Rh. a Rh-0 or a Rh-Ta contribution . From Table 9.2 it can be concluded that on the average each rhodium atom had approxi­

mately 7 .9 rhodium neighbors at 2.658 A, which is slightly shorter

Figure 9 .2 Rh/Ta20 5 after reduction at 858 K •

(a) Raw data (solid line) and calculated Rh-Rh EXAFS function

(dotted line)

(b) Imaginary parts of the k 3-weighted Fourier transform of the

raw data (solid line) and calculated Rh-Rh EXAFS function

(dotted line)

( c) Magnitude of the k 3-weighted Fourier transform of the raw

data (solid line) and calculated Rh- Rh EXAFS function (dot­

ted line)

(d) Raw data (solid line) and calculated Rh-(Rh+ T a) EXAFS

function (dotted line)

(e) Imaginary parts of the k 3-weighted Fourier transform of the

raw data (solid line) and calculated Rh- (Rh+ Ta) EXAFS

function (dotted line)

(f) Magnitude of the k 3-weighted Fourier transform of the raw

data (solid line) , calculated Rh-Rh EXAFS function (dotted

line) , the calculated Rh-Ta EXAFS functions (dashed lines)

and the sum of the calculated Rh-Rh and Rh-Ta contributions

(dash-dotted line)

Page 218: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 208

* 10-2 * 10-2

6~~~~~~~~~ 6~~~~~~~~~

a d

-6+-'-'"--'--'-t~~~'-t-'-~~-'-"--1 -6+-'-'"--'--'-t~~~'-t-'-~~-'-"--1

0 5 10 15 20 25 0 5 10 15 20 25

k [A- 1J k [A- 1J

30 b 30 e

-30 -30

40 c 40 f

20 20

0 ... ·.·.

0 2 4 2 4 6 0

R [A} R [A}

Page 219: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 209 Chapter 9

than the Rh-Rh bulk distance (the Rh-Rh distance in rhodium foil is 2.687 A) . Using the calibration procedure as described in ( 16), we

could estimate from the Rh-Rh coordination number that the parti­

cles were approximately 17 A in diameter and contained about 73 ± 5 rhodium atoms. Acco~ding to ( 16), the H/Rh value deter­

mined with hydrogen chemisorption should be about 0.95 according to both the experimental calibration method and the computer

model calculations. This agreed excellently with the measured value, H/Rh = 0.93, and therefore we conclude that the hydrogen

chemisorption capacity was 'normal' and thus, that the metal parti­cles were in the normal state.

In the EXAFS spectra of fully reduced supported metal catalysts, a metal-oxygen contribution from the metal atoms in the metal-support interface having oxygen neighbors has frequently been reported (4,5,12,17). In the Fourier transform, such a contri­

bution is situated at the left hand side of the main Rh-Rh peak. Since the relative extent of the metal-support interface decreases

with increasing particle size, the contribution from neighboring oxy­gen ions decreases with increasing particle size as well ( 12 .17). For

a 73 atom metal particle, about 30% of the rhodium atoms are situated in the metal support interface. When these rhodium atoms each have 2 to 3 oxygen neighbors, the average Rh0-02

- coordina­tion number is about 0.6 to 0.9 and. because of the low coordina­

tion number and the low backscattering amplitude of oxygen, the corresponding contribution in the Fourier transform should be

smaller by more than an order of magnitude compared to the Rh-Rh contribution. Accordingly, in the EXAFS spectrum of the sample

reduced at 523 K, no contribution from neighboring oxygen ions could be detected.

Page 220: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 210

Figure 9.3 Rh/Ta20 5 after reduction at 858 K and oxygen admission at 100 K:

(a)

(b)

(c)

Raw data (solid line) and calculated Rh-Rh EXAFS function (dotted line)

Imaginary parts of the k3-weighted Fourier transform of the raw data (solid line) and calculated Rh-Rh EXAFS function (dotted line)

Magnitude of the k3-weighted Fourier transform of the raw data (solid line) and calculated Rh-Rh EXAFS function (dot­ted line)

* 10-2 30 b 6 ,,--.-.-,,....,-.-,--,-,...T"T""T"""T""">,,....,~"T"""T"""~~

. a

40 c

20 - 6 +-'-~'-f-'....L...L..i..+..._.__..........,1-'-'-...J.....J...+-1--.._._'--l

0 5 10 15 20

k r A-1 J 25

0 +=~'--"----"-Jf----L~-+=~~~~ 0 2

R 4

f AJ 6

9.4.2 Rh/ Ta20 5 after Reduction at 858 K

After reduction at 773 K, the H/Rh value determined with hydrogen chemisorption decreased to 0.14, and after reduction at

873 K to 0.06. Clearly, after reduction at 858 K the metal particles

were in the SMSI state. The Rh- Rh coordination number remained

unchanged. Obviously, the basic structure of the metal particles

Page 221: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 211 Chapter 9

remained intact. In the earliest studies on SMSI. it was suggested

that in the SMSI state the metal particles might spread over the support and that ' pillboxes' would be formed ( 18.19). Such a spread

of the metal particles should be accompanied by a significant decrease in the Rh-Rh coordination number. We did not observe

such a decrease and therefore we conclude that in the case of

Rh/T a20 5 spreading of the rhodium particles did not occur. Another explanation was the formation of alloys ( 20). We did not observe contributions from neighboring tantalum atoms at dis­

tances in the range of 2.7-2.8 A (in the Rh3Ta bulk alloy, the Rh-Ta distance is 2.729 A (9)). Therefore, alloy formation can be ruled

out. The Rh-Rh coordination distance decreased by 0.008 A which is low for a metal particle which is not covered with adsorbed hydrogen. In ( 5) we reported a decrease of 0.05 A for very small rhodium metal particles. This contraction was due to the absence

of hydrogen on the surface of the metal particles in the SMSI state. The difference between the decrease observed for the Rh /T a20 5 catalyst and for the Rh/Ti02 catalyst in (5) can be explained as fol ­lows. In vacuum or under inert · gases, the first interatomic layer

distance decreases (21-23). Thus, for very small metal particles a considerable decrease in the average Rh-Rh distance is to be

expected and has indeed been reported (5,12,17). In general , how­ever , the second inter atomic layer distance expands, the third con­tracts, and so forth. The deeper the layers, the smaller the decrease or increase. As a result, for larger metal · particles the decrease in

the average Rh-Rh distance is less pronounced, as is the case for

the 17 A metal particles in the Rh/T a20 5 catalyst in this study.

In the spectrum, two more contributions were present which both originated from tantalum neighbors at notably short distances .

In the Fourier transforms, the peaks from both Rh-Ta contributions overlapped to a large extent with the major Rh-Rh contribution . As a result, th.e uncertainty in the Rh-Ta coordination number is larger

than the uncertainty in the Rh-Rh coordination number . Because the Rh-Rh contribution is much larger than the Rh-Ta contribu­

tions, the uncertainty in the Rh-Rh contribution is not affected. The

Page 222: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 212

estimated uncertainty in the 2.0 A Rh-Ta coordination number is

about 0 .2 and in the distance about 0 .1 A. The uncertainty in the 1. 7 A Rh-Ta contribution, however, is larger. An irregularity around

k = 9 A-1 was visible in the EXAFS spectra of the Rh/Ta20 5 sam­

ples reduced at 858 K (see Figures 9.2a and 9.3a) . This irregularity

gave rise to a peak in the Fourier transform centered at r = 1.6 A. In the Fourier transforms of the sample reduced at 523 K (see Fig­

ures 9.1b and 9.1c) , the contribution from this artefact was very

low and was clearly separated from the Rh-Rh peak and caused only

small deviations in the Fourier transform. The 1.7 A Rh-Ta contri­bution had a. main peak at 2.3-2 .4 A in the Fourier transform and

two sidelobes at 1.8 and 1.4 A (see Figure 9.2f). The main peak is shifted from the real Rh-Ta distance because the Fourier transform

was corrected for Rh-Rh phase shift and backscattering amplitude. The two sidelobes interfered with the artefact at 1.6 A and thus ,

only the main peak of this contribution could be used to determine the accompanying parameters. This caused an additional uncer­

tainty in the parameters for the 1.7 A Rh-Ta contribution : about 0.2 in the Rh-Ta coordination number and approximately 0.1 A in

the Rh-Ta coordination distance (additional to the uncertainty in the 2.0A Rh-Ta contribution). In the Rh-Ta distances, another

uncertainty has to be considered . The phase shift function for this absorber-scatterer pair has been composed from three phase shift

functions. Hence, the resulting uncertainty in the final phase shift function was the sum of the uncertainties in the three individual

phase shift functions . We estimated that based on this, the uncer­tainty in the resulting calculated Rh-Ta distances was about 0.1 A. In Table 9.2, the final uncertainties of all parameters are summar­ized .

Because the irregularity around k = 9 A-1

created problems, we tried to eliminate this from the spectra by means of the following

procedure. We subtracted the best fitting calculated spectra from

the measured data . The difference spectra contained mostly this artefact. Using a Hanning window between k= 8 and k= 9.5 A._-1, we

isolated this artefact from the difference file and subtracted this

Page 223: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 213 Chapter 9

from the measured data. The Fourier transform of these spectra

were indeed better than the spectra of the raw data : the artificial

peak around r = 1.6 A was almost completely removed. This is illustrated in Figure 9.4. However, this procedure induced new smaller artefacts around k= 8 and k= 9.5 A.- 1

. It was found impos­

sible to completely remove the artefact without introducing new artefacts. Nevertheless, our conclusion is, that the peak in the

Fourier transform around r=1.6A was indeed induced by the artefact around k= 8-9 A.- 1

.

Figure 9.4 The EXAFS function for the Rh/Tap5 sample reduced at 858 K (solid lines) and the spectrum for this sample. in which

the artefact has been 'removed· artificially (dotted lines)

(a) The EXAFS functions

(b) The k3-weighted Fourier transforms

(Dashed line : the calculated Rh-(Rh+ Ta) contribution)

* 10-2

6 ~~~~~~~~~~~~

a 40 b

20

- 6 4--L-..L..l....L+-'-'-L-'-l-_,_.._'-'-lf-L-L ......... +-'-J..-'-J'-l 0 +,..::..:::........:.'------'--+~----'---~~~~~

0 5 10 15 20 25 0 2 4 6

k rJ.- 11 R [A]

Altogether, the overall uncertainty in the Rh- Ta parameters

was quite large and made detailed interpretation meaningless. Nevertheless, there was no doubt that both contributions were present. Taking the uncertainties in the Rh-Ta distances into account, these results show that tantalum ions are present at

Page 224: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 214

distances ranging from 1.5 to 2.3 A. These are very short coordina­

tion distances and can arise only from Ta ions in almost direct con­

tact with rhodium atoms in the metal particles which are in contact

with tantalum (sub)oxide. They may therefore be located directly

underneath the rhodium metal particles. This indicates that indeed

the Ta20 5 support under the metal particles had been reduced.

However, the coordination numbers of the Rh- Ta contributions are

relatively high . When only the rhodium atoms in the metal-support

interface have Ta neighbors, and the support does not expose a

large amount of bare Ta ions, these coordination numbers cannot

exceed 0.3 (in the 73 atom metal particle, about 30% of the metal

atoms is in the metal-support interface and we assumed that each

interfacial rhodium atoms could have up to one tantalum neighbor

at such short distance) .. Therefore. we must conclude that also the

surface rhodium metal atoms, which are not in the metal-support

interface, must be in direct contact with Ta ions. We thus conclude that the metal particles were partly covered with reduced T a20 5.

The direct contact between rhodium atoms and tantalum ions after

reduction at high temperature could result in a strong interaction

between metal particle and support:

In (4,5) and in chapters 7 and 8, we reported Rh0-Ti distances

of 3.4 and 4.3 A. Such longer distances have not been observed for

the Rh/Ta20 5 catalyst in the SMSI state. The reason is the follow­

ing. Ti02 has a very simple and very regular structure. The 3.4 and 4 .3 A distances are therefore very well defined. For supports with a

more complicated crystal structure like Al 20 3 and Ta20 5 this is

(unfortunately) not the case. At short distances, the Rh-Ta (or

Rh-Al 3+) distances are well defined, but between about 3 and 5 A, many Rh-Ta distances can occur, each with a very low coordination

number. This makes makes it almost impossible to observe the

longer Rh-Ta distances and explains also why Rh-Al 3+ distances

have never been observed.

Page 225: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 215 Chapter 9

9.4.3 Rh/Ta20 5 after Admission of 0 2 in the SMSI State

After evacuating the sample at 523 K and admitting oxygen at

100 K, the major Rh-Rh contribution remained unaffected. Obvi­ously, just as for the Rh/Ti0 2 sample in the SMSI state (5), oxida­

tion had not taken place. Only a small but detectable change in the Rh-Ta parameters occurred. The distance in the most reliable Rh­

T a contribution increased by 0.1 A and the coordination number increased slightly, but the increase was less than the overall uncer­

tainty . The coordination number of the 1.7 A Rh-Ta contribution decreased. Regarding the uncertainties in these parameters. the only

conclusion we may draw is that no structural change has taken place and that possibly the reduced suboxide underneath the metal

particle might have started to reoxidize. The covering T a20 5 might be the reason why the oxidation was suppressed, but an electronic

effect might be the reason as well and may even be more likely. because we cannot estimate the extent of coverage (only a complete

coverage could suppress oxidation).

In chapter 7 and in ( 5), we did the same kind of oxygen

adsorption experiments on a Rh/Ti0 2 catalyst in the SMSI state. For the Rh/Ti02 sample, we found that oxygen could adsorb on the metal particles. Because the rhodium metai particles on the Ta20 5 support are much larger than those in the sample in chapter 7 and

in ( 5), and because the particles in Rh/Ta20 5 are partly covered with TaOX, the relative extent of the surface of the metal particles

exposed to the gas atrnosphere is very small, Therefore, for the Rh/T a20 5 sample it is impossible to detect oxygen absorbed on the

metal particles.

Page 226: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 216

9.5 Final Conclusions

With the Rh/Ti0 2 samples in chapters 7 and 8 we did not

observe any covering, but we could not exclude partial covering.

The fact that the rhodium metal particles in Rh/Ta20 5 are covered

to a possibly larger extent than the metal particles in Rh/Ti0 2 can

be explained by several reasons. First of all, the metal particles in

Rh/T a20 5 are much larger than the metal particles in the Rh/Ti02 samples in chapters 7 and 8 and coverage has up to now only been

reported in literature for larger metal particles. Another reason

might be the fact that the Rh/T a20 5 sample was reduced at much

higher temperatures (858 K) than the Rh/Ti0 2 samples (723 and 773 K). A third reason may be found in the preparation method :

the Rh/Ti02 samples in chapter 7 and 8 were prepared by exchang­

ing with a solution of Rh(N03h which had a relatively high pH.

The Rh/T a20 5 sample was prepared using the urea method and

therefore, the starting pH was low. Thus. during the preparation of

the Rh/Ta20 5 sample. some Ta20 5 might have dissolved and later

precipitated on top of the rhodium metal particles. This kind of

coverage has already been reported for Rh/V 20J/Si02 by Kip et al.

(24) and for Rh/V 20 3 by van der Lee et al. (25) and by Bastein

et al. . After reduction at high temperature, this Ta20 5 on top of

the metal particles will become reduced and may have an intimate

contact with the metal particle, giving rise to Rh-Ta bonding.

In summary, apart from the Rh-Rh contribution, no other con­

tributions could be detected In the EXAFS spectra of the Rh/Ta20 5 catalyst reduced at 523 K. From the Rh- Rh coordination number it

was concluded that the particles were about 17 A in diameter and

contained about 70 to 80 rhodium atoms . After reduction at 858 K

the sample was in the SMSI state. The Rh-Rh coordination number

remained unchanged. Thus, the basic structure of the metal parti­

cles remained intact. Any spread of the metal particles over the · support like a pillbox formation could be excluded . Alloy formation

had not taken place either. In the SMSI state the rhodium atoms in

Page 227: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 217 Chapter 9

the metal-support interface had tantalum ions as neighbors at very

short distances. ranging from 1.5 to 2.2 A. This indicated that the

Ta 20 5 oxide directly underneath the metal particles was reduced to

a lower oxide of Ta 20 5. In addition. the metal particles were at

least partly covered with reduced T a20 5. Since the only change in

the EXAFS spectra was from neighboring Ta ions. we conclude that

these neighboring Ta ions must be the origin for the SMSI state.

9.6 References.

1. Tauster, S J.; Fung, S. C.; Garten, R.L. J. Am. Chem. Soc. 1978. 100.170.

2. Tauster, S. J.; Fung, S. C. J. Catal. 1978, 55, 29.

3. Tauster, S. J.; Fung, S. C.; Baker, R. T. K.; Horsley, J A Science

(Washington, D.C.)1981, 211, 1121.

4. Koningsberger, D. C.; Martens, J. H. A.; Prins, R.; Short, D. R.;

Sayers, D. E. J. Phys. Chem. 1986. 90, 3047.

5. Martens, J. H. A.; Prins, R; Koningsberger, D. C. J. Phys. Chem.,

accepted for publication . (see chapter 7 of this thesis)

6. Teo. B. K.; Lee, P.A. 1. Am. Chem. Soc. 1979, 101, 2815.

7. Hermans, L. A. M.; Geus. J W. "Preparation of Catalysts 11 ": Del­

mon, B; Grange. P.; Jacobs, P A.; Poncelet. G., Eds.; Elsevier.

Amsterdam 1979. p. 113

8. Geus, J. W. "Preparation of Catal_ysts 111 ": Poncelet. G.; Grange,

P.; Jacobs, P.A., Eds.; Elsevier. Amsterdam 1983. p. 1

9. Crystal structures :

Rh metal Wyckhojf Crystal Strnctures 1963, 1. 10.

Rhp3 Structure Reports for 1974 1976. 40a, 301.

RhCl 3 Structure Reports for 1964 1972, 29, 275.

Structure Reports for 1964 1972. 29, 130.

Ta metal Wyckhojf Crystal Structures 1963, 1, 16.

TaCl 5 Structure Reports for 1958 1968, 22, 237.

Page 228: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 218

10. Citrin, P. H.; Eisenberger, P.; Kincaid, B. M. Phys. Rev. Let. 1976, 36, 1346

11. Sinfelt, J. H.; Via, G. H.; Lytle, F. W.; Greegor, R. B. J. Chem. Phys.

1980, 72(9 ), 4832

12. van Zon, J. B. A. D.; Koningsberger, D. C.; van 't Blik, H. F. J.;

Sayers, D. E. J. Chem. Phys. 1985, 12, 5742.

13. Duivenvoorden, F. B. M.; Koningsberger, D. C.; Uh, Y. S.; Gates, B.

C. J. Am. Chem. Soc. 1986 , 108, 6254.

14. van 't Blik, H. F. J.; van Zon, J. B. A. D.; Huizinga, T.; Vis, J. C.; Koningsberger, D. C.; Prins, R. J. Amer. Chem. Soc. 1985, 107, 3139.

15 Cook, J. W.; Sayers, D. E. J. Appl. Phys. 1981, 52, 5024.

16. Kip, B J.; Duivenvoorden, F. B. M.; Koningsberger, D. C.; Prins, R.

J. Catal. 1984, 105, 26.

17 Koningsberger, D. C.; van Zon, J. B. A. D.; van 't Blik, H. F. J.;

Visser, G J; Prins, R.; Mansour, A. N.; Sayers, D. E.; Short, D. R.;

Katzer, J. R. J. Phys. Chem. 1985, 89, 4075.

18. Baker, R. T. K.; Prestridge, E. B.; Garten, R. L. J. Catal. 1979, 56,

390.

19 Baker, R. T. K.; Prestridge, E. B.; Garten, R. L. J. Catal. 1979, 59,

293.

20. Beard, B. C.; Ross, P. N. J. Phys. Chem. 1984, 90, 6811.

21. Somorjai, G. A. "Chemistry in Two Dimensions", Cornell University

press, 1981

22. "The Chemical Physics of Solid Surfaces and Heterogeneous Catalysis", King, A. D.; Woodruff, D. P., Eds, Elseviers, Amsterdam

1981

23. Jona, F.; Marcus, P. M. Proceedings JCSOS II, Amsterdam, 1987, in

press

24. Kip, B. J.; Smeets, P. A. T.; van Grondelle, J.; Prins, R. Appl. Catal.

1987, 33, 181

25 van der Lee, G.; Schuller, B.; Post, H.; Favre, T. L. F.; Ponec, V. J. Catal. 1986, 98, 522

25 Bastein, A. G. T. M.; van der Boogert, W. J.; van der Lee, G.; Luo, H.; Schuller, B.; Ponec, V. Appl. Catal. 1987, 29, 243

Page 229: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 219 Chapter 10

Chapter 10

Concluding Remarks

In this thesis, we discussed several interesting aspects of sup­ported metal catalyst. Using the results from l\JMR, Raman and

adsorption experiments, we described a model that explained the reduction process of bimetallic Rh-Pt catalysts and the formation of

bimetallic particles. It appeared that adsorbed RhCl 3 complexes were reduced most easily and that the resulting rhodium atoms had

no significant interaction with the support. Therefore, these mobile rhodium atoms could migrate over the support and catalyze the

reduction of other adsorbed complexes. Since adsorbed RhCl 3 as well as adsorbed H2PtCl6 complexes were present, metal clusters

were formed that contained both platinum and rhodium. The com­position of these particles was therefore statistically determined. In

chapter 4, however , a completely different situation was encoun­tered : one of the two metals was iron, the other was a more noble

metal (Ru or Pt). It appeared that also in that case the iron was reduced to metallic iron and incorporated in bimetallic particles.

But not all of the iron was reduced, the other part remained in the Fe3+ state. In monometallic iron catalysts, however , iron can be

reduced to Fe2+ and further. Thus , noble metals seem to be able to either catalyze the reduction of a second component, or to suppress

the reduction of that second component . This laher phenomenon is up to now still not solved completely and additional research effort

has to be directed towards this problem. A solution may be found in the interaction between Fe3+ ions and the support. We could assume that Fe3+ in direct contact with the support or even in the surface of the support will not be reduced to Fe2+. Therefore , for

larger, three-dimensional iron oxide or iron hydroxide crystals , Fe3+ will always be present in the particle-support interface, even after

reduction , while the rest of the particle may become reduced. Thus, it may be that in bimetallic Fe-M catalyst systems the noble

Page 230: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Concluding remarks page 220

component plays a role in keeping the Fe3+ phase in a highly

dispersed form in which it cannot be reduced. And it has indeed

been shown by Mossbauer that the Fe3+ phase is a highly dispersed

phase. We could use the results from chapter 3 to put this model in

another light. Fe(N0 3)J is always used As precursor for iron

catalysts . It is common knowledge that these nitrates hydrolyze and

form large clusters in solution. This process can be suppressed in

acid media . Hence, when a chloride such as H2PtCl6 is used as a

precursor for the noble metal component, the acidity of the solution

will increase , hydrolyzation will be suppressed and iron complexes

may adsorb in a highly dispersed way on the support . This highly

dispersed iron is then ' immune' for reduction. This once more

confirms the importance of the first two steps in catalyst prepara­

tion, the choice of the metal precursors and the impregnation step,

and their impact on the final state of the catalyst.

Regeneration is an important aspect of a supported metal

catalyst : during processes such as Fischer T ropsch synthesis , the

activity of the catalyst slowly decreases . Therefore, after a period

of time, the catalyst has to be regenerated and oxidation is an

important step in the regeneration of a supported metal catalyst.

Oxidation is also important for storing a catalyst. Once a catalyst

has been reduced in hydrogen and is highly active, it cannot be

stored in air without further precautions. Passivation is a process

in which the small metal particles are oxidized in a slow and con­

trolled way in order to prevent sintering. Oxidation of small metal

particles is a process that has not been studied carefully up to now.

In chapter 5 we have described a model for the oxidation of small

rhodium metal particles based on data obtained with EXAFS.

EXAFS indicated that during the oxidation of the metal particles a

new metal-to-oxide interface was created : a rhodium-to-rhodium oxide interface. It was shown that during the oxidation process

rhodium oxide covered the metal particles. After careful oxidation

at room tempera~ure , a small kernel of the particles remained metal­lic and this metallic kernel was only partly covered with rhodium

oxide. Thus, a part of the metal phase was still exposed to the gas

Page 231: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 221 Chapter 10

atmosphere. This explains why passivated noble metal catalysts

can be reduced easily : the metallic kernel can adsorb and dissociate

hydrogen especially at low temperature and thus can catalyze the

reduction of the rhodium oxide. Another important aspect has been shown in chapter 5 : EXAFS proved to be capable of detecting

oxide on top of the metal particles, or covering metal particles .

This will be of interest in chapters 7 , 8 and 9, in which the question

is addressed whether metal particles supported on transition metal

oxide supports are covered with support material in the SMSI state.

The results of the ASED-MO computations in chapter 6 show

that this kind of modelling can indeed be of importance in determin­

ing structures in metal catalysts. It is, however, merely a star.t that indicates that a lot of useful information can be obtained in this

way . In order to ascertain that the results are indeed reliable, addi­

tional research will have to be done. The main question is, how

model-dependent the calculations are. For example, what would

happen if we did not incorporate any protons in the Al20 3 cluster ? We even could place the protons inside the Al20 3 support cluster.

What happens if we do not assume that the metal particle has a

rigid structure and we optimize the coordinates of each rhodium

atom in the metal particle separately ? One of the main problems

that has not been discussed in chapter 6 is of more practical nature : the computer and the (computer) time involved. For each calcula­

tion a memory capacity (storage capacity in I BIVI terminology) of 10

megabyte was needed . Since this is not at hand on commonly avail­

able mini computers, we had to install the software on the IBM 4381 main frame. Even on such a main frame, one CPU hour was

needed to calculate one point in the energy-distance curves. A com­

plete curve would typically take about 12 hours CPU time. On a

time-sharing base, this will take about a day or even more in real time. Hence, apart from the theoretical problems, a lot of practical

problems still have to be solved. But, once these are solved, a com­

plete new horizon of applications becomes available. For example,

we could perform these calculations for m~al particles supported on Ti02 and suboxides of Ti02. One could check whether coverage

Page 232: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Concluding remarks page 222

may occur, whether electronic influences are likely, what exactly

they are and how they influence the adsorption capacity of the metal particle, or we could find out if indeed alloy formation can

occur or whether the metal particles will spread on the reduced sup­

port.

In chapters 7, 8 and 9, the special properties of metal catalysts supported on transition metal oxides have been addressed. For

these metal catalyst , two states are attainable : a normal state and an SMSI state . In the normal state, they behave like metal particles

supported on inert oxides like Al 20 3, in the SMSI state their proper­ties change. Most pronounced is the decrease in hydrogen and car­

bon monoxide adsorption capacity. Several models have already been proposed to explain this phenomenon . The most important

models are ( i) the model of coverage, (ii) the electronic influence from the reduced support oxide on metal particle, (iii) a model of

'pill box ' formation , which assumes that the metal particles spread over the support, and (iv), the model of alloy formation. With two

Rh/Ti0 2 catalysts and a Rh/Ta 20 5 catalyst we found that in the SMSI state the metal particles rested on reduced support oxide.

With the Rh/Ti02 samples, we did not observe any covering oxide. We could, however , not exclude any loose coverage of metal parti­

cles in the Rh/Ti0 2 catalyst samples . But, with the Rh/Ta 20 5 we found that the metal particles were at least partly covered with a

suboxide of Ta 20 5. Several explanations are possible. First of all, the metal particles in the Rh/T a20 5 are much larger than the parti­

cles in the Rh/Ti0 2 catalysts and coverage has up to now mostly been reported for such larger metal particles. Why small metal par-'

ticles may not be covered epitaxially, however, is not clear . A pos­sible explanation is that larger metal particles expose more defined

crystal faces on which epitaxial growth of an oxide phase is likely. A second explanation may be found in the preparation method : the pH of the impregnating solution was much lower for the Rh/Ta 20 3 sample. Hence, some Ta20 5 may have been dissolved and during the drying procedure, this dissolved Ta 20 5 may have precipitated on top of the rhodium precursor particles. A third explanation can be

Page 233: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 223 Chapter 10

found in the reduction temperature : the Rh/Ta20 5 sample was

reduced at a much higher temperature than both Rh/Ti0 2 samples.

Thus, we could assume that the metal particles in the Rh/Ti0 2 samples in the SMSI state were indeed partly covered with a TiOx suboxide, but that this coverage very loose. it was not an epitaxially

grown oxide. This covering oxide will rearrange and cover the metal particles more perfectly only after reduction at higher temperatures

( > 773 K), as we have found for the Rh/T a20 5 catalyst. Note, that the covering oxide on top of the metal particles in the Rh/ Al 20 3 catalyst in chapter 5 was rhodium oxide which can grow easily in an epitaxial form on rhodium metal.

The metal particles in Rh/Ti02 and in Rh/Ta 20 5 in the SMSI state did not oxidize upon exposure to oxygen at 100 K. Under

these circumstances, a ' normal ' catalyst, Rh/ Al 20 3 in chapter 5, started to oxidize. For the Rh/Ti0 2 catalyst, we could show that

the metal particles were (partly) covered with adsorbed oxygen . For the Rh/T a20 5 catalyst , we did not observe adsorbed oxygen .

Form the hydrogen chemisorption experiments , however , we may conclude that still some surface area is exposed to the gas atmo­

sphere. Therefore, we may conclude that in the Rh/Ta 20 5 sample in the SMSI state and under oxygen , part of the surface of the

metal particles is in contact with oxygen . Thus , if there is no further influence from the support, oxidation should have started.

Since we did not observe this, we must conclude that for Rh/Ti0 2 and also for Rh /T a20 5, an electronic influence from the support is responsible for the suppression of the oxidation process. This elec­tronic influence may also alter any other property of the metal parti­

cles. However, research still has to be done in order to establish the nature of this electronic influence and its impact on the properties

of the metal particles. As indicated in the preceding paragraph, the ASED-MO method is very promising and may give valuable infor­

mation in this respect.

Page 234: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Concluding remarks page 224

Page 235: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 225 Summary

Summary

In this thesis we focussed on a few aspects which were impor­

tant for the structures in supported metal catalysts : the prepara­tion of supported metal catalysts and the reduction process in

which the inactive metal salts are reduced to active metal particles. We found that, when two metal salts were used to prepare a bime­

tallic catalyst, the specific combination of the two metals had a pro­nounced influence on the final state of the catalyst . When the -salts

of two noble metals were used, the noble metal that was reduced first aided the reduction of the second metal. However, when iron

was one of the two metals, the noble metal kept part of the iron in its highest valence state, Fe3+. The other part was reduced to

metallic iron which formed alloy particles with the noble metal. We also studied the process of passivating a catalyst. That process is

of importance for storing reduced (i.e., active) metal catalysts. In their active state, metal catalysts cannot be stored in air without

risking run-away-oxidation and consequently sintering of the metal particles. We used the ASED-MO theoretical method to ascertain

that the information from EXAFS about the metal-support interface was indeed reliable. The metal-support interface has received much

attention in this thesis, especially for catalyst systems that suffer from (strong) metal-support interactions (SMSI). In these systems,

the metal-support interface plays a key role and the answer for the anomalous properties of the metal particles m the SMSI state can

be found in the metal-support interface.

In chapter 3, the preparation of y-Al 20 3 supported Rh, Pt and

Rh-Pt catalysts has been investigated. From NMR and Raman spectroscopy experiments we concluded that during the pore

volume impregnation method H2PtCl6 could adsorb as separate monometallic complexes on the Al 20 3 support up to a loading of

about 0.18 mmol g- 1. Above this limit, H2PtCl6 crystals were

formed. Adsorption experiments indicated that for RhCl 3 the same

behavior could be expected. The maximum attainable coverage of

Page 236: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Summary page 226

monometallic complexes was even larger for RhCl 3 than for

H2PtCl 6. This maximum attainable coverage depended on the

amount of non protonated surface hydroxyl groups present on the

support during the adsorption process. Therefore, the adsorption

capacity decreased with decreasing pH value of the solution. Dur­

ing the reduction process, the monodispersed rhodium complexes

were reduced first and because of their mobility could catalyze the

reduction of the complexes and crystal I ites which are harder to

reduce. Thus. during the reduction process, metal clusters could

'wander' over the support and on their way, catalyze the reduction

of unreduced material. Consequently, as the reduction proceeded,

these metal clusters increased in size. When they encountered a

large crystal of metal salt, this crystal became reduced but, because

of the size of the cluster after the reduction process, the metal clus­

ter had no significant mobility. Hence, these larger crystallites

merely 'captured' the mobile metal clusters and smothered the

reduction process. When an alcohol was used as solute, mono­

disperse adsorption did not take place. During the drying process,

RhCl 3 and H2PtCl6 crystallized separately. Because there were no

monodisperse rhodium complexes and therefore no mobile rhodium

atoms formed during the reduction process, the RhCl3 and H2PtCl6

crystal! ites were reduced separately.

Chapter 4 describes ESR investigations of the presence of fer­ric iron (Fe3+) in reduced SiOrsupported Fe-M bimetallic catalysts.

With Mossbauer spectrocopy, the presence of ferric iron in reduced

Fe-Ru/Si02 and Fe-Pt/Si0 2 has been observed. However. from the

Mossbauer spectroscopy point of view, there might be doubts about

this assignment. In ESR, Fe3+ ions cannot be mistaken with other

Fen+ ions. Therefore, the ESR experiments clearly indicated that

the assignment made by Mossbauer spectroscopy was indeed

correct. Moreover, the amounts of ferric iron as determined by ESR

agreed very well with the amounts as found with Mossbauer spec­

troscopy. Therefore, these observations make the model in which

iron and the noble metal maintain each other in a highly dispersed

Page 237: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 227 Summary

state, a state in which iron cannot be reduced to ferrous iron, very

likely. Note the difference with the previous results for Rh-Pt

catalysts : in that case the more easily reduced component aided

the reduction of the other component.

In chapter 5, an alumina-supported rhodium catalyst has been

studied with EXAFS . After reduction and evacuation, oxygen was

admitted at 100 K and at 300 K. EXAFS spectra of the catalyst

after oxygen admission at 100 K indicated the beginning of oxida­

tion. At 300 K only a small part of the rhodium particles remained

metallic and this metallic 'kernel' was partly covered with rhodium

oxide. In the rhodium metal to rhodium oxide interface the same 2.7 A Rh0-0 2

- distances are present as in the metal-support interface. Using Rh0-Rh0. Rh0-02- and Rh"+ -02- coordination numbers deter­

mined with EXAFS. a model has been derived which describes the

oxidation of small rhodium metal particles supported on A120 3.

The interactions between a ten atom rhodium metal cluster and a 98 atom y-Al20 3 cluster serving as support have been stu­

died in chapter 6, using the ASED Molecular Orbital method. The

bonding of the metal particle was strongest when only surface

hydroxyl groups and no bare support oxygen ions were present

underneath the metal particle. The rhodium atoms in the metal­support interface rested preferable on three-fold coordinated sites.

The accompanying Rh-02- bond length was approximately 2.54-

2.57 A, which is in good agreement with the values reported with

EXAFS (2.6-2.8 A). When only bare oxygen ions · were present

underneath the metal particle. the interfacial rhodium atoms pre­ferred on-top sites and the Rh-0 bond length decreased to about

2.09-2.10 A. Thus, it appeared that the protons in the surface

hydroxyl groups played a key role in the bonding of the metal parti­

cle to the support. Protons disfavor on top coordination of the metal atoms and favor binding on three-fold coordination sites.

Page 238: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Summary page 228

In chapter 7 and 8, EXAFS and HRTEM have been used to

study the structure of two Rh/Ti0 2 catalysts . One of these sam­

ples was prepared in Eindhoven and measured in Daresbury, the

other sample was prepared at Yale university and was measured in Cornell (CHESS) . The results of both studies are very similar. For

the Eindhoven sample it was found that after reduction in H2 at

473 K (when the catalyst is in the normal state) the metal particles

contain on the average five rhodium atoms and are situated prefer­

ably on edges of the Ti02 crystallites, but also on [101] and to a

lesser extent on [001.] anatase crystal faces . Reduction in H2 at

723 K leads to the SMSI state. Besides oxygen neighbors from the

support, the rhodium metal atoms in the metal-support interface

have Tin+ neighbors at 3.4 and 4.3 A. These distances and their

coordination numbers fit well with a model in which the metal parti­cles rest on a Ti02 suboxide. This indicates that the supporting

oxide near the metal particle has been reduced to a suboxide of

Ti02. In the SMSI state no indication for coverage has been found

with either EXAFS or HRTEM. On the contrary, exposing the catalyst in the SMSI state to oxygen at 100 K resulted in changes

in the EXAFS spectrum due to physisorption of oxygen . Conse­

quently, in the SMSI state the particles are either not covered or are

incompletely covered with TiOx. Since a Rh/ Al20 3 catalyst under the same conditions became partly oxidized, it is evident that for

the Rh/Ti0 2 catalyst oxidation has been suppressed. This is most

probably the result of an electronic influence from the reduced sup­

porting oxide. Even after oxygen admission at room temperature,

the rhodium particles on the TiOx support remain in the metallic

state. The TiO" suboxide in the vicinity of the metal particles starts to re-oxidize and the metal-support interaction becomes weaker.

For the Rh/Ti0 2 catalyst prepared at Yale University , the

reduction process proceeded only very slowly because of the very

compact structure of the sample once it was pressed into a wafer. Even after reduction at 628 K rhodium oxide was present in the

sample. Because of the delayed reduction process, the SMSI state was only invoked at relatively high temperature. Consequently, at

Page 239: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 229 Summary

lower temperatures, the binding of the metal particles to the sup­

port was weak and sintering occurred. Once the SMSI state was established, sintering process stopped. As for the Rh/Ti0 2 sample

in chapter 7, it was found that in the 'normal' state the metal parti­cles rested on unreduced Ti0 2 and in the SMSI state on reduced

Ti0 2. Again, there was no evidence for coverage nor for alloy for­mation in the SMSI state.

In chapter 9, a Rh/T a20 5 catalyst has been studied with EXAFS. After reduction at 523 K in H2, the rhodium particles were fully reduced and in the 'normal' state. The metal particles con­

tained about 73 rhodium metal atoms and were approximately 17 A in diameter. Apart from a contribution from rhodium neighbors, no

other contribution was detected in the EXAFS spectrum. After reduction at 858 K the catalyst was in the SMSI state. In addition

to a contribution from rhodium nearest neighbors, two contributions from neighboring tantalum ions could be detected. The tantalum

ions were located in the reduced supporting oxide directly under­neath the rhodium metal particles and in tantalum oxide covering

the rhodium metal particles. Their were no indications for alloy for­mation, nor for the formation of pillbox or raftlike structures. After

admitting oxygen at 100 K to the catalyst, the metal particles remained unaffected and indications were found that the supporting

tantalum suboxide underneath the metal particles and covering the metal particles had started to reoxidize.

Concluding, for two different supported rhodium catalysts (Rh/Ti02 in chapter 7 and 8 and Rh/Ta20 5 in chapter 9) different

results were obtained : for Rh/Ti0 2 no evidence for coverage has been found, but some coverage could not be excluded while for

Rh/T a20 5 indications for coverage have indeed been found. How­ever, it is impossible to indicate to what extent coverage in the

Rh/T a20 5 catalyst had taken place. Note, that the Rh/Ta 20 5 sam­ple was reduced at a much higher temperature. Thus, a possible

explanation is that, because the Rh/Ti0 2 samples were reduced at

Page 240: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Summary page 230

lower temperatures, coverage was only loose and could not be

detected by EXAFS. Upon reduction at higher temperatures, how­

ever, this loose covering oxide may re-organize into epitaxially

grown (sub )oxide.

Page 241: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 231 Samenvatting

Samenvatting

In dit proef schrif t hebben we de aandacht gevestigd op struc­

turen in gedragen metaalkatalysatoren. We hebben een aantal

aspecten de revu laten passereri : de bereiding van katalysatoren,

het proces waarin de niet aktieve metaalzouten op het dragerma­

teriaal worden aangebracht en daarna worden gereduceerd tot

aktieve metalen. Wanneer twee metaalzouten aangebracht worden

op het dragermateriaal, dan blijkt dat de combinatie van die twee

metalen bepalend is voor de eindtoestand van de katalysator. In

sommige gevallen helpt het ene {meestal het meest edele metaal) bij

het reductieproces van het andere metaal, in andere gevallen onder­

drukt het meest edele metaal de reductie van het andere metaal.

We hebben ook het proces van passiveren bestudeerd. Oat proces

is van belang om katalysatorsystemen te kunnen bewaren. In hun

aktieve toestand is het niet mogelijk gedragen metaalkatalysatoren

in lucht te bewaren zonder het gevaar te lopen dat de metaaldeeltjes

groter worden en daardoor hun aktiviteit verliezen. We hebben

modelberekeningen gebruikt om aan te tonen dat de informatie die

ons de EXAFS-techniek verschafte over het grensvlak tussen het

metaaldeeltje en de drager betrouwbaar is. Oat metaal-drager

grensvlak is een aspect dat in het bijzonder is uitgediept in dit

proefschrift, en met name voor dragermaterialen die een bijzondere

invloed uit kunnen oefenen op de metaaldeeltjes die zich op dat

dragermateriaal bevinden. We hebben kunnen aantonen dat het

dragermateriaal de electronische eigenschappen van die metaal­

deeltjes kan beinvloeden en soms de metaaldeeltjes gedeeltelijk kan

bedekken.

In hoofdstuk 3 hebben we de bereiding van monometallische

Rh, Pt en bimetallische Rh-Pt katalysatoren gedragen op y-Al 20 3

besproken. Op grond van NMR en Raman experimenten hebben we

kunnen vaststellen dat tijdens porie-volume impregnatie het zout H2PtCl6 in de vorm van afzonderlijke molekulen aan het dragerma­

teriaal kan adsorberen tot aan een grens van ongeveer 0.18 mmol

Page 242: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Samenvatting page 232

per gram A1 20 3. Boven deze grens warden tijdens het draogproces

H2PtCl6 kristallietjes gevormd. Adsorptie-experimenten toonden

aan dat we voor RhCl3 hetzelfde gedrag mogen verwachten . De

maximale 'monodisperse' bedekkingsgraad voor RhCl3 is zelfs grater

dan voor H2PtCl6. Voor beide zouten wordt die maximale mono­

disperse bedekkingsgraad beinvloed door de zuurgraad van het

oplosmiddel waarmee de zouten op het dragermateriaal warden

aangebracht : bij een lagere pH kan er minder zout aan het drager­

materiaal adsorberen. Tijdens het reductiepraces warden de mono­

disperse rhodium-complexen als eerste gereduceerd en er ontstaan

rhodium-metaalatomen. Die atomen hebben in tegenstelling tot de

geadsorbeerde complexen nauwelijks een interactie met het drager­

materiaal en kunnen zich over het oppervlak van de drager bewegen.

Zodoende kunnen zij het reductiepraces van andere, moeilijker redu­

ceerbare platina complexen versnellen . Zo ontstaan er kleine

metaalclustertjes die zich over het oppervlak van de drager bewegen,

niet gereduceerde complexen reduceren en dus steeds grater warden.

Wanneer er grotere zoutkristallen op het dragermateriaal aanwezig

zijn, dan kunnen die de 'randzwervende' metaalatomen en kleine

metaalclusters invangen en zodoende het reductieproces afremmen.

Het is gebleken dat ook het oplosmiddel een belangrijke in vloed op

het bereidingspraces heeft. Gebruiken we een alkohol in plaats van

water (zoals in het hiervoorgaande het geval was) dan vindt er geen

adsorptie plaats en kristalliseren de zouten tijdens het draogpraces

afzonderlijk uit . Er ontstaan RhCl3 and H2PtCl6 kristallen die ook

afzonderlijk warden gereduceerd tijdens het reductieproces.

In hoof dstuk 4 hebben we met behulp van ESR-experimenten

kunnen aantonen dat Fe in bimetallische Fe- Pt en Fe-Ru katalysa­

toren voor een graot gedeelte in de vorm van Fe3+ aanwezig is . Er

waren al aanwijzingen vanuit de techniek van Mossbauer spectros­

copie dat een groot gedeelte van het ijzer driewaardig was, maar de

interpretatie van de spectra was niet eenduidig. ESR heeft nu

onomstotelijk kunnen aantonen dat dat wel het geval is , omdat we

met ESR Fe3+ ionen niet kunnen verwarren met andere Fe"+ ionen.

In katalysator-systemen waar ijzer aanwezig is zonder een ander

Page 243: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 233 Samenv att ing

3+ 2+ {edel)metaal, wordt Fe gereduceerd tot Fe en ook verder. Dat

betekent dat in dit geval het edelmetaal een heel aparte rol vervult :

het houdt het ijzer in een hoge valentietoestand , of het verhindert

dat het ijzer gereduceerd wordt. Dit geheel in tegenstelling tot de bevindingen in hoofdstuk 3, waar rhodium de reductie van platina

aanzienlijk kon versnellen. Het model voor de bimetallische Fe-M

katalysatoren is nu als volgt : beide metalen houden elkaar in een

hoog-disperse toestand en in die toestand kan ijzer niet meer gereduceerd worden.

In hoof dstuk 5 hebben we het passiveren van een Rh/ Al 20 3 katalysator besproken. Passiveren is voorzichtig oxideren, zodat we

de katalysator zonder problemen aan de lucht kunnen blootstellen . Nadat de katalysator was gereduceerd en geevacueerd hebben we

zuurstof toegelaten bij 100 K en bij 300 K. EXAFS spectra toonden

aan dat bij 100 K oxidatie begon en dat bij 300 K nog slechts een

kleine kern van de oorspronkelijke deeltjes metallisch is. De rest

van het rhodium is geoxideerd tot rhodium-oxide en bedekt voor een

gedeelte het metallische kerntje. De rhodiumatomen die in contact

zijn met het drager materiaal en de rhodiumatomen die in contact

zijn met het bedekkende rhodiumoxide hebben zuurstofburen op een

afstand van ongeveer 2.7 A. Op grond van Rh0-o2--coordinatie

getallen hebben we een eenvoudig model opgesteld waarmee we het beginnende oxidatieproces kunnen beschrijven.

Met behulp van een computerprogramma hebben we in

hoof dstuk 6 de interacties tussen een rhodium-metaaldeeltje en het

dragermateriaal y-Al 20 3 bestudeerd. Het programma was

gebaseerd op de ASED-MO methode. Wanneer het y -Al 20 3

oppervlak onder het metaaldeeltje volledig uit -OH groepen bestond,

een volledig gehydrateerd oppervlak, was de binding tussen metaal­

deeltje en drager energetisch het meest gunstig. De rhodiumatomen in het metaal-drager grensvlak prefereren dan sites waarin ze drie

-0 H groepen als naaste buren hebben . De Rh-0 af stand in het

metaal-drager grensvlak is ongeveer 2.5-2.6 A. Wanneer het

oppervlak onder het metaaldeeltje volledig uit -02- ionen bestaat ,

Page 244: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Samenvatting page 234

een volledig gedehydrateerd oppervlak, is de binding tussen metaal­deeltje en drager iets zwakker. Toch ziet de binding er totaal anders uit : de rhodiumatomen in het metaal-drager grensvlak prefereren nu posities boven op de zuurstof atomen en de Rh-0 afstand is nog maar 2.1 A. Het blijkt dus, dat de protonen in de hydroxylgroepen een belangrijke rol spelen in de binding tussen het metaaldeeltje en de drager. De protonen beinvloeden de on-top binding ongunstig en de binding in drievoudige sites daarentegen gustig. Een belangrijks resultaat in dit hoofdstuk was dat de metaal-zuurstof afstanden in het metaal-drager grensvlak zoals die gevonden zijn met EXAFS heel aardig kloppen met de afstanden die we op deze manier, op grond van theoretische berekeningen gevonden hebben .

In hoofdstukken 7 en 8 hebben we EXAFS en HRTEM gebruikt om de struktuur te bepalen van Rh/Ti0 2 katalysatoren. Deze katalysatoren kunnen in de SMSI toestand worden gebracht, een toestand waarin de eigenschappen van de metaaldeeltjes dras­tisch zijn veranderd. Een van de monsters was gemaakt in Eindho­ven en de EXAFS spectra zijn gemeten in Dares bury, het andere sample was gemaakt door de groep van prof. Haller in Yale Univer­sity (U.S.A.) en de EXAFS spectra van die groep zijn gemeten in Cornell. De resultaten van beide studies zijn vergelijkbaar . Voor het monster van Eindhoven vonden we dat na een reductie in H2 bij 473 K (de katalysator is dan in de 'normale ' toestand) de metaal­deeltjes gemiddeld ongeveer vijf rhodiumatomen bevatten en zich bevinden op hoeken van de TiOrkristallen en op 11011 en 10011 vlakken van de drager. Na een reductie in H2 bij 723 K (de kataly­sator is dan in de SMSl-toestand) zijn er nogal wat veranderingen in het EXAFS spectrum. De rhodium atomen in het metaal-drager grensvlak hebben nu niet alleen zuurstofburen van de drager, maar ook nog Ti"+ ionen van de drager op afstanden van 3.4 en 4.3 A. De coordinatie-getallen en -afstanden kloppen heel ardig met een model waarin de metaaldeeltjes rusten op gereduceerd Ti0 2. Hieruit blijkt dat inderdaad tijdens het reductieproces bij hogere tempera­tuur de drager in de buurt van het metaaldeeltje mee kan reduceren. We hebben geen aanwuzmgen gevonden voor bedekking van het

Page 245: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 235 Samenvatting

metaaldeeltje met dragermateriaal. In tegendeel, wanneer we

zuurstof toelaten bij de katalysator in de SMSI toestand kunnen we

in het EXAFS spectrum zien dat er zuurstof aan de metaaldeeltjes is

geadsorbeerd. Dus moet op zijn minst een gedeelte van het

oppervlak van de deeltjes onbedekt zijn. Ook werden de metaal­

deeltjes die bedekt waren met zuurstof niet geoxideerd, terwijl een

Rh/ A1 20 3 katalysator onder dezelfde omstandigheden wel wordt

geoxideerd. Hieruit blijkt , dat in de SMSl-toestand, als er al

bedekking optreedt, deze bedekking niet verantwoordelijk kan zijn

voor het abnormale gedrag van de metaaldeeltjes . We concluderen

dan ook dat een electronische interactie tussen drager en metaal­

deeltje verantwoordelijk moet zijn voor de verandering in de eigen­

schappen van de metaaldeeltjes in de SMSl-toestand .

De Rh/Ti0 2 katalysator die bereid was in, en waarvan de spec­

tra gemeten waren door, de groep van prof. Haller van Yale Univer-

. sity kon in de EXAFS-cel slechts moeilijk gereduceerd worden.

Zelfs na een reductie bij 628 K was er nog rhodiumoxide aanwezig.

Door die vertraging in het reductieproces kwam de katalysator ook pas bij een hogere temperatuur in de SMSl-toestand. Dat kan er de

oorzaak van zijn dat de metaaldeeltjes zijn gaan sinteren. Voordat sintering optrad, waren de deeltjes erg klein en ongeveer even groot

als de metaaldeeltjes in de Rh/Ti0 2 katalysator van Eindhoven : ongeveer 5 atomen per metaaldeeltje. Na sintering waren de

deeltjes aanzienlijk groter : ze bevatten elk ongeveer 15 rhodiumato­

men . Eenmaal in de SMSl-toestand is er een sterke interactie, een

sterke binding tussen drager en metaaldeeltje die ervoor kan zorgen

dat er geen sintering meer optreedt . Zoals voor de Rh/Ti0 2 kataly­

sator van Eindhoven, bleek dat ook voor deze katalysator in de 'nor­

male ' toestand de metaaldeeltjes op niet gereduceerd Ti0 2 en in de

SMSI toestand op gereduceerd Ti0 2 liggen . Wederom waren er geen

aanwijzingen voor bedekking. Ook nu moeten we concluderen dat

een interactie tussen metaaldeeltje en drager de oorzaak is van de

eigenschappen van de metaaldeeltjes in de SMSl-toestand.

Page 246: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

I I .

I

Samenvatting page 236

Tenslotte hebben we in hoofdstuk 9 een vergelijkbaar katalysa­torsysteem bestudeerd : Rh/Ta20 5. Na redutie in H2 bij 523 K waren de rhodiumdeeltjes volledig gereduceerd, er was geen rhodi­umoxide meer aanwezig. Na reductie bij 858 K was de katalysator in de SMSl-toestand. In de SMSl-toestand hadden de rhodium-atomen in het metaaldrager · grensvlak en de rhodiumatomen in het oppervlak van de metaaldeeltjes Tan+ ion en als buren op heel korte afstanden : 1.5-2.3 A. Dus waren deze metaaldeeltjes in de SMSl­toestand wel (gedeeltelijk) bedekt met gereduceerd dragermateriaal. We hebben verder geen aanwijzingen gevonden voor de vorming van een RhTa-legering. Ook hebben we kunnen uitsluiten dat in de SMSl-toestand de metaaldeeltjes zich spreiden over het dragerop­pervlak.

Concluderend, voor twee verschillende systemen (Rh/Ti02 en Rh /T a20 5) hebben we verschillende resultaten gezien wanneer de systemen in de SMSl-toestand gebracht werden . Voor de Rh/Ti0 2

katalysatoren vonden we geen aanwijzingen voor bedekking, voor de Rh/Ta20 5 katalysator wel. We konden echter niet uitsluiten dat de rhodium-metaaldeeltjes in de Rh/Ti0 2 katalysatoren in de SMSl­toestand gedeeltelijk, maar dan wel losjes bedekt waren met TiOx. De Rh/T a20 5 katalysator was echter bij veel hogere temperatuur gereduceerd dan de beide Rh/Ti02 katalysatoren (723 en 773 K). Het is heel goed mogelijk die losse bedekking na reductie bij hogere temperatuur overgaat in epitaxiale bedekking, en die bedekking kun­nen we met EXAFS wel aantonen.

Page 247: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 237 Dankwoord

Dankwoord

Een promotie-onderzoek is iets dat een promovendus zeker niet

in zijn eentje kan volbrengen. Aan het proef sch rift dat U nu in Uw handen heeft hebben dan ook heel veel mensen een steentje

bijgedragen. ledereen die ook maar enigszins heeft bijgedragen aan het voltooien van het mijn onderzoek en het tot stand komen van

dit. proefschrif t dank ik van ha rte .

Om te beginnen natuurlijk Roel Prins. Roel, je was een fantas­

tische begeleider. Zeker op wetenschappelijk gebied denk ik dat ik me geen betere eerste promotor had kunnen wensen . Zonder alle

verhelderende discussies, zonder jouw vele kritische kanttekeningen en zonder jouw snelle correcties op de vele proef versies van dit werk

zou dit boekje nooit tot stand zijn gekomen. Op een eerlijke tweede plaats, Diek Koningsberger. Beste Diek, jij bent degene die me de

techniek van EXAFS in alle geuren en kleuren hebt bijgebracht. Jij was degene die de aandacht wist te vestigen op kleine bijdragen in

de spectra, kleine verschillen die vaak grote gevolgen hadden. Jij bent dan ook degene die 'aan de wieg' heeft gestaan van een groot

aantal onderwerpen in dit proefschrif t.

Dan is er natuurlijk de rest van de vakgroep Anorganische

chemie, en in die vakgroep neemt de koffieclub van de groep metaal­katalyse de belangrijkste plaats in. De leden van die. koffieclub ben

in natuurlijk alle heel dankbaar voor de dagelijkse rustpunten. lk denk dat ik het zonder de gezellige koffiepauzes en de enervende

spelletjes bridge minder rustig zou hebben afgebracht . Maar natu­urlijk is hun bijdrage veel meer geweest. In de eerste plaats ben ik

veel dank verschuldigd aan het technisch vernuft van de groep metaalkatalyse, Joop van Grondelle, die konstant in de weer was om

alle apparatuur draaiende te houden. Dan natuurlijk mijn dagelijkse collega's, de mede-lotgenoten/promovendi. Op de eerste plaats Bert Kip, die samen met Joop van Grondelle al die jaren met mij een kamer heeft gedeeld . Frans Kampers, Fanny van Zon en Joop van

Page 248: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Dankwoord page 238

Grondelle hebben het allereerste begin van veel van de EXAFS stu­dies meegemaakt. Zonder hun hulp in Daresbury hadden we nooit zoveel metingen kunnen verrichten en was dit boekje aanzienlijk dunner geweest . Ad de Koster, niet in het minst omdat hij me op weg heeft geholpen met het ASED-MO programma. Rutger van Santen wil graag danken voor zijn inspanningen en wetenschap­pelijke bijdrage aan de ASED-MO berekeningen. Dick van Langeveld voor zijn altijd verhelderende discussies en kritische kanttekeningen . En zeker niet in de laatste plaats Hans Niemantsverdriet voor de prettige samenwerking tijdens de ESR experimenten.

Ook de hulp van collega's van de Rijks Universiteit in Leiden is onmisbaar geweest voor de metingen in Daresbury : met name Hans den Hartog. Marjan Botman en Frans Mijlhof ben ik hiervoor erken­telijk.

In het kader van hun praktikum of afstudeeronderzoek hebben een aantal studenten bijgedragen aan verschillende hoof dstukken van dit proefschrift : John Jansen, Theo van Dijk, Peter Wijnen, Mark Savelsberg, Ton Janssens, Arthur de Jong, Tom Janssens, Toine Ketelaars, Frank van Doormalen, Mark Brouwer, Frits de Koning en Jan van Casteren . Hen wil ik bedanken voor hun inzet, doorzettingsvermogen en enthousiasme.

T enslotte zijn er nog een aantal mensen in de vakgroep die ik zeker niet mag vergeten in dit dankwoord : Wout van Herpen voor zijn technische ondersteuning, Adelheid Elemans-Mehring voor de vele analyses, Frans Sanders voor leveren van al het glaswerk en andere magazijn artikelen en Henk van Lieshout voor alle adminis­tratieve beslommeringen die hij steeds snel en vakkundig wist af te handelen .

Dan ben ik, TGTAJM, de medewerkers in het rekencentrum veel dank verschuldigd voor hun niet aflatende inspanningen om alle computers optimaal te laten draaien . In de eerste plaats Gert Jan Visser, die alle software voor de EXAFS data analyse op touw heeft

Page 249: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 239 Dankwoord

gezet en voortdurend aan onze steeds veranderende eisen aanpast.

Dan, op de tweede plaats, Peter Bregman. de VAX/VMS systeem manager, die de VAXen ondanks de vele EXAFS data-analyses in

optima forma weet te houden en vele problemen wist op te lossen. Dan natuurlijk Frans Galle, die de zware taak heeft de VAX/UNIX

ondanks de vele en zware troff-processen draaiende te houden. Ook Tony van Langeveld voor zijn welwillende hulp bij het opstarten van

de ASED programma's op het IBM mainframe. En tenslotte alle medewerkers van de balie die niet alleen snel alle probelemen met het computerpark oplossen, maar ook dagelijks de vaak enorme sta­pels output van de printers sorteren en ervoor zorgen dat ze bij de

goede gebruikers terecht komen. Zonder jullie plezierige medewerk­ing zoud veel van het rekenwerk van TGTAJM niet zo succesvol zijn afgerond !

Many of the catalysts discussed in this thesis have been stu­

died with EXAFS. The EXAFS measurements have been performed at the Synchrotron Radiation Source in Daresbury (U. K.). I would like to express my gratitude to the staff of .the SRS Laboratory for their skilful! and ever ready assistance. Especially dr. G. Daikun,

station master of the EXAFS station 9.2, for his assistance during the set-up of the station to our needs , dr. N. Greeves, coordinator

of line 9 and Alf Neild, the beamline technician.

T enslotte, mijn vrouw Angeliene, wie ik meest dank verschul­

digd ben. Omdat ze ondanks de gezelligheid die ze , vooral gedurende het laatste jaar, 's avonds heeft moeten ontberen, al het geduld dat ze heeft moeten opbrengen, me altijd ter zijde staat. En ook mijn ouders, die me voortdurend hebben gestimuleerd tijdens mijn studie

en promotie, wil ik op deze plaats bedanken.

Th is dissertation has been supported by the Netherlands foun­

dation for Chemical Research (SON) with financial aid from the Organization for the Advancement of Pure Research (ZWO).

Page 250: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Curriculum Vitae page 240

Curriculum Vitae

Johannus Hubertus Anna Martens werd geboren in Elsloo (L) op 18 maart 1958. Na het be.halen van het eindexamen Gymnasium f3 aan de Scholengemeenschap Sint Michiel te Geleen in 1976 begon hij de studie Chemische T echnologie aan de T echnische Universiteit in Eindhoven . Na zijn afstudeer onderzoek, dat gericht was op de struktuur van kobalt en kobalt-rhodium katalysatoren gedragen op titaandioxide, onder leiding van prof. dr. R. Prins in de vakgroep Anorganische Chemie, studeerde hij af in augustus 1983. Op 1 sep­tember 1983 trad hij in dienst van de Nederlandse Organisatie voor Zuiver Wetenschappelijk Onderzoek en begon een promotieonder­zoek dat gericht was op het bestuderen van gedragen metaalka­talysatoren. Het onderzoek werd uitgevoerd onder leiding van prof. dr. R. Prins en prof. dr. ir. D. C. Koningsberger. De resultaten van dit onderzoek zijn in dit proefschrift beschreven . Op 21 juni 1986 trad hij in het huwelijk met Angelina Theresia Carolina van den Hof.

Page 251: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

page 241 Lijst van Publicaties

Lijst van Publicaties

1. Characterization of Supported Cobalt and Cobalt-Rhodium Catalysts : II. TPR and TPO of Co/Ti0 2 and Co-Rh/Ti02 Martens. J. H. A.: van 't Blik H. F. J.: Prins, R. }. Catal. 1986, 97, 200

2. Ferric Iron in Reduced Si02 Supported Fe-Ru and Fe-Pt Catalysts : Evidence from Mossbauer Spectrtoscopy and Elec­tron Spin Resonance Martens, J. H. A.: Prins, R. Niemantverdriet, J. W. }. Cata!. 1987. 108, 259

3. Preparation and Characterization of Very Highly Dispersed Iri­dium on Al20 3 and Si02 Catalysts Kip, B. J.: van Grondelle, J .: Martens. J. H. A.: Prins, R. Appl. Catal. 1986, 26, 353

4. The Stucture of a Rh/Ti02 Catalyst in the Strong Metal Sup­port Interaction State Determined by EXAFS Koningsberger, D. C.; Martens, J. H. A.; Prins, R.; Short, D. R.; Sayers. D. E. }. Phys. Chem. 1986, 90, 3047

5. The Structure of Rh/Ti02 in the Normal and the SMSI State as Determined by EXAFS and HRTEM Martens. J. H. A.; Prins. R.; Zandbergen, H.; Koningsberger, D. C. J. Phys. Chem., 1n press

6. Controlled Oxygen Chemisorption on an Alumina Supported Rhodium Catalyst : The Formation of a new Metal-Metal Oxide Interface Determined with EXAFS Martens, J. H. A.: Prins, R. Koningsberger, D .. C. J. Phys. Chem., to be published

Page 252: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Lijst van Publicaties page 242

7. Preparation of Monometallic Rh and Pt and Bimetallic Rh-Pt Catalysts Supported on y-Al 20 3 Martens. J. H. A.: Prins, R. to be published

8. The Structure of the Metal-Support Interface m Rh/ A1 20 3 Determined with ASED Martens, J. H. A.: van Santen, R. A.: Prins , R. to be published

Page 253: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

Stellingen

Behorende bij het proefschrift

A Spectroscopic Characterization of the

Structure of Supported Metal Catalysts

van

J. H. A. Martens

Page 254: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- II -

Stellingen

1. lnterne silanolgroepen in ZSM-5 met een hoog siliciumgehalte

zijn niet afkomstig van verbroken en gehydrateerde silicium­

zuurstof-siliciumbindingen in vierringen. zoals verondersteld door Nagy et al. en Boxhoorn et al. Men kan namelijk aantonen dat

vier van deze groepen vacatures omringen van T-atomen (T =Al,

Si), die willekeurig verspreid zijn door het rooster .

Nagy , J. B.; Gabelica , Z.; Derouane, E. G.; Jacobs . P A. Chem. Lett. 1982, 2003

Boxhoorn , G.; Kortbeek, A.G. T. G. ; Hays. G. R.; Alma . N. C. M. Zeolites 1984 4, 15

Kraushaar, B.; van de Ven, L. J . M.; de Haan , J . W.; van Hooff, J . H. C. •Studies in Surface Science and Catalysis". Elseviers Science Publishers, Amsterdam, 1988 (in press)

2. In tegenstelling tot hetgeen Rezek er al. veronderstellen. zijn de

door hen gepubliceerde data over het verband tussen laagdikte

en groeitijd van lnGaPAs kristallen, wel degelijk in overeenstem­

ming met normale diffusiebeperkte kristalgroei. als de transport­

snelheid maar in de berekeningen wordt meegenomen.

Rezek, A. E. ; Vojak , B. A.; Chin , R.; Holonyak , Jr . J . Electronic Mater.1981 , JO (1), 255

3. De door Williams en Nelson gevonden temperatuuraf­

hankelijkheid van de oppervlaktesegregatie in PtRh-legeringen is

waarschijnlijk geen intrinsieke eigenschap van het systeem.

Gezien de door de auteurs gebruikte experimentele condities kan

de invloed van verontreinigingen niet uitgesloten worden . Daar­

naast kan bij lage temperaturen door diff usielimitering het

preparaat zich in een niet-evenwichtstoestand bevinden.

Williams , F. L.; Nelson , G. C. Appl. Surf. Sci. 1979, 3, 409

van Delft , F. C. M. J . M. ; van Langeveld , A. D. ; Nieuwenhuys , B. E. Surf. Sci 1987. 1891190. 1129

Page 255: A spectroscopic characterization of the structure of ... · 2.4 Electron Spin Resonance Spectroscopy 18 2.5 Nuclear Magnetic Resonance Spectroscopy 20 2.6 Mossbauer Spectroscopy 21

- Ill -

4. Het verdient aanbeveling om in Raman-spectroscopische studies

naar de rek van hoge-modulus-vezels meer aandacht te besteden

aan de polarisatierichting van de invallende laserbundel ten

opzichte van de vezel.

Bool, R. P.; BretzlafL R. S.: Boyd, R. H. J. Pol. Sci., par1 B :

Polymer Physics 1986, 24, 1039

Robinson. J. M.; Yeung. P. H. J.; Galiotis, C.: Young. R. J.:

Batcheler. D N. J. Mal. Sci. 1986, 21, 3440

5. Bij het bestuderen van selectiviteiten naar zuurstofhoudende

produkten in de inloopperiode in de Fischer-T ropsch-synthese

wordt te weinig rekening gehouden met het feit dat dragerma­

terialen de gevorrnde verbindingen kunnen adsorberen. Met

name y-Al 20 3 kan alcoholen sterk adsorberen. De bevindingen

van Kip et al .. waarin veranderingen in de oxo-selectiviteiten

gerelateerd worden aan chloorgehaltes zijn waarschijnlijk geen

intrinsieke eigenschappen van het katalysatorsysteem. maar kun­

nen verklaard worden aan de hand van het adsorberend vermo­

gen van het dragermateriaal.

Kip, B. J.; Dirne, F. W. A.; van Grondelle. J.; Prins, R. Am.

Chem. Soc., Div. Pelr. Chem., Mechanisms of Fischer Trospch

Chemistry, 1986, 31, 43

Kip, B. J.; Smeets, P. A. T.; van Grondelle, J: Prins. R. Appl. Catal. 1987, 33, 181

6. De gevoeligheid en de kwaliteit van verschillende EXAFS­

opstellingen kan beter vergeleken worden aan de hand van nog

meetbare gewichtsfracties dan aan de hand van nog meetbare

concentraties van het adsorberend element.

7. Het feit dat EXAFS-metingen aan kobalt-molybdeensulfide

hydrotreatingkatalysatoren tot nu toe geen directe aanwijzingen

hebben gegeven voor een interactie tussen kobalt en molybdeen

kan teruggevoerd worden op hetzij een te lage signaal/ruis­

verhoud ing. hetzij een ontoereikende data-anal yseprocedure.


Recommended